This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The influence of crystalline electric field on the magnetic properties of CeCdX33{}_{3}X_{3} (XX = P and As)

Obinna P. Uzoh1, Suyoung Kim1, and Eundeok Mun1,2 1 Department of Physics, Simon Fraser University, 8888 University Drive, Burnaby, B.C. Canada 2 Center for Quantum Materials and Superconductivity (CQMS), Sungkyunkwan University, Suwon 16419, South Korea
Abstract

CeCd3P3 and CeCd3As3 compounds adopt the hexagonal ScAl3C3-type structure, where magnetic Ce ions on a triangular lattice order antiferromagnetically below TNT_{\text{N}}\sim0.42 K. Their crystalline electric field (CEF) level scheme has been determined by fitting magnetic susceptibility curves, magnetization isotherms, and Schottky anomalies in specific heat. The calculated results, incorporating the CEF excitation, Zeeman splitting, and molecular field, are in good agreement with the experimental data. The CEF model, with Ce3+ ions in a trigonal symmetry, explains the strong easy-plane magnetic anisotropy that has been observed in this family of materials. A detailed examination of the CEF parameters suggests that the fourth order CEF parameter B43B_{4}^{3} is responsible for the strong CEF induced magnetocrystalline anisotropy, with a large abab-plane moment and a small cc-axis moment. The reliability of our CEF analysis is assessed by comparing the current study with earlier reports of CeCd3As3. For both CeCdX33{}_{3}X_{3} (XX = P and As) compounds, less than 40 % of Rln(2)R\ln(2) magnetic entropy is recovered by TNT_{\text{N}} and full Rln(2)R\ln(2) entropy is achieved at the Weiss temperature θp\theta_{p}. Although the observed magnetic entropy is reminiscent of delocalized 4ff-electron magnetism with significant Kondo screening, the electrical resistivity of these compounds follows a typical metallic behavior. Measurements of thermoelectric power further validate the absence of Kondo contribution in CeCdX33{}_{3}X_{3}.

I Introduction

Several ff-electron materials with triangular lattices (TL) have shown rich phenomena, where the spin-orbit coupling enhances quantum fluctuations due to the highly anisotropic interactions between 4ff moments [1, 2, 3, 4, 5, 6]. For example, a spin liquid state has been proposed in insulating YbMgGaO4 [7, 8, 9, 10, 11] and NaYbS2[12, 13]. The low carrier density system YbAl3C3 shows a gap in the magnetic excitation spectrum due to the dimerization of the ff electrons in Yb3+ pairs [14, 15, 16, 17]. Of interest is the easy-plane antiferromagnets CeCdX33{}_{3}X_{3} (XX = P and As), a new class of TL system, with a low antiferromagnetic ordering temperature and extremely low carrier density [18, 19, 20, 21, 22].

Refer to caption
Figure 1: (a) Crystal structure of the hexagonal CeCdX33{}_{3}X_{3} (X=X= P and As). (b) Ce atoms forming two-dimensional triangular lattices. (c) Ce atom in a trigonal (D3dD_{3d}) point symmetry surrounded by P/As atoms.

The family of compounds CeM3X3M_{3}X_{3} (MM = Al, Cd, and Zn, XX = C, P, and As) have been investigated for their robust ground state properties arising from the interplay between crystalline electric field (CEF) and magnetic exchange interaction on the triangular lattice [16, 18, 19, 20, 21, 22, 23, 24]. CeCd3XX3 materials adopt the hexagonal ScAl3C3-type structure (space group P63/mmcP6_{3}/mmc) with the magnetic Ce3+ ions having a trigonal (D3dD_{3d}) point symmetry as shown in Fig. 1(c) [20, 21, 22, 23]. In this crystal structure, the Ce layers are well separated by the Cd and XX atoms, forming a layered 2D TL in the abab-plane, as depicted in Figs. 1(a) and (b) [20, 21, 22, 23]. Thermodynamic and transport property measurements have characterized CeCd3XX3 as an low carrier density system, with a strong easy plane magnetic anisotropy and antiferromagnetic ordering below TNT_{\text{N}} \sim0.42 K [19, 22]. The low carrier density means there are insufficient charge carriers to screen all local moments and to mediate Ruderman-Kittel-Kasuya-Yosida (RKKY) exchange interaction between moments. In this regard, both RKKY and Kondo interactions are expected to be weakened in these compounds. We note that the electrical resistivity measurement on CeCd3As3 grown by chemical vapor transport shows a semiconductor-like enhancement as decreasing temperature  [20], while the flux grown CeCd3As3 sample is metallic [19]. At the magnetic ordering temperature, roughly 40% of Rln(2)R\ln(2) entropy is recovered. This implies a doublet ground state resulting from CEF splitting of localized Ce ion energy levels. The highly enhanced specific heat below 10 K and the reduced magnetic entropy at TNT_{\text{N}} are reminiscent of Kondo lattice materials. However, the electrical resistivity of CeCdX33{}_{3}X_{3} shows no maximum or logarithmic upturns resulting from the Kondo scattering of conduction electrons from magnetic Ce ions. The resistivity of CeCdX33{}_{3}X_{3} is the same as that of LaCdX33{}_{3}X_{3}, implying an absence of Kondo screening in these materials [19, 22].

Here, we use the CEF scheme to clarify the anisotropic magnetic properties of CeCdX33{}_{3}X_{3} (XX = P and As), where the CEF analysis is performed using the PyCrystalFieldPyCrystalField package [26]. For CeCd3As3, the CEF analyses of two previous independent studies [20, 25], carried out only considering magnetic susceptibility and magnetization, have shown discrepancies in their energy level splittings and first excited state wave functions. Thus, we extend the CEF analysis to specific heat data to resolve these discrepancies. In contrast to CeCd3As3, no CEF analysis has been carried out for CeCd3P3. We show that the easy plane magnetic anisotropy observed in these compounds can be explained by the strong CEF acting on Ce3+ ions. In addition, we provide evidence of the lacking Kondo screening in these materials by way of thermoelectric power measurements.

II Experimental

For this work, single crystals of LaCdX33{}_{3}X_{3} and CeCdX33{}_{3}X_{3} (XX = P and As) were grown by high temperature ternary melt [27, 22, 19]. Magnetization measurements as a function of temperature and magnetic field were performed in a QD MPMS. Specific heat of these compounds was measured in QD PPMS. The obtained results are consistent with earlier reports [19, 22]. Thermoelectric power measurements were performed in a home made two thermometer and one heater setup. Detailed studies of thermodynamic and transport properties of these compounds are presented in Ref. [19, 22].

III Results & Discussion

Refer to caption
Refer to caption
Refer to caption
Figure 2: (a) Inverse magnetic susceptibility curves of CeCd3As3, taken at HH = 1 kOe for both HabH\parallel ab and HcH\parallel c. (b) Magnetization isotherms at 1.8 K for both HabH\parallel ab and HcH\parallel c. (c) Magnetic part of the specific heat at HH = 0 and 90 kOe along HabH\parallel ab. Open symbols are experimental data. Dashed lines represent CEF fits with no molecular field contribution. Solid lines are CEF fits with molecular field terms λab=5.7\lambda_{ab}~{}=~{}-5.7 mole/emu and λc=1.0\lambda_{c}~{}=~{}1.0 mole/emu.
Table 1: CeCd3As3: CEF parameters, energy eigenvalues, eigenstates, and molecular field parameters.
CEF and molecular field parameters
B20B_{2}^{0} (K) B40B_{4}^{0} (K) B43B_{4}^{3} (K) λi\lambda_{i} (mole/emu)
18.55 -0.08 23.02 λab=5.7\lambda_{ab}=-5.7,
λc=1.0\lambda_{c}=1.0
Energy eigenvalues and eigenstates
E (K) |52|-\frac{5}{2}\rangle |32|-\frac{3}{2}\rangle |12|-\frac{1}{2}\rangle |12|\frac{1}{2}\rangle |32|\frac{3}{2}\rangle |52|\frac{5}{2}\rangle
0 0.0 0.0 -0.898 0.0 0.0 0.440
0 0.440 0.0 0.0 0.898 0.0 0.0
242 0.0 -1.0 0.0 0.0 0.0 0.0
242 0.0 0.0 0.0 0.0 -1.0 0.0
553 -0.898 0.0 0.0 0.440 0.0 0.0
553 0.0 0.0 0.440 0.0 0.0 0.898

For CeCd3XX3 systems, Ce atoms occupy a single site at the 2aa Wyckoff position that has a trigonal point symmetry, shown in Fig. 1(c). In this point symmetry, the CEF Hamiltonian requires only three parameters [28, 29]: CEF=B20O20+B40O40+B43O43{\cal H}_{CEF}=B^{0}_{2}O^{0}_{2}+B^{0}_{4}O^{0}_{4}+B^{3}_{4}O^{3}_{4}, where OnmO^{m}_{n} are the Stevens operators and BnmB^{m}_{n} are known as the CEF parameters [30, 31]. It has to be noted here that thermodynamic and transport property measurements of these compounds indicate a possible structural phase transition below 132 K [16, 19, 22]. Additional CEF parameters below the transition temperature are probably required to account for the change in the local environment of the Ce atoms. However, the magnetic susceptibility curves show a smooth evolution through the transition temperature. Therefore, we assume in our CEF analysis that the Ce ions have the same trigonal symmetry below the transition temperature. To account for the Zeeman effect and the interaction between magnetic ions, the eigenvalues and eigenfunctions are determined by diagonalizing the total Hamiltonian: =CEFgJμBJi(Hi+λiMi){\cal H}={\cal H}_{CEF}-g_{J}\mu_{B}J_{i}(H_{i}+\lambda_{i}M_{i}) (ii = xx, yy, and zz), where gJg_{J} is the Landé gg-factor, μB\mu_{B} is the Bohr magneton, HiH_{i} is the applied field, JiJ_{i} is the angular momentum operator, and MiM_{i} is the magnetization. The second term is the Zeeman contribution and the third term represents the effective molecular interactions λi\lambda_{i}.

The point charge CEF model assumes that the magnetic ions are well localized with a stable valence state. The valence state of CeCd3XX3 can be deduced from the effective moment of Ce ions. Figures 2(a) and 3(a) show inverse magnetic susceptibility, χ1=H/M\chi^{-1}=H/M, curves of CeCd3As3 and CeCd3P3, respectively. At sufficiently high temperatures, magnetic susceptibility curves follow the Curie-Weiss (CW) behavior: χ(T)=C/(Tθp)\chi(T)=C/(T-\theta_{p}), where CC is the Curie constant and θp\theta_{p} is the Weiss temperature. The effective moments of CeCd3P3 and CeCd3As3 are estimated by applying the CW law to the polycrystalline averaged magnetic susceptibility to be μeff=2.51μB\mu_{eff}=2.51~{}\mu_{B} and μeff=2.54μB\mu_{eff}=2.54~{}\mu_{B}, respectively; which agree very well with the theoretical value μeff=2.54μB\mu_{eff}=2.54~{}\mu_{B}. From the effective moment values, it is reasonable to assume that Ce ions in these compounds are well localized with 3+ valence state. As shown in Ref. [32, 25] the B20B^{0}_{2} parameter mainly gives a measure of the magnetocrystalline anisotropy and can be expressed in terms of the Weiss temperatures [33, 34, 35]. From the Curie-Weiss fit (not shown), anisotropic Weiss temperatures of CeCd3As3 are estimated to be θab9.3\theta_{ab}\sim 9.3 K and θc=283\theta_{c}=-283 K. For CeCd3P3, θab=9.3\theta_{ab}=9.3 K and θc=248\theta_{c}=-248 K are obtained. The obtained θp\theta_{p} values are consistent with earlier reports [19, 22]. These values are used to estimate the leading CEF parameters: B2030.5B^{0}_{2}\sim 30.5 K for CeCd3As3 and B2026.7B^{0}_{2}\sim 26.7 K for CeCd3P3.

Figure 2 displays anisotropic magnetic susceptibility curves, magnetization isotherms, and specific heat of CeCd3As3 with the results of CEF analysis. Table 1 summarizes the obtained CEF parameters, including molecular field parameters, and the energy eigenstates and eigenvalues. The 2J+12J+1 degenerate levels for J=5/2J=5/2 of Ce3+ split into three Kramers doublets with energy level splittings Δ1\Delta_{1} = 242 K and Δ2\Delta_{2} = 553 K. The ground state and second excited state are in a mixture of |±5/2|\pm 5/2\rangle and |1/2|\mp 1/2\rangle states. However, the first excited state is a pure |±3/2|\pm 3/2\rangle state.

First, we discuss the CEF effects on thermodynamic properties of CeCd3As3 without the molecular field contribution. In Fig. 2, blue dashed-lines represent the CEF evaluations with λi\lambda_{i} = 0. The CEF fit generally agrees with the inverse magnetic susceptibility as the large anisotropy between crystallographic directions is captured. For HcH\parallel c the broad hump around 25 K is well reproduced. The CEF model aligns very well with the magnetization isotherm, M(H)M(H), for HcH\parallel c at 1.8 K, whereas there is an inconsistency between the CEF calculation and M(H)M(H) for HabH\parallel ab, as shown in Fig. 2(b). The magnetic contribution to the specific heat, CmC_{m}, at HH = 0 and 90 kOe is presented in Fig. 2(c), where CmC_{m} is estimated by subtracting the specific heat of non-magnetic analog LaCd3As3. In zero field, the broad maximum near 130 K in CmC_{m} can be reproduced by the CEF model. Note that the CEF calculation above the maximum cannot capture the experimental data, which is caused by the very large subtraction error as explained in Ref. [19]. Obviously, the sharp rise of CmC_{m} below 10 K and the sharp peak at TNT_{\text{N}} = 0.42 K [19] are not captured by the CEF model. At HH = 90 kOe for HabH\parallel ab, the overall shape of CmC_{m} is captured by the CEF model, but the maximum temperature is higher than that of experimental data.

The CEF model without λi\lambda_{i} does not adequately reproduce M(H)M(H) and CmC_{m} data for HabH\parallel ab. In order to account for this mismatch, the molecular field interactions between Ce3+ ions are incorporated. Red solid-lines in Fig. 2 represent the CEF model in the presence of the molecular field terms: λab=5.7\lambda_{ab}=-5.7 mole/emu and λc=1\lambda_{c}=1 mole/emu. The magnetic susceptibility curves with λi\lambda_{i} show a slightly better agreement than that with λi=0\lambda_{i}=0. However, as can be clearly seen in Fig. 2(b), the M(H)M(H) curve for HabH\parallel ab is well captured by the combination of the CEF and λi\lambda_{i}. Moreover, although the absolute value of the maximum in CmC_{m} at 90 kOe is slightly higher, the position of the maximum temperature is well reproduced by introducing λi\lambda_{i}. These results point to the importance of including the exchange interactions between Ce3+ magnetic ions.

Table 2: CeCd3P3: CEF parameters, energy eigenvalues, eigenstates, and molecular field parameters.
CEF and molecular field parameters
B20B_{2}^{0} (K) B40B_{4}^{0} (K) B43B_{4}^{3} (K) λi\lambda_{i} (mole/emu)
20.90 -0.03 26.00 λab=6.2\lambda_{ab}=-6.2,
λc=0.3\lambda_{c}=0.3
Energy eigenvalues and eigenstates
E (K) |52|-\frac{5}{2}\rangle |32|-\frac{3}{2}\rangle |12|-\frac{1}{2}\rangle |12|\frac{1}{2}\rangle |32|\frac{3}{2}\rangle |52|\frac{5}{2}\rangle
0 0.0 0.0 -0.897 0.0 0.0 0.442
0 0.442 0.0 0.0 0.897 0.0 0.0
257 0.0 -1.0 0.0 0.0 0.0 0.0
257 0.0 0.0 0.0 0.0 -1.0 0.0
621 -0.897 0.0 0.0 0.442 0.0 0.0
621 0.0 0.0 0.442 0.0 0.0 0.897
Refer to caption
Refer to caption
Refer to caption
Figure 3: (a) Inverse magnetic susceptibility curves of CeCd3P3, taken at HH = 1 kOe for both HabH\parallel ab and HcH\parallel c. (b) Magnetization isotherms at 1.8 K for both HabH\parallel ab and HcH\parallel c. (c) Magnetic part of the specific heat at HH = 0 and 90 kOe along HabH\parallel ab. Open symbols are experimental data. Dashed lines represent CEF fits with no molecular field contribution. Solid lines are CEF fits with molecular field terms λab=6.2\lambda_{ab}~{}=~{}-6.2 mole/emu and λc=0.3\lambda_{c}~{}=~{}0.3 mole/emu.

Now turning to the CEF analysis for CeCd3P3, following the same procedure applied to CeCd3As3. The isostructural compounds CeCd3P3 and CeCd3As3 show remarkably similar magnetic properties, implying that their local CEF environments are in close resemblance. Hence, it is expected that the CEF parameters will be quite similar for both compounds, giving rise to very similar CEF energy level splittings and eigenstates. Table 2 shows a summary of CEF fit results of CeCd3P3. As expected, the obtained CEF parameters, eigenstates, and eigenvalues for CeCd3P3 are very similar to those of CeCd3As3 (Table 1). The positive B20B_{2}^{0} term indicates that the magnetization lies in easy plane as seen in Fig. 3(b). The large B43B_{4}^{3} term implies a mixed ground state with |±5/2|\pm 5/2\rangle and |1/2|\mp 1/2\rangle, just like the case for CeCd3As3. The first excited state is in a pure state of |±3/2|\pm 3/2\rangle and the second excited state is in an admixture of |±5/2|\pm 5/2\rangle and |1/2|\mp 1/2\rangle states. The energy level splittings correspond to 257 K for the first excited state and 621 K for the second excited state.

Figure 3 shows magnetic susceptibility, magnetization, and specific heat curves of CeCd3P3 together with the CEF calculations. In the absence of molecular field terms, the CEF calculations (dashed-lines) agree very well with the experimental H/MH/M curves, as shown in Fig. 3(a). Also, in zero field CmC_{m} agrees with the CEF fit at high temperatures, as shown in Fig. 3(c), implying the high temperature broad maximum (Schottky anomaly) is due to the CEF effects. The inclusion of the molecular field terms greatly alters the M(H)M(H) curve calculation for HabH\parallel ab and adequately captures the experimental M(H)M(H) data. In addition, the CEF calculation with molecular field terms properly captures the low temperature maximum in CmC_{m} at HH = 90 kOe. Therefore, the CEF model in the present of the molecular field interaction should be used to describe the experimental data for CeCd3P3. As expected from the layered crystal structure, the absolute value of λab\lambda_{ab} is greater than λc\lambda_{c}, implying a strong exchange interaction in the abab-plane.

The obtained CEF parameters of CeCd3P3 are slightly larger than those of CeCd3As3. The slight difference in the values of their CEF parameters is quite surprising. When the distance between Ce and nearest As/P atoms is considered, the difference between Ce-As distance (3.05 Å) and Ce-P distance (2.96 Å) is about \sim9 pm [19, 20, 22], which might not be large enough to cause different CEF energy level splittings. Note that the validity of the CEF Hamiltonian must be verified below the (structural) phase transition temperature TsT_{s} = 127 K for CeCd3P3 [22] and TsT_{s} = 136 K for CeCd3As3 [19].

Refer to caption
Refer to caption
Figure 4: CEF profiles of CeCd3As3 for three independent studies. Study A is from Ref. [20] and Study B is from Ref. [25]. (a) CEF energy level splittings and eigenstates. Present study: α=0.44\alpha~{}=~{}0.44, β=0.90\beta~{}=~{}0.90, and γ=1\gamma~{}=~{}-1. Study A: α=0.32\alpha~{}=~{}0.32, β=0.95\beta~{}=~{}0.95, and γ=1\gamma~{}=~{}1. Study B: α=0.28\alpha~{}=~{}0.28, β=0.96\beta~{}=~{}0.96, and γ=1\gamma~{}=~{}1. (b) Comparison between CmC_{m} of present study and CEF calculations based on the three independent studies, where λi=0\lambda_{i}=0.

In Fig. 4(a) we compare the energy level splittings and eigenfunctions for three independent CEF analyses of CeCd3As3. For all three studies the CEF parameters are obtained from fits to the magnetic susceptibility curves. In this study, the Hamiltonian includes the molecular field term. In Ref. [20] (Study A), the CEF parameters are obtained in the presence of antiferromagnetic exchange interaction based on a mean-field approach. In Ref. [25] (Study B), the CEF parameters are evaluated by fitting the magnetic susceptibility curves to only CEF{\cal H}_{CEF} (without any interaction term).

For all three studies, the ground state (Γ1\Gamma_{1}) is in a mixed state of |±5/2|\pm 5/2\rangle and |1/2|\mp 1/2\rangle with a higher probability in the |1/2|\mp 1/2\rangle state. This requires a mixing angle θ\theta in the wave function: cos(θ)|±5/2+sin(θ)|1/2\cos(\theta)|\pm 5/2\rangle+\sin(\theta)|\mp 1/2\rangle. The obtained mixing angle is roughly similar for all three studies: 64 in this study, 72 in Ref. [20], and 74 in Ref. [25]. Unlike the ground state, the excited states Γ2\Gamma_{2} and Γ3\Gamma_{3} do not entirely agree for all three studies. In Ref. [20] there is a clear swap between Γ2\Gamma_{2} and Γ3\Gamma_{3} wave functions (Study A). This is caused by the relatively high |B40||B_{4}^{0}| value (= 1.4 K). In fact, we confirmed that any |B40|>0.7|B_{4}^{0}|>0.7 K would result in the Γ2\Gamma_{2} being in a mixed state and the Γ3\Gamma_{3} being in a pure state, just as the case in Ref. [20]. Another important distinction among these three studies is in the energy level splittings. The energy eigenvalue for the second excited state in Ref. [25] (Study B) is roughly two times smaller than that of the other two studies. However, the energy level splitting from the ground to the first excited state (\sim240 K) is comparable for all studies, clearly indicating that the ground state is well isolated from the excited states.

Figure 4 (b) shows the magnetic part of the specific heat, together with the calculated specific heat curves by using CEF parameters. Solid lines, dashed lines, and dotted lines are CEF calculations from the present study, Ref. [20], and Ref. [25], respectively. As shown in the figure, when subtle differences are ignored, the high temperature maximum is captured by all three calculated curves. This implies that the high temperature maximum in CmC_{m} is due to the CEF effect and the ground state doublet is well isolated from the excited states. When measurement uncertainty and different sample quality are considered, the best CEF parameters among three parameter sets cannot be selected from the comparison with CmC_{m}. Therefore, further measurements such as inelastic neutron scattering with Cd isotopes are necessary to distinctly specify the best CEF parameters in this system. In addition, the CEF scheme can be determined by optical spectroscopy techniques such as Raman scattering. Note that CEF parameters evaluated by the three independent studies provide a qualitatively good description of the experimental M/HM/H curves.

The significance of B43B_{4}^{3} CEF parameter has been observed in Ce-based antiferromagnets such as CeAuSn, CeIr3Ge7, and CeCdX33{}_{3}X_{3}, where Ce ions are in a trigonal environment [36, 28, 25]. These compounds indicate a large magnetic anisotropy with the abab-plane being the magnetic easy plane, which can be qualitatively explained by the CEF effect. A detailed CEF analysis based on both the magnetic susceptibility and inelastic neutron scattering data of hexagonal CeAuSn indicates a mixture of |±5/2|\pm 5/2\rangle and |1/2|\mp 1/2\rangle CEF ground state doublet, a pure |±3/2|\pm 3/2\rangle doublet as the first excited state at \sim345 K, and a mixture of |±5/2|\pm 5/2\rangle and |1/2|\mp 1/2\rangle doublet as the second excited state at \sim440 K [36, 28]. The anisotropy ratio of magnetic susceptibility between HabH\parallel ab and HcH\parallel c is found to be χab/χc15\chi_{ab}/\chi_{c}\sim 15 near TNT_{N}. Both magnetic susceptibility [28] and neutron scattering  [36] CEF evaluations clearly show a significant B43B_{4}^{3} contribution (\sim19 K) with a large mixing angle, which is consistent with the CEF analysis of CeCdX33{}_{3}X_{3}. The anisotropy ratio of CeCdX33{}_{3}X_{3} (χab/χc36\chi_{ab}/\chi_{c}\sim 36 at 1.8 K) is larger than that of CeAuSn because both B20B_{2}^{0} and B43B_{4}^{3} CEF parameters of CeCdX33{}_{3}X_{3} are larger than those of CeAuSn. The CEF investigation of the rhombohedral CeIr3Ge7 compound shows a very similar CEF eigenstates and eigenvalues with those of other compounds. However, in CeIr3Ge7, the reported CEF parameters (especially the term B43=67.3B_{4}^{3}=67.3 K) are larger than that of other compounds, thus inducing a huge energy level splittings of 374 K and 1398 K. [25]. It has been suggested that the exceptionally large CEF splitting can be related to the contribution of the 5d5d ligands of Ir atoms. [25]. In addition, the CEF analysis on rhombohedral CePtAl4Ge2 antiferromagnet has also been conducted [29]. Unlike the above mentioned compounds, the ground state and the second excited state of CePtAl4Ge2 are not in a mixed configuration of |5/2|5/2\rangle and |1/2|1/2\rangle states. The sign and magnitude of B20B_{2}^{0} (= 13.26 K) and B40B_{4}^{0} (= 0.3-0.3 K) CEF parameters in CePtAl4Ge2 are comparable to that of CeAuSn, CeIr3Ge7, and CeCd3X3 (XX = P and As). However, because the B43B_{4}^{3} term responsible for mixing is exceptionally small in CePtAl4Ge2 system (B43=0±0.02B_{4}^{3}=0\pm 0.02 K), the ground state and second excited state is in a pure |±1/2|\pm 1/2\rangle state and a pure |5/2|\mp 5/2\rangle state, respectively. We found that the ground and excited states are without mixing for any |B43|<1.1|B_{4}^{3}|<1.1 K. The small value of B20B_{2}^{0} and B43B_{4}^{3} implies a relatively smaller magnetic anisotropy in CePtAl4Ge2, which is clearly reflected on its magnetic susceptibility data (χab/χc4\chi_{ab}/\chi_{c}\sim 4 near TNT_{N}[29].

Refer to caption
Figure 5: Temperature dependence of the thermoelectric power, S(T)S(T), of RRCd3XX3 (RR = La and Ce, XX = P and As). Vertical arrows indicate the phase transition temperatures observed in electrical resistivity measurements [22, 19].

Although our CEF analysis on CeCd3XX3 (XX = P and As) provides a comprehensive picture at high temperatures, a number of unanswered questions remains at low temperatures. When the temperature is much lower than the CEF splitting, the lowest Kramers doublet is only relevant to explain the observed magnetic ordering at TN=0.42T_{\text{N}}=0.42 K and upturn in CmC_{m} below 10 K. It is obvious that the temperature dependence and absolute value of CmC_{m} below 10 K cannot be explained by the CEF effects as shown in Figs. 3(c) and 4(c).

Since magnetization isotherms at TT = 1.8 K for both compounds are reproduced well by CEF calculation with the ground state wave functions, the reduced magnetization value cannot be associated with the Kondo screening. This is consistent with the electrical resistivity results of CeCd3XX3 [19, 22]. The absence of Kondo effect in CeCd3XX3 is also confirmed from thermoelectric power (TEP) measurements, as shown in Fig. 5. The observed TEP value of LaCd3XX3 is an order of magnitude higher than that of typical metals, implying low carrier density system. The temperature dependence of the TEP, S(T)S(T), of both La- and Ce-based compounds shows a hump at low temperatures, which can be related to the phonon drag. In general, S(T)S(T) of typical metals shows a maximum, corresponding to the phonon-drag effect, where the maximum is expected to be located between ΘD/5\Theta_{D}/5 and ΘD/12\Theta_{D}/12 [37]. Many Ce- and Yb-based Kondo lattice systems have shown that S(T)S(T) indicates an extrema with enhanced value, corresponding to the Kondo effect in conjunction with CEF effect. The TEP of CeCd3XX3 exhibits behavior similar to that of LaCd3XX3, implying negligible Kondo contributions. Hence, as suggested in Refs. [19, 22], the enhancement of the specific heat below 10 K is likely related to either the magnetic frustration in triangular lattices [19, 22, 18, 20, 21] or simply the magnetic fluctuations observed in insulating antiferromagnets [38].

The electrical resistivity of CeCd3XX3 compounds exhibits a metallic behavior, suggesting that Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction may be responsible for the magnetic ordering. However, it would have to be mediated by an extremely small number of charge carriers [19, 22]. It has been qualitatively shown that the magnetic ordering temperature of non-Kondo materials scale with the distance between Ce ions, where the larger the Ce-Ce distance results in a smaller ordering temperature [39]. For example, a non-Kondo metal CeIr3Ge7 orders at an extremely low temperature TN=0.63T_{\text{N}}=0.63 K due to the large Ce-Ce distance (\sim6 Å). When the Ce-Ce distance (\sim4 Å) for CeCd3XX3 is considered, the magnetic ordering temperature is significantly suppressed compared to that of other Ce-based non-Kondo systems. On the contrary, it has been suggested that the superexchange interaction in low carrier density YbAl3C3 compound becomes dominant instead of the RKKY interaction, where the carrier concentration is estimated to be nn\sim0.01 per formula unit (f.u.) [14]. When the carrier concentrations of CeCd3As3 (nn\sim 0.003/f.u.) [19] and CeCd3P3 (nn\sim 0.002/f.u.) [22] compounds are considered, it is reasonable to assume that the superexchange interaction may be responsible for the antiferromagnetic ordering below 0.42 K. In addition, the partial HTH-T phase diagram of these compounds, especially the field-induced increase of TNT_{\text{N}}, is similar to that of 2D insulating triangular lattice systems with easy-plane anisotropy [41, 40]. Furthermore, the low temperature specific heat of CeCd3As3 has been explained by anisotropic exchange Hamiltonian for an insulating, layered triangular lattice [20].

IV Conclusion

At high temperatures, the observed magnetic properties of CeCd3XX3 triangular lattice compounds can be well understood by considering the CEF effects with molecular filed contributions. The large anisotropy in the magnetic susceptibility and magnetization and the electronic Schottky anomaly in the specific heat are explained by energy level splittings of the JJ = 5/2 Hund’s rule ground state of Ce3+ ions into three doublets. The striking similarity of the CEF profile of both compounds implies a very close resemblance of their crystal field environment. Three independent CEF analyses on CeCd3As3 indicate inconsistent CEF profiles, requiring further studies such as inelastic neutron scattering. When the temperature is well below the CEF splitting, the well-isolated Kramers’ doublet ground state is responsible for the antiferromagnetic ordering and the large enhancement of specific heat below 10 K. Further measurements such as magnetization, neutron scattering, and nuclear magnetic resonance are necessary to provide additional insight into the nature of magnetism below TNT_{\text{N}} and the role of anisotropic exchange interactions in the triangular motif.

Acknowledgements.
This work was supported by the Canada Research Chairs, Natural Sciences and Engineering Research Council of Canada, and Canada Foundation for Innovation program. EM was supported by the Korean Ministry of Science and ICT (No. 2021R1A2C2010925) and by BrainLink program funded by the Ministry of Science and ICT (2022H1D3A3A01077468) through the National Research Foundation of Korea.

References

  • [1] W.-J. Hu, S.-S. Gong, W. Zhu, and D. N. Sheng, Phys. Rev. B 92, 140403(R) (2015).
  • [2] Y.-D. Li, X. Wang, and G. Chen, Phys. Rev. B 94, 035107 (2016).
  • [3] Y. Iqbal, W.-J. Hu, R. Thomale, D. Poilblanc, and F. Becca, Phys. Rev. B 93, 144411 (2016).
  • [4] S.-S. Gong, W. Zhu, J.-X. Zhu, D. N. Sheng, and K. Yang, Phys. Rev. B 96, 075116 (2017).
  • [5] Z. Zhu, P. A. Maksimov, S. R. White, and A. L. Chernyshev, Phys. Rev. Lett. 120, 207203 (2018).
  • [6] J. G. Rau and M. J. P. Gingras, Phys. Rev. B 98, 054408 (2018).
  • [7] Y. Li, H. Liao, Z. Zhang, S. Li, F. Jin, L. Ling, L. Zhang, Y. Zou, L. Pi, Z. Yang, J. Wang, Z. Wu, and Q. Zhang, Sci. Rep. 5, 16419 (2015).
  • [8] Y. Li, G. Chen, W. Tong, L. Pi, J. Liu, Z. Yang, X. Wang, and Q. Zhang, Phys. Rev. Lett. 115, 167203 (2015).
  • [9] Y. Shen, Y.-D. Li, H. Wo, Y. Li, S. Shen, B. Pan, Q. Wang, H. Walker, P. Steffens, M. Boehm, Y. Hao, D. L. Quintero-Castro, L. W. Harriger, M. D. Frontzek, L. Hao, S. Meng, Q. Zhang, G. Chen, and J. Zhao, Nature 540, 559 (2016).
  • [10] Y. Li, D. Adroja, P. K. Biswas, P. J. Baker, Q. Zhang, J. Liu, A. A. Tsirlin, P. Gegenwart, and Q. Zhang, Phys. Rev. Lett. 117, 097201 (2016).
  • [11] J. A. Paddison, M. Daum, Z. Dun, G. Ehlers, Y. Liu, M. B. Stone, H. Zhou, and M. Mourigal, Nat. Phys. 13, 117 (2017).
  • [12] M. Baenitz, P. Schlender, J. Sichelschmidt, Y. A. Onykiienko, Z. Zangeneh, K. M. Ranjith, R. Sarkar, L. Hozoi, H. C. Walker, J.-C. Orain, H. Yasuoka, J. van den Brink, H. H. Klauss, D. S. Inosov, and T. Doert, Phys. Rev. B 98, 220409(R) (2018).
  • [13] J. Sichelschmidt, P. Schlender, B. Schmidt, M. Baenitz, and T. Doert, J. Phys.: Condens. Matter 31, 205601 (2019).
  • [14] A. Ochiai, T. Inukai, T. Matsumura, A. Oyamada, and K. Katoh, J. Phys. Soc. Jpn. 76, 123703 (2007).
  • [15] Y. Kato, M. Kosaka, H. Nowatari, Y. Saiga, A. Yamada, T. Kobiyama, S. Katano, K. Ohoyama, H. S. Suzuki, N. Aso, and K. Iwasa, J. Phys. Soc. Jpn. 77, 053701 (2008).
  • [16] A. Ochiai, K. Hara, F. Kikuchi, T. Inukai, E. Matsuoka, H. Onodera, S. Nakamura, T. Nojima, and K. Katoh, J. Phys.: Conf. Ser. 200, 022040 (2010).
  • [17] K. Hara, S. Matsuda, E. Matsuoka, K. Tanigaki, A. Ochiai, S. Nakamura, T. Nojima, and K. Katoh, Phys. Rev. B 85, 144416 (2012).
  • [18] Y. Liu, S. Zhang, J. Lv, S. Su, T. Dong, G. Chen, and N. Wang, arXiv preprint arXiv:1612.03720 (2016).
  • [19] S. R. Dunsiger, J. Lee, J. E. Sonier, and E. D. Mun, Phys. Rev. B 102, 064405 (2020).
  • [20] K. E. Avers, P. A. Maksimov, P. F. S. Rosa, S. M. Thomas, J. D. Thompson, W. P. Halperin, R. Movshovich, and A. L. Chernyshev, Phys. Rev. B 103, L180406 (2021).
  • [21] S. Higuchi, Y. Noshima, N. Shirakawa, M. Tsubota, and J. Kitagawa, Mater. Res. Express 3, 056101 (2016).
  • [22] J. Lee, A. Rabus, N. R. Lee-Hone, D. M. Broun, and E. Mun, Phys. Rev. B 99, 245159 (2019).
  • [23] S. S. Stoyko and A. Mar, Inorg. Chem. 50, 11152 (2011).
  • [24] A. Ochiai, N. Kabeya, K. Maniwa, M. Saito, S. Nakamura, and K. Katoh, Phys. Rev. B 104, 144420 (2021).
  • [25] J. Banda, B. K. Rai, H. Rosner, E. Morosan, C. Geibel, and M. Brando, Phys. Rev. B 98, 195120 (2018).
  • [26] A. Scheie, J. Appl. Crystallogr. 54, 356 (2021).
  • [27] P. C. Canfield and Z. Fisk, Philos. Mag. B 65, 1117 (1992).
  • [28] C. L. Huang, V. Fritsch, B. Pilawa, C. C. Yang, M. Merz, and H. v. Löhneysen, Phys. Rev. B 91, 144413 (2015).
  • [29] S. Shin, V. Pomjakushin, L. Keller, P. F. S. Rosa, U. Stuhr, C. Niedermayer, R. Sibille, S. Toth, J. Kim, H. Jang, S.-K. Son, H.-O. Lee, T. Shang, M. Medarde, E. D. Bauer, M. Kenzelmann, and T. Park, Phys. Rev. B 101, 224421 (2020).
  • [30] K. W. H. Stevens, Proc. Phys. Soc. A 65, 209 (1952).
  • [31] M. T. Hutchings, Solid State Phys. 16, 227 (1964).
  • [32] C. Klingner, C. Krellner, M. Brando, C. Geibel, and F. Steglich1, New J. Phys. 13 083024 (2011).
  • [33] Y.-L. Wang, Physics Letters A 35, 383 (1971).
  • [34] G. J. Bowden, D. S. P. Bunbury, and M. A. H. McCausland, J. Phys. C: Solid State Phys. 4 1840 (1971).
  • [35] N. Shohata, J. Phys. Soc. Jpn. 42, 1873 (1977).
  • [36] D. T. Adroja, B. D. Rainford, and A. J. Neville, J. of Phys.: Condens. Matter 9, L391 (1997).
  • [37] F. J. Blatt, P. A. Schroeder, C. L. Foiles, and D. Greig, Springer US, Boston, MA (1976).
  • [38] W. K. Robinson and S. A. Friedberg, Phys. Rev. 117, 402 (1960).
  • [39] Binod K. Rai, Jacintha Banda, Macy Stavinoha, R. Borth, D.-J. Jang, Katherine A. Benavides, D. A. Sokolov, Julia Y. Chan, M. Nicklas, Manuel Brando, C.-L. Huang, and E. Morosan, Phys. Rev. B 98, 195119 (2018).
  • [40] L. Seabra, T. Momoi, P. Sindzingre, and N. Shannon, Phys. Rev. B 84, 214418 (2011).
  • [41] D. H. Lee, J. D. Joannopoulos, J. W. Negele, and D. P. Landau, Phys. Rev. B 33, 450 (1986).