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

Complexity in the hybridization physics revealed by depth-resolved photoemission spectroscopy of single crystalline novel Kondo lattice systems, CeCuX2 (X = As/Sb)

Sawani Datta,1 Ram Prakash Pandeya,1 Arka Bikash Dey,2 A. Gloskovskii,2 C. Schlueter,2 T. R. F. Peixoto,2 Ankita Singh,1 A. Thamizhavel,1 and Kalobaran Maiti Corresponding author: [email protected] Department of Condensed Matter Physics and Materials Science, Tata Institute of Fundamental Research, Homi Bhabha Road, Colaba, Mumbai-400005, India.
2Deutsches Elektronen-Synchrotron DESY, 22607 Hamburg, Germany.
Abstract

We investigate the electronic structure of a novel Kondo lattice system CeCuX2 (X = As/Sb) employing high resolution depth-resolved photoemission spectroscopy of high quality single crystalline materials. CeCuSb2 and CeCuAs2 represent different regimes of the Doniach phase diagram exhibiting Kondo-like transport properties and CeCuSb2 is antiferromagnetic (TNT_{N}\sim 6.9 K) while CeCuAs2 does not show long-range magnetic order down to the lowest temperature studied. In this study, samples were cleaved in ultrahigh vacuum before the photoemission measurements and the spectra at different surface sensitivity establish the pnictogen layer having squarenet structure as the terminated surface which is weakly bound to the other layers. Cu 2pp and As 2pp spectra show spin-orbit split sharp peaks along with features due to plasmon excitations. Ce 3dd spectra exhibit multiple features due to the hybridization of the Ce 4ff/5dd states with the valence states. While overall lineshape of the bulk spectral functions look similar in both the cases, the surface spectra are very different; the surface-bulk difference is significantly weaker in CeCuAs2 compared to that observed in CeCuSb2. A distinct low binding energy peak is observed in the Ce 3dd spectra akin to the scenario observed in cuprates and manganites due to the Zhang-Rice singlets and/or high degree of itineracy of the conduction holes. The valence band spectra of CeCuSb2 manifest highly metallic phase. In CeCuAs2, intensity at the Fermi level is significantly small suggesting a pseudogap-type behavior. These results bring out an interesting scenario emphasizing the importance and subtlety of hybridization physics underlying the exoticity of this novel Kondo system.

I Introduction

Rare earth-based compounds exhibit several exotic properties due to the local character of the rare earth 4ff states and their hybridization with the conduction electrons [1, 2, 3]. In Ce-based materials, the occupied part of the 4ff band appears close to the Fermi level, ϵF\epsilon_{F} leading to strong hybridization of these states with the conduction electronic states and thereby these 4ff electrons gain significant itineracy in addition to their local character. Such dual properties of the Ce 4ff electrons leads to plethora of exotic properties and are among the most studied systems in this class [1, 2, 3, 4, 5, 6, 7, 8, 9]. Moreover, Ce 4ff orbitals have larger radial extension than the heavier rare-earth 4ff states and hence, Ce 4ff electrons are relatively more exposed to the crystal field that gives rise to additional complexity in the physics of these materials [10, 11, 9]. It has been found that the properties of the conduction electrons (correlation induced effect, spin-orbit coupling, etc.) also play key role in the exoticity of these materials [12]. For example, Ce2CoSi3 is a kondo material and does not show long-range magnetic order till the lowest temperature studied[13], while Ce2RhSi3 belonging to the same class show antiferromagnetic (AFM) ground state[14, 15] and considered to be an example of quantum critical behavior with spin density wave ground state [4].

Refer to caption
Figure 1: As 3dd, Ce 5ss and Ce 5dd survey spectra of CeCuAs2 collected using 2.5 keV (solid circles) and 6 keV (open circles) photon energies. Inset shows the crystal structure CeCuX2; X = pnictogen, As/Sb.

Here, we study the electronic structure of single crystalline CeCuX2 (X = As/Sb) employing high resolution depth-resolved photo-electron spectroscopy (DRPES) using hard xx-rays, where the substitution is done at the pnictogen sites. These materials form in ZrCuSi2-type layered tetragonal structure (space group P4/nmmP4/nmm) as shown in the inset of Fig. 1 [5, 8]. The pnictogen atoms X (X = As, Sb) have two different Wyckoff positions; X1 atoms form a squarenet structure and weakly coupled to Ce layers. X2 is strongly hybridized with Cu atoms along with reasonably strong Ce-X2 bonds. Larger atomic radius of Sb atoms leads to a slightly larger lattice parameters; aa = 4.018 and 4.337 Å, and cc = 10.104 and 10.233 Å for X = As and Sb, respectively. The Ce-Cu, Ce-X1 and Ce-X2 bondlengths in CeCuSb2 (= 3.3615, 3.3459 and 3.2297 Å) are larger than those in CeCuAs2 (= 3.2992, 3.1569 and 2.9279 Å). However, Cu-As2 (= 2.7718 Å) bond is weaker than Cu-Sb2 (= 2.6686 Å) bond. Since, X2-Cu-X2 layers are the conduction layers in these materials, such change in Cu-X2 bondlength has significant implication in the properties of the conduction electrons. Shorter Cu-Sb2 bondlength leads to a higher degree of metallicity of CeCuSb2 as also observed in the bulk properties study of these samples. The conduction electrons in As2-Cu-As2 layers are expected to be relatively less itinerant and are more strongly coupled to the Ce 4ff states than those in CeCuSb2.

The bulk properties of this class of materials have been studied extensively and most of the candidates show long range magnetic order [6, 7]. Among the two systems studied here, CeCuSb2 exhibits AFM ground state with Neél temperature, 6.9 K [8], whereas CeCuAs2 does not show long-range magnetic order down to the lowest temperature studied [5, 16]. Resistivity (ρ\rho) of both the materials increases logarithmically with cooling from room temperature similar to a Kondo system. A maxima is observed around 23 K (for J |||| [100]) in CeCuSb2 and subsequently, ρ\rho gradually decreases at lower temperatures [8]. In CeCuAs2, the resistivity maxima is observed at 20 K and the slope of the curve decreases at lower temperatures; such unusual behavior at low temperatures is not well understood yet [5]. The in-plane resistivity of CeCuAs2 is much higher than that for CeCuSb2, which suggests that CeCuSb2 is a better electronic conductor. The AFM ground state along with Kondo behavior in CeCuSb2 places this material in the spin density wave (SDW) type quantum critical regime such as Ce2RhSi3, where Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction mediates the magnetic order. In CeCuAs2, the Kondo behavior dominates giving rise to a paramagnetic ground state as observed in Ce2CoSi3. The uniqueness of these materials is that the tuning of the properties of conduction electrons occurs via substitution at the pnictogen site in contrast to the earlier cases where the transition metal site was used for substitution. Moreover, the pnictogen layer forming the squarenet structure hosts Dirac fermions and is in proximity to the Ce-layer while the pnictogen layer on the other side is strongly bonded to Ce and Cu layers. Thus, these materials provide a novel playground to study the complex physics involving behavior of Ce 4ff electron in proximity to topological states with varying Kondo coupling strength. We observe interesting features in the DRPES results revealing puzzling scenario in this system.

II Experiment

High-quality single crystals of CeCuSb2 and CeCuAs2 were prepared by flux method using Sb and As-Sn flux, respectively. Elemental analysis of the prepared samples was done by the energy dispersive analysis of xx-rays. Good crystallinity of the samples was confirmed by the Laue diffraction measurements. The hard xx-ray photoemission spectroscopy (HAXPES) measurements were performed at the P22 beamline of Petra III, DESY, Hamburg, Germany. We have used a high-resolution Phoibos electron analyzer; the energy resolution at 6 keV photon energy is close to 150 meV. The resolution becomes close to 200 meV at other photon energies. All the measurements were carried at a pressure lower than 1010\sim 10^{-10} Torr, immediately after cleaving the samples in ultrahigh vacuum employing a top-post removal method. The sample temperature was varied using an open cycle helium cryostat. To tune the surface sensitivity of the measurement technique, we have used three well-separated incident photon energies 2.5 keV, 6 keV and 8 keV.

In order to capture the properties of various features in the valence band, we calculated the electronic band structure were using full potential linearized augmented plane wave method (FLAPW) as implemented in the Wien2k software [17]. The Perdew-Burke-Ernzerhof generalized gradient approximation (GGA) [18] was used for the calculation of the density functionals [19]. We considered electron correlation strength, UU = 7.5 eV for 4ff electrons and spin-orbit coupling for the calculations. Lattice parameters were fixed at the experimentally derived values for our calculations [8, 5].

III Results and discussion

In Fig. 1, we show the survey scans consisting of As 3dd, Ce 5ss and Ce 5dd core level spectra of CeCuAs2 collected using 2.5 keV and 6 keV photon energies. The escape depths of these photoelectrons at 6 keV and 2.5 keV photon energies are close to 40 Å and 26 Å, respectively [20]. Hence, the surface sensitivity of the experiment will be significantly larger at 2.5 keV photon energy compared to the 6 keV case. In the figure, we normalized the spectral intensities by the intensity of the As 3dd peak height. This shows that the intensities of Ce core level peaks are significantly larger in the bulk sensitive 6 keV data. The intensity of these core level peaks depends essentially on (i) the photoemission cross-section and (ii) the number of electrons at the corresponding energy level. The cross-section ratio of As 3dd and Ce 5dd at 2.5 keV and 6 keV are very similar; I(As3d)/I(Ce5p)I(As3d)/I(Ce5p) = 2.78 (2.5 keV) and 2.66 (6 keV) using atomic photoemission cross-sections [21]. The occupancy of the energy levels in the ground state is the same in both the cases. Thus, the huge change in intensity observed in the experimental data in Fig. 1 can be attributed to the surface sensitivity of the technique and As layer appears to be the terminated surface of the sample.

From the crystal structure shown in the inset of Fig. 1, it is clear that As1 layers forming the squarenet structure are relatively far away from the Ce-As2-Cu-As2-Ce quintuple layers and are weakly bonded to the Ce-layers on both sides of the As1-plane. The covalent bonding of As2-layer with Ce and Cu layers is strong. From these structural parameters and the inter-layer bonding, the cleaving of the sample is expected to occur at the As1-layer leading to As1-layer or Ce-layer as the terminated surface. The experimental data shown in the figure suggests that the terminated layer in CeCuAs2 is the As1 layer for the cleaved sample used for our experiments. A similar scenario is observed in CeCuSb2 [22].

Refer to caption
Figure 2: As 2pp (xx-scale at the top axis) and Cu 2pp (xx-scale at the bottom axis). Signature of plasmon feature is shown in an enlarged intensity scale. As 2p3/2p_{3/2} peak at 2.5 keV (open circles), 6 keV (red line) and 8 keV (blue line) photon energy overlap on each other well. Line superimposed over the Cu 2pp spectrum is the fit curve; the component peaks are shown in the bottom panel.

As 2p3/2p_{3/2} spectra collected at 2.5 keV, 6 keV and 8 keV are shown in Fig. 2 exhibiting a sharp feature at 1323 eV binding energy in all the cases along with an asymmetry towards higher binding energies. This type of asymmetry in the spectral lineshape may arise due to the low energy excitations across the Fermi level along with the core level photoemission. The spectral functions around the main peak appear similar at all the photon energies probed. This suggests that the influence of surface-bulk difference, if there is any, on the deep core level, As 2pp photoemission is within the experimental limits. This is not unusual as these deep core levels are less exposed to the crystal field and the behavior is essentially like atomic levels. The scenario is similar for 2p1/2p_{1/2} region and not shown here for clarity of presentation. In addition to the main peak, we observe two broad humps at about 1339 eV and 1349 eV binding energies; the features are shown in an enlarged intensity scale in the same figure for clarity. These features are attributed to the plasmon excitations along with the photo-excitations of the As 2pp core electrons. The intensity of the 1349 eV peak is larger in the 2.5 keV data compared to the intensity of the other one suggesting it’s link to the surface electronic structure.

In the middle panel of Fig.2, we show the Cu 2pp core-level spectrum collected using 6 keV photon energy. Both the spin-orbit split features of Cu 2pp photoemission signal exhibit an asymmetric single peak structure as observed in the As 2pp case. The experimental data can be fitted using an asymmetric Gausian-Lorentzian function. The simulated envelope (red line) is superimposed over the experimental curve; the individual components are shown in the lower panel of Fig. 2. The sharp intense peaks of the spin-orbit split Cu 2pp features and the absence of satellites suggests close to monovalency of Cu with 3d10d^{10} electronic configuration [23]. In addition, there is a broad hump at about 16 eV away from the main spin-orbit split peaks as also observed in the As 2pp spectra which substantiates the attribution of this peak to plasmon excitations [24].

Refer to caption
Figure 3: (a) Ce 3dd spectra of CeCuAs2 collected at 45 K (open circles) and 120 K (solid line) using 6 keV (upper panel) and 2.5 keV (lower pannel) photon energies. The spin-orbit split 3d5/2d_{5/2} and 3d3/2d_{3/2} features are marked by the 1,2, … etc. and 1’, 2’, … etc. respectively. The extracted (b) bulk and (c) surface spectra (open circle). Lines superimposed over the experimental data are the fit curves. The component peaks are shown in the bottom panel. The plasmon feature is shown only in the bulk spectrum for clarity; the surface spectrum also has similar feature with relatively weaker intensity.

In Fig. 3 we study the Ce 3dd spectra obtained at varied experimental conditions. The experimental data are taken using 6 keV and 2.5 keV photon energies; the data at 45 K and 120 K sample temperatures are superimposed over each other in Fig. 3(a) exhibiting identical lineshape. There are several features in the spectral regions corresponding to spin-orbit split 3d5/2d_{5/2} and 3d3/2d_{3/2} photoemissions. The signature of distinct peaks are marked by 1, 2, … for 3d5/2d_{5/2} signal and those in 3d3/2d_{3/2} signal are marked by the primed numbers, 1, 2, …. Relative intensity of the features at different photon energies are somewhat different. For example, the intensity of the peak, 4 in the 6 keV data is enhanced in the 2.5 keV data relative to the intensity of the peak, 3, while the intensity of 5 is reduced. Relative intensities of the features, 1 and 2 and their primed counterparts also show subtle changes in intensity with the change in photon energy. Since the surface sensitivity of the technique enhances at 2.5 keV photon energy relative to that at 6 keV, the spectral changes observed at the two photon energies suggest different surface and bulk electronic structures.

The spectral functions corresponding to the surface and bulk electronic structures are extracted using the surface sensitivity of the technique following the well-established procedures[25]. Considering that the different surface electronic structure appears from the top few layers of thickness, dd and electron escape depth, λ\lambda, the spectral intensity, I(ϵ)I(\epsilon) can be expressed as, I(ϵ)=0dIs(ϵ)ex/λ𝑑x+dIb(ϵ)ex/λ𝑑xI(\epsilon)=\int_{0}^{d}I^{s}(\epsilon)e^{-x/\lambda}dx+\int_{d}^{\infty}I^{b}(\epsilon)e^{-x/\lambda}dx, where Is(ϵ)I^{s}(\epsilon) and Ib(ϵ)I^{b}(\epsilon) are the surface and bulk spectra. From the above equation the spectral intensity can be derived as, I(ϵ)=Is(ϵ)(1ed/λ)+Ib(ϵ)ed/λI(\epsilon)=I^{s}(\epsilon)(1-e^{-d/\lambda})+I^{b}(\epsilon)e^{-d/\lambda}. Since the core levels are fully occupied, the area under the spectra representing the spectral intensity will be same at all photon energies. Thus, we normalized the experimental spectra at 2.5 keV and 6 keV by the integrated area under the curve. Considering that the universal curve is applicable to scale the escape depth[20] and the spectral intensity is positive at all energies, we extracted the surface and bulk spectra functions using the above equation, and the spectra at 2.5 keV and 6 keV data.

The extracted bulk and surface 3dd spectra are shown in Fig.3(b) and (c), respectively; we have not shown the higher binding energy regime for better clarity as there are no distinct features in that region of the experimental data. To find out the peaks associated with different final states of the Ce 3dd photoemission, we have simulated both the spectra considering a Voigt function for each of the distinct feature in the experimental data as marked. The branching ratio (intensity ratio of the spin-orbit split features) are kept fixed to their degeneracy. The simulated data, shown by the red lines, are superimposed over the experimental extracted data in the figure and the individual peaks are shown in the lower panel. Peaks in the surface and bulk spectra are marked by s’s and b’s, respectively. Simulated data show a good representation of the experimental spectra.

In a correlated system, the eigenstates of the final state Hamiltonian will be different from those of the initial state Hamiltonian due to the presence of the photohole interacting with the other electrons. This allows transition to many final states leading to multiple features in the photoemission signal. The scenario can be represented by the configuration interaction model where each electronic configuration corresponds to a particular level of core-hole screening due to the hopping of electrons to the photoemission sites in such systems. The scenario is well captured by the Gunnarsson and Schonhammer’s (GS) approach using the single impurity Anderson model (SIAM) for the calculation of Ce 3dd final state [26]. According to the GS calculation, three final states configurations provide major contributions in the Ce 3dd core level spectrum: (a) f2f_{2} feature is the wellscreenedwell-screened final state, where the 3dd core hole is screened via hopping of a neighbor electron to the Ce 4ff level [27, 28]. The presence of this feature depends on the hybridization of Ce 4ff level with the valence states allowing transfer of electron. (b) f1f_{1} peak corresponds to the poorlyscreenedpoorly-screened state. Here, the final states possesses 4f1f^{1} electronic configuration and the core hole is not screened. (c) f0f_{0} feature relates to the final state where the Ce 4ff electron has hopped to the valence band leading to Ce4+ valency; this feature is often linked to the Kondo singlet which is a quantum entangled state of the Ce 4ff electron with the conduction electrons [29]. We did not observe signature of this feature in the present system. Each configuration consists of several multiplets; distinct signature of the multiplets depends on the eigenenergy separation of the multiplets.

In the Ce 3dd bulk spectra presented in Fig.3(b), the Ce 3d5/2d_{5/2} peaks are denoted by b’s and the spin-orbit split Ce 3d3/2d_{3/2} spectrum consists of peaks b’s. The spin-orbit splitting is found to be around 18.8 eV. From the binding energy considerations, we ascribe the b2 and b2{}_{2}^{\prime} peaks to f2f_{2} final state. The features, b3, b4 and b5 (b3{}_{3}^{\prime} b4{}_{4}^{\prime} and b5{}_{5}^{\prime} in the 3d3/2d_{3/2} region) peaks to f1f_{1} final state. The three-peak structure of the f1f_{1} final states is observable due to the multiplet splitting. This description is consistent with the calculated results within the Anderson impurity model [23]. We also observe a broad hump in the higher binding energy side of the main features representing the plasmon-induced loss feature.

The analysis of the surface spectrum is shown in Fig.3(c). Here, Ce 3d5/2d_{5/2} and Ce 3d3/2d_{3/2} peaks are denoted by s’s and s’s. Plasmon feature is weaker in the surface spectrum than that in the bulk. Similar to the bulk case, s2 and s2{}_{2}^{\prime} represent the well-screened f2f_{2} features. There are three multiplet features in the f1f_{1} signal denoted by s3, s4, s5 for 3d5/2d_{5/2} signal and s3{}_{3}^{\prime}, s4{}_{4}^{\prime}, s5{}_{5}^{\prime} for 3d3/2d_{3/2} signal. The intensity of the f2f_{2} peak relative to the f1f_{1} intensity is weaker in the surface spectrum compared to that in the bulk. The binding energy of the surface peak is reduced by about 0.3 eV from the corresponding bulk peaks. These observations suggest that Ce is slightly less positive at the surface relative to the bulk Ce-valency. This is consistent with the fact that the bulk Ce states are expected to have stronger hybridization with the valence states leading to larger extended character while the surface states will be more local[30]. This may also be a reason for the higher f2f_{2} intensity in the bulk spectra.

Refer to caption
Figure 4: Simulated Ce 3d5/2d_{5/2} spectra of CeCuAs2 using Imer and Wuilloud’s approach of GS calculation for (a) the bulk and (b) the surface (b) spectrum. Line superimposed over the experimental data is the simulated curve. The component peaks for the f1f_{1} and f2f_{2} final states are shown in the bottom panel.

In order to verify the above assertions with the existing models for such systems, we have simulated the Ce 3d5/2d_{5/2} spectrum using Imer and Wuilloud’s approach [31], which is a simplified version of the Gunnarsson and Schönhammer model and considers only the configurations without their multiplet splittings. The calculated plots are shown in Fig. 4(a) (bulk) and (b) (surface) by the red line superimposing over the experimental data (black open circle). Here, the experimental spectra are obtained by subtracting the contributions due to b1 and s1 signals. The f2f_{2} and f1f_{1} final states are shown by the green and blue lines, respectively. The simulated results show an excellent representation of the experimental spectra for the following set of parameters. For bulk: on-site Coulomb repulsion strength, UffU_{ff} = 8 eV, the core-hole potential UfcU_{fc} = 8.5 eV, energy of the unhybridized Ce 4ff level efe_{f} = -1.5 eV and the hybridization between the final states f0f_{0}, f1f_{1} and f2 states, denoted by Δ\Delta which is equal to 0.55 eV. Using this calculation we got the value for the ff occupancy of 0.92 for the bulk spectrum. The simulation of the surface spectrum requires slightly different efe_{f} (= -1.7 eV) keeping all the other parameters same. The ff-occupancy for surface Ce is found to be 0.94.

We note here that the Ce 3dd spectra analysed in Fig. 4 is similar to the reported Ce 3dd spectral lineshape for polycrystalline CeCuAs2[23]. It appears clear that the features, b1 and s1 have a different origin and they are revealed in the experimental data from single crystalline samples. Based on the discussion above and published literature, we assign these features to the well screened final states where the holes created at the ligand sites and/or conduction band are further stabilized via formation of singlets with a Cu 3dd hole and/or delocalization as observed in other materials [22, 32]. Similar effect is reported in Cu, Fe and Mn-based systems too [33, 34, 35, 36].

Refer to caption
Figure 5: (a) Cu 2pp spectra of CeCuAs2 ((open circles) and CeCuSb2 (solid line) collected using 6 keV photon energy. Inset shows the boxed part of the spectra. (b) Comparison of the Ce 3dd surface, bulk spectra of CeCuAs2(open circles) and CeCuSb2(solid line).

In order to study the underlying scenario further, we compare the results of CeCuAs2 with those of CeCuSb2 in Fig. 5. Cu 2pp spectra of CeCuAs2 (open circles) and CeCuSb2 (solid line) are shown in Fig. 5(a); in the case of CeCuSb2, Sb 3ss contributions are subtracted for one to one comparison. The lineshape of the main peak appear similar in both the cases except the fact that the spectral asymmetry is slightly larger in CeCuAs2. This is shown with clarity in the inset of Fig. 5(a). The plasmon features are more intense in CeCuSb2 presumably due to the highly conductive nature of CeCuSb2 among these two materials.

The surface and bulk Ce 3dd spectra are compared in Fig. 5(b). The surface spectrum of CeCuSb2 look very different from that of CeCuAs2. The electronic structure of Sb1 at the terminated surface are significantly different from the other Sb-layers [22]. Such difference for the pnictogen layers is less evident in CeCuAs2. This might be a reason for such a different Ce 3dd surface spectra. The bulk Ce 3dd spectra, however, appear similar in both the cases. The lowest binding energy feature, s1 is almost absent in CeCuSb2 surface spectra while that in CeCuAs2 is quite significant. In the bulk spectra too the intensity of the feature, b1 in CeCuAs2 is stronger than that in CeCuSb2.

To investigate this further, we compare various bond lengths in these two compounds which is directly linked to the hybridization parameters. The Ce-Cu bondlength is 3.2992 Å in CeCuAs2 and 3.3615 Å in CeCuSb2. Shorter bondlength in CeCuAs2 suggests stronger Ce-Cu hybridization compared to that in CeCuSb2. On the other hand, the Cu-X2 bond length is smaller in CeCuSb2 (2.6686 Å) than in CeCuAs2 (2.7718 Å). Hence, the X2-Cu-X2 layer is more tightly packed in CeCuSb2. Thus, the holes created in the conduction band due to Ce 3dd core-hole screening will have higher probability to propagate to the Cu-site in CeCuAs2 than in CeCuSb2 and form a singlet with the Cu 3dd-hole. While these considerations provide a qualitative picture of the scenario, we hope future theoretical studies would help to establish the properties of this feature better.

Refer to caption
Figure 6: Calculated partial density of states (PDOS) for (a) CeCuSb2 and (b) CeCuAS2. Insets show the Ce 4ff and Cu 3dd PDOS in an expanded PDOS scale.

In Fig. 6, we have shown the calculated density of states (DOS) of CeCuSb2 (Fig. 6(a)) and CeCuAs2 (Fig. 6(b)). In CeCuSb2, the major contribution in the bonding energy bands (-2.6 eV to -5.6 eV) are coming from the Cu 3dd and Sb2 5pp partial density of states (PDOS). Sb1 5pp contributions appear at higher energies compared to Sb2 5pp PDOS. The energy region between -3 eV and the Fermi level are contributed by Sb 5pp (both Sb layers), Ce 5dd and Cu 3dd states. Evidently, the hybridization of Sb2 5pp states with Ce 5dd and Cu 3dd PDOS is significant. The Ce 4ff PDOS primarily appear in the vicinity of the Fermi level, ϵF\epsilon_{F}. Contribution of Ce 4ff is quite low in the lower energy region, where Cu 3dd PDOS dominate; this is shown with clarity in the inset of Fig. 6(a). Large intensity in the vicinity of ϵF\epsilon_{F} suggests highly metallic character of the material.

On the other hand, Cu 3dd contributions appear much closer to the Fermi level in CeCuAs2. PDOS at about -1 eV appears to be negligible indicating signature of an energy gap there. Ce 5dd contributions are shifted towards lower energy leading to a larger hybridization with the Cu 3dd states. Ce 4ff states also contribute significantly between -1 to -2 eV suggesting relatively larger Ce - Cu hybridization as also predicted from the bondlength analysis. The difference between As1 and As2 5pp states are significantly reduced in CeCuAs2 compared to CeCuSb2. These results supports the observations in the core level spectra such as slightly larger asymmetry of Cu 2pp feature in CeCuAs2 arising from larger Cu 3dd PDOS near the Fermi level allowing larger low energy excitations. This also provides signature of larger Ce 4ff - Cu 3dd hybridization which might be a reason for larger intensity of b1 in CeCuAs2.

Refer to caption
Figure 7: Valence band spectra of CeCuAs2 and CeCuSb2 at 6 keV and 45 K temperature. Different spectral regions based on band structure results are shown by shaded areas. The data show negligible intensity at the Fermi level in CeCuAs2 while it is quite intense in CeCuSb2.

In Fig. 7 we show the valence band spectra of CeCuSb2 and CeCuAs2 measured using 6 keV photon energy at 45 k sample temperature. Comparison of these data with the calculated results shown in Fig. 6 suggests that the binding energy region between 1.5 eV and 6 eV is primarily contributed by As 4pp / Sb 5pp states. Cu 3dd contributions appear in the vicinity of 3 eV binding energy. The peak at about 5.5 eV is contributed by As2 4pp states while the Sb2 5pp contributions appear at slightly lower binding energy as also observed in the calculated data. Cu 3dd contributions appear to be much broader in CeCuAs2 compared to CeCuSb2 consistent with their PDOS. The energy region close to ϵF\epsilon_{F} is primarily contributed by Ce 5dd and 4ff PDOS with which are strongly hybridized with the As/Sb pp states. Interestingly, the spectral intensity at ϵF\epsilon_{F} is quite significant in CeCuSb2 suggesting highly metallic character of this material while the intensity in CeCuAs2 is negligibly small. This is consistent with the observation of pseudogap in earlier studies [23].

Refer to caption
Figure 8: (a) Valence band spectra of CeCuAs2 at 45 K using 2.5 keV (triangles), 6 keV (open circles) and 8 keV (solid circles) photon energies. (b) 6 keV valence band spectra at temperatures 45 K, 70 K and 230 K. (c) As 2pp spectra at 30 K and 250 K using 6 keV photon energy. (d) Cu 2pp and Ce 3dd spectra at 30 K and 250 K using 6 keV photon energy. The data at different temperatures show almost identical lineshape in all the cases.

In Fig. 8(a), we show the experimental valence band spectra collected at 45 K at different photon energies. The spectra collected using 6 keV and 8 keV photon energies overlap well except the intensity of the feature around 5.5 eV which is stronger in the 8 keV data. In the 2.5 keV data, the intensity of the feature around 4 eV enhances drastically compared to the intensity of the other features. The feature at 5.5 eV becomes much smaller in this spectrum. From the atomic cross-section data [21], we observe that the cross section of Cu 3dd at 1486.6 eV photon energy is about an order of magnitude higher than the cross section of As 4pp, Ce 4ff and 5dd states. The cross section reduces by 2 orders of magnitude at 8 keV photon energy in every case except Ce 4ff which reduces by 3 orders of magnitude. Thus, the change in intensity between 6 keV and 8 keV can be attributed essentially to the change in surface sensitivity and the feature at 5.5 eV constituted by As2 4pp PDOS is a bulk feature. The feature at 4 eV is primarily contributed by the Cu 3dd states. The intensity at ϵF\epsilon_{F} is weak in every case indicating a semi-metallic behavior of this material.

We now investigate the evolution of the spectral functions with temperature. The spectra collected using 6 keV photon energy at different temperatures are shown in Fig. 8(b), (c) and (d) for valence band, As 2pp, Cu 2pp and Ce 3dd photoemissions. We do not observe any distinct change in the spectral lineshape with temperature in any of the cases. This suggests that the thermal effect on the electronic structure of this system in the temperature range studied is not significant enough to be detected by these experiments.

IV Conclusion

In conclusion, we have investigated the electronic structure of ternary CeCu-pnictides (CeCuX2; X= Sb and As) employing depth-resolved hard xx-ray photoemission spectroscopy. We observe that these materials cleaved at the pnictogen layers leaving the terminated surface dominated by the pnictogen atoms forming a square-net structure. This has significant implication in the study and application of these materials as the square-net structure is known to host topologically protected states. In the present scenario, such states can be produced at the surface as well as at the X1 layer within the bulk of the material. CeCuAs2 is an unique Kondo material which does not show magnetic order down the lowest temperature studied while other materials in this class show magnetic ground state. Ce 3dd core level spectra do not show the signature of f0f_{0} feature in both the surface and bulk cases. We observe several features in the Ce 3dd spectra suggesting strong hybridization between the Ce 4ff/5dd states with the valence states. The feature at lowest binding energy of the Ce 3dd spectra is found to be strong; the intensity is relatively larger in CeCuAs2 which is an indication of stronger Cu-Ce hybridization in CeCuAs2. The valence band spectra is found to be consistent with the calculated results using density functional theory. The intensity close to the Fermi level show highly metallic ground state of CeCuSb2 while CeCuAs2 is semi-metallic. The spectral functions collected at different temperature show identical lineshape indicating that the temperature induced changes are subtle in these materials. Hybridization between various valence states are found to be complex and leads to significant change in the spectral functions, which might be responsible for the exoticity of this novel Kondo system.

V Acknowledgements

Authors acknowledge the financial support under India-DESY program and Department of Atomic Energy (DAE), Govt. of India (Project Identification no. RTI4003, DAE OM no. 1303/2/2019/R&D-II/DAE/2079 dated 11.02.2020). K. M. acknowledges financial support from BRNS, DAE, Govt. of India under the DAE-SRC-OI Award (grant no. 21/08/2015-BRNS/10977).

References

  • [1] G. R. Stewart, Rev. Mod. Phys. 56, 755 (1984); N. B. Brandt and V. V. Moshchalkov, Adv. Phys. 33, 373 (1984); F. Steglich, J. Magn. Magn. Mater 100, 186 (1991).
  • [2] J. C. Parlebas, Phys. Stat. Sol. (b) 160 11 (1990).
  • [3] A. C. Hewson, The Kondo Problem to Heavy Fermions (Cambridge University Press, Cambridge, 1993).
  • [4] S. Patil, K. K. Iyer, K. Maiti and E. V. Sampathkumaran, Phys. Rev. B 77, 094443 (2008). S. Patil, S. K. Pandey, V. R. R. Medicherla, R. S. Singh, R. Bindu, E. V. Sampathkumaran, and K. Maiti, J. Phys.: Condens. Matter 22, 255602 (2010); S. Patil, G. Adhikary, G. Balakrishnan, and K. Maiti, J. Phys.: Condensed Matter 23, 495601 (2011).
  • [5] K. Sengupta, E. V. Sampathkumaran, T. Nakano, M. Hedo, M. Abliz, N. Fujiwara, Y. Uwatoko, S. Rayaprol, K. Shigetoh, T. Takabatake, Th. Doert, and J. P. F. Jemetio, Phys. Rev. B 70, 064406 (2004).
  • [6] H. Sprenger, J. Less-Common Met. 34, 39 (1974); J. Stepien-Damm, D. Kaczorowski, and R. Troc, J. Less-Common Met. 132, 15 (1987); M. Brylak, M. H. Moller, and W. Jeitschko, J. Solid State Chem. 115, 305 (1995).
  • [7] D. Rutzinger, C. Bartsch, M. Doerr, H. Rosner, V. Neu, Th. Doert, and M. Ruck, J. Solid State Chem. 183, 510 (2010); M. Szlawska and D. Kaczorowski, J. Alloys Compd. 451, 464 (2008); E. V. Sampathkumaran, K. Sengupta, S. Rayaprol, K. K. Iyer, Th. Doert and J. P. F. Jemetio, Phys. Rev. Lett. 91, 036603 (2003).
  • [8] S. Araki, N. Metoki, A. Galatanu, E. Yamamoto, A. Thamizhavel, and Y. Onuki, Phys. Rev. B 68, 024408 (2003); A. Thamizhavel, T. Takeuchi, T. Okubo, M. Yamada, R. Asai, S. Kirita, A. Galatanu, E. Yamamoto, T. Ebihara, Y. Inada, R. Settai, and Y. Onuki, ibid. 68, 054427 (2003).
  • [9] S. Patil, V. R. R. Medicherla, R. S. Singh, S. K. Pandey, E. V. Sampathkumaran, and K. Maiti, Phys. Rev. B 82, 104428 (2010); K. Maiti, S. Patil, G. Adhikary, and G. Balakrishnan, J. Phys.: Conf. Ser. 273, 012042(2011); S. Patil, A. Generalov, M. Güttler, P. Kushwaha, A. Chikina, K. Kummer, T. C. Rödel, A. F. Santander-Syro, N. Caroca-Canales, C. Geibel, S. Danzenbächer, Yu. Kucherenko, C. Laubschat, J. W. Allen, and D. V. Vyalikh , Nat. Commun. 7, 11029 (2016).
  • [10] E. Weschke, C. Laubschat, R. Ecker, A. Höhr, M. Domke, G. Kaindl, L. Severin, and B. Johansson, Phys. Rev. Lett. 69, 1792 (1992).
  • [11] J. W. Allen, S. J. Oh, O. Gunnarsson, K. Schönhammer, M. B. Maple, M. S. Torikachvili, and I. Lindau, Adv. Phys. 35, 275 (1986).
  • [12] S. Patil, V. R. R. Medicherla, R. S. Singh, E. V. Sampathkumaran, and K. Maiti, EPL 97, 17004 (2012).
  • [13] R. A. Gordon, M. G. Alexander, C. J. Warren, F. J. DiSalvo and R. Pöttgen, J. Alloys Compounds 248, 24 (1997).
  • [14] B. Chevalier, P. Lejay, J. Etourneau, and P. Hagenmuller, SolidState Commun. 49, 753 (1984).
  • [15] M. Szlawska, D. Kaczorowski, A. Ślebarski, L. Gulay, and J. Stepień-Damm, Phys. Rev. B 79, 134435 (2009).
  • [16] K. Sengupta S.Rayaprol, E.V.Sampathkumaran, Th. Doert, J.P.F.Jemetio, Physica B 348, 465-474 (2004); E. V. Sampathkumaran, T. Ekino, R. A. Ribeiro, Kausik Sengupta, T. Nakano, M. Hedo, N. Fujiwara, M. Abliz, Y. Uwatoko, S. Rayaprol, Th. Doert, and J. P. F. Jemetio, Physica B 359-361 108 (2005)
  • [17] P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka, and J. Luitz, WIEN2K: An Augmented lane Wave plus local Orbitals Program for Calculating Crystal Properties. (Karlheinz Schwarz, Technische Universität Wien, Austria, 2001) ISBN 3-9501031-1-2 (2001)
  • [18] John P. Perdew, Kieron Burke, and Matthias Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • [19] P. Hohenberg and W. Kohn, Phys. Rev. 136, 864-871 (1964).
  • [20] M. P. Seah and W.A. Dench, Surface and Interface Analysis, 1, 2 (1979).
  • [21] Yeh and Lindau, Atomic Data And Nuclear Data Tables 32, l-l55 (1985).
  • [22] S. Datta, R. P. Pandeya, A. B. Dey, A. Gloskovskii, C. Schlueter, T. Peixoto, A. Singh, A. Thamizhavel, and K. Maiti, Phys. Rev. B 105, 205128 (2022).
  • [23] A. Chainani, M. Matsunami, M. Taguchi, R. Eguchi, Y. Takata, M. Oura, S. Shin, K. Sengupta, E. V. Sampathkumaran, Th. Doert, Y. Senba, H. Ohashi, K. Tamasaku, Y. Kohmura, M. Yabashi, and T. Ishikawa, Phys. Rev. B 89, 235117 (2014).
  • [24] D. Biswas, S. Thakur, G. Balakrishnan and K. Maiti, Sci. Rep. 5, 17351 (2015).
  • [25] K. Maiti, Priya Mahadevan, and D. D. Sarma, PRL 80, 2885 (1998).
  • [26] O. Gunnarsson and K. Schonhammer, Phys. Rev. B 28, 4315 (1983).
  • [27] K. Maiti, P. Mahadevan, and D. D. Sarma, Phys. Rev. B 59, 12457 (1999).
  • [28] M. Imada, A. Fujimori, and Y.Tokura, Rev. Mod. Phys. 70, 1039 (1998).
  • [29] O. Gunnarsson and K. Schonhammer, Phys. Rev. B 31, 4815 (1985).
  • [30] C. Laubschat, E. Weschke, C. Holtz, M. Domke, O. Strebel, and G. Kaindl, Phys. Rev. Lett. 65, 1639 (1990)
  • [31] J. M. Imer and E. Wuilloud, Z. Phys. B.-Condensed Matter 66, 153-160 (1987).
  • [32] S. Datta, R. P. Pandeya, A. B. Dey, A. Gloskovskii, C. Schlueter, T. R. F. Peixoto, A. Singh, A. Thamizhavel, and K. Maiti, J. Phys.: Condens. Matter 35, 235601 (2023).
  • [33] F. C. Zhang and T. M. Rice, Phys. Rev. B 37, 3759(R) (1988).
  • [34] T. Böske, K. Maiti, O. Knauff, K. Ruck, M. S. Golden, G. Krabbes, J. Fink, T. Osafune, N. Motoyama, H. Eisaki, and S. Uchida, Phys. Rev. B 57, 138 (1998).
  • [35] K. Horiba, M. Taguchi, A. Chainani et. al., Phys. Rev. Lett. 93, 236401 (2004).
  • [36] R. P. Pandeya A. P. Sakhya, S. Datta, T. Saha, G. D. Ninno, R. Mondal, C. Schlueter, A. Gloskovskii, P. Moras, M. Jugovac, C. Carbone, A. Thamizhavel, and K. Maiti, Phys. Rev. B 104, 094508 (2021).