Skip to content
Publicly Available Published by De Gruyter September 12, 2015

Equiatomic cerium intermetallics CeXX′ with two p elements

  • Rainer Pöttgen EMAIL logo and Bernard Chevalier

Abstract

The equiatomic CeXX′ phases (X and X′ = elements of the 3rd, 4th, or 5th main group) extend the large series of CeTX intermetallics (T = electron-rich transition metal). These phases crystallize with simple structure types, i.e. ZrNiAl, TiNiSi, CeScSi, α-ThSi2, AlB2, and GdSi2. In contrast to the CeTX intermetallics one observes pronounced solid solutions for the CeXX′ phases. The main influence on the magnetic ground states results from the absence of d electrons. All known CeXX′ phases show exclusively trivalent cerium and antiferro- or ferromagnetic ordering at low temperatures. The crystal chemical details and some structure-property relationships are reviewed.

1 Introduction

The equiatomic cerium intermetallics CeTX (T = transition metal; X = element of the 3rd, 4th, or 5th main group) have recently been reviewed with respect to their crystal chemistry and magnetic ground states [1–3]. The latter is a consequence of the Ce (4f)–T(nd) hybridization, leading to a variety of magnetically different materials: (i) Pauli paramagnets with essentially tetravalent cerium, (ii) compounds which display intermediate cerium valence, (iii) Kondo compounds with extremely low magnetic ordering temperatures, (iv) heavy Fermion systems, or (v) classical antiferro- or ferromagnetically ordering phases; there is a delicate interplay between diamagnetic Ce4+ with [Xe] configuration on the one and paramagnetic Ce3+ with [Xe]4f1 configuration on the other side. So far more than 120 CeTX compounds (35 with ZrNiAl type [1]; 33 with TiNiSi type [2]; 57 with other structure types [3]) have deeply been studied with respect to their structure–property relationships and they have been classified as a function of their Ce (4f)–T(nd) hybridization along with the Doniach diagram and its updated versions [4–7]. Several hundered publications have been devoted to this outstanding class of intermetallic compounds.

A distinct modification of the magnetic properties of the equiatomic cerium intermetallics is also possible through substitution of the transition metal by another p element, thus suppressing df hybridization. Also such CeXX′ phases show interesting structural chemistry and a manifold of magnetic properties, although this group is the smallest one of all equiatomic cerium intermetallics. The structure-property relations of these phases are reviewed herein and compared with that of the large family of CeTX intermetallics.

Some of the CeXX′ phases show extended solid solutions. Herein we only consider those parts of the solid solutions that keep the structure of the equiatomic CeXX′ phase.

2 Synthesis conditions

The syntheses of the CeXX′ phases depends on the boiling points of X and X′. If X and X′ have comparatively high boiling points, synthesis via arc-melting [8–10] is possible, even in quantities of several grams. Typical examples are the α-ThSi2-type phases Ce40Si27Ge32, Ce38Si48Ge13, or Ce37Si28Ga35 [11]. Zirconium metal was used as an additional getter material. The arc-melted buttons were subsequently milled, pressed, and a final thermal-pressing treatment (1470 K, 5 GPa for 240 s) was applied [11]. Several CeSi2–xGax samples were synthesized as 20 g batches for neutron diffraction studies [12]. A different approach has been used for the CeGe2–xGax samples [13]. CeGe2 and CeGa2 have first been arc-melted as master precursors. The latter were then weighed in the appropriate ratios and melted again.

Several other studies used conventional annealing in sealed evacuated silica ampoules after arc-melting in order to equilibrate the samples. Sometimes the samples were wrapped in molybdenum foil for protection against oxidation [14]. Depending on the sample composition, the annealing temperatures chosen vary from 670 to 1270 K with annealing times up 30 days. In the early work on the Ce-Si-Ge system cerium filings, silicon, and germanium powder were pressed to pellets and directly reacted in sealed silica tubes [15].

Samples in the Ce-Al-Si system were also prepared via a flux-assisted technique. The three elements were reacted in alumina crucibles under an eutectic KCl-LiCl flux as protective coating [16]. An aluminium flux was used for the CeAlxSi2–x samples [17]. The excess aluminum was removed by centrifugation at 970 K.

Magnesium-based CeMgX samples cannot be synthesized by arc-melting. The low boiling temperature of magnesium (1363 K [18]) induces strong evaporation and mass loss. Such samples are prepared in sealed, outgassed high melting metal tubes, mainly molybdenum, tantalum or niobium [19, 20]. The reaction proceeds in an induction furnace [21], sometimes followed by further annealing of the inductively melted samples in a muffle or tube furnace. All CeMgX (X = Ga, In, Tl, Sn or Pb) phases were synthesized through this procedure [19, 20, 22–25].

Besides preparation of pure polycrystalline powders, crystal growth experiments are important for diffraction studies and direction dependent physical property investigations. Single crystal X-ray diffraction studies need only crystals of mm size. For those purposes it is possible to anneal small arc-melted buttons in a special water-cooled sample chamber in a high-frequency furnace [26]. Larger single crystals can be obtained with the Bridgman technique, e.g. a 10 cm3 crystal of CeAlGa was grown in an alumina crucible [27].

3 Discussion

3.1 Crystal chemistry and chemical bonding

We start the discussion of the CeXX′ compounds with a review on the structural chemistry. The basic crystallographic data of the respective phases are summarized in Table 1. Many of these phases exhibit broad solid homogeneity ranges towards the binary phases CeX2 and CeX2. These homogeneity ranges are considered as far as the respective phases keep the structure of the equiatomic CeXX′ compound. The CeMgX phases with X = Ga, In, Tl, Sn or Pb are also included in this review since they are also transition metal free representatives.

Table 1

Lattice parameters of different CeXX′ phases.

CompoundStructure typeSpace groupa (pm)b (pm)c (pm)V (nm3)Magnetic behaviorRefs.
Magnesium compounds
 CeMgGaZrNiAlP6̅2m752.7(2)a454.8(1)0.2232TN = 3.1 K[22, 23]
 CeMgInZrNiAlP6̅2m774.9(3)a477.7(2)0.2484n.i.[24]
 CeMgTlZrNiAlP6̅2m774.1(1)a473.75(7)0.2458n.i.[25]
 CeMgSnTiNiSiPnma772.71(2)464.80(1)908.18(2)0.3262n.i.[19]
 CeMgSnTiNiSiPnma773.0(2)464.9(1)908.6(2)0.3266n.i.[32]
 CeMgSnTiNiSiPnma773.1(1)464.9(1)908.3(1)0.3264TN = 12 K[20]
 CeMgPbCeScSiI4/mmm455.7(1)a1640.5(1)0.3406TN = 7 K[40]
Aluminum gallides
 CeAlGaAlB2P6/mmm437.8a432.90.0719TN = 6 K[27]
 CeAl1.1Ga0.9AlB2P6/mmm437a4260.0705n.i.[30]
Gallium silicides
 CeGa0.7Si1.3α-ThSi2I41/amd423.7(4)a1410(2)0.2531TC = 12 K[41]
 CeGaSiα-ThSi2I41/amd424.12(5)a1436.4(3)0.2584n.i.[37]
 CeGaSiα-ThSi2I41/amd423.9(7)a1414(3)0.2541TC = 7 K[41]
 CeGaSiα-ThSi2I41/amd422.9(1)a1435.3(2)0.2567n.i.[11]
 CeGa1.3Si0.7α-ThSi2I41/amd424.4(2)a1415(3)0.2549TC = 2 K[41]
 CeGa1.4Si0.6AlB2P6/mmm418.8a433.60.0659TC = 2.3 K[31]
 CeGa1.5Si0.5AlB2P6/mmm422.5a434.20.0671<4.2 K[33]
 CeGa1.75Si0.25AlB2P6/mmm427.6a434.60.0688TN = 5 K[33]
Aluminum silicides
 CeAl1.2Si0.8α-ThSi2I41/amd427.41(1)a1473.2(1)0.2691TN = 4.2 K[14]
 CeAl1.02Si0.98α-ThSi2I41/amd424.9(1)a1451.08(4)0.2620TC = 10 K[17]
 CeAl1.5Si0.5α-ThSi2I41/amd428.0a14900.2728n.i.[29]
 CeAlSiα-ThSi2I41/amd424.2a1453.80.2616TC = 7.1 K[35, 39]
 CeAl1.2Si0.8α-ThSi2I41/amd428.0a1475.50.2703TC = 4.2 K[35]
 CeAl1.5Si0.5AlB2P6/mmm433.0a433.50.0704TC = 6.35 K[35]
Aluminum germanides
 CeAlGeα-ThSi2I41/amd426.7a1463.00.2664TN = 4.0 K[35, 39]
 CeAlGeα-ThSi2I41/amd428.89(9)a1472.1(5)0.2708n.i.[28]
 CeAlGeα-ThSi2I41/amd425.0(1)a1471.1(4)0.2701TC = 5.6 K[14]
 CeAl1.2Ge0.8α-ThSi2I41/amd429.8a1482.60.2739TN = 3.4 K[35]
 CeAl1.5Ge0.5AlB2P6/mmm435.4a434.20.0713TC = 6.15 K[35]
 CeAl1.62Ge0.38AlB2P6/mmm435.0a429.00.0703n.i.[42]
Aluminum stannides
 CeAl1.62Sn0.38AlB2P6/mmm445.4a428.40.0736n.i.[42]
Gallium germanides
 CeGaGeα-ThSi2I41/amd429.2(1)a1451.8(3)0.2674<4.2 K[36]
 CeGaGeα-ThSi2I41/amd427.4a1454.60.2657TC = 5.5 K[38]
Silicon germanides
 CeSiGeα-ThSi2I41/amd424.3(1)a1389.7(3)0.2502n.i.[11]
 CeSi1.8Ge0.2α-ThSi2I41/amd419.4(2)a1386(2)0.2438<4.2 K[34]
 CeSi1.6Ge0.4α-ThSi2I41/amd420.1(2)a1385(2)0.2444<4.2 K[34]
 CeSi1.5Ge0.5α-ThSi2I41/amd421.0(3)a1386(2)0.2457<4.2 K[34]
 CeSi1.4Ge0.6α-ThSi2I41/amd421.2(2)a1385(2)0.2457TC = 4.3 K[34]
 CeSi1.2Ge0.8α-ThSi2I41/amd422.9(3)a1386(2)0.2479TC = 8.1 K[34]
 CeSiGeGdSi2Imma424.5(3)423.0(3)1387(2)0.2491TC = 9.5 K[34]
 CeSi0.6Ge1.4GdSi2Imma428.4(3)423.6(3)1393(2)0.2528TC = 9.4 K[34]
 CeSi0.4Ge1.6GdSi2Imma430.5(3)424.7(3)1399(2)0.2558TC = 9.3 K[34]

TN, Néel temperature; TC, Curie temperature; n.i., not investigated. For samples that did not show magnetic ordering the lowest measuring temperature is given.

3.1.1 CeXX′ phases with α-ThSi2-type structure

Most of the equiatomic CeXX′ compound crystallize with the tetragonal α-ThSi2-type structure [43], space group I41/amd. As an example we present the CeGaSi structure [37] in Fig. 1. The gallium and silicon atoms statistically occupy the 8e silicon site of the α-ThSi2 type and build up the three-dimensional network of three-connected atoms with Ga/Si–Ga/Si distances of 233 and 246 pm. The latter are compatible with the sum of the covalent radii [18] of 242 pm for Ga + Si. Within the [GaSi] network one observes parallel zig-zag chains that alternately run in x and y direction. Due to the tetragonal distortion, the bond angles within the network (118.4 and 120.8°) deviate slightly from the ideal value of 120°.

Fig. 1: The crystal structures of CeGaSi and CeAlGa. Cerium and mixed occupied Ga/Si (Al/Ga) positions are drawn as light grey and magenta circles, respectively. The [GaSi] and [AlGa] networks are emphasized.
Fig. 1:

The crystal structures of CeGaSi and CeAlGa. Cerium and mixed occupied Ga/Si (Al/Ga) positions are drawn as light grey and magenta circles, respectively. The [GaSi] and [AlGa] networks are emphasized.

The near-neighbor coordination of the cerium atoms in CeGaSi is shown in Fig. 2. Each cerium atom has twelve nearest Ga/Si neighbors. The coordination geometry can be considered as two parallel distorted pentagons which are condensed by two further Ga/Si atoms. Additionally each cerium atom has eight cerium neighbors at 417 and 423 pm. These Ce–Ce distances are well above the Hill limit (340 pm) [44] for f-electron localization. This is consistent with the magnetic data discussed below. The crystal chemical description given for CeGaSi is also adequate for the isotypic ternaries listed in Table 1.

Fig. 2: Coordination of the cerium atoms in the structures of CeAlGa and CeGaSi. Cerium and Al/Ga (Ga/Si) positions are drawn as light grey and magenta circles, respectively. Site symmetries and relevant Ce–Ce distances are indicated.
Fig. 2:

Coordination of the cerium atoms in the structures of CeAlGa and CeGaSi. Cerium and Al/Ga (Ga/Si) positions are drawn as light grey and magenta circles, respectively. Site symmetries and relevant Ce–Ce distances are indicated.

At this point arises the question on a possible ordering of the X and X′ atoms, especially if they are different in size and electron configuration. The α-ThSi2 structure allows for a 1:1 ordering on the silicon site. This coloring pattern [45] is realized for the ternary silicide LaPtSi [46] (Fig. 3). The platinum-silicon ordering leads to a loss of the center of inversion. The space group symmetry is reduced via a translationengleiche transition of index 2 (t2) from I41/amd to I41md. The corresponding group subgroup scheme in the Bärnighausen formalism [47–50] is shown in Fig. 4. Although this symmetry reduction allows full ordering, even compounds with atoms of diffent chemical potential and size reveal residual disorder. An example is the structure refinement of CePt0.97(1)Si1.03(1) [51] which showed a small degree of mixing on both 4a sites.

Fig. 3: The crystal structure of LaPtSi, space group I41md. Lanthanum, platinum, and silicon atoms are drawn as light grey, blue, and magenta circles, respectively. The three-dimensional [PtSi] network is emphasized.
Fig. 3:

The crystal structure of LaPtSi, space group I41md. Lanthanum, platinum, and silicon atoms are drawn as light grey, blue, and magenta circles, respectively. The three-dimensional [PtSi] network is emphasized.

Fig. 4: Group-subgroup scheme in the Bärnighausen formalism [47–50] for α-ThSi2, GdSi2 and LaPtSi. The indices for the translationengleiche (t) symmetry reductions, the unit cell transformations and the evolution of the atomic coordinates are given.
Fig. 4:

Group-subgroup scheme in the Bärnighausen formalism [47–50] for α-ThSi2, GdSi2 and LaPtSi. The indices for the translationengleiche (t) symmetry reductions, the unit cell transformations and the evolution of the atomic coordinates are given.

The many tetragonal CeXX′ compounds described herein all contain the weak scatterers aluminum, gallium, silicon, and germanium which all have comparable covalent radii. Since the symmetry reduction is of the translationengleiche type, the ordering of the X and X′ atoms expresses itself only in small changes of the subcell intensities and this is hardly visible in the X-ray powder patterns. The similarity in size and electronegativity also explains the formation of extended solid solutions for the tetragonal CeXX′ phases.

A second kind of ordering is possible through an orthorhombic distortion which has first been observed for the gadolinium silicide GdSi2 [52]. The space group symmetry is reduced from I41/amd to Imma (Fig. 4) and the ordering of the X and X′ atoms is possible on two crystallographically independent 4e sites. This distortion has so far only been reported for CeSiGe [34]. The electronic reasons for this kind of distortion are not yet known in detail.

Assuming trivalent cerium (see the magnetic properties discussed below), the valence electron concentration (VEC) for the fully ordered CeXX′ phases ranges from 8 to 11. Those with the α-ThSi2 or GdSi2 structure have 10 and 11. CeGaSi, CeAlSi, CeAlGe, and CeGaGe (all with VEC 10) can be described as electron precise Zintl phases with three-bonded Al2–, Ga2–, Si, and Ge which are isoelectronic with phosphorus and arsenic. This electronic situation is similar to that of LaAlGe [53–55] which shows semimetal behavior (the DOS tends to zero at the Fermi level). The corresponding polyanionic networks show three two-electron two-center bonds with almost equal bond lengths. Within the solid solutions, the Fermi level is either lowered (higher content of the group III element, e.g. CeAl1.5Si0.5) or increased (higher content of the group IV element, e.g. CeSiGe), tending to higher density of states and consequently an increase of the metallic character of the sample. This tendency is similar for the pair LaAlGe/La2LiAlGe2 [55].

3.1.2 CeXX′ phases with AlB2-type structure

The second group of CeXX′ phases crystallizes with the AlB2-type structure, space group P6/mmm. The structure of CeAlGa is shown as an example in Fig. 1. All X-ray powder diffraction experiments revealed a statistical distribution of the group III and group IV elements on the honeycomb network. The near neighbor coordination of the cerium atoms in the α-ThSi2 and AlB2 type is compared in Fig. 2. In both structure types the cerium atoms have twelve p-element neighbors, however, in a different arrangement: (i) two planar hexagons in the AlB2-type phases and (ii) two planar, distorted pentagons which are condensed via two p-element atoms in the α-ThSi2-type phases. As a consequence of the different site symmetry, in both types one observes two sets of distances, e.g. 12 × 333 pm in CeAlGa as compared to 8 × 321 pm + 4 × 322 pm for CeGaSi. Each cerium atom has eight cerium neighbors, 6 + 2 in CeAlGa and 4 + 4 in CeGaSi, again a consequence of the change in site symmetry. The p-element substructures are both three-connected nets with different conjugations. Detailed studies of chemical bonding were performed by Zheng and Hoffmann [56]. They showed that filling of the π* band with more than 4.5 electrons per p-element atom leads to stabilization of the α-ThSi2 as compared to the AlB2 type.

The arguments for a p-element ordering also hold for the AlB2-type phases. The differences in the chemical potential should result in a coloring of the honeycomb networks. The many AlB2 superstructures have been discussed on the basis of group–subgroup schemes [45, 57]. The two possibilities for the phases discussed herein are shown in Fig. 5. The pure equiatomic phases can exhibit a 1:1 ordering. This leads to a lowering of the space group symmetry. Several samples within the solid solutions (Table 1) have compositions CeX1.5X0.5. This allows a 3:1 ordering with X6 rings that are condensed via the X′ atoms (Fig. 5). Although these two ordering patterns are possible, all phases have been described with the structure of the disordered aristotype.

Fig. 5: The 1:1 and 3:1 ordering (coloring) variants of honeycomb networks.
Fig. 5:

The 1:1 and 3:1 ordering (coloring) variants of honeycomb networks.

Most CeXX′ phases show pronounced homogeneity ranges, which were studied completely for the Al-Si [58], Al-Ge [59], and Al–Ga [60] systems where the whole ternary phase diagrams were constructed as isothermal sections.

3.1.3 CeMgX phases

The third group of CeXX′ phases covers the magnesium-based compounds CeMgX with X = Ga, In, Tl, Sn, or Pb. The gallide, indide, and thallide crystallize with the ZrNiAl-type structure. As an example we present the CeMgGa structure in Fig. 6. The magnesium and gallium atoms build up a three-dimensional [MgGa]δ– polyanionic network which consists of distorted Mg3Ga2 pentagons which are condensed via further gallium atoms. Within the polyanion each cerium atom is coordinated by two Mg3Ga2 pentagons and a bridging gallium atom. The coordination number with respect to the main group element atoms is 11 as compared to 12 in the AlB2- and α-ThSi2-type phases. The more electronegative gallium atoms are the nearest neighbors to cerium. Electronic structure calculations underline the strongly bonding Ce–Ga interactions [61, 62]. For chemical bonding in the indide and the thallide we can safely assume a rigid band model. Further crystal chemical details on ZrNiAl-type cerium intermetallics have been summarized in [1].

Fig. 6: The crystal structures of CeMgSn, CeMgGa and CeMgPb. Cerium, magnesium and tin (gallium, lead) atoms are drawn as light gray, blue and magenta circles, respectively. The three-dimensional [MgSn] and [MgGa] networks are emphasized. The layers of edge-sharing Ce4/4 tetrahedra and corner-sharing Mg4/2Pb2 octahedra are drawn for CeMgPb.
Fig. 6:

The crystal structures of CeMgSn, CeMgGa and CeMgPb. Cerium, magnesium and tin (gallium, lead) atoms are drawn as light gray, blue and magenta circles, respectively. The three-dimensional [MgSn] and [MgGa] networks are emphasized. The layers of edge-sharing Ce4/4 tetrahedra and corner-sharing Mg4/2Pb2 octahedra are drawn for CeMgPb.

CeMgSn adopts the orthorhombic TiNiSi type. The magnesium and tin atoms both have distorted tetrahedral coordination within the three-dimensional [MgSn]δ– polyanionic network (Fig. 6). The cerium atoms are coordinated by two strongly titled and othorhombically distorted Mg3Sn3 hexagons (290–304 pm Mg–Sn). The structural chemistry of CeMgSn very much resembles that of the isopointal CeTX phases [2]. As expected from the course of the electronegativities, the tin atoms are the next nearest neighbors to cerium (i.e. one observes the inverse coloring within the polyanionic network). In total, the cerium and magnesium atoms can deliver five valence electrons. A Sn4– stannide anion however can only carry four charges. This simple electron counting readily points to a metallic character for CeMgSn with the surplus electron within the conduction band. Accordingly, electronic structure calculations [63] point to negatively charged tin, underlining the stannide nature. Overlap population analyses show the Ce–Sn interactions as the strongest ones in the CeMgSn structure, as expected from the course of the electronegativities.

The last compound in this series is the plumbide CeMgPb which crystallizes with the tetragonal CeScSi-type structure, space group I4/mmm. From a purely geometrical point of view one can describe the CeMgPb structure with two basic building units, i.e. layers of corner-sharing Mg4/2Pb2 octahedra that are separated by layers of edge-sharing Ce4/4 tetrahedra (Fig. 6). As a consequence of the body-centered lattice, every other of these layers is shifted by a/2 + b/2 with respect to each other. Within the octahedral layer the magnesium square nets have Mg–Mg distances of 321 pm and the capping lead atoms are at 315 pm Mg–Pb. The cerium coordination is shown in Fig. 7. Each cerium atom has five nearest lead neighbors at Ce–Pb distances of 320 (1×) and 326 (4×) pm in form of a square pyramid. The first coordination sphere is in agreement with the course of the electronegativities. The square face of the pyramid is capped by a square of magnesium atoms in staggered conformation. Similar to CeMgSn, charge analyses [63] reveal plumbide character for CeMgPb and the Ce–Pb interactions as the strongest ones in the structure.

Fig. 7: Coordination of the cerium atoms in CeMgPb. Cerium, magnesium and lead atoms are drawn as light grey, blue and magenta circles, respectively. Relevant interatomic distances are indicated.
Fig. 7:

Coordination of the cerium atoms in CeMgPb. Cerium, magnesium and lead atoms are drawn as light grey, blue and magenta circles, respectively. Relevant interatomic distances are indicated.

3.2 Magnetic properties

In contrast to the many intermediate valence CeTX compounds discussed [1–3], one observes exclusively trivalent cerium in the CeXX′ phases summarized herein. The missing Ce(4f)-T(nd) hybridization seems to suppress the possibility for an enhanced cerium valence. Thus magnetic ordering of the cerium magnetic moments can be expected for all CeXX′ phases, even though the ordering temperature may be very low. This can be a consequence of structural disorder within the solid solutions.

We start the summary of the magnetic data with the magnesium containing phases CeMgX. With the triels, so far only magnetic data of CeMgGa have been reported [22]. The cerium magnetic moments show antiferromagnetic ordering at a Néel temperature of 3.1 K and a pronounced metamagnetic step at a critical field strength of 1 T. The saturation moment at 1.72 K and 5 T of 0.81 μB per cerium atom is significantly reduced from the theoretical saturation magnetization of 2.14 μB, a consequence of strong crystal field splitting.

The magnetic properties of CeMgSn have thoroughly been studied [19, 20, 64, 65]. Susceptibility measurements of CeMgSn show antiferromagnetic ordering at TN = 12 K [20], the highest magnetic ordering temperature among all CeXX′ phases. Magnetization measurements show a metamagnetic transition at 2 T. The saturation magnetization at 5 K and 9 T is almost 1.4 μB per cerium atom, distinctly higher than for CeMgGa. The magnetic structure of CeMgSn has been studied independently by Ritter et al. [19] and Lemoine et al. [65]. Both investigations have led to the same incommensurate sine-wave modulated magnetic structure with a propagation vector k = [0, 0.1886, 0.3384].

CeMgPb shows a slightly lower Néel temperature of 7 K [40] and also exhibits a metamagnetic transition at a slightly higher critical field of 2.8 T. The saturation magnetization at 5 K and 9 T is 1.1 μB per cerium atom. Neutron powder diffraction data taken at 2 K indicate a complex magnetic structure. The recorded pattern could be indexed by using two propagation vectors, a commensurate one k1 = [1/2, 1/2, 0] and and incommensurate one k2 = [0.448, 1/2, 0], leading to an antiferromagnetic and sine-modulated ordering of the magnetic moments.

The complete solid solution CeSi2–xGex has been studied with respect to the magnetic properties [34]. The low-temperature magnetic behavior is of a ferromagnetic type. The Curie temperature increases from CeGe2 (TC = 6.8 K) to a maximum of 9.5 K for CeSiGe (Fig. 8). Further increase of the silicon content leads to a drastic decrease of TC, driven by structural disorder. From ordered CeSiGe to CeSi2 one observes a huge increase of the Weiss constant from –16 to –286 K and the samples switch towards Fermi liquid behavior.

Fig. 8: Course of the Curie temperature TC as a function of the composition in the solid solution CeSi2–xGex (TC data taken from ref. [34]).
Fig. 8:

Course of the Curie temperature TC as a function of the composition in the solid solution CeSi2–xGex (TC data taken from ref. [34]).

The gallide CeGeGa shows Curie–Weiss behavior with an experimental magnetic moment of 2.56 μB per cerium atom [36]. CeGeGa orders ferromagnetically at TC = 5.5 K [38]. This is evident from magnetic susceptibility and specific heat data in the low temperature regime. The heat capacity data point to Kondo interactions and a Kondo temperature of 2.1 K was derived from the experimental data. CeGeGa shows a broad solid solution CeGe2–xGax in the range 0.7 ≤ x ≤ 1.3 [13]. Both p elements show certain flexibility for substitution while keeping the α-ThSi2 structure. However, the Ge/Ga ratio has a distinct influence on the magnetic ground state. CeGe1.3Ga0.7 orders ferromagnetically (similar to CeGeGa), while CeGe0.7Ga1.3 is an antiferromagnet.

The magnetic susceptibility and the magnetic structure of CeAlGa were thoroughly studied. CeAlGa [27, 65] orders antiferromagnetically at TN = 6 K. Magnetization measurements on a single crystalline specimen reveal two metamagnetic steps in the a and b direction, however, with a small magnetization value of only 0.28 μB per cerium atom. The magnetic structure is of an incommensurate modulated type with k = 0.156, 0.156, 0. A sample with composition CeAl0.2Ga1.8 shows the same type of magnetic structure (k = 0.14, 0.14, 0) but a distinctly higher Néel temperature of 11 K [66]. Inelastic neutron scattering experiments show transitions that are driven by the crystalline electric field (CEF). Fitting of the CEF Hamiltonian leads to similar values for the powder (B20=11 K,B40=0.33 K) and single crystalline (B20=10.3 K,B40=0.166 K,  B23=2 K) material.

The silicide CeAlSi and the germanide CeAlGe along with their solid solutions to the Al-, respectively Si(Ge)-richer regions were studied with respect to structural and magnetic data [14, 17, 35, 39, 67, 68]. The magnetic properties of CeAl1+xSi1–x are highly sensitive to the sample composition. Ferromagnetic ordering at 7.1 K was reported for the equiatomic compound CeAlSi [35], while a higher Curie temperature of 10 K was measured for CeAl1.02Si0.98 [17] (susceptibility and resistivity data). A further increase of the aluminum content leads to a decrease of TC. Two independent studies [14, 35] report TC = 4.2 K for CeAl1.2Si0.8. A discrepancy happens for CeAlGe. Dhar et al. report antiferromagnetic ordering at TN = 4.0 K [35] while Flandorfer et al. [14] observed a ferromagnetic ground state below TC = 5.6 K. These differences might be a consequence of different domain structures. The trivalent nature of cerium in CeAlSi and CeAlGe was substantiated from XAS spectra [14].

The most intensive studies were performed for CeSiGa and corresponding samples of the solid solution CeSi2–xGax in the range 0.5 ≤ x ≤ 1.3 [12, 31, 33, 41, 69–75]. Starting from binary CeSi2, the gallium for silicon substitution leads to a two-phase region along with a significant increase of the unit cell volume by around 3 %. This two-phase region extends up to a composition CeSi1.6Ga0.4. Since this structural behavior has not been observed for the corresponding lanthanum compound, it can safely be ascribed to the 4f electrons [69]. The course of the magnetic ordering temperature is not linear. First one observes an increase from x = 0.5 (TC ca. 12 K) to x = 1 (TC ca. 14 K) followed by a sharp decrease to ca 2 K for the x = 1.4 sample [31, 70, 73, 74]. The magnetic transitions were evident in the susceptibility and specific heat data as well. Slightly different (lower) transition temperatures were reported in [41]. This might be a consequence of a different domain composition. The crystal field parameters of equiatomic CeSiGa and members of the solid solution CeSi2–xGax (0.7 ≤ x ≤ 1.3) were determined by inelastic neutron scattering studies [12, 71, 72], indicating a crossover from a magnetic 4f state to a dense ferromagnetic Kondo system.

3.3 Electrical properties

The temperature dependence of the resistivity of CeMgGa [22] is indicative of metallic behavior. It can be fitted with a formula that includes terms that account for temperature-independent resistivity and spin-disorder resistivity as well as a Bloch-Grüneisen term for scattering of conduction electrons and a term for interband scattering processes. CeMgGa shows large and positive magnetoresistance with a value of 26 % at a field of 8 T.

Resistivity measurements on CeGeGa and its non-magnetic counterpart LaGeGa seriously suffered from large microcracks within the samples [38]. Both gallides showed relatively high values for the specific resistivities at room temperature and only a moderate decrease of the absolute values down to 1.5 K. A broad bump around 150 K was ascribed to crystal field splitting of the 4f level, and the magnetic ordering of the cerium magnetic moments was evident from a sharp drop around 8 K as a consequence of freezing of the spin-disorder scattering. Similar drops were observed for ferromagnetic CeGe1.3Ga0.7 and antiferromagnetic CeGe0.7Ga1.3 [13]. Also for the pair of compounds LaAlGa and CeAlGa, the influence of the magnetic cerium atoms on the resistivity behavior was studied in detail [27]. The magnetic contribution and the Néel temperature of 6 K were directly evident from a difference plot ρCeAlGa – ρLaAlGa.

Reduced spin-disorder scattering was observed for the CeSi2–xGax (0.7 ≤ x ≤ 1.3) samples along with a weak Kondo minimum [41]. Galvanometric studies revealed negative magnetoresistance behavior [75].

3.4 Thermoelectric properties

The α-ThSi2-type phases Ce40Si27Ge32, Ce38Si48Ge13 and Ce37Si28Ga35 have been studied in comparison with CeSi2 with respect to their thermoelectric properties [11] in the temperature regime from 240 to 380 K. The respective samples are moderately thermoelectric materials with the maximal ZT value for CeSi2. Formation of solid solutions through Si–Ge, respectively Si–Ga substitution worsens the thermoelectric parameters.

4 Summary

The family of CeXX′ intermetallics comprises 12 compounds along with their solid solutions with higher X or X′ content. Their crystal structures derive from the binary structure types AlB2, α-ThSi2, Fe2P, GdSi2, or La2Sb with partial or full ordering of the p elements. As a main difference to their transition metal containing counterparts, independent of the composition, all CeXX′ phases show stable trivalent cerium ground states. So far only few complete ranges of solid solutions were studied with respect to the property changes. One of the impressive examples is the solid solution CeSi2–xGex which shows a pronounced drop of the Curie temperature with increasing silicon content. Future investigations might focus on other series of solid solutions and different compositions. Many other combinations are conceivable. Until now no representatives with a pnictide component are known.


Corresponding author: Rainer Pöttgen, Institut für Anorganische und Analytische Chemie, Universität Münster, Corrensstrasse 30, 48149 Münster, Germany, e-mail:

Acknowledgments

This work was supported by the Deutsche Forschungsgemeinschaft. We thank Dr. R.-D. Hoffmann for discussion on the group-subgroup scheme.

References

[1] R. Pöttgen, B. Chevalier, Z. Naturforsch.2015, 70b, 289.Search in Google Scholar

[2] O. Janka, O. Niehaus, R. Pöttgen, B. Chevalier, Z. Naturforsch.2015, 70b, to be published.Search in Google Scholar

[3] O. Niehaus, O. Janka, R. Pöttgen, B. Chevalier, Z. Naturforsch.2015, 70b, to be published.Search in Google Scholar

[4] A. Szytuła, J. Leciejewicz, Handbook of Crystal Structures and Magnetic Properties of Rare Earth Intermetallics, CRC Press, Boca Raton, 1994.Search in Google Scholar

[5] S. Doniach, Physica1977, 91B, 231.10.1016/0378-4363(77)90190-5Search in Google Scholar

[6] J. R. Iglesias, C. Lacroix, B. Coqblin, Phys. Rev. B1997, 56, 11820.10.1103/PhysRevB.56.11820Search in Google Scholar

[7] Y.-F. Yang, Z. Fisk, H.-O. Lee, J. D. Thompson, D. Pines, Nature2008, 454, 611.10.1038/nature07157Search in Google Scholar

[8] R. Pöttgen, Th. Gulden, A. Simon, GIT Labor-Fachzeitschrift1999, 43, 133.Search in Google Scholar

[9] R. Ferro, A. Saccone, Intermetallic Chemistry, Elsevier, Amsterdam, 2008.Search in Google Scholar

[10] R. Pöttgen, D. Johrendt, Intermetallics, De Gruyter, Berlin, 2014.10.1524/9783486856187Search in Google Scholar

[11] A. V. Morozkin, V. A. Stupnikov, V. N. Nikiforov, N. Imaoka, I. Morimoto, J. Alloys Compd.2006, 415, 12.Search in Google Scholar

[12] K. R. Priolkar, M. N. Rao, R. B. Prahbu, P. R. Sarode, S. K. Paranjpe, P. Raj, A. Sathyamoorthy, J. Phys.: Condens. Matter1998, 10, 10557.10.1088/0953-8984/10/47/007Search in Google Scholar

[13] N. G. Patil, S. K. Dhar, Physica B1996, 223&224, 359.10.1016/0921-4526(96)00122-6Search in Google Scholar

[14] H. Flandorfer, D. Kaczorowski, J. Gröbner, P. Rogl, R. Wouters, C. Godart, A. Kostikas, J. Solid State Chem.1998, 137, 191.Search in Google Scholar

[15] H. Haschke, H. Nowotny, F. Benesovsky, Monatsh. Chem.1966, 97, 1452.Search in Google Scholar

[16] L. N. Altunina, E. I. Gladyshevskii, O. S. Zarechnyuk, I. F. Kolobnev, Russ. J. Inorg. Chem.1963, 8, 870.Search in Google Scholar

[17] S. Bobev, P. H. Tobash, V. Fritsch, J. D. Thompson, M. F. Hundley, J. L. Sarrao, Z. Fisk, J. Solid State Chem.2005, 178, 2091.Search in Google Scholar

[18] J. Emsley, The Elements, Oxford University Press, Oxford (U.K.) 1999.Search in Google Scholar

[19] C. Ritter, A. Provino, P. Manfrinetti, K. A. Gschneidner Jr., J. Alloys Compd.2011, 509, 9724.Search in Google Scholar

[20] P. Lemoine, A. Vernière, J. F. Marêchè, B. Malaman, J. Alloys Compd.2010, 508, 9.Search in Google Scholar

[21] R. Pöttgen, A. Lang, R.-D. Hoffmann, B. Künnen, G. Kotzyba, R. Müllmann, B. D. Mosel, C. Rosenhahn, Z. Kristallogr.1999, 214, 143.Search in Google Scholar

[22] R. Kraft, R. Pöttgen, D. Kaczorowski, Chem. Mater.2003, 15, 2998.Search in Google Scholar

[23] R. Kraft, M. Valldor, R. Pöttgen, Z. Naturforsch.2003, 58b, 827.Search in Google Scholar

[24] R. Kraft, M. Valldor, D. Kurowski, R.-D. Hoffmann, R. Pöttgen, Z. Naturforsch.2004, 59b, 513.Search in Google Scholar

[25] R. Kraft, R. Pöttgen, Z. Naturforsch.2005, 60b, 265.Search in Google Scholar

[26] D. Niepmann, Yu. M. Prots’, R. Pöttgen, W. Jeitschko, J. Solid State Chem.2000, 154, 329.Search in Google Scholar

[27] M. A. Fremy, D. Gignoux, D. Schmitt, A. Y. Takeuchi, J. Magn. Magn. Mater.1989, 82, 175.Search in Google Scholar

[28] J. T. Zhao, E. Parthé, Acta Crystallogr.1990, C46, 2276.Search in Google Scholar

[29] A. Raman, H. Steinfink, Inorg. Chem.1967, 6, 1789.Search in Google Scholar

[30] D. I. Dzyana, O. S. Zarechnyuk, Vis. Lviv Dez. Univ., Ser. Chim.1969, 11, 8.Search in Google Scholar

[31] K. R. Priolkar, S. M. Pattalwar, P. K. Mishra, P. Raj, A. Sathyamoorthy, S. K. Dhar, V. C. Sahni, P. R. Sarode, R. B. Prabhu, Solid State Commun.1997, 104, 71.Search in Google Scholar

[32] P. Manfrinetti, A. Provino, K. A. Gschneidner Jr., J. Alloys Compd.2009, 482, 81.Search in Google Scholar

[33] S. K. Dhar, S. M. Pattalwar, R. Vijavaraghavan, Solid State Commun.1993, 87, 409.Search in Google Scholar

[34] R. Lahiouel, R. M. Galera, J. Pierre, E. Siaud, Solid State Commun.1986, 58, 815.Search in Google Scholar

[35] S. K. Dhar, S. M. Pattalwar, J. Magn. Magn. Mater.1996, 152, 22.Search in Google Scholar

[36] Y. N. Grin, P. Rogl, B. Chevalier, A. A. Fedorchuk, I. A. Gryniv, J. Less-Common Met.1991, 167, 365.Search in Google Scholar

[37] J. Zhao, J. Chin. Rare Earth Soc.1999, 17, 169.Search in Google Scholar

[38] S. K. Dhar, S. M. Pattalwar, R. Vijayaraghavan, Physica B1993, 186–188, 491.10.1016/0921-4526(93)90613-BSearch in Google Scholar

[39] S. K. Dhar, S. M. Pattalwar, R. Vijayaraghavan, J. Magn. Magn. Mater.1992, 104–107, 1303.Search in Google Scholar

[40] P. Lemoine, A. Vernière, G. Venturini, J. F. Marèché, S. Capelli, B. Malaman, J. Magn. Magn. Mater.2012, 324, 2937.Search in Google Scholar

[41] K. R. Priolkar, R. B. Prahbu, P. R. Sarode, V. Ganesan, P. Raj, A. Sathyamoorthy, J. Phys.: Condens. Matter1998, 10, 4413.10.1088/0953-8984/10/20/009Search in Google Scholar

[42] A. Raman, Z. Metallkd.1967, 58, 179.Search in Google Scholar

[43] G. Brauer, A. Mitius, Z. Anorg. Allg. Chem.1942, 249, 325.Search in Google Scholar

[44] H. H. Hill, in Plutonium and Other Actinides, Nuclear Materials Series, AIME, Vol. 17 (Ed.: W. N. Mines), 1970, p. 2.Search in Google Scholar

[45] R. Pöttgen, Z. Anorg. Allg. Chem.2014, 640, 869.Search in Google Scholar

[46] K. Klepp, E. Parthé, Acta Crystallogr.1982, B38, 1105.Search in Google Scholar

[47] H. Bärnighausen, Commun. Math. Chem.1980, 9, 139.Search in Google Scholar

[48] U. Müller, Z. Anorg. Allg. Chem.2004, 630, 1519.Search in Google Scholar

[49] U. Müller, in International Tables for Crystallography, Vol. A1, Symmetry relations Between Space Groups, (Eds.: H. Wondratschek, U. Müller), John Wiley & sons, Ltd, 2nd Ed., Chichester, 2010, p. 44.10.1107/97809553602060000795Search in Google Scholar

[50] U. Müller, Symmetriebeziehungen zwischen verwandten Kristallstrukturen, Vieweg + Teubner Verlag, Wiesbaden, 2012.10.1007/978-3-8348-8342-1Search in Google Scholar

[51] W. Hermes, U. Ch. Rodewald, B. Chevalier, R. Pöttgen, Z. Naturforsch.2007, 62b, 613.Search in Google Scholar

[52] J. A. Perri, I. Binder, B. Post, J. Phys. Chem.1959, 63, 616.Search in Google Scholar

[53] A. M. Guloy, J. D. Corbett, Inorg. Chem.1991, 30, 4789.Search in Google Scholar

[54] J. D. Corbett, in Chemistry, Structure, and Bonding of Zintl Phases and Ions, (Ed.: S. M. Kauzlarich), VCH, Weinheim 1996, chapter 3, p. 139.Search in Google Scholar

[55] A. Stetskiv, I. Tarasiuk, V. Pavlyuk, Chem. Met. Alloys2012, 5, 148.Search in Google Scholar

[56] C. Zheng, R. Hoffmann, Inorg. Chem.1989, 28, 1074.Search in Google Scholar

[57] R.-D. Hoffmann, R. Pöttgen, Z. Kristallogr.2001, 216, 127.Search in Google Scholar

[58] P. Rogl, Aluminium-Cerium-Silicon, in Ternary Alloys, Vol. 4 (Eds.: G. Petzow, G. Effenberg), VCH, Weinheim, 1991, p. 100.Search in Google Scholar

[59] R Schmid-Fetzer, Aluminium-Cerium-Germanium, in Ternary Alloys, Vol. 4 (Eds.: G. Petzow, G. Effenberg), VCH, Weinheim, 1991, p. 67.Search in Google Scholar

[60] R Schmid-Fetzer, Aluminium-Cerium-Gallium, in Ternary Alloys, Vol. 4 (Eds.: G. Petzow, G. Effenberg), VCH, Weinheim, 1991, p. 64.Search in Google Scholar

[61] R. P. Singh, J. Magnesium Alloys2014, 2, 349.10.1016/j.jma.2014.10.004Search in Google Scholar

[62] S. F. Matar, J. Etourneau, R. Pöttgen, Solid State Sci.2015, 43, 28.Search in Google Scholar

[63] S. F. Matar, R. Pöttgen, Solid State Sci., in press. DOI 10.1016/j.solidstatesciences.2015.07.008.Search in Google Scholar

[64] P. Lemoine, A. Vernière, G. Venturini, B. Malaman, J. Tobola, Diffus. Defect Data, Pt. B2011, 170, 321.Search in Google Scholar

[65] P. Lemoine, A. Vernière, G. Venturini, S. Capelli, B. Malaman, J. Magn. Magn. Mater.2012, 324, 961.Search in Google Scholar

[66] M. Jerjini, M. Bonnet, P. Burlet, G. Lapertot, J. Rossat-Mignod, J. Y. Henry, D. Gignoux, J. Magn. Magn. Mater.1988, 76&77, 405.Search in Google Scholar

[67] A. A. Murav’eva, O. S. Zarechnyuk, Inorg. Mater.1970, 6, 933.Search in Google Scholar

[68] T. I. Yanson, O. S. Zarechnyuk, E. I. Gladyshevskii, Tezisy Dokl. – Vses. Konf. Kristallokhim. Intermet. Soedin1974, 35.Search in Google Scholar

[69] H. Mori, N. Sato, T. Satoh, Solid State Commun.1984, 49, 955.Search in Google Scholar

[70] V. V. Moshchalkov, O. V. Petrenko, M. K. Zalyalyutdinov, Physica B1990, 163, 395.10.1016/0921-4526(90)90222-GSearch in Google Scholar

[71] K. R. Priolkar, R. B. Prahbu, P. R. Sarode, M. N. Rao, S. K. Paranjpe, P. Raj, A. Sathyamoorthy, Solid State Commun.1997, 101, 825.Search in Google Scholar

[72] K. R. Priolkar, M. N. Rao, R. B. Prahbu, P. R. Sarode, S. K. Paranjpe, P. Raj, A. Sathyamoorthy, J. Magn. Magn. Mater.1998, 185, 375.Search in Google Scholar

[73] V. V. Moshchalkov, O. V. Petrenko, M. K. Zalyalyutdinov, I. Chirich, JETP Lett.1990, 51, 328.Search in Google Scholar

[74] V. V. Moshchalkov, O. V. Petrenko, M. V. Semenov, R. V. Lutsiv, O. I. Babich, Sov. J. Low Temp. Phys.1987, 13, 723.Search in Google Scholar

[75] N. B. Brandt, V. V. Moshchalkov, O. V. Petrenko, Yu. N. Grin, I. M. Gryniv, Sov. Phys. Solid State1988, 30, 192.Search in Google Scholar

Received: 2015-6-26
Accepted: 2015-7-7
Published Online: 2015-9-12
Published in Print: 2015-10-1

©2015 by De Gruyter

Downloaded on 27.4.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2015-0109/html
Scroll to top button