Skip to main content
ACS Omega logoLink to ACS Omega
. 2022 Jun 3;7(23):19048–19057. doi: 10.1021/acsomega.2c02152

Investigating the An+1BnX3n+1 Homologous Series: A New Platform for Studying Magnetic Praseodymium Based Intermetallics

Trent M Kyrk , Moises Bravo , Gregory T McCandless , Saul H Lapidus , Julia Y Chan †,*
PMCID: PMC9202054  PMID: 35721977

Abstract

graphic file with name ao2c02152_0006.jpg

The recent discovery of the An+1BnX3n+1 (A = lanthanide, B = transition metal, X = tetrel) homologous series provides a new platform to study the structure–property relationships of highly correlated electron systems. Several members of Cen+1ConGe3n+1 (n = 1, 4, 5, 6, and ∞) show evidence of heavy electron behavior with complex magnetic interactions. While the Ce analogues have been investigated, only n = 1, 2, and ∞ of Prn+1ConGe3n+1 have been synthesized, with n = 1 and 2 showing a nonsinglet magnetic ground state. The Pr analogues can provide a platform for direct comparison of highly correlated behavior. In this perspective, we discuss the impetus for synthesizing the Prn+1ConGe3n+1 members and present the structural characterization of the n = 3 and n = 4 members. We lay the foundation for future investigations of the Prn+1ConGe3n+1 family of compounds and highlight the importance of complementary methods to characterize new quantum materials.

1. Introduction

The discovery of novel emergent electronic states in magnetic highly correlated quantum materials1,2 is necessary to advance next generation technologies.3 These states allow the stabilization of a number of quasi-particles that can be taken advantage of for new technologies or advance our understanding of important physical phenomena, such as Majorana fermions in topological insulators for quantum computing,4,5 Weyl fermionic states in topological semimetals to study magnetotransport phenomena,6 and other highly correlated states.3,4,79 Herein, we focus on heavy fermions and their quantum critical points, around which unconventional superconductivity has been observed.10,11 From a materials perspective, our goals are to (1) connect highly correlated physics to the structural and electronic features of a material, (2) identify and synthesize quantum materials, and (3) understand the synthesis of quantum materials. These goals aim to discover new platforms of compounds with magnetic and transport properties that can be tuned to study quantum critical points.10 While external parameters such as temperature, pressure, and magnetic field are necessary in these investigations, additional parameters such as dimensionality and the structural environment of magnetic ions can enrich our understanding of structure–property relationships.

Occasionally, homologous series, families of related compounds with a chemical formula that changes as a function of an integer, are discovered that allow a robust method of tuning physical properties through discrete variations in chemical structure and composition. These families can be built from intergrowth structures, where each member is rationalized as an assembly of known crystal structures. In recent decades, several intermetallic homologous series have been discovered such as CenMIn3n+2,12,13 Th2(AuxSi1–x)[AuAl2]nSi2,14 Ce(PdAl)n,15 and [LnGe2]n[LnLi2Ge]m.16,17 The nature of a homologous series makes them specially systematic to study complex phenomena as linked to their repeating structural motifs, thus allowing novel connections to be drawn between chemical structure and properties in ways dissimilar compounds cannot achieve. The Ln2–yZn4As4·n(LnAs) (Ln = La–Nd, Sm, and Gd) series built from [LnAs] and [Zn2As2] slabs18 is an example of band gap tuning as a function of structure. Here, there is a gradual increase in electron density at the band gap as a function of the [LaAs] slabs’ thickness, directly attributed to the change in structure and composition. CenMIn3n+2 (n = 1, 2, and ∞; M = Co, Rh, and Ir) offered new potential in studying magnetically mediated superconductivity in the CeIn3 parent structure,19 CeMIn5,20,21 and Ce2MIn822 by considering the role of the CeIn3 and MIn2 layers and transition metal substitution. To evaluate the highly correlated nature of compounds, the Sommerfeld coefficient (γ), the electronic contribution to specific heat, and a measure of heavy electron behavior can be obtained from the following equation: Cp = γT + αT3, where T is the temperature and α is the phonon contribution to the specific heat. With application of pressure, the antiferromagnets emerge into a superconducting phase,13 with an enhanced Sommerfeld coefficient of γ ∼ 400 mJ/mol K2 for CenRhIn3n+2 (n = 1 and 2) and γ ∼ 700 mJ/mol K2 for CenIrIn3n+2 (n = 1 and 2) and, most prominently, with the antiferromagnetic CeIn3 (TN = 10 K) becoming superconducting with 27.5 kbar. The properties of this family of compounds led to the linking of dimensionality, in addition to the evaluation of the significance of the Ce local cuboctahedral environment on the influence of magnetic correlation. Given a new intermetallic homologous series, we present a robust platform to demonstrate the potential for the discovery of new phenomena and the study of exotic and emergent behavior as a function of dimensionality.

2. An+1ConGe3n+1

The An+1BnX3n+1 homologous series (A = lanthanide, B = transition metal, X = tetrel)23 holds promise as a new platform to study highly correlated emergent phenomena. The progression of the series can be understood as systematic additions of BaNiSn3/AuCu3-type disordered subunits intergrown with CeNiSi2-type subunits, where odd and even members crystallize in the orthorhombic Cmcm and Cmmm space groups, respectively. For each n member of the series, n – 1 BaNiSn3/AuCu3-type subunits are stacked along the b direction between CeNiSi2-type subunits, where the sequential stacking of BaNiSn3/AuCu3 subunits separates the CeNiSi2 subunits as a function of n. Therefore, the homologous series provides a systematic platform to study the effects of changing dimensionality and local environment on magnetism and electronic structure.

Cen+1ConGe3n+1 compounds n = 1,24 4,25,26 5,27 6,28 and ∞29 have been reported and can be used to study the interplay of local 4f1 electron systems and conduction electrons. Ce3+ ions are odd spin systems that follow Kramer’s theorem, which guarantees the formation of a doublet electronic ground state outside of an applied magnetic field due to the presence of time reversal symmetry. At sufficiently low temperatures, the doublet ground state allows for local 4f and conduction electron hybridization that results in the emergence of the magnetic Kondo effect, manifesting as heavy electron behavior, and eventually a valence transition from Ce3+ to Ce4+.30 As illustrated by the Doniach model,31 Kondo interactions are in direct competition with Ruderman–Kittel–Kasuya–Yosida (RKKY) interactions dependent upon the strength of magnetic exchange interactions, making compounds that exhibit both heavy electron behavior and magnetic ordering exciting platforms for studying quantum critical points. All characterized members of the Cen+1ConGe3n+1 family, n = 1, 4, 5, 6, and ∞, show evidence of heavy electron behavior, and the n = 4, 5, 6, and ∞ members demonstrate complex anisotropic competition between antiferromagnetic and ferromagnetic states.

The n = 1 member, CeCo0.89Ge2, is a Kondo compound (TK > 200 K) with no magnetic ordering down to 1.5 K.3234 However, patterns have emerged in the properties of the Cen+1ConGe3n+1 family for n = 4–6. The anisotropic magnetic properties of Ce5Co4+xGe13, Ce6Co5+xGe16, and Ce7Co6+xGe19 are similar to one another. Two of the five magnetic transitions of the n = 6 member overlap with two of the transitions in the n = 4 member, with the remaining three transitions overlapping with the transitions found in the n = 5 member, most likely due to their related local environments.2628 The n = ∞ member, CeCoGe3, is a Kondo antiferromagnet (TK ∼ 100 K) with three magnetic transitions at T = 21 K, T = 12 K, and T = 8 K35 that resemble the anisotropy of the other n members. The presence of CeCoGe3 in this series is particularly exciting, as the noncentrosymmetric superconductor has recently been theoretically predicted to be topologically nontrivial.3638

Topological phases containing local inversion symmetry breaking such as UTe239,40 and CeRh2As241,42 have gathered much attention in recent years. As such, the CeCoGe3 sublattice in Cen+1ConGe3n+1 creates a unique opportunity to understand the structure–property relationships of local inversion symmetry breaking on the bulk properties of a material as a function of n.43,44 With evidence of highly correlated phenomena, complex magnetism, and the potential for topological behavior, the Cen+1ConGe3n+1 family of compounds are candidates to study new states of matter, such as Weyl–Kondo semimetals,7,45 through the addition of CeCoGe3 layers. The potential of the An+1BnX3n+1 series is not limited to the Ce analogues, as the physical properties of Pr based intermetallic compounds have garnered increasing interest over the past three decades.

3. Prn+1ConGe3n+1

Ce and Pr behave fundamentally differently due to their respective 4f1 and 4f2 electronic configurations. For example, CeAlGe is an incommensurate antiferromagnetic type-I Weyl semimetal,46 whereas PrAlGe exhibits type-II Weyl semimetal behavior with an anomalous Hall conductivity of ∼680 Ω–1 cm–1 due to a ferromagnetic spin–glass-like ordering at 15 K.47,48 Observations of Weyl fermions through angle resolved photoelectron spectroscopy measurements and the intrinsic anomalous Hall response confirmed these differences in physical properties.49,50 Additionally, Si substitution in PrAlGe1–xSix (0 < x < 1) has shown tunable anomalous Hall effects, demonstrating the interplay between scattering mechanism and topological state. Magnetoresistance sensitivity was also shown to increase by as much as 100% from PrAlSi to PrAlGe.51

The 4f electronic ground states are responsible for differences in physical properties through changes in their interaction with conduction electrons. While Ce3+ ions follow Kramer’s theorem, Pr3+ ions have an even number of 4f electrons and are classified as non-Kramer’s ions. Here, the doublet ground state is no longer protected by time reversal symmetry, resulting in many Pr compounds adopting a singlet ground state and, thus, having neither heavy electron behavior nor magnetic ordering; for example, PrCoGe3 is non-Kondo with no magnetic ordering.52 However, the point group symmetry of the local environment of non-Kramer’s ions affords another path to achieving a doublet or quasi-doublet ground state through crystal electric field (CEF) splitting, inspiring the study of quantum fluctuations in spin systems and exotic quadrupolar phenomena.53 Crystal field schemes of Pr atoms in cubic environments have been the primary focus of the highly correlated behavior in Pr systems due to their predilection toward doublet ground states. The CEF stabilized non-Kramer’s doublet ground state results in a two-channel Kondo effect through electric quadrupolar ordering.54 Electric quadrupolar ordering has led to identification of new quantum critical points,55 non-Fermi liquid behavior,56 and unconventional superconductivity.57 PrInAg258,59 was the first heavy electron Pr compound discovered with the second largest Sommerfeld coefficient identified (γ ∼ 6.5 J/mol K2). Later, a large family of heavy superconducting compounds, PrM2X20 (M = Ti and V with X = Al;56 M = Ir and Rh with X = Zn;60,61 M = Ni with X = Cd62), was identified where the different M and X could be varied to tune the superconducting transition temperatures, quadrupolar order, and magnetic order.

Nonetheless, compounds with lower Pr point symmetry have recently shown that heavy electron behavior of a similar magnitude can be observed in quasi-doublet states. In PrSi,63 ferromagnetic ordering is observed despite a singlet ground state, due to quasi-doublet formation via a singlet ground state to singlet first excited state transition with a difference in energy of ΔT = 9 K. Low-lying crystal field excitations with ΔT = 12 K were also used to justify the elevated Sommerfeld coefficient of Pr2Rh3Ge5, γ ∼ 81 mJ/mol K2.64 The singlet–singlet quasi-doublets (ΔT = 63.9 K) in PrPdAl65 were recently found to result in a raised Sommerfeld coefficient of γ ∼ 940 mJ/mol K2. With these examples and others, such as Pr3Rh4Sn13 (γ ∼ 761 mJ/mol K2)66 and Pr5Ir4Sn10 (γ ∼ 500 mJ/mol K2),67,68 there is mounting interest in Pr based intermetallic compounds with noncubic point symmetries and highly correlated behavior that, when coupled with the modularity of a homologous series, supports investigation of the Prn+1ConGe3n+1 series.

In the Prn+1ConGe3n+1 series, n = 1, 2, and ∞ members have been reported. PrCo1–xGe2, n = 1, has one antiferromagnetic transition at 5 K and a potential metamagnetic transition in the low-field magnetization.69 For Pr3Co2+xGe7,70n = 2, magnetic susceptibility data at 0.1 T shows two antiferromagnetic transitions at 5.3 and 9.3 K and three metamagnetic transitions below an applied field of 3 T. These transitions seem to be suppressed with the introduction of Sn in Pr3Co2+xGe7–ySny,70 with only two metamagnetic transitions occurring at applied fields of <1 T. Both phases are far below the expected magnetization for Pr, approaching 1 μB/mol-Pr at 7 T and 4 K; however, the effective magnetic moment for Pr3Co2+xGe7–ySnyeff = 3.81 μB/Pr) is above the theoretical spin only moment of Pr3+ (3.58 μB/Pr3+). For comparison, Pr3Ni2–xSn771 has similar behavior, with a single metamagnetic transition occurring below 2 T and 2 K. With the identification of two newly synthesized members of the Prn+1ConGe3n+1 series, n = 3 and n = 4, we now have a platform to demonstrate the tunability of this series for future work (Figure 1).

Figure 1.

Figure 1

(a) Aggregate crystals containing Pr5Co4+xGe13–ySny before etching and (b) a fragment after etching. (c) Plates of Pr4Co3+xGe10–ySny.

The n = 3 member Pr4Co3+xGe10–ySny (see Figure 2 for structural representation and Table S1 for crystallographic data) is isostructural to Eu2Ni2–xSn5 (Eu4Ni4–xSn10, Cmcm).72 The atomic positions for both structures are shown in Table S2. Pr4Co3+xGe10–ySny can be described as a stacking of CeNiSi2 alternating with two slabs of BaNiSn3/AuCu3. The main differences between the Eu4Ni4–xSn10 and Pr4Co3+xGe10–ySny structures are (1) the complexity of the disorder and (2) the relative occupancy of the transition metals in the disordered components. In Eu4Ni4–xSn10, the disorder is composed of two components: AuCu3 and BaNiSn3 in one orientation. In contrast, the Pr4Co3+xGe10–ySny structure is composed of three disordered components, which are AuCu3 and two orientations of BaNiSn3 (right side up and down relative to the b axis).

Figure 2.

Figure 2

Crystal structure of Pr4Co3+xGe10–ySny, where yellow, blue, green, and gray spheres represent Pr, Co, Ge, and Sn, respectively.

Pr5Co4+xGe13–ySny (Figure 3 and Tables S1 and S3) of the Ce5Co4Ge13 structure type25 can be described as a Co incorporated Nb5Ga13-type structure composed of AlB2, BaNiSn3, and BaNiSn3/AuCu3 subunits. The use of AlB2 subunits is analogous to CeNiSi2, resulting in a stacking pattern of three BaNiSn3/AuCu3 subunits separated by one CeNiSi2 subunit. In the report of Ce5Co4Ge13, a disordered model with BaNiSn3/AuCu3 subunits was suggested, as a logical explanation of the observed residual electron densities. This disorder is also observed in the published structure of Ce5Co4+xGe13–ySny26 and in the model of Pr5Co4+xGe13–ySny presented in this manuscript. Like Pr4Co3+xGe10–ySny, there are also reduced Co occupancies relative to the Ge occupancies within the disordered components.

Figure 3.

Figure 3

Crystal structure of Pr5Co4+xGe13–ySny, where yellow, blue, green, and gray spheres represent Pr, Co, Ge, and Sn, respectively.

4. Challenges and Future Outlook

While single crystal diffraction provides great insights into the crystal structure, it is imperative to know if the single crystal is truly representative of the entire batch of crystals grown. With powder diffraction, the experiment can confirm the purity and homogeneity of the crystalline sample. For Pr4Co3+xGe10–ySny, while the structural models obtained through single crystal X-ray diffraction are in good agreement with measured intensities, significant differences arise when compared to the bulk sample, revealing the challenges of using coherent scattering methods to model disorder. The local environment of the Pr1 site has, based on single crystal models, five potential configurations over a length of approximately 5 Å which extends diffraction techniques to the limits of their capabilities when considering the similar X-ray scattering cross sections of Co and Ge, the similar contact distances seen in the literature from Co–Co, Co–Ge, and Ge–Ge contacts depending on the covalency, and the similar bonding behaviors of Sn and Ge. Thus, despite the acceptable refinement statistics and reasonable crystal chemistry of the Pr4Co3+xGe10–ySny model obtained from single crystal diffraction, the model must be refined further. While in many circumstances single crystal X-ray and powder diffraction are satisfactory to obtain exceptional models of disordered crystal structures,73 for the future of quantum materials, drawing attention to the use of complementary structure techniques for their characterization is extremely important. The electronic contributions of minor structural defects can result in significant differences in physical properties. The structural investigations of heavy fermion compound Ce2RhIn8 are a model learning tool in how complementary structure techniques come together.74 Incongruencies in Bragg peak intensities observed in X-ray powder diffraction inspired electron diffraction experiments ultimately leading to high-resolution synchrotron X-ray reciprocal space mapping of single crystals and high-resolution neutron powder diffraction experiments that identified unexpected two-phase mosaic disorder in a crystal structure otherwise believed to be ordered. The identification of complex disorder in Ce2RhIn8 went on to be used to rationalize the physical properties of other related heavy electron materials7578 and inspire more detailed synthetic investigations,79 again underscoring the importance of thorough structural characterization.

A similar approach could be used for the Prn+1ConGe3n+1 homologous series to clarify the number of apparent disorder configurations in the BaNiSn3/AuCu3-type subunits through short-range correlations, address the peak broadening of the (2 0 0) and (0 0 2) reflections, and identify the cause of non-Bragg scattering at the 10° to 11° 2θ range using long-range correlations. Techniques such as pair distribution function, neutron diffraction methods, electron microscopy, and electron diffraction methods provide alternative views of a material that, taken together, work to elucidate the structure of compounds from local disorder at specific atomic sites to macroscopic disorder such as stacking faults and mosaicity effects.

We have successfully synthesized two members of the Prn+1ConGe3n+1 homologous series: Pr4Co3+xGe10–ySny and Pr5Co4+xGe13–ySny. However, phase pure synthesis of a single homologous series member remains a significant challenge. Due to the similar stoichiometries based on the relative elemental ratios of the Prn+1ConGe3n+1 series members, altering the reactant ratios proved ineffective. As such, our remaining option has been to refine the temperature profile, but without comprehensive ternary or quaternary phase diagrams, there is little guidance for this approach. Unexpectedly, raising the Sn concentration of the reaction has been the most successful method of obtaining phase pure homologous series members, despite producing the byproduct Pr2Co3Ge5. This is similar to recent findings regarding Y4RuGe8.80 The possibility of altering the flux concentration to isolate different n members presents a promising new method of reaction tuning that has inspired an investigation into in situ X-ray powder diffraction methods8183 and thermal analysis84,85 to understand how changing flux conditions influence the formation of homologous series members and other related phases.86,87

Despite these structural challenges, the Prn+1ConGe3n+1 series presents new opportunities for 4f2 electron systems. Uncovering potential correlative effects between different crystallographic Pr sites, dimensionality, and local environment could impact local 4f hybridization with conduction electrons. Thus, there is strong motivation to measure the physical properties of the Prn+1ConGe3n+1 homologous series. Due diligence in developing new structural models of quantum materials is equally as important to the quantum community as their physical properties. We keep these considerations in mind for the synthesis of the Prn+1ConGe3n+1 homologous series, as understanding the Pr local environment is extremely important to relating crystal electric field to electronic and magnetic behavior at the intersection of topological materials and highly correlated electron systems.

5. Methods

5.1. Synthesis

Pr4Co3+xGe10–ySny and Pr5Co4+xGe13–ySny were prepared using the flux growth method with tin as a metallic flux. The elements were combined in the ratio 3 Pr:2 Co:7 Ge:50 Sn and 3 Pr:2 Co:7 Ge:20 Sn, respectively, with the weight of the Pr4Co3+xGe10–ySny reactants being 0.50 g, Sn flux 2.80 g Pr5Co4+xGe13–ySny reactants’ weight 0.25 g, and Sn flux 0.56 g. The elements were transferred to an alumina Canfield crucible88 and sealed in fused silica tubes under ∼1/3 atm of Ar gas. The Pr5Co4+xGe13–ySny ampule was placed in a programmable furnace at 300 °C and heated to 1175 °C at the rate 100 °C/h. The sample dwelled at 1175 °C for 24 h and was cooled to 815 °C at the rate 3 °C/h, where it dwelled for approximately 6 h before removal and centrifugation of excess Sn flux. Residual flux was removed via etching in dilute HCl. A layered aggregate of plates, approximately 2 mm × 4 mm × 7 mm in size, composed of Pr3Co2+xGe7–ySny, Pr4Co3+xGe10–ySny, and Pr5Co4+xGe13–ySny was obtained. The Pr4Co3+xGe10–ySny reaction followed similar parameters with a 4 h dwell time at 1175 °C and a 30 h dwell time at 815 °C. One large platelike crystal of Pr4Co3+xGe10–ySny, 1 mm × 3 mm × 4 mm in size, and some polyhedral shaped crystals of Pr2Co3Ge5, approximately 1 mm × 2 mm × 2 mm in size, were obtained. Single crystals of Pr4Co3+xGe10–ySny and Pr5Co4+xGe13–ySny were cut from their respective samples. The Pr4Co3+xGe10–ySny method was reproduced with a modified temperature profile for comparison to the synthesis of Pr5Co4+xGe13–ySny, where the dwell time at 1175 °C was increased to 24 h, and the sample was removed from the furnace upon reaching 815 °C with less than 1 h of dwell time. From this sample, Pr2Co3Ge5 was obtained as a dominant product. Small platelike crystals below 1 mm × 1 mm × 1 mm in size were also observed.

5.2. Laboratory Powder X-ray Diffraction

Powder X-ray diffraction data was collected on an aggregate containing Pr3Co2+xGe7–ySny, Pr4Co3+xGe10–ySny, and Pr5Co4+xGe13–ySny using a Bruker D8 Advance diffractometer operating at 40 kV/30 mA with Cu Kα radiation (λ = 1.54184 Å) equipped with a LYNXEYE detector. Data was collected in the 2θ range from 5.0° to 90°. Laboratory powder X-ray diffraction was performed on the aggregate sample that was used to collect Pr5Co4+xGe13–ySny single crystal data, and it was determined that the sample contained an intergrown mixture of series members; however, the presence of more than a single n member made accurate phase identification extremely difficult due to overlapping Bragg reflections and air scattering present in the laboratory powder diffraction. To overcome this problem, high-resolution powder diffraction data was collected at beamline 11-BM at the Advanced Photon Source of Argonne National Laboratory, where a combination of using a transmission mode, analyzer crystals, and high photon flux resulted in far higher signal-to-noise ratios.

5.3. High-Resolution Powder X-ray Diffraction

High-resolution powder X-ray diffraction data (λ = 0.458977 Å) was collected in the 2θ range from 0.5° to 43.5° for Pr5Co4+xGe13–ySny and from 0.5° to 48.8° for Pr4Co3+xGe10–ySny at the 11-BM beamline at the Advanced Photon Source of Argonne National Laboratory. All data modeling was carried out using TOPAS-Academic software. A two-phase Rietveld refinement using data collected on a ground fragment of the single phase plate of Pr4Co3+xGe10–ySny and Sn (Figure 4) was able to account for all visible Bragg reflections, showing the sample is phase pure. However, mismatched intensities and non-Bragg scattering reveal a more complex story than captured through single crystal methods. The intensities of the (0 2 0), (1 7 1), (1 11 0), and (0 14 0) reflections are different than the calculated model. Given the platelike morphology of crystals grown from this series and the large anisotropy of the lattice parameters, intensity corrections were attempted by considering the preferred orientation of different combinations of the following reflection families: (0 2n 0), (1 2n–1 0), (0 2n–1 1), and (1 2n–1 1), but the improvements on the refinement were insignificant. Anisotropic strain broadening was also evaluated with little improvement to the refinement. As all mismatched intensities have dependence on the b crystallographic direction, attempts were made to alter the disorder model obtained through single crystal diffraction. While the intensities were improved considerably, the resulting structural model disagreed with the compositions obtained from elemental analysis and single crystal models.

Figure 4.

Figure 4

High-resolution synchrotron X-ray powder diffraction collected at 11-BM at the Advanced Photon Source of Pr4Co3+xGe10–ySny shown in blue (λ = 0.458977 Å). The calculated fit from Rietveld refinement is shown in red, with the difference plot shown in black. Green and purple tick marks indicate Pr4Co3+xGe10–ySny and Sn reflections, respectively. Labeled reflections are those with mismatched intensities.

One factor that could be impacting the accuracy of the single crystal structure refinement is the lack of visibility for the (0 2 0) reflection. Due to the large b axis of Pr4Co3+xGe10–ySny, the (0 2 0) reflection appears at a low 2θ angle of 4.56° (Mo Kα, λ = 0.71073 Å). This is outside the 2θ collection range 10.06° to 59.74° due to shadowing from the beam stop either fully or partially blocking this reflection. Since the disorder is strongly correlated to the (0 2n 0) reflections, the (0 2 0) reflection is particularly sensitive to changes in the structural model. Without this reflection, the single crystal model lacks key intensity information against which to refine the structural model. This is an additional challenge when working with compounds with very large unit cells, especially when blocked reflections are critical to regions of disorder. Due to the mixed phase nature of the sample containing Pr5Co4+xGe13–ySny, a Pawley refinement was performed. From this data, low-angle (0 2n 0) reflections were visible (Figure 5), which were critical in identifying the members of the Prn+1ConGe3n+1 series due to their strong correlation with the systematically elongated b axis of the series. Yet, without a single phase sample we were limited in our ability to refine the crystal structure against the bulk powder diffraction pattern through Rietveld refinement. Instead, a Pawley refinement revealed a mixed phase system of n = 2, 3, and 4 (Table 1). Although the lattice parameters obtained for Pr3Co2+xGe7–ySny and Pr5Co4+xGe13–ySny agree with single crystal data,70 overlapping Bragg reflections remain a challenge with clear deviations in the refined a and c lattice parameters of Pr4Co3+xGe10–ySny from single crystal data and phase pure powder diffraction. From here, obtaining phase pure samples will be the only way to validate the homogeneity of the disorder observed in the single crystal model for Pr5Co4+xGe13–ySny.

Figure 5.

Figure 5

Comparison of laboratory (black, λ = 1.5406 Å) and high-resolution synchrotron (blue, λ = 0.458977 Å) X-ray powder diffraction data taken at 11-BM at the Advanced Photon Source. Colored lines and shapes indicate the (0 2n 0) reflections used to identify members of the homologous series Pr3Co2+xGe7–ySny (blue triangle), Pr4Co3+xGe10–ySny (yellow diamond), and Pr5Co4+xGe13–ySny (red pentagon).

Table 1. Lattice Parameters Obtained from the Pawley Refinement of Pr5Co4+xGe13–ySny.

  Space Group a (Å) b (Å) c (Å) V3)
Pr3Co2+xGe7–ySny Cmmm 4.25340(3) 26.0705(20) 4.29152(3) 475.881(7)
Pr4Co3+xGe10–ySny Cmcm 4.36710(14) 35.6931(63) 4.20920(14) 656.113(34)
Pr5Co4+xGe13–ySny Cmmm 4.28048(5) 45.4234(67) 4.30056(7) 836.175(21)

5.4. Single Crystal X-ray Diffraction

Single crystal X-ray diffraction data was obtained at room temperature using a Bruker D8 Quest Kappa single crystal X-ray diffractometer, equipped with an IμS microfocus source (Mo Kα, λ = 0.71073 Å), HELIOS optics monochromator, and PHOTON II CPAD detector. The Bruker SAINT program was used to integrate the diffraction data, and the intensities were corrected for absorption via a multiscan method implemented in SADABS 2016/2.89 Preliminary starting models were generated using the intrinsic phasing methods in SHELXT90 and subsequently anisotropically refined (full-matrix least-squares on F2) using SHELXL.91 Details of both structure refinements can be found in Tables S1–S3. Single crystals of Pr4Co3+xGe10–ySny (0.01 mm × 0.01 mm × 0.02 mm) and Pr5Co4+xGe13–ySny (0.02 mm × 0.12 mm × 0.14 mm) were cut from their respective samples and mounted on glass fibers using two-part epoxy.

The crystal structure of Pr4Co3+xGe10–ySny was preliminarily modeled in the noncentrosymmetric space group Amm2. However, additional symmetries were detected using the ADDSYM function in PLATON, resulting in the final selection of the centrosymmetric space group Cmcm, in agreement with the Eu4Co4–xSn10 structure type. In contrast to the Eu4Co4–xSn10 structure type, additional electron densities of 5.43 e3 and 4.53 e3 were observed in the Pr1 local environment at (0 0.6408 0) and (0 0.7082 0), respectively. To address the observed intensities, an additional Co1B–Ge1B dimer was added. In accordance with the elemental analysis results indicating the presence of Sn incorporation, the Sn1 site was considered, leading to marginally improved refinement statistics relative to the assignment of a Ge atom. Additionally, when compared to the lattice parameters of Pr4Co3+xGe10 grown in In flux (a = 4.2831(7) Å, b = 35.497(8) Å, c = 4.2533(10) Å), there is a difference in the b lattice parameter of approximately 0.3 Å corresponding to the incorporation of Sn. This effect was observed in Pr3Co2+xGe7–ySny and Pr3Co2+xGe7 as well. The Sn incorporation in the presence of Sn flux agrees with observations from Cen+1ConGe3n+1–ySny (n = 4, 5, and 6). At this point in the refinement, Co vacancies had not been considered, but a hole of −5.46 e3 was observed on the Co1A site. Thus, the Co1A, Co1B, and Co2 site occupancies were allowed to refine freely, resulting in a reduction in occupancy from ∼66% to ∼58% for Co1A, from ∼7.8% to ∼5.0% for Co1B, and from unity to 99% for Co2. The corresponding Ge sites’ occupancy saw an increase from ∼66% to ∼67% for Ge1A and from ∼7.8% to ∼8.4% for Ge1B. The fully occupied Co2 site and vacancies on the Co1A and Co1B sites agree with the crystal structure determination of Eu4Ni4–xSn10.

The crystal structure of Pr5Co4+xGe13–ySny was found to crystallize in the orthorhombic space group Cmmm. The disorder modeled in this structure is in extremely good agreement with the refinement of Ce5Co4+xGe13–ySny, with the Pr1 local environment consisting of one Co2–Ge4 dimer representing the BaNiSn3 structure type complemented by the Sn1 site representing the AuCu3 structure type and the Pr3 local environment consisting of two complementary Co3–Ge7 dimers with the Sn2 site occupying the remainder of the configurations. Unlike Ce5Co4+xGe13–ySny, upon refining the weighting scheme, a negative electron density of −6.00 e3 was observed on the Co3 site in the Pr3 local environment. To address this, the Co3 occupancy was allowed to refine freely, where its occupancy reduced from ∼35% to ∼28%. From here, the occupancies of Co1 of the Pr2 local environment and Co2 of the Pr1 local environment were refined freely, resulting in an occupancy of approximately 99% for Co1 and a reduction the Co2 occupancy from ∼87% to ∼83%. The Co1 occupancy was fixed to unity, but the Co2 and Co3 vacancies were retained. The freely refined Co2 and Co3 sites resulted in an occupancy increase from 87% to 88% for Ge4 and from 35% to 37% for Ge7.

5.5. Elemental Analysis

Single crystals of Pr4Co3Ge10–ySny and Pr5Co4Ge13–ySny were analyzed using energy dispersive (X-ray) spectroscopy on a VERSA 3D focused ion beam scanning electron microscope with an acceleration voltage of 20 kV. The weight percentages obtained from EDS (normalized to Pr) resulted in the atomic formulas Pr4.00(2)Co3.22(1)Ge9.34(7)Sn0.66(1) and Pr5.00(3)Co4.38(1)Ge12.27(5)Sn0.44(2), which are comparable to the single crystal refinement formulas Pr4Co3.25(3)Ge9.50(1)Sn0.469(4) and Pr5Co4.22(2)Ge12.48(1)Sn0.519(7), respectively.

Acknowledgments

This material is based upon work supported by the U.S. Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists, Office of Science Graduate Student Research (SCGSR) program. The SCGSR program is administered by the Oak Ridge Institute for Science and Education for the DOE under contract number DE–SC0014664. J.Y.C. gratefully acknowledges the support of the National Science Foundation (Grant No. DMR-2209804), the Welch Foundation (Grant No. AT-2056-20210327), and the Department of Energy Grant (No. DE-SC0022068). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science user facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.2c02152.

  • Additional details on the disorder model of Pr4Co3+xGe10–ySny and Pr5Co4+xGe13–ySny, the full structure refinement details of Pr4Co3+xGe10–ySny and Pr5Co4+xGe13–ySny (Table S1), a comparison of atomic positions for Pr4Co3+xGe10–ySny and Eu4Ni4–xSn10 (Table S2), and a comparison of atomic positions for Pr5Co4Ge13–ySny and Ce5Co4+xGe13–ySny (Table S3) (PDF)

  • CIF data for Pr4Co3+xGe10–ySny (CIF)

  • CIF data for Pr5Co4+xGe13–ySny(CIF)

The authors declare no competing financial interest.

Supplementary Material

ao2c02152_si_001.pdf (163.1KB, pdf)
ao2c02152_si_002.cif (347.6KB, cif)
ao2c02152_si_003.cif (519.9KB, cif)

References

  1. The Rise of Quantum Materials. Nat. Phys. 2016, 12, 105–105. 10.1038/nphys3668 [DOI] [Google Scholar]
  2. Ball P. Quantum Materials: Where Many Paths Meet. MRS Bull. 2017, 42, 698–705. 10.1557/mrs.2017.220. [DOI] [Google Scholar]
  3. Broholm C.; Fisher I.; Moore J.; Murnane M.; Moreo A.; Tranquada J.; Basov D.; Freericks J.; Aronson M.; MacDonald A.; Fradkin E.; Yacoby A.; Samarth N.; Stemmer S.; Horton L.; Horwitz J.; Davenport J.; Graf M.; Krause J.; Pechan M.; Perry K.; Rhyne J.; Schwartz A.; Thiyagarajan T.; Yarris L.; Runkles K.. Basic Research Needs Workshop on Quantum Materials for Energy Relevant Technology; USDOE Office of Science (SC), Basic Energy Sciences (BES): United States, 2016-02-10, 2016.
  4. Moore J. E. The Birth of Topological Insulators. Nature 2010, 464, 194–198. 10.1038/nature08916. [DOI] [PubMed] [Google Scholar]
  5. Giustino F.; Lee J. H.; Trier F.; Bibes M.; Winter S. M.; Valentí R.; Son Y.-W.; Taillefer L.; Heil C.; Figueroa A. I.; Plaçais B.; Wu Q.; Yazyev O. V.; Bakkers E. P. A. M.; Nygård J.; Forn-Díaz P.; De Franceschi S.; McIver J. W.; Torres L. E. F. F.; Low T.; Kumar A.; Galceran R.; Valenzuela S. O.; Costache M. V.; Manchon A.; Kim E.-A.; Schleder G. R.; Fazzio A.; Roche S. The 2021 Quantum Materials Roadmap. J. Phys. Mater. 2020, 3, 042006. 10.1088/2515-7639/abb74e. [DOI] [Google Scholar]
  6. Yan B.; Felser C. Topological Materials: Weyl Semimetals. Annu. Rev. Condens. Matter Phys. 2017, 8, 337–354. 10.1146/annurev-conmatphys-031016-025458. [DOI] [Google Scholar]
  7. Lai H.-H.; Grefe S. E.; Paschen S.; Si Q. Weyl-Kondo Semimetal in Heavy-Fermion Systems. Proc. Natl. Acad. Sci. U.S.A. 2018, 115, 93–97. 10.1073/pnas.1715851115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Tokura Y.; Kanazawa N. Magnetic Skyrmion Materials. Chem. Rev. 2021, 121, 2857–2897. 10.1021/acs.chemrev.0c00297. [DOI] [PubMed] [Google Scholar]
  9. Alfieri A.; Anantharaman S. B.; Zhang H.; Jariwala D. Nanomaterials for Quantum Information Science and Engineering. Adv. Mater. 2022, 2109621. 10.1002/adma.202109621. [DOI] [PubMed] [Google Scholar]
  10. Si Q.; Steglich F. Heavy Fermions and Quantum Phase Transitions. Science 2010, 329, 1161–1166. 10.1126/science.1191195. [DOI] [PubMed] [Google Scholar]
  11. Wirth S.; Steglich F. Exploring Heavy Fermions from Macroscopic to Microscopic Length Scales. Nat. Rev. Mater. 2016, 1, 16051. 10.1038/natrevmats.2016.51. [DOI] [Google Scholar]
  12. Grin Y. N.; Yarmolyuk Y. P.; Gladyshevskij E. I. Crystal structures of R2CoGa8 compounds (R = Sm, Gd, Tb, Dy, Ho, Er, Tm, Lu, Y) and RCoGa5 compounds (R = Gd, Tb, Dy, Ho, Er, Tm, Lu, Y). Kristallografiya 1979, 24, 242–246. [Google Scholar]
  13. Thompson J. D.; Movshovich R.; Fisk Z.; Bouquet F.; Curro N. J.; Fisher R. A.; Hammel P. C.; Hegger H.; Hundley M. F.; Jaime M.; Pagliuso P. G.; Petrovic C.; Phillips N. E.; Sarrao J. L. Superconductivity and Magnetism in a New Class of Heavy-Fermion Materials. J. Magn. Magn. Mater. 2001, 226–230, 5–10. 10.1016/S0304-8853(00)00602-8. [DOI] [Google Scholar]
  14. Latturner S. E.; Kanatzidis M. G. RE(AuAl2)nAl2(AuxSi1-x)2: A New Homologous Series of Quaternary Intermetallics Grown from Aluminum Flux. lnorg. Chem. 2008, 47, 2089–2097. 10.1021/ic701799z. [DOI] [PubMed] [Google Scholar]
  15. Tursina A.; Khamitcaeva E.; Gribanov A.; Gnida D.; Kaczorowski D. CePd2Al2, CePd3Al3, and CePd4Al4—A New Homologous Series Built of CaBe2Ge2- and CsCl-type Units. lnorg. Chem. 2015, 54, 3439–3445. 10.1021/ic503110c. [DOI] [PubMed] [Google Scholar]
  16. Guo S.-P.; You T.-S.; Bobev S. Closely Related Rare-Earth Metal Germanides RE2Li2Ge3 and RE3Li4Ge4 (RE = La-Nd, Sm): Synthesis, Crystal Chemistry, and Magnetic Properties. lnorg. Chem. 2012, 51, 3119–3129. 10.1021/ic202591j. [DOI] [PubMed] [Google Scholar]
  17. Guo S.-P.; You T.-S.; Jung Y.-H.; Bobev S. Synthesis, Crystal Chemistry, and Magnetic Properties of RE7Li8Ge10 and RE11Li12Ge16 (RE = La-Nd, Sm): New Members of the [REGe2]n[RELi2Ge]m Homologous Series. lnorg. Chem. 2012, 51, 6821–6829. 10.1021/ic300566x. [DOI] [PubMed] [Google Scholar]
  18. Lin X.; Tabassum D.; Mar A. Narrowing The Gap: From Semiconductor to Semimetal in The Homologous Series of Rare-Earth Zinc Arsenides RE2-yZn4As4·n(REAs) and Mn-Substituted Derivatives RE2-yMnxZn4-xAs4·n(REAs) (RE = La-Nd, Sm, Gd). Dalton Trans. 2015, 44, 20254–20264. 10.1039/C5DT03596A. [DOI] [PubMed] [Google Scholar]
  19. Mathur N. D.; Grosche F. M.; Julian S. R.; Walker I. R.; Freye D. M.; Haselwimmer R. K. W.; Lonzarich G. G. Magnetically Mediated Superconductivity in Heavy Fermion Compounds. Nature 1998, 394, 39–43. 10.1038/27838. [DOI] [Google Scholar]
  20. Park T.; Ronning F.; Yuan H. Q.; Salamon M. B.; Movshovich R.; Sarrao J. L.; Thompson J. D. Hidden Magnetism and Quantum Criticality in the Heavy Fermion Superconductor CeRhIn5. Nature 2006, 440, 65–68. 10.1038/nature04571. [DOI] [PubMed] [Google Scholar]
  21. Hegger H.; Petrovic C.; Moshopoulou E. G.; Hundley M. F.; Sarrao J. L.; Fisk Z.; Thompson J. D. Pressure-Induced Superconductivity in Quasi-2D CeRhIn5. Phys. Rev. Lett. 2000, 84, 4986–4989. 10.1103/PhysRevLett.84.4986. [DOI] [PubMed] [Google Scholar]
  22. Macaluso R. T.; Sarrao J. L.; Moreno N. O.; Pagliuso P. G.; Thompson J. D.; Fronczek F. R.; Hundley M. F.; Malinowski A.; Chan J. Y. Single-Crystal Growth of Ln2MIn8 (Ln = La, Ce; M = Rh, Ir): Implications for the Heavy-Fermion Ground State. Chem. Mater. 2003, 15, 1394–1398. 10.1021/cm021063e. [DOI] [Google Scholar]
  23. Weiland A.; Felder J. B.; McCandless G. T.; Chan J. Y. One Ce, Two Ce, Three Ce, Four? An Intermetallic Homologous Series to Explore: An+1BnX3n+1. Chem. Mater. 2020, 32, 1575–1580. 10.1021/acs.chemmater.9b04743. [DOI] [Google Scholar]
  24. Méot-Meyer M.; Venturini G.; Malaman B.; Roques B. De Nouveaux Isotypes Lacunaires de CeNiSi2: Les Germaniures RCoxGe2, R = Y, La-Sm, Gd-Lu, 0 < x ≤ 1. Mater. Res. Bull. 1985, 20, 1515–1521. 10.1016/0025-5408(85)90169-2. [DOI] [Google Scholar]
  25. Belan B.; Stȩpeń-damm J.; Gladyshevskii R.; Bodak O. The New Structure Type Ce5Co4Ge13. Chem. Met. Alloys 2008, 1, 43–48. 10.30970/cma1.0035. [DOI] [Google Scholar]
  26. Weiland A.; Wei K.; McCandless G. T.; Baumbach R. E.; Chan J. Y. Fantastic n = 4: Ce5Co4+xGe13-ySny of the An+1MnX3n+1 Homologous Series. J. Chem. Phys. 2021, 154, 114707. 10.1063/5.0045015. [DOI] [PubMed] [Google Scholar]
  27. Felder J. B.; Weiland A.; Hodovanets H.; McCandless G. T.; Estrada T. G.; Martin T. J.; Walker A. V.; Paglione J.; Chan J. Y. Law and Disorder: Special Stacking Units—Building the Intergrowth Ce6Co5Ge16. lnorg. Chem. 2019, 58, 6037–6043. 10.1021/acs.inorgchem.9b00350. [DOI] [PubMed] [Google Scholar]
  28. Weiland A.; Wei K.; McCandless G. T.; Felder J. B.; Eddy L. J.; Baumbach R. E.; Chan J. Y. Strongly Correlated Electron Behavior In a New Member of The An+1BnX3n+1 Homologous Series: Ce7Co6Ge19. Phys. Rev. Mater. 2020, 4, 074408. 10.1103/PhysRevMaterials.4.074408. [DOI] [Google Scholar]
  29. Venturini G.; Méot-Meyer M.; Malaman B.; Roques B. De Nouvelles Séries de Germaniures, Isotypes de Yb3Rh4Sn13 et BaNiSn3, dans les Systèmes Ternaires TR T Ge où TR est un Elément des Terres Rares et T = Co, Rh, Ir, Ru, Os. J. Less-Common Met. 1985, 113, 197–204. 10.1016/0022-5088(85)90277-2. [DOI] [Google Scholar]
  30. Weng Z.; Smidman M.; Jiao L.; Lu X.; Yuan H. Q. Multiple Quantum Phase Transitions and Superconductivity in Ce-Based Heavy Fermions. Rep. Prog. Phys. 2016, 79, 094503. 10.1088/0034-4885/79/9/094503. [DOI] [PubMed] [Google Scholar]
  31. Doniach S. The Kondo Lattice and Weak Antiferromagnetism. Physica B+C 1977, 91, 231–234. 10.1016/0378-4363(77)90190-5. [DOI] [Google Scholar]
  32. Pecharsky V. K.; Gschneidner K. A. CeCo0.89Ge2: A Heavy-Fermion System. Phys. Rev. B 1991, 43, 8238–8244. 10.1103/PhysRevB.43.8238. [DOI] [PubMed] [Google Scholar]
  33. Mun E. D.; Lee B. K.; Kwon Y. S.; Jung M. H. Kondo Ground State of CeCoGe2 With j = 5/2. Phys. Rev. B 2004, 69, 085113. 10.1103/PhysRevB.69.085113. [DOI] [Google Scholar]
  34. Ramachandran B.; Chang P. C.; Kuo Y. K.; Lue C. S. Electrical and Thermal Transport Properties of Intermetallic RCoGe2 (R= Ce and La) Compounds. J. Phys.: Condens. Matter 2014, 26, 255601. 10.1088/0953-8984/26/25/255601. [DOI] [PubMed] [Google Scholar]
  35. Thamizhavel A.; Takeuchi T.; D Matsuda T.; Haga Y.; Sugiyama K.; Settai R.; O̅nuki Y. Unique Magnetic Phases in an Antiferromagnet CeCoGe3. J. Phys. Soc. Jpn. 2005, 74, 1858–1864. 10.1143/JPSJ.74.1858. [DOI] [Google Scholar]
  36. Settai R.; Okuda Y.; Sugitani I.; O̅nuki Y.; Matsuda T. D.; Haga Y.; Harima H. Non-Centrosymmetric Heavy Fermion Superconductivity in CeCoGe3. Int. J. Mod. Phys. B 2007, 21, 3238–3245. 10.1142/S0217979207044287. [DOI] [Google Scholar]
  37. Frey N. C.; Horton M. K.; Munro J. M.; Griffin S. M.; Persson K. A.; Shenoy V. B. High-Throughput Search for Magnetic and Topological Order in Transition Metal Oxides. Sci. Adv. 2020, 6, eabd1076 10.1126/sciadv.abd1076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Xu Y.; Elcoro L.; Song Z.-D.; Wieder B. J.; Vergniory M. G.; Regnault N.; Chen Y.; Felser C.; Bernevig B. A. High-Throughput Calculations of Magnetic Topological Materials. Nature 2020, 586, 702–707. 10.1038/s41586-020-2837-0. [DOI] [PubMed] [Google Scholar]
  39. Ran S.; Eckberg C.; Ding Q.-P.; Furukawa Y.; Metz T.; Saha S. R.; Liu I.-L.; Zic M.; Kim H.; Paglione J.; Butch N. P. Nearly Ferromagnetic Spin-Triplet Superconductivity. Science 2019, 365, 684–687. 10.1126/science.aav8645. [DOI] [PubMed] [Google Scholar]
  40. Ishizuka J.; Sumita S.; Daido A.; Yanase Y. Insulator-Metal Transition and Topological Superconductivity in UTe2 from a First-Principles Calculation. Phys. Rev. Lett. 2019, 123, 217001. 10.1103/PhysRevLett.123.217001. [DOI] [PubMed] [Google Scholar]
  41. Khim S.; Landaeta J. F.; Banda J.; Bannor N.; Brando M.; Brydon P. M. R.; Hafner D.; Küchler R.; Cardoso-Gil R.; Stockert U.; Mackenzie A. P.; Agterberg D. F.; Geibel C.; Hassinger E. Field-induced Transition within the Superconducting Sate of CeRh2As2. Science 2021, 373, 1012–1016. 10.1126/science.abe7518. [DOI] [PubMed] [Google Scholar]
  42. Nogaki K.; Daido A.; Ishizuka J.; Yanase Y. Topological Crystalline Superconductivity in Locally Noncentrosymmetric CeRh2As2. Phys. Rev. Res. 2021, 3, L032071. 10.1103/PhysRevResearch.3.L032071. [DOI] [PubMed] [Google Scholar]
  43. Fischer M. H.; Loder F.; Sigrist M. Superconductivity and Local Noncentrosymmetricity in Crystal Lattices. Phys. Rev. B 2011, 84, 184533. 10.1103/PhysRevB.84.184533. [DOI] [Google Scholar]
  44. Yoshida T.; Sigrist M.; Yanase Y. Topological Crystalline Superconductivity in Locally Noncentrosymmetric Multilayer Superconductors. Phys. Rev. Lett. 2015, 115, 027001. 10.1103/PhysRevLett.115.027001. [DOI] [PubMed] [Google Scholar]
  45. Grefe S. E.; Lai H.-H.; Paschen S.; Si Q. Weyl-Kondo Semimetals in Nonsymmorphic Systems. Phys. Rev. B 2020, 101, 075138. 10.1103/PhysRevB.101.075138. [DOI] [Google Scholar]
  46. Puphal P.; Pomjakushin V.; Kanazawa N.; Ukleev V.; Gawryluk D. J.; Ma J.; Naamneh M.; Plumb N. C.; Keller L.; Cubitt R.; Pomjakushina E.; White J. S. Topological Magnetic Phase in the Candidate Weyl Semimetal CeAlGe. Phys. Rev. Lett. 2020, 124, 017202. 10.1103/PhysRevLett.124.017202. [DOI] [PubMed] [Google Scholar]
  47. Puphal P.; Mielke C.; Kumar N.; Soh Y.; Shang T.; Medarde M.; White J. S.; Pomjakushina E. Bulk Single-Crystal Growth of The Theoretically Predicted Magnetic Weyl Semimetals RAlGe (R = Pr, Ce). Phys. Rev. Mater. 2019, 3, 024204. 10.1103/PhysRevMaterials.3.024204. [DOI] [Google Scholar]
  48. Destraz D.; Das L.; Tsirkin S. S.; Xu Y.; Neupert T.; Chang J.; Schilling A.; Grushin A. G.; Kohlbrecher J.; Keller L.; Puphal P.; Pomjakushina E.; White J. S. Magnetism and Anomalous Transport in the Weyl Semimetal PrAlGe: Possible Route to Axial Gauge Fields. npj Quantum Mater. 2020, 5, 5. 10.1038/s41535-019-0207-7. [DOI] [Google Scholar]
  49. Sanchez D. S.; Chang G.; Belopolski I.; Lu H.; Yin J.-X.; Alidoust N.; Xu X.; Cochran T. A.; Zhang X.; Bian Y.; Zhang S. S.; Liu Y.-Y.; Ma J.; Bian G.; Lin H.; Xu S.-Y.; Jia S.; Hasan M. Z. Observation of Weyl Fermions in a Magnetic Non-Centrosymmetric Crystal. Nat. Commun. 2020, 11, 3356. 10.1038/s41467-020-16879-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Meng B.; Wu H.; Qiu Y.; Wang C.; Liu Y.; Xia Z.; Yuan S.; Chang H.; Tian Z. Large Anomalous Hall Effect in Ferromagnetic Weyl Semimetal Candidate PrAlGe. APL Mater. 2019, 7, 051110. 10.1063/1.5090795. [DOI] [Google Scholar]
  51. Yang H.-Y.; Singh B.; Lu B.; Huang C.-Y.; Bahrami F.; Chiu W.-C.; Graf D.; Huang S.-M.; Wang B.; Lin H.; Torchinsky D.; Bansil A.; Tafti F. Transition From Intrinsic to Extrinsic Anomalous Hall Effect in the Ferromagnetic Weyl Semimetal PrAlGe1-xSix. APL Mater. 2020, 8, 011111. 10.1063/1.5132958. [DOI] [Google Scholar]
  52. Puphal P.; Krebber S.; Suard E.; Cubitt R.; Wang C.; Shang T.; Ukleev V.; White J. S.; Pomjakushina E. Development of Magnetism in The Solid Solution of Ce1-xPrxAlGe: From Magnetic Topology to Spin Glass. Phys. Rev. B 2020, 101, 214416. 10.1103/PhysRevB.101.214416. [DOI] [Google Scholar]
  53. Onimaru T.; Kusunose H. Exotic Quadrupolar Phenomena in Non-Kramers Doublet Systems — The Cases of PrT2Zn20 (T = Ir, Rh) and PrT2Al20 (T = V, Ti) —. J. Phys. Soc. Jpn. 2016, 85, 082002. 10.7566/JPSJ.85.082002. [DOI] [Google Scholar]
  54. Van Dyke J. S.; Zhang G.; Flint R. Field-Induced Ferrohastatic Phase in Cubic Non-Kramers Doublet Systems. Phys. Rev. B 2019, 100, 205122. 10.1103/PhysRevB.100.205122. [DOI] [Google Scholar]
  55. Matsubayashi K.; Tanaka T.; Sakai A.; Nakatsuji S.; Kubo Y.; Uwatoko Y. Pressure-Induced Heavy Fermion Superconductivity in the Nonmagnetic Quadrupolar System PrTi2Al20. Phys. Rev. Lett. 2012, 109, 187004. 10.1103/PhysRevLett.109.187004. [DOI] [PubMed] [Google Scholar]
  56. Sakai A.; Nakatsuji S. Kondo Effects and Multipolar Order in the Cubic PrTr2Al20 (Tr = Ti, V). J. Phys. Soc. Jpn. 2011, 80, 063701. 10.1143/JPSJ.80.063701. [DOI] [Google Scholar]
  57. Patri A. S.; Kim Y. B.. Unconventional Superconductivity arising from Multipolar Kondo Interactions. SciPost Phys. 2022, 12, 10.21468/SciPostPhys.12.2.057 [DOI] [Google Scholar]
  58. Yatskar A.; Beyermann W. P.; Movshovich R.; Canfield P. C.; Panchula A.; Bud’ko S. L. Unexpected Heavy-Electron Behavior at Low Temperatures in PrInAg2. Phys. B: Condens. Matter 1997, 230–232, 46–48. 10.1016/S0921-4526(97)81582-7. [DOI] [Google Scholar]
  59. Yatskar A.; Beyermann W. P.; Movshovich R.; Canfield P. C. Possible Correlated-Electron Behavior from Quadrupolar Fluctuations in PrInAg2. Phys. Rev. Lett. 1996, 77, 3637–3640. 10.1103/PhysRevLett.77.3637. [DOI] [PubMed] [Google Scholar]
  60. Onimaru T.; Matsumoto K. T.; Inoue Y. F.; Umeo K.; Sakakibara T.; Karaki Y.; Kubota M.; Takabatake T. Antiferroquadrupolar Ordering in a Pr-Based Superconductor PrIr2Zn20. Phys. Rev. Lett. 2011, 106, 177001. 10.1103/PhysRevLett.106.177001. [DOI] [PubMed] [Google Scholar]
  61. Onimaru T.; Nagasawa N.; Matsumoto K. T.; Wakiya K.; Umeo K.; Kittaka S.; Sakakibara T.; Matsushita Y.; Takabatake T. Simultaneous Superconducting and Antiferroquadrupolar Transitions in PrRh2Zn20. Phys. Rev. B 2012, 86, 184426. 10.1103/PhysRevB.86.184426. [DOI] [Google Scholar]
  62. Yanagisawa T.; Hidaka H.; Amitsuka H.; Nakamura S.; Awaji S.; Green E. L.; Zherlitsyn S.; Wosnitza J.; Yazici D.; White B. D.; Maple M. B. Quadrupolar Susceptibility and Magnetic Phase Diagram of PrNi2Cd20 with Non-Kramers Doublet Ground State. Philos. Mag. 2020, 100, 1268–1281. 10.1080/14786435.2019.1709912. [DOI] [Google Scholar]
  63. Das P. K.; Bhattacharyya A.; Kulkarni R.; Dhar S. K.; Thamizhavel A. Anisotropic Magnetic Properties and Giant Magnetocaloric Effect of Single-Crystal PrSi. Phys. Rev. B 2014, 89, 134418. 10.1103/PhysRevB.89.134418. [DOI] [Google Scholar]
  64. Anand V. K.; Hossain Z.; Geibel C. Magnetic Order in Pr2Pd3Ge5 and Possible Heavy-Fermion Behavior in Pr2Rh3Ge5. Phys. Rev. B 2008, 77, 184407. 10.1103/PhysRevB.77.184407. [DOI] [Google Scholar]
  65. Wang Z.; Zhao H.; Lyu M.; Xiang J.; Isikawa Y.; Zhang S.; Sun P. Frustrated Antiferromagnetism and Heavy-Fermion-Like Behavior in PrPdAl. Phys. Rev. B 2022, 105, 125113. 10.1103/PhysRevB.105.125113. [DOI] [Google Scholar]
  66. Nair H. S.; Ogunbunmi M. O.; Ghosh S. K.; Adroja D. T.; Koza M. M.; Guidi T.; Strydom A. M. Absence of a Long-Range Ordered Magnetic Ground State in Pr3Rh4Sn13 Studied Through Specific Heat and Inelastic Neutron Scattering. J. Phys.: Condens. Matter 2018, 30, 145601. 10.1088/1361-648X/aab1e5. [DOI] [PubMed] [Google Scholar]
  67. Patil N. G.; Ramakrishnan S. Magnetism In The R5T4Sn10 (R = Ce, Pr, and Nd; T = Rh and Ir) system. Phys. Rev. B 1997, 56, 3360–3371. 10.1103/PhysRevB.56.3360. [DOI] [Google Scholar]
  68. Ramakrishnan S.; Mydosh J. A. Multiple Phase Transitions in Rare Earth R5T4M10 (T = Rh, Ir and M = Si, Ge, Sn). J. Magn. Magn. Mater. 2007, 310, 207–213. 10.1016/j.jmmm.2006.10.289. [DOI] [Google Scholar]
  69. Szytuła A.; Penc B.; Kaczorowski D.; Arulraj A.; Baran S.; Stüsser N.; Tomala K. Magnetic and Electronic Properties of RCoxGe2 (R = Pr, Nd) Compounds. J. Alloys Compd. 2008, 460, 120–124. 10.1016/j.jallcom.2007.06.089. [DOI] [Google Scholar]
  70. Khan M. A.; McCandless G. T.; Benavides K. A.; Martin T. J.; Palacios A. M.; Samuel A. W. B.; Young D. P.; Chan J. Y. Crystal Growth and Magnetic Properties of Pr3Co2+xGe7 and the Sn-Stabilized Ln3Co2+xGe7-ySny (Ln = Pr, Nd, Sm). Cryst. Growth. Des. 2018, 18, 6028–6034. 10.1021/acs.cgd.8b00868. [DOI] [Google Scholar]
  71. Lin X.; Bud’ko S. L.; Thimmaiah S.; Canfield P. C. Anisotropic Magnetization, Resistivity and Heat Capacity of Single Crystalline R3Ni2-xSn7 (R = La, Ce, Pr and Nd). J. Magn. Magn. Mater. 2013, 331, 53–61. 10.1016/j.jmmm.2012.11.011. [DOI] [Google Scholar]
  72. Harmening T.; Eul M.; Pöttgen R. Nickel-deficient Stannides Eu2Ni2-xSn5 - Structure, Magnetic Properties, and Mössbauer Spectroscopic Characterization. Z. Naturforsch. B: Chem. Sci. 2009, 64, 1107–1114. 10.1515/znb-2009-1003. [DOI] [Google Scholar]
  73. Weiland A.; Eddy L. J.; McCandless G. T.; Hodovanets H.; Paglione J.; Chan J. Y. Refine Intervention: Characterizing Disordered Yb0.5Co3Ge3. Cryst. Growth. Des. 2020, 20, 6715–6721. 10.1021/acs.cgd.0c00865. [DOI] [Google Scholar]
  74. Moshopoulou E. G.; Ibberson R. M.; Sarrao J. L.; Thompson J. D.; Fisk Z. Structure of Ce2RhIn8: An Example of Complementary Use of High-Resolution Neutron Powder Diffraction and Reciprocal-Space Mapping to Study Complex Materials. Acta Crystallogr. B 2006, 62, 173–189. 10.1107/S0108768106003314. [DOI] [PubMed] [Google Scholar]
  75. Hering E. N.; Borges H. A.; Ramos S. M.; Fontes M. B.; Baggio-Saitovich E.; Continentino M. A.; Bittar E. M.; Mendonça Ferreira L.; Lora-Serrano R.; Gandra F. C. G.; Adriano C.; Pagliuso P. G.; Moreno N. O.; Sarrao J. L.; Thompson J. D. Residual Superconducting Phases in The Disordered Ce2Rh1-xIrxIn8 Alloys. Phys. Rev. B 2010, 82, 184517. 10.1103/PhysRevB.82.184517. [DOI] [Google Scholar]
  76. Uhlířová K.; Prokleška J.; Sechovský V. Comment on `̀Emergence of a Superconducting State from an Antiferromagnetic Phase in Single Crystals of the Heavy Fermion Compound Ce2PdIn8’’. Phys. Rev. Lett. 2010, 104, 059701. 10.1103/PhysRevLett.104.059701. [DOI] [PubMed] [Google Scholar]
  77. Kaczorowski D.; Pikul A. P.; Gnida D.; Tran V. H. Emergence of a Superconducting State from an Antiferromagnetic Phase in Single Crystals of the Heavy Fermion Compound Ce2PdIn8. Phys. Rev. Lett. 2009, 103, 027003. 10.1103/PhysRevLett.103.027003. [DOI] [PubMed] [Google Scholar]
  78. Wirth S.; Prots Y.; Wedel M.; Ernst S.; Kirchner S.; Fisk Z.; Thompson J. D.; Steglich F.; Grin Y. Structural Investigations of CeIrIn5 and CeCoIn5 on Macroscopic and Atomic Length Scales. J. Phys. Soc. Jpn. 2014, 83, 061009. 10.7566/JPSJ.83.061009. [DOI] [Google Scholar]
  79. Uhlířová K.; Prokleška J.; Sechovský V.; Daniš S. Solution Growth of Ce-Pd-In Single Crystals: Characterization of The Heavy-Fermion Superconductor Ce2PdIn8. Intermetallics 2010, 18, 2025–2029. 10.1016/j.intermet.2010.03.017. [DOI] [Google Scholar]
  80. Bao J.-K.; Zheng H.; Wen J.; Ramakrishnan S.; Zheng H.; Jiang J. S.; Bugaris D.; Cao G.; Chung D. Y.; van Smaalen S.; Kanatzidis M. G.. Superconductivity in Y4RuGe8 with a Vacancy-Ordered CeNiSi2-Type Superstructure. Chem. Mater. 2021, 33, 7839. 10.1021/acs.chemmater.1c02488 [DOI] [Google Scholar]
  81. Wright C. J. C.; Dooryhée E.; Pressley L. A.; Phelan W. A.; Khalifah P. G.; Billinge S. J. L. Toward In Situ Synchrotron Mapping of Crystal Selection Processes during Crystal Growth. Chem. Mater. 2021, 33, 3359–3367. 10.1021/acs.chemmater.1c00602. [DOI] [Google Scholar]
  82. Shoemaker D. P.; Hu Y.-J.; Chung D. Y.; Halder G. J.; Chupas P. J.; Soderholm L.; Mitchell J. F.; Kanatzidis M. G. In Situ Studies of a Platform for Metastable Inorganic Crystal Growth and Materials Discovery. Proc. Natl. Acad. Sci. U.S.A. 2014, 111, 10922–10927. 10.1073/pnas.1406211111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Weiland A.; Frith M. G.; Lapidus S. H.; Chan J. Y. In Situ Methods for Metal-Flux Synthesis in Inert Environments. Chem. Mater. 2021, 33, 7657–7664. 10.1021/acs.chemmater.1c02413. [DOI] [Google Scholar]
  84. Schultze D. Applications of Differential Thermal Analysis to the Preparation of Single Crystals. Thermochim. Acta 1991, 190, 77–110. 10.1016/0040-6031(91)87140-R. [DOI] [Google Scholar]
  85. Janssen Y.; Angst M.; Dennis K. W.; McCallum R. W.; Canfield P. C. Differential Thermal Analysis and Solution Growth of Intermetallic Compounds. J. Cryst. Growth 2005, 285, 670–680. 10.1016/j.jcrysgro.2005.08.044. [DOI] [Google Scholar]
  86. Xu L.; Wang Z.; Chen J.; Chen S.; Yang W.; Ren Y.; Zuo X.; Zeng J.; Wu Q.; Sheng H. Folded Network and Structural Transition in Molten Tin. Nat. Commun. 2022, 13, 126. 10.1038/s41467-021-27742-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Latturner S. E. Clusters, Assemble: Growth of Intermetallic Compounds from Metal Flux Reactions. Acc. Chem. Res. 2018, 51, 40–48. 10.1021/acs.accounts.7b00483. [DOI] [PubMed] [Google Scholar]
  88. Canfield P.; Kong T.; Kaluarachchi U.; Jo N. H. Use of Frit-disc Crucibles for Routine and Exploratory Solution Growth of Single Crystalline Samples. Philos. Mag. 2016, 96, 84–92. 10.1080/14786435.2015.1122248. [DOI] [Google Scholar]
  89. Krause L.; Herbst-Irmer R.; Sheldrick G. M.; Stalke D. Comparison of Silver and Molybdenum Microfocus X-Ray Sources for Single-Crystal Structure Determination. J. Appl. Crystallogr. 2015, 48, 3–10. 10.1107/S1600576714022985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Sheldrick G. M. SHELXT - Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. A: Found. Adv. 2015, 71, 3–8. 10.1107/S2053273314026370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Sheldrick G. M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao2c02152_si_001.pdf (163.1KB, pdf)
ao2c02152_si_002.cif (347.6KB, cif)
ao2c02152_si_003.cif (519.9KB, cif)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES