Significance
A wide range of strongly correlated quantum materials, including some high-temperature superconductors, exhibit “electronic nematic” phases, in which the electronic properties spontaneously break the rotational symmetry of the crystal. However, the role that the corresponding nematic fluctuations play in these complicated systems is unclear, motivating the search for simpler model systems. Here, we identify a particular 4f intermetallic material which exhibits ferroquadrupole order, , as just such a model system. We also provide a robust and accurate method to probe the divergence of an important associated quantity, the quadrupole-strain susceptibility. The temperature dependence of this quantity provides insight into the nature of the interactions that lead to ferroquadrupole order, in this case, magnetoelastic coupling.
Keywords: electronic nematicity, strongly correlated systems, ferroquadrupolar order
Abstract
Ferroquadrupole order associated with local atomic orbitals of rare-earth ions is a realization of electronic nematic order. However, there are relatively few examples of intermetallic materials which exhibit continuous ferroquadrupole phase transitions, motivating the search for additional materials that fall into this category. Furthermore, it is not clear a priori whether experimental approaches based on transport measurements which have been successfully used to probe the nematic susceptibility in materials such as the Fe-based superconductors will be as effective in the case of intermetallic materials, for which the important electronic degrees of freedom are local rather than itinerant and are consequently less strongly coupled to the charge-carrying quasiparticles near the Fermi energy. In the present work, we demonstrate that the intermetallic compound exhibits a tetragonal-to-orthorhombic phase transition consistent with ferroquadrupole order of the Yb ions and go on to show that elastoresistivity measurements can indeed provide a clear window on the diverging nematic susceptibility in this system. This material provides an arena in which to study the causes and consequences of electronic nematicity.
Electronic nematic order in crystalline solids corresponds to a spontaneously broken discrete rotational symmetry driven by interactions between low-energy electronic degrees of freedom (1). The empirical observation that nematic order occurs in, or near to, superconducting phases in Fe-based materials (2–4) and possibly also cuprates (5–8) motivates a series of fundamental questions about the possible role(s) that nematicity might play in such systems (9–11). Fe-based and cuprate superconductors are complicated materials, with a variety of other intertwined electronic phases, making it highly desirable to identify simpler model systems that exhibit nematic order for which the underlying effective Hamiltonian is better understood. Ferroquadrupole order associated with local atomic orbitals of rare-earth ions is a realization of electronic nematic order (12) and in principle can provide just such a model system.
The effective Hamiltonian describing the low-energy properties of rare-earth ions incorporated in crystalline solids is understood in great detail. The electronic wavefunctions are spatially localized, and strong spin–orbit coupling yields an electronic multiplet that is described by a total angular momentum number . Due to the local nature of the electrons, the crystal electric field (CEF) from the surrounding ligands acts as a weak perturbation, splitting the degeneracy of the spherical harmonic basis states. Significantly, the low-energy states of these electronic multiplets can exhibit degeneracies or near degeneracies. In these degenerate or near degenerate manifolds, there can exist states with large multipolar moments. Both the coupling of the lattice with quadrupolar degrees of freedom and a coupling from the generalization of the Ruderman–Kittel–Kasuya–Yosida (RKKY) interaction mediated by conduction electrons can then provide energetically favorable conditions for multipolar instabilities to occur (13). A variety of multipolar behaviors have been observed in such materials. The specific case of ferroquadrupole order, in which each atomic site takes on a quadrupole moment of the same orientation, breaks the rotational symmetry of the point group of the crystal and hence provides an example of a lattice in which sites exhibit collective nematic order.
Several cubic rare-earth intermetallic phases are known to undergo ferroquadrupole order, including (14), (15), and (16). However, because of the presence of a cubic invariant in the free energy, such cubic cases cannot undergo a continuous cubic to tetragonal phase transition. In contrast, there are relatively few examples of intermetallic materials which exhibit continuous ferroquadrupole phase transitions, the primary examples being tetragonal-to-orthorhombic phase transitions observed for (17) and (18, 19), motivating the search for additional materials that fall into this category. The present work adds to that list.
has a tetragonal crystal structure at room temperature (space group /mmm), belonging to the same structure type as the familiar “122” Fe-pnictides such as . The electrons of the trivalent Yb ion form local multiplets with a total angular momentum quantum number J = 7/2. The CEF for the tetragonal point group symmetry results in four Kramers doublets, two with and two with character (20) (SI Appendix, section 1). The material displays some weak heavy fermion characteristics (21), but the low-temperature properties can be well understood in terms of these local orbitals (20). Analysis of heat capacity, susceptibility, and inelastic neutron-scattering measurements indicate that the CEF ground state is a quasi-quartet comprising states with almost pure () and () character, split by approximately 300 K from the other two doublets (20, 21). The material exhibits three continuous phase transitions at low temperature associated with the quasi-quartet: a nonmagnetic phase transition at = 10.2 K and two magnetic phase transitions [one at = 6.5 K, corresponding to the onset of a collinear amplitude-modulated antiferromagnet state, and a subsequent one at = 5.7 K, the origin of which is currently unknown (21, 22)]. Based on analysis of the multipole moments of the CEF ground state, combined with a mean-field treatment of the effective Hamiltonian for the system, it has been suggested that the origin of the nonmagnetic transition is a ferroquadrupolar ordering of the local electrons (ref. 21 and SI Appendix, section 3). However, to date no measurements have confirmed the ferroquadrupolar nature of the phase transition, nor has the quadrupole-strain susceptibility been investigated.
In the present work, we use high-resolution X-ray diffraction to reveal that the nonmagnetic phase transition corresponds to a tetragonal-to-orthorhombic phase transition, consistent with spontaneous ferroquadrupole order with symmetry. We then present elastoresistivity measurements which reveal a divergence of the quadrupole-strain susceptibility (i.e., the nematic susceptibility) in the same () symmetry channel, . Comparison of the divergence of and the critical temperature for the ferroquadrupole order allows evaluation of the Levy criterion (defined below), from which it is inferred that the ferroquadrupole transition in is driven primarily by magnetoelastic coupling. Our measurements establish as a model system to explore electronic nematicity, for example as a possible starting point from which the ferroquadrupole order could be continuously suppressed toward a nematic quantum phase transition.
We start by describing the quadrupole-strain susceptibility (i.e., the nematic susceptibility for the quadrupolar system) and its significance for ferroquadrupole order.
A minimal Hamiltonian can be written which describes the quadrupolar behavior of the electrons in :
[1] |
The first term in the sum is the bilinear magnetoelastic quadrupole-strain coupling, and the second term is a mean-field quadrupole interaction term originating from an effective RKKY coupling mediated by the conduction electrons (13). The sum is over the three relevant symmetry channels for the tetragonal system, , , and . are the quadrupole operators of the th symmetry channel (the well-known Stevens operators) and is the corresponding strain. describes the crystal electric field from the surrounding Ru and Ge atoms that produces the ground-state quasi-quartet (SI Appendix, section 1). The gap between the ground-state doublet and the first excited doublet has been proposed to be from entropy considerations (21). The other doublets have been shown to be much higher in energy () and can be ignored for any low-temperature physics.
Strain , induced by an external stress, admixes and splits the CEF eigenstates (Fig. 1), resulting in an induced quadrupole moment. The associated quadrupole-strain susceptibility can be defined for each symmetry channel . Considering first the case in which = 0, the quadrupole-strain susceptibility has the form
[2] |
where is the magnitude of the quadrupole moment (SI Appendix, section 3). In the limit , this reduces to an approximate Curie form , which becomes exact for .
Adding quadrupole interactions to this model renormalizes the quadrupole-strain susceptibilities. Considering again the limit , this results in a Curie–Weiss functional form:
[3] |
Thus, the Weiss temperature of the quadrupole-strain susceptibility in a given symmetry channel is proportional to the strength of the quadrupole–quadrupole interaction term and the saturated quadrupole moment in that channel. If the lattice were infinitely stiff (i.e., strain is treated as a fixed parameter), the system would have a tendency to undergo quadrupolar order at this “bare” critical temperature.
Of course, the lattice is not infinitely stiff, and the magnetoelastic coupling renormalizes the critical temperature. Minimizing the free energy with respect to strain, including the elastic energy terms, the ordering temperature is found to be
[4] |
where B and C are the magnetoelastic and elastic constants, respectively, of the ordering symmetry channel i, and G represents the renormalized quadrupolar interaction coefficient.
The relative magnitudes of and dictate whether the nematic phase transition is driven by strain, in the traditional Cooperative Jahn–Teller sense (23), or by electronic quadrupolar fluctuations, as inferred for certain ferroquadrupole systems (13, 17, 18) and underdoped 122 Fe-pnictides (4, 24). Thus, a precise measurement that determines the ratio of these two components, known as the Levy criterion, is an essential development to understanding the nature of quadrupolar phase transitions in any given system.
While the quadrupole-strain susceptibility is frequently invoked from a theoretical perspective (13), it has rarely been determined experimentally. Simple consideration of thermodynamic relations implies that can be inferred from softening of the elastic constants, magnetostriction, and measurements of the third-order magnetic susceptibility (13), and all of these approaches have been used previously in other candidate systems (19). However, while measurements of the elastic moduli provide a direct perspective on the lattice symmetry modes, a large temperature-dependent background and a strict reliance on a suitable geometry of the sample make it difficult in practice to obtain the temperature dependence of with quantitative accuracy. Similarly, magnetic measurements face difficulties in obtaining the temperature dependence of , in part because of issues with demagnetization fields and in part because quadrupolar-field susceptibilities must be untangled from the quadrupole-strain susceptibilities to properly infer them. Here, rather than inferring based on a combination of thermodynamic measurements, we outline how this quantity can be determined via elastoresistance measurements.
Although resistivity is not a thermodynamic measurement, it is a sensitive measure of the electronic scattering rate in a crystal. Considering intermetallic systems, the local electronic sites serve as strong scatterers of the conduction electrons. Anisotropic scattering from these sites and their associated quadrupolar moments has previously been described for cubic materials (25) but not investigated in the context of quadrupolar phase transitions in tetragonal materials. Under two conditions, which can be readily justified for , we can directly relate a change in the normalized anisotropic resistivity (divided by the isotropic resistivity) to a change in the corresponding quadrupole moment (SI Appendix, section 5 and ref. 26). The first condition is that the resistivity is dominated by scattering of the conduction electrons from the lattice of electronic mutliplets. As we demonstrate, this is indeed the case for . The second condition is that the induced quadrupole moment is perturbative, which is by definition the case for the technique that we use to determine the elastoresistivity.
When these two conditions are satisfied, an elastoresistivity technique, in which a tunable strain is applied to a material while measuring the induced anisotropic resistivity (illustrated schematically in Fig. 1), provides an ideal method to directly probe the temperature dependence of the material’s quadrupole-strain susceptibility. For example, considering the symmetry channel (Fig. 1, middle row), symmetry strain (caused by some external stresses) induces a finite symmetry change in the resistivity , where and are measured in the presence of the strain and is the unstrained isotropic in-plane resistivity. If the scattering rate is dominated by scattering from the 4f orbitals, and if the effective mass of the conduction electrons is not strongly affected by the induced strain, then
[5] |
Taking the appropriate derivatives with respect to strain gives
[6] |
Thus, measuring the symmetry decomposed elements of the elastoresistivity tensor provides direct measurements of the corresponding quadrupole-strain susceptibilities. Appropriate orientations of crystal axes, normal strain axes, and contact placement necessary to measure each symmetry channel are illustrated in Fig. 1 and described in greater detail in ref. 27.
Elastoresistance techniques have previously been used to determine the nematic susceptibility of materials for which the nematicity derives from itinerant electronic states, including Fe-based superconductors (24, 27, 29–31) and (32). Here, we demonstrate that the technique is also appropriate to determine of materials that undergo ferroquadrupole order of local orbitals. Since for such systems can in principle be large in several symmetry channels , we introduce additional means (described in Materials and Methods) to ensure appropriate symmetry decomposition, using a focused ion beam to prepare the sample for each measurement (Fig. 2).
Results
Splitting of the (6 0 0) Bragg peak was observed below 10.2 K along the (1 1 0) and (1 −1 0) directions, with the new peaks indicative of a orthorhombic structural distortion. The temperature dependence of the orthorhombicity parameter of is shown in Fig. 3, together with the anticipated temperature dependence of a mean-field Ising order parameter that onsets at 10.2 K (SI Appendix). No evidence for a similar phase transition is found for the nonmagnetic analog (ref. 33 and Fig. 4, Inset). The data for are consistent with a continuous ferroquadrupole phase transition, as predicted previously based on analysis of the CEF parameters (21). Of the two possibilities (corresponding to either or symmetry), these measurements determine that the ferroquadrupolar state has a symmetry.
The temperature dependence of the in-plane resistivity of unstrained (i.e., free-standing) is shown in Fig. 4. The data are consistent with previously published measurements, revealing a Kondo-like rise below 50 K and sharp features (seen more clearly in the derivative ) signifying the ferroquadrupole and magnetic phase transitions. Similar measurements of the nonmagnetic, iso-structural, iso-electronic analog, , also shown in Fig. 4, reveal a considerably smaller in-plane resistivity and no signatures of any phase transitions. Since the band structure of the two materials is presumably quite similar, the large difference in the resistivity can be attributed to scattering from the local orbitals in . Significantly, since the resistivity of in the temperature window above (blue shading in Fig. 4) is dominated by scattering, the elastoresistivity (described below) can be directly related to the quadrupole-strain susceptibility in the same symmetry channels, as described previously.
Elastoresistivity measurements were performed for multiple samples. Responses in the and symmetry channels were always found to be linear. Representative data for both of these symmetry channels (taken for samples of equal dimensions) are shown in Fig. 5 as a function of temperature. Elastoresistivity data for the symmetry channel were also recorded during these measurements; while these data do not correspond to a susceptibility of a symmetry-breaking quadrupole order, the measurements are presented in SI Appendix, section 7 for completeness. The response in the channel was found to be considerably larger than that in the channel. If the proportionality constants relating the resistivity anisotropy to the local quadrupole moment (SI Appendix, section 5) are similar for the two symmetry channels, which would be the case for an isotropic Fermi surface but is certainly not guaranteed by symmetry, then the observed large anisotropy in the elastoresistivity would imply that the magnetoelastic coupling coefficient for is significantly larger than that for the symmetry channel. This would certainly be consistent with the observed symmetry of the orthorhombic distortion, but needs to be verified by other thermodynamic measurements. The and elastoresistivity coefficients were also measured for . No significant temperature dependence for either channel was found within experimental error, as anticipated given the absence of partially filled orbitals in this compound.
Since the resistivity of in this temperature range is dominated by scattering from orbitals (Fig. 4), the temperature dependence of the elastoresistivity provides a good measure of the temperature dependence of the quadrupole-strain susceptibility. Motivated by the previous discussion, we fitted the elastoresistivity to a Curie–Weiss functional form:
[7] |
The quality of fit over the full temperature range, right down to (Fig. 5 and Fig. 5, Inset), indicates that the splitting of the quasi-quartet in the tetragonal state is less than in magnitude, consistent with previous estimates from inelastic neutron scattering (22). Significantly, with reference to Eq. 2, since the Weiss temperature is found to be close to 0 K, it is clear that quadrupolar interactions mediated by the conduction electrons do not play an important role in the eventual ferroquadrupole phase transition. In other words, referring back to Eq. 4 and the Levy criterion, we find that for in the symmetry channel and hence deduce that the ferroquadrupole transition must be driven primarily by magnetoelastic coupling in this material.
Summary and Discussion
is a counterpoint to underdoped Fe-based superconductors, such as , in terms of how the nematic phase transition is driven. Numerous measurements, including elastoresistivity (4, 24, 27, 29–31), elastic moduli (34, 35), Raman scattering (36–38), and NMR (39), have shown that the nematic susceptibility of and its analogs follows a Curie–Weiss functional form, with a Weiss temperature that is only slightly lower than the critical temperature for the coupled nematic/structural phase transition. Thus, while nematic–elastic coupling is still a necessary factor in driving the structural phase transition in those materials, the critical temperature is predominantly determined by electronic coupling between the nematic degrees of freedom (13). In contrast, presents a case in which magnetoelastic coupling is the dominant driving force for the phase transition. Even so, the material is clearly nematic: A separate degree of freedom (the local quadrupole moment) that couples bilinearly to elastic strain with the same symmetry is evident from the temperature dependence of the quadrupole-strain susceptibility, and its divergence drives a pseudo-proper ferroelastic phase transition. The lattice provides the effective coupling between the local quadrupole moments, but does not itself harbor any tendency toward a structural instability, as demonstrated by the absence of a comparable phase transition in the nonmagnetic analog .
The significance of the present observation is twofold. First, we have shown that elastoresistance measurements can be used to probe the nematic susceptibility (quadrupole-strain susceptibility) for local systems, at least in this specific case of . This provides an additional avenue to characterize a critically important characteristic of intermetallic quadrupolar systems. Since the quadrupole-strain susceptibility can, at least in principle, be large in several symmetry channels, we adopted a methodology of microstructuring the crystal using a focused ion beam (FIB) to guarantee symmetry decomposition of the resulting elastoresistivity.
Second, we have established as a model system to explore effects of nematicity in metals. Even though the phase transition is driven by magnetoelastic effects, fluctuations of the local quadrupoles couple directly to ungapped quasiparticles at the Fermi energy. Hence, if the ferroquadrupole order can be suppressed toward zero temperature, and if the phase transition remains continuous, is potentially an ideal system in which to explore the effects of nematic fluctuations on the properties of the metal in the quantum critical regime, including the possibility of emergent electronic phases. Methods of suppressing ferroquadrupole order using doping, magnetic field, or orthogonal antisymmetric strain (12) are readily experimentally accessible.
Materials and Methods
Single crystals of were grown from a high-temperature indium flux, as described elsewhere (40) (SI Appendix, section 9). The resulting crystals could be cleaved to a thickness of 20 m or less and were identified and oriented using single-crystal X-ray diffraction.
Low-temperature, high-resolution, X-ray diffraction measurements were performed on beamline A2 at Cornell High Energy Synchrotron Source (CHESS). The intensity profiles of the (6 0 0) Bragg peak were mapped out in momentum space from 40 K to 6.6 K. The orthorhombicity parameter was obtained by finding the first moment of the intensity (integrating counts multiplied by position) of line scans of the orthorhombic Bragg peaks (SI Appendix, section 2).
Flat samples were bonded to 50-m-thick Si substrates using Angstrom Bond. Hall bar patterns were then etched into the sample using a focused ion beam (see SI Appendix, section 8 for beam parameters) to precisely define current directions and contact geometry (Fig. 1). The Si substrates were glued onto the side of a piezoelectric stack, and strains experienced by the sample as a consequence of varying the voltage to the stack were estimated via resistive strain gauges, both on the side of the stack and on top of bare Si substrates (24).
Detailed elastoresistivity measurements were taken using an amplitude demodulation technique, following the method described in ref. 31. An ac voltage was applied to the piezoelectric stack, as well as an ac current applied to the sample, and the combination of mechanical and electrical excitations produced voltage signals at the sum and difference of the two frequencies, proportional to the unstrained resistivity of the sample multiplied by its elastoresistivity (31).
Supplementary Material
Acknowledgments
We acknowledge fruitful conversations with J. Straquadine and S. Kivelson. E.W.R. and I.R.F. were supported by the Gordon and Betty Moore Foundation Emergent Phenomena in Quantum Systems Initiative through Grant GBMF4414. Part of this work was performed at the Stanford Nano Shared Facilities, supported by the National Science Foundation (NSF) under Award ECCS-1542152. J.-H.C. acknowledges support from the State of Washington-funded Clean Energy Institute. Research conducted at the Cornell High Energy Synchrotron Source is supported by the NSF and NIH/National Institute of General Medical Sciences via NSF Award DMR-1332208. A.T.H. is supported by a NSF Graduate Research Fellowship under Grant DGE-114747.
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1818910116/-/DCSupplemental.
References
- 1.Fradkin E, Kivelson SA, Lawler MJ, Eisenstein JP, Mackenzie AP. Nematic Fermi fluids in condensed matter physics. Annu Rev Condens Matter Phys. 2010;1:153–178. [Google Scholar]
- 2.Fisher IR, Degiorgi L, Shen ZX. In-plane electronic anisotropy of underdoped “122” Fe-arsenides revealed by measurements of detwinned single crystals. Rep Prog Phys. 2011;74:124506. [Google Scholar]
- 3.Fernandes RM, Schmalian J. Manifestations of nematic degrees of freedom in the magnetic, elastic, and superconducting properties of the iron pnictides. Supercond Sci Technol. 2012;25:084005. [Google Scholar]
- 4.Chu JH, Kuo HH, Analytis JG, Fisher IR. Divergent nematic susceptibility in an iron arsenide superconductor. Science. 2012;337:710–712. doi: 10.1126/science.1221713. [DOI] [PubMed] [Google Scholar]
- 5.Ando Y, et al. Electrical resistivity anisotropy from self-organized one dimensionality in high-temperature superconductors. Phys Rev Lett. 2002;88:137005. doi: 10.1103/PhysRevLett.88.137005. [DOI] [PubMed] [Google Scholar]
- 6.Daou R, et al. Broken rotational symmetry in the pseudogap phase of a high-Tc superconductor. Nature. 2010;463:519–522. doi: 10.1038/nature08716. [DOI] [PubMed] [Google Scholar]
- 7.Cyr-Choiniere O, et al. Two types of nematicity in the phase diagram of the cuprate superconductor . Phys Rev B. 2015;92:224502. [Google Scholar]
- 8.Chang J, et al. Nernst effect in the cuprate superconductor : Broken rotational and translational symmetries. Nat Commun. 2016;7:11494. [Google Scholar]
- 9.Lederer S, Schattner Y, Berg E, Kivelson SA. Enhancement of superconductivity near a nematic quantum critical point. Phys Rev Lett. 2015;114:097001. doi: 10.1103/PhysRevLett.114.097001. [DOI] [PubMed] [Google Scholar]
- 10.Lederer S, Schattner Y, Berg E, Kivelson SA. Superconductivity and non-Fermi liquid behavior near a nematic quantum critical point. Proc Natl Acad Sci USA. 2017;114:4905–4910. doi: 10.1073/pnas.1620651114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Vojta M. Lattice symmetry breaking in cuprate superconductors: Stripes, nematics, and superconductivity. Adv Phys. 2009;58:699–820. [Google Scholar]
- 12.Maharaj A, et al. Transverse fields to tune an Ising-nematic quantum phase transition. Proc Natl Acad Sci USA. 2017;114:13430–13434. doi: 10.1073/pnas.1712533114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Morin P, Schmitt D. Quadrupolar interactions and magneto-elastic effects in rare earth intermetallic compounds. In: Wohlfarth EP, editor. Handbook of Ferromagnetic Materials. Vol 5. North-Holland; Amsterdam: 1990. pp. 1–132. [Google Scholar]
- 14.Bucher E, Birgeneau RJ, Maita JP, Felcher GP, Brun TO. Magnetic and structural phase transition in DySb. Phys Rev Lett. 1972;28:746–749. [Google Scholar]
- 15.Sato TJ, et al. Ferroquadrupolar ordering in . Phys Rev B. 2012;86:184419. [Google Scholar]
- 16.Shimura Y, et al. Field-induced quadrupolar quantum criticality in . Phys Rev B. 2015;91:241102(R). [Google Scholar]
- 17.Morin P, Rouchy J. Quadrupolar ordering in tetragonal . Phys Rev B. 1993;48:256–268. doi: 10.1103/physrevb.48.256. [DOI] [PubMed] [Google Scholar]
- 18.Kosaka M, et al. Quadrupolar ordering and magnetic properties of tetragonal . Phys Rev B. 1998;58:6339–6345. [Google Scholar]
- 19.Morin P, Kazei Z, Lejay P. Quadrupolar couplings and magnetic phase diagrams in tetragonal . J Phys Condens Matter. 1999;11:1305–1319. [Google Scholar]
- 20.Takimoto T, Thalmeier P. Theory of induced quadrupolar order in tetragonal . Phys Rev B. 2008;77:045105. [Google Scholar]
- 21.Jeevan HS, Geibel C, Hossain Z. Quasiquartet crystal-electric-field ground state with possible quadrupolar ordering in the tetragonal compound . Phys Rev B. 2006;73:020407(R). [Google Scholar]
- 22.Jeevan HS, et al. Muon spin relaxation and neutron diffraction investigations of quadrupolar and magnetically ordered states of . Phys Rev B. 2011;84:184405. [Google Scholar]
- 23.Gehring GA, Gehring KA. Co-operative Jahn-Teller effects. Rep Prog Phys. 1975;38:1–89. [Google Scholar]
- 24.Kuo HH, et al. Ubiquitous signatures of nematic quantum criticality in optimally doped Fe-based superconductors. Science. 2016;352:958–962. doi: 10.1126/science.aab0103. [DOI] [PubMed] [Google Scholar]
- 25.Sablik M, Levy PM. Anisotropy from quadrupole scattering in magnetically ordered rare-earth compounds. J Appl Phys. 1978;49:2171–2173. [Google Scholar]
- 26.Friederich A, Fert A. Electron scattering by the electronic quadrupole moment of rare-earth impurities. Phys Rev Lett. 1974;33:1214–1216. [Google Scholar]
- 27.Shapiro MC, Hristov AT, Palmstrom JC, Chu JH, Fisher IR. Measurement of the and components of the elastoresistivity tensor for tetragonal materials via transverse resistivity configurations. Rev Sci Instrum. 2016;87:063902. doi: 10.1063/1.4953334. [DOI] [PubMed] [Google Scholar]
- 28.Walmsley P, Fisher IR. Determination of the resistivity anisotropy of orthorhombic materials via transverse resistivity measurements. Rev Sci Instrum. 2017;88:043901. doi: 10.1063/1.4978908. [DOI] [PubMed] [Google Scholar]
- 29.Kuo HH, Shapiro MC, Riggs SC, Fisher IR. Measurement of the elastoresistivity coefficients of the underdoped iron arsenide Ba(Fe0.975Co0.025)2As2. Phys Rev B. 2013;88:085113. [Google Scholar]
- 30.Palmstrom JC, et al. Critical divergence of the symmetric () nonlinear elastoresistance near the nematic transition in an iron-based superconductor. Phys Rev B. 2017;96:205133. [Google Scholar]
- 31.Hristov AT, et al. Measurement of elastoresistivity at finite frequency by amplitude demodulation. Rev Sci Instrum. 2018;89:103901. doi: 10.1063/1.5031136. [DOI] [PubMed] [Google Scholar]
- 32.Riggs SC, et al. Evidence for a nematic component to the hidden-order parameter in from differential elastoresistance measurements. Nat Commun. 2015;6:6425. doi: 10.1038/ncomms7425. [DOI] [PubMed] [Google Scholar]
- 33.Bouvier M, et al. Anomalous low-temperature magnetic behaviour in and . Europhys Lett. 1996;33:647–652. [Google Scholar]
- 34.Fernandes RM, et al. Effects of nematic fluctuations on the elastic properties of iron arsenide superconductors. Phys Rev Lett. 2010;105:157003. doi: 10.1103/PhysRevLett.105.157003. [DOI] [PubMed] [Google Scholar]
- 35.Bohmer AE, et al. Nematic susceptibility of hole-doped and electron-doped iron-based superconductors from shear modulus measurements. Phys Rev Lett. 2014;112:047001. doi: 10.1103/PhysRevLett.112.047001. [DOI] [PubMed] [Google Scholar]
- 36.Gallais Y, et al. Observation of incipient charge nematicity in Ba(Fe1−xCox)2As2. Phys Rev Lett. 2013;111:267001. doi: 10.1103/PhysRevLett.111.267001. [DOI] [PubMed] [Google Scholar]
- 37.Thorsmolle VK, et al. Critical quadrupole fluctuations and collective modes in iron pnictide superconductors. Phys Rev B. 2016;93:054515. [Google Scholar]
- 38.Kretzschmar F, et al. Critical spin fluctuations and the origin of nematic order in Ba(Fe1−xCox)2As2. Nat Phys. 2016;12:560–563. [Google Scholar]
- 39.Kissikov T, et al. NMR study of nematic spin fluctuations in a detwinned single crystal of underdoped Ba(Fe1−xCox)2As2. Phys Rev B. 2016;94:165123. [Google Scholar]
- 40.Jeevan HS. 2010 Crystal growth and investigation of and : Competition/co-existence of superconducting, dipolar and quadrupolar order. Available at Core.ac.uk. Accessed January 24, 2018.
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.