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

\marginsize

2.0 cm2.0 cm1 cm1 cm

Van Vleck excitons in Ca2RuO4

P. M. Sarte California NanoSystems Institute, University of California, Santa Barbara, CA 93106-6105, USA Materials Department, University of California, Santa Barbara, CA 93106-5050, USA    C. Stock School of Physics and Astronomy, University of Edinburgh, Edinburgh EH9 3FJ, United Kingdom    B. R. Ortiz California NanoSystems Institute, University of California, Santa Barbara, CA 93106-6105, USA Materials Department, University of California, Santa Barbara, CA 93106-5050, USA    K. H. Hong School of Chemistry, University of Edinburgh, Edinburgh EH9 3FJ, United Kingdom Centre for Science at Extreme Conditions, University of Edinburgh, Edinburgh EH9 3FD, United Kingdom    S. D. Wilson California NanoSystems Institute, University of California, Santa Barbara, CA 93106-6105, USA Materials Department, University of California, Santa Barbara, CA 93106-5050, USA
Abstract

A framework is presented for modeling and understanding magnetic excitations in localized, intermediate coupling magnets where the interplay between spin-orbit coupling, magnetic exchange, and crystal field effects are known to create a complex landscape of unconventional magnetic behaviors and ground states. A spin-orbit exciton approach for modeling these excitations is developed based upon a Hamiltonian which explicitly incorporates single-ion crystalline electric field and spin exchange terms. This framework is then leveraged to understand a canonical Van Vleck j=eff0j\rm{{}_{eff}}=0 singlet ground state whose excitations are coupled spin and crystalline electric field levels. Specifically, the anomalous Higgs mode [Jain et al. Nat. Phys. 13, 633 (2017)], spin-waves [S. Kunkemöller et al. Phys. Rev. Lett. 115, 247201 (2015)], and orbital excitations [L. Das et al. Phys. Rev. X 8, 011048 (2018)] in the multiorbital Mott insulator Ca2RuO4 are captured and good agreement is found with previous neutron and inelastic x-ray spectroscopic measurements. Furthermore, our results illustrate how a crystalline electric field-induced singlet ground state can support coherent longitudinal, or amplitude excitations, and transverse wavelike dynamics. We use this description to discuss mechanisms for accessing a nearby critical point.

I Introduction

The magnetism inherent to Hund’s metals Georges et al. [2013] and their parent multiorbital Mott states Rau et al. [2016] is believed to play a crucial role in many of their anomalous electronic properties, Witczak-Krempa et al. [2014] ranging from unconventional superconductivity, Yin et al. [2011] to violations of Fermi liquid theory, to the recent unveiling of novel nonequilibrium states. Zhao et al. [2019] This is perhaps most clearly illustrated in the range of emergent phenomena that appear in 4d4d transition metal oxides—where an intermediate coupling regime manifests. In this regime, spin-orbit coupling, magnetic exchange, and crystalline electric field energies compete with one another on equal footing in determining a material’s ground state properties.

Due to the interplay of these competing energy scales, modeling the excitations out of the unusual ground states realized in this intermediate coupling regime and ultimately understanding their microscopic Hamiltonians is an enduring challenge. Magnetic excitations in this space become intertwined with other degrees of freedom as local orbital degeneracies are quenched via both the electrostatic crystal field potential and via spin-orbit coupling. This often yields a complex excitation spectrum reflective of transitions between exchange-coupled single-ion states and an “excitonic” energy landscape which is difficult to experimentally interpret.

One example of this richness appears in the unusual magnetic ground state of the multiorbital Mott insulator Ca2RuO4Cao et al. [1997]; Nakatsuji et al. [1997a]; Nakatsuji and Maeno [2000a, b]; Nakatsuji et al. [2003]; Anisimov et al. [2002]; Nakamura et al. [2002]; Porter et al. [2018] Ca2RuO4 possesses a distorted K2NiF4 structure Braden et al. [1998]; Steffens et al. [2005]; Pincini et al. [2018]; Friedt et al. [2001] (Fig. 1) with Ru4+ cations in a t2g4t^{4}_{2g} orbital configuration present in a strong crystalline electric field. In this setting, Ru4+ cations possess a spin angular momentum S=1S=1 and an effective orbital angular momentum of ll=1, yielding a nonmagnetic singlet j=eff0j\rm{{}_{eff}}=0 ground state. Neutron scattering measurements nevertheless observe long-range ordered antiferromagnetism, albeit with a reduced ordered moment \simμB\mu_{B}. The result is an unusual manifestation of a spin-orbit induced mixing of higher energy crystal field levels Feldmaier et al. [2020] that stabilizes a static magnetic moment in a naively j=eff0j\rm{{}_{eff}}=0 singlet ground state—a higher order state mixing leading to analogies with Van Vleck susceptibility.Khaliullin [2013] The ground state in Ca2RuO4 is different from the weakly magnetic ground states observed in compounds based on Kramers ions (e.g.e.g. CeRhSi3 Pásztorová et al. [2019] and YbRh2Si2 Stock et al. [2012]) that result from the near cancellation of the elastic magnetic cross section from contributions from differing members of the ground state doublet.

Recent experiments have since confirmed this mixed level structure in Ca2RuO4,Gretarsson et al. [2019] harking back to previous effects observed in rare earth intermetallic compounds. Wang and Cooper [1968] Compounds such as PrTl3 Birgeneau et al. [1971]; Holden and Buyers [1974] and TbSb Hsieh and Blume [1972] possess similar singlet magnetic non-Kramers ground states yet nevertheless also exhibit coherent sharp spin-waves Birgeneau et al. [1972] with ferromagnetically polarized ground states.Cooper and Vogt [1971]; Pink [1968] Ca2RuO4 is analogous Fang et al. [2004] in this regard given the apparent contradiction of its non-magnetic j=eff0j\rm{{}_{eff}}=0 ground state hosting weak antiferromagnetic order along with highly dispersive, coherent spin excitations.Kunkemöller et al. [2015] Notably however, the spin dynamics arising from the weak magnetic order in Ca2RuO4 have been so far analyzed through conventional Holstein-Primakoff approaches Jain et al. [2017] rather than employing the spin-orbit exciton framework Peschel et al. [1972] endemic to more strongly spin-orbit coupled rare earth systems.

Here we adapt the spin-orbit exciton framework to intermediate coupling oxides by modeling the low energy collective and single-ion magnetic excitations in Ca2RuO4. Comparison is made with previously reported spectroscopic data with the goal of understanding the origin of the anomalous transverse and longitudinal spin fluctuations reported in neutron scattering measurements.Jain et al. [2017]; Kunkemöller et al. [2017] The model further accounts for the single-ion physics and crystal field levels reported with inelastic x-rays Das et al. [2018] by including the Heisenberg interaction between spins and the single-ion terms in the Hamiltonian equally. This formalism implicitly includes the multiorbital nature of the Ru4+ ion in Ca2RuO4’s Mott state where intermediate coupling requires orbital and spin degrees of freedom are necessarily linked in the Hamiltonian through the spin-orbit interaction term. In using this minimal model and alternative approach, we are able to reproduce the anomalous collective transverse and longitudinal/amplitude excitations observed with neutrons as well as the higher energy spin-orbit transitions reported from inelastic x-ray experiments. This approach and its ability to quantitatively parameterize the magnetic Hamiltonian of Ca2RuO4 suggests its broader utility to model other multiorbital Mott states in the intermediate coupling regime as well its broader relevance for understanding the parent magnetic instabilities governing the behaviors of Hunds’ metals.

This paper is divided into five sections including this introduction. In section two, we first state the definitions of the problem in terms of response functions and the Hamiltonian under consideration. In section three, we establish the theoretical framework of the spin-orbit exciton model. In the fourth section, we then utilize the exciton model to account for previously reported inelastic neutronJain et al. [2017] and x-ray spectroscopic data,Das et al. [2018] yielding optimized parameters in the model Hamiltonian whose values are directly compared to their corresponding physical quantities reported in literature. In the fifth and final section, we infer necessary conditions for a crystalline electric field-induced singlet ground state to host both longitudinal excitations and transverse wavelike dynamics, while providing possible mechanisms to achieve quantum criticality.

II Definitions: Correlation, response functions, and scattering cross sections

In this section, we present the theoretical framework and the definitions of the spin-orbit exciton model and its ability to directly parametrize the magnetic neutron scattering response in Ca2RuO4.

Previously utilized to address the temperature dependence of the low energy magnetic fluctuations in PrTl3 by Buyers et al. Buyers et al. [1975] and more recently to understand the complex excitation spectra in CoO,Sarte et al. [2019] the spin-orbit exciton model employs the direct proportionality of the magnetic neutron cross section and the magnetic dynamic structural factor S(𝐐,ω)S({\bf{Q}},\omega) given by

S(𝐐,ω)=gL2f2(𝐐)αβ(δαβQ^αQ^β)Sαβ(𝐐,ω),S({\bf{Q}},\omega)=g_{L}^{2}f^{2}({\bf{Q}})\sum_{\alpha\beta}(\delta_{\alpha\beta}-\hat{Q}_{\alpha}\hat{Q}_{\beta})S^{\alpha\beta}({\bf{Q}},\omega),

corresponding to a product of the Landé gg factor gLg_{L}, the magnetic form factor f(𝐐)f(\mathbf{Q}), a polarization factor providing sensitivity to the component exclusively perpendicular to the momentum transfer 𝐐\mathbf{Q}, and the dynamic spin structure factor Sαβ(𝐐,ω)S^{\alpha\beta}({\bf{Q}},\omega). Corresponding to the Fourier transform of the spin-spin correlations

Sαβ(𝐐,ω)=12π𝑑teiωtS^α(𝐐,t)S^β(𝐐,0),S^{\alpha\beta}({\bf{Q}},\omega)=\frac{1}{2\pi}\int dte^{i\omega t}\langle\hat{S}^{\alpha}({\bf{Q}},t)\hat{S}^{\beta}(-{\bf{Q}},0)\rangle,

where α,β=x,y,z\alpha,\beta=x,y,z, Sαβ(𝐐,ω)S^{\alpha\beta}({\bf{Q}},\omega) as written above considers only the spin contribution to the neutron scattering cross section, a valid approximation given that the expectation value of the orbital angular momentum 𝐋^\langle{\hat{\bf{L}}}\rangle\equiv 0 via quenching for dd-orbitals,Yosida [1996]. In the next section, we will show that the orbital contribution to the scattering cross section does exist, and is enabled through a spin-orbit (𝐋^𝐒^\hat{\mathbf{L}}\cdot\hat{\mathbf{S}}) coupling term in the model Hamiltonian.

The relation of the structure factor Sαβ(𝐐,ω)S^{\alpha\beta}({\bf{Q}},\omega) to the response function is given by the fluctuation-dissipation theorem

Sαβ(𝐐,ω)=1π11exp(ω/kTB)Gαβ(𝐐,ω),S^{\alpha\beta}({\bf{Q}},\omega)=-\frac{1}{\pi}\frac{1}{1-\exp(\omega/k{\rm{{}_{B}}}T)}\Im{G^{\alpha\beta}(\bf{Q},\omega)}, (1)

and allows the magnetic neutron cross section to be defined in terms of a Green’s response function Gαβ(𝐐,ω)G^{\alpha\beta}(\bf{Q},\omega).Zubarev [1960] Recognizing that the neutron response function is proportional to the temperature dependent Bose factor multiplied by the Fourier transform of the retarded Green’s function

Gαβ(ij,t)=G(S^α(i,t),S^β(j,0))=iΘ(t)[S^α(i,t),S^β(j,0)],\begin{split}G^{\alpha\beta}(ij,t)&=G(\hat{S}^{\alpha}(i,t),\hat{S}^{\beta}(j,0))\\ &=-i\Theta(t)\langle[\hat{S}^{\alpha}(i,t),\hat{S}^{\beta}(j,0)]\rangle,\end{split}

where S^\hat{S} here denotes a generic angular momentum operator, it can be shown that the application of appropriate boundary conditions onto the time Fourier transform of the first time derivative of Gαβ(ij,t)G^{\alpha\beta}(ij,t), yields an equation-of-motion of the general form

ωG(a^,a^,ω)=[a^,a^]+G([a^,^],a^,ω),\omega G(\hat{a},\hat{a}^{\prime},\omega)=\langle[\hat{a},\hat{a}^{\prime}]\rangle+G([\hat{a},\hat{\mathcal{H}}],\hat{a}^{\prime},\omega), (2)

for a magnetic Hamiltonian ^\hat{\mathcal{H}} and a generic component of a general angular momentum operator a^\hat{a}. The presence of the [a^,^][\hat{a},\hat{\mathcal{H}}] in Eq. 2 demonstrates that the employment of response function Gαβ(𝐐,ω)G^{\alpha\beta}(\bf{Q},\omega) allows for the magnetic neutron cross section to be directly parametrized via the individual contributing terms to ^\hat{\mathcal{H}}. This equation-of-motion provides a direct connection between the model microscopic Hamiltonian of interest to the neutron scattering cross section that is measured experimentally.

In the case of magnetic fluctuations that stem from localized magnetic moments on site ii with spin 𝐒^(i)\hat{\mathbf{S}}(i), each with a single-ion crystal field (CF) contribution, and coupled to each other by Heisenberg exchange between sites ii and jj defined by J(ij)J(ij), ^\hat{\mathcal{H}} can be written as

^=^CF+ijJ(ij)𝐒^(i)𝐒^(j).\begin{split}\hat{\mathcal{H}}&=\hat{\mathcal{H}}_{CF}+\sum\limits_{ij}J(ij)\hat{\mathbf{S}}(i)\cdot\hat{\mathbf{S}}(j).\end{split} (3)

At temperatures T<TNT<T\rm{{}_{N}} (or TCT\rm{{}_{C}}), a molecular field stemming from the assumption of long range magnetic order will be present at each given site ii. This collective effect can be accounted for by a Zeeman-like term in the Hamiltonian given by

^MF(i)=iHMF(i)S^z(i),\hat{\mathcal{H}}_{MF}(i)=\sum\limits_{i}H_{MF}(i)\hat{S}_{z}(i), (4)

where the molecular field is

HMF(i)=2i>jJ(ij)S^z(j).H_{MF}(i)=2\sum\limits_{i>j}J(ij)\langle\hat{S}_{z}(j)\rangle. (5)

Using these definitions, it can be shown Buyers et al. [1975]; Sarte et al. [2019] that ^\hat{\mathcal{H}} can be divided into a sum of a single-ion (^1\hat{\mathcal{H}}_{1}) and an inter-ion (^2\hat{\mathcal{H}}_{2}) term given by

^1=i^CF(i)+i^MF(i),\hat{\mathcal{H}}_{1}=\sum\limits_{i}\hat{\mathcal{H}}_{CF}(i)+\sum\limits_{i}\hat{\mathcal{H}}_{MF}(i),

and

^2=ijJ(ij)S^z(i)[S^z(j)2S^z(j)]+12ijJ(ij)[S^+(i)S^(j)+S^(i)S^+(j)],\begin{split}\hat{\mathcal{H}}_{2}&=\sum\limits_{ij}J(ij)\hat{S}_{z}(i)[\hat{S}_{z}(j)-2\langle\hat{S}_{z}(j)\rangle]\\ &+\frac{1}{2}\sum\limits_{ij}J(ij)[\hat{S}_{+}(i)\hat{S}_{-}(j)+\hat{S}_{-}(i)\hat{S}_{+}(j)],\end{split}

respectively.

In the spin-orbit exciton model for magnetic excitations, the single-ion term is first diagonalized for a given molecular field to provide the basis states |n|n\rangle. In second quantization formalism, the diagonalization of ^1\hat{\mathcal{H}}_{1} is written as

^1|n=ωn|n,\hat{\mathcal{H}}_{1}|n\rangle=\omega_{n}|n\rangle, (6)

where ωn\omega_{n} corresponds to the energy eigenvalue of the |n|n\rangle Fock state. Such a diagonalization facilitates a redefinition of ^1\hat{\mathcal{H}}_{1} in terms of ladder operators C(i)C(i) and C(i)C^{\dagger}(i), such that

^1=niωnCn(i)Cn(i),\hat{\mathcal{H}}_{1}=\sum_{n}\sum_{i}\omega_{n}C^{\dagger}_{n}(i)C_{n}(i), (7)

where C(i)C(i) and C(i)C^{\dagger}(i) satisfy the commutation relations [Cn(i),Cm(j)]=δijδnm[C_{n}(i),C^{\dagger}_{m}(j)]=\delta_{ij}\delta_{nm}. It should be noted that for the purposes of conciseness, the circumflex accent ^\hat{} in the case of operators have not been included, as will remain convention for the remainder of this section.

Since the equation-of-motion (Eq. 2) involves the commutator of angular momentum operators with ^\hat{\mathcal{H}} (Eq. 3), the evaluation of the response function requires a common basis. In the spin-orbit exciton model, the Fock states {|n}\{|n\rangle\} accompanying the diagonalization of ^1\hat{\mathcal{H}}_{1} (Eqs. 6-7) provides a natural basis for such a task. By noting that the inter-ion term ^2\hat{\mathcal{H}}_{2} is itself a function of angular momentum operators, it becomes clear that the evaluation of the equation-of-motion requires the rotation of the components of the angular momentum operator 𝐒^\hat{\mathbf{S}} onto the {|n}\{|n\rangle\} basis. Such a coordinate rotation is given by

S(±,z)=mnS(±,z)mnCmCn,S_{(\pm,z)}=\sum\limits_{mn}S_{(\pm,z)mn}C^{\dagger}_{m}C_{n},

utilizing the same ladder operators that were previously defined in Eq. 7. Such an approach is analogous to the bosonic approach of SU(NN) spin-wave theory previously discussed in the context of Kramers j=eff1/2j\rm{{}_{eff}}=1/2 magnets. Dong et al. [2018]; Muniz et al. [2014] However, our excitonic description includes the crystalline electric field contribution explicitly in this analysis to calculate the uncoupled single-ion states. In later sections, we will use this excitonic formalism to investigate how the crystalline electric field can be tuned to access nearby critical points and induce anomalous excitations.

As shown by Buyers et al. Buyers et al. [1975] in the context of PrTl3, and further applied to CoO by Sarte et al.,Sarte et al. [2019, 2020] the evaluation of the equation-of-motion begins by first defining an inter-level susceptibility G~\tilde{G}

Gαβ(i,j,ω)=mnSαmnG~β(m,n,i,j,ω).G^{\alpha\beta}(i,j,\omega)=\sum\limits_{mn}S_{\alpha mn}\tilde{G}^{\beta}(m,n,i,j,\omega).

The use of Gαβ(i,j,ω)G^{\alpha\beta}(i,j,\omega), in combination with the projection of the full magnetic Hamiltonian ^=^1+^2\hat{\mathcal{H}}=\hat{\mathcal{H}}_{1}+\hat{\mathcal{H}}_{2} onto the {|n}\{|n\rangle\} basis, reduces the second term of the equation-of-motion (Eq. 2) to three sets of commutators, termed diagonal, transverse, and longitudinal, with each commutator involving spin operators that are written in terms of ladder operators. When combined Sarte et al. [2019] with the random phase decoupling methodWolff [1960]; Cooke [1973]; Yamada and Shimizu [1966, 1967] the evaluation of the three commutators in the T0T\rightarrow 0 K limit reduces the Fourier transform of the equation-of-motion (Eq. 2) into a set of coupled linear and homogeneous equations given by:

Gαβ(𝐐,ω)=gαβ(E)+gα+(ω)J(𝐐)Gβ(𝐐,ω)+gα(ω)J(𝐐)G+β(𝐐,ω)+2gαz(ω)J(𝐐)Gzβ(𝐐,ω),G^{\alpha\beta}(\mathbf{Q},\omega)=g^{\alpha\beta}(E)+g^{\alpha+}(\omega)J(\mathbf{Q})G^{-\beta}(\mathbf{Q},\omega)+g^{\alpha-}(\omega)J(\mathbf{Q})G^{+\beta}(\mathbf{Q},\omega)+2g^{\alpha z}(\omega)J(\mathbf{Q})G^{z\beta}(\mathbf{Q},\omega), (8)

describing the coupling of the single-site response function

gαβ(ω)=n{Sα0nSβn0ωωn0Sαn0Sβ0nω+ωn0},g^{\alpha\beta}(\omega)=\sum\limits_{n}\left\{\frac{S_{\alpha 0n}S_{\beta n0}}{\omega-\omega_{n0}}-\frac{S_{\alpha n0}S_{\beta 0n}}{\omega+\omega_{n0}}\right\}, (9)

by the Fourier transform of the exchange interaction

J(𝐐)=ijJijei𝐐𝐝ij.J(\mathbf{Q})=\sum\limits_{i\neq j}J_{ij}e^{i\mathbf{Q}\cdot\mathbf{d}_{ij}}. (10)

By noting that the non-zero single-site response functions in a highly symmetric local octahedral coordination environment are restricted to ++-, +-+, or zzzz combinations for αβ\alpha\beta, Eq. 8, upon the inclusion of site indices, can be simplified to

Refer to caption
Figure 1: (a, top) Isometric view of the orthorhombic PbcaPbca (S.G. #61) unit cell of Ca2RuO4, consisting of (a, bottom) quasi-two-dimensional psuedo-square lattice layers of corner-sharing distorted RuO6 octahedra.Braden et al. [1998] (b) Calculated energy eigenvalues as a function of the tetragonal distortion and the magnetic molecular field perturbations applied to the individual jeffj\rm{{}_{eff}} manifolds of the ground-state crystal-field triplet 3T1g (ll=1, SS=1) of Ru4+ in octahedral coordination. Both the eigenvalues and individual parameters: Γ\Gamma and HMFH\rm{{}_{MF}} have been normalized, and are presented to scale. (c) Calculated Tanabe-Sugano diagram for d4d^{4} in perfect octahedral coordination with a C/BC/B ratio of 4.6. Shaded rectangle corresponds to the 10Dq10~{}Dq/BB regime of interest for Ca2RuO4.Cao and DeLong [2013]
Gijαβ(𝐐,ω)=δijgiαβ(ω)+0kj1giαβ(ω)ΦJi,i+k(𝐐)Gi+k,jαβ(𝐐,ω),\begin{split}&G^{\alpha\beta}_{ij}(\mathbf{Q},\omega)=\\ &\delta_{ij}g^{\alpha\beta}_{i}(\omega)+\sum\limits_{0\leq k\leq j-1}g^{\alpha\beta}_{i}(\omega)\Phi J_{i,i+k}(\mathbf{Q})G^{\alpha\beta}_{i+k,j}(\mathbf{Q},\omega),\end{split} (11)

where αβ={+,+,zz}\alpha\beta=\{+-,-+,zz\}, and the prefactor Φ=2\Phi=2 when α\alpha = zz, and 1 otherwise.

Finally, the sum of Eq. 11 over both αβ\alpha\beta and ijij combinations yields the total response function

G(𝐐,ω)=αβijGijαβ=G+(𝐐,ω)+G+(𝐐,ω)+Gzz(𝐐,ω),\begin{split}&G(\mathbf{Q},\omega)=\sum\limits_{\alpha\beta}\sum\limits_{ij}G^{\alpha\beta}_{ij}\\ &=G^{+-}(\mathbf{Q},\omega)+G^{-+}(\mathbf{Q},\omega)+G^{zz}(\mathbf{Q},\omega),\end{split} (12)

whose imaginary component, by the fluctuation-dissipation theorem (Eqs. 1), is proportional to the total dynamic structure factor, and thus the low temperature magnetic neutron cross section.

III Microscopic model: single-ion terms and Heisenberg spin exchange

The simplification of the equation-of-motion (Eq. 2) to a set of coupled linear and homogeneous equations (Eq. 8, or equivalently Eq. 11), yields a model whose ω\omega- and 𝐐\mathbf{Q}-dependence is explicitly parametrized by the single-site response function gαβ(ω)g^{\alpha\beta}(\omega) coupled by the Fourier transform of the exchange interaction J(𝐐)J(\mathbf{Q}), respectively. This explicit paramtrization corresponds to one of the main advantages of the excitonic approach in addressing the scattering cross section as the single-ion physics, corresponding to the effects of the crystalline electric field, is directly incorporated through the uncoupled single site response function g(ω)g(\omega). In this section, we will discuss both the individual contributions to, and evaluation of, both these two terms in the equation-of-motion.

III.1 Single site gαβ(ω)g^{\alpha\beta}(\omega): Parameters & Approximations

With a sample temperature (55 K) much smaller than the energy transfers of interest (ω>15\hbar\omega>15 meV), it is a valid approximation to restrict the discussion exclusively to the T0T\rightarrow 0 K limit. In this limit, the single-site response function gαβ(ω)g^{\alpha\beta}(\omega) (Eq. 9) is solely a function of the single-ion Hamiltonian ^1\hat{\mathcal{H}}_{1} (Eq. II).

As illustrated in Fig. 1(b), there are four terms comprising the single-ion Hamiltonian for Ru4+ in a local octahedral crystalline electric field

^1=^CF+^MF=(^CEF+^SO+^dis)+^MF,\begin{split}\hat{\mathcal{H}}_{1}&=\hat{\mathcal{H}}_{CF}+\hat{\mathcal{H}}_{MF}=\\ &(\hat{\mathcal{H}}_{CEF}+\hat{\mathcal{H}}_{SO}+\hat{\mathcal{H}}_{dis})+\hat{\mathcal{H}}_{MF},\end{split}

corresponding to the individual contributions from the octahedral crystalline electric field ^CEF\hat{\mathcal{H}}_{CEF}, spin-orbit ^SO\hat{\mathcal{H}}_{SO}, the structural distortion ^dis\hat{\mathcal{H}}_{dis} away from ideal octahedral coordination, and a mean molecular field ^MF\hat{\mathcal{H}}_{MF} stemming from long range magnetic order. Additional contributions such as hyperfine nuclear splitting exhibit much weaker energy scales (μ\sim\mueV), and are thus neglected in the current discussion. Chatterji and Schneider [2009] We will now discuss each contributing term to this single-ion Hamiltonian, and thus the terms that ultimately parametrize the uncoupled single site susceptibilities g(ω)g(\omega).

The Octahedral Crystalline Electric Field, ^CEF\hat{\mathcal{H}}_{CEF}. In the case of the 4d4d^{4} ion Ru4+ in an octahedral environment surrounded by six oxygens, with a reported crystal field strength (10DqDq) of \sim 4 eV, Cao and DeLong [2013]; Gretarsson et al. [2019] it cannot be assumed that the dd-orbital splitting induced by the crystalline electric field is small in comparison to the energy cost of violating the Pauli principle and pairing electrons in individual dd-orbitals. This contrasts with the case of 3d3d ions such as Co2+ where a weak crystalline electric field is present, allowing both Hund’s rules and the Pauli exclusion principle to be applied amongst the five degenerate dd orbitals. Cowley et al. [2013] With a strong crystalline electric field, the basis is taken to be {|t,|e}\{|t\rangle,|e\rangle\}, corresponding to the dd orbital states that have been split electrostatically into a low energy triplet |t|t\rangle and higher energy doublet |e|e\rangle states. McClure [1959]; Tanabe and Sugano [1954a, b], separated in energy by a value of Δ=10Dq\Delta=10Dq.

The electronic configuration of the ground state can be determined by Hund’s rules combined with the Pauli principle, albeit in the context of a large gap Δ10Dq\Delta\equiv 10Dq. Whereas in the case of weak crystal field theory, where the values of both SS and LL are those for the free-ion states that are obtained from the direct application of Hund’s rules and Pauli principle on 5 degenerate dd-orbitals, in the case of Hund’s first rule and the spin angular momentum quantum number SS, its value is still maximized, but bound by the restriction that the |t|t\rangle manifold must be fully populated before proceeding to populating the higher energy |e|e\rangle manifold. Based on such a restriction, Hund’s first rule yields two unpaired electrons for a d4d^{4} free-ion configuration, corresponding to a total spin angular momentum quantum number SS=mS\sum m_{S}=1. This gives a |t4|e0|t^{4}\rangle|e^{0}\rangle configuration.

In the case of the total orbital angular momentum number LL, it is not clear on what this value should be simply based on Hund’s second rule since the dxyd_{xy}, dxzd_{xz}, and dyzd_{yz} orbitals which make up the |t|t\rangle manifold are mixtures of uncoupled |m|m\rangle states which are eigenstates of the LzL_{z}. Given the application of the Pauli principle discussed in the previous paragraph, we expect the orbital ground state to be triply degenerate given the |t4|e0|t^{4}\rangle|e^{0}\rangle electron configuration and only one of the |t|t\rangle orbitals are fully filled. This is confirmed by the Sugano-Tanabe diagram Tanabe and Sugano [1954a, b] presented in Fig. 1(c), that was calculated using the electrostatic matrices and character tables supplied in Refs. McClure, 1959 in the strong crystal field basis with the assumption that the ratio of the Racah parameters CB4.6\frac{C}{B}\sim 4.6. In the strong crystal field limit (10Dq>B10Dq>B) the ground state is a T1g3{}^{3}T_{1g} orbital triplet, itself stemming from the H3{}^{3}H (LL=5, SS=1) free-ion state (10Dq010Dq\rightarrow 0).

Spin-Orbit Coupling, ^SO\hat{\mathcal{H}}_{SO}. Corresponding to the relativistic interaction between the spin and orbital degrees of freedom, spin-orbit coupling is given by

^SO=λ𝐋^𝐒^,\hat{\mathcal{H}}_{SO}=\lambda\hat{\mathbf{L}}\hat{\cdot\mathbf{S}}, (13)

where λ\lambda is the spin-orbit coupling constant and expected to increase with atomic number as Z2\sim Z^{2}Landau and Lifshitz [1977] The inclusion of spin-orbit coupling in the magnetic Hamiltonian yields a non-zero [Sz,][S_{z},\mathcal{H}], and as a result, the expectation value of SzS_{z} is not conserved. This lack of conservation is in stark contrast with an exclusively Heisenberg magnetic Hamiltonian, and enables the possibility of an amplitude, or longitudinal zzzz, mode to exist. Pekker and Varma [2015] In the case of Ca2RuO4, |λ|0.1|\lambda|\sim 0.1 eV is an order of magnitude smaller than 10Dq10Dq Das et al. [2018]; Mizokawa et al. [2001]; Fatuzzo et al. [2015], and thus spin-orbit coupling can be considered as a perturbation to the crystal field states defined by ^CEF\hat{\mathcal{H}}_{CEF} in the {|t,|e}\{|t\rangle,|e\rangle\} basis.

The treatment of this particular perturbation can be simplified by noting that the magnetic fluctuations of interest originate exclusively from the T1g3{}^{3}T_{1g} crystal field ground state presented in Fig. 1 (c)(c), being already accessible with neutron incident energies Ei|λ|E_{i}\lesssim|\lambda| incident on a sample at 5 K. This exclusivity in the determination of the magnetic properties of Ca2RuO4, allows one to confine the current discussion to the triply degenerate crystal field ground state. Requiring a projection from the original {|t,|e}\{|t\rangle,|e\rangle\} basis onto the smaller |l=1,ml|l=1,m_{l}\rangle basis that defines the subspace spanned by the T1g3{}^{3}T_{1g} orbital triplet, the spin-orbit Hamiltonian (Eq. 13) can be rewritten as

^SO=αλ𝐥^𝐒^,\hat{\mathcal{H}}_{SO}=\alpha^{\prime}\lambda\hat{\mathbf{l}}\hat{\cdot\mathbf{S}}, (14)

consisting of new orbital angular momentum operators that act on the new |l=1,ml|l=1,m_{l}\rangle basis, accompanied by a scalar projection factor α\alpha^{\prime}. In contrast to magnets located in the weak-intermediate field regime, the determination of the scalar α\alpha^{\prime} for Ca2RuO4 is not particularly straightforward. In the case of the weak crystal field limit, the projection is between bases with good quantum numbers LL (mLm_{L}) or ll (ml), and SS (mSm_{S}). Possessing fixed orbital and spin angular momentum values, the projection is amenable to methods based on the matrix representation of the angular momentum operators, significantly simplifying the process for determining α\alpha^{\prime} as was done for CoO. Sarte et al. [2019] For magnetic ions located in the strong crystal field regime such as is the case for Ca2RuO4, the value of LL (mLm_{L}) was addressed by Griffith Griffith [1960] using the TT-PP equivalence relation

𝐋^(t)=𝐋^(p).\hat{\mathbf{L}}(t)=-\hat{\mathbf{L}}(p).

Valid in the Δλ\Delta\gg\lambda regime where minimal mixing between tt and ee states occur, the projection of the L=2L=2 tt onto the L=1L=1 pp states greatly simplifies the process of calculating α\alpha^{\prime} directly using representation theory. In the case of the orbital triplet ground state of the d4d^{4} configuration, Griffith Griffith [1960] and Moffitt Moffitt et al. [1959] determined a value of 12-\frac{1}{2} for α\alpha^{\prime} in the pure LL-SS Russell-Saunders coupling scheme.

Having projected 𝐋^\hat{\mathbf{L}} onto a fictitious operator 𝐥^\hat{\mathbf{l}}, the basis of the new spin-orbit Hamiltonian ^SO\hat{\mathcal{H}}_{SO} (Eq. 14) is now comprised of the 9 |l=1,ml;S=1,mS|l=1,m_{l};S=1,m_{S}\rangle states. Based on both the Landé interval rule and the addition theorem, ^SO\hat{\mathcal{H}}_{SO} yields three unique effective total angular momentum 𝐣^=𝐥^+𝐒^\hat{\mathbf{j}}=\hat{\mathbf{l}}+\hat{\mathbf{S}} manifolds, corresponding to j=eff0j\rm{{}_{eff}}=0, 1, and 2, with energy eigenvalues

E=(αλ2)[j(j+1)S(S+1)l(l+1)].E=\left(\frac{\alpha^{\prime}\lambda}{2}\right)[j(j+1)-S(S+1)-l(l+1)]. (15)

By employing both the projection constant α=12\alpha^{\prime}=-\frac{1}{2} previously determined by Griffith, and the reported value of 75 meV for |λ||\lambda|Das et al. [2018] the diagonalization of ^SO\hat{\mathcal{H}}_{SO}

diag(^SO)=[7500000000037.500000000037.500000000037.500000000037.500000000037.500000000037.500000000037.500000000037.5],\begin{split}&diag(\hat{\mathcal{H}}_{SO})=\\ &\left[\begin{array}[]{ccccccccc}-75&0&0&0&0&0&0&0&0\\ 0&-37.5&0&0&0&0&0&0&0\\ 0&0&-37.5&0&0&0&0&0&0\\ 0&0&0&-37.5&0&0&0&0&0\\ 0&0&0&0&37.5&0&0&0&0\\ 0&0&0&0&0&37.5&0&0&0\\ 0&0&0&0&0&0&37.5&0&0\\ 0&0&0&0&0&0&0&37.5&0\\ 0&0&0&0&0&0&0&0&37.5\\ \end{array}\right],\end{split}

confirms a singlet ground state, separated from triplet and pentet excited manifolds by Δ\Delta(singlet\rightarrowtriplet) = 37.537.5 meV αλ\equiv\alpha^{\prime}\lambda, and Δ\Delta(singlet\rightarrowpentet) = 112.5112.5 meV 3αλ\equiv 3\alpha^{\prime}\lambda (Fig. 1(b)), in agreement with Eq. 15. Confirmation of the assignment of j=eff0j\rm{{}_{eff}}=0, 1, and 2 to the singlet, triplet, and pentet manifolds, respectively, is accomplished by the projection of the components of the effective total angular momentum operator 𝐣^=𝐥^+𝐒^\hat{\mathbf{j}}=\hat{\mathbf{l}}+\hat{\mathbf{S}} onto the subspaces that are spanned by the three individual manifolds defined by ^SO\hat{\mathcal{H}}_{SO}. Such a projection corresponds to a rotation of the individual angular momentum operators from the original |l,ml,S,mS|l,m_{l},S,m_{S}\rangle basis to a basis |ϕSO|\phi{\rm{{}_{SO}}}\rangle, consisting of the individual eigenvectors of ^SO\hat{\mathcal{H}}_{SO}. In the matter of a generic angular momentum operator 𝒪^\hat{\mathcal{O}}, this particular rotation is achieved by

𝒪^|ϕSO=𝒞1𝒪^|l,ml𝒞,\hat{\mathcal{O}}_{|\phi_{SO}\rangle}=\mathcal{C}^{-1}\hat{\mathcal{O}}_{|l,m_{l}\rangle}\mathcal{C}, (16)

corresponding to the matrix multiplication of 𝒪^\hat{\mathcal{O}} by a transformation matrix 𝒞\mathcal{C} (and its inverse) which consists of the individual eigenvectors ϕSO\phi\rm{{}_{SO}} that are arranged in order of increasing energy. In the case of zz-component jz^\hat{j_{z}}, the rotation given by Eq. 16 yields

𝒞1jz^𝒞=[𝟎000000000𝟏𝟎𝟎000000𝟎𝟎𝟎000000𝟎𝟎𝟏000000000𝟐𝟎𝟎𝟎𝟎0000𝟎𝟏𝟎𝟎𝟎0000𝟎𝟎𝟎𝟎𝟎0000𝟎𝟎𝟎𝟏𝟎0000𝟎𝟎𝟎𝟎𝟐],\mathcal{C}^{-1}\hat{j_{z}}\mathcal{C}=\left[\begin{array}[]{c|ccc|ccccc}\mathbf{0}&0&0&0&0&0&0&0&0\\ \hline\cr 0&\mathbf{-1}&\mathbf{0}&\mathbf{0}&0&0&0&0&0\\ 0&\mathbf{0}&\mathbf{0}&\mathbf{0}&0&0&0&0&0\\ 0&\mathbf{0}&\mathbf{0}&\mathbf{1}&0&0&0&0&0\\ \hline\cr 0&0&0&0&\mathbf{-2}&\mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{0}\\ 0&0&0&0&\mathbf{0}&\mathbf{-1}&\mathbf{0}&\mathbf{0}&\mathbf{0}\\ 0&0&0&0&\mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{0}\\ 0&0&0&0&\mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{1}&\mathbf{0}\\ 0&0&0&0&\mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{0}&\mathbf{2}\\ \end{array}\right],

corresponding to a matrix whose top 1 ×\times 1, middle 3 ×\times 3, and bottom 5 ×\times 5 block matrices are identical to the Jz^\hat{J_{z}} operator in |j=eff0,mj,eff|j{\rm{{}_{eff}}}=0,m_{j,{\rm{eff}}}\rangle, |j=eff1,mj,eff|j{\rm{{}_{eff}}}=1,m_{j,{\rm{eff}}}\rangle, and |j=eff2,mj,eff|j{\rm{{}_{eff}}}=2,m_{j,{\rm{eff}}}\rangle bases, respectively. By performing the same projection for the jx^\hat{j_{x}} and jy^\hat{j_{y}} operators, it can be shown that these block matrices satisfy the canonical commutation relations of angular momentum 𝐣^=i𝐣^×𝐣^\hat{\mathbf{j}}=i\hat{\mathbf{j}}\times\hat{\mathbf{j}}, thus confirming that these block matrices do indeed correspond to valid angular momentum operators.

The Distortion Hamiltonian, ^dis\hat{\mathcal{H}}_{dis}: Employing single crystal neutron diffraction, Braden et al. Braden et al. [1998] were the first to identify that the d4d^{4} Ca2RuO4 exhibits a strong cooperative Jahn-Teller distortion away from an ideal octahedral environment. Over the next two decades, a plethora of extensive studies would establish that the distortion accompanies orbital ordering, corresponding to the driving mechanism for the metal-to-insulator transition at TMIT\rm{{}_{MI}}=357 K Alexander et al. [1999]; Zegkinoglou et al. [2005]; Qi et al. [2010]; Lee et al. [2002] and the presence of the “Higgs” mode. Zhang and Pavarini [2020]

To account for such a distortion in our model of the single-ion eigenstates, deviations of the crystalline electric field away from ideal local octahedral coordination are considered. By noting that the triplet-doublet gap Δ\Delta induced by a crystalline electric field in close proximity to the orbital cross-over value 10Dq/B10Dq/B\sim2.7 (Fig. 1(b)) is identical (in magnitude) to the splitting observed near the free-ion limit, it is a valid assumption that in the case of Ca2RuO4, the undistorted crystalline electric field ^CEF\hat{\mathcal{H}}_{CEF} can be written in terms of Stevens operators 𝒪^40\hat{\mathcal{O}}^{0}_{4} and 𝒪^44\hat{\mathcal{O}}^{4}_{4} as

^CEF=B40(𝒪^40+5𝒪^44).\hat{\mathcal{H}}_{CEF}=B^{0}_{4}\left(\hat{\mathcal{O}}^{0}_{4}+5\hat{\mathcal{O}}^{4}_{4}\right).

The Stevens parameter BlmB^{m}_{l} prefactor is a numerical coefficient given by Hutchings [1964]; Bauer and Rotter [2010]

Blm=|e|plmγlmrlΘl,B^{m}_{l}=-|e|p^{m}_{l}\gamma^{m}_{l}\langle r^{l}\rangle\Theta_{l}, (17)

where Θl\Theta_{l} corresponds to projection constants accompanying the conversion from Cartesian coordinates to angular momentum operators via the Wigner-Eckart theorem, and is commonly denoted as α\alpha and β\beta for l=2l=2 and l=4l=4, respectively. rl\langle r^{l}\rangle denotes the expectation values of the radial wavefunction and correspond to 3.319a02a^{2}_{0} and 20.22a04a^{4}_{0} for l=2l=2 and l=4l=4, respectively. Abragam and Bleaney [1986]; Hotta [2006] plmp^{m}_{l} are the scalar coefficients of the corresponding Tesseral functions ZlmZ^{m}_{l}. In the point charge approximation, where the charge density ρ(𝐫)δ(𝐫)\rho(\mathbf{r})\propto\delta(\mathbf{r}), the Tesseral functions are incoroporated into the γlm\gamma^{m}_{l} term given by

γlm=12l+1iqiZlm(xi,yi,zi)ϵ0ril+1,\gamma^{m}_{l}=\frac{1}{2l+1}\sum\limits_{i}\frac{q_{i}Z^{m}_{l}(x_{i},y_{i},z_{i})}{\epsilon_{0}r^{l+1}_{i}},

where qiq_{i} denotes the ithi\rm{{}^{th}} charge.

As a first approximation, we have considered the simplest case of a uniaxial distortion along zz. Given the definition above of the crystalline electric field in terms of the Stevens operator equivalents, an equivalent expression for a tetragonal distortion is given by Walter [1960]

^dis=B20𝒪^20=Γ(l^z223),\hat{\mathcal{H}}_{dis}=B^{0}_{2}\hat{\mathcal{O}}^{0}_{2}=\Gamma\left(\hat{l}^{2}_{z}-\frac{2}{3}\right),

and parametrized by Γ\Gamma, whose sign and magnitude are determined by B20B^{0}_{2}.

Table 1: Atomic coordinates of ions constituting a single distorted octahedron about a central magnetic Ru4+ used in the determination of the Stevens parameters for Ca2RuO4 in the point charge limit.
  Ion   xx   yy   zz
Ru4+ 0 0.5 0.5
O2-(1) 0.3058 0.7004 0.5225
O2-(2) 0.1942 0.2004 0.5225
O2-(3) -0.3058 0.2996 0.4775
O2-(4) -0.1942 0.7996 0.4775
O2-(5) -0.0611 0.5152 0.6667
O2-(6) 0.0611 0.4848 0.3333

To obtain an estimate for the energy scales and the signs of the Stevens coefficients to guide the analysis below, we have considered the simplest case for a point charge model. Corresponding to a single electron in a dd-orbital (LL=2), where α=221\alpha=-\frac{2}{21} and β=+263\beta=+\frac{2}{63}Hutchings [1964] Eq. 17 for a single distorted octahedra about a magnetic Ru4+ (Tab. 1) in the point charge limit produces Stevens parameters B40B^{0}_{4} = +1.32+1.32 meV>0>0 and B20B^{0}_{2}=+0.52+0.52 meV>0>0, consistent with what is expected for the ground state orbital triplet configuration T1g3{}^{3}T_{1g} of Ru4+. Consisting of one fully filled and two partially filled |t|t\rangle orbitals, the electronic configuration of Ru4+ is analogous to that of d2d^{2} V3+ placed in a weak octahedral crystalline electric field, where the sign for B40B^{0}_{4} and B20B^{0}_{2} are both positive. Abragam and Bleaney [1986]

The Molecular Field Hamiltonian, ^MF\hat{\mathcal{H}}_{MF}: Corresponding to the final perturbative term to ^CEF\hat{\mathcal{H}}\rm{{}_{CEF}} in the single-ion Hamiltonian, ^MF\hat{\mathcal{H}}_{MF} (Eq. 4) addresses the effect of the mean molecular field stemming from the assumption of long range magnetic order by the Ru4+ moments below T=N110T\rm{{}_{N}}=110 K. In the simplest case where a single dominant Ru4+-O2--Ru4+ superexchange pathway with an isotropic magnetic exchange constant J1J_{1} between z1z_{1} nearest neighbors is considered, ^MF\hat{\mathcal{H}}_{MF} (Eq. 4) reduces to

^MF=iHMF(i)S^z=2z1J1Sz^S^z.\hat{\mathcal{H}}_{MF}=\sum_{i}H_{MF}(i)\hat{S}_{z}=2z_{1}J_{1}\langle\hat{S_{z}}\rangle\hat{S}_{z}.

As illustrated in Fig. 1(b), ^MF\hat{\mathcal{H}}_{MF} corresponds to a Zeeman-like term that removes time-reversal symmetry, splitting originally degenerate jeffj\rm{{}_{eff}} levels. In the case where |J||λ||J|\rightarrow|\lambda| (or equivalently |H|MF/|αλ|1|H{\rm{{}_{MF}}}|/|\alpha^{\prime}\lambda|\rightarrow 1), the corresponding increase in splitting induced by ^MF\hat{\mathcal{H}}_{MF} results in the significant entanglement between individual jeffj\rm{{}_{eff}} levels.Sarte et al. [2018a, b] The presence of such a strong admixture renders the modeling of the these systems difficult using psuedo-bosonic (Holstein-Primakoff) approaches that are based on conventional linear spin-wave theory, ultimately making it necessary to employ alternative methods such as the multi-level spin-orbit exciton model discussed here.

Table 2: Displacement vectors 𝐝m,ij\mathbf{d}_{m,ij} for each Ru4+ constituting the ss- and dd-sublattices within the first two coordination shells mm of Ca2RuO4. All vectors and their magnitudes were calculated using the VESTA visualization software package, Momma and Izumi [2011] employing the reported unit cell parameters at T=11T=11 K. Braden et al. [1998] Numbers in parentheses indicate errors.
   mm    |𝐝m,ij||\mathbf{d}_{m,ij}| (Å)    Number of Neighbors    𝐝m,ij\mathbf{d}_{m,ij} in ss-Sublattice (aa)    𝐝m,ij\mathbf{d}_{m,ij} in dd-Sublattice (aa)
1 3.8618(5) 4 (12120)\left(\frac{1}{2}~{}\frac{1}{2}~{}0\right)
(12120)\left(\frac{1}{2}~{}-\frac{1}{2}~{}0\right)
(12120)\left(-\frac{1}{2}~{}\frac{1}{2}~{}0\right)
(12120)\left(-\frac{1}{2}~{}-\frac{1}{2}~{}0\right)
2 5.5150(11) 4 (1 0 0)
(1-1 0 0)
(0 1 0)
(0 1-1 0)
Refer to caption
Figure 2: Non-primitive pseudo-tetragonal (red) and primitive square (black) (a) direct and (b) reciprocal space unit cells of Ca2RuO4 viewed in the (a,b)(a,b) and (H,K)(H,K) planes, respectively. The antiferromagnetic unit cell in the (a,b)(a,b) basal plane coincides with the non-primitive pseudo-tetragonal unit cell. (c) High symmetry directions between high symmetry points of the antiferromagnetic reciprocal unit cell (red) that has been symmetrized with respect to the square lattice (black).

III.2 J(Q): Parameters & Approximations

Having already discussed the single-ion Hamiltonian and its role in determining both the eigenstates basis and the single site susceptibility g(ω)g(\omega), the discussion now shifts to addressing the coupling of these individual sites. By enabling the coupling of gαβ(ω)g^{\alpha\beta}(\omega), the Fourier transform of the exchange interaction J(𝐐)J(\mathbf{Q}) uniquely specifies the 𝐐\mathbf{Q}-dependence of the response function G(𝐐,ωG(\mathbf{Q},\omega). Being itself parametrized by both JijJ_{ij} and 𝐝ij\mathbf{d}_{ij}, corresponding to the magnetic exchange constant and displacement vector between moments located at sites ii and jj, respectively, an analytical expression for J(𝐐)J(\mathbf{Q}) requires detailed knowledge of both the nuclear and magnetic structures of the system under investigation.

In the case of Ca2RuO4,Braden et al. [1998]; Nakatsuji et al. [1997b] its orthorhombic PbcaPbca unit cell (aa=5.4074 Å, bb=5.5150 Å, cc=11.90520 Å) is a result of the reduction of symmetry of a I4/mmmI4/mmm unit cell through a combination of the rotation and tilting of the compressed [RuO6]8\left[{\rm{RuO}}_{6}\right]^{8-} octahedra about the cc axis and (a,b)(a,b) plane, respectively. Corresponding to the ideal K2NiF4 structure (Fig. 1), a structure type commonly observed among the cuprates and other high TcT_{c} superconductors, the derivation of the Ca2RuO4 PbcaPbca unit cell from I4/mmmI4/mmm suggests that the spin-orbit exciton model may utilize certain approximations commonly employed with these superconductors. Zhou et al. [2018]

The large inter-plane distance illustrated in Fig. 1(a), combined with a Néel state consisting of magnetic moments that lie almost exclusively in the (a,b)(a,b) plane, suggests the restriction of the spin-orbit exciton model in the case of Ca2RuO4 to a single layer in the (a,b)(a,b) basal plane defined by a pseudo-tetragonal unit cell (abca\simeq b\neq c) that is illustrated in Fig. 2. Such quasi-two dimensionality (Jc0J_{c}\approx 0) has been experimentally validated with reported inelastic spectra (Fig. 3) exhibiting an XYXY-like dispersion with a maximum at (H,K)=(0,0)(H,K)=(0,0)Kunkemöller et al. [2017]; Jain et al. [2017]; Kunkemöller et al. [2015]

As illustrated in Fig. 2(d) and summarized in Tab. 2, a 2d2d collinear antiferromagnet such as Ca2RuO4 can be reduced to two unique site indices corresponding to sublattices consisting of moments that are aligned anti-parallel and parallel relative to a reference moment. By exclusively considering the antiferromagnetically ordered Ru4+ moments contained in the first two coordination shells within the (a,b)(a,b) basal plane that are coupled with isotropic magnetic exchange constants J1J_{1} and J2J_{2}, Eq. 10 yields two unique expressions for the Fourier transform of the exchange constants:

Jd(𝐐)=2J1(cos(π(H+K))+cos(π(HK))),J_{d}({\bf{Q}})=2J_{1}(\cos(\pi(H+K))+\cos(\pi(H-K))), (18)

and

Js(𝐐)=2J2(cos(2π(H))+cos(2π(K))),J_{s}({\bf{Q}})=2J_{2}(\cos(2\pi(H))+\cos(2\pi(K))), (19)

for Ji,i+k(𝐐)J_{i,i+k}(\mathbf{Q}) in Eq. 12, where labels 𝐝\mathbf{d} and 𝐬\mathbf{s} denote different ii+ki\neq i+k and same i=i+ki=i+k site indices, respectively.

III.3 Model: Numerical Details

In the case of the 2d2d collinear antiferromagnet Ca2RuO4, the restriction of the site indices i,ji,j to two unique values reduces Eq. 12 to a series of four coupled linear equations given by:

G11+(𝐐,E)\displaystyle G^{+-}_{11}(\mathbf{Q},E) =g1+(E)+g1+(E)Js(𝐐)G11+(𝐐,E)\displaystyle=g^{+-}_{1}(E)+g^{+-}_{1}(E)J_{s}(\mathbf{Q})G^{+-}_{11}(\mathbf{Q},E)
+g1+(E)Jd(𝐐)G21+(𝐐,E)\displaystyle+g^{+-}_{1}(E)J_{d}(\mathbf{Q})G^{+-}_{21}(\mathbf{Q},E)
G21+(𝐐,E)\displaystyle G^{+-}_{21}(\mathbf{Q},E) =g2+(E)Js(𝐐)G21+(𝐐,E)\displaystyle=g^{+-}_{2}(E)J_{s}(\mathbf{Q})G^{+-}_{21}(\mathbf{Q},E)
+g2+(E)Jd(𝐐)G11+(𝐐,E)\displaystyle+g^{+-}_{2}(E)J_{d}(\mathbf{Q})G^{+-}_{11}(\mathbf{Q},E)
G12+(𝐐,E)\displaystyle G^{+-}_{12}(\mathbf{Q},E) =g1+(E)Js(𝐐)G12+(𝐐,E)\displaystyle=g^{+-}_{1}(E)J_{s}(\mathbf{Q})G^{+-}_{12}(\mathbf{Q},E)
+g1+(E)Jd(𝐐)G22+(𝐐,E)\displaystyle+g^{+-}_{1}(E)J_{d}(\mathbf{Q})G^{+-}_{22}(\mathbf{Q},E)
G22+(𝐐,E)\displaystyle G^{+-}_{22}(\mathbf{Q},E) =g2+(E)+g2+(E)Js(𝐐)G22+(𝐐,E)\displaystyle=g^{+-}_{2}(E)+g^{+-}_{2}(E)J_{s}(\mathbf{Q})G^{+-}_{22}(\mathbf{Q},E)
+g2+(E)Jd(𝐐)G12+(𝐐,E)\displaystyle+g^{+-}_{2}(E)J_{d}(\mathbf{Q})G^{+-}_{12}(\mathbf{Q},E)

and

G11zz(𝐐,E)\displaystyle G^{zz}_{11}(\mathbf{Q},E) =g1zz(E)+2g1zz(E)Js(𝐐)G11zz(𝐐,E)\displaystyle=g^{zz}_{1}(E)+2g^{zz}_{1}(E)J_{s}(\mathbf{Q})G^{zz}_{11}(\mathbf{Q},E)
+2g1zz(E)Jd(𝐐)G21zz(𝐐,E)\displaystyle+2g^{zz}_{1}(E)J_{d}(\mathbf{Q})G^{zz}_{21}(\mathbf{Q},E)
G21zz(𝐐,E)\displaystyle G^{zz}_{21}(\mathbf{Q},E) =2g2zz(E)Js(𝐐)G21zz(𝐐,E)\displaystyle=2g^{zz}_{2}(E)J_{s}(\mathbf{Q})G^{zz}_{21}(\mathbf{Q},E)
+2g2zz(E)Jd(𝐐)G11zz(𝐐,E)\displaystyle+2g^{zz}_{2}(E)J_{d}(\mathbf{Q})G^{zz}_{11}(\mathbf{Q},E)
G12zz(𝐐,E)\displaystyle G^{zz}_{12}(\mathbf{Q},E) =2g1zz(E)Js(𝐐)G12zz(𝐐,E)\displaystyle=2g^{zz}_{1}(E)J_{s}(\mathbf{Q})G^{zz}_{12}(\mathbf{Q},E)
+2g1zz(E)Jd(𝐐)G22zz(𝐐,E)\displaystyle+2g^{zz}_{1}(E)J_{d}(\mathbf{Q})G^{zz}_{22}(\mathbf{Q},E)
G22zz(𝐐,E)\displaystyle G^{zz}_{22}(\mathbf{Q},E) =g2zz(E)+2g2zz(E)Js(𝐐)G22zz(𝐐,E)\displaystyle=g^{zz}_{2}(E)+2g^{zz}_{2}(E)J_{s}(\mathbf{Q})G^{zz}_{22}(\mathbf{Q},E)
+2g2zz(E)Jd(𝐐)G12zz(𝐐,E)\displaystyle+2g^{zz}_{2}(E)J_{d}(\mathbf{Q})G^{zz}_{12}(\mathbf{Q},E)

where \hbar has been set to 1, and thus ω\omega being relabeled as E=ωE=\hbar\omega. We note here that we have assumed the different single-ion levels are coupled with the same J(𝐐)J(\mathbf{Q}). We make this assumption for simplicity and test this below against data in the next section. Solving these four coupled equations yields:

G+(𝐐,E)ijGij+(𝐐,E)=g1+(E)+g2+(E)+2g1+(E)g2+(E)[Jd(𝐐)Js(𝐐)][1g1+(E)Js(𝐐)][1g2+(E)Js(𝐐)]g1+(E)g2+(E)[Jd(𝐐)]2,Gzz(𝐐,E)ijGijzz(𝐐,E)=g1zz(E)+g2zz(E)+4g1zz(E)g2zz(E)[Jd(𝐐)Js(𝐐)][12g1zz(E)Js(𝐐)][12g2zz(E)Js(𝐐)]4g1zz(E)g2zz(E)[Jd(𝐐)]2,\displaystyle\begin{split}G^{+-}(\mathbf{Q},E)\equiv\sum\limits_{ij}G^{+-}_{ij}(\mathbf{Q},E)=\frac{g^{+-}_{1}(E)+g^{+-}_{2}(E)+2g^{+-}_{1}(E)g^{+-}_{2}(E)[J_{d}(\mathbf{Q})-J_{s}(\mathbf{Q})]}{[1-g^{+-}_{1}(E)J_{s}(\mathbf{Q})]\cdot[1-g^{+-}_{2}(E)J_{s}(\mathbf{Q})]-g^{+-}_{1}(E)g^{+-}_{2}(E)[J_{d}(\mathbf{Q})]^{2}},\\ \\ G^{zz}(\mathbf{Q},E)\equiv\sum\limits_{ij}G^{zz}_{ij}(\mathbf{Q},E)=\frac{g^{zz}_{1}(E)+g^{zz}_{2}(E)+4g^{zz}_{1}(E)g^{zz}_{2}(E)[J_{d}(\mathbf{Q})-J_{s}(\mathbf{Q})]}{[1-2g^{zz}_{1}(E)J_{s}(\mathbf{Q})]\cdot[1-2g^{zz}_{2}(E)J_{s}(\mathbf{Q})]-4g^{zz}_{1}(E)g^{zz}_{2}(E)[J_{d}(\mathbf{Q})]^{2}},\end{split} (20)

where G+(𝐐,E)G^{-+}(\mathbf{Q},E) has the same form as G+(𝐐,E)G^{+-}(\mathbf{Q},E) with indices ++ \longleftrightarrow -. Here, the energy EE was redefined as E+iδE+i\delta, where δ\delta is a positive infinitesimal to ensure analyticity of gαβg^{\alpha\beta} (Eq. 9). Its value was set to 50% of the experimental elastic resolution width (HWHM) on the ARCS time-of-flight neutron spectrometer (SNS, ORNL) set to the experimental parameters employed by Jain et alJain et al. [2017]

In the T0T\rightarrow 0 K limit where the Bose factor n(E)1n(E)\simeq 1, the fluctuation-dissipation theorem (Eq. 1) is reduced to

S(𝐐,E)G(𝐐,E),S(\mathbf{Q},E)\propto-\Im G(\mathbf{Q},E),

where the magnetic dynamic structure factor is directly proportional to the imaginary component of the total response function G(𝐐,E)G(\mathbf{Q},E). Since the dynamic structure factor is directly proportional to the magnetic neutron cross section d2σdΩdE\frac{d^{2}\sigma}{d\Omega dE} in the T0T\rightarrow 0 K limit, plus the addition of the square of the magnetic form factor, an expression for the unnormalized raw inelastic neutron scattering intensity is given by

I(𝐐,E)𝒜f2(𝐐)G(𝐐,E),I(\mathbf{Q},E)\simeq-\mathcal{A}f^{2}(\mathbf{Q})\Im G(\mathbf{Q},E), (21)

where the magnetic form factor has been approximated by the isotropic magnetic form factor f(𝐐)f(\mathbf{Q}) and the pre-factor 𝒜\mathcal{A} is a scalar corresponding to a combination of conversion and scale constants. The combination of the definition of G(𝐐,E)G(\mathbf{Q},E) (Eq. 12) with its individual components Gαβ(𝐐,E)G^{\alpha\beta}(\mathbf{Q},E) (Eq. 20) reduces the inelastic neutron scattering intensity given in Eq. 21 to a closed-form analytic expression describing the coupling of gαβ(E)g^{\alpha\beta}(E) by J(𝐐)J(\mathbf{Q}). Since gαβ(E)g^{\alpha\beta}(E) (Eq. 9) is function of λ\lambda, HMFH_{MF}, Γ\Gamma, while J(𝐐)J(\mathbf{Q}) (Eq. 10) is a function of J1J_{1}, and J2J_{2}, the analytic expression for the scattering intensity, for a fixed 𝐐\mathbf{Q} and EE, is itself a closed-form function of five distinct parameters, making both parametrization and the subsequent optimization readily amenable to numerical calculation methods. Sarte et al. [2020]

Table 3: Refined parameter values of the spin-orbit exciton model for Ca2RuO4 obtained from a least squares minimization (Eq. 22) performed throughout a 5dd-hyperdimensional parameter space whose individual axes were defined by a range of values roughly centered about each parameter’s initial value. All values are reported in meV and numbers in parentheses indicate calculated errors.
Parameter Initial Value   Range Refined Value
αλ\alpha^{\prime}\lambda 37.5 [0,70] 39(4)
Γ\Gamma 14.9 [0,25] 19(2)
J1J_{1} 1.88 [0,5] 2.1(2)
J2J_{2} 0.425-0.425 [2-2,2] 0.37(6)-0.37(6)
HMFH_{MF} 11.6 [0,25] 0.1(1)0.1(1)

The determination of optimal parameter values for: λ\lambda, HMFH_{MF}, Γ\Gamma, J1J_{1}, and J2J_{2} was accomplished numerically in MATLAB by solving the least squares minimization problem

β=T,L𝐐[E(λ,HMF,Γ,J1,J2)𝐐,βEdata,𝐐,β]2,\sum\limits_{\beta=T,L}\sum\limits_{\mathbf{Q}}\left[E(\lambda,{\rm{H}}_{MF},\Gamma,J_{1},J_{2})_{\mathbf{Q},\beta}-E_{data,\mathbf{Q},\beta}\right]^{2}, (22)

where E()𝐐,βE(\dots)_{\mathbf{Q},\beta} and Edata,𝐐,βE_{data,\mathbf{Q},\beta} denote the calculated and measured energy transfer possessing maximum intensity for a fixed 𝐐\mathbf{Q} and branch β\beta, respectively. The methods used for the determination of the initial values varied significantly from parameter to parameter. In the case of αλ\alpha^{\prime}\lambda and J1J_{1}, the values of 37.5 meV and 1.88 meV simply correspond to their respective values reported in literature. Das et al. [2018]; Mizokawa et al. [2001]; Veenstra et al. [2014]; Fatuzzo et al. [2015]; Cao and DeLong [2013]; Rho et al. [2003, 2005]; Sugai et al. [1990] An initial estimate for the mean molecular field HMFH_{MF} was determined by first extracting the value for i>jzijJij\sum\limits_{i>j}z_{ij}J_{ij} from the experimentally determined Nakatsuji and Maeno [2000a, b] Curie-Weiss temperature θCW90\theta{\rm{{}_{CW}}}\sim-90 K (7.76-7.76 meV) via its mean field definition

θ=CW23S(S+1)i>jzijJij.\theta{\rm{{}_{CW}}}=-\frac{2}{3}S(S+1)\sum\limits_{i>j}z_{ij}J_{ij}.
Refer to caption
Figure 3: (a) Magnetic dynamic structure factor S(𝐐,E)S(\mathbf{Q},E) of Ca2RuO4 calculated in the T0T\rightarrow 0 K limit using a mean-field multilevel spin-orbit exciton model with refined parameters summarized in Tab. 3. (b) Comparison of the dispersion relation for the longitudinal (LL) and transverse modes (TT, TT^{\prime}) calculated using a spin-orbit exciton model (red) and spin-wave theory by Jain et al. Jain et al. [2017] employing a phenomenological Hamiltonian (black). (c) Comparison of 𝐐\mathbf{Q}-integrated cuts of data measured at 5 K by Jain et al.Jain et al. [2017] and calculated using the spin-orbit exciton model for various 𝐐\mathbf{Q} \in [(0,0), (π\pi,0)] along [HH,HH]. For the purposes of comparison, the energy transfer for the maximum of the longitudinal mode that was previously calculated by spin-wave theoryJain et al. [2017] and spin-orbit exciton model have been both labeled explicitly for each 𝐐\mathbf{Q}-integrated cut.
Refer to caption
Figure 4: (a) Individual αβ\alpha\beta components of the magnetic dynamic structure factor SS(𝐐\mathbf{Q},EE) along [HH,0] calculated using the spin-orbit exciton model. (b)-(e) Illustration of the effects of spin-orbit coupling αλ\alpha^{\prime}\lambda, tetragonal distortion Γ\Gamma, next nearest neighbor magnetic exchange J2J_{2}, and the mean molecular field HMFH\rm{{}_{MF}} parameters, respectively, on the total calculated magnetic dynamic structure factor. Unless otherwise stated, the spin-orbit exciton model’s parameters were fixed to their refined values that are listed in Tab. 3.

The insertion of the extracted value of 5.82 meV for i>jzijJij\sum\limits_{i>j}z_{ij}J_{ij} into the definition of HMFH_{MF} (Eq. 5) yields an initial estimate of 11.6 meV for HMFH_{MF}. Furthermore, since the spin-orbit exciton model considered here is restricted to the first two coordination shells, i>jzijJij\sum\limits_{i>j}z_{ij}J_{ij} is reduced to 4J1+4J24J_{1}+4J_{2}. By combining the extracted value of 5.82 meV for i>jzijJij\sum\limits_{i>j}z_{ij}J_{ij} with the initial estimate of 1.881.88 meV for J1J_{1}, the value of 0.425-0.425 meV is obtained for an initial estimate of J2J_{2}. A negative value whose magnitude is significantly smaller than J1J_{1}, consistent with a system that assumes long range antiferromagnetic order in the mean field limit. Finally, the initial estimate of 14.9 meV for Γ\Gamma was determined by the scaling of the distortion parameter reported for KCoF3 Buyers et al. [1971] by an empirical factor of 0.02/0.00197=10.15 corresponding to the ratio of their respective tetragonal distortions δc/c\delta c/c, while its positive sign corresponds to the restriction previously established in the point charge calculation that Γ>0\Gamma>0.

As is the case for all derivative-free methods, including the simplex search method specifically employed by MATLAB, the parameter values determined by the minimization algorithm do not necessarily yield the global minimum, or even a local minimum. This is particularly true when the initial values are too far removed from the true optimal values. As an attempt to address such a concern, the least squares minimization problem given by Eq. 22 was solved for various initial test values for the five parameters. These test values define a hyperdimensional 10d10^{d}, d=5d=5 parameter space, where each axis corresponds to a linear distribution of 10 values over a specified range that is roughly centered about the specific parameter’s initial value. The set of optimized values for these five parameters that yield the minimum of the 10510^{5} solutions to the least squares minimization problem was defined as parameters’ refined values. The initial values, distribution ranges in the 55 dimensional parameter space, and the final refined values of the five parameters are summarized in Tab. 3.

IV Calculated Results

Having established the underlying theoretical foundation and the corresponding physical parameters that constitute our model, we now present a direct comparison of the experimental data reported by Jain et al. Jain et al. [2017] to our calculated parametrization based on the spin-orbit exciton approach.

As illustrated in Fig. 3(a), by employing the refined parameters in Tab. 3 that were obtained through a least squares minimization, the spin-orbit exciton model yields two distinct modes (denoted as LL and TT). We note that we will address the TT^{\prime} mode present in Fig. 3 at the end of this section. Fig. 4(a) illustrates that the strongly dispersive low energy TT mode corresponds to transverse fluctuations along the (a,c)(a,c) plane (αβ\alpha\beta = ++- and +-+), while the longitudinal zzzz fluctuations along the bb axis uniquely constitute the second LL mode located at higher energy transfers. The dispersion relation for both modes throughout the Brillouin zone presented in Fig. 3(b) are in excellent agreement with their respective counterparts previously calculated by Jain et al. Jain et al. [2017] applying linear spin-wave theory to a phenomenological Hamiltonian in the Γλ\Gamma\ll\lambda limit. As illustrated in Fig. 5, constant energy and (𝐐,E)(\mathbf{Q},E) slices along select high symmetry directions identified the presence of minor discrepancies between the calculated model and experimental data that are predominately limited to a region in the Brillouin zone between (π,0)(\pi,0) and (π2,π2)\left(\frac{\pi}{2},\frac{\pi}{2}\right) along [0,K][0,-K] at energy transfers \sim35-40 meV. Such discrepancies were previously noted by Jain et al.Jain et al. [2017] and it is suspected that in the case of the current model, this particular discrepancy may stem from a complex further neighbor exchange that has not been accounted for in our J(𝐐)J({\bf{Q}}) (Eqs. 18 and 19) which has been restricted to employing isotropic magnetic exchange constants that span only over the first two coordination shells in the (H,K)(H,K) plane of the pseudo-tetragonal unit cell.

The influences of the spin-orbit exciton model’s individual parameters, each corresponding to physically measurable quantities, on both the TT and LL modes are summarized in Figs. 4(b)-(e). By defining the splitting in energy between different jeffj\rm{{}_{eff}} manifolds of ^SO\hat{\mathcal{H}}\rm{{}_{SO}} (Fig. 1(b)), αλ\alpha^{\prime}\lambda determines the energy scale of interest. In the case of a fixed value for αλ\alpha^{\prime}\lambda, the magnetic exchange constants J1J_{1} and J2J_{2} both determine the modes’ dispersion relation and bandwidth, while the value of the tetragonal distortion Γ\Gamma dictates the gap in energy between the TT and LL modes. In the case of Γ\Gamma, a positive value yields an LL mode higher in energy relative to TT, while a negative value simply reverses the order. Finally, for fixed values of αλ\alpha^{\prime}\lambda, Γ\Gamma, J1J_{1}, and J2J_{2}, the magnitude and sign of the mean molecular field HMFH_{MF} determines the separation and relative order in energy, respectively, between the individual components (++- and +-+) that constitute the transverse TT mode.

Refer to caption
Figure 5: (a) Comparison of constant energy slices in the (H,K)(H,K) plane measured at 5 K by Jain et al. Jain et al. [2017] and calculated using the spin-orbit exciton model with corresponding (𝐐\mathbf{Q},EE)-slices along the (b) [HH,0], (c) [HH,12\frac{1}{2}], (d) [HH,HH], and (e) [HH,HH-11] high symmetry directions. For the purposes of clarity, the origin in reciprocal space for each (𝐐,E)(\mathbf{Q},E) slice has been stated explicitly.
Refer to caption
Figure 6: Comparison of the dispersion relation calculated using the spin-orbit exciton model (red) and spin-wave theory Jain et al. [2017] employing ^phen\hat{\mathcal{H}}\rm{{}_{phen}} (black) with the kinematically permissable (𝐐\mathbf{Q},EE) region for the two-magnon continuum.

As summarized in Tab. 3, the refined values for four (out of the five) parameters: αλ\alpha^{\prime}\lambda, Γ\Gamma , J1J_{1}, and J2J_{2} are in good agreement with their initial values. While the refined value of 39(4) meV for αλ\alpha^{\prime}\lambda agrees within error with the value previously deduced from RIXS, the slight deviation of the refined tetragonal distortion parameter Γ\Gamma of 19(2) meV from its initial value of 14.9 meV may be attributed to the oversimplification of the reportedBraden et al. [1998] monoclinic distortion into one that is bound uniaxially along cc. In the case of the magnetic exchange constants: J1J_{1} and J2J_{2}, both the magnitude and sign of their refined values agree with their respective initial values. With refined values J1J_{1} = 1.88 meV >|J2=0.425|>|J_{2}=-0.425| meV, mean field theory suggests that Ca2RuO4 would assume antiferromagnetic long range order at TN23S(S+1){4J1+4J2}=108(2)T_{N}\sim\frac{2}{3}S(S+1)\{4J_{1}+4J_{2}\}=108(2) K, all consistent with both previously reported physical property and neutron diffraction measurements.Braden et al. [1998]; Nakatsuji and Maeno [2000b, a]; Cao et al. [1997]; Bertinshaw et al. [2019] Intuitively, the presence of long range antiferromagnetic order in Ca2RuO4 at first appears somewhat perplexing considering the presence of a j=eff0j\rm{{}_{eff}}=0 singlet ground state for ^SO\hat{\mathcal{H}}_{SO}. An explanation for the apparent contradiction is the presence of a tetragonal distortion that enables coupling and admixture between higher lying spin-orbit manifolds and the singlet ground state. Furthermore, with a value for the tetragonal distortion parameter Γ=19(2)\Gamma=19(2) meV being smaller than the energy scales of interest defined by αλ\alpha^{\prime}\lambda, the spectrum would appear to originate from a system consisting of a non-distorted j=eff0j\rm{{}_{eff}}=0. This seemingly non-magnetic singlet ground state would be consistent with the refined value of HMFH_{MF}=0 meV instead of the initial mean field value listed in Tab. 3 that was obtained from the Curie-Weiss temperature θCW\theta\rm{{}_{CW}}.

It was this presence of a j=eff0j\rm{{}_{eff}}=0 singlet ground state that made Ca2RuO4 such an attractive candidate in the search of a condensed-matter analog of the much-celebrated Higgs mode. Higgs [1964] As was previously determined in the orignal study by Jain et al.,Jain et al. [2017] our spin-orbit exciton model produced a longitudinally polarized LL mode that remains well-defined throughout the Brillouin zone (Figs. 3(a,b)). Possessing 33(5)% of the intensity of the corresponding transverse mode at 𝐐=(0,0)\mathbf{Q}=(0,0) (Fig. 3(c)), this LL mode corresponds to amplitude fluctuations of the magnetic moment of a system of interacting spins that is located near a quantum critical point, consistent with what one would expect for the Higgs mode.

The importance for the presence of the Higgs mode analog in a condensed matter system is that it provides a unique platform in the study of the decay processes of a particle (and its properties through inference) that has been postulated to play a key role in the determination of masses in the Standard model. According to earlier theoretical treatments, it has been postulated that the Higgs mode decays into a pair of Goldstone modes. Munehisa [2015]; Rose et al. [2015] To pursue such a possibility, the kinematically accessible phase space permitted for such a decay process was calculated based on energy and momentum conservation given by Huberman et al. [2005]; Stock et al. [2018]; Songvilay et al. [2018]

G(𝐐,E)=𝐐1,𝐐2δ(𝐐𝐐1𝐐2)δ(EE𝐐1E𝐐2),\begin{split}&G(\mathbf{Q},E)=\\ &\sum\limits_{\mathbf{Q}_{1},\mathbf{Q}_{2}}\delta(\mathbf{Q}-\mathbf{Q}_{1}-\mathbf{Q}_{2})\delta(E-E_{\mathbf{Q}_{1}}-E_{\mathbf{Q}_{2}}),\end{split}

where E𝐐1,2E_{\mathbf{Q}_{1,2}} are the energies of transverse excitations at a given momentum transfer 𝐐\mathbf{Q}. As illustrated in Fig. 6, the kinematically allowed region overlaps in both momentum and energy with the antiferromagnetic wave vector (π,π)(\pi,\pi). The predicted overlap naturally lends itself to the possibility of coupling between the Higgs LL mode and the “multi-magnon” continuum. Such coupling was required in the previous theoretical treatment to address the broad scattering that was experimentally observed at (π,π)(\pi,\pi), yet clearly absent at (0,0)(0,0).

Refer to caption
Figure 7: (a) Magnetic dynamic structure factor S(𝐐,E)S(\mathbf{Q},E) of Ca2RuO4 calculated using the spin-orbit exciton model in the T0T\rightarrow 0 K limit exhibits a high energy mode (EE\sim116 meV) associated with the dipolar forbidden j=eff02j{\rm{{}_{eff}}}=0\rightarrow 2 transition. (b) 𝐐\mathbf{Q}-integrated cuts at various high symmetry points throughout the Brillouin zone reveal that the mode is minimally dispersive, (c) in close agreement with a spin-orbital excitation previously measured with RIXS.Das et al. [2018]

As illustrated in Fig. 7, the extension of the spin-orbit exciton model to higher energy transfers (E0.1E\gtrsim 0.1 eV) identified an additional transverse mode centered at E116E\sim 116 meV. Corresponding to dipolar forbidden j=eff0j=eff2j{\rm{{}_{eff}}}=0\rightarrow j\rm{{}_{eff}}=2 transitions, the minimally dispersive mode exhibits an intensity three orders of magnitude lower than the corresponding TT mode at lower energy transfers. We note that while this mode is dipolar forbidden in the case of a perfect octahedral field, its corresponding transition is allowed in the case of Ca2RuO4 owing to the presence of a weak structural distortion. Such a spin-orbital excitonic origin is consistent with previous theoretical approaches Das et al. [2018]; Souliou et al. [2017] addressing a minimally dispersive mode of magnetic origin that was measured at approximately the same energy transfer (Fig. 7(c)) with both RIXS and Raman spectroscopy.

Refer to caption
Figure 8: (a) Effective (second) Brillouin zone (b) backfolded onto the first Brillouin zone demonstrates that Γ(π,π)M(π,0)\Gamma_{(\pi,\pi)}\rightarrow M_{(\pi,0)} (blue) in the second Brillouin zone is symmetrically equivalent to Γ(0,0)M(π,0)\Gamma_{(0,0)}\rightarrow M_{(\pi,0)} (green) in the first Brillouin zone, Γ(0,0)Γ(π,π)\Gamma_{(0,0)}\rightarrow\Gamma_{(\pi,\pi)} (brown) is reversed, while M(π,0)X(π2,π2)M_{(\pi,0)}\rightarrow X_{\left(\frac{\pi}{2},\frac{\pi}{2}\right)} (gray) is unaffected. (c) New high symmetry directions between high symmetry points of the antiferromagnetic reciprocal unit cell (red) that has been symmetrized with respect to the square lattice (black) and whose zone center is now at Γ(π,π)\Gamma_{(\pi,\pi)}. Colors in (b) and (c) illustrate the relationship between specific dispersion relations for T and T modes.

We now shift the discussion to address the additional transverse TT^{\prime} mode present in Fig. 3. Despite all the success of the spin-orbit exciton model to account for the experimental data presented so far (Figs. 3(a,b) and 5), constant-𝐐\mathbf{Q} cuts (Fig. 3(c)) revealed that the combined intensities of the LL and TT modes alone could not account for all the scattering intensity that was observed experimentally. Closer inspection of the unpolarized and polarized inelastic spectra previously measured on ARCS and PUMA, Jain et al. [2017] respectively, revealed that the discrepancy in intensity was limited to transverse fluctuations and were extended throughout the Brillouin zone with energy transfers other than the TT and LL modes, including the presence of additional weak scattering present at the antiferromagnetic center (π,π)(\pi,\pi), confirming the need for an additional transverse mode, termed TT^{\prime}, that is absent in the spin-orbit exciton model.

The particular mode is weak in intensity, where its presence at both low energy and high energy transfers in Fig. 5 was concealed by the dominant TT and LL modes, respectively. Possessing both an identical energy bandwidth and clear similarities to the dispersion relation exhibited by the TT mode throughout the entire Brillouin zone, the mode in question is consistent with one that originates from backfolding onto the first Brillouin zone, as originally proposed by Jain et al. Jain et al. [2017] Fig. 8 illustrates that backfolding yields a mode whose dipsersion along Γ(π,π)M(π,0)\Gamma_{(\pi,\pi)}\rightarrow M_{(\pi,0)} is replaced by Γ(0,0)M(π,0)\Gamma_{(0,0)}\rightarrow M_{(\pi,0)} (and vice versa), the dispersion along Γ(0,0)Γ(π,π)\Gamma_{(0,0)}\rightarrow\Gamma_{(\pi,\pi)} is reversed, while the dispersion along M(π,0)X(π2,π2)M_{(\pi,0)}\rightarrow X_{\left(\frac{\pi}{2},\frac{\pi}{2}\right)} is unaffected. Such a mode is illustrated in Fig. 3(b), demonstrating clear agreement with the dispersion relation reported in the experimental data, supporting the attribution of the TT^{\prime} mode to backfolding. By considering the scattering intensity of the TT^{\prime} mode as originating from a reciprocal unit cell that is identical to that illustrated in Fig. 2, but whose nuclear zone center coincides with Γ(π,π)\Gamma_{(\pi,\pi)}, the spin-orbit exciton model accounts for a significant portion of the intensity deficiency first identified in Fig. 3(c).

V Discussion and Concluding Remarks

To summarize, by extending previous theoretical approaches,Buyers et al. [1975]; Sarte et al. [2019] we have established the theoretical framework for a spin-orbit exciton model that accounts for the low energy magnetic excitations in the layered perovskite Ca2RuO4. In particular, the model successfully reproduces the longitudinal Higgs excitation. This mode has been theoretically predicted near quantum critical points in bilayer magnets, Chubukov and Morr [1995] and has been experimentally identified in a plethora of systems spanning condensed matter physics including dimerized TlCuCl3 near a pressure induced quantum critical point, Rüegg et al. [2008]; Merchant et al. [2014] and in superconductors Anderson [2015]; Sherman et al. [2015] including 2H-NbSe2Méasson et al. [2014]

In contrast to conventional pseudo-bosonic approaches that employ the Holstein-Primakoff transformation,Majlis [2007] the spin-orbit exciton approach presented here employs a minimalist Hamiltonian (Eq. 3), enabling not only the explicit, but also equal weighted incorporation of the individual contributions to the crystalline electric field with magnetic superexchange. As illustrated in Fig. 4, a clear advantage for such an approach is that it facilitates the quantization of the influences for each individual contribution. By understanding how each term in the Hamiltonian influences the excitation spectrum, the spin-orbit exciton model represents a tool that may be used to identify possible candidates possessing the appropriate conditions to host exotic excitations, such as the Higgs mode.

In case of the longitudinally polarized Higgs mode, its existence is precluded when no orbital degree of freedom is present, such as is the case for S=52S=\frac{5}{2} in an undistorted octahedral weak ligand field. In this case, orthogonality ensures that matrix elements of the form m|Jz|0\langle m|J_{z}|0\rangle are zero, and thus by Eq. 9 ensuring gzzg_{zz} is zero as well. In the case of Ca2RuO4, an orbital degree of freedom is introduced in the spin-orbit exciton model by spin-orbit coupling ^SO\hat{\mathcal{H}}_{SO} (Eq. 14), enabling mixing of the crystalline electric field eigenstates. Such an effect can be rationalized by noting that the inclusion of a 𝐥𝐒\mathbf{l}\cdot\mathbf{S} term in the total magnetic Hamiltonian \mathcal{H} results in a non-zero [Sz,][S_{z},\mathcal{H}], and by the Heisenberg equation-of-motion, Sz\langle S_{z}\rangle is not conserved, providing the possibility of a longitudinal mode to exist. Even in the case that zzzz mode does exist, its dispersion is not necessarily unique compared to its transverse ++- and +-+ counterparts, as is the case for an undistorted octahedral ligand field. As summarized in Fig. 4, it is the introduction of a uniaxial distortion ^dis\hat{\mathcal{H}}_{dis} (Eq. III.1) along zz that results in the separation of the zzzz mode from the transverse modes, with the sign of Γ\Gamma determining the relative order in energy. This term in the Hamiltonian also controls the magnetic anisotropy which has been suggested to be important for stabilizing the Higgs mode. Su et al. [2020]

In addition to a structural distortion dis\mathcal{H}_{dis}, another contribution that has a particularly strong influence on the relative separation between the longitudinal and transverse branches is the molecular mean field. Having a disproportionately larger influence on the transverse ++- and +-+ modes relative to their longitudinal counterpart, the value and sign of ^MF\hat{\mathcal{H}}_{MF} (Eq. 4) determines the relative separation of the two transverse modes. An interesting result is that the least squares optimization of the spin-orbit exciton model yielded a zero molecular field HMF(i)H_{MF}(i) for all sites ii. This is consistent with the single-ion prediction of a jeffj\rm{{}_{eff}}=0 ground state, but is not consistent with the presence of an ordered magnetic moment characterized by a magnetic Bragg peak in the neutron scattering response.Braden et al. [1998]; Jain et al. [2017]; Pincini et al. [2018] This discrepancy can be reconciled by noting that the value of molecular field was determined by fitting excitations with energies equal or greater than the distortion energy, and thus, at the energy transfers of interest, Ca2RuO4 could be approximated as a pure j=eff0j\rm{{}_{eff}}=0 magnet. One possible origin for the presence of an elastic Bragg peak in the neutron response despite a singlet ground state is the reportedBraden et al. [1998] presence of a prominent distortion of the local octahedral coordination environment. However, other theoretical ideas have been proposed that may also be consistent with our refinement of the molecular field, including Hund’s coupling Svoboda et al. [2017] and triplon condensation. Chen and Balents [2011]

In the discussion presented so far, key inferences concerning the properties of the magnetic excitation spectrum of Ca2RuO4 have been limited to the separate distinct influences for each of the spin-orbit exciton model’s individual parameters. By considering a combination of these parameters, inferences with possibly large widespread applicability may be addressed, with one such example would be quantum criticality. A quantum critical point has been predicted to exist when the longitudinal response can be driven to zero energy Rancon and Dupuis [2014] or when the longitudinal and transverse modes become degenerate. Oitmaa [2018] In such a situation the spectral weight in the longitudinal channel would be expected to diverge. This can also be seen by applying the “multi-magnon” formalism discussed in the case of classical two-dimensional magnets. Huberman et al. [2005] As illustrated in Fig. 4, the degeneracy of the longitudinal and transverse branches occurs in the absence of a distortion away from an ideal octahedra arrangement, while the shift of the zzzz mode to lower energy transfers can be accomplished by an appropriate ratio of the exchange constants J1J_{1}, J2J_{2} to the spin-orbit coupling constant λ\lambda. Possible mechanisms to achieve such conditions could involve strain or pressure, whose influence is not only restricted to the structural distortion of local coordination octahedra,Ding et al. [2006] but extends to the magnetic superexchange constants.Blanco-Canosa et al. [2007]; Zhang et al. [2006]; Rocquefelte et al. [2012]

In conclusion, we have established the theoretical framework for a spin-orbit exciton model where the use of a minimalist Hamiltonian enables for the direct and equal weighted incorporation of the individual contributions to the crystalline electric field with magnetic superexchange. Such an excitonic approach was then used to model and understand the magnetic excitations originating from the coupled spin and crystalline electric fields of a canonical Van Vleck j=eff0j\rm{{}_{eff}}=0 ground state in Ca2RuO4. The anomalous longitudinally polarized Higgs mode,Jain et al. [2017] transverse spin-waves,Kunkemöller et al. [2015] and orbital excitationsDas et al. [2018] were successfully captured, and in good agreement with previously reported neutron and inelastic x-ray spectroscopic measurements. The framework established here illustrates how a crystalline electric field-induced singlet ground state can support coherent longitudinal excitations, and transverse wavelike dynamics, while providing possible mechanisms for accessing a nearby quantum critical point.

VI Acknowledgements

We acknowledge useful conversations with W.J.L. Buyers, R.A. Cowley, H. Lane, K.J. Camacho, C. Schwenk, Y. Wolde-Mariam, and A. Reyes. P.M.S. and B.R.O. acknowledge financial support from the University of California, Santa Barbara through the Elings Prize Fellowship. C.S. and K.H.H. would like to acknowledge the ERC, the EPSRC, the STFC, and the Carnegie Trust for the Universities of Scotland for financial support. Finally, this material is based upon work supported by the National Science Foundation’s Q-AMASE-i initiative under award DMR-1906325.

References