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

Fragment-orbital-dependent spin fluctuations in the single-component molecular conductor [Ni(dmdt)2]

Taiki Kawamura and Akito Kobayashi Department of Physics, Nagoya University, Nagoya, Aichi 464-8602, Japan
Abstract

Motivated by recent nuclear magnetic resonance experiments, we calculated the spin susceptibility, Knight shift, and spin-lattice relaxation rate (1/T1T1/T_{1}T) of the single-component molecular conductor [Ni(dmdt)2] using the random phase approximation in a multi-orbital Hubbard model describing the Dirac nodal line electronic system in this compound. This Hubbard model is composed of three fragment orbitals and on-site repulsive interactions obtained using ab initio many-body perturbation theory calculations. We found fragment-orbital-dependent spin fluctuations with the momentum q=0 and an incommensurate value of the wavenumber q=Q at which a diagonal element of the spin susceptibility is maximum. The q=0 and Q responses become dominant at low and high temperatures, respectively, with the Fermi-pocket energy scale as the boundary. We show that 1/T1T1/T_{1}T decreases with decreasing temperature but starts to increase at low temperature owing to the q=0 spin fluctuations, while the Knight shift keeps monotonically decreasing. These properties are due to the intra-molecular antiferromagnetic fluctuations caused by the characteristic wave functions of this Dirac nodal line system, which is described by an nn-band (n3n\geq 3) model. We show that the fragment orbitals play important roles in the magnetic properties of [Ni(dmdt)2].

I introduction

Dirac electron systems in solids are of interest to many researchers because of not only their quantum transport phenomena T.Ando1998 ; V.P.Gusynin2006 ; S.Murakami2004 ; D.Hsieh2008 and large diamagnetism, H.Fukuyama1970 ; H.Fukuyama2007 but also their unusual effects induced by the Coulomb interaction.A.A.Abrikosov1971 ; J.Gonzalez1994 ; J.Gonzalez1999 ; V.N.Kotov2012 ; T.O.Wehling2014 ; W.Witczak-Krempa2014

Dirac electron systems in molecular conductors, such as α\alpha-(BEDT-TTF)2I3, provide suitable platforms for studying the effect of interaction because the electron hopping integrals between neighboring molecules are smaller than the on-site repulsive interactions reflecting the weak inter-molecular coupling. K.Kajita1992 ; N.Tajima2000 ; A.Kobayashi2004 ; S.Katayama2006 ; A.Kobayashi2007 ; M.O.Goerbig2008 ; K.Kajita2014 At high pressure, α\alpha-(BEDT-TTF)2I3 is a massless Dirac electron system. However, at low pressure, a charge-ordered state appears presumably due to nearest-neighbor Coulomb repulsions, H.Seo2000 ; T.Takahashi2003 ; T.Takiuchi2007 where anomalous spin–charge separation on spin gaps K.Ishibashi2016 ; Y.Katayama2016 and transport phenomena occur.R.Beyer2016 ; D.Lu2016 ; D.Ohki2019 In addition, the long-range Coulomb interaction reshapes the Dirac cone because of a logarithmic velocity renormalization, which induces an anomalous magnetic response.M.Hirata2016 ; M.Hirata2017 ; A.Kobayashi2013 Moreover, ferrimagnetism and spin-triplet excitonic fluctuations are observed.D.Ohki2020 ; M.HirataRPP

The Dirac electron system in α\alpha-(BEDT-TTF)2I3 is two-dimensional K.Kajita2014 because it is a layered molecular conductor and the hopping of electrons from one conducting layer to the neighboring one over the insulating anion layer is incoherent. By contrast, if the electron hopping perpendicular to the main conducting layer were coherent, the Dirac point would be connected and draw lines (rings) in the three-dimensional momentum space, which are called the Dirac nodal lines (rings). A.Burkov2011 ; C.-K.Chiu2014 ; C.Fang2015 ; Z.Gao2016

Such kinds of Dirac nodal line (ring) systems have indeed been found in graphite, P.R.Wallace1947 transition-metal monophosphates,H.Weng2015 Cu3N,Y.Kim2015 antiperovskites,R.Yu2015 perovskite iridates,J.-M.Carter2012 and hexagonal pnictides with the composition CaAgX (X = P, As),A.Yamakage2016 as well as in the single-component molecular conductors [Pd(dddt)2], R.Kato2017_A ; R.Kato2017_B ; Y.Suzumura2017_A ; Y.Suzumura2017_B ; Y.Suzumura2018_A ; Y.Suzumura2018_B ; T.Tsumuraya2018 ; Y.Suzumura2019 [Pt(dmdt)2],B.Zhou2019 ; R.Kato2020 ; T.Kawamura2020 ; T.Kawamura2021 ; A.Kobayashi2021 and [Ni(dmdt)2].T.Kawamura2021 ; A.Kobayashi2021

The Dirac nodal line (ring) systems exhibit not only the properties in common with two-dimensional Dirac electron systems, e.g., the in-plane conductivityB.Zhou2019 , but also the characteristic electronic properties such as non-dispersive Landau levels,J.-W.Rhim2015 Kondo effect,A.K.Mitchell2015 quasi-topological electromagnetic responses,S.T.Ramamurthy2017 and edge magnetismT.Kawamura2021 because of the three-dimensionality. However, the electron correlation effects on the Fermi surface in the Dirac nodal line systems have not yet been elucidated.

The prime focus of this study is such Dirac nodal line system in [M(dmdt)2] (M = Pt, Ni), which is a single-component molecular conductor that consists of the M(dmdt)2 molecules, where the bracket [\cdots] stand for a crystal. This material is a triclinic system, as shown in Fig. 1, and has space-inversion symmetry. One unit cell contains one M(dmdt)2 molecule. In previous studies, the electronic properties of [M(dmdt)2] were studied using density functional theory (DFT), and tight-binding models were constructed on the basis of the extended Hückel method and DFT.B.Zhou2019 ; R.Kato2020 ; T.Kawamura2020 ; T.Kawamura2021 These investigations showed that [M(dmdt)2] is a Dirac nodal line system. Furthermore, electronic resistivity measurements using conventional four-probe methods were performed and showed that the resistivity of [M(dmdt)2] hardly depends on the temperature (TT), which is consistent with the property of the Dirac electron system.B.Zhou2019 That is the universal conductivity.N.H.Shon1998 In addition, we previously suggested that the nesting between the Fermi arcs localized at the edge and the electronic correlation induce a helical spin density wave (SDW) at the edge.T.Kawamura2021

Refer to caption
Refer to caption
Figure 1: (a) Crystal structure in the bbcc plane of [M(dmdt)2]. This material consists of M(dmdt)2 molecules. (b) Crystal structure along to the bb-axis. The red, blue, brown, and cyan balls represent Ni, S, C, and H atoms, respectively. The black frames represent the unit cell.

Recently, the spin-lattice relaxation rate 1/T1T1/T_{1}T, probing the low-energy spin dynamics, and the Knight shift, scaling to the spin susceptibility, of [Ni(dmdt)2] were observed in a 13C nuclear magnetic resonance (13C-NMR) experiment.T.Sekine.Private At high temperature, 1/T1T1/T_{1}T decreases with cooling and is almost proportional to T2T^{2}. However, at low temperature, it starts to increase with decreasing temperature and exhibits a peak structure at approximately 3030 K. Meanwhile, the Knight shift is almost proportional to TT because of the linear energy dispersion and does not increase. The mechanism of this anomalous temperature dependence of the spin fluctuations has not been elucidated.

In the present study, we theoretically investigate the electron correlation using the Fermi surface in [Ni(dmdt)2] to elucidate the mechanism of this anomalous temperature dependence of the spin fluctuations. We calculate the spin susceptibility, Knight shift, and 1/T1T1/T_{1}T using the random phase approximation (RPA) in a three-orbital Hubbard model describing [Ni(dmdt)2], which is obtained using ab initio many-body perturbation theory calculations.

The electronic state of a molecular conductor is described by the molecular orbitals, which are linear combinations of the atomic orbitals in a molecule. The molecular orbital that has the highest energy and is fully occupied by electrons is called the highest occupied molecular orbital (HOMO), whereas the one having the lowest energy with no electrons is called the lowest unoccupied molecular orbital (LUMO). HOMO and LUMO are also called frontier orbitals. The electronic states of single-component molecular conductors, e.g., [M(tmdt)2] (M = Ni, Au, Cu) and [M(dmdt)2] (M = Pt, Ni), are described by multiple molecular orbitals localized in a part of the molecule.H.Seo2008 ; H.Seo2013 ; M.Tsuchiizu2012 ; T.Kawamura2020 ; T.Kawamura2021 These molecular orbitals are the energy eigenstates obtained using first-principles calculations and are called “fragment orbitals”. The fragment orbitals are transformed into HOMO and LUMO by a high-symmetry unitary transform.

Based on the band parameters determined from first-principle calculations, Seo et al. have constructed a Hubbard model of [M(tmdt)2](M=Ni, Au, Cu), which is described by the fragment orbitals.H.Seo2008 ; H.Seo2013 The on-site repulsion acts between the same fragment orbitals that have spins of opposite signs, which is similar to the case of the present study. They have investigated the ordered state by calculating the electron density and spin density using mean-field approximation. By contrast, in this study, we will investigate the spin fluctuations by calculating the spin susceptibility on the basis of RPA. Furthermore, we show that the idea of the fragment orbitals is important for the physical properties of the Dirac nodal line systems in the single-component molecular conductor. As the other previous study of the fragment orbitals, some charge-transfer complexes such as (TTM-TTP)I3 are also modeled by fragment orbitals.M.Tsuchiizu2011 ; M.Tsuchiizu2012

We found that the q=0 spin fluctuations are enhanced in two out of the three fragment orbitals, while an enhancement at an incommensurate wavenumber vector develops in the third orbital. Detailed analysis showed that these q=0 spin fluctuations do not correspond to a simple ferromagnetic fluctuations; rather, they are linked to an intra-molecular antiferromagnetic fluctuations. This implies that the spins of the fragment orbitals within the same molecule are inversely correlated. Further, q=0 implies a direct correlation of the spins between molecules. Using RPA, we determined that the 1/T1T1/T_{1}T starts to increase at a low temperature by the q=0 spin fluctuations. By contrast, the Knight shift does not increase upon cooling because of the intra-molecular antiferromagnetic fluctuations.

At high temperature, an incommensurate spin fluctuations dominate the temperature dependence of 1/T1T1/T_{1}T. These magnetic responses are associated with the geometry of the Fermi surface and the characteristic wave functions of the nn-orbital (n3n\geq 3) Dirac nodal line system. Thus, it is expected that other Dirac nodal line systems described by multiple-orbital models may have similar magnetic properties.

The remainder of this paper is organized as follows. In Section II, we introduce the spin susceptibility based on RPA and formulate 1/T1T1/T_{1}T and the Knight shift. In Section III, we calculate the band structure, spin susceptibility, Knight shift, 1/T1T1/T_{1}T, and so on in the absence of interaction. In Section IV, we calculate the Stoner factor, Knight shift, and 1/T1T1/T_{1}T in the presence of interaction by applying RPA to a Hubbard model. Section V draws conclusions.

II Formulation

We calculate the spin susceptibility, which incorporates the electron correlation effect within perturbation theory to investigate the enhancement of spin fluctuations in [Ni(dmdt)2]. Furthermore, we calculate the Knight shift and the spin-lattice relaxation rate 1/T1T1/T_{1}T, which are the physical quantities observed using NMR. We apply RPA to the Hubbard model to calculate the spin susceptibility in [Ni(dmdt)2]. Although calculations incorporating the self-energy, e.g., FLEX, are better approximations than RPA, RPA is more suitable for investigating spin fluctuations because the self-energy suppresses spin susceptibility.

The Hubbard Hamiltonian that we employ is given by

H=i,α;j,β,σti,α;j,βci,α,σcj,β,σ+i,αUαni,α,ni,α,,H=\sum_{\left<i,\alpha;j,\beta\right>,\sigma}t_{i,\alpha;j,\beta}c^{\dagger}_{i,\alpha,\sigma}c_{j,\beta,\sigma}+\sum_{i,\alpha}U_{\alpha}n_{i,\alpha,\uparrow}n_{i,\alpha,\downarrow}, (1)

where ii and jj are the unit-cell indices, and σ\sigma is the spin index. Here, ti,α;j,βt_{i,\alpha;j,\beta} is a transfer integral defined between the orbital α\alpha in the unit cell ii and the orbital β\beta in the cell jj, and UαU_{\alpha} represents the on-site repulsive interaction on the orbital α\alpha, with α\alpha and β\beta standing for one of the three fragment orbitals in the unit cell (A, B, and C in Fig. 2). The indices α\alpha and β\beta represent the three fragment orbitals A, B, and C. \sum_{\left<\cdots\right>} represents a summation that runs only for the hoppings that have a large energy scale than the cutoff (set to be 0.0100.010 eV in this study).

The electronic states of [Ni(dmdt)2] near the Fermi energy EFE_{F} are described by three fragment-decomposed Wannier orbitals(fragment orbitals) dubbed orbitals A, B, and C as illustrated in Fig. 2. They are obtained using Wannier fitting to three isolated energy bands near EFE_{F}, which were previously obtained using first-principles calculations. T.Kawamura2021

The Wannier fitting and first-principles calculations were performed using the programs respackK.Nakamura2020 and Quantum EspressoP.Giannozzi2017 , respectively.T.Kawamura2021 respack was also used for calculating the Coulomb interaction and other factors. Further, Quantum Espresso was used for first-principles calculations based on the pseudopotential method. Figure 2(a) and (b) show a side and vertical-axis view, respectively, of the molecule. The orbital B in Fig. 2(a) may look like a dd orbital, but is a pp orbital of the S atoms, as is evident from Fig. 2(b). The dd orbital of Ni is localized near the Ni atom, and its contribution to the orbital B is small. The orbitals A and C have pp-orbital-like shapes and are equivalent because of space-inversion symmetry. In the present study, we assume that these three fragment orbitals sit in the same molecule and in the same unit cell.

Refer to caption
Refer to caption
Figure 2: Schematic of a Ni(dmdt)2 molecule and the fragment orbitals A, B, and C. The red, blue, brown, and cyan balls represent Ni, S, C, and H atoms, respectively. The red dots in orbitals A, B, and C indicate the location of the Ni atom as a guide to the eye. (a) Side view of the molecule. The black solid line represents the Ni(dmdt)2 molecule. (b) Vertical-axis view of the molecule. The red dashed circles indicate C atoms, which are replaced with 13C in the 13C-NMR experiment.T.Sekine.Private These figures are plotted by VESTA. Different colors represent different signs of the wave functions. The cut-off of the normalized wave functions is 0.01.

By performing a Fourier transform, the Hamiltonian (Eq. 1) is rewritten as

H=k,α,β,σHαβ,σ0(k)ck,α,σc,k,β,σ+1NLk,k,q,αUααck+q,α,ckq,α,ck,α,ck,α,,\begin{split}H&=\sum_{\textbf{k},\alpha,\beta,\sigma}H^{0}_{\alpha\beta,\sigma}(\textbf{k})c^{\dagger}_{\textbf{k},\alpha,\sigma}c_{,\textbf{k},\beta,\sigma}\\ &+\frac{1}{N_{L}}\sum_{\textbf{k},\textbf{k}^{\prime},\textbf{q},\alpha}U_{\alpha\alpha}c^{\dagger}_{\textbf{k}+\textbf{q},\alpha,\uparrow}c^{\dagger}_{\textbf{k}^{\prime}-\textbf{q},\alpha,\downarrow}c_{\textbf{k}^{\prime},\alpha,\downarrow}c_{\textbf{k},\alpha,\uparrow},\end{split} (2)

where k, k\textbf{k}^{\prime}, and q are the wavenumber vectors. NLN_{L} is the number of the unit cells in the system. The first term corresponds to the unperturbed Hamiltonian, and the second term is treated as a perturbed Hamiltonian. Here, Hαβ,σ0(k)H^{0}_{\alpha\beta,\sigma}(\textbf{k}) is defined as

Hαβ,σ0(k)=𝜹tαβ,σ,𝜹eik𝜹,H^{0}_{\alpha\beta,\sigma}(\textbf{k})=\sum_{\left<{\bm{\delta}}\right>}t_{\alpha\beta,\sigma,{\bm{\delta}}}e^{i\textbf{k}\cdot{\bm{\delta}}}, (3)

where 𝜹{\bm{\delta}} is a lattice vector connecting the neighbor unit cell. Further, tαβ,σ,𝜹t_{\alpha\beta,\sigma,{\bm{\delta}}} is the transfer integral between the fragment orbitals α\alpha and β\beta, which are separated by the lattice vector 𝜹{\bm{\delta}} and have the spin σ\sigma. We allot tαβ,σ,𝜹t_{\alpha\beta,\sigma,{\bm{\delta}}} to the transfer integrals t1t_{1}, t2t_{2}, …, t12t_{12} and the site potential Δ\Delta in Table 1, where ΔtBB,σ,0tAA,σ,0\Delta\equiv t_{BB,\sigma,\textbf{0}}-t_{AA,\sigma,\textbf{0}}. Note that these transfer integrals were obtained from the Wannier fitting. And we omit the small hoppings whose sizes are less than a cutoff energy of 0.0100.010 eV to make the analysis simple (see Fig. 3 for the schematic illustration of such a hopping network). Note that t1t_{1} connects the nearest-neighbor fragment orbitals within a molecule, while t2t_{2}, t3t_{3}, …, t8t_{8} connect those between molecules in the crystalline bbcc plane, which corresponds to the kbk_{b}kck_{c} plane in momentum space, where Dirac cones exist. We point out that t1t_{1}, t2t_{2}, and t3t_{3} are the three essential transfer integrals that are needed to create the Dirac cones, while t9t_{9} creates Fermi surfaces, and t10t_{10}, t11t_{11}, and t12t_{12} wind the Dirac nodal lines.

        transfer integrals (eV)
t1t_{1}       -0.2372
t2t_{2}       -0.1840
t3t_{3}       -0.2080
t4t_{4}       0.0302
t5t_{5}       0.0326
t6t_{6}       -0.0389
t7t_{7}       0.0103
t8t_{8}       -0.0144
t9t_{9}       -0.0140
t10t_{10}      -0.0541
t11t_{11}      -0.0534
t12t_{12}      0.0116
         site potential (eV)
Δ\Delta       0.0429
Table 1: Transfer integrals and site potential of [Ni(dmdt)2].
Refer to caption
Refer to caption
Figure 3: Hopping networks between fragment orbitals in tight-binding model of [Ni(dmdt)2]. (a) Schematic illustration of the 2D network of the transfer integrals (shown by double–headed arrows) in the crystalline bcbc–plane. The dashed square stands for the unit cell. (b) Schematic illustration of the 3D hopping network including the transfer integrals along the aa–direction. The black chain lines and vertical bold lines in (b) are guides to the eyes: The former lines are parallel to the aa–direction, while the latter lines connect the molecules in the same bcbcplane.

Previously, we calculated some quantities of [M(dmdt)2] (M = Pt, Ni) considering the spin–orbit interaction (SOI) as a parameter. There, we found that the SOI can reduce the Fermi surface in the bulk and induce helical edge modes.T.Kawamura2020 ; T.Kawamura2021 However, because the energy scale of SOI in this material (0.0016\sim 0.0016 eV) seems to be considerably smaller than the energy scale of the Fermi surface (0.01\sim 0.01 eV),T.Kawamura2021 in reality SOI should not be large enough to significantly reduce the size of the Fermi surface. Therefore, we will omit the influence of SOI on the Knight shift and 1/T1T1/T_{1}T.

From the definition in Eq. 3, Hαβ,σ0(k)H^{0}_{\alpha\beta,\sigma}(\textbf{k}) are expressed by the following equations.

HAA,σ0(k)=2t9coska,HAB,σ0(k)=t12ei(ka+kb+kc)+t1+t5ei(kb+kc)+t4eikc,HAC,σ0(k)=t10ei(ka+kb+kc)+t11ei(ka+kc)+t6+t3eikc+t2ei(kb+kc),HBB,σ0(k)=Δ+2t7cos(kb+kc)+2t8coskc,HBC,σ0(k)=t12ei(ka+kb+kc)+t1+t5ei(kb+kc)+t4eikc,HCC,σ0(k)=2t9coska.\begin{split}H^{0}_{AA,\sigma}(\textbf{k})&=2t_{9}\cos{k_{a}},\\ H^{0}_{AB,\sigma}(\textbf{k})&=t_{12}e^{i(-k_{a}+k_{b}+k_{c})}+t_{1}\\ &+t_{5}e^{i(k_{b}+k_{c})}+t_{4}e^{ik_{c}},\\ H^{0}_{AC,\sigma}(\textbf{k})&=t_{10}e^{i(-k_{a}+k_{b}+k_{c})}+t_{11}e^{i(k_{a}+k_{c})}+t_{6}\\ &+t_{3}e^{ik_{c}}+t_{2}e^{i(k_{b}+k_{c})},\\ H^{0}_{BB,\sigma}(\textbf{k})&=\Delta+2t_{7}\cos{(k_{b}+k_{c})}+2t_{8}\cos{k_{c}},\\ H^{0}_{BC,\sigma}(\textbf{k})&=t_{12}e^{i(-k_{a}+k_{b}+k_{c})}+t_{1}\\ &+t_{5}e^{i(k_{b}+k_{c})}+t_{4}e^{ik_{c}},\\ H^{0}_{CC,\sigma}(\textbf{k})&=2t_{9}\cos{k_{a}}.\end{split} (4)

Here, for simplicity, we set all the lattice constants to be unity. Two of the three bands for which we performed Wannier fitting are occupied.T.Kawamura2020 Thus, the tight-binding model obtained using the Wannier fitting is also 2/32/3 filling. The unperturbed Hamiltonian H^σ0(k)\hat{H}^{0}_{\sigma}(\textbf{k}) satisfies the eigenvalue equation

H^σ0(k)|k,n,σ=En,σ(k)|k,n,σ,\hat{H}^{0}_{\sigma}(\textbf{k})\ket{{\textbf{k},n,\sigma}}=E_{n,\sigma}(\textbf{k})\ket{{\textbf{k},n,\sigma}}, (5)
|k,n,σ=(dA,n,σ(k)dB,n,σ(k)dC,n,σ(k)).\ket{{\textbf{k},n,\sigma}}=\left(\begin{array}[]{c}d_{A,n,\sigma}(\textbf{k})\\ d_{B,n,\sigma}(\textbf{k})\\ d_{C,n,\sigma}(\textbf{k})\end{array}\right). (6)

where En,σ(k)E_{n,\sigma}(\textbf{k}) is the eigenvalue and |k,n,σ\ket{{\textbf{k},n,\sigma}} is the eigen vector; nn is the band index; and dα,n,σ(k)d_{\alpha,n,\sigma}(\textbf{k}) denotes the wave functions. H^σ0(k)\hat{H}^{0}_{\sigma}(\textbf{k}) in Eq. 5 consists of the matrix elements Hαβ,σ0(k)H^{0}_{\alpha\beta,\sigma}(\textbf{k}) in Eqs. 2 and 4. Reflecting the 2/32/3-filling, the chemical potential μ\mu is determined by

1NLk,n,σfk,n,σ=4,\frac{1}{N_{L}}\sum_{\textbf{k},n,\sigma}f_{\textbf{k},n,\sigma}=4, (7)
ϵk,n,σEn,σ(k)μ,\epsilon_{\textbf{k},n,\sigma}\equiv E_{n,\sigma}(\textbf{k})-\mu, (8)

fk,n,σf_{\textbf{k},n,\sigma}=1/[1+exp(ϵk,n,σ/T)]1/\left[1+\exp{(\epsilon_{\textbf{k},n,\sigma}/T)}\right] is the Fermi distribution function, and we have μ\mu=EFE_{F} at TT=0.

As to the second term in the Hamiltonian Eq. 2, we introduce the on-site repulsive interaction UααU_{\alpha\alpha} defined on the fragment orbital α\alpha, which is defined as the diagonal element of the interaction matrix as follows

U^\displaystyle\hat{U} =\displaystyle= (UAA000UBB000UCC)\displaystyle\left(\begin{array}[]{ccc}U_{AA}&0&0\\ 0&U_{BB}&0\\ 0&0&U_{CC}\\ \end{array}\right) (12)
=\displaystyle= (λU000U000λU),\displaystyle\left(\begin{array}[]{ccc}\lambda U&0&0\\ 0&U&0\\ 0&0&\lambda U\\ \end{array}\right), (16)

Here, because of the inversion symmetry, we have UAAU_{AA}=UCCU_{CC}, and we use the relative size of UAAU_{AA} to UBBU_{BB}, λ\lambda=UAA/UBBU_{AA}/U_{BB}, as a control parameter of this model. According to respack, we found UU=6.726.72 eV and λ\lambda=0.790.79 in the unscreened case, and UU=2.682.68 eV and λ\lambda=0.950.95 in the screened case. In this study, we set λ\lambda=0.950.95 and use a value of UU less than 2.682.68 eV because U^\hat{U} tends to be overestimated in RPA.

The longitudinal and transverse spin susceptibilities are defined as follows:

χ^zz(q,iωl)1201/T𝑑τeiωlτTτS^qz(τ)S^qz(0),\displaystyle\hat{\chi}^{zz}(\textbf{q},i\omega_{l})\equiv\frac{1}{2}\int^{1/T}_{0}d\tau e^{i\omega_{l}\tau}\left<T_{\tau}\hat{S}^{z}_{\textbf{q}}(\tau)\hat{S}^{z}_{-\textbf{q}}(0)\right>, (17)
S^qz=1NLk(c^k+q,c^k,c^k+q,c^k,),\displaystyle\hat{S}^{z}_{\textbf{q}}=\frac{1}{N_{L}}\sum_{\textbf{k}}\left(\hat{c}^{\dagger}_{\textbf{k}+\textbf{q},\uparrow}\hat{c}_{\textbf{k},\uparrow}-\hat{c}^{\dagger}_{\textbf{k}+\textbf{q},\downarrow}\hat{c}_{\textbf{k},\downarrow}\right), (18)
χ^±(q,iωl)01/T𝑑τeiωlτTτS^q+(τ)S^q(0),\displaystyle\hat{\chi}^{\pm}(\textbf{q},i\omega_{l})\equiv\int^{1/T}_{0}d\tau e^{i\omega_{l}\tau}\left<T_{\tau}\hat{S}^{+}_{\textbf{q}}(\tau)\hat{S}^{-}_{-\textbf{q}}(0)\right>, (19)
S^q+=1NLkc^k+q,c^k,,\displaystyle\hat{S}^{+}_{\textbf{q}}=\frac{1}{N_{L}}\sum_{\textbf{k}}\hat{c}^{\dagger}_{\textbf{k}+\textbf{q},\uparrow}\hat{c}_{\textbf{k},\downarrow}, (20)
S^q=1NLkc^k,c^k+q,.\displaystyle\hat{S}^{-}_{-\textbf{q}}=\frac{1}{N_{L}}\sum_{\textbf{k}}\hat{c}^{\dagger}_{\textbf{k},\downarrow}\hat{c}_{\textbf{k}+\textbf{q},\uparrow}. (21)

Here, iωli\omega_{l}=2liπT2li\pi T (lZ)(l\in\textbf{Z}) is the Matsubara frequency and τ\tau is the imaginary time. S^qz\hat{S}^{z}_{\textbf{q}}, S^q+\hat{S}^{+}_{\textbf{q}}, and S^q\hat{S}^{-}_{\textbf{q}} are the spin operators. S^qz(τ)\hat{S}^{z}_{\textbf{q}}(\tau), S^q+(τ)\hat{S}^{+}_{\textbf{q}}(\tau), and S^q(τ)\hat{S}^{-}_{\textbf{q}}(\tau) are described in the Heisenberg picture. The spin susceptibility is the proportionality coefficient of the magnetization to the infinitesimal magnetic field. It represents the degree of the “spin fluctuations”, because spins in the system sensitively respond to the infinitesimal magnetic field when spin susceptibility is large.
By performing a perturbation expansion of Eqs. 17 and 19, we obtain the non-interacting longitudinal spin susceptibility χ^zz,0(q,iωl)\hat{\chi}^{zz,0}(\textbf{q},i\omega_{l}) and non-interacting transverse spin susceptibility χ^±,0(q,iωl)\hat{\chi}^{\pm,0}(\textbf{q},i\omega_{l}) as the zeroth-order perturbation terms. In this study, SU(2) symmetry is protected. Therefore, we define χ^zz,0(q,iωl)\hat{\chi}^{zz,0}(\textbf{q},i\omega_{l})=χ^±,0(q,iωl)\hat{\chi}^{\pm,0}(\textbf{q},i\omega_{l})\equiv χ^0(q,iωl)\hat{\chi}^{0}(\textbf{q},i\omega_{l}). Its matrix elements are written as

χαβ0(q,iωl)=TNLk,mGαβ0(k+q,iω~m+iωl)Gβα0(k,iω~m),\begin{split}&\chi^{0}_{\alpha\beta}(\textbf{q},i\omega_{l})\\ &=-\frac{T}{N_{L}}\sum_{\textbf{k},m}G^{0}_{\alpha\beta}(\textbf{k}+\textbf{q},i\tilde{\omega}_{m}+i\omega_{l})G^{0}_{\beta\alpha}(\textbf{k},i\tilde{\omega}_{m}),\end{split} (22)
Gαβ,σ0(k,iω~l)=ndα,n,σ(k)dβ,n,σ(k)1iω~lϵk,n,σ.\begin{split}G^{0}_{\alpha\beta,\sigma}(\textbf{k},i\tilde{\omega}_{l})=\sum_{n}d_{\alpha,n,\sigma}(\textbf{k})d^{*}_{\beta,n,\sigma}(\textbf{k})\frac{1}{i\tilde{\omega}_{l}-\epsilon_{\textbf{k},n,\sigma}}.\end{split} (23)

Eq. 23 expresses the matrix elements of the non-interacting Green’s function. iω~li\tilde{\omega}_{l}=(2l+1)iπT(2l+1)i\pi T is the Matsubara frequency of the fermions. The spin index σ\sigma is omitted in Eq. 22 because we have ϵk,n,\epsilon_{\textbf{k},n,\uparrow}=ϵk,n,\epsilon_{\textbf{k},n,\downarrow} in Eq. 23. The longitudinal and transverse spin susceptibilities in RPA are represented by the Feynman diagrams shown in Fig. 4.

Refer to caption
Refer to caption
Figure 4: (A) Feynman diagram of χ^zz,s(q,iωl)\hat{\chi}^{zz,s}(\textbf{q},i\omega_{l}). (B) Feynman diagram of χ^±,s(q,iωl)\hat{\chi}^{\pm,s}(\textbf{q},i\omega_{l}). The solid lines represent the non-interacting Green’s functions, which describe the quasi-particles. The dashed lines represent the interactions. The open circles are the vertexes connecting the non-interacting Green’s functions and the interaction. The black dots represent the spin operators.

The first terms in the right-hand sides of diagrams (A) and (B) correspond to the terms χ^zz,0(q,iωl)\hat{\chi}^{zz,0}(\textbf{q},i\omega_{l})=χ^±,0(q,iωl)\hat{\chi}^{\pm,0}(\textbf{q},i\omega_{l})== χ^0(q,iωl)\hat{\chi}^{0}(\textbf{q},i\omega_{l}) (Eq. 22). Because the interacting longitudinal and transverse spin susceptibilities are represented by summations of the series of U^χ^0(q,iωl)\hat{U}\hat{\chi}^{0}(\textbf{q},i\omega_{l}) in RPA, they are written as

χ^zz,s(q,iωl)\displaystyle\hat{\chi}^{zz,s}(\textbf{q},i\omega_{l}) =\displaystyle= χ^±,s(q,iωl)χ^s(q,iωl)\displaystyle\hat{\chi}^{\pm,s}(\textbf{q},i\omega_{l})\equiv\hat{\chi}^{s}(\textbf{q},i\omega_{l})
=\displaystyle= χ^0(q,iωl)[I^U^χ^0(q,iωl)]1\displaystyle\hat{\chi}^{0}(\textbf{q},i\omega_{l})[\hat{I}-\hat{U}\hat{\chi}^{0}(\textbf{q},i\omega_{l})]^{-1}
,

where I^\hat{I} is the unit matrix.

Here we introduce the Stoner factor ξs(q)\xi_{s}(\textbf{q}) representing the degree of enhancement of the spin fluctuations. The Stoner factor ξs(q)\xi_{s}(\textbf{q}) is defined as the maximum eigenvalue of U^χ^0(q,0)\hat{U}\hat{\chi}^{0}(\textbf{q},0). The relation between ξs(q)\xi_{s}(\textbf{q}) and χ^s(q,0)\hat{\chi}^{s}(\textbf{q},0) in the three-orbital model is given by

χ^s(q,0)=1(1ξs(q))χ^0(q,0)P^(q)(1ϕ1(q))(1ϕ2(q)),\displaystyle\hat{\chi}^{s}(\textbf{q},0)=\frac{1}{(1-\xi_{s}(\textbf{q}))}\frac{\hat{\chi}^{0}(\textbf{q},0)\hat{P}(\textbf{q})}{(1-\phi_{1}(\textbf{q}))(1-\phi_{2}(\textbf{q}))}, (25)

where ξs(q)\xi_{s}(\textbf{q}), ϕ1(q)\phi_{1}(\textbf{q}), and ϕ2(q)\phi_{2}(\textbf{q}) are the maximum and other eigenvalues of U^χ^0(q,0)\hat{U}\hat{\chi}^{0}(\textbf{q},0). P^(q)\hat{P}(\textbf{q}) is the adjugate matrix of I^U^χ^0(q,0)\hat{I}-\hat{U}\hat{\chi}^{0}(\textbf{q},0). The eigenvalues of U^χ^0(q,0)\hat{U}\hat{\chi}^{0}(\textbf{q},0) is smaller than 11 in the paramagnetic regime. When ξs(q)1\xi_{s}(\textbf{q})\rightarrow 1, the spin susceptibility χ^s(q,0)\hat{\chi}^{s}(\textbf{q},0) diverges and induces a magnetic order, corresponding to the wavenumber q.

Within the framework of linear response theory, the Knight shift, KK, and the spin-lattice relaxation rate, (1/T1T)(1/T_{1}T), for the orbital α\alpha are given byS.Katayama2009

Kα\displaystyle K_{\alpha} \displaystyle\propto βRe[χαβzz(0,0)],\displaystyle\sum_{\beta}{\rm Re}\left[\chi^{zz}_{\alpha\beta}(\textbf{0},0)\right], (26)
(1/T1T)α\displaystyle\left(1/T_{1}T\right)_{\alpha} \displaystyle\propto limω+0[1NLqImχαα±(q,ω)ω]\displaystyle\lim_{\omega\rightarrow+0}\left[\frac{1}{N_{L}}\sum_{\textbf{q}}\frac{{\rm Im}\chi^{\pm}_{\alpha\alpha}(\textbf{q},\omega)}{\omega}\right] (27)

Here, note that Eq. II is satisfied. According to Eqs. 25 and 27, all the q components for which ξs(q)\xi_{s}(\textbf{q}) becomes close to unity make a dominant contribution to 1/T1T1/T_{1}T because they lead to a large value in the spin susceptibility. By contrast, the Knight shift is solely affected by the q=0 component that satisfies ξs(q)1\xi_{s}(\textbf{q})\sim 1 (see Eq. 26).

Because the spin susceptibility in the real-frequency representation χ^s(q,ω)\hat{\chi}^{s}(\textbf{q},\omega) is necessary for 1/T1T1/T_{1}T, we obtain χ^s(q,ω)\hat{\chi}^{s}(\textbf{q},\omega) by performing an analytic continuation of Eq. II. In this way, we use the real-frequency representation depending on the physical quantities.

III result in the absence of UU

In this section, we calculate the electronic state, spin susceptibility, Knight shift, and 1/T1T1/T_{1}T in the absence of the repulsive interaction UU. Further, we find that the spin susceptibility in [Ni(dmdt)2] greatly depends on the fragment orbitals and will explain the relationship between the spin susceptibilities and the wave functions.

III.1 Electronic state and spin susceptibility in the absence of UU

We first calculate the energy dispersion of [Ni(dmdt)2] by diagonalizing the unperturbed Hamiltonian H^σ0(k)\hat{H}^{0}_{\sigma}(\textbf{k}) in Eq. 5. The resulting energy bands are depicted in Fig. 5 (a), where the dispersion seen in the kbk_{b}kck_{c} plane at kak_{a}=π/2-\pi/2 is shown. Inside the 2D first Brillouin zone, two pairs of gapless Dirac cones appear between the first and second top bands near EFE_{F} (the Band 1 and 2 in Fig. 5 (a), where EFE_{F} is chosen as the energy origin) and between the second and third bands rather beneath EFE_{F} (the Bands 2 and 3), where these band-crossing points are protected by space-inversion symmetry.

Dirac points between bands 1 and 2 in the kbk_{b}kck_{c} plane draw the Dirac nodal lines in the kak_{a} direction. The inset of Fig. 5 (b) shows the Dirac nodal line in the first Brillouin zone. Upon changing the momentum along the kak_{a} direction, the positions of the Dirac points near EFE_{F} move in the kbk_{b}kck_{c} plane and eventually form a pair of so-called Dirac nodal lines in the 3D Brillouin zone [see the inset of Fig. 5 (b)]. The band-crossing points accordingly move up and down across EFE_{F} with changing the value of kak_{a} as illustrated in Fig. 5 (c), which generates electron and hole pockets around these nodal lines [see Fig. 5 (b), where the electron and hole pockets are illustrated as thin magenta and green strips, respectively]. Figure 5(b) shows the Fermi surface in [Ni(dmdt)2]. The energy scale of such Fermi pockets is approximately 0.010 eV.

The corresponding density of states (DOS), Dtot(ω)D_{\rm tot}(\omega), was also calculated, which is given by a sum of a fragment-orbital dependent DOS, Dα(ω)D_{\alpha}(\omega) [α\alpha=A(=C) and B], that is given by

Dα(ω)=1πNLk,σImGαα,σR,0(k,ω),\displaystyle D_{\alpha}(\omega)=-\frac{1}{\pi N_{L}}\sum_{\textbf{k},\sigma}{\rm Im}G^{R,0}_{\alpha\alpha,\sigma}(\textbf{k},\omega), (28)
Gαβ,σR,0(k,ω)=ndα,n,σ(k)dβ,n,σ(k)1ωϵk,n,σ+iη.\displaystyle G^{R,0}_{\alpha\beta,\sigma}(\textbf{k},\omega)=\sum_{n}d_{\alpha,n,\sigma}(\textbf{k})d^{*}_{\beta,n,\sigma}(\textbf{k})\frac{1}{\omega-\epsilon_{\textbf{k},n,\sigma}+i\eta}. (29)

Here, G^R,0(k,ω)\hat{G}^{R,0}(\textbf{k},\omega) is the non-interacting retarded Green’s function, where η\eta=+0+0. The resulting Dtot(ω)D_{\rm tot}(\omega) and Dα(ω)D_{\alpha}(\omega) are depicted in Fig. 6. Note that the DOS has linear ω\omega dependence near EFE_{F} (corresponding to ω\omega=0 in Fig. 6) because the energy dispersion of this system near EFE_{F} is close to that of a two-dimensional Dirac electron system, and the three-dimensional effect is only an addition of a small dispersion along a kak_{a} direction. The finite DOS at EFE_{F} in this material is ascribed to the presence of the Fermi pockets induced by that kak_{a}-axis dispersion.

Before moving to the analysis of the spin susceptibility, further comments are needed on the fragment-orbital-dependent characters of this system. Figure 7 shows the momentum dependence of the squared wave function for the orbital B projected onto the second top band [Band 2 in Fig. 5(a)], |dB,2,σ(k)|2|d_{B,2,\sigma}(\textbf{k})|^{2}, plotted as a function of kbk_{b} and kck_{c} at kak_{a}=π/2-\pi/2. Notably, the line segments that connect the positions of these Dirac points (illustrated with black lines in Fig. 7) have a vanishing amplitude, |dB,2,σ(k)|2|d_{B,2,\sigma}(\textbf{k})|^{2}=0, which we call the “zero region” in this paper. By contrast, the wave functions of the orbitals A and C do not have such ZR. The presence of a similar ZR was previously found in other nn-band (n3n\geq 3) Dirac electron systems, such as the organic conductors α\alpha-(BEDT-TTF)2I3 (nn=44)A.Kobayashi2013 .

Refer to caption
Refer to caption
Refer to caption
Figure 5: (a) Energy dispersion of [Ni(dmdt)2] in the kbk_{b}kck_{c} plane, where kak_{a}=π/2-\pi/2. Dirac cones exist between each pair of bands. Those between bands 1 and 2 draw the Dirac nodal line. (b) Fermi surface in the first Brillouin zone. The electron and hole pockets are drawn in magenta and green, respectively. The inset shows the Dirac nodal line in the first Brillouin zone. (c) Schematic of the relationship among the Dirac nodal line, Fermi surface, and wavenumber kak_{a}. The red curved line shows the Dirac nodal line. The dotted transverse line shows EFE_{F}.
Refer to caption
Figure 6: The local density of state Dα(ω)D_{\alpha}(\omega) for the fragment orbital α\alpha=A, B, and C in the unit cell of [Ni(dmdt)2]. The red dotted and blue dashed lines show DA(ω)D_{A}(\omega)(=DC(ω)D_{C}(\omega)) and DB(ω)D_{B}(\omega), respectively. The black solid line shows the total density of states Dtot(ω)D_{{\rm tot}}(\omega)=DA(ω)+DB(ω)+DC(ω)D_{A}(\omega)+D_{B}(\omega)+D_{C}(\omega). The integral value of Dtot(ω)D_{{\rm tot}}(\omega) orver ω\omega is equal to 66 due to the three orbitals and the spins. Its unit is the number of electrons.
Refer to caption
Figure 7: Absolute squared wave function of orbital B in band 2, |dB,2,σ(k)|2|d_{B,2,\sigma}(\textbf{k})|^{2}, in the kbk_{b}kck_{c} plane, where kak_{a}=π/2-\pi/2. The up arrows indicate the corresponding positions of the Dirac points formed between the Bands 1 and 2 in Fig. 5 (a), while the down arrows indicate those between the Bands 2 and 3. The color bar represents the magnitude of |dB,2,σ(k)|2|d_{B,2,\sigma}(\textbf{k})|^{2}.

Second, we calculate the non-interacting spin susceptibility χ^0(q,ω)\hat{\chi}^{0}(\textbf{q},\omega) to elucidate spin fluctuations, which can be enhanced in this material.

Refer to caption
Figure 8: The momentum dependences of the diagonal elements of the spin susceptibility in the absence of UU. (a) χAA0(q,0)\chi^{0}_{AA}(\textbf{q},0) in the qbq_{b}qcq_{c} plane, where qaq_{a}=0. (b) χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) in the qbq_{b}qcq_{c} plane, where qaq_{a}=0.2π0.2\pi. (c) Im[χAA0(q,ω0)]{\rm Im}[\chi^{0}_{AA}(\textbf{q},\omega_{0})] in the qbq_{b}qcq_{c} plane, where qaq_{a}=0. (d) Im[χBB0(q,ω0)]{\rm Im}[\chi^{0}_{BB}(\textbf{q},\omega_{0})] in the qbq_{b}qcq_{c} plane, where qaq_{a}=0.2π0.2\pi. The temperature TT=0.0030.003 eV.

Figure 8 (a), (b), (c), and (d) show the diagonal elements of the non-interacting spin susceptibility χAA0(q,0)\chi^{0}_{AA}(\textbf{q},0), χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0), Im[χAA0(q,ω0)]{\rm Im}[\chi^{0}_{AA}(\textbf{q},\omega_{0})], and Im[χBB0(q,ω0)]{\rm Im}[\chi^{0}_{BB}(\textbf{q},\omega_{0})], respectively, at TT=0.0030.003 eV. These quantities slightly increase with raising temperature for q=0, while the magnitude relation χAA0(q0,0)\chi^{0}_{AA}(\textbf{q}\sim\textbf{0},0)<<χBB0(qQ,0)\chi^{0}_{BB}(\textbf{q}\sim\textbf{Q},0) and Im[χAA0(q0,ω0)]{\rm Im}\left[\chi^{0}_{AA}(\textbf{q}\sim\textbf{0},\omega_{0})\right]>>Im[χBB0(qQ,ω0)]{\rm Im}\left[\chi^{0}_{BB}(\textbf{q}\sim\textbf{Q},\omega_{0})\right] do not change with temperature. Here, χαα0(q,0)\chi^{0}_{\alpha\alpha}(\textbf{q},0) is a real number. We set ω0\omega_{0}=0.0010.001 eV because the imaginary part of the spin susceptibility at the infinitesimal frequency is essential to solve Eq. 27. One of qaq_{a}, qbq_{b}, and qcq_{c} must be fixed to show the spin susceptibilities in the three-dimensional figures. We fix qaq_{a}=0 in Fig. 8 (a) and (c) and qaq_{a}=0.2π0.2\pi in Fig. 8 (b) and (d) because χAA0(q,0)\chi^{0}_{AA}(\textbf{q},0) and χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) have the maximum value at the commensurate wavenumber q=0 and the incommensurate wavenumber q=Q=(0.20π,0.73π,0.58π)(0.20\pi,0.73\pi,0.58\pi) at TT=0.0030.003 eV, respectively. We define Q as the wavenumber at which χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) has the maximum value. Further, Q varies slightly with temperature. χCC0(q,ω)\chi^{0}_{CC}(\textbf{q},\omega) is equivalent to χAA0(q,ω)\chi^{0}_{AA}(\textbf{q},\omega) owing to space-inversion symmetry. In Fig. 8 (a) and (c), χAA0(q,0)\chi^{0}_{AA}(\textbf{q},0) and Im[χAA0(q,ω0)]{\rm Im}[\chi^{0}_{AA}(\textbf{q},\omega_{0})] have their maximum value at q=0. In Fig. 8 (b) and (d), χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) and Im[χBB0(q,ω0)]{\rm Im}[\chi^{0}_{BB}(\textbf{q},\omega_{0})] have their maximum value at q=Q. In addition, Im[χBB0(q,ω0)]{\rm Im}[\chi^{0}_{BB}(\textbf{q},\omega_{0})] has some peaks other than that at q=Q.

The difference between χAA0(q,0)\chi^{0}_{AA}(\textbf{q},0) and χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) implies that [Ni(dmdt)2] has two candidates for magnetic order that can be induced in the bulk. They are the q=0 magnetic order and SDW. To explain the mechanism of the fragment-orbital-dependent spin susceptibility, we calculate the spectral weights on the Fermi surface. The spectral weight is given by

ρα(k,ω)=1πImGααR,0(k,ω).\rho_{\alpha}(\textbf{k},\omega)=-\frac{1}{\pi}{\rm Im}G^{R,0}_{\alpha\alpha}(\textbf{k},\omega). (30)

The spin index σ\sigma is omitted in Eq. 30, and k=(ka,kb,kc)(k_{a},k_{b},k_{c}) is the wavenumber. Eq. 30 with ω\omega=0 yields the spectral weight at EFE_{F}. The spectral weight shows the weights of the respective fragment orbitals for the energy ω\omega and the wavenumber k because GααR,0(k,ω)G^{R,0}_{\alpha\alpha}(\textbf{k},\omega) in Eq. 30 contains the absolute square of the wave function |dα,n(k)|2|d_{\alpha,n}(\textbf{k})|^{2}. Furthermore, the relationship 1NLkρα(k,ω)\frac{1}{N_{L}}\sum_{\textbf{k}}\rho_{\alpha}(\textbf{k},\omega)=Dα(ω)D_{\alpha}(\omega) is satisfied. Figure 9 (a) and (b) show the spectral weight on the cross-section where the Fermi surface in Fig. 5(b) is cut on the kak_{a}=π\pi plane, which corresponds to the hole pocket. We set kak_{a}=π\pi because the hole pockets are important for χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0), as we discuss below. Figure 9 (a) and (b), respectively, show ρA(k,0)\rho_{A}(\textbf{k},0) and ρB(k,0)\rho_{B}(\textbf{k},0) in the kbk_{b}kck_{c} plane. In both figures, 0.65π<kb<0.50π-0.65\pi<k_{b}<-0.50\pi and 0.20π<kc<0.30π0.20\pi<k_{c}<0.30\pi. The spectral weight of A is not zero on the Fermi surface, but that of B has the appearance of a crescent moon because of a ZR. This difference in spectral weights results in the fragment-orbital-dependent spin susceptibilities.

Refer to caption
Figure 9: (a) Spectral weight ρA(k,0)\rho_{A}(\textbf{k},0). It is not zero on the Fermi surface. (b) Spectral weight ρB(k,0)\rho_{B}(\textbf{k},0). A part of it is zero because of a ZR. In the both figures, kak_{a}=π\pi, 0.65π<kb<0.50π-0.65\pi<k_{b}<-0.50\pi, and 0.20π<kc<0.30π0.20\pi<k_{c}<0.30\pi. Color bars represent the magnitude of the spectral weights ρα(π,kb,kc,0)\rho_{\alpha}(\pi,k_{b},k_{c},0) and ρα(0.8π,kb,kc,0)\rho_{\alpha}(0.8\pi,k_{b},k_{c},0). The yellow region indicates that the spectral weight at EFE_{F} is high.

After performing summation over iω~mi\tilde{\omega}_{m} in Eq. 22, the non-interacting spin susceptibility is written as

χα,β0(q,iωl)=1NLk,m,nfk+q,mfk,nϵk+q,mϵk,niωl×dα,m(k+q)dβ,m(k+q)dβ,n(k)dα,n(k),\begin{split}&\chi^{0}_{\alpha,\beta}(\textbf{q},i\omega_{l})=-\frac{1}{N_{L}}\sum_{\textbf{k},m,n}\frac{f_{\textbf{k}+\textbf{q},m}-f_{\textbf{k},n}}{\epsilon_{\textbf{k}+\textbf{q},m}-\epsilon_{\textbf{k},n}-i\omega_{l}}\\ &\times d_{\alpha,m}(\textbf{k}+\textbf{q})d^{*}_{\beta,m}(\textbf{k}+\textbf{q})d_{\beta,n}(\textbf{k})d^{*}_{\alpha,n}(\textbf{k}),\end{split} (31)

where the spin index σ\sigma is omitted. The terms in which the denominator is close to 0 and the numerator is close to 11 in Eq. 31 contribute to χαβ0(q,0)\chi^{0}_{\alpha\beta}(\textbf{q},0). Such terms are given by the wavenumber k+q\textbf{k}+\textbf{q} and k near the Fermi surface. Thus, the vector q connecting the Fermi surface is important for the spin susceptibility. This wavenumber is called the nesting vector. q=Q in Fig. 8 corresponds to the nesting vector Q in Fig. 10. q=Q connects the regions where the spectral weight of orbital B is high on the hole pockets. Q is the wavenumber at which χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) has the maximum value. The nesting between the electron pockets is relatively weaker than that between the hole pockets. Note that we solved Eq. 22 using a fast Fourier transform, instead of solving Eq. 31.

Refer to caption
Figure 10: Nesting vector Q in the momentum space. Q connects the regions where the spectral weight of orbital B is high (in the right figure). Color bars represent the magnitude of the spectral weight ρB(π,kb,kc,0)\rho_{B}(\pi,k_{b},k_{c},0). The electron and hole pockets are drawn in magenta and green, respectively (in the left figure).

Next, we calculate the temperature dependence of the matrix elements of χ^0(0,0)\hat{\chi}^{0}(\textbf{0},0) using Eq. 22 because it is essential to the following calculations. They are real numbers because q=0 and ω\omega=0. χAA0(0,0)\chi^{0}_{AA}(\textbf{0},0)=χCC0(0,0)\chi^{0}_{CC}(\textbf{0},0), χAB0(0,0)\chi^{0}_{AB}(\textbf{0},0)=χBA0(0,0)\chi^{0}_{BA}(\textbf{0},0)=χBC0(0,0)\chi^{0}_{BC}(\textbf{0},0)=χCB0(0,0)\chi^{0}_{CB}(\textbf{0},0), and χAC0(0,0)\chi^{0}_{AC}(\textbf{0},0)=χCA0(0,0)\chi^{0}_{CA}(\textbf{0},0) are satisfied owing to space-inversion symmetry and time-reversal symmetry. Figure 11 shows the temperature dependence of the matrix elements of χ^0(0,0)\hat{\chi}^{0}(\textbf{0},0). The diagonal element χAA0(0,0)\chi^{0}_{AA}(\textbf{0},0) is almost constant. Furthermore, χBB0(0,0)\chi^{0}_{BB}(\textbf{0},0) decreases slowly with decreasing temperature. However, the off-diagonal elements χAB0(0,0)\chi^{0}_{AB}(\textbf{0},0) and χAC0(0,0)\chi^{0}_{AC}(\textbf{0},0) are negative and decrease as the temperature decreases. To show what determines the sign of χAB0(0,0)\chi^{0}_{AB}(\textbf{0},0) and χAC0(0,0)\chi^{0}_{AC}(\textbf{0},0), we define the band-dependent spin susceptibility as

χαβ,mn0(q,iωl)=TNLk,lGαβ,m0(k+q,iω~l+iωl)Gβα,n0(k,iω~l),\begin{split}&\chi^{0}_{\alpha\beta,mn}(\textbf{q},i\omega_{l})\\ &=-\frac{T}{N_{L}}\sum_{\textbf{k},l^{\prime}}G^{0}_{\alpha\beta,m}(\textbf{k}+\textbf{q},i\tilde{\omega}_{l^{\prime}}+i\omega_{l})G^{0}_{\beta\alpha,n}(\textbf{k},i\tilde{\omega}_{l^{\prime}}),\end{split} (32)
Gαβ,m,σ0(k,iω~l)=dα,m,σ(k)dβ,m,σ(k)1iω~lϵk,m,σ,\begin{split}G^{0}_{\alpha\beta,m,\sigma}(\textbf{k},i\tilde{\omega}_{l})=d_{\alpha,m,\sigma}(\textbf{k})d^{*}_{\beta,m,\sigma}(\textbf{k})\frac{1}{i\tilde{\omega}_{l}-\epsilon_{\textbf{k},m,\sigma}},\end{split} (33)

where mm and nn are the band indices. Eq. 32 satisfies m,nχαβ,mn0(q,iωl)\sum_{m,n}\chi^{0}_{\alpha\beta,mn}(\textbf{q},i\omega_{l})= χαβ0(q,iωl)\chi^{0}_{\alpha\beta}(\textbf{q},i\omega_{l}), where χαβ0(q,iωl)\chi^{0}_{\alpha\beta}(\textbf{q},i\omega_{l}) is given by Eq. 22. The inset of Fig. 11 shows the temperature dependence of χAB,120(0,0)\chi^{0}_{AB,12}(\textbf{0},0) and χAC,120(0,0)\chi^{0}_{AC,12}(\textbf{0},0). They are negative. Such terms render the off-diagonal elements of spin susceptibility, χAB0(0,0)\chi^{0}_{AB}(\textbf{0},0) and χAC0(0,0)\chi^{0}_{AC}(\textbf{0},0), negative.

Refer to caption
Figure 11: Temperature dependence of χαβ0(0,0)\chi^{0}_{\alpha\beta}(\textbf{0},0). The red dashed, blue solid, green dotted, and purple chain lines represent χAA0(0,0)\chi^{0}_{AA}(\textbf{0},0), χBB0(0,0)\chi^{0}_{BB}(\textbf{0},0), χAB0(0,0)\chi^{0}_{AB}(\textbf{0},0), and χAC0(0,0)\chi^{0}_{AC}(\textbf{0},0), respectively. The inset shows the band-dependent spin susceptibilities χAB,120(0,0)\chi^{0}_{AB,12}(\textbf{0},0) and χAC,120(0,0)\chi^{0}_{AC,12}(\textbf{0},0) by the green dotted and purple chain lines, respectively.

III.2 Knight shift and 1/T1T1/T_{1}T in the absence of UU

We calculate the Knight shift in the absence of UU using Eq. 26. The fragment orbital components of the Knight shift can be given by KAK_{A}=χAA0(0,0)+χAB0(0,0)+χAC0(0,0)\chi^{0}_{AA}(\textbf{0},0)+\chi^{0}_{AB}(\textbf{0},0)+\chi^{0}_{AC}(\textbf{0},0) and KBK_{B}=χBB0(0,0)+2χAB0(0,0)\chi^{0}_{BB}(\textbf{0},0)+2\chi^{0}_{AB}(\textbf{0},0) using the spin susceptibilities in Fig. 11 due to the space-inversion symmetry. Figure 12 shows the Knight shift for the fragment orbitals A(=C) and B in the absence of UU. TT=TT^{*}\sim0.010.01eV is the energy scale where the Fermi surface affects the Knight shift and 1/T1T1/T_{1}T. TT^{*}\sim0.010.01eV is consistent with the energy scale of the Fermi surface. For temperatures higher than TT^{*}, the Knight shift and 1/T1T1/T_{1}T are affected by the linear energy dispersion. The Knight shift of the Dirac electron system in the absence of the interaction is given by KαDα(ω)(df(ω)dω)𝑑ωK_{\alpha}\simeq\int^{\infty}_{-\infty}D_{\alpha}(\omega)\left(-\frac{df(\omega)}{d\omega}\right)d\omega.S.Katayama2009 f(ω)f(\omega) is the Fermi distribution function for the energy ω\omega. Because Dα(ω)ωD_{\alpha}(\omega)\propto\omega near the Femi energy in Fig. 6, KαK_{\alpha} is proportional to TT for TTT\gtrsim T^{*} in Fig. 12. In the two-dimensional Dirac electron system, the Knight shift is proportional to TT at low temperature and becomes zero at TT=0 because DOS is zero at EFE_{F}. However, KαK_{\alpha} in this material is not proportional to TT for TTT\lesssim T^{*} because Dα(0)0D_{\alpha}(0)\neq 0 in Fig. 6. This is the effect of the Fermi surface. The magnitude relationship KB>KAK_{B}>K_{A} results from DB(ω)>DA(ω)D_{B}(\omega)>D_{A}(\omega) near the Fermi energy.

Refer to caption
Figure 12: Temperature dependence of KαK_{\alpha} in the absence of UU. The red dashed and blue dotted lines show KAK_{A} and KBK_{B}, respectively. The black solid line shows the total Knight shift KtotalK_{total}=KA+KB+KCK_{A}+K_{B}+K_{C}. The dotted longitudinal line indicates TT=TT^{*}\sim0.010.01 eV. The green thin line which is proportional to TT is drawn for the eye guide.

Next, we calculate the spin-lattice relaxation rate 1/T1T1/T_{1}T in the absence of UU using Eq. 27. 1/T1T1/T_{1}T is determined by qIm[χαα0(q,ω0)]\sum_{\textbf{q}}{\rm Im}[\chi^{0}_{\alpha\alpha}(\textbf{q},\omega_{0})]. Im[χAA0(q,ω0)]{\rm Im}[\chi^{0}_{AA}(\textbf{q},\omega_{0})] and Im[χBB0(q,ω0)]{\rm Im}[\chi^{0}_{BB}(\textbf{q},\omega_{0})] are shown in Fig. 8. Fig. 13 shows the temperature dependence of 1/T1T1/T_{1}T in the absence of UU. (1/T1T)α(1/T_{1}T)_{\alpha} is proportional to T2T^{2} for TTT\gtrsim T^{*} in Fig. 12 because the 1/T1T1/T_{1}T of the Dirac electron system is given by (1/T1T)α[Dα(ω)]2(df(ω)dω)𝑑ω(1/T_{1}T)_{\alpha}\simeq\int^{\infty}_{-\infty}[D_{\alpha}(\omega)]^{2}\left(-\frac{df(\omega)}{d\omega}\right)d\omega.S.Katayama2009 It means that q=0 components of the imaginary parts of the spin susceptibilities contribute to 1/T1T1/T_{1}T. For TTT\lesssim T^{*}, (1/T1T)α(1/T_{1}T)_{\alpha} is not proportional to T2T^{2}, because of the Fermi surface.

Refer to caption
Figure 13: Temperature dependence of (1/T1T)α(1/T_{1}T)_{\alpha} in the absence of UU. The red solid and blue dotted lines show (1/T1T)A(1/T_{1}T)_{A} and (1/T1T)B(1/T_{1}T)_{B}, respectively. The dashed longitudinal line indicates TT=TT^{*}\sim0.010.01 eV. The green thin line which is proportional to T2T^{2} is drawn for the eye guide.

IV result in the presence of UU

In this section, we calculate spin fluctuations, the Knight shift, and 1/T1T1/T_{1}T in the presence of UU. It is shown that electron correlation effects are important for explaining the observed temperature dependence of the Knight shift and 1/T1T1/T_{1}T reported by 13C-NMR experiments.T.Sekine.Private According to Eq. 25, the Stoner factor ξs(q)\xi_{s}(\textbf{q}) that has a value close to unity gives major contributions to the spin susceptibility and therefore to the Knight shift and 1/T1T1/T_{1}T. ξs(q=0)1\xi_{s}(\textbf{q}=\textbf{0})\approx 1 contributes to the Knight shift (Eq. 26) and 1/T1T1/T_{1}T. ξs(q=Q)1\xi_{s}(\textbf{q}=\textbf{Q})\approx 1 does not contribute to the Knight shift but to 1/T1T1/T_{1}T (Eq. 27).

IV.1 Spin fluctuations

We calculate the temperature dependence of the Stoner factor ξs(q)\xi_{s}(\textbf{q}) because ξs(q)\xi_{s}(\textbf{q}) is an important measurement of spin fluctuations as discussed in Section II. Figure 14 shows the temperature dependence of ξs(q)\xi_{s}(\textbf{q}), where UU=0.8020.802 and λ\lambda=0.950.95. λ\lambda=0.950.95 was obtained in the calculation of the screened Coulomb interaction using respack.

Refer to caption
Figure 14: Temperature dependence of the Stoner factors ξs(0)\xi_{s}(\textbf{0}) and ξs(Q)\xi_{s}(\textbf{Q}). The red solid and blue dotted lines show ξs(0)\xi_{s}(\textbf{0}) and ξs(Q)\xi_{s}(\textbf{Q}), respectively.

In the case of λ\lambda=0.950.95, ξs(0)\xi_{s}(\textbf{0}) is maximum in the momentum space and increases as TT decreases. The combination of λ\lambda=0.950.95 and UU=0.8020.802 yields ξs(0)\xi_{s}(\textbf{0})=0.9990.999 at TT=0.0030.003 eV. On the other hand, ξs(Q)\xi_{s}(\textbf{Q}) decreases slowly with decreasing TT and is less than ξs(0)\xi_{s}(\textbf{0}). Q is the wavenumber at which χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) has the maximum value. χBBs(q,0)\chi^{s}_{BB}(\textbf{q},0) also has the maximum value at q=Q. The magnitude relationship ξs(0)>ξs(Q)\xi_{s}(\textbf{0})>\xi_{s}(\textbf{Q}) implies that the q=0 magnetic order is easier to induce than SDW at q=Q. At a low temperature of TT\lesssim0.0030.003eV, it is difficult for ξs(q)\xi_{s}(\textbf{q}) to converge in the numerical calculation because enormous Matsubara frequencies and wavenumbers are required.

We explain why ξs(0)\xi_{s}(\textbf{0}) increases at low temperature. In fact, ξs(q)\xi_{s}(\textbf{q}) isn’t directly determined by the DOS because the maximum eigenvalue of U^χ^0(q,0)\hat{U}\hat{\chi}^{0}(\textbf{q},0) contains the products of the χαβ0(q,0)\chi^{0}_{\alpha\beta}(\textbf{q},0). It means that the electron correlation effect is important. We calculate the first- and second-order perturbation terms in Eq. (A) and (B) in Fig. 4. In this study, Eq. (A) and (B) in Fig. 4 are equivalent. Their matrix elements χαβs,1st(q,iωl)\chi^{s,1st}_{\alpha\beta}(\textbf{q},i\omega_{l}) and χαβs,2nd(q,iωl)\chi^{s,2nd}_{\alpha\beta}(\textbf{q},i\omega_{l}) are written as

χαβs,1st(q,iωl)\displaystyle\chi^{s,1st}_{\alpha\beta}(\textbf{q},i\omega_{l}) =\displaystyle= γχαγ0(q,iωl)Uγγχγβ0(q,iωl),\displaystyle\sum_{\gamma}\chi^{0}_{\alpha\gamma}(\textbf{q},i\omega_{l})U_{\gamma\gamma}\chi^{0}_{\gamma\beta}(\textbf{q},i\omega_{l}), (34)
χαβs,2nd(q,iωl)\displaystyle\chi^{s,2nd}_{\alpha\beta}(\textbf{q},i\omega_{l}) =\displaystyle= γ,γχαγ0(q,iωl)Uγγχγγ0(q,iωl)\displaystyle\sum_{\gamma,\gamma^{\prime}}\chi^{0}_{\alpha\gamma}(\textbf{q},i\omega_{l})U_{\gamma\gamma}\chi^{0}_{\gamma\gamma^{\prime}}(\textbf{q},i\omega_{l})
×\displaystyle\times Uγγχγβ0(q,iωl).\displaystyle U_{\gamma^{\prime}\gamma^{\prime}}\chi^{0}_{\gamma^{\prime}\beta}(\textbf{q},i\omega_{l}).

They are the first- and second-order perturbation terms in RPA and correspond to the second and third terms in Eq. (A) or (B) in Fig. 4, respectively. Figure 15 shows the temperature dependence of χAAs,1st(0,0)\chi^{s,1st}_{AA}(\textbf{0},0) and χAAs,2nd(0,0)\chi^{s,2nd}_{AA}(\textbf{0},0). They increase as TT decreases. Note that the zero-order perturbation term is identical to the red dashed line in Fig. 11. Since the off-diagonal elements of χ^0(0,0)\hat{\chi}^{0}(\textbf{0},0) are negative and decrease as TT decreases in Fig. 11, their absolute squares increase as TT decreases. Thus, terms such as χAB0(0,0)UBBχBA0(0,0)\chi^{0}_{AB}(\textbf{0},0)U_{BB}\chi^{0}_{BA}(\textbf{0},0) in Eq. 34 and χAC0(0,0)UCCχCC0(0,0)UCCχCA0(0,0)\chi^{0}_{AC}(\textbf{0},0)U_{CC}\chi^{0}_{CC}(\textbf{0},0)U_{CC}\chi^{0}_{CA}(\textbf{0},0) in Eq. IV.1 increase χAAs,1st(0,0)\chi^{s,1st}_{AA}(\textbf{0},0) and χAAs,2nd(0,0)\chi^{s,2nd}_{AA}(\textbf{0},0) at low temperature. The other higher-order perturbation terms behave similarly. Therefore, ξs(0)\xi_{s}(\textbf{0}) increases as TT decreases.

Refer to caption
Figure 15: First- and second-order perturbation terms of χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0) shown by the brown solid and magenta dotted line, respectively. The horizontal axis represents the temperature.

Fragment orbitals A and C are sensitive to ξs(0)\xi_{s}(\textbf{0}). On the other hand, fragment orbital B is insensitive to ξs(0)\xi_{s}(\textbf{0}) but sensitive to ξs(Q)\xi_{s}(\textbf{Q}). We calculate the matrix elements of χ^s(0,0)\hat{\chi}^{s}(\textbf{0},0) and χ^s(Q,0)\hat{\chi}^{s}(\textbf{Q},0) to exhibit this phenomenon using Eqs. 22 and II. Figure 16(a) shows the temperature dependence of χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0), χBBs(0,0)\chi^{s}_{BB}(\textbf{0},0), χABs(0,0)\chi^{s}_{AB}(\textbf{0},0), and χACs(0,0)\chi^{s}_{AC}(\textbf{0},0). They are real numbers because q=0 and ω\omega=0. χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0)=χCCs(0,0)\chi^{s}_{CC}(\textbf{0},0), χABs(0,0)\chi^{s}_{AB}(\textbf{0},0)=χBAs(0,0)\chi^{s}_{BA}(\textbf{0},0)=χBCs(0,0)\chi^{s}_{BC}(\textbf{0},0)=χCBs(0,0)\chi^{s}_{CB}(\textbf{0},0), and χACs(0,0)\chi^{s}_{AC}(\textbf{0},0)=χCAs(0,0)\chi^{s}_{CA}(\textbf{0},0) are satisfied because of space-inversion symmetry and time-reversal symmetry. The inset shows an enlarged view of the region around χBBs(0,0)\chi^{s}_{BB}(\textbf{0},0) and χABs(0,0)\chi^{s}_{AB}(\textbf{0},0). χBBs(0,0)\chi^{s}_{BB}(\textbf{0},0) and χABs(0,0)\chi^{s}_{AB}(\textbf{0},0) are difficult to increase at low temperature, whereas χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0) sharply increases. Moreover, χACs(0,0)\chi^{s}_{AC}(\textbf{0},0) is negative and sharply decreases as TT decreases. ξs(0)\xi_{s}(\textbf{0}) easily affects fragment orbitals A and C but not fragment orbital B. The situation of χACs(0,0)\chi^{s}_{AC}(\textbf{0},0)<<0 in Fig. 16(a) implies intra-molecular antiferromagnetic fluctuations. The negative off-diagonal elements of the spin susceptibility are due to the characteristic wave function of the Dirac nodal line system.

Figure 16(b) shows the temperature dependence of Re[χAAs(Q,0)]{\rm Re}\left[\chi^{s}_{AA}(\textbf{Q},0)\right], Re[χBBs(Q,0)]{\rm Re}\left[\chi^{s}_{BB}(\textbf{Q},0)\right], Re[χABs(Q,0)]{\rm Re}\left[\chi^{s}_{AB}(\textbf{Q},0)\right], and Re[χACs(Q,0)]{\rm Re}\left[\chi^{s}_{AC}(\textbf{Q},0)\right]. Re[χ^s(Q,0)]{\rm Re}\left[\hat{\chi}^{s}(\textbf{Q},0)\right] reflects ξs(Q)\xi_{s}(\textbf{Q}) and slowly varies with temperature. At q=Q, Re[χBBs(Q,0)]{\rm Re}\left[\chi^{s}_{BB}(\textbf{Q},0)\right] is the largest of all matrix elements and increases with temperature. Fragment orbital B is sensitive to ξs(Q)\xi_{s}(\textbf{Q}). Because Re[χABs(Q,0)]{\rm Re}\left[\chi^{s}_{AB}(\textbf{Q},0)\right]<<0 and Re[χACs(Q,0)]{\rm Re}\left[\chi^{s}_{AC}(\textbf{Q},0)\right]>>0, the spins of the fragment orbitals B and A(C) within a molecule are inversely correlated.

Figure 17 schematically illustrates the spin polarization pictures that we obtained from the calculated results in Fig. 16 for the paramagnetic regime. Figure 17(a) corresponds to the case in Fig. 16 (a), where we found intra-molecular antiferromagnetic spin fluctuations, which are commensurate(q=0) between molecules. The solid arrows in Fig. 17 (a) represents a tendency where an infinitesimally small downward local magnetic field at the orbital C(A) induces an upward spin polarization at the orbital A(C) by the linear response relation MA(C)M_{A(C)}=χAC(CA)sHC(A)\chi^{s}_{AC(CA)}H_{C(A)}, respectively [MM is the magnetization and HH is the infinitesimal magnetic field]. Figure 17(b), which is derived from Fig. 16(b), stand for the spin fluctuations within a molecule that are incommensurate (q=Q) between molecules. In this case, the spins at the orbitals A(=C) and B tend to be inversely correlated within a molecule.

Refer to caption
Refer to caption
Figure 16: (a) Temperature dependence of χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0), χBBs(0,0)\chi^{s}_{BB}(\textbf{0},0), χABs(0,0)\chi^{s}_{AB}(\textbf{0},0), and χACs(0,0)\chi^{s}_{AC}(\textbf{0},0) shown by the red dashed, blue solid, green dotted, and purple chain lines, respectively. The inset shows an enlarged view of the region around χBBs(0,0)\chi^{s}_{BB}(\textbf{0},0) and χABs(0,0)\chi^{s}_{AB}(\textbf{0},0). (b) Temperature dependence of Re[χAAs(Q,0)]{\rm Re}\left[\chi^{s}_{AA}(\textbf{Q},0)\right], Re[χBBs(Q,0)]{\rm Re}\left[\chi^{s}_{BB}(\textbf{Q},0)\right], Re[χABs(Q,0)]{\rm Re}\left[\chi^{s}_{AB}(\textbf{Q},0)\right], and Re[χACs(Q,0)]{\rm Re}\left[\chi^{s}_{AC}(\textbf{Q},0)\right]. The combination of matrix elements and lines is the same as in (a).
Refer to caption
Figure 17: Schematic illustrations of spin polarization. (a) intra-molecular antiferromagnetic spin fluctuations linked to the q=0 response shown in Fig. 16(a). (b) Spin correlations within a molecule given by the q=Q response in Fig. 16(b).

Next, we show the diagonal elements of the spin susceptibilities χααs(q,ω)\chi^{s}_{\alpha\alpha}(\textbf{q},\omega) at TT=0.0030.003 eV. Figure 18 (a), (b), (c), and (d) show χAAs(q,0)\chi^{s}_{AA}(\textbf{q},0), χBBs(q,0)\chi^{s}_{BB}(\textbf{q},0), Im[χAAs(q,ω0)]{\rm Im}[\chi^{s}_{AA}(\textbf{q},\omega_{0})], and Im[χBBs(q,ω0)]{\rm Im}[\chi^{s}_{BB}(\textbf{q},\omega_{0})] in the qbq_{b}-qcq_{c} plane, respectively. χααs(q,0)\chi^{s}_{\alpha\alpha}(\textbf{q},0) is a real number. We fix qaq_{a}=0 in Fig. 18 (a) and (c) and qaq_{a}=0.2π0.2\pi in Fig. 18 (b) and (d). Furthermore, ω0\omega_{0} is equal to 0.0010.001 eV. Re[χAAs(q,0)]{\rm Re}[\chi^{s}_{AA}(\textbf{q},0)] and Im[χAAs(q,ω0)]{\rm Im}[\chi^{s}_{AA}(\textbf{q},\omega_{0})] have very large values at q=0 because ξs(0)\xi_{s}(\textbf{0})=0.9990.999. However, the BB component is difficult to be affected by ξs(0)\xi_{s}(\textbf{0}) but easily affected by ξs(Q)\xi_{s}(\textbf{Q}). Q is the wavenumber at which χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) and χBBs(q,0)\chi^{s}_{BB}(\textbf{q},0) have the maximum values. χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0) and Im[χAAs(0,ω0)]{\rm Im}[\chi^{s}_{AA}(\textbf{0},\omega_{0})] are much larger than χBBs(Q,0)\chi^{s}_{BB}(\textbf{Q},0) and Im[χBBs(Q,ω0)]{\rm Im}[\chi^{s}_{BB}(\textbf{Q},\omega_{0})] because ξs(0)>ξs(Q)\xi_{s}(\textbf{0})>\xi_{s}(\textbf{Q}) and the spin susceptibility obtained using RPA is determined by 1/(1ξs(q))1/\left(1-\xi_{s}(\textbf{q})\right). χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0) and Im[χAAs(0,ω0)]{\rm Im}[\chi^{s}_{AA}(\textbf{0},\omega_{0})] in Fig. 18 (a) and (c) decrease with temperature. On the other hand, χBBs(Q,0)\chi^{s}_{BB}(\textbf{Q},0) and Im[χBBs(Q,ω0)]{\rm Im}[\chi^{s}_{BB}(\textbf{Q},\omega_{0})] in Fig. 18 (b) and (d) slowly increase and the peaks become broad with temperature. These behavior result from temperature dependence of ξs(q)\xi_{s}(\textbf{q}) in Fig. 14

Refer to caption
Figure 18: The momentum dependences of the diagonal elements of the spin susceptibility in the presence of UU (a) χAAs(q,0)\chi^{s}_{AA}(\textbf{q},0) in the qbq_{b}qcq_{c} plane, where qaq_{a}=0. (b) χBBs(q,0)\chi^{s}_{BB}(\textbf{q},0) in the qbq_{b}qcq_{c} plane, where qaq_{a}=0.2π0.2\pi. (c) Im[χAAs(q,ω0)]{\rm Im}[\chi^{s}_{AA}(\textbf{q},\omega_{0})] in the qbq_{b}qcq_{c} plane, where qaq_{a}=0. (d) Im[χBBs(q,ω0)]{\rm Im}[\chi^{s}_{BB}(\textbf{q},\omega_{0})] in the qbq_{b}qcq_{c} plane, where qaq_{a}=0.2π0.2\pi. The temperature TT=0.0030.003 eV.

IV.2 Knight shift and 1/T1T1/T_{1}T in the presence of UU

In this subsection, we solve Eq. 26 and 27 to investigate the effects of the fluctuations on the Knight shift and 1/T1T1/T_{1}T.

Fig. 19 shows the temperature dependence of the Knight shift, where UU=0.8020.802 and λ\lambda=0.950.95. KBK_{B} in the presence of UU is larger than that in the absence of UU. However, KAK_{A} and KCK_{C} in the presence of UU are smaller than those in the absence of UU. In the case of UU=0.8020.802 and λ\lambda=0.950.95, KAK_{A} and KCK_{C} are negative, while KBK_{B} and Ktot(=KA+KB+KC)K_{tot}(=K_{A}+K_{B}+K_{C}) are positive. Similar behavior was previously observed in the organic conductor α\alpha-(BEDT-TTF)2I3.A.Kobayashi2013

KAK_{A} and KCK_{C} don’t increase at low temperature, although the Stoner factor ξs(0)\xi_{s}(\textbf{0}) is almost 11 at TT=0.0030.003 eV. The behavior of the Knight shift is understood by considering the off-diagonal elements of χ^s(0,0)\hat{\chi}^{s}(\textbf{0},0). Because χAAs(0,0)\chi^{s}_{AA}(\textbf{0},0) and χACs(0,0)\chi^{s}_{AC}(\textbf{0},0) have opposite signs in Fig. 16(a), their cancelation prevents the increase of the Knight shift in Eq. 26. In other words, the q=0 spin fluctuations are not observed in the Knight shift because of the intra-molecular antiferromagnetic fluctuations.

Refer to caption
Figure 19: Temperature dependence of the Knight shift KαK_{\alpha} for UU=0.8020.802 and λ\lambda=0.950.95. The red dashed and blue dotted lines show KAK_{A} and KBK_{B}, respectively. The black solid line shows the total Knight shift KtotK_{\rm tot}=KA+KB+KCK_{A}+K_{B}+K_{C}.

Next, we solve Eq. 27. The spin-lattice relaxation rate, 1/T1T1/T_{1}T, is determined by qIm[χααs(q,ω0)]\sum_{\textbf{q}}{\rm Im}[\chi^{s}_{\alpha\alpha}(\textbf{q},\omega_{0})]. Im[χAAs(q,ω0)]{\rm Im}[\chi^{s}_{AA}(\textbf{q},\omega_{0})] and Im[χBBs(q,ω0)]{\rm Im}[\chi^{s}_{BB}(\textbf{q},\omega_{0})] are shown in Fig. 18. Figure 20 shows the temperature dependence of 1/T1T1/T_{1}T, where λ\lambda=0.950.95 and UU=0.8020.802.

Refer to caption
Figure 20: Temperature dependence of (1/T1T)α(1/T_{1}T)_{\alpha} for UU=0.8020.802 and λ\lambda=0.950.95. The red solid and blue dotted lines show (1/T1T)A(1/T_{1}T)_{A} and (1/T1T)B(1/T_{1}T)_{B}, respectively.

At high temperature, 1/T1T1/T_{1}T values for all orbitals decrease with decreasing TT. However, at low temperature, the (1/T1T)A(1/T_{1}T)_{A} starts to increase because orbital A is easily affected by ξs(0)\xi_{s}(\textbf{0}). The behavior of (1/T1T)C(1/T_{1}T)_{C} is identical to that of (1/T1T)A(1/T_{1}T)_{A} because of space-inversion symmetry. Although (1/T1T)B(1/T_{1}T)_{B} is difficult to be affected by ξs(0)\xi_{s}(\textbf{0}) in Fig. 20, ξs(0)1\xi_{s}(\textbf{0})\rightarrow 1 makes (1/T1T)B(1/T_{1}T)_{B}\rightarrow\infty. This is because ξs(0)1\xi_{s}(\textbf{0})\rightarrow 1 makes 1/(1ξs(0))1/\left(1-\xi_{s}(\textbf{0})\right)\rightarrow\infty and RPA imposes the factor 1/(1ξs(0))1/\left(1-\xi_{s}(\textbf{0})\right) on the spin susceptibility (Eq. 25). For TT\gtrsim0.0050.005 eV, (1/T1T)B(1/T_{1}T)_{B} is more dominant than (1/T1T)A(1/T_{1}T)_{A} and (1/T1T)C(1/T_{1}T)_{C} because ξs(Q)\xi_{s}(\textbf{Q}) slowly increases with TT and orbital B is easier to be affected by ξs(Q)\xi_{s}(\textbf{Q}). Q is the wavenumber at which χBB0(q,0)\chi^{0}_{BB}(\textbf{q},0) and χBBs(q,0)\chi^{s}_{BB}(\textbf{q},0) have the maximum values.

In the case of a small λ\lambda, such as λ\lambda=0.790.79, ξs(Q)\xi_{s}(\textbf{Q}) is larger than ξs(0)\xi_{s}(\textbf{0}), and SDW can be induced. However, the incommensurate spin fluctuations in this material are suppressed at low temperature in Fig. 14. Similar behavior occurs in the case of a small λ\lambda. If UU is sufficiently large, ξs(Q)\xi_{s}(\textbf{Q}) reaches 11 at low temperature. However, the magnetic transition to the SDW phase would have already been induced at a high temperature. Thus, it is difficult to explain the upturn of the 1/T1T1/T_{1}T curve near 3030 K, which is observed in the 13C-NMR experiment, by the incommensurate spin fluctuations. We calculate λ\lambda using the respack program. Because λ\lambda=0.790.79 is obtained as the ratio of diagonal elements with the unscreened on-site Coulomb interaction while λ\lambda=0.950.95 is obtained as that with the screened on-site Coulomb interaction, we consider that λ\lambda=0.950.95 is the more realistic ratio.

In this subsection, we showed that the Knight shift does not increase at low temperature because of the intra-molecular antiferromagnetic fluctuations. We also showed that the (1/T1T)A(1/T_{1}T)_{A} and (1/T1T)C(1/T_{1}T)_{C} start to increase at low temperature because of the behavior of ξs(0)\xi_{s}(\textbf{0}). They are dominant in TT\lesssim0.0050.005 eV, whereas (1/T1T)B(1/T_{1}T)_{B} is dominant in TT\gtrsim0.0050.005 eV because of the temperature dependence of ξs(Q)\xi_{s}(\textbf{Q}) and fragment-orbital-dependence of spin susceptibilities. Fig. 21 summarizes the fragment-orbital dependence of the Fermi surface, non-interacting spin susceptibilities, Stoner factors, and 1/T1T1/T_{1}T. Fig. 21 shows the important factors for orbitals A and B. ξs(0)\xi_{s}(\textbf{0}) is the main contributor to orbital A and causes the upturn of the (1/T1T)A(1/T_{1}T)_{A} curve at low temperature. However, the contribution of ξs(0)\xi_{s}(\textbf{0}) to orbital B is small. Orbital B is sensitive to ξs(Q)\xi_{s}(\textbf{Q}), which dominantly contributes to (1/T1T)B(1/T_{1}T)_{B} at high temperature but does not contribute to orbital A as much. These fragment-orbital-dependent magnetic properties are caused by the presence of ZR because ZR biases ρB(k,0)\rho_{B}(\textbf{k},0) in Eq. 30 to a part of the Fermi surface.

Refer to caption
Figure 21: Correspondence table between the fragment orbitals, shape of the Fermi surface, momenta at which the non-interacting spin susceptibilities have peaks, Stoner factors, and 1/T1T1/T_{1}T.

V conclusion

In this study, we found that multiple fragment orbitals play important roles in the magnetic properties of [Ni(dmdt)2]. On orbitals A and C, which are unevenly distributed toward one side of a molecule, the q=0 spin fluctuations are enhanced, whereas the incommensurate spin fluctuations are enhanced on orbital B, which is centered on the Ni atom. Because of the q=0 spin fluctuations, the A and C components of 1/T1T1/T_{1}T start to increase as TT decreases at low temperature. However, the q=0 spin fluctuations do not affect the Knight shift, because they are intra-molecular antiferromagnetic fluctuations. The reason why the q=0 spin fluctuations is enhanced is understood from the perturbation process and the off-diagonal elements of the non-interacting spin susceptibility. The incommensurate spin fluctuations dominantly contribute to the B component of 1/T1T1/T_{1}T at high temperature. These fragment-orbital-dependent quantities result from the presence of ZR. If no ZR exists in the Brillouin zone, the BB components of the spin susceptibilities do not have the maximum value at the incommensurate wavenumber, because the spectral weight of fragment orbital B at EFE_{F} may be similar to those of A and C. Because ZR biases ρB(k,0)\rho_{B}(\textbf{k},0) in Eq. 30 to a part of the Fermi surface, the BB component of the spin susceptibility has a maximum value at the incommensurate wavenumber q=Q\textbf{q}=\textbf{Q}. Thus, the wavenumber dependence of the spin susceptibilities is different between the BB component and AA(CC) component. ZR is a characteristic of materials with a Dirac nodal line system described by an nn-band model(n3n\geq 3). Thus, it is expected that the fragment-orbital-dependent properties due to ZR will be found in other Dirac nodal line systems. Moreover, it is predicted that transition-metal substitution in the Ni(dmdt)2 molecule controls spin fluctuations because it changes λ\lambda and UU. In the two-dimensional Dirac electron system under the charge-neutral condition, the spin fluctuations are weak because the Fermi surface is identical to the Dirac points. On the other hand, the spin fluctuations are enhanced in [Ni(dmdt)2] by the Fermi surface. The Fermi surface arises from transfer integrals in the nodal line direction. This is the three-dimensionality of this material.

In the 13C-NMR experiment, 1/T1T1/T_{1}T has a peak at TT\sim3030 K. The experiment was performed for a sample in which C atoms were replaced with 13C. Fig. 2(b) shows a Ni(dmdt)2 molecule, and the red dashed circles surround 13C atoms. Therefore, orbitals A and C mainly contribute to the physical quantities observed in the 13C-NMR experiment. Thus, our calculation of 1/T1T1/T_{1}T is almost consistent with the experiment. Furthermore, we expect that the B component of 1/T1T1/T_{1}T can be observed by experiments using a sample in which 12C atoms near the Ni atom are substituted by 13C. On the other hand, the A and C components of the Knight shift obtained in our calculation are negative, but the Knight shift observed in the 13C-NMR experiment is positive. Therefore, we consider the following possible electronic states at TT\lesssim3030 K. The first is the q=0 magnetic ordered state, which is the intra-molecular antiferromagnetism is realized at TT\lesssim3030 K. In this case, it is considered that the Knight shift observed in the experiment is attributed to the sum of KAK_{A} and KBK_{B}, which is positive. In the second possible electronic state, UU is not so large that KAK_{A} is negative. In this case, it is considered that another ordered state is induced at TT\lesssim3030 K and that the 1/T1T1/T_{1}T curve is upturned by the fluctuations corresponding to the order. Examples of such orders are bond order and topological order.

Acknowledgements.
The authors thank T. Sekine, T. Hatamura, K. Sunami, K. Miyagawa, K. Kanoda, B. Zhou, Akiko Kobayashi, and K. Yoshimi for fruitful discussions. This work was financially supported by MEXT (JP) JSPS (Grant No. 15K05166) and JST SPRING (Grant No. JPMJSP2125). T. Kawamura acknowledges support from the “Interdisciplinary Frontier Next-Generation Researcher Program of the Tokai Higher Education and Research System.” The computation in this work was performed using the facilities of the Supercomputer Center, Institute for Solid State Physics, University of Tokyo.

References

  • (1) T. Ando, T. Nakanishi, and R. Saito, J. Phys. Soc. Jpn. 67, 2857 (1998).
  • (2) V. P. Gusynin and S. G. Sharapov, Phys. Rev. B 73, 245411 (2006).
  • (3) S. Murakami, N. Nagaosa, and S.-C. Zhang, Phys. Rev. Lett. 93, 156804 (2004).
  • (4) D. Hsieh, D. Qian, L. Wray, Y. Xia, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Nature 452, 970–974 (2008).
  • (5) H. Fukuyama and R. Kubo, J. Phys. Soc. Jpn. 28, 570 (1970).
  • (6) H. Fukuyama, J. Phys. Soc. Jpn. 76, 043711 (2007).
  • (7) A. A. Abrikosov and S. D. Beneslavskii, Zh. Eksp. Teor. Fiz. 59, 1280 (1971).
  • (8) J. González, F. Guinea, and M. A. H. Vozmediano, Nucl. Phys., Sect. B 424, 595 (1994).
  • (9) J. González, F. Guinea, and M. A. H. Vozmediano, Phys. Rev.B 59, R2474(R) (1999).
  • (10) V. N. Kotov, B. Uchoa, V. M. Pereira, F. Guinea, and A. H. Castro Neto, Rev. Mod. Phys. 84, 1067 (2012).
  • (11) T. O. Wehling, A. M. Black-Schaffer, and A. V. Balatsky, Adv. Phys. 63, 1 (2014).
  • (12) W. Witczak-Krempa, G. Chen, Y. B. Kim, and L. Balents, Annu. Rev. Condens. Matter Phys. 5, 57 (2014).
  • (13) K. Kajita, T. Ojiro, H. Fujii, Y. Nishio, H. Kobayashi, A. Kobayashi, and R. Kato, J. Phys. Soc. Jpn. 61, 23 (1992).
  • (14) N. Tajima, M. Tamura, Y. Nishio, K. Kajita, and Y. Iye, J. Phys. Soc. Jpn. 69, 543 (2000).
  • (15) A. Kobayashi, S. Katayama, K. Noguchi, and Y. Suzumura, J. Phys. Soc. Jpn. 73, 3135 (2004).
  • (16) S. Katayama, A. Kobayashi, and Y. Suzumura, J. Phys. Soc. Jpn. 75, 054705 (2006).
  • (17) A. Kobayashi, S. Katayama, Y. Suzumura, and H. Fukuyama, J. Phys. Soc. Jpn. 76, 034711 (2007).
  • (18) M. O. Goerbig, J. N. Fuchs, G. Montambaux, and F. Piéchon, Phys. Rev. B 78, 045415 (2008).
  • (19) K. Kajita, Y. Nishio, N. Tajima, Y. Suzumura, and A. Kobayashi, J. Phys. Soc. Jpn. 83, 072002 (2014).
  • (20) H. Seo, J. Phys. Soc. Jpn. 69, 805 (2000).
  • (21) T. Takahashi, Synth. Met. 133-134, 261 (2003).
  • (22) T. Kakiuchi, Y. Wakabayashi, H. Sawa, T. Takahashi, and T. Nakamura, J. Phys. Soc. Jpn. 76, 113702 (2007).
  • (23) K. Ishikawa, M. Hirata, D. Liu, K. Miyagawa, M. Tamura, and K. Kanoda, Phys. Rev. B 94, 085154 (2016).
  • (24) Y. Tanaka and M. Ogata, J. Phys. Soc. Jpn. 85, 104706 (2016).
  • (25) R. Beyer, A. Dengl, T. Peterseim, S. Wackerow, T. Ivek, A. V. Pronin, D. Schweitzer, and M. Dressel, Phys. Rev. B 93, 195116 (2016).
  • (26) D. Liu, K. Ishikawa, R. Takehara, K. Miyagawa, M. Tamura, and K. Kanoda, Phys. Rev. Lett. 116, 226401 (2016).
  • (27) D. Ohki, Y. Omori, and A. Kobayashi, Phys. Rev. B 100, 075206 (2019).
  • (28) M. Hirata, K. Ishikawa, K. Miyagawa, M. Tamura, C. Berthier, D. Basko, A. Kobayashi, G. Matsuno, and K. Kanoda, Nat. Commun. 7, 12666 (2016).
  • (29) M. Hirata, K. Ishikawa, G. Matsuno, A. Kobayashi, K. Miyagawa, M. Tamura, C. Berthier, and K. Kanoda, Science 358, 1403 (2017).
  • (30) A. Kobayashi and Y. Suzumura, J. Phys. Soc. Jpn. 82, 054715 (2013).
  • (31) D. Ohki, M. Hirata, T. Tani, K. Kanoda, and A. Kobayashi, Phys. Rev. Res. 2, 033479 (2020).
  • (32) M. Hirata, A. Kobayashi, C. Berthier, and K. Kanoda, Rep. Prog. Phys. 84, 036502 (2021).
  • (33) A. Burkov, M. D. Hook, and L. Balents, Phys. Rev. B 84, 235126 (2011).
  • (34) C. K. Chiu and A. P. Schnyder, Phys. Rev. B 90, 205136 (2014).
  • (35) C. Fang, Y. Chen, H. Y. Kee, and L. Fu, Phys. Rev. B 92, 081201 (2015).
  • (36) Z. Gao, M. Hua, H. Zhang, and X. Zhang, Phys. Rev. B 93, 205109 (2016).
  • (37) P. R. Wallace, Phys. Rev. 71, 622 (1947).
  • (38) H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and X. Dai, Phys. Rev. X 5, 011029 (2015).
  • (39) Y. Kim, B. J. Wieder, C. L. Kane, and A. M. Rappe, Phys. Rev. Lett. 115, 036806 (2015).
  • (40) R. Yu, H. Weng, Z. Fang, X. Dai, and X. Hu, Phys. Rev. Lett. 115, 036807 (2015).
  • (41) J. M. Carter, V. V. Shankar, M. A. Zeb, and H. Y. Kee, Phys. Rev. B 85, 115105 (2012).
  • (42) A. Yamakage, Y. Yamakawa, Y. Tanaka, and Y. Okamoto, J. Phys. Soc. Jpn. 85, 013708 (2016).
  • (43) R. Kato, H. Cui, T. Tsumuraya, T. Miyazaki, and Y. Suzumura, J. Am. Chem. Soc. 139, 1770 (2017).
  • (44) R. Kato and Y. Suzumura, J. Phys. Soc. Jpn. 86, 064705 (2017).
  • (45) Y. Suzumura, J. Phys. Soc. Jpn. 86, 124710 (2017).
  • (46) Y. Suzumura and R. Kato, Jpn. J. Appl. Phys. 56, 05FB02 (2017).
  • (47) Y. Suzumura, H. Cui, and R. Kato, J. Phys. Soc. Jpn. 87, 084702 (2018).
  • (48) Y. Suzumura and A. Yamakage, J. Phys. Soc. Jpn. 87, 093704 (2018).
  • (49) T. Tsumuraya, R. Kato, and Y. Suzumura, J. Phys. Soc. Jpn. 87, 113701 (2018).
  • (50) Y. Suzumura, T. Tsumuraya, R. Kato, H. Matsuura, and M. Ogata, J. Phys. Soc. Jpn. 88, 124704 (2019).
  • (51) B. Zhou, S. Ishibashi, T. Ishii, T. Sekine, R. Takehara, K. Miyagawa, K. Kanoda, E. Nishibori, and A. Kobayashi, Chem. Commun. 55, 3327 (2019).
  • (52) R.Kato and Y.Suzumura, J. Phys. Soc. Jpn. 89, 044713 (2020).
  • (53) T. Kawamura, D. Ohki, B. Zhou, A. Kobayashi, and A. Kobayashi, J. Phys. Soc. Jpn. 89, 074704 (2020).
  • (54) T. Kawamura, B. Zhou, A. Kobayashi, and A. Kobayashi, J. Phys. Soc. Jpn. 90, 064710 (2021).
  • (55) A. Kobayashi, B. Zhou, R. Takagi, K. Miyagawa, S. Ishibashi, A. Kobayashi, T. Kawamura, E. Nishibori, and K. Kanoda, Bull. Chem. Soc. Jpn. 94, 2540–2562 (2021).
  • (56) J. W. Rhim and Y. B. Kim, Phys. Rev. B 92, 045126 (2015).
  • (57) A. K. Mitchell and L. Fritz, Phys. Rev. B 92, 121109 (2015).
  • (58) S. T. Ramamurthy and T. L. Hughes, Phys. Rev. B 95, 075138 (2017).
  • (59) N. H. Shon and T. Ando, J. Phys. Soc. Jpn. 67, 2421 (1998).
  • (60) T. Sekine, T. Hatamura, K. Sunami, K. Miyagawa, and K. Kanoda, Priv. Commun.
  • (61) H. Seo, S. Ishibashi, Y. Okano, H. Kobayashi, A. Kobayashi, H. Fukuyama, and K. Terakura, J. Phys. Soc. Jpn. 77, 023714 (2008).
  • (62) H. Seo, S. Ishibashi, Y. Otsuka, H. Fukuyama, and K. Terakura, J. Phys. Soc. Jpn. 82, 054711 (2013).
  • (63) M. Tsuchiizu, Y. Omori, Y. Suzumura, M.-L. Bonnet, and V. Robert, J. Chem. Phys. 136, 044519 (2012).
  • (64) M. Tsuchiizu, Y. Omori, Y. Suzumura, M-L. Bonnet, V. Robert, S. Ishibashi, and H. Seo, J. Phys. Soc. Jpn. 80, 013703 (2011).
  • (65) K. Nakamura, Y. Yoshimoto, Y. Nomura, T. Tadano, M. Kawamura, T. Kosugi, K. Yoshimi, T. Misawa, and Y. Motoyama, arXiv:2001.02351.
  • (66) P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. Buongiorno Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, M. Cococcioni, N. Colonna, I. Carnimeo, A. Dal Corso, S. de Gironcoli, P. Delugas, R. A. DiStasio Jr., A. Ferretti, A. Floris, G. Fratesi, G. Fugallo, R. Gebauer, U. Gerstmann, F. Giustino, T. Gorni, J. Jia, M. Kawamura, H.-Y. Ko, A. Kokalj, E. Küçükbenli, M. Lazzeri, M. Marsili, N. Marzari, F. Mauri, N. L. Nguyen, H.-V. Nguyen, A. Otero-de-la-Roza, L. Paulatto, S. Poncé, D. Rocca, R. Sabatini, B. Santra, M. Schlipf, A. P. Seitsonen, A. Smogunov, I. Timrov, T. Thonhauser, P. Umari, N. Vast, X. Wu, and S. Baroni, J. Phys. Condens. Matter 29, 465901 (2017).
  • (67) S. Katayama, A. Kobayashi, and Y. Suzumura, Eur. Phys. J. B. 67, 139-148 (2009).