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

Odd-parity spin-loop-current order mediated by transverse spin fluctuations
in cuprates and related electron systems

Hiroshi Kontani, Youichi Yamakawa, Rina Tazai, and Seiichiro Onari Department of Physics, Nagoya Furo-cho, Nagoya 464-8602, Japan.
Abstract

Unconventional symmetry-breaking phenomena due to nontrivial order parameters attract increasing attention in strongly correlated electron systems. Here, we predict theoretically the occurrence of nanoscale spontaneous spin-current, called the spin loop-current (sLC) order, as a promising origin of the pseudogap and electronic nematicity in cuprates. We reveal that the sLC is driven by the odd-parity electron-hole condensation that are mediated by transverse spin fluctuations around the pseudogap temperature TT^{*}. At the same temperature, odd-parity magnon pair condensation occurs. The sLC order is “hidden” in that neither internal magnetic field nor charge density modulation is induced, whereas the predicted sLC with finite wavenumber naturally gives the Fermi arc structure. In addition, the fluctuations of sLC order work as attractive pairing interaction between adjacent hot spots, which enlarges the dd-wave superconducting transition temperature TcT_{\rm c}. The sLC state will be a key ingredient in understanding the pseudogap, electronic nematicity as well as superconductivity in cuprates and other strongly correlated metals.

I Introduction

Various unconventional symmetry-breaking phenomena, such as violations of rotational and parity symmetries, have been discovered in many strongly correlated metals recently. This fact strongly indicates the emergence of exotic density-wave orders, which are totally different from usual spin/charge density waves. Various exotic symmetry-breaking phenomena, such as violations of rotational and parity symmetries, are the central issues in cuprate high-TcT_{\rm c} superconductors. However, their microscopic mechanisms still remain as unsolved issues. Figure 1 (a) shows a schematic phase diagram of cuprate superconductors. Below TCDW200T_{\rm CDW}\sim 200K, a stripe charge-channel density-wave emerges at finite wavevector 𝒒(π/2,0){{\bm{q}}}\approx(\pi/2,0) in many compounds Y-Xray1 ; Bi-Xray1 ; STM-Kohsaka ; STM-Fujita , which produces the Fermi arc structure and causes a reduction in the density-of-states (DOS). However, it cannot be the origin of the pseudogap temperature TT^{*} since T>TCDWT^{*}>T_{\rm CDW}. Short quasiparticle lifetime due to spin or charge fluctuations could reduce the DOS Tremblay ; Scalapino ; Moriya .

Refer to caption
Figure 1: (a) Possible phase-diagram of hole-doped cuprate superconductors. The sLC phase is obtained by the present study. (b) Current of spin (σ=±1{\sigma}=\pm 1) from site 1 to site 2. σ=±1{\sigma}=\pm 1 is the spin of the electron. (c)(d) Schematic pictures of the diagonal sLC at 𝒒sLC=(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,\pi/2) and the axial sLC at 𝒒sLC=(π/2,0){{\bm{q}}}_{\rm sLC}=(\pi/2,0), respectively.

Recently, much experimental evidence for the phase transition at TT^{*} has been accumulated RUS ; ARPES-Science2011 ; Y-Sato ; Hg-Murayama ; Fujimori-nematic ; Shibauchi-nematic . Various fascinating order parameters have been proposed and actively investigated, such as the CDW or bond-order (BO) Bulut ; Chubukov ; Chubukov-AL ; Sachdev ; Metzner ; DHLee-PNAS ; Kivelson-NJP ; Yamakawa-CDW ; Tsuchiizu-CDW ; Kawaguchi-CDW , the pair-density-wave PALee ; Agterberg , and the charge loop-current (cLC) order Varma ; Affleck ; FCZhang ; Schultz . At present, the symmetries of the hidden order in the pseudo-gap phase are not yet confirmed experimentally. Thus, it is necessary to study various possibilities without prejudice based on advanced many-body theories Chubukov ; Chubukov-AL ; Sachdev ; Metzner ; DHLee-PNAS ; Kivelson-NJP ; Yamakawa-CDW ; Tsuchiizu-CDW ; Kawaguchi-CDW ; PALee ; Agterberg ; Varma ; Affleck ; FCZhang ; Schultz .

Let us discuss the symmetry breaking in the correlated hopping between sites ii and jj; ti,jti,j+δti,jt_{i,j}\rightarrow t_{i,j}+\delta t_{i,j}, where δti,j(=δtj,i)\delta t_{i,j}(=\delta t_{j,i})^{*} is the order parameter. Then, the BO is given by a real and even-parity δti,j\delta t_{i,j} Bulut ; Chubukov ; Sachdev ; Metzner ; DHLee-PNAS ; Kivelson-NJP ; Yamakawa-CDW ; Tsuchiizu-CDW ; Kawaguchi-CDW . A spin-fluctuation mechanism Tsuchiizu-CDW ; Kawaguchi-CDW predicts the ferro (𝒒=𝟎{{\bm{q}}}={\bm{0}}) dd-wave BO state at TT^{*} and stripe (𝒒(π/2,0){{\bm{q}}}\approx(\pi/2,0)) BO at TCDWT_{\rm CDW}. The former order explains the experimental nematic transition Y-Sato ; Shibauchi-nematic . However, simple translational symmetry preserving ferro-BO does not explain the pseudogap formation. OAlso, the cLC order is given by a pure imaginary and odd-parity δti,j\delta t_{i,j} Varma ; Affleck ; FCZhang ; Schultz . Both order parameters have been actively investigated.

In contrast, spin current flows if pure imaginary order parameter is odd under space and spin inversions; δti,jσ=δtj,iσ=δti,jσ\delta t_{i,j}^{{\sigma}}=-\delta t_{j,i}^{{\sigma}}=-\delta t_{i,j}^{-{\sigma}} as shown in Fig. 1 (b) Schultz ; Nersesyan ; Ozaki ; Ikeda ; Fujimoto ; Sr2IrO4 . Here, σ=±1{\sigma}=\pm 1 represents the spin of the electron. Figures 1 (c) and (d) depict the spin loop-current (sLC) order at the wavevectors 𝒒sLC=(δ,δ){{\bm{q}}}_{\rm sLC}=(\delta,\delta) and that at 𝒒sLC=(δ,0){{\bm{q}}}_{\rm sLC}=(\delta,0) with δ=π/2\delta=\pi/2, respectively. The sLC is a hidden order in the sense that no internal magnetic field appears, and charge density modulation is quite small. Nonetheless, the sLC is very attracting since the pseudogap and Fermi surface (FS) reconstruction are induced by band-folding if 𝒒sLC𝟎{{\bm{q}}}_{\rm sLC}\neq{\bm{0}}.

In this paper, we discover the emergence of “hidden symmetry breaking” accompanied by finite spin current at 𝒒sLC(π/2,π/2){{\bm{q}}}_{\rm sLC}\approx(\pi/2,\pi/2). This sLC order originates from the spin-flipping magnon-exchange process, called the Aslamazov-Larkin (AL) process Onari-SCVC ; Onari-FeSe ; Yamakawa-FeSe ; Chubukov-AL . The sLC order is hidden in that neither internal magnetic field nor charge density modulation is induced, while the band-folding by these sLC orders produces the Fermi arc structure and pseudogap in the DOS ARPES-Science2011 ; Yoshida-arc . The derived transition temperature TsLCT_{\rm sLC} is higher than that of the stripe-BO, and comparable to that of the ferro-BO. The sLC order will be responsible for the pseudogap and electronic nematicity not only in cuprates, but also in iridates and ff-electron systems Sr2IrO4 ; Sr2IrO4-ARPES ; Ikeda ; Fujimoto .

The emergence of the sLC has been discussed in various electronic systems Schultz ; Ozaki ; Nersesyan ; Sr2IrO4 ; Ikeda . From the microscopic viewpoint, however, the mechanism of the sLC is highly nontrivial, since the realization condition of the sLC order is very severe in the extended UU-VV-JJ Hubbard model within the mean-field theory Nersesyan . In addition, only the case 𝒒sLC=(π,π){{\bm{q}}}_{\rm sLC}=(\pi,\pi) was analyzed in previous works. The present study can explain the sLC order based on a simple Hubbard model with on-site UU, without assuming the wavevector 𝒒sLC{{\bm{q}}}_{\rm sLC}.

II Spin-fluctuation-driven unconventional orders

II.1 Model Hamiltonian

Here, we analyze the single-orbital square-lattice Hubbard model

H=𝒌,σϵ𝒌c𝒌σc𝒌σ+Uinini.\displaystyle H=\sum_{{{\bm{k}}},{\sigma}}{\epsilon}_{{\bm{k}}}c_{{{\bm{k}}}{\sigma}}^{\dagger}c_{{{\bm{k}}}{\sigma}}+U\sum_{i}n_{i\uparrow}n_{i\downarrow}. (1)

We denote the hopping integrals (t1,t2,t3)=(1,1/6,1/5)(t_{1},t_{2},t_{3})=(-1,1/6,-1/5), where tlt_{l} is the ll-th nearest hopping integral Kontani-ROP ; Springer . Hereafter, we set the unit of energy as |t1|=1|t_{1}|=1, which corresponds to 4000\sim 4000 [K] in cuprates, and fix the temperature T=0.05(200K)T=0.05\ (\sim 200{\rm K}). The FS at filling n=0.85n=0.85 is given in Fig. 2 (a). The spin susceptibility in the random-phase-approximation (RPA) is χs(q)=χ0(q)/(1Uχ0(q))\chi^{s}(q)=\chi^{0}(q)/(1-U\chi^{0}(q)), where χ0(q)\chi^{0}(q) is the irreducible susceptibility without UU and q(𝒒,ωl)q\equiv({{\bm{q}}},{\omega}_{l}). The spin Stoner factor is defined as αSmaxq{Uχ0(q)}=Uχ0(𝑸s,0){\alpha}_{S}\equiv\max_{q}\{U\chi^{0}(q)\}=U\chi^{0}({{\bm{Q}}}_{s},0). Figure 2 (b) shows the obtained χs(q)\chi^{s}(q) at αS=0.99{\alpha}_{S}=0.99 (U=3.27U=3.27). Here, χs(𝑸s,0)30\chi^{s}({{\bm{Q}}}_{s},0)\sim 30 [1/t11/{t_{1}}] 80\sim 80 [μB2/eV\mu_{\rm B}^{2}/{\rm eV}], which is still smaller than Imχs(𝑸s,E=31meV)200\chi^{s}({{\bm{Q}}}_{s},E=31{\rm meV})\sim 200 [μB2/eV\mu_{\rm B}^{2}/{\rm eV}] at T200T\sim 200K in 60K YBCO neutron . Thus, αS>0.99{\alpha}_{S}>0.99 in real compounds. Owing to the Mermin-Wagner theorem, the relation αS1{\alpha}_{S}\lesssim 1 is naturally satisfied for U3.3U\gg 3.3 without any fine tuning of UU by considering the spin-fluctuation-induced self-energy self-consistently Kontani-ROP .

Refer to caption
Figure 2: (a) The FS of the present model at n=0.85n=0.85. 𝑸s{{\bm{Q}}}_{s} is the major nesting vectors. 𝑸d=(δ,δ){{\bm{Q}}}_{\rm d}=(\delta,\delta) and 𝑸a=(δ,0){{\bm{Q}}}_{\rm a}=(\delta,0) (δδFS\delta\approx\delta_{\rm FS}) are minor nesting vectors. They correspond to the sLC order wavelength in the present theory. (b) χs(𝒒)\chi^{s}({{\bm{q}}}) given by the RPA. It shows the incommensurate peak at 𝒒=𝑸s{{\bm{q}}}={{\bm{Q}}}_{s}. (c) Obtained eigenvalue λ𝒒\lambda_{{\bm{q}}} for the BO at n=0.800.88n=0.80\sim 0.88. They have peaks at 𝒒=𝟎{{\bm{q}}}={\bm{0}} and 𝒒=𝑸a{{\bm{q}}}={{\bm{Q}}}_{\rm a}. (d) Relations αS=10.444p2{\alpha}_{S}=1-0.444p^{2} (full line) and αS=1.010.2p{\alpha}_{S}=1.01-0.2p (broken line). (e) Obtained dd-wave form factor f𝒒(𝒌)f_{{\bm{q}}}({{\bm{k}}}) for 𝒒=𝟎{{\bm{q}}}={\bm{0}} together with the FS.

II.2 Introduction of singlet and triplet DW equations

From now on, we investigate possible exotic density-wave (DW) states for both charge- and spin-channels with general wavevector (𝒒{{\bm{q}}}), which is generally expressed as Nersesyan

D𝒒σρ(𝒌)\displaystyle D_{{\bm{q}}}^{{\sigma}\rho}({{\bm{k}}}) =\displaystyle= c𝒌,σc𝒌+,ρc𝒌,σc𝒌+,ρ0\displaystyle\langle c_{{{\bm{k}}}_{-},{\sigma}}^{\dagger}c_{{{\bm{k}}}_{+},\rho}\rangle-\langle c_{{{\bm{k}}}_{-},{\sigma}}^{\dagger}c_{{{\bm{k}}}_{+},\rho}\rangle_{0} (2)
=\displaystyle= d𝒒c(𝒌)δσ,ρ+𝒅𝒒s(𝒌)𝝈σ,ρ\displaystyle d^{c}_{{\bm{q}}}({{\bm{k}}})\delta_{{\sigma},\rho}+\bm{d}^{s}_{{{\bm{q}}}}({{\bm{k}}})\cdot\bm{{\sigma}}_{{\sigma},\rho}

where 𝒌±𝒌±𝒒/2{{\bm{k}}}_{\pm}\equiv{{\bm{k}}}\pm{{\bm{q}}}/2, and d𝒒c(𝒌)d^{c}_{{\bm{q}}}({{\bm{k}}}) (𝒅𝒒s(𝒌))(\bm{d}^{s}_{{{\bm{q}}}}({{\bm{k}}})) is the charge (spin) channel order parameter. It induces the symmetry breaking in the self-energy:

ΔΣ𝒒σρ(𝒌)\displaystyle\Delta\Sigma_{{\bm{q}}}^{{\sigma}\rho}({{\bm{k}}}) =\displaystyle= f𝒒(𝒌)δσ,ρ+𝒈𝒒(𝒌)𝝈σ,ρ\displaystyle f_{{\bm{q}}}({{\bm{k}}})\delta_{{\sigma},\rho}+\bm{g}_{{{\bm{q}}}}({{\bm{k}}})\cdot\bm{{\sigma}}_{{\sigma},\rho} (3)

which we call the form factors in this paper. Below, we assume 𝒈𝒒(𝒌)=g𝒒(𝒌)𝒆z\bm{g}_{{{\bm{q}}}}({{\bm{k}}})=g_{{{\bm{q}}}}({{\bm{k}}}){\bm{e}}_{z} without losing generality. The DW is interpreted as the electron-hole pairing condensation Nersesyan .

Here, f𝒒(𝒌)f_{{\bm{q}}}({{\bm{k}}}) is given by the Fourier transformation of the spin-independent hopping modulation 𝒓i,𝒓jδti,jei(𝒓i𝒓j)𝒌ei(𝒓i+𝒓j)𝒒/2\sum_{{{\bm{r}}}_{i},{{\bm{r}}}_{j}}\delta t_{i,j}e^{i({{\bm{r}}}_{i}-{{\bm{r}}}_{j})\cdot{{\bm{k}}}}e^{i({{\bm{r}}}_{i}+{{\bm{r}}}_{j})\cdot{{\bm{q}}}/2}. When δti,j=±δtj,i\delta t_{i,j}=\pm\delta t_{j,i}, the relation f𝒒(𝒌)=±f𝒒(𝒌)f_{{\bm{q}}}({{\bm{k}}})=\pm f_{{\bm{q}}}(-{{\bm{k}}}) holds. Also, g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}}) is given by the spin-dependent modulation δti,j=δti,j\delta t_{i,j}^{\uparrow}=-\delta t_{i,j}^{\downarrow}. The even-parity f𝒒(𝒌)f_{{\bm{q}}}({{\bm{k}}}) and the odd-parity g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}}) respectively correspond to the BO state and the sLC state. Both states preserve the time-reversal symmetry.

To find possible DW in an unbiased way, we generalize the DW equation Kawaguchi-CDW for both spin/charge channels:

λ𝒒f𝒒(k)=TNpI𝒒c(k,p)G(p)G(p+)f𝒒(p),\displaystyle\!\!\!\!\lambda_{{{\bm{q}}}}f_{{\bm{q}}}(k)=-\frac{T}{N}\sum_{p}I_{{\bm{q}}}^{c}(k,p)G(p_{-})G(p_{+})f_{{\bm{q}}}(p), (4)
η𝒒g𝒒(k)=TNpI𝒒s(k,p)G(p)G(p+)g𝒒(p),\displaystyle\!\!\!\!\eta_{{{\bm{q}}}}g_{{\bm{q}}}(k)=-\frac{T}{N}\sum_{p}I_{{\bm{q}}}^{s}(k,p)G(p_{-})G(p_{+})g_{{\bm{q}}}(p), (5)

where λ𝒒\lambda_{{{\bm{q}}}} (η𝒒\eta_{{\bm{q}}}) is the eigenvalue that represents the charge (spin) channel DW instability, k(𝒌,ϵn)k\equiv({{\bm{k}}},{\epsilon}_{n}), p(𝒑,ϵm)p\equiv({{\bm{p}}},{\epsilon}_{m}) (ϵn{\epsilon}_{n}, ϵm{\epsilon}_{m} are fermion Matsubara frequencies). These DW equations are interpreted as the “spin/charge channel electron-hole pairing equations”.

The charge (spin) channel kernel function is I𝒒c(s)=I𝒒,+()I𝒒,I_{{\bm{q}}}^{c(s)}=I_{{\bm{q}}}^{\uparrow,\uparrow}+(-)I_{{\bm{q}}}^{\uparrow,\downarrow}; I𝒒σ,ρI_{{\bm{q}}}^{{\sigma},\rho} at 𝒒=0{{\bm{q}}}=0 is given by the Ward identity δΣσ(k)/δGρ(k)-\delta\Sigma_{\sigma}(k)/\delta G_{\rho}(k^{\prime}), which is composed of one single-magnon exchange term and two double-magnon exchange ones: The former and the latter are called the Maki-Thompson (MT) term and the AL terms; see Fig. 7 in Appendix A. The lowest order Hartree term Uδσ,ρ-U\delta_{{\sigma},\rho} in I𝒒σ,ρI_{{\bm{q}}}^{{\sigma},\rho} gives the RPA contribution. while the AL terms are significant for αS1{\alpha}_{S}\lesssim 1 in various strongy correlated systems Onari-SCVC ; Kawaguchi-CDW ; Tazai-CeCu2Si2 ; Tazai-CeB6 . The significance of the AL processes have been revealed by the functional-renormalization-group (fRG) study, in which higher-order vertex corrections are produced in an unbiased way Tsuchiizu-CDW ; Tsuchiizu-PRL ; Tazai-FRG ; Tazai-kappa . Note that the MT term is important for the superconducting gap equation, transport phenomena Kontani-ROP , and cLC order Tazai-cLC .

Figure 2 (c) shows the charge-channel eigenvalue λ𝒒\lambda_{{{\bm{q}}}} derived from the DW eq. (4) Kawaguchi-CDW ; Onari-B2g ; Onari-AFB . Hereafter, we put UU to satisfy the relation αS=10.444p2{\alpha}_{S}=1-0.444p^{2} with p1np\equiv 1-n, shown as full line in Fig. 2 (d). The obtained form factor f𝒒(𝒌)f_{{\bm{q}}}({{\bm{k}}}) at 𝒒=𝟎,𝑸d{{\bm{q}}}={\bm{0}},{{{\bm{Q}}}_{\rm d}}, shown in Fig. 2 (e), belongs to B1gB_{1g} symmetry BO, consistently with previous studies Kawaguchi-CDW ; Tsuchiizu-CDW . As we discuss in Appendix B, the large eigenvalue in Fig. 2 (c) (and Fig. 3 (a)) is strongly suppressed to O(1)O(1) by considering the small quasiparticle weight z=m/mO(101)z=m/m^{*}\sim O(10^{-1}) due to the self-energy in cuprates Kawaguchi-CDW ; Onari-B2g .

Refer to caption
Figure 3: (a) Obtained eigenvalue η𝒒\eta_{{\bm{q}}} for spin-channel DW at n=0.800.88n=0.80\sim 0.88. They have peaks at 𝒒=𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm d} and 𝑸a{{\bm{Q}}}_{\rm a}. (b) Form factor of the diagonal sLC order g𝑸d(𝒌)g_{{{\bm{Q}}}_{\rm d}}({{\bm{k}}}). We also show the shifted FSs given by μ=ϵ𝒌±𝑸d/2\mu={\epsilon}_{{{\bm{k}}}\pm{{\bm{Q}}}_{\rm d}/2}. (c) Img𝑸a(𝒓){\rm Im}g_{{{\bm{Q}}}_{\rm a}}({\bm{r}}), which is even (odd) with respect to x+yx+y (xyx-y).

III Derivation of sLC order based on triplet DW equation

III.1 Origin of sLC order

Next, we discuss the spin-fluctuation-driven sLC order, which is the main issue of this manuscript. Figure 3 (a) exhibits the spin-channel eigenvalue η𝒒\eta_{{{\bm{q}}}} derived from the DW eq. (5). Peaks of η𝒒\eta_{{{\bm{q}}}} are located at the nesting vectors 𝒒=𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm d} (diagonal) and 𝒒=𝑸a{{\bm{q}}}={{\bm{Q}}}_{\rm a} (axial). The obtained form factor g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}}) at 𝒒=𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm d} (diagonal sLC) is shown in Fig. 3 (b). The odd-parity solution g𝒒(𝒌)=g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}})=-g_{{\bm{q}}}(-{{\bm{k}}}) means the emergence of the sLC order. The reason for large η𝑸𝒅\eta_{{{\bm{Q}}}_{\bm{d}}} is that all hot spots contribute to the diagonal sLC as shown in Fig. 3 (b). Figure 3 (c) shows the form factor in real space Img𝑸a(𝒓){\rm Im}g_{{{\bm{Q}}}_{\rm a}}({\bm{r}}) with 𝒓=(x,y){\bm{r}}=(x,y). Here, δti,jσ=σg𝑸a(𝒓ij)cos(𝒓i+j𝑸a/2)\delta t_{i,j}^{\sigma}={\sigma}g_{{{\bm{Q}}}_{\rm a}}({\bm{r}}_{i-j})\cos({\bm{r}}_{i+j}\cdot{{\bm{Q}}}_{\rm a}/2).

To understand why sLC state is obtained, we simplify Eq. (5) by taking the Matsubara summation analytically by approximating that I𝒒sI_{{\bm{q}}}^{s} and g𝒒(k)g_{{\bm{q}}}(k) are static:

η𝒒g𝒒(𝒌)=1N𝒑I𝒒s(𝒌,𝒑)F𝒒(𝒑)g𝒒(𝒑),\displaystyle\eta_{{{\bm{q}}}}g_{{\bm{q}}}({{\bm{k}}})=\frac{1}{N}\sum_{{{\bm{p}}}}I_{{\bm{q}}}^{s}({{\bm{k}}},{{\bm{p}}})F_{{\bm{q}}}({{\bm{p}}})g_{{\bm{q}}}({{\bm{p}}}), (6)

where F𝒒(𝒑)TmG(p+)G(p)=n(ϵ𝒑+)n(ϵ𝒑)ϵ𝒑ϵ𝒑+\displaystyle F_{{\bm{q}}}({{\bm{p}}})\equiv-T\sum_{m}G(p_{+})G(p_{-})=\frac{n({\epsilon}_{{{\bm{p}}}_{+}})-n({\epsilon}_{{{\bm{p}}}_{-}})}{{\epsilon}_{{{\bm{p}}}_{-}}-{\epsilon}_{{{\bm{p}}}_{+}}} is a positive function, and n(ϵ)n({\epsilon}) is Fermi distribution function; see Appendix A. In general, the peak positions of η𝒒\eta_{{{\bm{q}}}} in Eq. (6) are located at 𝒒=𝟎{{\bm{q}}}={\bm{0}} and/or nesting vectors with small wavelength (𝒒=𝑸a,𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm a},{{\bm{Q}}}_{\rm d} in the present model). The reason is that I𝒒Tpχs(𝒑+)χs(𝒑)I_{{\bm{q}}}\sim T\sum_{p}\chi^{s}({{\bm{p}}}_{+})\chi^{s}({{\bm{p}}}_{-}) by AL terms is large for small |𝒒||{{\bm{q}}}|, and F𝒒(𝒑)F_{{\bm{q}}}({{\bm{p}}}) is large for wide area of 𝒑{{\bm{p}}} when 𝒒{{\bm{q}}} is a nesting vector.

To understand why odd-parity form factor is obtained, we show the spin-channel “electron-hole pairing interaction” I𝒒=𝟎s(𝒌,𝒌)I^{s}_{{{\bm{q}}}={\bm{0}}}({{\bm{k}}},{{\bm{k}}}^{\prime}) on the FS in Fig. 4 (a). The charge-channel one I𝒒=𝟎c(𝒌,𝒌)I^{c}_{{{\bm{q}}}={\bm{0}}}({{\bm{k}}},{{\bm{k}}}^{\prime}) is also shown in Fig.4 (b). Here, θ\theta represents the position of 𝒌{{\bm{k}}} shown in Figs. 4 (c) and (d). I𝒒s(𝒌,𝒌)I_{{\bm{q}}}^{s}({{\bm{k}}},{{\bm{k}}}^{\prime}) in Fig. 4 (a) gives large attractive interaction for 𝒌𝒌{{\bm{k}}}\approx{{\bm{k}}}^{\prime} and large repulsive one for 𝒌𝒌{{\bm{k}}}\approx-{{\bm{k}}}^{\prime}. In this case, we naturally obtain pp-wave form factor g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}}) shown in Fig. 3 (b), as we explain in Fig. 4 (c). Here, red (blue) arrows represent the attractive (repulsive) interaction.

The strong 𝒌,𝒌{{\bm{k}}},{{\bm{k}}}^{\prime}-dependence of I𝒒=𝟎s(𝒌,𝒌)I^{s}_{{{\bm{q}}}={\bm{0}}}({{\bm{k}}},{{\bm{k}}}^{\prime}) originates from the AL1 and AL2 terms in Fig. 4 (e), or Fig. 7 (a) in Appendix A. Owing to the spin-conservation law, AL terms in Is=I,I,I^{s}=I^{\uparrow,\uparrow}-I^{\uparrow,\downarrow} originates from the spin-flipping processes due to transverse spin fluctuations in Fig. 4 (e), in proportion to χ±s(𝑸s)χ±s(𝑸s)\chi^{s}_{\pm}({{\bm{Q}}}_{s})\chi^{s}_{\pm}({{\bm{Q}}}_{s}). (In IsI^{s}, the spin non-flipping AL processes in proportion to χzs(𝑸s)χzs(𝑸s)\chi^{s}_{z}({{\bm{Q}}}_{s})\chi^{s}_{z}({{\bm{Q}}}_{s}) are exactly canceled out.) Therefore, Is=[AL1][AL2]I^{s}=\mbox{[AL1]}-\mbox{[AL2]}. The AL1 term with the p-h (anti-parallel) pair Green functions causes large attractive interaction for 𝒌𝒌{{\bm{k}}}\approx{{\bm{k}}}^{\prime}, and the AL2 term with the p-p (parallel) ones does for 𝒌𝒌{{\bm{k}}}\approx-{{\bm{k}}}^{\prime}, as explained in Ref. Onari-B2g in detail. Thus, θ,θ\theta,\theta^{\prime}-dependence in Fig. 4 (a) and resultant odd-parity solution is understood naturally.

In contrast, the charge channel kernel I𝒒c(𝒌,𝒌)I_{{\bm{q}}}^{c}({{\bm{k}}},{{\bm{k}}}^{\prime}) gives an attractive interaction for both 𝒌±𝒌{{\bm{k}}}\approx\pm{{\bm{k}}}^{\prime} as shown in Fig. 4 (b), beucase Ic=3([AL1]+[AL2])/2I^{c}=3(\mbox{[AL1]}+\mbox{[AL2]})/2. Then, we obtain dd-wave form factor f𝒒(𝒌)f_{{\bm{q}}}({{\bm{k}}}) in Fig. 2 (e), as we explain in Fig. 4 (d) Kawaguchi-CDW ; Onari-B2g .

In the present transverse spin fluctuation mechanism, the 𝒈\bm{g}-vector will be parallel to zz-direction when χx(y)s(𝑸s)>χzs(𝑸s)\chi^{s}_{x(y)}({{\bm{Q}}}_{s})>\chi^{s}_{z}({{\bm{Q}}}_{s}) (XY-anisotropy) due to the spin-orbit interaction (SOI). When the XY-anisotropy of χμs(𝑸s)\chi^{s}_{\mu}({{\bm{Q}}}_{s}) is very large, IcI^{c} due to AL terms is multiplied by 2/32/3 whereas IsI^{s} is unchanged, so it is suitable condition for the sLC order.

Refer to caption
Figure 4: (a) Spin-channel and (b) charge-channel kernel functions on the FS, I𝒒=𝟎s,c(θ,θ)I_{{{\bm{q}}}={\bm{0}}}^{s,c}(\theta,\theta^{\prime}), where θ\theta represents the position of 𝒌{{\bm{k}}}. We see that I𝟎sI_{{\bm{0}}}^{s} and I𝟎cI_{{\bm{0}}}^{c} has large negative and positive values for θθ+π\theta^{\prime}\approx\theta+\pi, respectively, due to the p-p channel (= Cooper channel) in AL2. (c) Origin of pp-wave sLC order and (d) that of dd-wave BO. Red (blue) color arrows represent the attractive (repulsive) interaction. (e) Spin-flipping AL-processes in I𝒒s(𝒌,𝒑)I_{{\bm{q}}}^{s}({{\bm{k}}},{{\bm{p}}}) that give the sLC order. The wavy lines are transverse spin susceptibilities. (Spin non-flipping AL processes caused by longitudinal susceptibility are cancelled out in Is=I,I,I^{s}=I^{\uparrow,\uparrow}-I^{\uparrow,\downarrow}.) The AL1 term with a anti-parallel (p-h) pair gives red line on θθ\theta\approx\theta^{\prime} in (a). Also, the AL2 term with a parallel (p-p) pair gives blue line on θθ+π\theta\approx\theta^{\prime}+\pi in (a).

III.2 Filling dependences of sLC/BO instabilities

Refer to caption
Figure 5: (a) Obtained eigenvalues of sLC and BO as function of p=1np=1-n. (b) αS{\alpha}_{S}-dependences of the eigenvalues at p=0.16p=0.16. λSC\lambda_{\rm SC} is the eigenvalue of the superconducting gap equation. (c) Diagrammatic expression for the odd/even parity magnon-pair condensation, which is the physical origin of the sLC/BO.

The sLC and BO eigenvalues are summarized in Fig. 5 (a). The relation η𝑸d>λ𝑸a\eta_{{{\bm{Q}}}_{\rm d}}>\lambda_{{{\bm{Q}}}_{\rm a}} around the optimal doping (p0.15p\sim 0.15) means that the sLC transition temperature TsLCT_{\rm sLC} is higher than TCDWT_{\rm CDW}, as in Fig. 1 (a). The robustness of Fig. 5 (a) is verified in Appendix B. We also verify in Fig. 5 (b) that both η𝑸d\eta_{{{\bm{Q}}}_{\rm d}} and λ𝑸a\lambda_{{{\bm{Q}}}_{\rm a}} are larger than the dx2y2d_{x^{2}-y^{2}}-wave superconducting eigenvalue λSC\lambda_{\rm SC} for αS0.98{\alpha}_{S}\gtrsim 0.98. Here, λSC\lambda_{\rm SC} is derived from the gap equation

λSCΔ(k)=TpVSC(k,p)|G(p)|2Δ(p),\displaystyle\lambda_{\rm SC}\Delta(k)=T\sum_{p}V^{\rm SC}(k,p)|G(p)|^{2}\Delta(p), (7)

where VSC(k,p)=32U2χs(kp)+12U2χc(kp)UV^{\rm SC}(k,p)=-\frac{3}{2}U^{2}\chi^{s}(k-p)+\frac{1}{2}U^{2}\chi^{c}(k-p)-U is the MT-type kernel d-wave . Note that η𝒒,λ𝒒<λSC\eta_{{\bm{q}}},\lambda_{{\bm{q}}}<\lambda_{\rm SC} if AL terms are dropped Norman . The large eigenvalues in Fig. 5 are suppressed to O(1)O(1) by small zz; see Refs. Kawaguchi-CDW ; Onari-B2g ; Kontani-ROP and Appendix B. We stress that sLC is not suppressed by the ferro-BO that induces neither Fermi arc nor pseudogap, as explained in Appendix B.

Here, we analyzed the sLC/BO in terms of the electron-hole pairing. Another physical interpretation of the sLC/BO is the “condensation of odd/even parity magnon-pairs”, which is the origin of the nematic order in quantum spin systems Andreev ; Coleman ; Shannon . In fact, the two-magnon propagator shown in Fig. 5 (c) diverges when the eigenvalue of DW equation reaches unity, as we explain in Appendix C. That is, triplet (singlet) magnon-pair condensation occurs at T=TsLC(TCDW)T=T_{\rm sLC}(T_{\rm CDW}). Thus, the sLC/BO discussed here and the spin nematic order in quantum spin systems are essentially the same phenomenon.

III.3 Fermi arc and pseudogap under sLC order

Now, we discuss the band-folding and hybridization gap due to the diagonal sLC order. Figures 6 (a) and (b) show the Fermi arc structures in the cases of (a) single-𝒒{{\bm{q}}} and (b) double-𝒒{{\bm{q}}} orders. We set gmaxmax𝒌{g𝑸d(𝒌)}=0.1g^{\rm max}\equiv\max_{{\bm{k}}}\{g_{{{\bm{Q}}}_{\rm d}}({{\bm{k}}})\}=0.1. Here, the folded band structure under the sLC order with finite 𝒒sLC{{\bm{q}}}_{\rm sLC} is “unfolded” into the original Brillouin zone by following Ref. Ku to make a comparison with ARPES results. The Fermi arc due to the single-𝒒{{\bm{q}}} order in Fig. 6 (a) belongs to B2gB_{2g} symmetry. In contrast, the Fermi arc due to the double-𝒒{{\bm{q}}} order in Fig. 6 (b) preserves the C4C_{4} symmetry. The resultant pseudogap in the DOS is shown in Fig. 6 (c). The unfolded band structure in the single-𝑸d{{\bm{Q}}}_{\rm d} sLC order is displayed in Fig. 12 in Appendix B.

Refer to caption
Figure 6: (a) Fermi arc structure due to the single-𝒒{{\bm{q}}} order, (b) that due to the double-𝒒{{\bm{q}}} order, and (c) pseudogap in the DOS due to the diagonal sLC order (𝒒=𝑸d{{\bm{q}}}={\bm{Q}}_{\rm d}). In calculating (a)-(c), we introduced BCS-type cutoff energy ωc=0.5{\omega}_{c}=0.5 for the band-hybridization by g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}}).

Recent magnetic torque measurements revealed the B1gB_{1g} symmetry breaking at T=TT=T^{*} occurs in YBCO Y-Sato , while B2gB_{2g} one appears in Hg-based cuprates Hg-Murayama . To understand different symmetry breakings, we examine the t3t_{3} dependence of DW equation solution in Appendix B 2: As shown in Figs. 9 (b)-(e), the sLC wavevector 𝒒sLC{{\bm{q}}}_{\rm sLC} changes from 𝑸d{{\bm{Q}}}_{\rm d} to 𝑸a{{\bm{Q}}}_{\rm a} for larger t3/t1t_{3}/t_{1}. When 𝒒sLC=𝑸d{{\bm{q}}}_{\rm sLC}={{\bm{Q}}}_{\rm d}, the symmetry of the Fermi arc is B2gB_{2g} in Fig. 6 (a). On the other hand, the Fermi arc has B2gB_{2g} symmetry in the axial sLC order in Fig. 1 (d), as we show in Fig. 10 (b).

Since the van Vleck susceptibility becomes anisotropic when C4C_{4} symmetry of the FS is broken Kawaguchi-B2g , the reported compound-dependent symmetry breaking Y-Sato ; Hg-Murayama would be explained by the sLC order scenario. This is an important future issue.

III.4 Spin-current pattern under sLC order

Next, we investigate the spin current in real space, which is driven by a fictitious Peierls phase due to the “spin-dependent self-energy” δti,jσ=σgi,j\delta t_{i,j}^{\sigma}={\sigma}g_{i,j}. As shown in Fig. 3 (c), δti,j\delta t_{i,j} is purely imaginary and odd with respect to iji\leftrightarrow j. The conservation law n˙iσ=jji,jσ\dot{n}_{i}^{\sigma}=\sum_{j}j_{i,j}^{\sigma} directly leads to the definition the spin current operator from site jj to site ii as ji,jσ=iσσ(hi,jσciσcjσ(ij))j_{i,j}^{\sigma}=-i\sum_{{\sigma}}{\sigma}(h_{i,j}^{\sigma}c_{i{\sigma}}^{\dagger}c_{j{\sigma}}-(i\leftrightarrow j)), where hi,jσ=ti,j+δti,jσh_{i,j}^{\sigma}=t_{i,j}+\delta t_{i,j}^{\sigma}. Then, the spontaneous spin current from jj to ii is Ji,js=ji,jsh^σJ_{i,j}^{s}=\langle j_{i,j}^{s}\rangle_{\hat{h}^{\sigma}}.

Here, we calculate the spin current for the commensurate sLC order at 𝒒sLC=(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,\pi/2), which is achieved by putting n=1.0n=1.0. Then, the unit cell are composed four sites A-D. In Fig. 14 in Appendix D, we show the obtained spin current Ji,jsJ_{i,j}^{s} from the center site (j=Aj={\rm A}-B{\rm B}) to different site in Fig. 1 (c). The derived spin current pattern between the nearest and second-nearest sites is depicted in Fig. 1 (c).

The Fermi arc structure and the DOS in Fig. 6 are independent of the phase shift g𝒒eiψg𝒒g_{{\bm{q}}}\rightarrow e^{i\psi}g_{{\bm{q}}}. In contrast, the real space current pattern depends on the phase shift. We discuss other possible sLC patterns in Appendix D. The charge modulation due to the sLC is just |Δni|5×104|\Delta n_{i}|\sim 5\times 10^{-4} for gmax=0.1g^{\rm max}=0.1 since |Δni|(gmax)2|\Delta n_{i}|\propto(g^{\rm max})^{2}. Thus, experimental detection of translational symmetry breaking by sLC order may be difficult. However, the cLC is induced by applying uniform magnetic field parallel to 𝒈i,j{\bm{g}}_{i,j}. In the present sLC state, under 10 T magnetic field, the induced cLC gives ΔH±0.1\Delta H\sim\pm 0.1 Oe when m/m10m^{*}/m\sim 10, which may be measurable by NMR or μ\muSR study.

III.5 sLC fluctuation mediated superconductivity

Finally, we discuss that the sLC fluctuations can contribute to the dx2y2d_{x^{2}-y^{2}}-wave pairing mechanism. The sLC fluctuations connects the close hot spots at 𝑷\bm{P} and 𝑷=𝑷+𝒒\bm{P^{\prime}}=\bm{P}+{{\bm{q}}} in Fig. 2 (a). At both points, the dx2y2d_{x^{2}-y^{2}}-wave gap function Δ𝒌\Delta_{{\bm{k}}} has the same sign. The pairing interaction mediated by the spin-channel sLC fluctuations between singlet pairs (𝑷,𝑷)(\bm{P},-\bm{P}) and (𝑷,𝑷)(\bm{P^{\prime}},-\bm{P^{\prime}}) is

VsLC(𝑷,𝑷)g𝒒(𝒌)g𝒒(𝒌)(χsLC(𝒒)),\displaystyle V^{\rm sLC}(\bm{P},\bm{P^{\prime}})\propto g_{{{\bm{q}}}}({{\bm{k}}})g_{{{\bm{q}}}}(-{{\bm{k}}})\cdot(-\chi_{\rm sLC}({{\bm{q}}})), (8)

where 𝒌=(𝑷+𝑷)/2{{\bm{k}}}=(\bm{P}+\bm{P^{\prime}})/2 and 𝒒=𝑷𝑷𝑸a{{\bm{q}}}=\bm{P}-\bm{P^{\prime}}\sim{{\bm{Q}}}_{\rm a}. χsLC(𝒒)(>0)\chi_{\rm sLC}({{\bm{q}}})\ (>0) is the sLC susceptibility and g𝒒(𝒌)=g𝒒(𝒌)g_{{{\bm{q}}}}({{\bm{k}}})=-g_{{{\bm{q}}}}(-{{\bm{k}}}) is its form factor. Thus, Eq. (8) give positive (=attractive) pairing interaction between close hot spots. The derivation of Eq. (8) is given in Appendix A and in SM F of Ref. Onari-AFB . This mechanism may be important for slightly over-doped cuprates with TsLCTcT_{\rm sLC}\lesssim T_{\rm c}.

IV Summary

In summary we proposed a novel and long-sought formation mechanism for the nanoscale spin-current order, which violates the parity and translational symmetry while time-reversal symmetry is preserved. It was revealed that the formation of triplet odd-parity electron-hole pairs that is mediated by spin fluctuations, and therefore the spontaneous sLC is established at T=TsLCT=T_{\rm sLC}. In the present spin-fluctuation mechanism, the condensation of odd-parity magnon-pairs occurs simultaneously. The band-folding by the sLC orders results in the formations of the Fermi arc structure and pseudogap at TTT\sim T^{*}. In the sLC state, a staggered moment is expected to appear under the uniform magnetic field. The sLC order will be a key ingredient in understanding pseudogap phase and electronic nematicity not only in cuprates, but also in iridates Sr2IrO4 ; Sr2IrO4-ARPES and heavy-fermion compound. It is an important future issue to incorporate the self-energy effect into the present theory.

Acknowledgements.
The authors are grateful to Y. Matsuda, K. Yamada, T. Moriya, and A. Kobayashi for the fruitful comments and discussions. This work was supported by the “Quantum Liquid Crystals” No. JP19H05825 KAKENHI on Innovative Areas from JSPS of Japan, and JSPS KAKENHI (JP18H01175, JP20K22328, JP20K03858, JP17K05543).

Appendix A Derivation of singlet and triplet DW equations

Here, we discuss the linearized density-wave (DW) equation driven by spin fluctuations. For this purpose, we introduce the irreducible four-point vertex function I𝒒σ,ρ(k,k)I_{{\bm{q}}}^{{\sigma},\rho}(k,k^{\prime}). It is given by the Ward identity at 𝒒=0{{\bm{q}}}=0, that is, I𝒒σ,ρ(k,k)δΣσ(k)/δGρ(k)I_{{\bm{q}}}^{{\sigma},\rho}(k,k^{\prime})\equiv-\delta\Sigma_{\sigma}(k)/\delta G_{\rho}(k^{\prime}). Here, we use the one-loop self-energy given as

Σσ(k)=TqU2χLσ(q)Gσ(k+q)\displaystyle\!\!\!\!\!\!\!\!\Sigma_{\sigma}(k)=T\sum_{q}U^{2}\chi_{\rm L}^{\sigma}(q)G_{\sigma}(k+q)
+TpU2(χTσ(q)χT(0)σ(q))Gσ(k+q),\displaystyle+T\sum_{p}U^{2}(\chi_{\rm T}^{\sigma}(q)-\chi_{\rm T}^{(0){\sigma}}(q))G_{-{\sigma}}(k+q), (9)

where χLσ(q)\chi_{\rm L}^{\sigma}(q) and χTσ(q)\chi_{\rm T}^{\sigma}(q) are longitudinal and transverse susceptibilities. They are given as

χLσ(q)\displaystyle\chi_{\rm L}^{{\sigma}}(q) =\displaystyle= χL(0)σ(q)(1U2χL(0)σ(q)χL(0)σ(q))1,\displaystyle\chi_{\rm L}^{(0){\sigma}}(q)(1-U^{2}\chi_{\rm L}^{(0){\sigma}}(q)\chi_{\rm L}^{(0)-{\sigma}}(q))^{-1}, (10)
χTσ(q)\displaystyle\chi_{\rm T}^{\sigma}(q) =\displaystyle= χT(0)σ(q)(1UχT(0)σ(q))1,\displaystyle\chi_{\rm T}^{(0){\sigma}}(q)(1-U\chi_{\rm T}^{(0){\sigma}}(q))^{-1}, (11)

where χL(0)σ(q)=TpGσ(p)Gσ(p+q)\chi_{\rm L}^{(0){\sigma}}(q)=-T\sum_{p}G_{{\sigma}}(p)G_{{\sigma}}(p+q) and χT(0)σ(q)=TpGσ(p)Gσ(p+q)\chi_{\rm T}^{(0){\sigma}}(q)=-T\sum_{p}G_{{\sigma}}(p)G_{-{\sigma}}(p+q) are longitudinal and transverse irreducible susceptibilities. Then, the irreducible vertex function I𝒒σ,ρ(k,k)I_{{\bm{q}}}^{{\sigma},\rho}(k,k^{\prime}) given by the Ward identity is composed of one MT term and two AL terms in Fig. 7 (a). Note that I𝒒σ,ρI_{{\bm{q}}}^{{\sigma},\rho} contains the lowest order Hartree term Uδσ,ρ-U\delta_{{\sigma},\rho}.

First, we derive the charge-channel (singlet) DW equation in the absence of the magnetic field, where the form factor is independent of spin: f𝒒(k)=f𝒒(k)=f𝒒(k)f_{{\bm{q}}}^{\uparrow}(k)=f_{{\bm{q}}}^{\downarrow}(k)=f_{{\bm{q}}}(k). The singlet DW equation was introduced in the study of Fe-based superconductors Onari-B2g and cuprate superconductors Kawaguchi-CDW . It is given as

λ𝒒f𝒒(k)\displaystyle\lambda_{{{\bm{q}}}}f_{{\bm{q}}}(k) =\displaystyle= TNkI𝒒c(k,k)G(k)G(k+)f𝒒(k),\displaystyle-\frac{T}{N}\sum_{k^{\prime}}I_{{\bm{q}}}^{c}(k,k^{\prime})G(k_{-}^{\prime})G(k_{+}^{\prime})f_{{\bm{q}}}(k^{\prime}), (12)

which is shown in Fig. 7 (b), and 𝒌±𝒌±𝒒/2{{\bm{k}}}_{\pm}\equiv{{\bm{k}}}\pm{{\bm{q}}}/2. Here, I𝒒c(k,k)=I𝒒σ,σ(k,k)+I𝒒σ,σ(k,k)I_{{\bm{q}}}^{c}(k,k^{\prime})=I_{{\bm{q}}}^{{\sigma},{\sigma}}(k,k^{\prime})+I_{{\bm{q}}}^{{\sigma},-{\sigma}}(k,k^{\prime}). It is given as

I𝒒c(k,k)=32Vs(kk)12Vc(kk)\displaystyle\!\!\!\!\!\!\!\!\!\!I_{\bm{q}}^{c}(k,k^{\prime})=-\frac{3}{2}V^{s}(k-k^{\prime})-\frac{1}{2}V^{c}(k-k^{\prime})
+TNp[32Vs(p+)Vs(p)+12Vc(p+)Vc(p)]\displaystyle+\frac{T}{N}\sum_{p}[\frac{3}{2}V^{s}(p_{+})V^{s}(p_{-})+\frac{1}{2}V^{c}(p_{+})V^{c}(p_{-})]
×G(kp)G(kp)\displaystyle\ \ \ \ \ \times G(k-p)G(k^{\prime}-p)
+TNp[32Vs(p+)Vs(p)+12Vc(p+)Vc(p)]\displaystyle+\frac{T}{N}\sum_{p}[\frac{3}{2}V^{s}(p_{+})V^{s}(p_{-})+\frac{1}{2}V^{c}(p_{+})V^{c}(p_{-})]
×G(kp)G(k+p),\displaystyle\ \ \ \ \ \times G(k-p)G(k^{\prime}+p), (13)

where p=(𝒑,ωl)p=({{\bm{p}}},{\omega}_{l}), V^s(q)=U+U2χ^s(q)\hat{V}^{s}(q)=U+U^{2}\hat{\chi}^{s}(q), and V^c(q)=U+U2χ^c(q)\hat{V}^{c}(q)=-U+U^{2}\hat{\chi}^{c}(q). The first, the second, the third terms in Eq. (13) corresponds to the MT, AL1 and AL2 terms in Fig. 7 (a).

In cuprates, Eq. (12) gives even-parity solution with wavevector 𝒒=𝟎{{\bm{q}}}={\bm{0}} and 𝒒(π/2,0){{\bm{q}}}\approx(\pi/2,0). This singlet and even-parity electron-hole condensation is interpreted as the BO.

Refer to caption
Figure 7: (a) Irreducible four-point vertex I𝒒σ,ρ(k,k)I_{{\bm{q}}}^{{\sigma},\rho}(k,k^{\prime}) composed of one MT term and two AL terms. (b) Linearized singlet DW equation with the kernel IcIσ,σ+Iσ,σI^{c}\equiv I^{{\sigma},{\sigma}}+I^{{\sigma},-{\sigma}}. (c) Linearized triplet DW equation with the kernel IsIσ,σIσ,σI^{s}\equiv I^{{\sigma},{\sigma}}-I^{{\sigma},-{\sigma}}. (d) A three-magnon exchange term, which is less important. (e) Full four-point vertex function Γ𝒒s(c)(k,k)\Gamma^{s(c)}_{{\bm{q}}}(k,k^{\prime}) given by solving the DW equation. The sLC order (BO) emerges when Γ𝒒s(c)(k,k)\Gamma^{s(c)}_{{\bm{q}}}(k,k^{\prime}) diverges.

Next, we derive the spin-channel (triplet) DW equation in the absence of the magnetic field, the spin-dependent form factor is g𝒒(k)g𝒒(k)=g𝒒(k)g_{{\bm{q}}}(k)\equiv g_{{\bm{q}}}^{\uparrow}(k)=-g_{{\bm{q}}}^{\downarrow}(k). It is given as

η𝒒g𝒒(k)\displaystyle\eta_{{{\bm{q}}}}g_{{\bm{q}}}(k) =\displaystyle= TNkI𝒒s(k,k)G(k)G(k+)g𝒒(k),\displaystyle-\frac{T}{N}\sum_{k^{\prime}}I_{{\bm{q}}}^{s}(k,k^{\prime})G(k_{-}^{\prime})G(k_{+}^{\prime})g_{{\bm{q}}}(k^{\prime}), (14)

which is shown in Fig. 7 (c). Here, I𝒒s(k,k)=I𝒒σ,σ(k,k)I𝒒σ,σ(k,k)I_{{\bm{q}}}^{s}(k,k^{\prime})=I_{{\bm{q}}}^{{\sigma},{\sigma}}(k,k^{\prime})-I_{{\bm{q}}}^{{\sigma},-{\sigma}}(k,k^{\prime}). It is given as

I𝒒s(k,k)=12Vs(kk)12Vc(kk)\displaystyle\!\!\!\!\!\!\!\!\!\!I_{\bm{q}}^{s}(k,k^{\prime})=\frac{1}{2}V^{s}(k-k^{\prime})-\frac{1}{2}V^{c}(k-k^{\prime})
+TNp[Vs(p+)Vs(p)+12Vs(p+)Vc(p)\displaystyle+\frac{T}{N}\sum_{p}[V^{s}(p_{+})V^{s}(p_{-})+\frac{1}{2}V^{s}(p_{+})V^{c}(p_{-})
+12Vc(p+)Vs(p)]G(kp)G(kp)\displaystyle\ \ \ \ \ +\frac{1}{2}V^{c}(p_{+})V^{s}(p_{-})]G(k-p)G(k^{\prime}-p)
+TNp[Vs(p+)Vs(p)+12Vs(p+)Vc(p)\displaystyle+\frac{T}{N}\sum_{p}[-V^{s}(p_{+})V^{s}(p_{-})+\frac{1}{2}V^{s}(p_{+})V^{c}(p_{-})
+12Vc(p+)Vs(p)]G(kp)G(k+p),\displaystyle\ \ \ \ \ +\frac{1}{2}V^{c}(p_{+})V^{s}(p_{-})]G(k-p)G(k^{\prime}+p), (15)

where the first, the second, the third terms in Eq. (13) corresponds to the MT, AL1 and AL2 terms in Fig. 7 (a). The AL terms with Vs(p+)Vs(p)V^{s}(p_{+})V^{s}(p_{-}) are shown in Fig. 4 (e). In cuprates, Eq. (14) gives the odd-parity solution at wavevector 𝒒=(π/2,π/2){{\bm{q}}}=(\pi/2,\pi/2) and (π/2,0)(\pi/2,0). This triplet and odd-parity electron-hole pairing is interpreted as the spin-loop-current (sLC).

In both Eqs. (13) and (15), the AL terms are proportional to ϕ𝒒(2)T𝒑1,𝒑2Vs(𝒑1)Vs(𝒑2)δ𝒑1+𝒑2,𝒒\phi^{(2)}_{{\bm{q}}}\equiv T\sum_{{{\bm{p}}}_{1},{{\bm{p}}}_{2}}V^{s}({{\bm{p}}}_{1})V^{s}({{\bm{p}}}_{2})\cdot\delta_{{{\bm{p}}}_{1}+{{\bm{p}}}_{2},{{\bm{q}}}}. The AL terms are significant when the spin fluctuations are large, since both Vs(𝒑1)V^{s}({{\bm{p}}}_{1}) and Vs(𝒒𝒑1)V^{s}({{\bm{q}}}-{{\bm{p}}}_{1}) take large value simultaneously when 𝒑1𝑸s{{\bm{p}}}_{1}\approx{{\bm{Q}}}_{s} in the case of 𝒒𝟎{{\bm{q}}}\approx{\bm{0}}. If we put Vs(𝒑)ξ2/(1+ξ2(𝒑𝑸s)2)V^{s}({{\bm{p}}})\propto\xi^{2}/(1+\xi^{2}({{\bm{p}}}-{{\bm{Q}}}_{s})^{2}) at zero Matsubara frequency, where ξ(1)\xi\ (\gg 1) is the magnetic correlation length, ϕ𝒒=𝟎(2)Tξ2\phi^{(2)}_{{{\bm{q}}}={\bm{0}}}\propto T\xi^{2} in two-dimensional systems. Therefore, double-magnon exchange (AL) terms induce not only BO, but also the sLC order when ξ1\xi\gg 1. A three-magnon exchange term shown in Fig. 7 (d) is proportional to ϕ𝒒(3)T2𝒑1,𝒑2,𝒑3Vs(𝒑1)Vs(𝒑2)Vs(𝒑3)δ𝒑1+𝒑2+𝒑3,𝒒\phi^{(3)}_{{\bm{q}}}\equiv T^{2}\sum_{{{\bm{p}}}_{1},{{\bm{p}}}_{2},{{\bm{p}}}_{3}}V^{s}({{\bm{p}}}_{1})V^{s}({{\bm{p}}}_{2})V^{s}({{\bm{p}}}_{3})\cdot\delta_{{{\bm{p}}}_{1}+{{\bm{p}}}_{2}+{{\bm{p}}}_{3},{{\bm{q}}}}. Then, ϕ𝒒=𝟎(3)T2ξ2\phi^{(3)}_{{{\bm{q}}}={\bm{0}}}\propto T^{2}\xi^{2} in two-dimensional systems for 𝒒𝑸s{{\bm{q}}}\sim{{\bm{Q}}}_{s}, which is smaller than ϕ𝒒=𝟎(2)\phi^{(2)}_{{{\bm{q}}}={\bm{0}}} at low temperatures TEFT\ll E_{\rm F}. Thus, the AL process would be the most significant, which is also indicated by functional-renormalization-group studies Tsuchiizu-CDW .

The electron-hole pairing order is generally expressed in real space as follows Nersesyan :

Di,jσ,ρ\displaystyle D_{i,j}^{{\sigma},\rho} \displaystyle\equiv ciσcjρciσcjρ0\displaystyle\langle c_{i{\sigma}}^{\dagger}c_{j\rho}\rangle-\langle c_{i{\sigma}}^{\dagger}c_{j\rho}\rangle_{0} (16)
=\displaystyle= di,jcδσ,ρ+𝒅i,js𝝈σ,ρ,\displaystyle d^{c}_{i,j}\delta_{{\sigma},\rho}+\bm{d}^{s}_{i,j}\cdot\bm{{\sigma}}_{{\sigma},\rho},

where Di,jσ,ρ={Dj,iρ,σ}D_{i,j}^{{\sigma},\rho}=\{D_{j,i}^{\rho,{\sigma}}\}^{*}, and di,jcd^{c}_{i,j} (𝒅i,js{\bm{d}}^{s}_{i,j}) is spin singlet (triplet) pairing. It induces the symmetry breaking in the self-energy:

ΔΣi,jσρ\displaystyle\Delta\Sigma_{i,j}^{{\sigma}\rho} =\displaystyle= fi,jδσ,ρ+𝒈i,j𝝈σ,ρ\displaystyle f_{i,j}\delta_{{\sigma},\rho}+\bm{g}_{i,j}\cdot\bm{{\sigma}}_{{\sigma},\rho} (17)

which we call the form factors in this paper. The BO is given by real even-parity function fi,j=fj,if_{i,j}=f_{j,i}, and the sLC is given by pure imaginary odd-parity vector 𝒈i,j=𝒈j,i{\bm{g}}_{i,j}=-{\bm{g}}_{j,i}. Both orders preserve the time-reversal symmetry. Note that f𝒒(k)f_{{\bm{q}}}(k) and g𝒒(k)g_{{\bm{q}}}(k) in Eqs. (12) and (14) correspond to fi,jf_{i,j} and gi,jzg_{i,j}^{z}, respectively.

Finally, we discuss the effective interaction driven by the BO/sLC fluctuations. By solving the DW equation (12), we obtain the full four-point vertex function Γ𝒒c(k,k)\Gamma^{c}_{{\bm{q}}}(k,k^{\prime}) that is composed of I𝒒cI_{{\bm{q}}}^{c} and G(k+)G(k)G(k_{+})G(k_{-}) shown in Fig. 7 (e), which increases in proportion to (1η𝒒)1(1-\eta_{{\bm{q}}})^{-1}. Thus, we obtain the relation Γ𝒒c(k,k)f𝒒(k){f𝒒(k)}I¯𝒒c(1λ𝒒)1\Gamma^{c}_{{\bm{q}}}(k,k^{\prime})\approx f_{{\bm{q}}}(k)\{f_{{\bm{q}}}(k^{\prime})\}^{*}{\bar{I}}_{{\bm{q}}}^{c}(1-\lambda_{{\bm{q}}})^{-1}, which is well satisfied when λ𝒒\lambda_{{\bm{q}}} is close to unity. Here, I¯𝒒c(s)T2k,k{f𝒒(k)}I𝒒c(s)(k,k)f𝒒(k)/Tk|f𝒒(k)|2{\bar{I}}_{{\bm{q}}}^{c(s)}\equiv T^{2}\sum_{k,k^{\prime}}\{f_{{\bm{q}}}(k)\}^{*}I_{{\bm{q}}}^{c(s)}(k,k^{\prime})f_{{\bm{q}}}(k^{\prime})/T\sum_{k}|f_{{\bm{q}}}(k)|^{2}. In the same way, we obtain the relation Γ𝒒s(k,k)g𝒒(k){g𝒒(k)}I¯𝒒s(1η𝒒)1\Gamma^{s}_{{\bm{q}}}(k,k^{\prime})\approx g_{{\bm{q}}}(k)\{g_{{\bm{q}}}(k^{\prime})\}^{*}{\bar{I}}_{{\bm{q}}}^{s}(1-\eta_{{\bm{q}}})^{-1}. Thus, it is apparent that the sLC order gg (BO ff) emerges when Γ𝒒s(c)(k,k)\Gamma^{s(c)}_{{\bm{q}}}(k,k^{\prime}) diverges.

The pairing interaction due to the sLC fluctuations is given by the full four-point vertex. It is approximately expressed as VSC(𝒌+,𝒌)=Γ𝒒s(𝒌,𝒌)g𝒒(𝒌){g𝒒(𝒌)}(1η𝒒)1V^{\rm SC}({{\bm{k}}}_{+},{{\bm{k}}}_{-})=-\Gamma_{{\bm{q}}}^{s}({{\bm{k}}},-{{\bm{k}}})\propto-g_{{{\bm{q}}}}({{\bm{k}}})\{g_{{{\bm{q}}}}(-{{\bm{k}}})\}^{*}(1-\eta_{{\bm{q}}})^{-1}. Since gg is odd-function, the sLC fluctuations cause attractive interaction: VSC(𝒌+,𝒌)|g𝒒(𝒌)|2(1η𝒒)1V^{\rm SC}({{\bm{k}}}_{+},{{\bm{k}}}_{-})\propto-|g_{{{\bm{q}}}}({{\bm{k}}})|^{2}(1-\eta_{{\bm{q}}})^{-1}.

Appendix B Additional numerical results of DW equations

B.1 Enhancement of the sLC instability under the finite ferro BO

In Figs. 5 (a) and (b) in the main text, the sLC eigenvalue η𝒒=𝑸d\eta_{{{\bm{q}}}={{\bm{Q}}}_{\rm d}} is comparable to the BO eigenvalue λ𝒒=𝟎\lambda_{{{\bm{q}}}=\bm{0}} for a wide doping range. This result means that the sLC order at 𝒒sLC=𝑸d{{\bm{q}}}_{\rm sLC}={{\bm{Q}}}_{\rm d} and the ferro-BO occur at almost the same temperature T\sim T^{*}. Here, we discuss the possibility of coexistence of sLC order and ferro-BO.

Refer to caption
Figure 8: Obtained η𝒒\eta_{{\bm{q}}} for n=0.85n=0.85 (αS=0.99{\alpha}_{S}=0.99) under the ferro-BO with f𝒒=𝟎max=0, 0.01, 0.03f_{{{\bm{q}}}=\bm{0}}^{\rm max}=0,\ 0.01,\ 0.03. Thus, the ferro-BO does not prohibit the emergence of the sLC order.

Since the ferro-BO does not induce the band-folding and pseudogap, the sLC order will emerge even if the ferro-BO transiting temperature is higher. To verify this expectation, we calculated the triplet DW equation (14) under the ferro-BO with f𝒒=𝟎max=0, 0.01, 0.03f_{{{\bm{q}}}=\bm{0}}^{\rm max}=0,\ 0.01,\ 0.03. Figure 8 shows the eigenvalue of sLC as function of 𝒒{{\bm{q}}} for n=0.85n=0.85 and U=3.27U=3.27 (αS=0.99{\alpha}_{S}=0.99) under the ferro-BO form factor obtained by the spin-singlet DW equation (12). It is verified that the ferro-BO does not prohibit the emergence of the sLC order. The eigenvalue η𝒒\eta_{{\bm{q}}} slightly increases with f𝒒=𝟎maxf_{{{\bm{q}}}=\bm{0}}^{\rm max}, since the spin Stoner factor αS{\alpha}_{S} is enlarged by the ferro-BO Kawaguchi-CDW ; Tsuchiizu-CDW .

B.2 Change in the phase diagram by t3t_{3}

In the main text, we show that the sLC eigenvalue at 𝒒=𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm d} develops as large as the ferro-BO eigenvalue near the optimally-doping case (p0.15p\sim 0.15), based on the Hubbard model with the hopping integrals (t1,t2,t3)=(1.0,1/6,1/5)(t_{1},t_{2},t_{3})=(-1.0,1/6,-1/5). The obtained Fermi surface (FS) has the flat part near the Brillouin zone boundary, which captures the characteristic of YBCO compounds.

Here, we examine a key model parameter for the phase diagram, and reveal that the sLC instability is sensitively controlled by t3t_{3}. Figure 9 (a) shows the FSs for t3=0.200.25t_{3}=-0.20\sim 0.25 in the case of t1=1.0t_{1}=-1.0 and t2=1/6t_{2}=1/6. The shape of the flat part of the FS near the BZ, which is significant for the density wave (DW) instabilities at 𝒒=𝑸a,𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm a},{{\bm{Q}}}_{\rm d}, is sensitively modified by t3t_{3}.

Refer to caption
Figure 9: (a) FSs for t3=0.200.25t_{3}=-0.20\sim-0.25 at n=0.85n=0.85, in the case of t1=1.0t_{1}=-1.0 and t2=1/6t_{2}=1/6. The set t3=0.20t_{3}=0.20 in the main text. (b)-(e) Obtained eigenvalue η𝒒\eta_{{\bm{q}}} and λ𝒒\lambda_{{\bm{q}}} in the cases of t3=0.23t_{3}=-0.23 ((b) and (c)) and t3=0.25t_{3}=-0.25 ((d) and (e)).

Figures 9 (b)-(e) shows the obtained spin-channel eigenvalue η𝒒\eta_{{\bm{q}}} and the charge-channel one λ𝒒\lambda_{{\bm{q}}} in the cases of t3=0.23t_{3}=-0.23 ((b) and (c)) and t3=0.25t_{3}=-0.25 ((d) and (e)). (We set the condition αS=10.444p2{\alpha}_{S}=1-0.444p^{2} by following the main text.) With increasing |t3||t_{3}|, the peak of sLC instability is found to shift to the under-doped region. Interestingly, the SLC eigenvalue at 𝒒=𝑸a{{\bm{q}}}={{\bm{Q}}}_{\rm a} becomes larger than that at 𝒒=𝑸d{{\bm{q}}}={{\bm{Q}}}_{\rm d} for t3=0.25t_{3}=-0.25. Its spin current pattern in real space is shown in Fig. 10 (a).

We note that recent experiments indicate that the phase diagram of cuprate superconductors is very diverse and rich. For example, the in-plain anisotropy of the ferro-magnetic susceptibility at T=TT=T^{*} is B1gB_{1g} in YBCO Y-Sato , whereas B2gB_{2g} in Hg-compound Hg-Murayama . Also, the antiferro-magnetic susceptibility in slightly under-doped YBCO exhibits a clear symmetry breaking at a temperate between TT^{*} and TCDWT_{\rm CDW} Keimer . The present sensitive t3t_{3} dependent sLC may give important hint to understand diverse symmetry breaking in cuprates.

Refer to caption
Figure 10: (a) Schematic spin current pattern due to the sLC order at 𝒒=𝑸a{{\bm{q}}}={{\bm{Q}}}_{\rm a}. (b) Fermi arc structure due to the single-𝒒{{\bm{q}}} order, (c) that due to the double-𝒒{{\bm{q}}} order, and (d) pseudogap in the DOS due to the axial sLC order (𝒒=𝑸a{{\bm{q}}}={\bm{Q}}_{\rm a}).

Here, we discuss the band-folding and hybridization gap in the axial sLC phase. Figures 10 (b) and (c) show the Fermi arc structures induced by the axial sLC order in the cases of (a) single-𝒒{{\bm{q}}} and (b) double-𝒒{{\bm{q}}} orders. We set gmaxmax𝒌{g𝑸d(𝒌)}=0.1g^{\rm max}\equiv\max_{{\bm{k}}}\{g_{{{\bm{Q}}}_{\rm d}}({{\bm{k}}})\}=0.1. Here, the folded band structure under the sLC order with finite 𝒒sLC{{\bm{q}}}_{\rm sLC} is “unfolded” to make a comparison with ARPES results. The Fermi arc due to the single-𝒒{{\bm{q}}} order in Fig. 10 (b) belongs to the B1gB_{1g} symmetry. The resultant pseudogap in the DOS is shown in Fig. 10 (d).

B.3 Reduction of eigenvalues by z<1z<1

In the present work, we study the mechanism of exotic DW orders due to the interference between paramagnons based on the linearized DW equation. The obtained form factor represents the characteristics and the symmetry of the DW, and the eigenvalue expresses the strength of the DW instability. In the present numerical study, we drop the self-energy in the DW equation. Then, the obtained eigenvalues shown in Fig. 5 and Fig. 9 are much larger than unity. In addition, the eigenvalue of the superconducting gap equation, λSC{\lambda}_{\rm SC}, shown in Fig. 5 (b) is also very large.

The self-energy gives the quasiparticle weight as z(1ReΣ(ϵ)/ϵ|ϵ=μ)1(<1)z\equiv(1-{\rm Re}{\partial}\Sigma({\epsilon})/{\partial}{\epsilon}|_{{\epsilon}=\mu})^{-1}\ (<1), and z1(>1)z^{-1}\ (>1) is the mass-enhancement. The effect of the self-energy in the DW equation has been studied in Ref Onari-AFB for Fe-based superconductors (z1/3z\sim 1/3), and it was revealed that the eigenvalue of the orbital fluctuations is reduced by the self-energy, and the orbital order temperature is reduced to 100\sim 100K, consistently with experiments.

Refer to caption
Figure 11: Obtained eigenvalues (a) η𝒒\eta_{{\bm{q}}} and (b) λ𝒒\lambda_{{\bm{q}}} as functions of zz.

Here, we study the effect of the renormalization factor zz on the eigenvalues in the present single-orbital Hubbard model. Then, the Green function is given as

Gz(k)=1iϵn/zμϵ𝒌.\displaystyle G^{z}(k)=\frac{1}{i{\epsilon}_{n}/z-\mu-{\epsilon}_{{\bm{k}}}}. (18)

First, we discuss the effect of zz on λSC{\lambda}_{\rm SC}, by replacing two GG’s in Eq. (7) with GzG^{z} given by Eq. (18). On the other hand, we do not include zz in the susceptibilities χs,c\chi^{s,c} in Vs,cV^{s,c} in the pairing interaction, in order not to change the Stoner factor αS{\alpha}_{S}. Then, it is well known that λSC{\lambda}_{\rm SC} is reduced as zλSCz\cdot{\lambda}_{\rm SC} based on the Eliashberg equation Allen .

Next, we discuss the effect of zz on λ𝒒\lambda_{{\bm{q}}} (η𝒒\eta_{{\bm{q}}}) by replacing four GG’s in Eqs. (12) and (13) (in Eqs. (14) and (15)) with GzG^{z} given by Eq. (18). The obtained zz-dependences of the eigenvalues at T=0.03T=0.03 are shown in Fig. 11. It is verified that both λ𝒒\lambda_{{\bm{q}}} and η𝒒\eta_{{\bm{q}}} are reduced in proportion to zz. Although this approximation may be justified only for z1z\lesssim 1, the obtained results strongly indicate that both λ𝒒\lambda_{{\bm{q}}} and η𝒒\eta_{{\bm{q}}} are reduced to O(1)O(1) in the case of z0.2z\lesssim 0.2, which is realized in cuprate superconductors.

B.4 Unfolded band structure under the sLC order

Refer to caption
Figure 12: The unfolded band structure in the single-𝑸d{{\bm{Q}}}_{\rm d} sLC order corresponds to Fig. 6 (a) in the main text.

Here, we examine the experimentally observed band structure in the sLC ordered state, by applying the unfolding procedure proposed in Ref. Ku . Figure 12 shows the “unfolded” band structure in the single-𝑸d{{\bm{Q}}}_{\rm d} sLC order at gmax=0.1g^{\rm max}=0.1, which corresponds to Fig. 6 (a) in the main text. The pseudogap closes on the X-Y line owing to the odd-parity form factor. This Dirac point which will be smeared out for TT(TCDW)T\sim T^{*}\ (\gg T_{\rm CDW}) because of very large inelastic scattering at the hot spot Kontani-ROP ; Tremblay ; Scalapino ; Moriya . In addition, the Dirac point should be masked by the dd-wave BO below TCDWT_{\rm CDW}.

Appendix C BO/sLC order as magnon-pair condensation

We explain that the sLC order is exactly the same as the magnon-pair condensation. The following spin quadrupole order occurs owing to the magnon-pair condensation Andreev :

Ki,jα,β\displaystyle K_{i,j}^{{\alpha},{\beta}} \displaystyle\equiv siαsjβsiαsjβ0,\displaystyle\langle s_{i}^{\alpha}s_{j}^{\beta}\rangle-\langle s_{i}^{\alpha}s_{j}^{\beta}\rangle_{0}, (19)

where α,β=x,y,z{\alpha},{\beta}=x,y,z, and the relation Ki,jα,β=Kj,iβ,αK_{i,j}^{{\alpha},{\beta}}=K_{j,i}^{{\beta},{\alpha}} holds. We will explain that the even-parity function ai,jKi,jα,α/3a_{i,j}\equiv K_{i,j}^{{\alpha},{\alpha}}/3  (with ai,j=aj,ia_{i,j}=a_{j,i}) corresponds the BO state, and the odd-parity function bi,jαiϵαβγKi,jβ,γ/2b_{i,j}^{\alpha}\equiv i{\epsilon}_{{\alpha}{\beta}{\gamma}}K_{i,j}^{{\beta},{\gamma}}/2  (with bi,jα=bj,iαb_{i,j}^{\alpha}=-b_{j,i}^{\alpha}) corresponds the sLC order.

Refer to caption
Figure 13: Diagrammatic expression of Γ¯𝒒s(c)(p,p)\bar{\Gamma}^{s(c)}_{{\bm{q}}}(p,p^{\prime}), which represents the scattering process of the magnon pair through the interaction J𝒒s(c)(p,p)J_{{\bm{q}}}^{s(c)}(p,p^{\prime}). Mathematically, Γ¯𝒒s(c)(p,p)\bar{\Gamma}^{s(c)}_{{\bm{q}}}(p,p^{\prime}) diverges when magnon pairs with momentum 𝒒{{\bm{q}}} condense. Thus, sLC/BO is interpreted as the condensation of odd/even parity magnon pairs.

Here, we explain that Γ𝒒s,c(k,k)\Gamma^{s,c}_{{\bm{q}}}(k,k^{\prime}) due to the AL processes represents the scattering between two-magnons. To simplify the discussion, we drop the MT term, and consider only AL terms with two χs\chi^{s}s. Then, we define Γ¯𝒒s(c)(p,p)\bar{\Gamma}^{s(c)}_{{\bm{q}}}(p,p^{\prime}) by the following relation; Γ𝒒c(s)(k,k)=T2p,p[G(kp)+()G(k+p)]Γ¯𝒒c(s)(p,p)G(kp)\Gamma^{c(s)}_{{\bm{q}}}(k,k^{\prime})=T^{2}\sum_{p,p^{\prime}}[G(k-p)+(-)G(k+p)]\bar{\Gamma}^{c(s)}_{{\bm{q}}}(p,p^{\prime})G(k^{\prime}-p). Figure 13 shows the diagrammatic expression of Γ¯𝒒s,c(p,p)\bar{\Gamma}^{s,c}_{{\bm{q}}}(p,p^{\prime}), which represents the scattering process of magnon pair amplitude bzb^{z} (aa) through the interaction J𝒒s(c)(p,p)J_{{\bm{q}}}^{s(c)}(p,p^{\prime}), which is a moderate function of TT. With decreasing temperatures, Γ¯𝒒c(s)(p,p)\bar{\Gamma}^{c(s)}_{{\bm{q}}}(p,p^{\prime}) diverges when singlet (triplet) magnon pairs with momentum 𝒒{{\bm{q}}} condensate, and the critical temperature corresponds to λ𝒒=1\lambda_{{\bm{q}}}=1 (η𝒒=1\eta_{{\bm{q}}}=1).

Here, we introduce f¯𝒒(k)TpH𝒒(k,p)f𝒒(p)\bar{f}_{{\bm{q}}}(k)\equiv T\sum_{p}H_{{\bm{q}}}(k,p)f_{{\bm{q}}}(p) and g¯𝒒(k)TpH𝒒(k,p)g𝒒(p)\bar{g}_{{\bm{q}}}(k)\equiv T\sum_{p}H_{{\bm{q}}}(k,p)g_{{\bm{q}}}(p), where Hq(k,p)=G(p+)G(p)G(pk)H_{q}(k,p)=G(p_{+})G(p_{-})G(p-k), and f𝒒(p)f_{{\bm{q}}}(p) and g𝒒(p)g_{{\bm{q}}}(p) are form factors of the DW equations. Then, the DW equations are rewritten as

λ𝒒f¯𝒒(k)\displaystyle\lambda_{{\bm{q}}}\bar{f}_{{\bm{q}}}(k) =\displaystyle= TpJ𝒒c(k,p)χs(p+)χs(p)f¯𝒒(p)\displaystyle T\sum_{p}J^{c}_{{\bm{q}}}(k,p)\chi^{s}(p_{+})\chi^{s}(p_{-})\bar{f}_{{\bm{q}}}(p) (20)
η𝒒g¯𝒒(k)\displaystyle\eta_{{\bm{q}}}\bar{g}_{{\bm{q}}}(k) =\displaystyle= TpJ𝒒s(k,p)χs(p+)χs(p)g¯𝒒(k)\displaystyle T\sum_{p}J^{s}_{{\bm{q}}}(k,p)\chi^{s}(p_{+})\chi^{s}(p_{-})\bar{g}_{{\bm{q}}}(k) (21)

where the kernel function J𝒒c,s(k,p)J^{c,s}_{{\bm{q}}}(k,p) is given in Fig. 13. These equations means that f¯𝒒(k)\bar{f}_{{\bm{q}}}(k) (g¯𝒒(k)\bar{g}_{{\bm{q}}}(k)) corresponds to the singlet (triplet) magnon pair condensation. Therefore, their Fourier transformations correspond to ai,ja_{i,j} and bi,jzb_{i,j}^{z}, respectively.

To summarize, in the present double spin-flip mechanism, magnon-pair condensation a,bz0a,b^{z}\neq 0 occurs at T=TsLCT=T_{\rm sLC}. Therefore, the sLC/BO given by the present mechanism is exactly the same as “condensation of odd/even parity magnon pairs”.

Appendix D Spontaneous spin current in the sLC phase

D.1 Calculation of the spin current

Here, we investigate the spin current in real space due to the “spin-dependent self-energy” δti,jσ=σgi,j\delta t_{i,j}^{\sigma}={\sigma}g_{i,j} shown in Fig. 3 (c), which is purely imaginary and odd with respect to iji\leftrightarrow j. The spin current operator from site jj to site ii is ji,jσ=iσσ(hi,jσciσcjσ(ij))j_{i,j}^{\sigma}=-i\sum_{{\sigma}}{\sigma}(h_{i,j}^{\sigma}c_{i{\sigma}}^{\dagger}c_{j{\sigma}}-(i\leftrightarrow j)), where hi,jσ=ti,j+δti,jσh_{i,j}^{\sigma}=t_{i,j}+\delta t_{i,j}^{\sigma}. Then, the spin current from jj to ii is given as Ji,js=ji,jsh^σJ_{i,j}^{s}=\langle j_{i,j}^{s}\rangle_{\hat{h}^{\sigma}}.

Refer to caption
Figure 14: Obtained spin current from the center site (A-D) to different sites under the diagonal sLC with period 4aCuCu4a_{\rm Cu-Cu}. The real space pattern is depicted in Fig. 1 (c), with sites A-D in a unit cell.

Here, we calculate the spin current for the commensurate sLC order at 𝒒sLC=(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,\pi/2), which is achieved by putting n=1.0n=1.0. Then, the unit cell are composed four sites A-D. Figure 14 shows the obtained spin current Ji,jsJ_{i,j}^{s} from the center site (j=Aj={\rm A}-B{\rm B}) to different site in Fig. 1 (c), by setting gmax=0.1g^{\rm max}=0.1. The obtained current is |Ji,js|102|J_{i,j}^{s}|\sim 10^{-2} in unit |t1|/|t_{1}|/\hbar. The derived spin current pattern between the nearest and second-nearest sites is depicted in Fig. 1 (c). The spin current is exactly conserved at each site.

D.2 Possible diagonal sLC patterns in real space

Refer to caption
Figure 15: Examples of the diagonal sLC pattern in real space for 𝒒sLC=(π/2,π/2),(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,\pi/2),(\pi/2,-\pi/2). (a) Single-𝒒{{\bm{q}}} sLC pattern for ψ=π/2\psi=\pi/2. (b)-(e) Four examples of double-𝒒{{\bm{q}}} sLC patterns.

Next, we explain that the spin current pattern derived from the form factor g𝒒(𝒌)g_{{\bm{q}}}({{\bm{k}}}) in Fig. 3 (b) is not uniquely determined. In fact, the form factor in real space is given as iIm{gi,jeiψ}iIm{ei𝒒(𝒓i+𝒓j)/2eiψ}i{\rm Im}\{g_{i,j}e^{i\psi}\}\sim i{\rm Im}\{e^{i\bm{q}\cdot(\bm{r}_{i}+\bm{r}_{j})/2}e^{i\psi}\}, where ψ\psi is an arbitrary phase. Here, we discuss other possible spin current patterns by choosing ψ\psi.

First, we discuss the real space pattern for 𝒒sLC=(π/2,π/2),(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,\pi/2),(\pi/2,-\pi/2). We assume that Fig. 1 (c) corresponds to ψ=0\psi=0. Then, the single-𝒒{{\bm{q}}} spin current pattern for ψ=π/2\psi=\pi/2 is given in Fig. 15 (a). The double-𝒒{{\bm{q}}} spin current order is given by the combination of the sLC order at 𝒒sLC=(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,\pi/2) and that at 𝒒sLC=(π/2,π/2){{\bm{q}}}_{\rm sLC}=(\pi/2,-\pi/2) with arbitrary phase factors. Figure 15 (b)-(c) are given by the combination of Fig. 1 (c) with its π/2\pi/2-rotation, and Figs. 15 (d)-(e) are given by the combination of Figs. 15 (a) with its π/2\pi/2-rotation. We stress that the magnitude of spin current |Ji,js||J^{s}_{i,j}| in Fig. 15 (b)-(c) has C4C_{4} symmetry, whereas that in Fig. 15 (d)-(e) breaks the C4C_{4} symmetry.

References

  • (1) G. Ghiringhelli, M. L. Tacon, M. Minola, S. Blanco-Canosa, C. Mazzoli, N. B. Brookes, G. M. D. Luca, A. Frano, D. G. Hawthorn, F. He, T. Loew, M. M. Sala, D. C. Peets, M. Salluzzo, E. Schierle, R. Sutarto, G. A. Sawatzky, E. Weschke, B. Keimer, and L. Braicovich, Long-range incommensurate charge fluctuations in (Y,Nd)Ba2Cu3O6+x, Science 337, 821 (2012).
  • (2) R. Comin, A. Frano, M. M. Yee, Y. Yoshida, H. Eisaki, E. Schierle, E. Weschke, . Sutarto, F. He, A. Soumyanarayanan, Y. He, M. L. Tacon, I. S. Elfimov, J. E. Hffman, G. A. Sawatzky, B. Keimer, and A. Damascelli, Charge order driven by Fermi-arc instability in Bi2Sr2-xLaxCuO, Science 343, 390 (2014).
  • (3) Y. Kohsaka, T. Hanaguri, M. Azuma, M. Takano, J. C. Davis, and H. Takagi, Visualization of the emergence of the pseudogap state and the evolution to superconductivity in a lightly hole-doped Mott insulator, Nature Physics 8, 534 (2012).
  • (4) K. Fujita, M. H. Hamidian, S. D. Edkins, C. K. Kim, Y. Kohsaka, M. Azuma, M. Takano, H. Takagi, H. Eisaki, S. Uchida, A. Allais, M. J. Lawler, E.-A. Kim, S. Sachdev, and J. C. S. Davis, Direct phase-sensitive identification of a dd-form factor density wave in underdoped cuprates, Proc. Natl. Acad. Sci. USA, 111, E3026 (2014).
  • (5) B. Kyung, S. S. Kancharla, D. Senechal, A.-M. S. Tremblay, M. Civelli, and G. Kotliar, Pseudogap induced by short-range spin correlations in a doped Mott insulator, Phys. Rev. B 73, 165114 (2006).
  • (6) T.A. Maier, M.S. Jarrell, and D.J. Scalapino, Understanding high-temperature superconductors with quantum cluster theories, Physica C, 460-462, 13 (2007).
  • (7) T. Moriya and K. Ueda, Spin fluctuations and high temperature superconductivity, Adv. Phys. 49, 555 (2000).
  • (8) A.Shekhter, B. J. Ramshaw, R. Liang, W. N. Hardy, D. A. Bonn, F. F. Balakirev, R. D. McDonald, J. B. Betts, S. C. Riggs, and A. Migliori, Bounding the pseudogap with a line of phase transitions in YBa2Cu3O6+δ, Nature 498, 75 (2013).
  • (9) R.-H. He, M. Hashimoto, H. Karapetyan, J. D. Koralek, J. P. Hinton, J. P. Testaud, V. Nathan, Y. Yoshida, H. Yao, K. Tanaka, W. Meevasana, R. G. Moore, D. H. Lu, S.-K. Mo, M. Ishikado, H. Eisaki, Z. Hussain, T. P. Devereaux, S. A. Kivelson, J. Orenstein, A. Kapitulnik, and Z.-X. Shen, From a Single-Band Metal to a High-Temperature Superconductor via Two Thermal Phase Transitions, Science 331, 1579 (2011).
  • (10) Y. Sato, S. Kasahara, H. Murayama, Y. Kasahara, E. -G. Moon, T. Nishizaki, T. Loew, J. Porras, B. Keimer, T. Shibauchi, and Y. Matsuda, Thermodynamic evidence for a nematic phase transition at the onset of the pseudogap in YBa2Cu3Oy, Nature Physics 13, 1074 (2017).
  • (11) H. Murayama, Y. Sato, R. Kurihara, S. Kasahara, Y. Mizukami, Y. Kasahara, H. Uchiyama, A. Yamamoto, E. -G. Moon, J. Cai, J. Freyermuth, M. Greven, T. Shibauchi, and Y. Matsuda, Diagonal nematicity in the pseudogap phase of HgBa2CuO4+δ, Nat. Commun. 10, 3282 (2019).
  • (12) S. Nakata, M. Horio, K. Koshiishi, K. Hagiwara, C. Lin, M. Suzuki, S. Ideta, K. Tanaka, D. Song, Y. Yoshida, H. Eisaki, and A. Fujimori, Nematicity in the pseudogap state of cuprate superconductors revealed by angle-resolved photoemission spectroscopy, arXiv:1811.10028.
  • (13) K. Ishida, S. Hosoi, Y. Teramoto, T. Usui, Y. Mizukami, K. Itaka, Y. Matsuda, T. Watanabe, and T. Shibauchi, Divergent nematic susceptibility near the pseudogap critical point in a cuprate superconductor, J. Phys. Soc. Jpn. 89, 064707 (2020).
  • (14) S. Bulut, W.A. Atkinson and A.P. Kampf, Spatially modulated electronic nematicity in the three-band model of cuprate superconductors Phys. Rev. B 88, 155132 (2013).
  • (15) Y. Wang and A.V. Chubukov, Charge-density-wave order with momentum (2Q,0)(2Q,0) and (0,2Q)(0,2Q) within the spin-fermion model: Continuous and discrete symmetry breaking, preemptive composite order, and relation to pseudogap in hole-doped cuprates, Phys. Rev. B 90, 035149 (2014).
  • (16) R.-Q. Xing, L. Classen, and A. V. Chubukov, Orbital order in FeSe: The case for vertex renormalization, Phys. Rev. B 98, 041108(R) (2018).
  • (17) M.A. Metlitski and S. Sachdev, Instabilities near the onset of spin density wave order in metals, New J. Phys. 12, 105007 (2010); S. Sachdev and R. La Placa, Bond Order in Two-Dimensional Metals with Antiferromagnetic Exchange Interactions, Phys. Rev. Lett. 111, 027202 (2013).
  • (18) W. Metzner, M. Salmhofer, C. Honerkamp, V. Meden, and K. Schönhammer, Functional renormalization group approach to correlated fermion systems, Rev. Mod. Phys. 84, 299 (2012); C. Husemann and W. Metzner, Incommensurate nematic fluctuations in the two-dimensional Hubbard model, Phys. Rev. B 86, 085113 (2012); T. Holder and W. Metzner, Incommensurate nematic fluctuations in two-dimensional metals, Phys. Rev. B 85, 165130 (2012); C. Honerkamp, Charge instabilities at the metamagnetic transition of itinerant electron systems, Phys. Rev. B 72, 115103 (2005).
  • (19) J. C. S. Davis and D.-H. Lee, Concepts relating magnetic interactions, intertwined electronic orders, and strongly correlated superconductivity, Proc. Natl. Acad. Sci. USA, 110, 17623 (2013).
  • (20) E. Berg, E. Fradkin, S. A. Kivelson, and J. M. Tranquada, Striped superconductors: how spin, charge and superconducting orders intertwine in the cuprates, New J. Phys. 11, 115004 (2009).
  • (21) Y. Yamakawa, and H. Kontani, Spin-Fluctuation-Driven Nematic Charge-Density Wave in Cuprate Superconductors: Impact of Aslamazov-Larkin Vertex Corrections, Phys. Rev. Lett. 114, 257001 (2015).
  • (22) M. Tsuchiizu, K. Kawaguchi, Y. Yamakawa, and H. Kontani, Multistage electronic nematic transitions in cuprate superconductors: A functional-renormalization-group analysis, Phys. Rev. B 97, 165131 (2018).
  • (23) K. Kawaguchi, Y. Yamakawa, M. Tsuchiizu, and H. Kontani, Competing Unconventional Charge-Density-Wave States in Cuprate Superconductors: Spin-Fluctuation-Driven Mechanism, J. Phys. Soc. Jpn. 86, 063707 (2017).
  • (24) P. A. Lee, Amperean Pairing and the Pseudogap Phase of Cuprate Superconductors, Phys. Rev. X 4, 031017 (2014).
  • (25) D. F. Agterberg, J. C. S. Davis, S. D. Edkins, E. Fradkin, D. J. Van Harlingen, S. A. Kivelson, P. A. Lee, L. Radzihovsky, J. M. Tranquada, and Y. Wang, The Physics of Pair Density Waves, Annu. Rev. Condens. Matter Phys. 11, 231 (2020).
  • (26) C. M. Varma, Non-Fermi-liquid states and pairing instability of a general model of copper oxide metals, Phys. Rev. B 55, 14554 (1997).
  • (27) I. Affleck and J. B. Marston, Large-nn limit of the Heisenberg-Hubbard model: Implications for high-TcT_{c} superconductors, Phys. Rev. B 37, 3774(R) (1988).
  • (28) F. C. Zhang, Superconducting instability of staggered-flux phase in the t-J model, Phys. Rev. Lett. 64, 974 (1990).
  • (29) H. J. Schulz, Fermi-surface instabilities of a generalized two-dimensional Hubbard model, Phys. Rev. B 39, 2940(R) (1989).
  • (30) A. A. Nersesyan, G. I. Japaridze, and I. G. Kimeridze, Low-temperature magnetic properties of a two-dimensional spin nematic state, J. Phys.: Condens. Matter 3, 3353 (1991).
  • (31) M. Ozaki, Broken Symmetry Solutions of the Extended Hubbard Model, Int. J. Quantum. Chem. 42, 55 (1992).
  • (32) S. Zhou, K. Jiang, H. Chen, and Z. Wang, Correlation Effects and Hidden Spin-Orbit Entangled Electronic Order in Parent and Electron-Doped Iridates Sr2IrO4, Phys. Rev. X 7, 041018 (2017).
  • (33) H. Ikeda and Y. Ohashi, Theory of Unconventional Spin Density Wave: A Possible Mechanism of the Micromagnetism in U-based Heavy Fermion Compounds, Phys. Rev. Lett. 81, 3723 (1998).
  • (34) S. Fujimoto, Spin Nematic State as a Candidate of the Hidden Order Phase of URu2Si2, Phys. Rev. Lett. 106, 196407 (2011).
  • (35) S. Onari and H. Kontani, Self-consistent Vertex Correction Analysis for Iron-based Superconductors: Mechanism of Coulomb Interaction-Driven Orbital Fluctuations, Phys. Rev. Lett. 109, 137001 (2012).
  • (36) S. Onari, Y. Yamakawa, and H. Kontani, Sign-Reversing Orbital Polarization in the Nematic Phase of FeSe due to the C2C_{2} Symmetry Breaking in the Self-Energy, Phys. Rev. Lett. 116, 227001 (2016).
  • (37) Y. Yamakawa, S. Onari, and H. Kontani, Nematicity and Magnetism in FeSe and Other Families of Fe-Based Superconductors, Phys. Rev. X 6, 021032 (2016).
  • (38) T. Yoshida, M. Hashimoto, I. M. Vishik, Z. X. Shen, and A. Fujimori, Pseudogap, Superconducting Gap, and Fermi Arc in High-TcT_{c} Cuprates Revealed by Angle-Resolved Photoemission Spectroscopy, J. Phys. Soc. Jpn. 81, 011006 (2011).
  • (39) Y. K. Kim, O. Krupin, J. D. Denlinger, A. Bostwick, E. Rotenberg, Q. Zhao, J. F. Mitchell, J. W. Allen, an B. J. Kim, Fermi arcs in a doped pseudospin-1/2 Heisenberg antiferromagnet, Science 345, 187 (2014).
  • (40) H. Kontani, Anomalous Transport Phenomena in Fermi Liquids with Strong Magnetic Fluctuations, Rep. Prog. Phys. 71, 026501 (2008);
  • (41) H. Kontani, Transport Phenomena in Strongly Correlated Fermi Liquids (Springer-Verlag Berlin and Heidelberg GmbH & Co. K, 2013).
  • (42) C. Stock, W. J. L. Buyers, R. Liang, D. Peets, Z. Tun, D. Bonn, W. N. Hardy, and R. J. Birgeneau, Dynamic stripes and resonance in the superconducting and normal phases of YBa2Cu3O6.5 ortho-II superconductor, Phys. Rev. B 69, 014502 (2004).
  • (43) Hexadecapole Fluctuation Mechanism for s-wave Heavy Fermion Superconductor CeCu2Si2: Interplay between Intra- and Inter-Orbital Cooper Pairs J. Phys. Soc. Jpn. 88, 063701 (2019); R. Tazai, Y. Yamakawa, and H. Kontani, Plain s-wave superconductivity near the magnetic criticality: Enhancement of attractive electron-boson coupling vertex corrections, J. Phys. Soc. Jpn. 86, 073703 (2017).
  • (44) R. Tazai and H. Kontani, Multipole fluctuation theory for heavy fermion systems: Application to multipole orders in CeB6, Phys. Rev. B 100, 241103(R) (2019).
  • (45) M. Tsuchiizu, Y. Ohno, S. Onari, and H. Kontani, Orbital Nematic Instability in the Two-Orbital Hubbard Model: Renormalization-Group + Constrained RPA Analysis, Phys. Rev. Lett. 111, 057003 (2013).
  • (46) R. Tazai, Y. Yamakawa, M. Tsuchiizu, and H. Kontani, Functional renormalization group study of orbital fluctuation mediated superconductivity: Impact of the electron-boson coupling vertex corrections, Phys. Rev. B 94, 115155 (2016).
  • (47) R. Tazai, Y. Yamakawa, M. Tsuchiizu, and H. Kontani, Prediction of dd-wave bond-order and pseudogap in organic superconductor κ\kappa-(BEDT-TTF)2X: Similarities to cuprate superconductors, arXiv:2010.15516.
  • (48) R. Tazai, Y. Yamakawa, and H. Kontani, Emergence of Charge Loop Current in Geometrically Frustrated Hubbard Model: Functional Renormalization Group Study, arXiv:2010.16109.
  • (49) S. Onari and H. Kontani, Origin of diverse nematic orders in Fe-based superconductors: 45 degree rotated nematicity in AFe2As2 (A=Sc,Rb), Phys. Rev. B 100, 020507(R) (2019).
  • (50) S. Onari and H. Kontani, Hidden antiferro-nematic order in Fe-based superconductor BaFe2As2 and NaFeAs above TST_{S}, Phys. Rev. Research 2, 042005 (2020).
  • (51) D. J. Scalapino, E. Loh, and J. E. Hirsch, d-wave pairing near a spin-density-wave instability, Phys. Rev. B 34, 8190 (1986).
  • (52) V. Mishra and M. R. Norman, Strong coupling critique of spin fluctuation driven charge order in underdoped cuprates, Phys. Rev. B 92, 060507(R) (2015).
  • (53) A. F. Andreev and I. A. Grishchuk, Spin nematics, Zh. Eksp. Teor. Fiz. 87, 467 (1984).
  • (54) P. Chandra and P. Coleman, Quantum spin nematics: Moment-free magnetism, Phys. Rev. Lett. 66, 100 (1991).
  • (55) N. Shannon, T. Momoi, and P. Sindzingre, Nematic Order in Square Lattice Frustrated Ferromagnets, Phys. Rev. Lett. 96, 027213 (2006).
  • (56) W. Ku, T. Berlijn, and C.-C. Lee, Unfolding First-Principles Band Structures, Phys. Rev. Lett. 104, 216401 (2010).
  • (57) K. Kawaguchi and H. Kontani, Spin-fluctuation-driven B1gB_{1g} and B2gB_{2g} Bond Order and Induced In-plane Anisotropy in Magnetic Susceptibility in Cuprate Superconductors: Impact of Hot-/Cold-spot Structure on Bond-order Symmetry, J. Phys. Soc. Jpn. 89, 124704 (2020).
  • (58) V. Hinkov, D. Haug, B. Fauque, P. Bourges, Y. Sidis, A. Ivanov, C. Bernhard, C. T. Lin, and B. Keimer, Electronic Liquid Crystal State in the High-Temperature Superconductor YBa2Cu3O6.45, Science 319, 597 (2008).
  • (59) P. B. Allen and B. Mitrovic, Theory of Superconducting TcT_{c}, Solid State Physics, 37, 1 (1983)