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

Coexistence of ss- and dd-wave gaps due to pair-hopping and exchange interactions

Shigeru Koikegami Second Lab, LLC, 19-27 Inarimae, Tsukuba 305-0061, Japan
Abstract

I investigate the superconductivity of the three-band ttJJUU model derived from the three-band Hubbard model using the Schrieffer–Wolff transformation. My model is designed considering the hole-doped high-TcT_{\mathrm{c}} superconducting cuprate. The model does not exclude the double occupancy of Cu sites by dd electrons, and there is a pair-hopping interaction between the dd and pp bands together with the exchange interaction. I analyse the superconducting transition temperature, electronic state, and superconducting gap function based on strong coupling theory and find that the superconductivity emerges due to the pair-hopping and exchange interactions via the Suhl-Kondo mechanism. In the superconducting state, the extended ss- and dx2y2d_{x^{2}-y^{2}}-wave superconducting gaps coexist, where both charge fluctuations and ddpp band hybridization are key ingredients.

  • Received 12 April 2021, revised 26 June 2021

1 Introduction

The ttJJ model is one of the model Hamiltonians that form the basis of many theoretical studies of strongly correlated electron systems [1, 2]. The ttJJ model can also be derived as the low-energy effective Hamiltonian of the two-dimensional (2D) multiband Hubbard model [3, 4, 5], regarded as the fundamental model Hamiltonian for the high-TcT_{\mathrm{c}} superconducting cuprate (HTSC). Many theoretical studies of HTSC to date use the ttJJ model as the model Hamiltonian [6, 7, 8, 10, 9, 11, 12, 13]. These studies often exclude the double occupancy of Cu sites by dd electrons, considering that the on-site Coulomb repulsion between dd orbitals is much larger than the transfer energy between the dd and pp orbitals. As a result, the ttJJ model contains only one electron (or hole) band and a localized spin.

However, the double occupancy of Cu sites need not necessarily be excluded when the on-site Coulomb repulsion UU is comparable to the transfer energy. Relaxing the single occupancy constraint and explicitly considering UU instead results in the ttJJUU model that includes both the ttJJ model and the single-band Hubbard model as one of its limits [14, 15, 16, 17, 18, 19, 20]. Thus, the ttJJUU model serves as an interpolation between the ttJJ model and the single-band Hubbard model and is able to account for more properties caused by strong correlation. However, the charge transfer gap should be comparable to tt in the charge transfer regime. In this case, pp electron scattering by dd electrons cannot be negligible, and both pp and dd electrons must be considered.

In this paper, I derive the three-band ttJJUU model from the 2D three-band Hubbard model as its effective Hamiltonian by using the Schrieffer–Wolff (SW) transformation [21] and assume that double occupancy is not excluded. In my model, the pair-hopping interaction between the dd and pp bands exists separately from the exchange interaction. Treating these interactions using iterative perturbation theory (IPT) approximation, I investigate the superconductivity of the model in a strong coupling framework. The results show that the multicomponent superconductivity emerges with the hole doping, which introduces the dd-pp band hybridization through exchange and pair-hopping interactions. This emergence of the superconductivity is due to the pair-hopping and exchange interactions via the Suhl-Kondo (SK) mechanism [22, 23, 24], which stabilizes the superconducting gaps with different signs in a multiband system. In the superconducting state, the extended ss- and dx2y2d_{x^{2}-y^{2}}-wave superconducting gaps coexist, and the ss- and dd-wave gaps emerge due to the pair-hopping and exchange interactions, respectively.

2 Formulation

Consider the three-band Hubbard model [25] that expresses the Hamiltonian as =0+α1α{\mathcal{H}}={\mathcal{H}}_{0}+\sum_{\alpha}{\mathcal{H}}_{1}^{\alpha}, where

0=εdjσdjσdjσ+εpα𝐤σp𝐤σαp𝐤σα+Ujdjdjdjdj{\mathcal{H}}_{0}=\varepsilon_{d}\sum_{j\sigma}d_{j\sigma}^{\dagger}d_{j\sigma}+\varepsilon_{p}\sum_{\alpha}\sum_{{\mathbf{k}}\sigma}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}\sigma}^{\alpha}+U\sum_{j}d_{j\uparrow}^{\dagger}d_{j\uparrow}d_{j\downarrow}^{\dagger}d_{j\downarrow} (1)

and

1α=1Nj𝐤σ(Vα𝐤ei𝐤𝐑jp𝐤σαdjσ+H.c.).{\mathcal{H}}_{1}^{\alpha}=\frac{1}{\sqrt{N}}\sum_{j}\sum_{{\mathbf{k}}\sigma}\left(V_{\alpha{\mathbf{k}}}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}d_{j\sigma}+{\mathrm{H.c.}}\right). (2)

Here, α{x,y}\alpha\in\{x,y\}; djσ(djσ)d_{j\sigma}(d_{j\sigma}^{\dagger}) is the annihilation (creation) operator for the dd electron of spin σ\sigma at Cu site jj; p𝐤σα(p𝐤σα)p_{{\mathbf{k}}\sigma}^{\alpha}(p_{{\mathbf{k}}\sigma}^{\alpha\dagger}) is the annihilation (creation) operator for pαp^{\alpha} electrons of spin σ\sigma with momentum 𝐤{\mathbf{k}}, based on oxygen sites in real space; εd\varepsilon_{d} and εp\varepsilon_{p} are the dd and pp electron site energies, respectively; UU is the on-site Coulomb repulsion between dd orbitals; and NN is the number of k-space points in the first Brillouin zone (FBZ). The lattice constant of the square lattice of Cu sites is the length unit. Thus, Vx𝐤=2itpdsinkx2V_{x{\mathbf{k}}}=2{\mathrm{i}}t_{pd}\sin\frac{k_{x}}{2} and Vy𝐤=2itpdsinky2V_{y{\mathbf{k}}}=-2{\mathrm{i}}t_{pd}\sin\frac{k_{y}}{2}, where tpdt_{pd} is the transfer energy between the dd orbital and the neighbouring pαp^{\alpha} orbital.

In order to derive the effective Hamiltonian for \mathcal{H}, I adopt the SW transformation as follows:

eα𝒮αeβ𝒮β\displaystyle e^{\sum_{\alpha}{\mathcal{S}}^{\alpha}}{\mathcal{H}}e^{-\sum_{\beta}{\mathcal{S}}^{\beta}} =\displaystyle= 0+α1α+α[𝒮α,0]+αβ[𝒮α,1β]+12αβ[𝒮α,[𝒮β,0]]+\displaystyle{\mathcal{H}}_{0}+\sum_{\alpha}{\mathcal{H}}_{1}^{\alpha}+\sum_{\alpha}\left[{\mathcal{S}}^{\alpha},{\mathcal{H}}_{0}\right]+\sum_{\alpha\beta}\left[{\mathcal{S}}^{\alpha},{\mathcal{H}}_{1}^{\beta}\right]+\frac{1}{2}\sum_{\alpha\beta}\left[{\mathcal{S}}^{\alpha},\left[{\mathcal{S}}^{\beta},{\mathcal{H}}_{0}\right]\right]+\ldots (3)
=\displaystyle= 0+12αβ[𝒮α,1β]+,\displaystyle{\mathcal{H}}_{0}+\frac{1}{2}\sum_{\alpha\beta}\left[{\mathcal{S}}^{\alpha},{\mathcal{H}}_{1}^{\beta}\right]+\ldots,

using 1α+[𝒮α,0]=0{\mathcal{H}}_{1}^{\alpha}+\left[{\mathcal{S}}^{\alpha},{\mathcal{H}}_{0}\right]=0 and

𝒮α=1Nj𝐤σ(Vα𝐤ei𝐤𝐑jΔpdUndjσp𝐤σαdjσ+Vα𝐤ei𝐤𝐑jΔpd(1ndjσ)p𝐤σαdjσ)H.c.{\mathcal{S}}^{\alpha}=\frac{1}{\sqrt{N}}\sum_{j}\sum_{{\mathbf{k}}\sigma}\left(\frac{V_{\alpha{\mathbf{k}}}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}}{\mathit{\Delta}_{pd}-U}\,n_{d\,j-\sigma}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}d_{j\sigma}+\frac{V_{\alpha{\mathbf{k}}}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}}{\mathit{\Delta}_{pd}}(1-n_{d\,j-\sigma})p_{{\mathbf{k}}\sigma}^{\alpha\dagger}d_{j\sigma}\right)-{\mathrm{H.c.}} (4)

Here, Δpdεpεd{\mathit{\Delta}}_{pd}\equiv\varepsilon_{p}-\varepsilon_{d}, ndjσdjσdjσn_{d\,j\sigma}\equiv d_{j\sigma}^{\dagger}d_{j\sigma}, and H.c. indicates the Hermitian conjugate of the terms already written. The observable ndjσn_{d\,j\sigma} has 0 or 11 as its eigenvalue for each jj and σ\sigma. Using Eqs. (2) and (4), the following results:

[𝒮α,1β]\displaystyle\left[{\mathcal{S}}^{\alpha},{\mathcal{H}}_{1}^{\beta}\right] =δαβNjj𝐤σ(ndjσΔpdU+1ndjσΔpd)Vα𝐤Vβ𝐤ei𝐤𝐑jei𝐤𝐑jdjσdjσ\displaystyle=-\frac{\delta_{\alpha\beta}}{N}\sum_{jj^{\prime}}\sum_{{\mathbf{k}}\sigma}\left(\frac{n_{d\,j-\sigma}}{\mathit{\Delta}_{pd}-U}+\frac{1-n_{d\,j-\sigma}}{\mathit{\Delta}_{pd}}\right)V_{\alpha{\mathbf{k}}}V_{\beta{\mathbf{k}}}^{*}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}e^{{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j^{\prime}}}d_{j^{\prime}\sigma}^{\dagger}d_{j\sigma}
+1Nj𝐤𝐤σ(ndjσΔpdU+1ndjσΔpd)Vα𝐤Vβ𝐤ei𝐤𝐑jei𝐤𝐑jp𝐤σαp𝐤σβ\displaystyle+\frac{1}{N}\sum_{j}\sum_{{\mathbf{k}}{\mathbf{k}}^{\prime}\sigma}\left(\frac{n_{d\,j-\sigma}}{\mathit{\Delta}_{pd}-U}+\frac{1-n_{d\,j-\sigma}}{\mathit{\Delta}_{pd}}\right)V_{\alpha{\mathbf{k}}}V_{\beta{\mathbf{k}}^{\prime}}^{*}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}e^{{\mathrm{i}}{\mathbf{k}}^{\prime}\cdot{\mathbf{R}}_{j}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}^{\prime}\sigma}^{\beta}
1Nj𝐤𝐤σ(1ΔpdU1Δpd)Vα𝐤Vβ𝐤ei𝐤𝐑jei𝐤𝐑jp𝐤σαp𝐤σβdjσdjσ\displaystyle-\frac{1}{N}\sum_{j}\sum_{{\mathbf{k}}{\mathbf{k}}^{\prime}\sigma}\left(\frac{1}{\mathit{\Delta}_{pd}-U}-\frac{1}{\mathit{\Delta}_{pd}}\right)V_{\alpha{\mathbf{k}}}V_{\beta{\mathbf{k}}^{\prime}}^{*}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}e^{{\mathrm{i}}{\mathbf{k}}^{\prime}\cdot{\mathbf{R}}_{j}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}^{\prime}-\sigma}^{\beta}d_{j\,-\sigma}^{\dagger}d_{j\sigma}
1Nj𝐤𝐤σ(1ΔpdU1Δpd)Vα𝐤Vβ𝐤ei𝐤𝐑jei𝐤𝐑jp𝐤σαp𝐤σβdjσdjσ+H.c.\displaystyle-\frac{1}{N}\sum_{j}\sum_{{\mathbf{k}}{\mathbf{k}}^{\prime}\sigma}\left(\frac{1}{\mathit{\Delta}_{pd}-U}-\frac{1}{\mathit{\Delta}_{pd}}\right)V_{\alpha{\mathbf{k}}}V_{\beta{\mathbf{k}}^{\prime}}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}}e^{-{\mathrm{i}}{\mathbf{k}}^{\prime}\cdot{\mathbf{R}}_{j}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}^{\prime}-\sigma}^{\beta\dagger}d_{j\,-\sigma}d_{j\sigma}+{\mathrm{H.c.}}

Hereafter, I consider only the first two terms of the right-hand side of Eq. (3), i.e., up to the second order of tpdt_{pd}. Now, I assume that the distribution of the dd electron is spatially uniform in the ground state and that the ground state is paramagnetic. Thus, ndj+ndj0nd\langle n_{d\,j\uparrow}+n_{d\,j\downarrow}\rangle_{0}\equiv n_{d} and ndj0=ndj0\langle n_{d\,j\uparrow}\rangle_{0}=\langle n_{d\,j\downarrow}\rangle_{0} for any jj where ndn_{d} is a c-number equal to the number of dd electrons in the ground state, where 0\langle\cdots\rangle_{0} indicates the average in the ground state. I apply this approximation to Eqs. (3) and (LABEL:eq:06) and treat ndn_{d} as a parameter that should be determined self-consistently. When I set εp\varepsilon_{p} to zero, i.e., Δpd=εd\mathit{\Delta}_{pd}=-\varepsilon_{d}, and omit the constant terms, I obtain the effective Hamiltonian:

eff=HF+ex+pair+U.{\mathcal{H}}_{\mathrm{eff}}={\mathcal{H}}_{\mathrm{HF}}+{\mathcal{H}}_{\mathrm{ex}}+{\mathcal{H}}_{\mathrm{pair}}+{\mathcal{H}}_{U}^{\prime}. (6)

HF{\mathcal{H}}_{\mathrm{HF}} is the Hartree-Fock approximation of 0{\mathcal{H}}_{0}:

HF\displaystyle{\mathcal{H}}_{\mathrm{HF}} =\displaystyle= 𝐤σεd𝐤d𝐤σd𝐤σ+αβ𝐤σεαβ𝐤p𝐤σαp𝐤σβ,\displaystyle\sum_{{\mathbf{k}}\sigma}\varepsilon_{d{\mathbf{k}}}d_{{\mathbf{k}}\sigma}^{\dagger}d_{{\mathbf{k}}\sigma}+\sum_{\alpha\beta}\sum_{{\mathbf{k}}\sigma}\varepsilon_{\alpha\beta{\mathbf{k}}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}\sigma}^{\beta}, (7)

where d𝐤σ=1Njdjσei𝐤𝐑j,d𝐤σ=1Njdjσei𝐤𝐑jd_{{\mathbf{k}}\sigma}^{\dagger}=\frac{1}{\sqrt{N}}\sum_{j}d_{j\sigma}^{\dagger}e^{{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}},d_{{\mathbf{k}}\sigma}=\frac{1}{\sqrt{N}}\sum_{j}d_{j\sigma}e^{-{\mathrm{i}}{\mathbf{k}}\cdot{\mathbf{R}}_{j}},

εd𝐤=εd+U2nd+t(vx𝐤vx𝐤+vy𝐤vy𝐤),\varepsilon_{d{\mathbf{k}}}=\varepsilon_{d}+\frac{U}{2}n_{d}+t\left(v_{x{\mathbf{k}}}v_{x{\mathbf{k}}}^{*}+v_{y{\mathbf{k}}}v_{y{\mathbf{k}}}^{*}\right), (8)

and

εαβ𝐤=(Jndt)vα𝐤vβ𝐤,\varepsilon_{\alpha\beta{\mathbf{k}}}=\left(Jn_{d}-t\right)v_{\alpha{\mathbf{k}}}v_{\beta{\mathbf{k}}}^{*}, (9)

with vx𝐤=isinkx2v_{x{\mathbf{k}}}={\mathrm{i}}\sin\frac{k_{x}}{2}, vy𝐤=isinky2v_{y{\mathbf{k}}}=-{\mathrm{i}}\sin\frac{k_{y}}{2},

t=4tpd2(ndεd+U+1ndεd),t=4\,t_{pd}^{2}\left(\frac{n_{d}}{\varepsilon_{d}+U}+\frac{1-n_{d}}{\varepsilon_{d}}\right), (10)

and

J=2tpd2(1εd+U1εd).J=2\,t_{pd}^{2}\left(\frac{1}{\varepsilon_{d}+U}-\frac{1}{\varepsilon_{d}}\right). (11)

ex{\mathcal{H}}_{\mathrm{ex}} is an exchange interaction term:

ex=JNαβ𝐤𝐤σ𝐪vα𝐤vβ𝐤p𝐤σαp𝐤σβd𝐤+𝐪σd𝐤+𝐪σ+H.c.{\mathcal{H}}_{\mathrm{ex}}=\frac{J}{N}\sum_{\alpha\beta}\sum_{{\mathbf{k}}{\mathbf{k}}^{\prime}\sigma}\sum_{{\mathbf{q}}}v_{\alpha{\mathbf{k}}}v_{\beta{\mathbf{k}}^{\prime}}^{*}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}^{\prime}-\sigma}^{\beta}d_{{\mathbf{k}}^{\prime}+{\mathbf{q}}\,-\sigma}^{\dagger}d_{{\mathbf{k}}+{\mathbf{q}}\,\sigma}+{\mathrm{H.c.}} (12)

pair{\mathcal{H}}_{\mathrm{pair}} is a pair-hopping term:

pair=JNαβ𝐤𝐤σ𝐪vα𝐤vβ𝐤p𝐤σαp𝐤σβd𝐤𝐪σd𝐤+𝐪σ+H.c.{\mathcal{H}}_{\mathrm{pair}}=\frac{J}{N}\sum_{\alpha\beta}\sum_{{\mathbf{k}}{\mathbf{k}}^{\prime}\sigma}\sum_{{\mathbf{q}}}v_{\alpha{\mathbf{k}}}v_{\beta{\mathbf{k}}^{\prime}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}p_{{\mathbf{k}}^{\prime}-\sigma}^{\beta\dagger}d_{{\mathbf{k}}^{\prime}-{\mathbf{q}}\,-\sigma}d_{{\mathbf{k}}+{\mathbf{q}}\,\sigma}+{\mathrm{H.c.}} (13)

U{\mathcal{H}}_{U}^{\prime} is the Coulomb interaction term excluding the component with 𝐪=𝟎{\mathbf{q}}={\mathbf{0}}:

U=UN𝐤𝐤𝐪𝟎d𝐤+𝐪d𝐤d𝐤𝐪d𝐤.{\mathcal{H}}_{U}^{\prime}=\frac{U}{N}\sum_{{\mathbf{k}}{\mathbf{k}}^{\prime}}\sum_{{\mathbf{q}}\neq{\mathbf{0}}}d_{{\mathbf{k}}+{\mathbf{q}}\,\uparrow}^{\dagger}d_{{\mathbf{k}}\uparrow}d_{{\mathbf{k}}^{\prime}-{\mathbf{q}}\,\downarrow}^{\dagger}d_{{\mathbf{k}}^{\prime}\downarrow}. (14)

As a consequence, eff{\mathcal{H}}_{\mathrm{eff}} [Eq. (6)] can be characterized by the three parameters tt [Eq. (10)], JJ [Eq. (11)], and UU, and it can be regarded as the three-band ttJJUU model.

Here, tt in Eq. (10) is positive near the half-filling in the charge-transfer regime, i.e., U>εd>0U>-\varepsilon_{d}>0. For instance, in the case εd=U/2\varepsilon_{d}=-U/2, t>0t>0 for nd>0.5n_{d}>0.5, and the dd electron band dispersion εd𝐤\varepsilon_{d{\mathbf{k}}} in Eq. (8) is the same as that for the single-band Hubbard model on a square lattice. JJ in Eq. (11) is always positive in the charge-transfer regime. Thus, ex{\mathcal{H}}_{\mathrm{ex}} in Eq. (12) describes the transverse component of the antiferromagnetic exchange interaction between the dd and pp electrons, while the longitudinal component of this interaction narrows the bandwidth of εαβ𝐤\varepsilon_{\alpha\beta{\mathbf{k}}} in Eq. (9) from tt to tJndt-Jn_{d}. Further, ex{\mathcal{H}}_{\mathrm{ex}} indicates that the pp electron is affected by the spin fluctuation of the dd electron. As will be shown later, the dd-wave superconducting gap composed of dd and pp electrons emerges from ex{\mathcal{H}}_{\mathrm{ex}}. pair{\mathcal{H}}_{\mathrm{pair}} in Eq. (13) appears for the first time by considering the double occupancy of Cu sites. The pair-hopping term is not included in the single-band ttJJ model if double occupancy is excluded. In the model that includes the pair-hopping interaction, electrons favour pair formation [26]. This is also true in the presence of the on-site interaction [27, 28] and in the zero-bandwidth limit [29]. Thus, the pair-hopping term in my model is expected to provide superconductivity in another way.

I introduce another assumption according to the speculation about the ground state of the three-band Hubbard model [30]. In the normal ground state, the dd and pp electrons should be combined to construct coherent quasi-particles through hybridization. The matrix elements of the hybridization between the dd and pp electrons can be found in the components with 𝐪=𝟎{\mathbf{q}}={\mathbf{0}} in Eqs. (12) and (13) as follows. Defining

hpd=iNα𝐤[vα𝐤p𝐤αd𝐤0vα𝐤d𝐤p𝐤α0]=iNα𝐤[vα𝐤p𝐤αd𝐤0vα𝐤d𝐤p𝐤α0],\displaystyle h_{pd}=-\frac{{\mathrm{i}}}{N}\sum_{\alpha}\sum_{{\mathbf{k}}}\left[v_{\alpha{\mathbf{k}}}\langle p_{{\mathbf{k}}\uparrow}^{\alpha\dagger}d_{{\mathbf{k}}\uparrow}\rangle_{0}-v_{\alpha{\mathbf{k}}}^{*}\langle d_{{\mathbf{k}}\uparrow}^{\dagger}p_{{\mathbf{k}}\uparrow}^{\alpha}\rangle_{0}\right]=-\frac{{\mathrm{i}}}{N}\sum_{\alpha}\sum_{{\mathbf{k}}}\left[v_{\alpha{\mathbf{k}}}\langle p_{{\mathbf{k}}\downarrow}^{\alpha\dagger}d_{{\mathbf{k}}\downarrow}\rangle_{0}-v_{\alpha{\mathbf{k}}}^{*}\langle d_{{\mathbf{k}}\downarrow}^{\dagger}p_{{\mathbf{k}}\downarrow}^{\alpha}\rangle_{0}\right],
(15)

eff{\mathcal{H}}_{\mathrm{eff}} can be rewritten as

eff=0+ex+pair+U,{\mathcal{H}}_{\mathrm{eff}}={\mathcal{H}}_{0}^{\prime}+{\mathcal{H}}_{\mathrm{ex}}^{\prime}+{\mathcal{H}}_{\mathrm{pair}}^{\prime}+{\mathcal{H}}_{U}^{\prime}, (16)

where

0=HF+iJhpdα𝐤σ(vα𝐤p𝐤σαd𝐤σvα𝐤d𝐤σp𝐤σα).{\mathcal{H}}_{0}^{\prime}={\mathcal{H}}_{\mathrm{HF}}+{\mathrm{i}}Jh_{pd}\sum_{\alpha}\sum_{{\mathbf{k}}\sigma}\left(v_{\alpha{\mathbf{k}}}p_{{\mathbf{k}}\sigma}^{\alpha\dagger}d_{{\mathbf{k}}\sigma}-v_{\alpha{\mathbf{k}}}^{*}d_{{\mathbf{k}}\sigma}^{\dagger}p_{{\mathbf{k}}\sigma}^{\alpha}\right). (17)

Here, A0\langle A\rangle_{0} in Eq. (15) means the expectation value of AA in the ground state of 0{\mathcal{H}}_{0}^{\prime}. ex{\mathcal{H}}_{\mathrm{ex}}^{\prime} and pair{\mathcal{H}}_{\mathrm{pair}}^{\prime} indicate the exchange interaction and pair-hopping terms excluding the component with 𝐪=𝟎{\mathbf{q}}={\mathbf{0}} from Eqs. (12) and (13), respectively. Thus, in the ground state of 0{\mathcal{H}}_{0}^{\prime}, the dd and pp electrons are combined to construct the coherent quasi-particles when hpd>0h_{pd}>0.

Hereafter, I treat 0{\mathcal{H}}_{0}^{\prime} as the unperturbed part of eff{\mathcal{H}}_{\mathrm{eff}} on the assumption that hpd>0h_{pd}>0. I diagonalize 0{\mathcal{H}}_{0}^{\prime} and derive the unperturbed Green functions as follows:

Gdd0(𝐤,iϵn)=iϵn+μεxx𝐤εyy𝐤(iϵn+με𝐤+)(iϵn+με𝐤),\displaystyle G_{dd}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})=\frac{{\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{xx{\mathbf{k}}}-\varepsilon_{yy{\mathbf{k}}}}{({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{+})({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{-})}, (18)
Gdα0(𝐤,iϵn)=iJhpdvα𝐤(iϵn+με𝐤+)(iϵn+με𝐤),\displaystyle G_{d\alpha}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})=\frac{{\mathrm{i}}Jh_{pd}v_{\alpha\mathbf{k}}}{({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{+})({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{-})}, (19)
Gαd0(𝐤,iϵn)=iJhpdvα𝐤(iϵn+με𝐤+)(iϵn+με𝐤),\displaystyle G_{\alpha d}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})=\frac{-{\mathrm{i}}Jh_{pd}v_{\alpha\mathbf{k}}^{*}}{({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{+})({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{-})}, (20)

and

(Gxx0(𝐤,iϵn)Gxy0(𝐤,iϵn)Gyx0(𝐤,iϵn)Gyy0(𝐤,iϵn))=1(iϵn+μ)(iϵn+με𝐤+)(iϵn+με𝐤)\displaystyle\left(\begin{array}[]{cc}G_{xx}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})&G_{xy}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})\\ G_{yx}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})&G_{yy}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n})\\ \end{array}\right)=\frac{1}{({\mathrm{i}}\epsilon_{n}+\mu)({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{+})({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{{\mathbf{k}}}^{-})} (23)
×((iϵn+μεd𝐤)(iϵn+μεyy𝐤)J2hpd2vy𝐤vy𝐤(iϵn+μεd𝐤)εyx𝐤+J2hpd2vy𝐤vx𝐤(iϵn+μεd𝐤)εxy𝐤+J2hpd2vx𝐤vy𝐤(iϵn+μεd𝐤)(iϵn+μεxx𝐤)J2hpd2vx𝐤vx𝐤).\displaystyle\times\left(\begin{array}[]{cc}({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{d{\mathbf{k}}})({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{yy{\mathbf{k}}})-J^{2}h_{pd}^{2}v_{y{\mathbf{k}}}v_{y{\mathbf{k}}}^{*}&({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{d{\mathbf{k}}})\varepsilon_{yx{\mathbf{k}}}+J^{2}h_{pd}^{2}v_{y{\mathbf{k}}}v_{x{\mathbf{k}}}^{*}\\ ({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{d{\mathbf{k}}})\varepsilon_{xy{\mathbf{k}}}+J^{2}h_{pd}^{2}v_{x{\mathbf{k}}}v_{y{\mathbf{k}}}^{*}&({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{d{\mathbf{k}}})({\mathrm{i}}\epsilon_{n}+\mu-\varepsilon_{xx{\mathbf{k}}})-J^{2}h_{pd}^{2}v_{x{\mathbf{k}}}v_{x{\mathbf{k}}}^{*}\\ \end{array}\right). (26)

Here, I use the fermion Matsubara frequencies, ϵn=πT(2n+1)\epsilon_{n}=\pi T(2n+1), with integer nn and temperature TT. μ\mu is the chemical potential and

ε𝐤±=εd𝐤+εxx𝐤+εyy𝐤2±(εd𝐤εxx𝐤εyy𝐤2)2+J2hpd2(vx𝐤vx𝐤+vy𝐤vy𝐤).\varepsilon_{{\mathbf{k}}}^{\pm}=\frac{\varepsilon_{d{\mathbf{k}}}+\varepsilon_{xx{\mathbf{k}}}+\varepsilon_{yy{\mathbf{k}}}}{2}\pm\sqrt{\left(\frac{\varepsilon_{d{\mathbf{k}}}-\varepsilon_{xx{\mathbf{k}}}-\varepsilon_{yy{\mathbf{k}}}}{2}\right)^{2}+J^{2}h_{pd}^{2}\left(v_{x{\mathbf{k}}}v_{x{\mathbf{k}}}^{*}+v_{y{\mathbf{k}}}v_{y{\mathbf{k}}}^{*}\right)}. (28)

For hpd>0h_{pd}>0, Eq. (15) can be rewritten as

1=2JN𝐤vx𝐤vx𝐤+vy𝐤vy𝐤ε𝐤+ε𝐤{θ(ε𝐤+μ)θ(ε𝐤μ)},1=\frac{2J}{N}\sum_{{\mathbf{k}}}\frac{v_{x{\mathbf{k}}}v_{x{\mathbf{k}}}^{*}+v_{y{\mathbf{k}}}v_{y{\mathbf{k}}}^{*}}{\varepsilon_{{\mathbf{k}}}^{+}-\varepsilon_{{\mathbf{k}}}^{-}}\,\left\{\theta(\varepsilon_{{\mathbf{k}}}^{+}-\mu)-\theta(\varepsilon_{{\mathbf{k}}}^{-}-\mu)\right\}, (29)

where θ(x)\theta(x) means the Heaviside step function.

In order to investigate the superconductivity in a strong coupling framework, I start with the Dyson-Gor’kov equations:

Gμν(𝐤,iϵn)=Gμν0(𝐤,iϵn)+Gμκ0(𝐤,iϵn)Σκλ(𝐤,iϵn)Gλν(𝐤,iϵn)+Gμκ0(𝐤,iϵn)Φκλ(𝐤,iϵn)Fλν(𝐤,iϵn),\displaystyle G_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})=G_{\mu\nu}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n})+G_{\mu\kappa}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}){\mathit{\Sigma}}_{\kappa\lambda}({\mathbf{k}},\mathrm{i}\epsilon_{n})G_{\lambda\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})+G_{\mu\kappa}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}){\mathit{\Phi}}_{\kappa\lambda}({\mathbf{k}},\mathrm{i}\epsilon_{n})F_{\lambda\nu}^{\dagger}(-{\mathbf{k}},-\mathrm{i}\epsilon_{n}),
(30)
Fμν(𝐤,iϵn)=Gμκ0(𝐤,iϵn)Σκλ(𝐤,iϵn)Fλν(𝐤,iϵn)+Gμκ0(𝐤,iϵn)Φκλ(𝐤,iϵn)Gλν(𝐤,iϵn),\displaystyle F_{\mu\nu}^{\dagger}({\mathbf{k}},\mathrm{i}\epsilon_{n})=G_{\mu\kappa}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}){\mathit{\Sigma}}_{\kappa\lambda}({\mathbf{k}},\mathrm{i}\epsilon_{n})F_{\lambda\nu}^{\dagger}({\mathbf{k}},\mathrm{i}\epsilon_{n})+G_{\mu\kappa}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}){\mathit{\Phi}}_{\kappa\lambda}^{*}({\mathbf{k}},\mathrm{i}\epsilon_{n})G_{\lambda\nu}(-{\mathbf{k}},-\mathrm{i}\epsilon_{n}), (31)
Fμν(𝐤,iϵn)=Gμκ0(𝐤,iϵn)Σκλ(𝐤,iϵn)Fλν(𝐤,iϵn)+Gμκ0(𝐤,iϵn)Φκλ(𝐤,iϵn)Gλν(𝐤,iϵn).\displaystyle F_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})=G_{\mu\kappa}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}){\mathit{\Sigma}}_{\kappa\lambda}({\mathbf{k}},\mathrm{i}\epsilon_{n})F_{\lambda\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})+G_{\mu\kappa}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}){\mathit{\Phi}}_{\kappa\lambda}({\mathbf{k}},\mathrm{i}\epsilon_{n})G_{\lambda\nu}(-{\mathbf{k}},-\mathrm{i}\epsilon_{n}). (32)

The orbital indices μ\mu, ν\nu, κ\kappa, and λ\lambda run over dd, xx, and yy, and I adopt the Einstein summation convention. Gμν(𝐤,iϵn)G_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) and Fμν(𝐤,iϵn)F_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) represent the normal and anomalous Green functions, respectively, and Σμν(𝐤,iϵn){\mathit{\Sigma}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) and Φμν(𝐤,iϵn){\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) correspond to the normal and anomalous self-energies, respectively. When ex+pair+U{\mathcal{H}}_{\mathrm{ex}}^{\prime}+{\mathcal{H}}_{\mathrm{pair}}^{\prime}+{\mathcal{H}}_{U}^{\prime} in Eq. (16) is treated as a perturbation, the normal self-energies up to the second order of JJ and UU are evaluated by the IPT approximation as follows:

Σdd(𝐤,iϵn)=TN𝐤n[J2χJG(𝐤𝐤,iϵniϵn)Gpp0(𝐤,iϵn)+U2χU(𝐤𝐤,iϵniϵn)Gdd0(𝐤,iϵn)],\displaystyle{\mathit{\Sigma}}_{dd}(\mathbf{k},\mathrm{i}\epsilon_{n})=\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[J^{2}\chi_{J}^{G}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})G_{pp}^{0}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})+U^{2}\chi_{U}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})G_{dd}^{0}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right], (36)
Σdα(𝐤,iϵn)=vα𝐤TN𝐤n[J2χJG(𝐤𝐤,iϵniϵn)Gpd0(𝐤,iϵn)],\displaystyle{\mathit{\Sigma}}_{d\alpha}(\mathbf{k},\mathrm{i}\epsilon_{n})=-v_{\alpha\mathbf{k}}^{*}\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[J^{2}\chi_{J}^{G}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})G_{pd}^{0}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right],
Σαd(𝐤,iϵn)=vα𝐤TN𝐤n[J2χJG(𝐤𝐤,iϵniϵn)Gdp0(𝐤,iϵn)],\displaystyle{\mathit{\Sigma}}_{\alpha d}(\mathbf{k},\mathrm{i}\epsilon_{n})=-v_{\alpha\mathbf{k}}\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[J^{2}\chi_{J}^{G}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})G_{dp}^{0}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right],
Σαβ(𝐤,iϵn)=vα𝐤vβ𝐤TN𝐤n[J2χJG(𝐤𝐤,iϵniϵn)Gdd0(𝐤,iϵn)].\displaystyle{\mathit{\Sigma}}_{\alpha\beta}(\mathbf{k},\mathrm{i}\epsilon_{n})=v_{\alpha\mathbf{k}}v_{\beta\mathbf{k}}^{*}\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[J^{2}\chi_{J}^{G}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})G_{dd}^{0}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right].

The IPT approximation was first applied in the study of the half-filled single-impurity Anderson model [31, 32], and it was adopted to solve the effective impurity model in the study of the d=d=\infty Hubbard model [33, 34]. In these works, it was shown that the second order perturbation theory in large energy scale UU could reproduce not only the coherent band but also the lower and upper incoherent bands. In a later section, it will be shown that my approach can reproduce similar band structure to be justified as the theory for the 2D three-band ttJJUU model.

The anomalous self-energies up to the second order of JJ and UU are evaluated as follows:

Φdd(𝐤,iϵn)=TN𝐤n[{J+J2χJF(𝐤𝐤,iϵniϵn)}Fpp(𝐤,iϵn)+{U+U2χU(𝐤𝐤,iϵniϵn)}Fdd(𝐤,iϵn)],\displaystyle{\mathit{\Phi}}_{dd}(\mathbf{k},\mathrm{i}\epsilon_{n})=-\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[\left\{J+J^{2}\chi_{J}^{F}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})\right\}\!F_{pp}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})+\left\{U+U^{2}\chi_{U}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})\right\}\!F_{dd}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right],
(37)
Φdα(𝐤,iϵn)=vα𝐤TN𝐤n[{J+J2χJF(𝐤𝐤,iϵniϵn)}Fpd(𝐤,iϵn)],\displaystyle{\mathit{\Phi}}_{d\alpha}(\mathbf{k},\mathrm{i}\epsilon_{n})=v_{\alpha-\mathbf{k}}\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[\left\{J+J^{2}\chi_{J}^{F}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})\right\}\!F_{pd}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right], (38)
Φαd(𝐤,iϵn)=vα𝐤TN𝐤n[{J+J2χJF(𝐤𝐤,iϵniϵn)}Fdp(𝐤,iϵn)],\displaystyle{\mathit{\Phi}}_{\alpha d}(\mathbf{k},\mathrm{i}\epsilon_{n})=v_{\alpha\mathbf{k}}\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[\left\{J+J^{2}\chi_{J}^{F}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})\right\}\!F_{dp}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right], (39)
Φαβ(𝐤,iϵn)=vα𝐤vβ𝐤TN𝐤n[{J+J2χJF(𝐤𝐤,iϵniϵn)}Fdd(𝐤,iϵn)].\displaystyle{\mathit{\Phi}}_{\alpha\beta}(\mathbf{k},\mathrm{i}\epsilon_{n})=-v_{\alpha\mathbf{k}}v_{\beta-\mathbf{k}}\frac{T}{N}\sum_{{\mathbf{k}}^{\prime}n^{\prime}}\left[\left\{J+J^{2}\chi_{J}^{F}({\mathbf{k}}-{\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n}-\mathrm{i}\epsilon_{n^{\prime}})\right\}\!F_{dd}({\mathbf{k}}^{\prime},\mathrm{i}\epsilon_{n^{\prime}})\right]. (40)

Here, the orbital indices α\alpha and β\beta run over xx and yy, and

χJG(𝐪,iωm)=χdd,ppG(𝐪,iωm)χdp,dpG(𝐪,iωm)χpd,pdG(𝐪,iωm)+χpp,ddG(𝐪,iωm),\displaystyle\chi_{J}^{G}({\mathbf{q}},\mathrm{i}\omega_{m})=\chi_{dd,pp}^{G}({\mathbf{q}},\mathrm{i}\omega_{m})-\chi_{dp,dp}^{G}({\mathbf{q}},\mathrm{i}\omega_{m})-\chi_{pd,pd}^{G}({\mathbf{q}},\mathrm{i}\omega_{m})+\chi_{pp,dd}^{G}({\mathbf{q}},\mathrm{i}\omega_{m}), (41)
χJF(𝐪,iωm)=χdd,ppF(𝐪,iωm)χdp,pdF(𝐪,iωm)χpd,dpF(𝐪,iωm)+χpp,ddF(𝐪,iωm),\displaystyle\chi_{J}^{F}({\mathbf{q}},\mathrm{i}\omega_{m})=\chi_{dd,pp}^{F}({\mathbf{q}},\mathrm{i}\omega_{m})-\chi_{dp,pd}^{F}({\mathbf{q}},\mathrm{i}\omega_{m})-\chi_{pd,dp}^{F}({\mathbf{q}},\mathrm{i}\omega_{m})+\chi_{pp,dd}^{F}({\mathbf{q}},\mathrm{i}\omega_{m}), (42)
χU(𝐪,iωm)=χdd,ddG(𝐪,iωm)+χdd,ddF(𝐪,iωm),\displaystyle\chi_{U}({\mathbf{q}},\mathrm{i}\omega_{m})=\chi_{dd,dd}^{G}({\mathbf{q}},\mathrm{i}\omega_{m})+\chi_{dd,dd}^{F}({\mathbf{q}},\mathrm{i}\omega_{m}), (43)
χμν,κλG(𝐪,iωm)=TN𝐤nGμν0(𝐪+𝐤,iωm+iϵn)Gκλ0(𝐤,iϵn),\displaystyle\chi_{\mu\nu,\kappa\lambda}^{G}({\mathbf{q}},\mathrm{i}\omega_{m})=-\frac{T}{N}\sum_{{\mathbf{k}}n}G_{\mu\nu}^{0}({\mathbf{q}}+{\mathbf{k}},\mathrm{i}\omega_{m}+\mathrm{i}\epsilon_{n})G_{\kappa\lambda}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n}), (44)
χμν,κλF(𝐪,iωm)=TN𝐤nFμν(𝐪+𝐤,iωm+iϵn)Fκλ(𝐤,iϵn),\displaystyle\chi_{\mu\nu,\kappa\lambda}^{F}({\mathbf{q}},\mathrm{i}\omega_{m})=-\frac{T}{N}\sum_{{\mathbf{k}}n}F_{\mu\nu}({\mathbf{q}}+{\mathbf{k}},\mathrm{i}\omega_{m}+\mathrm{i}\epsilon_{n})F_{\kappa\lambda}^{\dagger}({\mathbf{k}},\mathrm{i}\epsilon_{n}), (45)
Gpp0(𝐤,iϵn)=αβvα𝐤vβ𝐤Gαβ0(𝐤,iϵn),\displaystyle G_{pp}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha\beta}v_{\alpha\mathbf{k}}^{*}v_{\beta\mathbf{k}}G_{\alpha\beta}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n}), (46)
Gdp0(𝐤,iϵn)=αvα𝐤Gdα0(𝐤,iϵn),\displaystyle G_{dp}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha}v_{\alpha\mathbf{k}}G_{d\alpha}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n}), (47)
Gpd0(𝐤,iϵn)=αvα𝐤Gαd0(𝐤,iϵn),\displaystyle G_{pd}^{0}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha}v_{\alpha\mathbf{k}}^{*}G_{\alpha d}^{0}(\mathbf{k},\mathrm{i}\epsilon_{n}), (48)
Fpp(𝐤,iϵn)=αβvα𝐤vβ𝐤Fαβ(𝐤,iϵn),\displaystyle F_{pp}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha\beta}v_{\alpha\mathbf{k}}^{*}v_{\beta-\mathbf{k}}^{*}F_{\alpha\beta}(\mathbf{k},\mathrm{i}\epsilon_{n}), (49)
Fdp(𝐤,iϵn)=αvα𝐤Fdα(𝐤,iϵn),\displaystyle F_{dp}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha}v_{\alpha-\mathbf{k}}^{*}F_{d\alpha}(\mathbf{k},\mathrm{i}\epsilon_{n}), (50)
Fpd(𝐤,iϵn)=αvα𝐤Fαd(𝐤,iϵn),\displaystyle F_{pd}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha}v_{\alpha\mathbf{k}}^{*}F_{\alpha d}(\mathbf{k},\mathrm{i}\epsilon_{n}), (51)
Fpp(𝐤,iϵn)=αβvα𝐤vβ𝐤Fαβ(𝐤,iϵn),\displaystyle F_{pp}^{\dagger}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha\beta}v_{\alpha\mathbf{k}}v_{\beta-\mathbf{k}}F_{\alpha\beta}^{\dagger}(\mathbf{k},\mathrm{i}\epsilon_{n}), (52)
Fdp(𝐤,iϵn)=αvα𝐤Fdα(𝐤,iϵn),and\displaystyle F_{dp}^{\dagger}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha}v_{\alpha-\mathbf{k}}F_{d\alpha}^{\dagger}(\mathbf{k},\mathrm{i}\epsilon_{n}),{\mathrm{and}} (53)
Fpd(𝐤,iϵn)=αvα𝐤Fαd(𝐤,iϵn),\displaystyle F_{pd}^{\dagger}({\mathbf{k}},\mathrm{i}\epsilon_{n})=\sum_{\alpha}v_{\alpha\mathbf{k}}F_{\alpha d}^{\dagger}(\mathbf{k},\mathrm{i}\epsilon_{n}), (54)

using the boson Matsubara frequencies, ωm=2mπT\omega_{m}=2m\pi T with integer mm. In Eqs. (44) and (45), μ\mu, ν\nu, κ\kappa, and λ\lambda denote dd or pp, respectively. Note that ndn_{d}, tt, hpdh_{pd}, and the chemical potential μ\mu must be determined self-consistently in the ground state of 0{\mathcal{H}}_{0}^{\prime} through Eqs. (8)–(10), (28), and (29). To this end, I approximate ndn_{d} by the number of dd electrons in the ground state of 0{\mathcal{H}}_{0}^{\prime}:

nd=22N𝐤1ε𝐤+ε𝐤{(ε𝐤+εxx𝐤εyy𝐤)θ(ε𝐤+μ)(ε𝐤εxx𝐤εyy𝐤)θ(ε𝐤μ)}.n_{d}=2-\frac{2}{N}\sum_{\mathbf{k}}\frac{1}{\varepsilon_{{\mathbf{k}}}^{+}-\varepsilon_{{\mathbf{k}}}^{-}}\left\{(\varepsilon_{{\mathbf{k}}}^{+}-\varepsilon_{xx{\mathbf{k}}}-\varepsilon_{yy{\mathbf{k}}})\theta(\varepsilon_{{\mathbf{k}}}^{+}-\mu)-(\varepsilon_{{\mathbf{k}}}^{-}-\varepsilon_{xx{\mathbf{k}}}-\varepsilon_{yy{\mathbf{k}}})\theta(\varepsilon_{{\mathbf{k}}}^{-}-\mu)\right\}. (55)

Specifically, I regard ndn_{d} as a given parameter and solve Eqs. (8)–(10), (28), (29), and (55) to determine tt, hpdh_{pd}, and the number of doped holes δh0\delta_{\mathrm{h}}^{0} for the ground state of 0{\mathcal{H}}_{0}^{\prime}, where

δh0=2N𝐤[θ(ε𝐤+μ)+θ(ε𝐤μ)+θ(μ)]1.\delta_{\mathrm{h}}^{0}=\frac{2}{N}\sum_{\mathbf{k}}\left[\theta(\varepsilon_{{\mathbf{k}}}^{+}-\mu)+\theta(\varepsilon_{{\mathbf{k}}}^{-}-\mu)+\theta(-\mu)\right]-1.

Once tt, hpdh_{pd}, and δh0\delta_{\mathrm{h}}^{0} are determined for the ground state of 0{\mathcal{H}}_{0}^{\prime}, I treat tt, hpdh_{pd}, and δh0\delta_{\mathrm{h}}^{0} as temperature independent parameters, whose values do not change from those at T=0T=0. Then, Eqs. (8)–(54) are solved in a fully self-consistent manner to obtain Σμν(𝐤,iϵn){\mathit{\Sigma}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) and Φμν(𝐤,iϵn){\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}). To determine the transition temperature TcT_{\mathrm{c}}, I perform these calculations in two steps. First, Σμν(𝐤,iϵn){\mathit{\Sigma}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) is calculated with Φμν(𝐤,iϵn)=0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})=0, and μ\mu is self-consistently determined so that δh\delta_{\mathrm{h}} obtained from Gμν(𝐤,iϵn)G_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) becomes equal to δh0\delta_{\mathrm{h}}^{0}. In the first step, μ\mu is correctly adjusted to compensate the temperature-dependent shift by Σμν(𝐤,iϵn){\mathit{\Sigma}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) with Φμν(𝐤,iϵn)=0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})=0. Here, δh=ndh+nph1\delta_{\mathrm{h}}=n_{d\mathrm{h}}+n_{p\mathrm{h}}-1, where

ndh\displaystyle n_{d\mathrm{h}} =\displaystyle= 22TN𝐤nGdd(𝐤,iϵn)eiϵn0+,\displaystyle 2-2\,\frac{T}{N}\sum_{{\mathbf{k}}n}G_{dd}({\mathbf{k}},\mathrm{i}\epsilon_{n})e^{\mathrm{i}\epsilon_{n}{0^{+}}}, (56)
nph\displaystyle n_{p\mathrm{h}} =\displaystyle= 42TN𝐤n[Gxx(𝐤,iϵn)+Gyy(𝐤,iϵn)]eiϵn0+,\displaystyle 4-2\,\frac{T}{N}\sum_{{\mathbf{k}}n}\left[G_{xx}({\mathbf{k}},\mathrm{i}\epsilon_{n})+G_{yy}({\mathbf{k}},\mathrm{i}\epsilon_{n})\right]e^{\mathrm{i}\epsilon_{n}{0^{+}}}, (57)

and ndhn_{d\mathrm{h}} and nphn_{p\mathrm{h}} are the number of dd and pp holes, respectively. Next, using the determined μ\mu, fully self-consistent calculations are performed to obtain Σμν(𝐤,iϵn){\mathit{\Sigma}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) and Φμν(𝐤,iϵn){\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}). At this time, only the temperature-dependent shift by Φμν(𝐤,iϵn){\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) is reflected in δh\delta_{\mathrm{h}} obtained from Gμν(𝐤,iϵn)G_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}). That is, if δh\delta_{\mathrm{h}} deviates from δh0\delta_{\mathrm{h}}^{0}, Φμν(𝐤,iϵn)0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})\neq 0. Therefore, the temperature at which δh\delta_{\mathrm{h}} deviates from δh0\delta_{\mathrm{h}}^{0} is TcT_{\mathrm{c}}. Also in the second step, μ\mu can be self-consistently determined so that δh\delta_{\mathrm{h}} obtained from Gμν(𝐤,iϵn)G_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) becomes equal to δh0\delta_{\mathrm{h}}^{0} with Φμν(𝐤,iϵn)0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})\neq 0. In this case, the temperature at which μ\mu deviates from the value with Φμν(𝐤,iϵn)=0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})=0 is TcT_{\mathrm{c}}, which is consistent with the temperature at which δh\delta_{\mathrm{h}} deviates from δh0\delta_{\mathrm{h}}^{0} with fixed μ\mu.

3 Results and discussion

To perform the numerical calculations, I divide the FBZ into a 64×6464\times 64 meshes and prepare 20482048 or 40964096 Matsubara frequencies. I commonly use tpd=10000Kt_{pd}=10000\,{\mathrm{K}} for my calculations, and here, I only consider the case εd=U/2\varepsilon_{d}=-U/2. For this case, we have J=tpdJ=t_{pd} when U=8tpdU=8\,t_{pd}. I find fully self-consistent solutions with hpd>0h_{pd}>0 in δh0.117\delta_{\mathrm{h}}\geq 0.117. The ones in 0.117δh0.1390.117\leq\delta_{\mathrm{h}}\leq 0.139 have Φμν(𝐤,iϵn)=0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})=0 and the others in 0.166δh0.2850.166\leq\delta_{\mathrm{h}}\leq 0.285 have Φμν(𝐤,iϵn)0{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n})\neq 0. The former solutions correspond to metallic phase and the latter to superconducting phase. Although I find other fully self-consistent solutions with hpd=0h_{pd}=0 in δh0.031\delta_{\mathrm{h}}\leq 0.031, which correspond to insulating phase, I cannot find any solutions in 0.031<δh<0.1170.031<\delta_{\mathrm{h}}<0.117. The absence of solutions in this doping range indicates that some of my assumptions break down. In particular, it is difficult to achieve the spatially uniform distribution of the dd electron in this range. For instance, the chemical potential shift suppression is observed in La2-xSrxCuO4 (0<x<0.120<x<0.12) by photoemission spectroscopy [35, 36]. This suppression suggests the possibility of electronic phase separation between the insulating phase and the superconducting phase [37], where the electrons are inhomogeneously distributed due to the strong electron correlation. Therefore, the theory in 0.031<δh<0.1170.031<\delta_{\mathrm{h}}<0.117 should consider the possibility of the spatially non-uniform distribution of the dd electron.

02040608010012014016018000.050.050.10.10.150.150.20.20.250.250.30.300.20.40.60.81.0IMSRefer to caption

Tc(K)T_{\mathrm{c}}\,({\mathrm{K}})

hpdh_{pd}

δh\delta_{\mathrm{h}}TcT_{\mathrm{c}}Tc0T_{\mathrm{c}}^{0}hpdh_{pd}
Figure 1: Doping dependences of TcT_{\mathrm{c}} and hpdh_{pd}. "I", "M" and "S" indicate insulating, metallic and superconducting phases, respectively. The shaded region indicates 0.031<δh<0.1170.031<\delta_{\mathrm{h}}<0.117 in which any solutions cannot be found. Tc0T_{\mathrm{c}}^{0} is the temperature at which the divergence of the Cooper susceptibility occurs. δh\delta_{\mathrm{h}} for Tc0T_{\mathrm{c}}^{0} and hpdh_{pd} are evaluated at T=170KT=170\,\mathrm{K}.

Figure 1 summarizes these results with the doping dependences of TcT_{\mathrm{c}} and hpdh_{pd}. Comparing TcT_{\mathrm{c}} with Tc0T_{\mathrm{c}}^{0}, at which the divergence of the Cooper susceptibility occurs, TcT_{\mathrm{c}} is higher than Tc0T_{\mathrm{c}}^{0} by 1014K10\sim 14\,\mathrm{K} since TcT_{\mathrm{c}} reflects the fluctuation of Φμν(𝐤,iϵn){\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}). While hpdh_{pd} increases monotonically with δh\delta_{\mathrm{h}}, TcT_{\mathrm{c}} reaches its maximum, 157K157\,\mathrm{K}, at δh=0.209\delta_{\mathrm{h}}=0.209 and then decreases. This doping dependence of TcT_{\mathrm{c}} reproduces the dome-shaped superconducting phase that is typical for the hole-doped HTSC [38, 39]. This behavior is related to the doping dependence of the density of states, and it will be explained later.

Figure 2 shows the temperature dependences of δhδh0\delta_{\mathrm{h}}-\delta_{\mathrm{h}}^{0} for every δh0\delta_{\mathrm{h}}^{0}, which are used to determine TcT_{\mathrm{c}}. Here, I define the temperature at which δhδh0\delta_{\mathrm{h}}-\delta_{\mathrm{h}}^{0} jumps as TcT_{\mathrm{c}}. The jumps of δhδh0\delta_{\mathrm{h}}-\delta_{\mathrm{h}}^{0} at TcT_{\mathrm{c}} in the underdoped regime, δh00.190\delta_{\mathrm{h}}^{0}\leq 0.190 [Fig. 2(a)], are larger than those in the overdoped regime, δh00.205\delta_{\mathrm{h}}^{0}\geq 0.205 [Fig. 2(b)]. In other words, while strong coupling superconductivity is established in the underdoped regime, the superconductivity in the overdoped regime remains with weak coupling. This tendency must be reflected in the superconducting gap magnitude, which has been shown to decrease with doping by the low-temperature specific heats of La2-xSrxCuO4 [40, 41].

00.020.020.040.040.060.060.080.080.10.100.020.020.040.040.060.060.080.080.10.100.020.020.040.040.060.060.080.089010011012013014015016017000.020.020.040.040.060.060.080.08Refer to caption

δhδh0\delta_{\mathrm{h}}-\delta_{\mathrm{h}}^{0}

\makebox(0.0,0.0)[]{{} }

(a)δh0=0.166\delta_{\mathrm{h}}^{0}=0.166δh0=0.170\delta_{\mathrm{h}}^{0}=0.170δh0=0.178\delta_{\mathrm{h}}^{0}=0.178δh0=0.182\delta_{\mathrm{h}}^{0}=0.182δh0=0.190\delta_{\mathrm{h}}^{0}=0.190

δhδh0\delta_{\mathrm{h}}-\delta_{\mathrm{h}}^{0}

\makebox(0.0,0.0)[]{{} }

T(K)T\,({\mathrm{K}})(b)δh0=0.205\delta_{\mathrm{h}}^{0}=0.205δh0=0.221\delta_{\mathrm{h}}^{0}=0.221δh0=0.236\delta_{\mathrm{h}}^{0}=0.236δh0=0.248\delta_{\mathrm{h}}^{0}=0.248δh0=0.266\delta_{\mathrm{h}}^{0}=0.266
Figure 2: Temperature dependences of δhδh0\delta_{\mathrm{h}}-\delta_{\mathrm{h}}^{0}: (a) δh00.190\delta_{\mathrm{h}}^{0}\leq 0.190 and (b) δh00.205\delta_{\mathrm{h}}^{0}\geq 0.205.

The electronic states of the obtained solutions are reconstructed from the unperturbed ground state. Figure 3 shows the doping dependences of ndhn_{d\mathrm{h}}, nphn_{p\mathrm{h}}, ndh0n_{d\mathrm{h}}^{0}, and nph0n_{p\mathrm{h}}^{0} at T=170KT=170\,\mathrm{K}. Here, ndh0=2ndn_{d\mathrm{h}}^{0}=2-n_{d} and nph0=δh0+nd1n_{p\mathrm{h}}^{0}=\delta_{\mathrm{h}}^{0}+n_{d}-1, and ndh0n_{d\mathrm{h}}^{0} and nph0n_{p\mathrm{h}}^{0} are the numbers of dd and pp holes in the unperturbed ground state, respectively.

00.20.20.40.40.60.60.80.8111.21.20.10.10.150.150.20.20.250.250.30.300.20.20.40.40.60.60.80.8111.21.2Refer to caption

Number of holes

\makebox(0.0,0.0)[]{{} }

δh\delta_{\mathrm{h}}ndhn_{d{\mathrm{h}}}nphn_{p{\mathrm{h}}}ndh0n_{d{\mathrm{h}}}^{0}nph0n_{p{\mathrm{h}}}^{0}
Figure 3: Doping dependences of ndhn_{d\mathrm{h}}, nphn_{p\mathrm{h}}, ndh0n_{d\mathrm{h}}^{0}, and nph0n_{p\mathrm{h}}^{0} at T=170KT=170\,\mathrm{K}.

As shown in Fig. 3, holes are transferred from the dd band to the pp band due to the charge fluctuations χJG(𝐪,iωm)\chi_{J}^{G}({\mathbf{q}},\mathrm{i}\omega_{m}) via the normal self-energies Σμν(𝐤,iϵn){\mathit{\Sigma}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) in Eqs. (36)–(36). As a consequence, while nphn_{p\mathrm{h}} mainly increases with δh\delta_{\mathrm{h}}, ndh<1n_{d\mathrm{h}}<1, which means that the dd band is always electron doped. Since the dd band deviates from the half-filling due to the charge fluctuations, there is room for the pair-hopping interaction in Eq. (13) to work effectively between the pp and dd electrons despite the strong correlations among dd electrons. Later, I show how the pair-hopping interaction works for the superconductivity in the analysis of the superconducting gap function.

Figures 4 and 5 show the doping dependences of ρd(ε)\rho_{d}(\varepsilon) and ρp(ε)\rho_{p}(\varepsilon) at T=170KT=170\,\mathrm{K}, which elucidate how the dome-shaped superconducting phase develops.

00.20.20.40.40.60.60.80.8111.21.21.41.40246810121400.20.20.40.40.60.60.80.8111.21.202468101200.20.20.40.40.60.60.80.8111.21.202468101200.20.20.40.40.60.60.80.8111.21.202468101200.20.20.40.40.60.60.80.8111.21.25-54-43-32-21-101122334455024681012Refer to caption

ρd(ε)\rho_{d}(\varepsilon)(arbitrary unit)

ρp(ε)\rho_{p}(\varepsilon)(arbitrary unit)

(a) δh=0.123\delta_{\mathrm{h}}=0.123ρd(ε)\rho_{d}(\varepsilon)ρp(ε)\rho_{p}(\varepsilon)

ρd(ε)\rho_{d}(\varepsilon)(arbitrary unit)

ρp(ε)\rho_{p}(\varepsilon)(arbitrary unit)

(b) δh=0.166\delta_{\mathrm{h}}=0.166ρd(ε)\rho_{d}(\varepsilon)ρp(ε)\rho_{p}(\varepsilon)

ρd(ε)\rho_{d}(\varepsilon)(arbitrary unit)

ρp(ε)\rho_{p}(\varepsilon)(arbitrary unit)

(c) δh=0.205\delta_{\mathrm{h}}=0.205ρd(ε)\rho_{d}(\varepsilon)ρp(ε)\rho_{p}(\varepsilon)

ρd(ε)\rho_{d}(\varepsilon)(arbitrary unit)

ρp(ε)\rho_{p}(\varepsilon)(arbitrary unit)

(d) δh=0.266\delta_{\mathrm{h}}=0.266ρd(ε)\rho_{d}(\varepsilon)ρp(ε)\rho_{p}(\varepsilon)

ρd(ε)\rho_{d}(\varepsilon)(arbitrary unit)

ρp(ε)\rho_{p}(\varepsilon)(arbitrary unit)

ε\varepsilon(eV)(e) δh=0.293\delta_{\mathrm{h}}=0.293ρd(ε)\rho_{d}(\varepsilon)ρp(ε)\rho_{p}(\varepsilon)
Figure 4: ρd(ε)\rho_{d}(\varepsilon) and ρp(ε)\rho_{p}(\varepsilon) at T=170KT=170\,\mathrm{K}: (a) δh=0.123\delta_{\mathrm{h}}=0.123, (b) δh=0.166\delta_{\mathrm{h}}=0.166, (c) δh=0.205\delta_{\mathrm{h}}=0.205, (d) δh=0.266\delta_{\mathrm{h}}=0.266, and (e) δh=0.293\delta_{\mathrm{h}}=0.293.
00.10.10.20.20.30.30.40.40.50.50.60.60.70.70.80.800.10.10.20.20.30.30.40.40.50.50.60.60.70.70.80.8024681012-0.6-0.5-0.4-0.3-0.2-0.10.00.10.2024681012Refer to caption

ρd(ε)\rho_{d}(\varepsilon)(arbitrary unit)

\makebox(0.0,0.0)[]{{} }

(a)δh=0.123\delta_{\mathrm{h}}=0.123δh=0.166\delta_{\mathrm{h}}=0.166δh=0.205\delta_{\mathrm{h}}=0.205δh=0.266\delta_{\mathrm{h}}=0.266δh=0.293\delta_{\mathrm{h}}=0.293

ρp(ε)\rho_{p}(\varepsilon)(arbitrary unit)

\makebox(0.0,0.0)[]{{} }

ε\varepsilon(eV)(b)δh=0.123\delta_{\mathrm{h}}=0.123δh=0.166\delta_{\mathrm{h}}=0.166δh=0.205\delta_{\mathrm{h}}=0.205δh=0.266\delta_{\mathrm{h}}=0.266δh=0.293\delta_{\mathrm{h}}=0.293
Figure 5: Doping dependences of ρd(ε)\rho_{d}(\varepsilon) and ρp(ε)\rho_{p}(\varepsilon) at T=170KT=170\,\mathrm{K} around the coherent band: (a) ρd(ε)\rho_{d}(\varepsilon) and (b) ρp(ε)\rho_{p}(\varepsilon).

Here,

ρd(ε)\displaystyle\rho_{d}(\varepsilon) =\displaystyle= 1πN𝐤ImGdd(𝐤,iϵn)|iϵnε+iη,\displaystyle-\,\frac{1}{\pi N}\sum_{{\mathbf{k}}}\left.{\mathrm{Im}}\,G_{dd}({\mathbf{k}},\mathrm{i}\epsilon_{n})\right|_{\mathrm{i}\epsilon_{n}\rightarrow\varepsilon+i\eta}, (58)
ρp(ε)\displaystyle\rho_{p}(\varepsilon) =\displaystyle= 1πN𝐤[ImGxx(𝐤,iϵn)|iϵnε+iη+ImGyy(𝐤,iϵn)|iϵnε+iη].\displaystyle-\,\frac{1}{\pi N}\sum_{{\mathbf{k}}}\left[\left.{\mathrm{Im}}\,G_{xx}({\mathbf{k}},\mathrm{i}\epsilon_{n})\right|_{\mathrm{i}\epsilon_{n}\rightarrow\varepsilon+i\eta}+\left.{\mathrm{Im}}\,G_{yy}({\mathbf{k}},\mathrm{i}\epsilon_{n})\right|_{\mathrm{i}\epsilon_{n}\rightarrow\varepsilon+i\eta}\right]. (59)

iϵnε+iη\mathrm{i}\epsilon_{n}\rightarrow\varepsilon+i\eta indicates the performance of analytic continuation, for which I use the Pade´\acute{\mathrm{e}} approximation [42] and η=0.04tpd\eta=0.04\,t_{pd}. ρd(ε)\rho_{d}(\varepsilon) and ρp(ε)\rho_{p}(\varepsilon) represent the density of states (DOS) of the dd and pp bands, respectively. It has been confirmed that the peak positions of ρd(ε)\rho_{d}(\varepsilon) and ρp(ε)\rho_{p}(\varepsilon) hardly change even if η\eta is changed to 0.02tpd0.02\,t_{pd}. The three blocks appearing in ρd(ε)\rho_{d}(\varepsilon) correspond to the lower Hubbard band, coherent band, and upper Hubbard band. The coherent band is split due to the hybridization with the pp band, and the higher peak energy approaches the Fermi level with the hole doping [Fig. 5(a)]. In contrast, ρp(ε)\rho_{p}(\varepsilon) is large in the coherent band only. Reflecting that the holes are mainly doped into the pp band, as shown in Fig. 3, the peak energy moves away from the Fermi level with the hole doping [Fig. 5(b)]. Due to the competitive effect of these changes in DOS in the coherent band, there is a dome-shaped superconducting phase.

The superconducting gap function, given in matrix form by [Δ^(𝐤,ε)]μνΔμν(𝐤,ε)[\hat{\mathit{\Delta}}({\mathbf{k}},\varepsilon)]_{\mu\nu}\equiv{\mathit{\Delta}}_{\mu\nu}({\mathbf{k}},\varepsilon), is defined as follows:

Δ^(𝐤,ε)=iϵnImG^(𝐤,iϵn)Φ^(𝐤,iϵn)|iϵnε+iη,\hat{\mathit{\Delta}}({\mathbf{k}},\varepsilon)=\left.\mathrm{i}\epsilon_{n}\,{\mathrm{Im}}\,\hat{G}({\mathbf{k}},\mathrm{i}\epsilon_{n})\cdot\hat{\mathit{\Phi}}(\mathbf{k},\mathrm{i}\epsilon_{n})\right|_{\mathrm{i}\epsilon_{n}\rightarrow\varepsilon+i\eta}, (60)

where [G^(𝐤,iϵn)]μνGμν(𝐤,iϵn)[\hat{G}({\mathbf{k}},\mathrm{i}\epsilon_{n})]_{\mu\nu}\equiv G_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}) and [Φ^(𝐤,iϵn)]μνΦμν(𝐤,iϵn)[\hat{\mathit{\Phi}}({\mathbf{k}},\mathrm{i}\epsilon_{n})]_{\mu\nu}\equiv{\mathit{\Phi}}_{\mu\nu}({\mathbf{k}},\mathrm{i}\epsilon_{n}). Here, I use the Pade´\acute{\mathrm{e}} approximation for analytic continuation and η=0.04tpd\eta=0.04\,t_{pd}. It has been confirmed that Δμν(𝐤,0){\mathit{\Delta}}_{\mu\nu}({\mathbf{k}},0) hardly changes even if η\eta is changed to 0.02tpd0.02\,t_{pd}. The components of the superconducting gap function are classified into two classes. The first class is composed of Δdα(𝐤,ε){\mathit{\Delta}}_{d\alpha}({\mathbf{k}},\varepsilon) and Δαd(𝐤,ε){\mathit{\Delta}}_{\alpha d}({\mathbf{k}},\varepsilon), where α\alpha runs over xx and yy. The real parts of these components with ε=0\varepsilon=0 are shown in Fig. 6. The imaginary parts of these components with ε=0\varepsilon=0 are all zero. One can see that ReΔdα(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{d\alpha}({\mathbf{k}},0) [Fig. 6(a)] and ReΔαd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{\alpha d}({\mathbf{k}},0) [Fig. 6(b)] are roughly proportional to sinkα2\sin\frac{k_{\alpha}}{2}. These momentum dependences are derived from the first-order terms of JJ in Eqs. (38) and (39), which originate from the exchange interaction in Eq. (12). Thus, Δdα(𝐤,ε){\mathit{\Delta}}_{d\alpha}({\mathbf{k}},\varepsilon) and Δαd(𝐤,ε){\mathit{\Delta}}_{\alpha d}({\mathbf{k}},\varepsilon) emerge due to the exchange interaction via the SK mechanism. It can be verified that the SK mechanism can work effectively with the exchange interaction only if hpd>0h_{pd}>0. Moreover, the signs of ReΔdy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dy}({\mathbf{k}},0) and ReΔyd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yd}({\mathbf{k}},0) differ from the signs of ReΔdx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dx}({\mathbf{k}},0) and ReΔxd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xd}({\mathbf{k}},0), respectively. Therefore, as shown in Fig. 6(c), the linear combination Re{Δdx(𝐤,0)+Δdy(𝐤,0)+Δxd(𝐤,0)+Δyd(𝐤,0)}{\mathrm{Re}}\{{\mathit{\Delta}}_{dx}({\mathbf{k}},0)+{\mathit{\Delta}}_{dy}({\mathbf{k}},0)+{\mathit{\Delta}}_{xd}({\mathbf{k}},0)+{\mathit{\Delta}}_{yd}({\mathbf{k}},0)\} has line nodes at kx=kyk_{x}=k_{y} and kx=kyk_{x}=-k_{y} and behaves like a nodal dx2y2d_{x^{2}-y^{2}}-wave superconducting gap.

(a)Refer to captionReΔdx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dx}({\mathbf{k}},0)ReΔdy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dy}({\mathbf{k}},0)(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}100-10050-5005050100100

Energy (meV)

(b)Refer to captionReΔxd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xd}({\mathbf{k}},0)ReΔyd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yd}({\mathbf{k}},0)(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}100-10050-5005050100100

Energy (meV)

(c)Refer to caption500-50(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}100-10050-5005050100100

Energy (meV)

Figure 6: Δdα(𝐤,0){\mathit{\Delta}}_{d\alpha}({\mathbf{k}},0) and Δαd(𝐤,0){\mathit{\Delta}}_{\alpha d}({\mathbf{k}},0) at δh=0.237\delta_{\mathrm{h}}=0.237 and T=110KT=110\,{\mathrm{K}}: (a) ReΔdx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dx}({\mathbf{k}},0) and ReΔdy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dy}({\mathbf{k}},0), (b) ReΔxd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xd}({\mathbf{k}},0) and ReΔyd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yd}({\mathbf{k}},0), and (c) Re{Δdx(𝐤,0)+Δdy(𝐤,0)+Δxd(𝐤,0)+Δyd(𝐤,0)}{\mathrm{Re}}\{{\mathit{\Delta}}_{dx}({\mathbf{k}},0)+{\mathit{\Delta}}_{dy}({\mathbf{k}},0)+{\mathit{\Delta}}_{xd}({\mathbf{k}},0)+{\mathit{\Delta}}_{yd}({\mathbf{k}},0)\}.

The second class is composed of Δdd(𝐤,ε){\mathit{\Delta}}_{dd}({\mathbf{k}},\varepsilon) and Δαβ(𝐤,ε){\mathit{\Delta}}_{\alpha\beta}({\mathbf{k}},\varepsilon), where α\alpha and β\beta run over xx and yy. The real part of these components with ε=0\varepsilon=0 are shown in Fig. 7. The imaginary part of these components with ε=0\varepsilon=0 are all zero. ReΔαβ(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{\alpha\beta}({\mathbf{k}},0) [Fig. 7(b) and (c)] is roughly proportional to sinkα2sinkβ2\sin\frac{k_{\alpha}}{2}\sin\frac{k_{\beta}}{2}. This momentum dependence is derived from the first-order term of JJ in Eq. (40), which originates from the pair-hopping interaction in Eq. (13). ReΔdd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dd}({\mathbf{k}},0) [Fig. 7(a)] has the momentum dependence of an extended ss-wave, and its sign differs from the signs of ReΔxx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xx}({\mathbf{k}},0) and ReΔyy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yy}({\mathbf{k}},0). Thus, Δdd(𝐤,ε){\mathit{\Delta}}_{dd}({\mathbf{k}},\varepsilon)–as well as Δαβ(𝐤,ε){\mathit{\Delta}}_{\alpha\beta}({\mathbf{k}},\varepsilon)–emerges due to the pair-hopping interaction via the SK mechanism, although it is affected by the terms of UU and U2U^{2} in Eq. (2). It can be verified that the SK mechanism can work with the pair-hopping interaction even if hpd=0h_{pd}=0. Moreover, the absolute values of ReΔdd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dd}({\mathbf{k}},0) are larger than those of ReΔαβ(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{\alpha\beta}({\mathbf{k}},0) for all α\alpha, β\beta, and 𝐤{\mathbf{k}}. Therefore, as shown in Fig. 7(d), the linear combination Re{Δdd(𝐤,0)+Δxx(𝐤,0)+Δxy(𝐤,0)+Δyx(𝐤,0)+Δyy(𝐤,0)}{\mathrm{Re}}\{{\mathit{\Delta}}_{dd}({\mathbf{k}},0)+{\mathit{\Delta}}_{xx}({\mathbf{k}},0)+{\mathit{\Delta}}_{xy}({\mathbf{k}},0)+{\mathit{\Delta}}_{yx}({\mathbf{k}},0)+{\mathit{\Delta}}_{yy}({\mathbf{k}},0)\} behaves like an extended ss-wave superconducting gap.

(a)Refer to captionReΔdd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dd}({\mathbf{k}},0)(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}40-4030-3020-2010-100

Energy (meV)

(b)Refer to captionReΔxx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xx}({\mathbf{k}},0)ReΔxy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xy}({\mathbf{k}},0)(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}20-2010-10010102020

Energy (meV)

(c)Refer to captionReΔyx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yx}({\mathbf{k}},0)ReΔyy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yy}({\mathbf{k}},0)(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}20-2010-10010102020

Energy (meV)

(d)Refer to caption-20-25-30(2π,0)(2\pi,0)kxk_{x}(2π,2π)(2\pi,2\pi)kyk_{y}40-4020-200

Energy (meV)

Figure 7: Δdd(𝐤,0){\mathit{\Delta}}_{dd}({\mathbf{k}},0) and Δαβ(𝐤,0){\mathit{\Delta}}_{\alpha\beta}({\mathbf{k}},0) at δh=0.237\delta_{\mathrm{h}}=0.237 and T=110KT=110\,{\mathrm{K}}: (a) ReΔdd(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{dd}({\mathbf{k}},0), (b) ReΔxx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xx}({\mathbf{k}},0) and ReΔxy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{xy}({\mathbf{k}},0), (b) ReΔyx(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yx}({\mathbf{k}},0) and ReΔyy(𝐤,0){\mathrm{Re}}{\mathit{\Delta}}_{yy}({\mathbf{k}},0), and (d) Re{Δdd(𝐤,0)+Δxx(𝐤,0)+Δxy(𝐤,0)+Δyx(𝐤,0)+Δyy(𝐤,0)}{\mathrm{Re}}\{{\mathit{\Delta}}_{dd}({\mathbf{k}},0)+{\mathit{\Delta}}_{xx}({\mathbf{k}},0)+{\mathit{\Delta}}_{xy}({\mathbf{k}},0)+{\mathit{\Delta}}_{yx}({\mathbf{k}},0)+{\mathit{\Delta}}_{yy}({\mathbf{k}},0)\}.

I have shown that the coexistence of extended ss- and dd-wave gaps is theoretically possible in the three-band ttJJUU model. The coexistence of ss- and dd-wave gaps was originally proposed to explain the apparently conflicting results of scanning tunnelling spectroscopy in HTSC [43]. So far, the experiments on Bi2Sr2CaCu2O8+δ (Bi2212) utilizing tunneling effect in the superconducting phase, which include cc-axis twist Josephson experiments [44, 45, 46, 47, 48, 49, 50], cc-axis scanning tunnelling microscopy [51, 52, 53], and intrinsic Josephson junction terahertz emission [54, 55], provide clear evidences that the superconducting gap has ss-wave symmetry. These experiments can directly observe the superconducting gap without breaking the gap into quasiparticles, and this result is also reasonable for the coexistence of ss- and dd-wave gaps. In the superconducting phase, where ss- and dd-wave gaps coexist, the ss-wave gap is dominant over the dd-wave gap in the energy |ε|<Δs|\varepsilon|<{\mathit{\Delta}}_{s}, where Δs{\mathit{\Delta}}_{s} indicates the ss-wave gap magnitude. On the other hand, only when the dd-wave gap magnitude Δd{\mathit{\Delta}}_{d} satisfies Δd>Δs{\mathit{\Delta}}_{d}>{\mathit{\Delta}}_{s}, the dd-wave gap becomes dominant over the ss-wave gap in the energy |ε|>Δs|\varepsilon|>{\mathit{\Delta}}_{s}.

In contrast, the quasiparticles from the dd-wave gap can be observed in the energy |ε|<Δs|\varepsilon|<{\mathit{\Delta}}_{s}, where their excitation energies are always smaller than those of the quasiparticles from the ss-wave gap. Thus, the experimental method breaking the gap into quasiparticles does mainly observe the dd-wave gap. For example, both temperature and magnetic field dependences of low-temperature specific heat indicate that the dd-wave superconducting gap exists in near optimally doped Bi2Sr2-xLaxCuO(x0.4)6+δ{}_{6+\delta}\,(x\sim 0.4) [56].

The above discussion holds even if the dd-wave gap is not a superconducting gap. The angle-resolved photoemission spectroscopy (ARPES) experiment on Bi2212 shows the marked change of temperature dependence of spectral intensity across critical value pc0.19p_{c}\sim 0.19 with hole doping [57]. This change with hole doping pp can be interpreted as a result of the coexistence of ss- and dd-wave gaps when we replace energy with temperature in the above discussion. For p<pcp<p_{c}, the dd-wave gap affects the electronic structure above TcT_{c} if Δd>Δs{\mathit{\Delta}}_{d}>{\mathit{\Delta}}_{s}. The electronic structure affected by the dd-wave gap is called pseudogap. However, for p>pcp>p_{c}, both the ss- and dd-wave gaps do not affect the electronic structure above TcT_{c} if Δd<Δs{\mathit{\Delta}}_{d}<{\mathit{\Delta}}_{s}. Therefore, the pseudogap disappears across pcp_{c} with hole doping, which has also been observed by the ARPES experiment [57].

Furthermore, Raman spectroscopy [58] and the magnetic field penetration depth measurement by muon-spin rotation [59, 60, 61] have provided evidence that supports the coexistence of ss- and dd-wave gaps in hole-doped HTSC. In theoretical work, the possibility of the coexistence of an extended ss- and dd-wave superconducting state has been shown with the analysis of the 2D ttJJ model considering fluctuation effects [13], and further experimental and theoretical research that assumes such coexistence is desired in the future.

I conclude by comparing the obtained superconducting state to that found in other theoretical work. The dx2y2d_{x^{2}-y^{2}}-wave superconducting gap composed of Δdα(𝐤,ε){\mathit{\Delta}}_{d\alpha}({\mathbf{k}},\varepsilon) and Δαd(𝐤,ε){\mathit{\Delta}}_{\alpha d}({\mathbf{k}},\varepsilon), which emerges due to the exchange interaction via the SK mechanism, corresponds to the one mediated by antiferromagnetic spin fluctuations (AFSF) [62]. This is clear because the superexchange interaction among dd electrons, which is responsible for the AFSF, can be derived from the exchange interaction between dd and pp electrons. In general, once the superexchange interaction acts between charge carriers, the dx2y2d_{x^{2}-y^{2}}-wave superconductivity can emerge [63]. Moreover, the dx2y2d_{x^{2}-y^{2}}-wave superconductivity in my model can emerge only with the dd-pp band hybridization. Therefore, it must be important that the dd electron is implicitly hybridized with the pp electron in the AFSF-mediated superconductivity. This speculation is supported by the studies of Kondo lattice models proposed for copper oxide [64, 65, 66, 67, 68, 69, 70, 71]. The Kondo interaction between localized dd spin and pp electron in Kondo lattice models corresponds to the exchange interaction between dd and pp electrons in the large-UU limit of my model. The studies of Kondo lattice models indicate that superconductivity emerges due to the Kondo effect, the compensation for the localized dd spin by the pp electrons via the Kondo interaction. As the Kondo effect corresponds to the formation of a Fermi liquid state through the dd-pp band hybridization [30], the superconductivity in Kondo lattice models is consistent with the dx2y2d_{x^{2}-y^{2}}-wave superconductivity in my model.

The extended ss-wave superconducting gap composed of Δdd(𝐤,ε){\mathit{\Delta}}_{dd}({\mathbf{k}},\varepsilon) and Δαβ(𝐤,ε){\mathit{\Delta}}_{\alpha\beta}({\mathbf{k}},\varepsilon), which emerges due to the pair-hopping interaction via the SK mechanism, corresponds to the kinetic-energy-driven superconductivity of the single-band ttJJ model [72, 73, 74, 75, 76, 77, 78, 79, 80]. In the kinetic-energy-driven superconductivity, the charge carriers form the superconducting pairs to gain kinetic energy. This energy gain can be derived from the pair-hopping interaction between pp and dd electrons, which works to form the extended ss-wave superconducting gap in my model.

4 Summary

In summary, the three-band ttJJUU model is derived assuming that the double occupancy by dd electrons is not excluded. When the dd electron is hybridized with the pp electron through exchange and pair-hopping interactions, the dome-shaped superconducting phase can be reproduced despite the strong correlations among dd electrons. In the superconducting phase, the extended ss- and dx2y2d_{x^{2}-y^{2}}-wave superconducting gaps coexist. The extended ss-wave gap emerges due to the pair-hopping interaction via the SK mechanism, which works effectively due to the charge fluctuations. In contrast, the dx2y2d_{x^{2}-y^{2}}-wave gap emerges due to the exchange interaction via the SK mechanism, which can effectively work only with the dd-pp band hybridization. The obtained superconducting state is consistent with those in other theoretical work, which include AFSF-mediated superconductivity and kinetic-energy-driven superconductivity.

Acknowledgements

The author would like to thank Prof. T. Tohyama and Prof. H. Yamase for their invaluable comments. The author is also grateful to anonymous reviewers for providing informations on many important references and insightful comments.

References

References

  • [1] Chao K A, Spałek J and Oleś A M 1977 J. Phys. C: Solid State Phys.10 L271
  • [2] Chao K A, Spałek J and Oleś A M 1978 Phys. Rev.B 18 3453
  • [3] Zhang F C and Rice T M 1988 Phys. Rev.B 37 3759
  • [4] Zaanen J and Oleś A M 1988 Phys. Rev.B 37 9423
  • [5] Matsukawa H and Fukuyama H 1989 J. Phys. Soc. Japan58 2845
  • [6] Mori M, Tohyama T and Maekawa S 2002 Phys. Rev.B 66 064502
  • [7] Plakida N M 2002 Cond. Matter Phys. 5 707
  • [8] Rösch O and Gunnarsson O 2004 Phys. Rev. Lett.92 146403
  • [9] Ishihara S and Nagaosa N 2004 Phys. Rev.B 69 144520
  • [10] Tohyama T 2004 Phys. Rev.B 70 174517
  • [11] Mishchenko A S and Nagaosa N 2004 Phys. Rev. Lett.93 036402
  • [12] Mishchenko A S and Nagaosa N 2006 Phys. Rev.B 73 092502
  • [13] Mallik Aabhaas V, Gupta Gaurav K, Shenoy Vijay B and Krishnamurthy H R 2020 Phys. Rev. Lett.124 147002
  • [14] Zhang F C 2003 Phys. Rev. Lett.90 207002
  • [15] Yuan F, Yuan Q and Ting C S 2005 Phys. Rev.B 71 104505
  • [16] Wang Q H, Wang Z D, Chen Y and Zhang F C 2006 Phys. Rev.B 73 092507
  • [17] Abram M, Kaczmarczyk J, Jędrak J and Spałek J 2013 Phys. Rev.B 88 094502
  • [18] Spałek J, Zegrodnik M and Kaczmarczyk J 2017 Phys. Rev.B 95 024506
  • [19] Zegrodnik M and Spałek J 2017 Phys. Rev.B 96 054511
  • [20] Fidrysiak M and Spałek J 2020 Phys. Rev.B 102 014505
  • [21] Schrieffer J R and Wolff P A 1966 Phys. Rev.149 491
  • [22] Suhl H, Matthias B T and Walker L R 1959 Phys. Rev. Lett.3 552
  • [23] Kondo J 1963 Prog. Theor. Phys. 29 1
  • [24] Kondo J 2002 J. Phys. Soc. Japan71 1353
  • [25] Emery V J 1987 Phys. Rev. Lett.58 2794
  • [26] Penson K A and Kolb M 1986 Phys. Rev.B 33 1663
  • [27] Belkasri A and Buzatu F D 1996 Phys. Rev.B 53 7171
  • [28] Robaszkiewicz S and Bułka B D 1999 Phys. Rev.B 59 6430
  • [29] Kapcia K J 2014 Acta Phys. Pol. A 126 53
  • [30] Yamada K, Yosida K and Hanzawa K 1992 Prog. Theor. Phys. Suppl. No. 108 141
  • [31] Yamada K 1975 Prog. Theor. Phys. 53 970
  • [32] Yosida K and Yamada K 1975 Prog. Theor. Phys. 53 1286
  • [33] Georges A and Kotliar G 1992 Phys. Rev.B 45 6479
  • [34] Zhang X Y, Rozenberg M J and Kotliar G 1993 Phys. Rev. Lett.70 1666
  • [35] Ino A, Mizokawa T, Fujimori A, Tamasaku K, Uchida S, Kimura T, Sasagawa T and Kishio K 1997 Phys. Rev. Lett.79 2101
  • [36] Fujimori A, Ino A, Mizokawa T, Kim C, Shen Z X, Sasagawa T, Kimura T, Kishio K, Takaba M, Tamasaku K, Eisaki H and Uchida S 1998 J. Phys. Chem. Solids 59 1892
  • [37] Fujimori A, Ino A, Yoshida T, Mizokawa T, Shen Z X, Kim C, Kakeshita T, Eisaki H and Uchida S 2001, Open Problems in Strongly Correlated Electron Systems ed J. Bonca et al. (Dordrecht: KluIr Academic Pub.) p 119
  • [38] Tsukada A, Yamamoto H and Naito M 2006 Phys. Rev.B 74 174515
  • [39] Yoshida T, Zhou X J, Tanaka K, Yang W L, Hussain Z, Shen Z X, Fujimori A, Sahrakorpi S, Lindroos M, Markiewicz R S, Bansil A, Komiya Seiki, Ando Yoichi, Eisaki H, Kakeshita T and Uchida S 2006 Phys. Rev.B 74 224510
  • [40] Wen H-H, Shan L, Wen X G, Wang Y, Gao H, Liu Z Y, Zhou F, Xiong J W and Ti W X 2005 Phys. Rev.B 72 134507
  • [41] Wang Y, Yan J, Shan L, Wen H-H, Tanabe Y, Adachi T and Koike Y 2007 Phys. Rev.B 76 064512
  • [42] Vidberg H J and Serene J W 1977 J. Low. Temp. Phys. 29 179
  • [43] Müller K A 1995 Nature 377 133
  • [44] Li Qiang, Tsay Y N, Suenaga M, Klemm R A, Gu G D and Koshizuka N 1999 Phys. Rev. Lett.83 4160
  • [45] Takano Y, Hatano T, Fukuyo A, Ishii A, Ohmori M, Arisawa S, Togano K and Tachiki M 2002 Phys. Rev.B 65 140513(R)
  • [46] Takano Y, Hatano T, Ohmori M, Kawakami S, Ishii A, Arisawa S, Kim S-J, Yamashita T, Togano K and Tachiki M 2003 J. Low. Temp. Phys. 131 533
  • [47] Takano Y, Hatano T, Kawakami S, Ohmori M, Ikeda S, Nagao M, Inomata K, Yun K S, Ishii A, Tanaka A, Yamashita T and Tachiki M 2004 Physica C 408-410 296
  • [48] Latyshev Y I, Orlov A P, Nikitina A M, Monceau P and Klemm R A 2004 Phys. Rev.B 70 094517
  • [49] Klemm R A 2005 Phil. Mag. 85 801
  • [50] Zhu Y, Liao M, Zhang Q, Xe H-Y, Meng F, Liu Y, Bai Z, Ji S, Zhang J, Jiang K, Zhong R, Schneeloch J, Gu G, Gu L, Ma X, Zhang D and Xue Q-K 2021 arXiv:1903.07965v2
  • [51] Misra S, Oh S, Hornbaker D J, DiLuccio T, Eckstein J N and Yazdani A 2002 Phys. Rev. Lett.89 087002
  • [52] Hoogenboom B W, Kadowaki K, Revaz B and Fischer Ø 2003 Physica C 391 376
  • [53] Zhong Y, Wang Y, Han S, Lv Y-F, Wang W-L, Zhang D, Ding H, Zhang Y-M, Wang L, He K, Zhong R, Schneeloch J, Gu G-D, Song C-L, Ma X-C and Xue Q-K 2016 Sci. Bull. 61 1239
  • [54] Kashiwagi T, Sakamoto K, Kubo H, Shibano Y, Enomoto T, Kitamura T, Asanuma K, Yasui T, Watanabe C, Nakade K, Saiwai Y, Katsuragawa T, Tsujimoto M, Yoshizaki R, Yamamoto T, Minami H, Klemm R A and Kadowaki K 2015 Appl. Phys. Lett. 107 082601
  • [55] Kashiwagi T, Yamamoto T, Minami H, Tsujimoto M, Yoshizaki R, Delfanazari K, Kitamura T, Watanabe C, Nakade K, Yasui T, Asanuma K, Saiwai Y, Shibano Y, Enomoto T, Kubo H, Sakamoto K, Katsuragawa T, Marković B, Mirković J, Klemm R A and Kadowaki K 2015 Phys. Rev.Appl. 4 054018
  • [56] Wang Y, Liu Z-Y, Lin C T and Wen H-H 2011 Phys. Rev.B 83 054509
  • [57] Chen S-D, Hashimoto M, He Y, Song D, Xu K-J, He J-F, Devereaux T P, Eisaki H, Lu D-H, Zaanen J, Shen Z-X 2019 Science 366 1099
  • [58] Masui T, Limonov M, Uchiyama H, Lee S, Tajima S and Yamanaka A 2003 Phys. Rev.B 68 060506(R)
  • [59] Khasanov R, Shengelaya A, Maisuradze A, La Mattina F, Bussmann-Holder A, Keller H and Müller K A 2007 Phys. Rev. Lett.98 057007
  • [60] Khasanov R, Strässle S, Di Castro D, Masui T, Miyasaka S, Tajima S, Bussmann-Holder A and Keller H 2007 Phys. Rev. Lett.99 237601
  • [61] Khasanov R, Shengelaya A, Bussmann-Holder A, Karpinski J, Keller H and Müller K A 2008 J. Supercond. Nov. Magn. 21 81
  • [62] Moriya T and Ueda K 2000 Adv. Phys. 49 555 and references therein
  • [63] Que W-M and Kirczenow G 1987 Solid State Commun.64 1052
  • [64] Prelovšek P 1988 Phys. Lett.A 126 287
  • [65] Ramšak A and Prelovšek P 1989 Phys. Rev.B 40 2239
  • [66] Castellani C, Di Castro C and Grilli M 1988 Physica C 153-155 1659
  • [67] Castellani C, Di Castro C and Grilli M 1988 Int. J. Mod. Phys. B 2 659
  • [68] Cancrini N, Caprara S, Castellani C, Di Castro C, Grilli M and Raimondi R 1991 Europhys. Lett. 14 597
  • [69] Kamimura H 1988 Int. J. Mod. Phys. B 2 699
  • [70] Andrei N and Coleman P 1989 Phys. Rev. Lett.62 595
  • [71] Hatsugai Y, Imada M and Nagaosa N 1989 J. Phys. Soc. Japan58 1347
  • [72] Hirsch J E 1992 Physica C 201 347
  • [73] Tsunetsugu H and Imada M 1998 J. Phys. Soc. Japan67 1864
  • [74] Tsunetsugu H and Imada M 1999 J. Phys. Soc. Japan68 3162
  • [75] Imada M and Onoda S 2001, Open Problems in Strongly Correlated Electron Systems ed J. Bonca et al. (Dordrecht: KluIr Academic Pub.) p 69
  • [76] Sarker S K and Lovorn T 2010 Phys. Rev.B 82 014504
  • [77] Sarker S K and Lovorn T 2012 Phys. Rev.B 85 144502
  • [78] Feng S, Zhao H and Huang Z 2012 Phys. Rev.B 85 054509
  • [79] Feng S, Kuang L and Zhao H 2015 Physica C 517 5
  • [80] Gao D, Mou Y and Feng S 2018 J. Low Temp. Phys. 192 19