Introduction

The symmetry principle is one of the most powerful tools to diagnose low-energy electronic band structures, lattice vibrations, and linear responses1, and is also valuable to explore various symmetry-breaking ordered phases such as magnetism, charge/spin density-wave, nematicity and superconductivity2. The crystal symmetry of a solid-state system dictates the normal band structures it hosts near the Fermi level, which could in turn determine the most favorable superconducting pairing symmetry3,4. This symmetry principle for superconductors (SC) is recently extended to investigate multi-band unconventional superconductivity5,6,7. Interestingly, the orbital-independent and orbital-dependent pairings that belong to the same symmetry representation may coexist with each other8. Such orbital-dependent pairings have been studied in a wide variety of systems with multi-band character, including Sr2RuO49, iron-chalcogenide SCs10,11,12,13, Cu-doped Bi2Se314 and half-Heusler compounds15,16,17,18, from which the guiding principle by symmetry is crucial to understanding the nature of unconventional superconductivity.

A few specific systems can be effectively characterized by a general normal-state model Hamiltonian that contains both spin ({, }) and pseudospin ({1, 2}) degrees of freedom, where pseudospin could originate from two atomic orbitals, two sublattices, two layers, or two valleys6. We start from a spin-singlet centrosymmetric SC to explore the existence of even-parity pseudospin-triplet pairings, for example, c1,(k)c2,( − k) + c2,(k)c1,( − k) − c1,(k)c2,( − k) − c2,(k)c1,( − k), and further investigate their valuable roles in tailoring anisotropic pairing gap functions with or without nodes19. Different from spin-triplet pairings, spin-singlet pseudospin-triplet pairings have not been much explored in real materials since such pairings are usually considered to be energetically unfavorable. This is partly due to the common belief that the double degeneracy of the two orbitals is lifted by orbital hybridization so that the orbital-dependent pairing would be severely suppressed under crystal field splitting or electron-electron repulsive interaction. One aim of this work is concerned with the possible condition for the existence of even-parity spin-singlet orbital-dependent pairings, and possible applications to real materials.

On the other hand, the effects of symmetry breaking in unconventional SCs is an important topic that has attracted tremendous interest. The symmetry could be broken explicitly by external fields or strain, or be broken spontaneously from many-body interactions. Two typical examples are rotational symmetry breaking20,21 and time-reversal-symmetry (TRS) breaking22,23,24,25,26. Besides, the interplay between nematicity and superconductivity is yet to be fully understood in some real materials, such as FeSe12,27, where gap functions can be highly anisotropic. These systems are all multi-band SCs, while symmetry-reducing signatures are experimentally observed above the superconducting transition temperature, which is mainly caused by both crystal field splittings and interaction-induced order parameters (e.g. nematicity). Thus, discovering the coexistence of nematicity and superconductivity in these multi-band systems can shed new light on understanding the underlying favorable pairing symmetries.

The main finding of this work is that the anisotropic gap functions with or without nodes could be attributed to the mixing of isotropic s-wave pairing and even-parity spin-singlet pseudospin-triplet pairing, even though both of them belong to the A1g symmetry representation. For technical conveniences, we adopt an orbital do(k)-vector notations11 to describe the pairing matrix and similarly a go(k)-vector for orbital hybridization in the two-orbital subspace ({1, 2}). Solving linearized gap equations, we show that the presence of go-vector generally suppresses the superconductivity with orbital do-vector except for do(k)∥go(k), which is consistent with the concept of superconducting fitness6. This sets up weak-coupling criteria for A1g-type orbital-dependent pairings that could naturally give rise to anisotropic gap functions in real superconducting materials. Moreover, we reveal a deep connection between two-orbital nematic SC and pseudospin-triplet pairings. Within the mean-field theory for electron-electron repulsive interactions, the nematic order develops in the orbital subspace at T < Tnem, which also contributes to the total orbital hybridization, gtot = go + gnem. This leads to the stabilization of a nematic orbital do-vector for do(k)∥gtot(k), indicating the coexistence of nematicity and superconductivity. The direct applications to FeSe12,27 are also discussed. We also generalize it to a two-valley system with C6 breaking terms (e.g., Kekulé distortion). In the end, we also predict an orbital-polarized superconducting state.

Results

Classification of Spin-singlet Orbital-triplet pairings

To explore the weak-coupling criterion for the energetically favorable even-parity spin-singlet pseudospin-triplet pairing, we consider the mean-field pairing Hamiltonian,

$${{{{{{\mathcal{H}}}}}}}_{\Delta }=\mathop{\sum}\limits_{{{{{{\bf{k}}}}}}}\mathop{\sum}\limits_{{s}_{1}a,{s}_{2}b}{{{\Delta }}}_{{s}_{1},{s}_{2}}^{a,b}({{{{{\bf{k}}}}}}){F}_{{s}_{1}a,{s}_{2}b}^{{{\dagger}} }({{{{{\bf{k}}}}}})+\,{{\mbox{h.c.}}}\,,$$
(1)

where \({F}_{{s}_{1}a,{s}_{2}b}^{{{\dagger}} }({{{{{\bf{k}}}}}})={c}_{{s}_{1}a}^{{{\dagger}} }({{{{{\bf{k}}}}}}){c}_{{s}_{2}b}^{{{\dagger}} }(-{{{{{\bf{k}}}}}})\) is the creation operator of Cooper pairs, s1, s2 are indices for spins and a, b are for pseudospins (e.g., two orbitals {1, 2}). A general pairing potential of a two-band model is a four-by-four matrix6. In particular, the spin-singlet pairing function \({{{\Delta }}}_{{s}_{1},{s}_{2}}^{a,b}({{{{{\bf{k}}}}}})=f({{{{{\bf{k}}}}}}){M}_{a,b}({{{{{\bf{k}}}}}}){(i{\sigma }_{2})}_{{s}_{1},{s}_{2}}\) consists of the angular form factor f(k) and Ma,b(k) in the orbital channel. The spin-singlet pairings are not mixed with spin-triplet pairings in the absence of spin-orbit coupling (SOC). In analogy to spin-triplet SCs, for the technical convenience, we then use an orbital do(k)-vector for the spin-singlet orbital-dependent pairing potential11,

$${\hat{{{\Delta }}}}_{tot}({{{{{\bf{k}}}}}})=[{{{\Delta }}}_{s}{{{\Psi }}}_{s}({{{{{\bf{k}}}}}}){\tau }_{0}+{{{\Delta }}}_{o}({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})](i{\sigma }_{2}),$$
(2)

where Δs and Δo are pairing strengths in orbital-independent and orbital-dependent channels, respectively. Here τ and σ are Pauli matrices acting on the orbital and spin subspace, respectively, and τ0 is a 2-by-2 identity matrix. When both Δs and Δo are real, a real orbital do(k)-vector preserves TRS while a complex one spontaneously breaks TRS (\({{{{{\mathcal{T}}}}}}=i{\tau }_{0}{\sigma }_{2}{{{{{\mathcal{K}}}}}}\) with \({{{{{\mathcal{K}}}}}}\) being complex conjugate). The Fermi statistics requires Ψs(k) = Ψs( − k), \({d}_{o}^{1,3}({{{{{\bf{k}}}}}})={d}_{o}^{1,3}(-{{{{{\bf{k}}}}}})\) and \({d}_{o}^{2}({{{{{\bf{k}}}}}})=-{d}_{o}^{2}(-{{{{{\bf{k}}}}}})\). In other words, \({d}_{o}^{2}({{{{{\bf{k}}}}}})\) describes odd-parity spin-singlet orbital-singlet pairings and the other two are for even-parity spin-singlet orbital-triplet pairings. Moreover, we provide an alternative definition of orbital do-vectors in Supplementary Note 1. Even though the orbital-independent part Ψs(k) is also “orbital-triplet” by statistics, it is completely trivial. Hereafter, we only refer to \({d}_{o}^{1}({{{{{\bf{k}}}}}})\) and \({d}_{o}^{3}({{{{{\bf{k}}}}}})\) as orbital-triplet pairings28.

In addition, the basis functions for both Ψs(k) and orbital do(k)-vectors in Eq. (2) could be classified by crystalline symmetry.

Under the action of an n-fold rotation operator Cn about the z-axis, the pairing potential \(\hat{{{\Delta }}}({{{{{\bf{k}}}}}})\) transforms as

$${{{{{\mathcal{D}}}}}}[{C}_{n}]\,{\hat{{{\Delta }}}}_{J}({{{{{\bf{k}}}}}})\,{({{{{{\mathcal{D}}}}}}[{C}_{n}])}^{T}={e}^{i\frac{2\pi }{n}J}\,{\hat{{{\Delta }}}}_{J}({C}_{n}^{-1}{{{{{\bf{k}}}}}}),$$
(3)

where \({{{{{\mathcal{D}}}}}}[{C}_{n}]\) is the corresponding matrix representation, J is the orbital angular momentum quantum number, and also labels the irreducible representations of the Cn point group. For example, J = 0 is for A representation and J = 2 is for B representation. Firstly, the TRS requires the coexistence of \({\hat{{{\Delta }}}}_{J}\) and \({\hat{{{\Delta }}}}_{-J}\) with equal weight. If the rotation symmetry Cn is further imposed, then J and − J have to be equivalent modulo n, i.e. J ≡ − J mod n. The results for the basis functions of Ψs(k) and do(k) are summarized in Table 1 for a two-band SC with the {dxz, dyz}-orbitals. In this case, \({{{{{\mathcal{D}}}}}}[{C}_{n}]=[\cos (\frac{2\pi }{n}){\tau }_{0}-i\sin (\frac{2\pi }{n}){\tau }_{2}]\otimes {\sigma }_{0}\). For instance, \({{{{{\mathcal{D}}}}}}[{C}_{4}]=-i{\tau }_{2}\otimes {\sigma }_{0}\) explains that both Δoτ1 and Δoτ3 are d-wave-like pairing states29.

Table 1 Classification of spin-singlet pairing potentials for Eq. (2).

At the mean-field level, the Bogoliubov de-Gennes (BdG) Hamiltonian is given by

$$\begin{array}{r}{{{{{{\mathcal{H}}}}}}}_{{{{{{\rm{BdG}}}}}}}=\left(\begin{array}{rc}{{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})&{\hat{{{\Delta }}}}_{tot}({{{{{\bf{k}}}}}})\\ {\hat{{{\Delta }}}}_{tot}^{{{\dagger}} }({{{{{\bf{k}}}}}})&-{{{{{{\mathcal{H}}}}}}}_{0}^{* }(-{{{{{\bf{k}}}}}})\end{array}\right),\end{array}$$
(4)

where \({{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})\) represents a two-band normal-state Hamiltonian with both spin and pseudospin degrees of freedom.

In general, the BdG Hamiltonian is also invariant under the Cn rotation symmetry, i.e., \({{{{{{\mathcal{D}}}}}}}_{{{{{{\rm{BdG}}}}}}}[{C}_{n}]\,{{{{{{\mathcal{H}}}}}}}_{{{{{{\rm{BdG}}}}}}}({{{{{\bf{k}}}}}})\,{({{{{{{\mathcal{D}}}}}}}_{{{{{{\rm{BdG}}}}}}}[{C}_{n}])}^{{{\dagger}} }={{{{{{\mathcal{H}}}}}}}_{{{{{{\rm{BdG}}}}}}}({C}_{n}^{-1}{{{{{\bf{k}}}}}})\) when we define \({{{{{{\mathcal{D}}}}}}}_{{{{{{\rm{BdG}}}}}}}[{C}_{n}]=\left(\begin{array}{rc}{{{{{\mathcal{D}}}}}}[{C}_{n}]&0\\ 0&{e}^{i\frac{2\pi }{n}J}{({{{{{\mathcal{D}}}}}}[{C}_{n}])}^{* }\end{array}\right)\) based on Eq. (3).

Here we assume both inversion and time-reversal symmetries are preserved. To be specific, we consider a SOC-free Hamiltonian,

$${{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})=\epsilon ({{{{{\bf{k}}}}}}){\tau }_{0}{\sigma }_{0}+{\lambda }_{o}({{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){\sigma }_{0},$$
(5)

where the basis is \({\psi }_{{{{{{\bf{k}}}}}}}^{{{\dagger}} }=({c}_{1,\uparrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{1,\downarrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{2,\uparrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{2,\downarrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}))\), \(\epsilon ({{{{{\bf{k}}}}}})=({k}_{x}^{2}+{k}_{y}^{2})/2m-\mu \) is the band energy measured relative to the chemical potential μ, m is the effective mass, λo represents the orbital hybridization and go(k) = (g1(k), g2(k), g3(k)). And the g3-component leads to the different effective masses of different orbitals. As mentioned earlier, this vector notation is just for the technical convenience. Besides, the g1 and g2 components are determined by symmetries. For example, TRS requires g1,3(k) = g1,3( − k) and g2(k) = − g2( − k). If inversion symmetry (IS) is present, g2(k) (or g1(k)) must vanish for \({{{{{\mathcal{I}}}}}}={\tau }_{0}{\sigma }_{0}\) (or \({{{{{\mathcal{I}}}}}}={\tau }_{3}{\sigma }_{0}\)), which is the same as the constraint for the orbital do-vector. The more explicit form of go(k) is determined by other crystal symmetries.

In general, the pseudospin-triplet (i.e. orbital-triplet) pairing state shares some similarities with the spin-triplet pairing state30. To show that, we first discuss the superconducting quasi-particle spectrum of orbital-triplet SCs in the absence of band-splitting caused by orbital hybridizations, i.e., go(k) = 0 for Eq. (5). In this case, the superconducting gaps on the Fermi surface are

$$E({{{{{\bf{k}}}}}})=\pm | {{{\Delta }}}_{o}| \sqrt{| {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}}){| }^{2}\pm | {{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\times {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})| },$$
(6)

for the Δs = 0 limit. This indicates that there are two distinct gaps if TRS is spontaneously broken. In the following, we mainly focus on the time-reversal-invariant superconducting states, i.e., real do-vectors, for which the classification of pairing potentials is shown in Table 1 based on Eq. (3). We will show the interplay between Δs and Δo can lead to anisotropic superconducting gaps on different Fermi surfaces. Moreover, its stability against orbital-hybridization, electron-electron interactions, and applications to real materials will be discussed in detail as follows. We will also briefly comment on the effects of TRS-breaking in the end.

Stability for spin-singlet orbital-triplet pairings

We apply the weak-coupling scheme6 for spin-singlet orbital-triplet pairings against crystal field splittings, which cause orbital hybridizations [i.e. the go(k) term in Eq. (5)]. We analytically calculate the superconductivity instability for the orbital do-vector by BCS decoupling scheme. The superconducting transition temperature Tc of orbital-dependent pairing channels is calculated by solving the linearized gap equation,

$${{{\Delta }}}_{{s}_{1},{s}_{2}}^{a,b}({{{{{\bf{k}}}}}})= -\frac{1}{\beta }\mathop{\sum}\limits_{{\omega }_{n}}\mathop{\sum}\limits_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}{V}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}^{{s}_{1}a,{s}_{2}b}({{{{{\bf{k}}}}}},{{{{{{\bf{k}}}}}}}^{{\prime} })\\ \times {\left[{G}_{e}({{{{{{\bf{k}}}}}}}^{{\prime} },i{\omega }_{n})\hat{{{\Delta }}}({{{{{{\bf{k}}}}}}}^{{\prime} }){G}_{h}(-{{{{{{\bf{k}}}}}}}^{{\prime} },i{\omega }_{n})\right]}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }},$$
(7)

where β = 1/kBT, \({G}_{e}({{{{{\bf{k}}}}}},i{\omega }_{n})={[i{\omega }_{n}-{{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})]}^{-1}\) is the Matsubara Green’s function for electrons with ωn = (2n + 1)π/β and \({G}_{h}({{{{{\bf{k}}}}}},i{\omega }_{n})=-{G}_{e}^{* }({{{{{\bf{k}}}}}},i{\omega }_{n})\). We expand the attractive interactions as \({V}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}^{{s}_{1}a,{s}_{2}b}({{{{{\bf{k}}}}}},{{{{{{\bf{k}}}}}}}^{{\prime} })=-{\sum }_{\Gamma ,l}{[{v}^{\Gamma }{{{{{{\bf{d}}}}}}}_{o}^{\Gamma ,l}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}i{\sigma }_{2}]}_{{s}_{1}a,{s}_{2}b}{[{{{{{{\bf{d}}}}}}}_{o}^{\Gamma ,l}({{{{{{\bf{k}}}}}}}^{{\prime} })\cdot {{{{{\boldsymbol{\tau }}}}}}i{\sigma }_{2}]}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}\) with vΓ > 0. Here Γ labels the irreducible representation with \(l=1,2,...,{{{{{\rm{Dim}}}}}}\,{{\Gamma }}\). In this work we focus on 1d representations, i.e. Dim Γ = 1, which already include many interesting cases and are sufficient for the applications discussed in later sections. Due to the possible existence of multiple pairing channels belonging to different representations, each channel has its own critical temperature \({T}_{c}^{\Gamma }\), the largest of which becomes the actual critical temperature of the system. In the weak-coupling theory, \({T}_{c}^{\Gamma }\) follows the standard BCS form and is solely determined by the corresponding pairing interaction vΓ in that particular channel. To the leading order of \({\lambda }_{o}{k}_{F}^{2}/\mu \) (\({k}_{F}=\sqrt{2m\mu }\)), the equation for Tc for the channel Γ reads (see details in the Methods section),

$$\begin{array}{r}\ln \left(\frac{{T}_{c}^{\Gamma }}{{T}_{c0}}\right)={\int}_{\!\!S}d{{\Omega }}\,{{{{{{\mathcal{C}}}}}}}_{0}({T}_{c})\left({\left\vert {{{{{{\bf{d}}}}}}}_{o}^{\Gamma }\right\vert }^{2}-{\left\vert {{{{{{\bf{d}}}}}}}_{o}^{\Gamma }\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}\right\vert }^{2}\right),\end{array}$$
(8)

where Tc0 is the critical temperature for λo = 0, Ω is the solid angle of k, \({\hat{{{{{{\bf{g}}}}}}}}_{o}={{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})/| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| \) are normalized vectors. Here we take \({\int}_{\!\!S}d{{\Omega }}\,| {{{{{{\bf{d}}}}}}}_{o}^{\Gamma }{| }^{2}=1\). And \({{{{{{\mathcal{C}}}}}}}_{0}({T}_{c})=\,{{\mbox{Re}}}\,[{\psi }^{(0)}(\frac{1}{2})-{\psi }^{(0)}(\frac{1}{2}+i\frac{{\lambda }_{o}| {{{{{\bf{g}}}}}}({{{{{\bf{k}}}}}})| }{2\pi {k}_{B}{T}_{c}})]\), where ψ(0)(z) is the digamma function.

We now discuss its implications. In general, the λo-term describes a pair-breaking term, since \({{{{{{\mathcal{C}}}}}}}_{0}({T}_{c})\le 0\) and it monotonically decreases as λo increases, hence the right-hand side of Eq. (8) suppresses Tc in general. However, if we focus on one-dimensional representations, i.e. Dim Γ = 1, it is straightforward to see that \({{{{{{\bf{d}}}}}}}_{o}^{\Gamma }\parallel {{{{{{\bf{g}}}}}}}_{o}\) can lead to Tc = Tc0 for any value of λo, which indicates that the orbital do-vector that is parallel with go is unaffected by the orbital hybridizations. It is worth mentioning that due to the possible suppression of Tc, depending on the relation between \({{{{{{\bf{d}}}}}}}_{o}^{\Gamma }\) and go the leading instability channel at λo = 0 could be suppressed more than some of the other coexisting channels and may eventually become sub-leading. This interesting behavior is discussed further in Supplementary Note 3. For notional simplicity, we will drop the representation index Γ when there is no danger of confusion. Choosing \({{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})=(2{k}_{x}{k}_{y},0,{k}_{x}^{2}-{k}_{y}^{2})\), the numerical results are shown in Fig. 1. The black line confirms that Tc is unaffected as \({\lambda }_{o}{k}_{F}^{2}/{k}_{B}{T}_{c0}\) increases for \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})={k}_{F}^{-2}(2{k}_{x}{k}_{y},0,{k}_{x}^{2}-{k}_{y}^{2})\), which is the unconventional A1g pairing. However, Tc for other do-vectors are severely suppressed. The light-blue line is for \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})=\frac{1}{\sqrt{2}}(1,0,1)\), and the light-orange line for \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})={k}_{F}^{-2}({k}_{x}^{2}-{k}_{y}^{2},0,-2{k}_{x}{k}_{y})\). Therefore, we conclude that the orbital do-vector could exist in SCs with two active orbitals that are not fully degenerate. This is similar to spin-triplet SCs, where the A1g-type spin ds-vector could exist in noncentrosymmetric SCs because dsgs is optimally satisfied4,6.

Fig. 1: Stability of orbital do-vectors vs orbital hybridization λo in Eq. (5).
figure 1

It shows the transition temperature Tc/Tc0 as a function of \({\lambda }_{o}{k}_{F}^{2}/{k}_{B}{T}_{c0}\) for \({{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})=(2{k}_{x}{k}_{y},0,{k}_{x}^{2}-{k}_{y}^{2})\). Tc0 is Tc at λo = 0. The curves from top to bottom correspond to \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})={k}_{F}^{-2}(2{k}_{x}{k}_{y},0,{k}_{x}^{2}-{k}_{y}^{2})\), \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})=\frac{1}{\sqrt{2}}(1,0,1)\), and \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})={k}_{F}^{-2}({k}_{x}^{2}-{k}_{y}^{2},0,-2{k}_{x}{k}_{y})\), respectively.

It is worth mentioning that the results presented above is using a continuum form of the Hamiltonian based on k ⋅ p theory. For real materials, given the interaction on the lattice, the components of the interaction in terms of the basis functions of the representations might not be exactly the same with the form of the vector go. As a result, the parallel condition presented above may not be exactly satisfied. However, the theory developed in this work is generally applicable and the extend to which the parallel condition holds can still be a useful criterion for the most favorable pairing.

Next, we include Δs, and investigate the coupling between Ψs and do. Solving the coupled linearized gap equations up to \({({\lambda }_{o}{k}_{F}^{2}/\mu )}^{2}\) order (see details in Supplementary Note 3), we find that the results from Eq. (8) are still correct. Besides, the magnitude of orbital do-vectors might be determined as \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})={{{\Psi }}}_{s}({{{{{\bf{k}}}}}}){\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\). It implies that Ψs and do belong to the same representation of crystalline groups. Therefore, the stability of orbital do-vector by Eq. (8) indicates the symmetry principle for spin-singlet orbital-triplet pairings.

We now explain Eq. (8) from the band picture. Within the band basis, the pairing potential in the orbital subspace becomes \({\hat{{{\Delta }}}}_{{{{{{\rm{band}}}}}}}({{{{{\bf{k}}}}}})={U}^{{{\dagger}} }({{{{{\bf{k}}}}}})\left[{{{\Delta }}}_{s}{{{\Psi }}}_{s}({{{{{\bf{k}}}}}}){\tau }_{0}+{{{\Delta }}}_{o}({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})\right]U({{{{{\bf{k}}}}}})\), where U(k) is the unitary matrix in the orbital subspace, U(k)[ϵ(k)τ0 + λo(go(k) ⋅ τ)]U(k) = Diag[E+(k), E(k)], with the normal band dispersion

$${E}_{\pm }({{{{{\bf{k}}}}}})=\epsilon ({{{{{\bf{k}}}}}})\pm {\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| .$$
(9)

The intra-orbital pairing naturally gives rise to the intra-band pairing. However, it is different for orbital-dependent pairings. To show that, we decompose the orbital do-vector, \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})={d}_{\parallel }({{{{{\bf{k}}}}}}){\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})+{{{{{{\bf{d}}}}}}}_{\perp }({{{{{\bf{k}}}}}})\), where \({d}_{\parallel }({{{{{\bf{k}}}}}})={{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\) and \({{{{{{\bf{d}}}}}}}_{\perp }({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})=0\). We find that the d-part gives rise to the intra-band pairing, while the d-part leads to the inter-band pairing (see Supplementary Note 4). If the band splitting is much larger than the pairing gap (\({\lambda }_{o}{k}_{F}^{2}\gg {{{\Delta }}}_{o}\)), the inter-band pairing is not energetically favorable in the weak-coupling pairing limit. It means that the inter-band pairing will be severely suppressed if we increase the orbital hybridization λo, consistent with Eq. (8) and results in Fig. 1. Now if we again include the orbital-independent pairing part ΔsΨs(kτ0iσ2), the relation between do and Ψs(k) obtained previously from solving the coupled linearized gap equation (see Supplementary Note 3) can also be reproduced in the band picture by considering the maximization of the condensation energy. The total condensation energy per volume and per spin of the two intra-band pairings is given by

$$\delta E= \,{N}_{+}\mathop{\sum}\limits_{{{{{{\bf{k}}}}}}\in {{{\mbox{FS}}}}_{+}}{\left({{{\Delta }}}_{s}{{{\Psi }}}_{s}({{{{{\bf{k}}}}}})+{{{\Delta }}}_{o}{d}_{\parallel }({{{{{\bf{k}}}}}})\right)}^{2}\\ +{N}_{-}\mathop{\sum}\limits_{{{{{{\bf{k}}}}}}\in {{{\mbox{FS}}}}_{-}}{\left({{{\Delta }}}_{s}{{{\Psi }}}_{s}({{{{{\bf{k}}}}}})-{{{\Delta }}}_{o}{d}_{\parallel }({{{{{\bf{k}}}}}})\right)}^{2},$$
(10)

where N± are the density of states on the two Fermi surfaces (E±). And ΔsΨs(k) ± Δod(k) are the pairing gaps on these two Fermi surfaces. In order to maximize δE, we have d(k) = sign[(N+ − NsΔos(k) (See Supplementary Note 4 for details). Even though the intra-orbital pairing and the orbital-triplet pairing belong to the same symmetry representation, the different k-dependencies of Ψs(k) and d(k) can naturally lead to the anisotropic superconducting gap on the Fermi surface observed in experiments.

Applications to superconductors with/without nodes

As a consequence of the mixing of the orbital-independent pairing (Δs) and orbital-dependent pairing (Δo) discussed in the previous section, there could be a nodal SC. In this section, we apply the results of the previous section to study superconductors with two orbitals, where Δs and Δo coexist. It is shown that the anisotropic gap functions with/without nodes depend on the ratio of Δs and Δo superconducting order parameters. Our weak-coupling theory might have potential applications to some of the nodal/nodeless SCs in the iron-chalcogenides family. For example, the angle-resolved photoemission spectroscopy (ARPES) measurements indicate a nontrivial superconducting gap anisotropy for the monolayer FeSe thin film31. The penetration depth measurements on both LaFePO32 and LiFeP33 show a linear dependence on T, suggesting the presence of superconducting gap nodes.

As an example, we consider the pairing potential in Eq. (2) for monolayer FeSe, where there is no hole pocket around the Γ-point, and a two-spin two-orbital model has been shown to be a good approximation around the electron pockets near the M point of the Brillouin zone (two Fe unit cell). The density functional theory calculations show that there are four bands around the M point, giving rise to only two electron pockets. In the one Fe unit cell, there is one pocket near the X and Y points, respectively. After folding with respect to the unit cell with two Fe, we obtain two pockets around the M point. Considering spin degrees of freedom, it naturally resembles a C4z-invariant two-orbital model34,

$${{{{{{\mathcal{H}}}}}}}_{M}({{{{{\bf{k}}}}}})= \, \left[\epsilon ({{{{{\bf{k}}}}}}){\tau }_{0}+A{k}_{x}{k}_{y}{\tau }_{z}\right]{\sigma }_{0} +{v}_{{{{{{\rm{so}}}}}}}{\tau }_{x}\left[{k}_{x}{\sigma }_{y}+{k}_{y}{\sigma }_{x}\right],$$
(11)

where \(\epsilon ({{{{{\bf{k}}}}}})=({k}_{x}^{2}+{k}_{y}^{2})/(2m)-\mu \) with m > 0 the effective mass, A leads to the anisotropic effective mass (i.e., orbital hybridization), and vso represents SOC that still preserves inversion symmetry. These four states are degenerate at the M point since they form the four-dimensional representation of the space group No. 129 (P4/nmm)35. We take the parameters for the FeSe thin film as μ = 55 meV, 1/(2m) = 1375 meV ⋅ Å2, A = 600 meV ⋅ Å2 and vso≤15 meV ⋅ Å34. The SOC is very weak to open a tiny gap along the kx = 0 and ky = 0 lines, shown in Fig. 2a. As what we expect, it shows two C4z rotational-invariant Fermi surfaces, and the maximal gap, which is induced by the z-component of the go vector, is around 12 meV along the (11) and (1\(\bar{1}\)) directions. This is larger than the typical superconducting gaps in iron-chalcogenide SCs (~4 meV), implying that the effect of the orbital hybridization on the pairing symmetries should not be neglected.

Fig. 2: The application to iron-chalcogenide superconductors with/without linear Dirac nodes.
figure 2

In (a), the two-electron pockets around the M point. For zero spin-orbit coupling, vso = 0, (b) shows the phase diagram as a function of the intra-orbital pairing Δs and the inter-orbital pairing Δo. For the gap parameters represented by the green dot in (b), the nodal superconductor is exhibited in (c), where the eight dark red points represent the chiral symmetry-protected Dirac nodes.

We now use the criteria derived above (Eq. (8)) to examine the superconducting states. Specifically, the weak-coupling criterion indicates that the most favorable pairing to characterize the anisotropic superconducting gap is the A1g-type s-wave pairing symmetry,

$$\hat{{{\Delta }}}({{{{{\bf{k}}}}}})=\left[{{{\Delta }}}_{s}{\tau }_{0}+{{{\Delta }}}_{o}{k}_{x}{k}_{y}{\tau }_{z}\right]\left(i{\sigma }_{2}\right).$$
(12)

The ratio between Δs and Δo determines the superconducting nodal structure. To simplify the analysis, we turn off the weak SOC. In the band basis, the dispersion of \({{{{{{\mathcal{H}}}}}}}_{M}({{{{{\bf{k}}}}}})\) is \({\epsilon }_{\pm }({{{{{\bf{k}}}}}})=({k}_{x}^{2}+{k}_{y}^{2})/(2m)\pm A| {k}_{x}{k}_{y}| -\mu \). Here ± label the band index. Projecting \(\hat{{{\Delta }}}({{{{{\bf{k}}}}}})\) onto the bands leads to Δ± = Δs ± Δokxky∣. Given that Δs, Δo > 0, nodal points can only appear for Δ on the “ − ” band. The nodal condition would be ∣kxky∣ = Δso has solution on the FS given by ϵ(k) = 0. By using the mathematical inequality \({k}_{x}^{2}+{k}_{y}^{2}\le 2| {k}_{x}{k}_{y}| \), it can be shown that the nodal condition is given by,

$$\frac{{{{\Delta }}}_{s}}{{{{\Delta }}}_{o}}\le \frac{\mu }{1/m-A},$$
(13)

which is shown in Fig. 2b. In general, the ratio Δso should depend on both interaction strength in each pairing channel and the orbital hybridization strength. This gives rise to the condition of nodal A1g-type s-wave superconducting states. Therefore, it could not only explain the anisotropic gap functions observed in the FeSe thin film (fully gapped) but also the nodal superconductivity in LaFePO and LiFeP. Around one linear Dirac node, the effective Hamiltonian up to linear-k can be mapped out as

$${{{{{{\mathcal{H}}}}}}}_{D}={k}_{1}{\tilde{\sigma }}_{0}{\tilde{\tau }}_{z}+{k}_{2}{\tilde{\sigma }}_{y}{\tilde{\tau }}_{y},$$
(14)

where k1, k2 are liner combinations of kx and ky. All the other Dirac nodes are related to this one by reflection symmetries. Then, we only need to focus on \({{{{{{\mathcal{H}}}}}}}_{D}\), which is a Dirac Hamiltonian with topological charge (winding number) ± 2, whose node is protected by the chiral symmetry (i.e., the product of time-reversal symmetry and particle-hole symmetry). The \(2{\mathbb{Z}}\) winding number is due to the presence of inversion symmetry and time-reversal symmetry. To analytically show the topology of Dirac nodes, we apply perturbation analysis with respect to PT symmetry (i.e., the product of time-reversal symmetry and inversion symmetry) and Chiral symmetry. Note that the PT symmetry can be also C2zT symmetry for a 2D or quasi-2D SC. The projected symmetry representations are given by \(\,{{\mbox{PT}}}\,={\tilde{\sigma }}_{y}{\tilde{\tau }}_{0}\) and \({{{{{\mathcal{C}}}}}}={\tilde{\sigma }}_{y}{\tilde{\tau }}_{x}\). As expected, the PT symmetry commutes with \({{{{{{\mathcal{H}}}}}}}_{D}\), while the Chiral symmetry anti-commutes with \({{{{{{\mathcal{H}}}}}}}_{D}\). Then, local perturbations preserving PT and Chiral are

$${{{{{{\mathcal{H}}}}}}}_{D}^{{\prime} }={m}_{1}{\tilde{\sigma }}_{0}{\tilde{\tau }}_{y}+{m}_{2}{\tilde{\sigma }}_{y}{\tilde{\tau }}_{z},$$
(15)

where m1 and m2 represent perturbation strengths or mass terms. The spectrum of \({{{{{{\mathcal{H}}}}}}}_{D}+{{{{{{\mathcal{H}}}}}}}_{D}^{{\prime} }\) are given by

$$E=\pm \sqrt{{k}_{1}^{2}+{k}_{2}^{2}+{m}_{1}^{2}+{m}_{2}^{2}\pm 2| {m}_{1}{k}_{2}+{m}_{2}{k}_{1}| },$$
(16)

which indicates that the Dirac nodes are movable but not removable. For example, k1 = 510.7kx + 76.5ky and k2 = − 14.7kx − 40.9ky around one Dirac node. Then, turning on the SOC vso = 15 meV ⋅ Å, we numerically confirm the nodal SC phase with Δs = 3 meV and Δo = 200 meV ⋅ Å2, shown in Fig. 2c, where the logarithm of superconducting gaps are plotted. The eight dark red points are the linear Dirac nodes. Based on the topology-protection argument, the interplay between intra- and inter-orbital pairings for nodal superconductivity is robust against local perturbations. Note that our results are different from a previous work34, in which the d-wave pairing symmetry induced nodal SC. In experiments, the nodal gap structure could be detected by measuring the temperature dependence of physical quantities like specific heat and penetration depth at low temperatures. A power law dependence usually indicates the existence of nodal structures (point nodes or line nodes), whereas exponential dependence implies the SC is fully gapped3.

Applications to superconductors with nematic order

In addition to the crystal field splitting, the many-body electron-electron interactions may also lead to orbital hybridization, such as the nematic ordering in the normal states (See Supplementary Note 5 for details). The rotational symmetry reduction could either be from interaction-induced spontaneous symmetry breaking or from explicit symmetry breaking from, say, adding external strain. Then the natural question to ask is whether it is still possible to have an orbital-dependent pairing order characterized by some do-vector. Interestingly, we find that the orbital-dependent pairing can coexist with the electronic nematic ordering as long as do is parallel to the gtot, which is an effective orbital-hybridization vector that also contains the nematic order. This establishes a deep connection between SCs with nematic order and spin-singlet orbital-triplet pairings. In the following, we study two typical examples.

  • For case A [two-orbital system], we apply the theory to fit the anisotropic superconducting gap of the hole pocket in the bulk FeSe measured by the quasiparticle interference imaging12.

  • For case B [two-valley system], we use a toy model to demonstrate the possible existence of s + d-like nematic nodal superconductor in two-valley systems on a honeycomb lattice. We also show the transition between U-shaped and V-shaped quasi-particle density-of-state by tuning the chemical potential.

Case A: two-orbital model for the bulk FeSe SC

We discussed the possible anisotropic A1g-type s-wave pairing states for the C4-symmetric iron-chalcogenide SCs including fully gapped FeSe thin film and nodal SC in LiFeP and LaFePO. Here we investigate the C4-breaking nematic SC in bulk FeSe. Let us revisit the iron-based SC with a well-established nematic ordering. We consider \({{{{{{\mathcal{H}}}}}}}_{{{{{{\rm{int}}}}}}}={v}_{1}{\hat{n}}_{1}({{{{{\bf{r}}}}}}){\hat{n}}_{2}({{{{{\bf{r}}}}}})\), where \({\hat{n}}_{i}\) is electron density operator for the i-atomic orbital. If \(\langle {\hat{n}}_{1}\rangle \ne \langle {\hat{n}}_{2}\rangle \), Cn (n > 2) is spontaneously broken down to C2 and we have the nematic order. The intra-orbital interaction does not alter the mean-field results for nematic orders (See Supplementary Note 5). The total inter-orbital hybridization contains two parts,

$${{{{{{\bf{g}}}}}}}_{{{{{{\rm{tot}}}}}}}({{{{{\bf{k}}}}}})={{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})+{{{{{{\bf{g}}}}}}}_{{{{{{\rm{nem}}}}}}},$$
(17)

where go(k) is caused by the crystal field splitting and gnem = (0, 0, Φ) is induced due to the nematicity \({{\Phi }}={v}_{1}\langle {\hat{n}}_{1}-{\hat{n}}_{2}\rangle \), which is momentum-independent if translation symmetry is to be preserved. Hereafter, we focus on the hole pockets around the Γ point to fit the experimental data of superconducting gap functions12. We will see that even this simplified weak-coupling model, where the coupling between the hole pockets at the Γ point and the electron pockets at the M point is ignored, can produce a descent fit the experimental data. A similar result is expected for the electron pockets near the M point. Replacing go with gtot in Eq. (5), we can still use Eq. (8) to investigate the interplay between superconductivity and nematic order, thus the orbital do-vector satisfying dogtot leads to the nematic superconductivity. Thus, it generally shows the A1g-type s-wave spin-singlet orbital-triplet pairings in nematic SCs.

This scenario can be adopted to study the quasi-two dimensional bulk FeSe, where superconductivity (Tc ~ 8 K) emerges inside a well-developed nematic phase (transition temperature Tnem ~ 90 K36), shown in Fig. 3a. For a minimal two-band model37 for the bulk FeSe with {dxz, dyz}-orbitals, \({{{{{{\bf{g}}}}}}}_{o}=(2{k}_{x}{k}_{y},0,{k}_{x}^{2}-{k}_{y}^{2})\) and gnem = (0, 0, Φ)38,39. Therefore, the nematic orbital do-vector with dogtot breaks C4 (see Supplementary Note 5 for more details). The projected pairing gap function on the large Fermi surface is given by

$${{{\Delta }}}_{{{{{{\rm{FS}}}}}}}({{{{{\bf{k}}}}}})={{{\Delta }}}_{s}+{{{\Delta }}}_{o}\sqrt{{(-{\lambda }_{o}({k}_{x}^{2}-{k}_{y}^{2})+{{\Phi }})}^{2}+{(2{\lambda }_{o}{k}_{x}{k}_{y})}^{2}}.$$
(18)

If Φ = 0, ΔFS(k) is reduced to \({{{\Delta }}}_{s}+{{{\Delta }}}_{o}| {\lambda }_{o}| ({k}_{x}^{2}+{k}_{y}^{2})\) that is in the isotropic limit. The presence of Φ is the driving force for the anisotropy of ΔFS(k). When the nematicity Φ is strong enough, the orbital do-vector will be pinned along the z-axis, resulting in the so-called orbital-selective pairing states. We adopt the realistic parameters for the bulk FeSe SC from Ref. 39 to calculate the superconducting gap measured by the quasiparticle interference imaging12. In Fig. 3b, we show the angular dependence of the pairing gap around the hole pocket at the Γ-point of FeSe in the presence of nematic order. Our theory provides an equally decent fit to recent experimental data12 as the intra-orbital s + d-pairing theory proposed by Kang et al.39, even though our work uses a simplified model without considering the coupling to the other two electron pockets. Our theory shows more clearly the role of nematic order on the pairing symmetries. Therefore, the theory developed in this work may alternatively explain the experimental evidence of orbital-selective pairings of the FeSe SC in refs. 12,27, and reveal a deep connection between nematic SC and spin-singlet orbital-triplet pairings. It has to be mentioned that here we only focused on the hole pockets around the Γ point and discussed the nematicity-induced gap anisotropy around the hole FS. There are other possible mechanisms for gap anisotropy in Fe-based SCs. For example, a previous work40 discussed, among other things, the anisotropy/isotropy of the SC gap around the electron pockets at the M point, where the degree of anisotropy depends on the J1-J2 magnetic frustration in the proposed five-orbital t-J1-J2 microscopic model.

Fig. 3: The application to bulk FeSe superconductors with nematicity.
figure 3

a Schematic phase diagram vs temperature T for normal metal (T > Tnem), nematic metal (T < Tnem), and nematic superconductivity (T < Tc). b Angular dependence of the superconducting pairing gap: comparison between experiment data (black dots) by Sprau et al. ref. 12, our theory (red line) and the theory proposed by Kang et al.39 (blue line). Fitting parameters used for our model: Δs = 2.6, Δo = − 0.055 in Eq. (18). All the other parameters used are the same39, including the chemical potential, effective mass, orbital hybridization, and nematic order.

Case B: two-valley system superconductivity

Similar to the two-orbital systems considered above, we discuss in this section superconductivity in two-valley systems, like single layer graphene SC41 or transition metal dichalcogenide (TMD)42, where the pairing can be between opposite valleys K±. The spin-singlet pairing is merely characterized by the orbital do-vector with Δs = 0 in Eq. (2). For the single-particle Hamiltonian, the inter-valley hopping is naturally forbidden by translational symmetry, namely, λo = 0 in Eq. (5). Then, we consider the inter-valley scattering Hamiltonian, \({{{{{{\mathcal{H}}}}}}}_{{{{{{\rm{int}}}}}}}={\sum }_{{{{{{\bf{k}}}}}},{{{{{{\bf{k}}}}}}}^{{\prime} },\sigma }V({{{{{\bf{k}}}}}}-{{{{{{\bf{k}}}}}}}^{{\prime} }){c}_{+,\sigma }^{{{\dagger}} }({{{{{\bf{k}}}}}}){c}_{+,\sigma }({{{{{{\bf{k}}}}}}}^{{\prime} }){c}_{-,\sigma }^{{{\dagger}} }({{{{{{\bf{k}}}}}}}^{{\prime} }){c}_{-,\sigma }({{{{{\bf{k}}}}}})\). It generates the inter-valley coupling gint by defining the order parameter \({{\Phi }}({{{{{\bf{k}}}}}})={\sum }_{{{{{{{\bf{k}}}}}}}^{{\prime} },\sigma }V({{{{{\bf{k}}}}}}-{{{{{{\bf{k}}}}}}}^{{\prime} })\langle {c}_{+,\sigma }({{{{{{\bf{k}}}}}}}^{{\prime} }){c}_{-,\sigma }^{{{\dagger}} }({{{{{{\bf{k}}}}}}}^{{\prime} })\rangle \) that spontaneously breaks the translational symmetry,

$${{{{{{\bf{g}}}}}}}_{{{{{{\rm{int}}}}}}}({{{{{\bf{k}}}}}})=({g}_{{{{{{\rm{int}}}}}},1}({{{{{\bf{k}}}}}}),{g}_{{{{{{\rm{int}}}}}},2}({{{{{\bf{k}}}}}}),0),$$
(19)

where gint,1(k) = Re[Φ(k)] and gint,2(k) = − Im[Φ(k)]. In this case, TRS is \({{{{{\mathcal{T}}}}}}=i{\tau }_{1}{\sigma }_{2}{{{{{\mathcal{K}}}}}}\) and IS is \({{{{{\mathcal{I}}}}}}={\tau }_{1}{\sigma }_{0}\). The do-vector is manifested as do = (d1(k), id2(k), 0) with d1(k) = d1( − k) and d2(k) = − d2( − k). Both d1(k) and d2(k) are real to preserve TRS. As for the interaction-induced gint, \({{{{{\mathcal{T}}}}}}\) and \({{{{{\mathcal{I}}}}}}\) require gint,1(k) = gint,1( − k) and gint,2(k) = 0. By symmetry, there are two general possibilities. One is gint,1(k) = 1, so \({C}_{3}\times {{{{{\mathcal{I}}}}}}={C}_{6}\) is preserved, and it describes the charge-density-wave order43,44. The other one is \({g}_{{{{{{\rm{int}}}}}},1}({{{{{\bf{k}}}}}})\in \{{k}_{x}{k}_{y},{k}_{x}^{2}-{k}_{y}^{2}\}\) that spontaneously breaks C6 down to C2, forming a nematic order. This is experimentally possible for the strain-induced Kekul’e distortion (i.e., \(\sqrt{3}\times \sqrt{3}\) type).

We next discuss superconductivity in the presence of inter-valley couplings, by replacing the go-vector in Eq. (5) with the interaction-induced gint. As a result, Eq. (8) is still applicable. It is similar to a recent work45 where the charge order coexists with a sublattice-selective non-unitary pairing state.

The nematic inter-valley coupling is represented as \({g}_{{{{{{\rm{int}}}}}},1}({{{{{\bf{k}}}}}})=1+2{t}_{1}{k}_{x}{k}_{y}+{t}_{2}({k}_{x}^{2}-{k}_{y}^{2})\), which requires that \({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})=(1+2{t}_{1}{k}_{x}{k}_{y}+{t}_{2}({k}_{x}^{2}-{k}_{y}^{2}),0,0)\) (see Supplementary Note 6). Here the normalization factor has been dropped without changing the essential physics. The system is fully gapped if the s-wave gap is dominant (\(1\gg \sqrt{{t}_{1}^{2}+{t}_{2}^{2}}\)), otherwise, it is a d-wave dominant nodal SC (\(1\ll \sqrt{{t}_{1}^{2}+{t}_{2}^{2}}\)).

As a concrete toy model, we look at superconductivity on a generic honeycomb lattice with two valleys K±, with the Hamiltonian around the two valleys given by,

$${{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})=\epsilon ({{{{{\bf{k}}}}}}){\tau }_{0}{\sigma }_{0}+\alpha \left({k}_{x}^{3}-3{k}_{x}{k}_{y}^{2}\right){\tau }_{3}{\sigma }_{0},$$
(20)

where the two-valley basis used here is given by \({\psi }_{{{{{{\bf{k}}}}}}}^{{{\dagger}} }=({c}_{{K}_{+},\uparrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{{K}_{+},\downarrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{{K}_{-},\uparrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{{K}_{-},\downarrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}))\) and ϵ(k) takes the same form as in Eq. (5). The parameter α determines the C3 anisotropy of the continuum model around each valley. This Hamiltonian was used as an effective model46 to study twisted bilayer graphene.

Including the inter-valley scattering effects, the one-band model is given by

$${{{{{\mathcal{H}}}}}}({{{{{\bf{k}}}}}})= \,\epsilon ({{{{{\bf{k}}}}}}){\tau }_{0}{\sigma }_{0}+\alpha \left({k}_{x}^{3}-3{k}_{x}{k}_{y}^{2}\right){\tau }_{3}{\sigma }_{0} +{\lambda }_{{{{{{\rm{int}}}}}}}[{{{{{{\bf{g}}}}}}}_{{{{{{\rm{int}}}}}}}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}]{\sigma }_{0},$$
(21)

where the λint determines the strength of the inter-valley scattering. In Fig. 4, we present representative numerical results for Eq. (21). Panels (a–c) illustrate Fermi surfaces with varying parameters, while panels (d–f) depict the corresponding quasi-particle density of states (DOS).

Fig. 4: Fermi surfaces (FSs) at the K± valleys and the quasi-particle density of states (DOS).
figure 4

Panels (ac) show FSs with varying parameters, while panels (df) exhibit the corresponding quasi-particle DOS. The C6 symmetric FS without inter-valley scattering is shown in (a) and its DOS with an isotropic s-wave pairing is given in (d). b Shows C6-breaking FSs due to the inter-valley scattering, together with the nodal lines of nematic pairing. There are no nodes on the FSs and the corresponding DOS is shown in (e). c Is similar to (b) but with chemical potential μ adjusted so that the nodal lines intersect the FSs, hence a V-shaped DOS is obtained as in (f). Parameters used are the following, the C3 anisotropy α = 0.2, the coefficients for basis functions t1 = 0.15, t2 = 0.25, the orbitial-dependent pairing gap Δo = 0.5. For (a) and (d) μ = 1.5 (chemical potential), the inter-valley coupling gint = 0; for (b) and (e) μ = 1.5, gint = 0.7; for (c) and (f) μ = 2.7, gint = 0.7.

In the absence of inter-valley scattering (λint = 0), the Fermi surfaces (FSs) around the two K± valleys are plotted in Fig. 4a. As expected, with a fully symmetric s-wave pairing characterized by do = (1, 0, 0), a fully gapped or U-shaped quasi-particle density-of-states (DOS) is obtained and shown in Fig. 4d. Then we include the aforementioned inter-valley scattering gint that breaks C6 down to C2. As a result, our theory implies that the effective nematicity generated will favor a nematic pairing characterized by dogint. Consider the generic form \({{{{{{\bf{d}}}}}}}_{o}={{{{{{\bf{g}}}}}}}_{{{{{{\rm{int}}}}}}}=(1+2{t}_{1}{k}_{x}{k}_{y}+{t}_{2}({k}_{x}^{2}-{k}_{y}^{2}),0,0)\), the resulting C6-breaking FS are shown in Fig. 4b, c, where the nodal lines of the pairing are also shown. By tuning the chemical potential μ, the FSs and nodal lines can go from non-intersecting in Fig. 4b to intersecting in Fig. 4c, leading to the corresponding evolution from the gapped U-shaped DOS in Fig. 4e to the gapless V-shaped DOS in Fig. 4f. Our results may explain the experimental observations in magic-angle twist bilayer graphene that reports the nematic order47, and V-shaped DOS48 at the specific doping level.

Discussions

We briefly discuss the difference between our theory and the previous studies21 for nematic SCs. One example is a pairing state belonging to a 2D irreducible representation (Irrep), e.g., the E-pairing in Cu or Nb-doped Bi2Se349,50 and UPt351,52. A real order parameter vector (ΔE,1, ΔE,2) spontaneously breaks C3, leading to nematic superconductivity. Alternatively, a nematic SC can be formed by mixing two different 1D-Irrep-pairing channels. In FeSe53,54, the nematic order breaks the C4 down to C2, which mixes the s-wave and d-wave pairing channels. However, Tc of the (s + d) orbital-independent pairing state could be generally affected by increasing nematicity, because of the significant change in the density of states at the Fermi energy. In our theory, the (s + d)-like nematic do-vector coexists with the nematic order, so Tc is almost unaffected by increasing nematicity. Therefore, it may help to distinguish our results from previous proposals in experiments, where one may use the chemical or physical pressures to tune the nematicity and measure Tc as a function of pressure55. Nevertheless, more efforts are necessary to test the results established in this work for nematic SCs.

In addition to the above discussions for the time-reversal-invariant superconducting states, we also comment on the effects of the spontaneous TRS-breaking, where a complex orbital do-vector generates the orbital orderings as Mo = − iγ1/αM(d × d*), of which only the y-component breaks TRS (see details in Supplementary Note 7), as illustrated in Fig. 5a. Alternatively, the corresponding quasi-particle spectrum in Fig. 5b shows the two distinct gaps, similar to the range given by Eq. (6). More explicitly, we schematically plot the atomic orbital-polarized density of states (DOS) by defining \(\left\vert \pm \right\rangle =\left\vert 1\right\rangle +i\left\vert 2\right\rangle \) for complex orbitals, where D+ ≠ D at finite energy clearly indicates that the DOS is orbital-polarized, which is consistent with the Ginzburg-Landau theory, shown in Supplementary Note 2. Moreover, we also find that the orbital-spin conversion would lead to the spin-polarized DOS56.

Fig. 5: Schematic diagrams for the time-reversal symmetry (TRS) breaking effects.
figure 5

a, b Are for orbital-triplet superconductors (SCs), while (c) and (d) are for spin-triplet SCs. As for orbital-triplet SCs characterized by a do-vector, (a) shows a complex orbital do-vector that spontaneously breaks TRS and results in the TRS-breaking orbital-polarization \({{{{{{\bf{M}}}}}}}_{o}\propto i{{{{{{\bf{d}}}}}}}_{o}\times {{{{{{\bf{d}}}}}}}_{o}^{* }\); and (b) shows the quasi-particle spectrum along kx and the orbital-polarized density of states (DOS) D± with \(\left\vert \pm \right\rangle \) representing \(\left\vert 1\right\rangle \pm i\left\vert 2\right\rangle \). As a comparison, in spin-triplet SCs, (c) shows the superconductivity-induced spontaneous spin-polarization \({{{{{{\bf{M}}}}}}}_{s}\propto i{{{{{{\bf{d}}}}}}}_{s}\times {{{{{{\bf{d}}}}}}}_{s}^{* }\); and (d) shows the two distinct gaps of the quasi-particle spectrum along kx and the spin-polarized DOS Dσ with σ = {, }. The gapped spectrum is plotted for kz ≠ 0 and the node in DOS profile is due to the nodal line at kz = 0.

The above result for orbital-triplet pairings is similar to the superconducting gaps for non-unitary spin-triplet SCs3. By symmetry, the Gizburg-Landau free energy is the same. To show the similarity, for the single-band spin-triplet SCs57, the spin-triplet pairing potential is generally given by \(\hat{{{\Delta }}}({{{{{\bf{k}}}}}})={{{\Delta }}}_{0}[{{{{{{\bf{d}}}}}}}_{s}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\sigma }}}}}}](i{\sigma }_{2})\), where Δ0 is the pairing strength and σ are Pauli matrices in the spin subspace. Due to the Fermi statistics, the spin ds(k)-vector has to satisfy ds(k) = − ds( − k). The ds-vector formalism is firstly developed for He3 superfluid58. And it also occurs in noncentrosymmetric SCs, the spin ds(k)-vector is usually pinned along a certain crystal axis since superconductivity is non-suppressed only for ds(k)∥gs(k), where gs(k) represents the Rashba spin-orbit coupling (SOC)4,6. Besides, there is intrinsic spontaneous spin-polarization induced by the non-unitary pairing, ds(k) = kz(1, − iη0, 0) with real η0. Fig. 5c shows the spin expectation value of the Cooper pairs (\({{{{{{\bf{M}}}}}}}_{s}\propto i{{{{{{\bf{d}}}}}}}_{s}^{* }({{{{{\bf{k}}}}}})\times {{{{{{\bf{d}}}}}}}_{s}({{{{{\bf{k}}}}}})=2{\eta }_{0}{k}_{z}^{2}{\overrightarrow{e}}_{z}\)). It is an equal-spin pairing so that σ3 is conserved, and non-zero Ms leads to two distinct superconducting gaps of the quasi-particle spectrum59, shown in Fig. 5d. In addition, the density of states (DOS) is spin-polarized, namely, D ≠ D at finite energy ω, as illustrated in Fig. 5d.

To summarize, we have derived a general weak-coupling criterion to investigate the spin-singlet orbital-triplet pairings in nematic SCs. For technical convenience, we adopt the orbital do-vector to describe the spin-singlet orbital-dependent pairing states and the go-vector for the orbital hybridizations. The main results of this work include, first, we demonstrate that an orbital do-vector that is parallel with go-vector for orbital hybridizations is possible to be realized in real superconducting materials. Second, the interplay between intra-orbital and orbital-dependent pairings that belong to the same symmetry representation can explain the observation of robust Dirac nodes in the quasi-2D iron-based SCs. Remarkably, we find that do-vectors could even coexist with many-body interaction-induced nematic orders or charge-density-wave orders when do ∝ gtot = go + gnem (or gint). Moreover, our theory discovers the important role of nematic orders in SC pairing symmetry, which builds a possible bridge between repulsive interaction-induced nematic orders and nematic superconductivity and also reveals a deep connection between spin-singlet orbital-triplet pairings in nematic SCs. Our results may be helpful in understanding the nematic superconductivity in bulk FeSe. Our work will motivate more theoretical and experimental efforts to search for spin-singlet orbital-triplet SCs, even for topological superconductivity, which might contribute to further understanding the effects of spontaneous symmetry breaking on superconductivity.

Methods

Here we present the derivation for the main result Eq. (8), which is first order in λo, by solving the linearized gap equation. The second-order results are presented in Supplementary Note 3. The general k ⋅ p normal Hamiltonian considered in the main text reads,

$${{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})=\epsilon ({{{{{\bf{k}}}}}}){\tau }_{0}{\sigma }_{0}+{\lambda }_{o}({{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){\sigma }_{0},$$
(22)

where the electronic basis is made of {1, 2}-orbitals \({{{\Psi }}}_{{{{{{\bf{k}}}}}}}^{{{\dagger}} }=({c}_{1,\uparrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{1,\downarrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{2,\uparrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}),{c}_{2,\downarrow }^{{{\dagger}} }({{{{{\bf{k}}}}}}))\), \(\epsilon ({{{{{\bf{k}}}}}})=({k}_{x}^{2}+{k}_{y}^{2})/2m-\mu \) is the band energy measured relative to the chemical potential μ, λo represents the orbital hybridization and go(k) = (g1(k), g2(k), g3(k)). The TRS \({{{{{\mathcal{T}}}}}}=i{\sigma }_{2}{\tau }_{0}{{{{{\mathcal{K}}}}}}\) requires g1,3(k) = g1,3( − k) and g2(k) = − g2( − k). It leads that

$${{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}={\left[{{{{{{\bf{g}}}}}}}_{o}(-{{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}\right]}^{* }.$$
(23)

Besides, we set λo > 0 without loss of generality. The Matsubara Green’s function for electrons is \({G}_{e}({{{{{\bf{k}}}}}},i{\omega }_{n})={[i{\omega }_{n}-{{{{{{\mathcal{H}}}}}}}_{0}({{{{{\bf{k}}}}}})]}^{-1}\) and that for holes is \({G}_{h}({{{{{\bf{k}}}}}},i{\omega }_{n})=-{G}_{e}^{* }({{{{{\bf{k}}}}}},i{\omega }_{n})\). Here β = 1/kBT and ωn = (2n + 1)π/β with n integer. Therefore,

$${G}_{e}({{{{{\bf{k}}}}}},i{\omega }_{n})= \frac{{{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})}{i{\omega }_{n}-\epsilon ({{{{{\bf{k}}}}}})+{\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| }+\frac{{{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}})}{i{\omega }_{n}-\epsilon ({{{{{\bf{k}}}}}})-{\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| }\\ \triangleq {G}_{e}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n}){{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})+{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}}),$$
(24)
$${G}_{h}(-{{{{{\bf{k}}}}}},i{\omega }_{n})= \frac{{{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})}{i{\omega }_{n}+\epsilon ({{{{{\bf{k}}}}}})-{\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| }+\frac{{{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}})}{i{\omega }_{n}+\epsilon ({{{{{\bf{k}}}}}})+{\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| }\\ \triangleq {G}_{h}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n}){{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})+{G}_{h}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}}),$$
(25)

where \({{{{{{\mathcal{P}}}}}}}_{\pm }({{{{{\bf{k}}}}}})=\frac{1}{2}(1\pm {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})\) with \({\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})={{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})/| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| \). Here \({G}_{e}^{\pm }({{{{{\bf{k}}}}}},i{\omega }_{n})=\frac{1}{i{\omega }_{n}-\epsilon ({{{{{\bf{k}}}}}})\mp {\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| }\) and \({G}_{h}^{\pm }({{{{{\bf{k}}}}}},i{\omega }_{n})=\frac{1}{i{\omega }_{n}+\epsilon ({{{{{\bf{k}}}}}})\pm {\lambda }_{o}| {{{{{{\bf{g}}}}}}}_{o}({{{{{\bf{k}}}}}})| }\). We expand the attractive interactions as

$${V}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}^{{s}_{1}a,{s}_{2}b}({{{{{\bf{k}}}}}},{{{{{{\bf{k}}}}}}}^{{\prime} })= -\mathop{\sum}\limits_{\Gamma ,l}{v}_{0}^{\Gamma }{[{{{{{{\bf{d}}}}}}}_{o}^{\Gamma ,l}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}i{\sigma }_{2}]}_{{s}_{1}a,{s}_{2}b}\\ \times {[{{{{{{\bf{d}}}}}}}_{o}^{\Gamma ,l}({{{{{{\bf{k}}}}}}}^{{\prime} })\cdot {{{{{\boldsymbol{\tau }}}}}}i{\sigma }_{2}]}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }},$$
(26)

where \({v}_{0}^{\Gamma } > 0\) is the interaction strength of the irreducible representation channel Γ of the crystalline group, and \(l=1,2,...,{{{{{\rm{Dim}}}}}}\,{{\Gamma }}\). Each pairing channel Γ gives rise to an SC critical temperature \({T}_{c}^{\Gamma }\), and the actual transition temperature of the system is given by the largest of these critical temperatures. In our work, we mainly focus on the case where \({{{{{\rm{Dim}}}}}}\,{{\Gamma }}=1\), which is sufficient for the applications discussed in the main text. The coupling between orbital-dependent pairings and orbital-independent pairings will be discussed in detail later. The transition temperature \({T}_{c}^{\Gamma }\) of orbital-dependent pairing channels is calculated by solving the linearized gap equation,

$${{{\Delta }}}_{{s}_{1},{s}_{2}}^{a,b}({{{{{\bf{k}}}}}})= -\frac{1}{\beta }\mathop{\sum}\limits_{{\omega }_{n}}\mathop{\sum}\limits_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}{V}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }}^{{s}_{1}a,{s}_{2}b}({{{{{\bf{k}}}}}},{{{{{{\bf{k}}}}}}}^{{\prime} })\\ \times {\left[{G}_{e}({{{{{{\bf{k}}}}}}}^{{\prime} },i{\omega }_{n})\hat{{{\Delta }}}({{{{{{\bf{k}}}}}}}^{{\prime} }){G}_{h}(-{{{{{{\bf{k}}}}}}}^{{\prime} },i{\omega }_{n})\right]}_{{s}_{1}^{{\prime} }{a}^{{\prime} },{s}_{2}^{{\prime} }{b}^{{\prime} }},$$
(27)

which is reduced to \({v}_{0}^{\Gamma }{\chi }^{\Gamma }(T)-1=0\) with the superconductivity susceptibility χΓ(T) in the channel Γ defined as,

$${\chi }^{\Gamma }(T)=-\frac{1}{\beta }\mathop{\sum}\limits_{{{{{{\bf{k}}}}}},{\omega }_{n}}\,{{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}^{\Gamma }({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}i{\sigma }_{2})}^{{{\dagger}} }{G}_{e}({{{{{\bf{k}}}}}},i{\omega }_{n})({{{{{{\bf{d}}}}}}}_{o}^{\Gamma }({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}i{\sigma }_{2}){G}_{h}(-{{{{{\bf{k}}}}}},i{\omega }_{n})\right],$$
(28)
$$= -\frac{2}{\beta }\mathop{\sum}\limits_{{{{{{\bf{k}}}}}},{\omega }_{n}}\mathop{\sum}\limits_{\alpha ,\beta }{G}_{e}^{\alpha }({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{\beta }({{{{{\bf{k}}}}}},i{\omega }_{n})\\ \times \,{{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}^{\Gamma }({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})}^{{{\dagger}} }{{{{{{\mathcal{P}}}}}}}_{\alpha }({{{{{\bf{k}}}}}})({{{{{{\bf{d}}}}}}}_{o}^{\Gamma }({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){{{{{{\mathcal{P}}}}}}}_{\beta }({{{{{\bf{k}}}}}})\right],$$
(29)

where α, β ∈ { + , − }. For notional simplicity, the superscript Γ will be dropped when there is no danger of confusion. Firstly, let us calculate the trace part. In the following calculation, we will use

$$ {{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})}^{{{\dagger}} }{{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}})({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}})\right]\\ +\,{{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})}^{{{\dagger}} }{{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})\right]\\ =\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)+2\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right.\\ \left.\;\;\;\;\;\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right).$$
(30)

And,

$$ {{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})}^{{{\dagger}} }{{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}})({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})\right] \\ +\,{{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})}^{{{\dagger}} }{{{{{{\mathcal{P}}}}}}}_{-}({{{{{\bf{k}}}}}})({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){{{{{{\mathcal{P}}}}}}}_{+}({{{{{\bf{k}}}}}})\right]\\ = \left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)-2\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\\ \;\;\;\;+\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right).$$
(31)

Therefore, we arrive at

$$ {{\mbox{Tr}}}\,\left[{({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}})}^{{{\dagger}} }{{{{{{\mathcal{P}}}}}}}_{\alpha }({{{{{\bf{k}}}}}})({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {{{{{\boldsymbol{\tau }}}}}}){{{{{{\mathcal{P}}}}}}}_{\beta }({{{{{\bf{k}}}}}})\right]\\ = \frac{1}{2}\left[\right.\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)+i\alpha \left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot \left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\times {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right)\\ +i\beta \left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot \left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\times {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right)\\ + \alpha \beta \left(2\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right.\\ \left.-\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right)\left]\right..$$
(32)

Then we have

$$\chi (T) = -\frac{1}{\beta }\mathop{\sum}\limits_{{{{{{\bf{k}}}}}},{\omega }_{n}}\mathop{\sum}\limits_{\alpha ,\beta }{G}_{e}^{\alpha }({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{\beta }({{{{{\bf{k}}}}}},i{\omega }_{n})\left[\right.\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\\ +i\alpha \left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot \left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\times {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right)+i\beta \left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot \left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\times {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right)\\ + \alpha \beta \left(2\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right.\\ \left.-\left({{{{{{\bf{d}}}}}}}_{o}^{* }({{{{{\bf{k}}}}}})\cdot {{{{{{\bf{d}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\left({\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}({{{{{\bf{k}}}}}})\right)\right)\left]\right..$$
(33)

Next, we calculate the integration for \({\sum }_{{{{{{\bf{k}}}}}},{\omega }_{n}}\) by using,

$$\mathop{\sum}\limits_{{{{{{\bf{k}}}}}},{\omega }_{n}}\to \frac{{N}_{0}}{4}\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon {\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }\,\mathop{\sum}\limits_{{\omega }_{n}},$$
(34)

where N0 is the density of states at Fermi surface and Ω is the solid angle of k on Fermi surfaces. Then,

$$ -\frac{{N}_{0}}{\beta }\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon {\int}_{\!\!S}\frac{d{{\Omega }}}{2\pi }\mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ =-\frac{{N}_{0}}{\beta }\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon {\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }\mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ \equiv {\chi }_{0}(T),$$
(35)

On one hand,

$$\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}} d\epsilon \mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ =\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon \mathop{\sum}\limits_{{\omega }_{n}}\frac{1}{i{\omega }_{n}+\epsilon }\frac{1}{i{\omega }_{n}-\epsilon }\\ =\beta \int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon \frac{\tanh \frac{\beta \epsilon }{2}}{2\epsilon }\\ =\beta \int\nolimits_{0}^{\beta {\omega }_{D}/2}dx\frac{\tanh x}{x}\approx \beta \ln \left(\frac{2{e}^{\gamma }{\omega }_{D}}{\pi {k}_{B}T}\right),$$
(36)

where the approximation is done at low temperature when β → .

On the other hand, we could find a series representation for χ0, which also applies to the case where λo ≠ 0, so that χ0 ≡ χ(λo = 0) and χ(λo ≠ 0) can be related by a simple relation. The way to do it is to perform the integration in ϵ first. More precisely,

$$ \int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon \mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ =\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}\mathop{\sum}\limits_{{\omega }_{n}}\frac{1}{i{\omega }_{n}+\epsilon }\frac{1}{i{\omega }_{n}-\epsilon }\\ =2{{{{{\rm{Re}}}}}}\mathop{\sum}\limits_{n\ge 0}\int\nolimits_{-{\omega }_{D}}^{{\omega }_{D}}d\epsilon \frac{1}{i{\omega }_{n}+\epsilon }\frac{1}{i{\omega }_{n}-\epsilon }\\ =2\beta {{{{{\rm{Re}}}}}}\mathop{\sum}\limits_{n\ge 0}\int\nolimits_{-\beta {\omega }_{D}}^{\beta {\omega }_{D}}d\epsilon \frac{1}{i2\pi (n+1/2)+\epsilon }\frac{1}{i2\pi (n+1/2)-\epsilon }\\ \approx 2\beta {{{{{\rm{Re}}}}}}\mathop{\sum}\limits_{n\ge 0}\int\nolimits_{-\infty }^{\infty }d\epsilon \frac{1}{i2\pi (n+1/2)+\epsilon }\frac{1}{i2\pi (n+1/2)-\epsilon }\\ =\beta {{{{{\rm{Re}}}}}}\mathop{\sum}\limits_{n\ge 0}\frac{1}{n+1/2},$$
(37)

where the low-temperature limit is again assumed and the integration is done using the residue theorem. In the same spirit, we have,

$$ \int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}} d\epsilon \mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ =\beta {{{{{\rm{Re}}}}}}\mathop{\sum}\limits_{n\ge 0}\frac{1}{n+1/2+i\frac{{\lambda }_{o}| {{{{{\bf{{g}}}}}_{o}}}({{{{{\bf{k}}}}}})| }{2\pi {k}_{B}T}},$$
(38)

Now by introducing the digamma function defined on the complex plane,

$${\psi }^{(0)}(z)=-\gamma +\mathop{\sum}\limits_{n\ge 0}\left(\frac{1}{n+1}-\frac{1}{n+z}\right),$$
(39)

we have the following relation,

$$ \int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon \mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ -\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon \mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ =\beta \,{{\mbox{Re}}}\, \left[{\psi }^{(0)}\left(\frac{1}{2}\right)-{\psi }^{(0)}\left(\frac{1}{2}+i\frac{{\lambda }_{o}| {{{{{\bf{g}}}}}}({{{{{\bf{k}}}}}})| }{2\pi {k}_{B}T}\right)\right]\\ \equiv \beta {{{{{{\mathcal{C}}}}}}}_{0}(T),$$
(40)

where \({\chi }_{0}(T)={N}_{0}\ln \left(\frac{2{e}^{\gamma }{\omega }_{D}}{\pi {k}_{B}T}\right)\), γ = 0.57721 ⋯   is the Euler-Mascheroni constant and ωD is the Debye frequency.

Therefore,

$$ -\frac{{N}_{0}}{\beta }\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon {\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }\mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ =-\frac{{N}_{0}}{\beta }\int\nolimits_{-{\omega }_{D}}^{+{\omega }_{D}}d\epsilon {\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }\mathop{\sum}\limits_{{\omega }_{n}}{G}_{e}^{+}({{{{{\bf{k}}}}}},i{\omega }_{n}){G}_{h}^{-}({{{{{\bf{k}}}}}},i{\omega }_{n})\\ ={\chi }_{0}(T)+{N}_{0}{\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }{{{{{{\mathcal{C}}}}}}}_{0}(T).$$
(41)

Now we can proceed to calculate χ(T) given in Eq. (33),

$$\chi (T)= {\chi }_{0}(T){\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }{\left\vert {{{{{{\bf{d}}}}}}}_{o}\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}\right\vert }^{2}$$
(42)
$$\quad +{\chi }_{0}(T){\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }\left(| {{{{{{\bf{d}}}}}}}_{o}{| }^{2}-{\left\vert {{{{{{\bf{d}}}}}}}_{o}\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}\right\vert }^{2}\right)$$
(43)
$$\quad +{N}_{0}{\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }{{{{{{\mathcal{C}}}}}}}_{0}(T)\left(| {{{{{{\bf{d}}}}}}}_{o}{| }^{2}-{\left\vert {{{{{{\bf{d}}}}}}}_{o}\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}\right\vert }^{2}\right)$$
(44)
$$={\chi }_{0}(T)+{N}_{0}{\int}_{\!\!\!\!S}\frac{d{{\Omega }}}{2\pi }\,{{{{{{\mathcal{C}}}}}}}_{0}(T)\left({\left\vert {{{{{{\bf{d}}}}}}}_{o}\right\vert }^{2}-{\left\vert {{{{{{\bf{d}}}}}}}_{o}\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}\right\vert }^{2}\right).$$
(45)

In the calculation, we use normalized gap functions with \({\int}_{\!\!S}\frac{d{{\Omega }}}{2\pi }\,{{{{{{\bf{d}}}}}}}_{o}^{* }\cdot {{{{{{\bf{d}}}}}}}_{o}=1\). It leads to,

$$\ln \left(\frac{{T}_{c}}{{T}_{c0}}\right)={\int}_{\!\!S}\frac{d{{\Omega }}}{2\pi }\,{{{{{{\mathcal{C}}}}}}}_{0}({T}_{c})\left({\left\vert {{{{{{\bf{d}}}}}}}_{o}\right\vert }^{2}-{\left\vert {{{{{{\bf{d}}}}}}}_{o}\cdot {\hat{{{{{{\bf{g}}}}}}}}_{o}\right\vert }^{2}\right),$$
(46)

where Tc0 is Tc for λo = 0 case by solving v0χ0(Tc0) = 1. This is the Eq. (8) in the main text. In general, the right-hand side of Eq. (46) suppresses Tc. It clearly indicates that Tc would not be suppressed by orbital hybridization once dogo for all k. So we conclude that the orbital do-vector is possible to be stabilized in materials.