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

Local condensation of charge-4e4e superconductivity at a nematic domain wall

Matthias Hecker School of Physics and Astronomy, University of Minnesota, Minneapolis 55455 MN, USA    Rafael M. Fernandes School of Physics and Astronomy, University of Minnesota, Minneapolis 55455 MN, USA School of Physics and Astronomy, University of Minnesota, Minneapolis 55455 MN, USA School of Physics and Astronomy, University of Minnesota, Minneapolis 55455 MN, USA
(September 7, 2025)
Abstract

In the fluctuation regime that precedes the onset of pairing in multi-component superconductors, such as nematic and chiral superconductors, the normal state is generally unstable towards the formation of charge-4e4e order – an exotic quantum state in which electrons form coherent quartets rather than Cooper pairs. However, charge-4e4e order is often suppressed by other competing composite orders, such as nematics. Importantly, the formation of nematic domains is unavoidable due to the long-range strains generated, leading to one-dimensional regions where the competing nematic order is suppressed. Here, we employ a real-space variational approach to demonstrate that, in such nematic domain walls, charge-4e4e order is locally condensed via a vestigial-order mechanism. We explore the experimental manifestations of this effect and discuss materials in which it can be potentially observed.

Introduction.

Shortly after the development of the BCS model for superconductivity, it was recognized that a gas of bosons could also form a coherent state of pairs of bosons before the Bose-Einstein condensation of individual particles, provided that strong enough attractive interactions are present [1, 2]. In condensed matter systems, pair condensation of bosonic quasiparticles has been studied in various settings, from biexcitons in semiconductors [3, 4] to two-magnon bound states in frustrated magnets [5, 6]. A fascinating possibility is the emergence, in superconducting materials, of a coherent state of pairs of Cooper pairs, dubbed quartets [7, 8, 9, 10, 11, 12]. Theoretically, a charge-4e4e superconducting state is expected to display gapless excitations [13, 14] and half flux-quantum vortices [10]. Experimentally, charge-4e4e bound states have been recently invoked to explain puzzling magneto-transport data in kagome superconductors [15, 16].

In the case of bosonic particles, the state with paired bosons is thermodynamically stable only when there is more than one bosonic “flavor” available for condensation (e.g. spin-1 bosons) [17]. This suggests that “multi-flavor” superconductors are a promising setting to search for charge-4e4e superconductivity. Indeed, theoretical proposals for quartet formation have included spin-3/2 systems [18], spinor condensates [19], multi-band superconductors [20], pair-density waves [10, 21, 22, 23, 24, 25, 26, 27], and multi-component superconductors [28, 29, 30, 31, 32, 33]. In the latter case, the superconductor is described by multiple gap functions related by lattice symmetries, 𝚫=(Δ1,Δ2,)\boldsymbol{\Delta}=\left(\Delta_{1},\,\Delta_{2},\cdots\right); in group-theory jargon, 𝚫\boldsymbol{\Delta} transforms as a multi-dimensional irreducible representation (IR) of the point group. A broad range of pairing states belong to this category, including several versions of pp-wave and dd-wave states in tetragonal, hexagonal, and cubic lattices [34, 35]. More importantly, there is experimental evidence for the realization of multi-component pairing in various systems of interest, from heavy-fermions [36, 37, 38] to moiré superlattices [39] to doped topological insulators [40, 41, 42].

Charge-4e4e order emerges in multi-component superconductors via the condensation of a complex-valued composite order parameter 𝚫𝚫0\left\langle\boldsymbol{\Delta}\cdot\boldsymbol{\Delta}\right\rangle\neq 0 (distinct from the real-valued composite 𝚫𝚫\left\langle\boldsymbol{\Delta}^{\dagger}\cdot\boldsymbol{\Delta}\right\rangle), while the superconducting order parameter itself remains zero, 𝚫=0\left\langle\boldsymbol{\Delta}\right\rangle=0, such that the transition temperature of the composite order, T4eT_{4e}, is larger than the superconducting one, TcT_{c} [28, 43]. This spontaneous symmetry-breaking, which is driven by fluctuations and thus not captured by mean-field approaches, lowers the U(1)U(1) gauge symmetry to Z2Z_{2}; the latter is further broken if 𝚫\left\langle\boldsymbol{\Delta}\right\rangle becomes non-zero. It is said then that the charge-4e4e and charge-2e2e superconducting states are intertwined, and that the former is a vestigial order of the latter [44, 45].

The main obstacle for the stabilization of charge-4e4e vestigial order is the competition with other vestigial phases, most notably nematic and ferromagnetic. Indeed, besides U(1)U(1) symmetry, the ground state of a multi-component superconductor also breaks either time-reversal (chiral superconductor) or rotational symmetry (nematic superconductor) [46]. These additional symmetries can be broken before the onset of superconductivity via the condensation of real-valued composite order parameters. Large-NN and variational calculations found that the corresponding vestigial nematic and ferromagnetic orders generally preempt the onset of charge-4e4e order [47, 48] – except for the special case of a hexagonal nematic superconductor [28].

\ffigbox

[][]Refer to caption

Figure 1: (a) Real-space sketch of our main result, the emergence of charge-4e4e order at a nematic domain wall. (b) Phase diagram of the vestigial orders supported by the nematic superconducting ground state of a tetragonal system. Here, uu, vv and ww are the Landau parameters of Eq. (4). The different shaded regions indicate which vestigial channel is attractive: none (green), nematic (blue), and leading nematic with subleading charge-4e4e (red).

In this work, we investigate whether the local suppression of the leading nematic or ferromagnetic vestigial order enables the local condensation of charge-4e4e order. As a motivation, we first discuss a 2D triplet superconductor without spin-orbit coupling (SOC), whose multi-component gap function has a continuous SU(2) symmetry. In this case, the continuous symmetry globally forbids quasi-long-range order of any composite order parameter, except for charge-4e4e. We then study the more realistic situation of discrete multi-component superconductors in tetragonal and hexagonal systems. While long-range order in the competing vestigial channel is unavoidable, it is locally suppressed due to the formation of domains. Focusing on a nematic superconductor on the tetragonal lattice, we employ a real-space variational approach that treats all composite orders on an equal footing. We find a wide range of parameters for which charge-4e4e order is condensed at the nematic domain walls while the superconducting order parameter remains uncondensed, as illustrated schematically in Fig. 1(a). We finish by discussing the experimental manifestations of this local condensation and candidate materials.

Global suppression of the competing vestigial phases.

To set the stage, we discuss a special case in which, for symmetry reasons alone, the only order that can be stabilized is charge-4e4e. Consider a 2D system in which the electrons experience negligible SOC (e.g. graphene) and, in addition to spin, have another pseudospin degree of freedom (e.g. sublattice or valley). Enforcing the pairing state to be momentum-independent and pseudospin-singlet, the gap function must be spin-triplet and represented by an order parameter 𝚫=(Δ1,Δ2,Δ3)T\boldsymbol{\Delta}=(\Delta_{1},\,\Delta_{2},\,\Delta_{3})^{T} that transforms as a vector in SU(2) spin space. This is nothing but the 𝒅\boldsymbol{d}-vector of a triplet superconductor [34, 35], albeit even in momentum. While there is no indication of valley-singlet spin-triplet superconductivity in graphene, this type of state has been studied in twisted moiré systems [49, 50, 51, 52] and Bernal bilayer graphene [32, 53]. The superconducting action is given by

𝒮\displaystyle\mathcal{S} =𝗊χ𝗊1|𝚫𝗊|2+𝗑(u|𝚫𝗑|4+v|𝚫𝗑𝚫𝗑|2),\displaystyle=\int_{\mathsf{q}}\chi_{\mathsf{q}}^{-1}\left|\boldsymbol{\Delta}_{\mathsf{q}}\right|^{2}+\int_{\mathsf{x}}\Big{(}u\left|\boldsymbol{\Delta}_{\mathsf{x}}\right|^{4}+v\left|\boldsymbol{\Delta}_{\mathsf{x}}\cdot\boldsymbol{\Delta}_{\mathsf{x}}\right|^{2}\Big{)}, (1)

with (bare) superconducting susceptibility χ𝗊1\chi_{\mathsf{q}}^{-1} and 𝗊=(𝐪,ωn)\mathsf{q}=\left(\mathbf{q},\omega_{n}\right) denoting momentum and Matsubara frequency. The interaction part has Landau coefficients uu and vv, where 𝗑=(𝐱,τ)\mathsf{x}=\left(\mathbf{x},\tau\right) denotes position and imaginary time. The mean-field phase diagram of this model is well-established [54, 55], displaying different types of unitary and non-unitary pairing depending on the sign of vv.

To describe the vestigial orders, however, it is necessary to go beyond mean-field [45]; for our purposes, a group-theoretical analysis is sufficient. There are four symmetry-breaking bilinear combinations of 𝚫\boldsymbol{\Delta}: two real-valued composites 𝚿(l=1)=𝗂𝚫×𝚫¯\boldsymbol{\Psi}^{(l=1)}=\mathsf{i}\boldsymbol{\Delta}\times\bar{\boldsymbol{\Delta}} and Ψμν(l=2)=12(ΔμΔ¯ν+ΔνΔ¯μ)13δμν|𝚫|2\Psi_{\mu\nu}^{(l=2)}=\frac{1}{2}\left(\Delta_{\mu}\bar{\Delta}_{\nu}+\Delta_{\nu}\bar{\Delta}_{\mu}\right)-\frac{1}{3}\,\delta_{\mu\nu}\left|\boldsymbol{\Delta}\right|^{2} and two complex-valued ones, ψ(l=0)=𝚫𝚫\psi^{(l=0)}=\boldsymbol{\Delta}\cdot\boldsymbol{\Delta} and ψμν(l=2)=ΔμΔν13δμν(𝚫𝚫)\psi_{\mu\nu}^{(l=2)}=\Delta_{\mu}\Delta_{\nu}-\frac{1}{3}\,\delta_{\mu\nu}\left(\boldsymbol{\Delta}\cdot\boldsymbol{\Delta}\right), where z¯z\bar{z}\equiv z^{*}. The superscript indicates the transformation properties in SU(2) spin-space, corresponding to a scalar (l=0l=0), a vector (l=1l=1), or a tensor (l=2l=2). Each composite has a clear physical interpretation: 𝚿(l=1)\boldsymbol{\Psi}^{(l=1)} corresponds to time-reversal symmetry-breaking and, thus, vestigial ferromagnetic order. Ψμν(l=2)\Psi_{\mu\nu}^{(l=2)} is associated with rotational symmetry-breaking in spin-space, and therefore denotes (spin-)nematic vestigial order [56, 57, 31]. Finally, ψ(l=0)\psi^{(l=0)} and ψμν(l=2)\psi_{\mu\nu}^{(l=2)} correspond to ss-wave and dd-wave charge-4e4e vestigial orders, respectively.

Because 𝚿(l=1)\boldsymbol{\Psi}^{(l=1)}, Ψμν(l=2)\Psi_{\mu\nu}^{(l=2)}, ψμν(l=2)\psi_{\mu\nu}^{(l=2)}, and 𝚫𝚫(l=1)\boldsymbol{\Delta}\equiv\boldsymbol{\Delta}^{(l=1)} itself transform non-trivially in SU(2) spin-space (i.e. they are at least Heisenberg-type order parameters), none of them can sustain (quasi-)long-range order at non-zero temperatures in 2D. On the other hand, because ψ(l=0)\psi^{(l=0)} is a complex scalar (i.e. XY-type), it can establish quasi-long-range order through a BKT transition. Therefore, the only state allowed to develop quasi-long-range order in this model is the vestigial charge-4e4e phase. We note that similar 2D and 1D models for triplet superconductors [31, 32, 33] and spinor condensates [19] have been previously studied and shown to support charge-4e4e order.

Local suppression of the competing vestigial phases.

Despite illuminating, the simple model above is not representative of realistic multi-component superconductors, where either SOC is not negligible or singlet states are realized. Yet, it highlights an efficient strategy to realize charge-4e4e order: suppression of the other, leading, vestigial phases. While this generally cannot be accomplished globally via Mermin-Wagner’s theorem, vestigial nematic or ferromagnetic states tend to form domains to minimize the elastic or magnetic dipolar energies. To explore this idea, we consider a generic two-component superconducting order parameter 𝚫=(Δ1,Δ2)\boldsymbol{\Delta}=(\Delta_{1},\Delta_{2}), which could describe (px,py)(p_{x},p_{y})-wave or (dxz,dyz)(d_{xz},d_{yz})-wave states in tetragonal and hexagonal lattices [34, 35], or (dx2y2,dxy)(d_{x^{2}-y^{2}},d_{xy})-wave pairing in hexagonal systems and 4545^{\circ}-twisted bilayer tetragonal dd-wave superconductors [58]. In contrast to the previous example, 𝚫\boldsymbol{\Delta} now transforms as a two-dimensional IR of a discrete point group. While we will focus on tetragonal (D4h\mathrm{D_{4h}}) superconductors, where 𝚫\boldsymbol{\Delta} transforms as the IR EgE_{g} or EuE_{u}, the conclusions apply to all other cases.

We start by classifying all possible composite order parameters. As shown in [43], seven different bilinear combinations can be formed. Apart from the symmetry-preserving bilinear ΨA1g=𝚫τ0𝚫\Psi^{A_{1g}}=\boldsymbol{\Delta}^{\dagger}\tau^{0}\boldsymbol{\Delta}, with Pauli matrices τi\tau^{i} acting on the 𝚫\boldsymbol{\Delta} subspace, there are three additional real-valued bilinears:

ΨA2g\displaystyle\Psi^{A_{2g}} =𝚫τy𝚫,\displaystyle=\boldsymbol{\Delta}^{\dagger}\tau^{y}\boldsymbol{\Delta}, ΨB1g\displaystyle\Psi^{B_{1g}} =𝚫τz𝚫,\displaystyle=\boldsymbol{\Delta}^{\dagger}\tau^{z}\boldsymbol{\Delta}, ΨB2g\displaystyle\Psi^{B_{2g}} =𝚫τx𝚫.\displaystyle=\boldsymbol{\Delta}^{\dagger}\tau^{x}\boldsymbol{\Delta}. (2)

Here, ΨA2g\Psi^{A_{2g}} breaks time-reversal-symmetry and causes ferromagnetism, while ΨB1g\Psi^{B_{1g}} and ΨB2g\Psi^{B_{2g}} break tetragonal symmetry and cause nematicity. The three complex-valued bilinears are given by

ψA1g\displaystyle\psi^{A_{1g}} =𝚫Tτ0𝚫,\displaystyle=\boldsymbol{\Delta}^{T}\tau^{0}\boldsymbol{\Delta}, ψB1g\displaystyle\psi^{B_{1g}} =𝚫Tτz𝚫,\displaystyle=\boldsymbol{\Delta}^{T}\tau^{z}\boldsymbol{\Delta}, ψB2g\displaystyle\psi^{B_{2g}} =𝚫Tτx𝚫,\displaystyle=\boldsymbol{\Delta}^{T}\tau^{x}\boldsymbol{\Delta}, (3)

and describe, respectively, ss-wave, dx2y2d_{x^{2}-y^{2}}-wave, and dxyd_{xy}-wave charge-4e4e superconductivity. In our notation, Ψn\Psi^{n} (ψn\psi^{n}) denotes real-valued (complex-valued) bilinears, whereas the superscript nn indicates the IR according to which the composite transforms. The superconducting action is 𝒮=𝗑r0|𝚫𝗑|2+𝒮grad+𝒮int,\mathcal{S}=\int_{\mathsf{x}}r_{0}\left|\boldsymbol{\Delta}_{\mathsf{x}}\right|^{2}+\;\mathcal{S}^{\mathrm{grad}}\;+\;\mathcal{S}^{\mathrm{int}}\,, where r0=a0(TT0)r_{0}=a_{0}(T-T_{0}) denotes the bare SC transition (a0,T0>0a_{0},T_{0}>0) and 𝒮grad\mathcal{S}^{\mathrm{grad}} contains the symmetry-allowed gradient terms [35]. The interaction part is given by [43]

𝒮int\displaystyle\mathcal{S}^{\mathrm{int}} =𝗑[u(Ψ𝗑A1g)2+v(Ψ𝗑A2g)2+w(Ψ𝗑B1g)2],\displaystyle=\int_{\mathsf{x}}\Big{[}u\,(\Psi_{\mathsf{x}}^{A_{1g}})^{2}+v\,(\Psi_{\mathsf{x}}^{A_{2g}})^{2}+w\,(\Psi_{\mathsf{x}}^{B_{1g}})^{2}\Big{]}, (4)

and contains three independent interaction parameters u>0u>0 and v,w>uv,w>-u. The mean-field phase diagram in the (vu,wu)\left(\frac{v}{u},\,\frac{w}{u}\right) parameter space is well-established, displaying chiral and two types of nematic superconductivity [35, 10, 46].

The vestigial orders associated with each mean-field ground state were analyzed in Ref. [43] via a variational approach. The leading vestigial instability is always that of a real-valued composite (nematic or ferromagnetic) whereas the vestigial charge-4e4e orders are always subleading. While our conclusions hold across the entire phase diagram, hereafter we focus on the v>0>wv>0>w region [Fig. 1(b)], where the superconducting ground state is nematic and the competing vestigial phases are dx2y2d_{x^{2}-y^{2}} nematic (ΨB1g\Psi^{B_{1g}}) and ss-wave charge-4e4e (ψA1g\psi^{A_{1g}}). In bulk, the only vestigial order realized is the nematic one [43]. However, because the effective interaction in the charge-4e4e channel is attractive, the system could gain energy by condensing this mode in regions where nematic order is suppressed. Due to its Ising-like character, ΨB1g\Psi^{B_{1g}} can sustain long-range order at non-zero temperatures. However, because of the linear coupling between ΨB1g\Psi^{B_{1g}} and strain, nematic domains must form to minimize the elastic energy and accommodate long-range lattice deformations. This opens up the possibility of ψA1g\psi^{A_{1g}} condensation at nematic domain walls.

Charge-4e condensation at the domain wall.

To proceed, we employ a real-space Gaussian variational approach, which treats all vestigial channels equally, on a 1D grid of length LL and i=1,,Ni=1,\dots,N sites. The variational ansatz consists of a trial action 𝒮0\mathcal{S}_{0} [59, 60, 61, 43], which in our case is 𝒮0=12LTi𝚫^iGi1𝚫^i+𝒮grad\mathcal{S}_{0}=\frac{1}{2}\frac{L}{T}\sum_{i}\hat{\boldsymbol{\Delta}}_{i}^{\dagger}\,G_{i}^{-1}\,\hat{\boldsymbol{\Delta}}_{i}+\mathcal{S}^{\mathrm{grad}}. Represented in the Nambu basis 𝚫^i=(𝚫i,𝚫¯i)\hat{\boldsymbol{\Delta}}_{i}=(\boldsymbol{\Delta}_{i},\bar{\boldsymbol{\Delta}}_{i}), the local inverse Green’s function

Gi1\displaystyle G_{i}^{-1} =(Ri+ΦiB1gΦiB2g𝗂ΦiA2gϕiA1g+ϕiB1gϕiB2gRiΦiB1gϕiB2gϕiA1gϕiB1gRi+ΦiB1gΦiB2g+𝗂ΦiA2gH.c.RiΦiB1g),\displaystyle=\left(\begin{smallmatrix}R_{i}+\Phi_{i}^{B_{1g}}&\Phi_{i}^{B_{2g}}-\mathsf{i}\Phi_{i}^{A_{2g}}&\phi_{i}^{A_{1g}}+\phi_{i}^{B_{1g}}&\phi_{i}^{B_{2g}}\\ &R_{i}-\Phi_{i}^{B_{1g}}&\phi_{i}^{B_{2g}}&\phi_{i}^{A_{1g}}-\phi_{i}^{B_{1g}}\\ &&R_{i}+\Phi_{i}^{B_{1g}}&\Phi_{i}^{B_{2g}}+\mathsf{i}\Phi_{i}^{A_{2g}}\\ \mathrm{H.c.}&&&R_{i}-\Phi_{i}^{B_{1g}}\end{smallmatrix}\right), (5)

contains the real-valued (Φin\Phi_{i}^{n}) and complex-valued (ϕin,ϕ¯in)(\phi_{i}^{n},\bar{\phi}_{i}^{n}) variational parameters, and the mass renormalization parameter Ri=r0+ΦiA1gR_{i}=r_{0}+\Phi_{i}^{A_{1g}}. Because 𝒮0\mathcal{S}_{0} is Gaussian, it is straightforward to compute the variational free energy

Fv\displaystyle F_{v} =TlogZ0+T𝒮𝒮00,\displaystyle=-T\log Z_{0}+T\langle\mathcal{S}-\mathcal{S}_{0}\rangle_{0}, (6)

where Z0=D(𝚫,𝚫¯)e𝒮0Z_{0}=\int D(\boldsymbol{\Delta},\bar{\boldsymbol{\Delta}})\,e^{-\mathcal{S}_{0}} is the partition function of the trial action. The detailed evaluation of Eq. (6) is shown in the Supplementary Material (SM). Importantly, the original Landau coefficients uu, vv, and ww appear in FvF_{v} as different combinations in each symmetry channel, corresponding to effective interactions UA1g=3u+v+wU_{A_{1g}}=3u+v+w, UA2g=u+3vw,U_{A_{2g}}=u+3v-w, UB1g=uv+3wU_{B_{1g}}=u-v+3w, UB2g=uvwU_{B_{2g}}=u-v-w, uA1g=uv+wu_{A_{1g}}=u-v+w, uB1g=u+v+wu_{B_{1g}}=u+v+w, and uB2g=u+vwu_{B_{2g}}=u+v-w [43]. An attractive interaction (Un,un<0U_{n},u_{n}<0) indicates a potential instability, signaled by the condensation of the corresponding variational order parameter (Φin,ϕin\Phi_{i}^{n},\,\phi_{i}^{n}). Importantly, a vestigial phase only emerges if superconductivity is not immediately triggered by Φin,ϕin0\Phi_{i}^{n},\,\phi_{i}^{n}\neq 0. In the variational approach, this condition can be verified by confirming that the variational superconducting susceptibility χ\chi, which is given by a combination of RiR_{i}, Φin\Phi_{i}^{n}, and ϕin\phi_{i}^{n} (see SM for details), remains finite.

Refer to caption
Figure 2: Variational nematic order parameter (ΦB1g\Phi^{B_{1g}}), inverse superconducting susceptibility (χ1\chi^{-1}), and variational charge-4e4e order parameter (ϕA1g\phi^{A_{1g}}) obtained from the numerical minimization of the variational free energy (6) across two nematic domains (all in units of the gradient-term stiffness t0t_{0}). Each panel corresponds to a different temperature, parametrized by r0TT0r_{0}\propto T-T_{0}. The parameters used here are w=0.5uw=-0.5u and v=0.1uv=0.1u [red cross in Fig. 1(b)]. Here and in Fig. 3, we set t0/uT0/L=20/7t_{0}/\sqrt{uT_{0}/L}=20/7.
Refer to caption
Figure 3: Same variational parameters as in Fig. 2, but evaluated for the parameters w=0.05uw=-0.05u and v=2.1uv=2.1u [purple cross in Fig. 1(b)], corresponding to a subleading attractive charge-4e4e channel (uA1g<0u_{A_{1g}}<0). Local condensation of ϕA1g\phi^{A_{1g}} is observed at the nematic domain wall.

In bulk, for most of the (vu,wu)\left(\frac{v}{u},\,\frac{w}{u}\right) phase diagram, there are two competing attractive vestigial-order channels, corresponding to a real-valued composite (nematic/ferromagnetic) and a complex-valued one (ss-wave/dd-wave charge-4e4e). As demonstrated in Ref. [43], the nematic/ferromagnetic instability is always the leading one in bulk [Fig. 1(b)]. Our goal is to determine the fate of these phases along a nematic domain wall. We therefore numerically minimize the free energy (6) for the NN-site 1D grid with domain-wall boundary conditions Φ1B1g=ΦNB1g=Φ0B1g\Phi_{1}^{B_{1g}}=-\Phi_{N}^{B_{1g}}=\Phi_{0}^{B_{1g}}, where Φ0B1g\Phi_{0}^{B_{1g}} is the (self-consistently obtained) bulk value of the nematic order parameter (see SM for additional details). Consider first the phase diagram region where the only attractive vestigial-order channel is the nematic, i.e. UB1g<0U_{B_{1g}}<0 but uA1g>0u_{A_{1g}}>0 [red cross in Fig. 1(b)]. The results are shown in Fig. 2. As the control parameter r0TT0r_{0}\propto T-T_{0} is decreased, a nematic domain emerges below a temperature that coincides with the bulk nematic critical temperature. As expected, the domain wall becomes sharper as the temperature is lowered, since the wall width scales as t0/|Φ0B1g|t_{0}/|\Phi_{0}^{B_{1g}}| where t0t_{0} is the gradient-term stiffness. At the domain-wall center, the superconducting susceptibility χ\chi is suppressed, consistent with the fact that vestigial order enhances the superconducting transition. Interestingly, χ1\chi^{-1} has a non-monotonic spatial dependence, displaying a dip at the domain-wall boundaries, which can lead to local condensation of superconductivity [yellow region of Fig. 2(c)]. No sign of charge-4e4e order is observed, consistent with a repulsive effective interaction in this channel.

Consider now the phase-diagram region where the charge-4e4e vestigial channel is attractive, but subleading to the nematic, UB1g<uA1g<0U_{B_{1g}}<u_{A_{1g}}<0 [purple cross in Fig. 1(b)] . As shown in Fig. 3, upon decreasing r0r_{0}, nematic order emerges first, in two domains. However, in contrast to Fig. 2, the charge-4e4e order parameter ϕA1g\phi^{A_{1g}} condenses inside the domain wall while the superconducting susceptibility χ\chi remains finite. This is the main result of the paper. Because the domain wall is one-dimensional, this should be understood as a local condensation, since phase slips will destroy long-range order along the wall. Note, even though the dd-wave charge-4e4e channel is repulsive in this phase-diagram region, the order parameter ϕB1g\phi^{B_{1g}} condenses due to the trilinear coupling between ΦB1g\Phi^{B_{1g}}, ϕA1g\phi^{A_{1g}}, and ϕ¯B1g\bar{\phi}^{B_{1g}} (see the inset) [43]. We verified that this behavior is not particular to this set of parameters, but occurs in any region of the (vu,wu)\left(\frac{v}{u},\,\frac{w}{u}\right) phase diagram where ss-wave/dd-wave vestigial charge-4e4e order is sufficiently attractive but subleading to nematic/ferromagnetic vestigial order.

Discussion.

In this work, we employed a variational approach to demonstrate that charge-4e4e order locally condenses at nematic domain walls that emerge in the normal-state of nematic superconductors upon approaching the superconducting transition. Before the onset of superconductivity, the system is unstable towards the formation of vestigial nematic and charge-4e4e orders arising from the fluctuation-induced condensation of composite superconducting order parameters. Because nematic order generally wins over charge-4e4e order, the suppression of nematic order along the domain wall enables the system to further minimize the free energy by condensing charge-4e4e order. Since an analogous mechanism also applies for chiral superconductors, our results point to a wide class of systems – multi-component superconductors – where local charge-4e4e order can potentially emerge. Among the materials for which there is strong experimental evidence for nematic superconductivity, doped Bi2Se3\mathrm{Bi_{2}Se_{3}} [62, 40, 42, 41] and twisted bilayer graphene (TBG) [39, 63, 64] are the most promising candidates to display this effect. Indeed, a vestigial nematic phase exists in doped Bi2Se3\mathrm{Bi_{2}Se_{3}} [65, 66], whereas in TBG, normal-state nematic order appears close to the superconducting dome [39]. There are also several chiral-superconductor candidates [67], including the widely studied heavy-fermion UPt3 [36, 37, 38].

An important question is how to experimentally detect this effect. Since charge-4e4e order emerges at nematic domain walls, local probes such as scanning tunneling microscopy (STM) and scanning near-field optical microscopy (SNOM) are ideal. Because the charge-4e4e state is expected to be gapless [13, 14], its density-of-states (DOS) profile, which can be accessed in a standard STM measurement, will likely differ from the normal-state DOS only in subtle ways. On the other hand, Josephson-STM, in which a superconducting tip is used [68, 69], could provide more direct evidence for quartets. The SNOM technique has the unique capability of probing the local optical response with a few nanometers resolution, from which the properties of the local optical conductivity σ(ω,𝒓)\sigma(\omega,\boldsymbol{r}) can be inferred [70, 71]. Because the charge-4e4e state has a non-zero superfluid density [14], Imσ(ω,𝒓)1/ω\mathrm{Im}\,\sigma(\omega,\boldsymbol{r})\sim 1/\omega is expected to emerge at low frequencies near nematic domain walls. While this behavior could also be due to a charge-2e2e superconducting filament, only in the charge-4e4e case this behavior would be accompanied by a gapless DOS, which can be probed via STM. These results also reveal the tantalizing possibility of using uniaxial strain to control the charge-4e4e phase, since beyond a critical strain value, the sample becomes mono-nematic-domain and local charge-4e4e order disappears.

Acknowledgements.
We thank J. Schmalian, L. Fu, and R. Willa for fruitful discussions, and particularly Y. Wang for in-depth discussions about the properties of charge-4e4e superconductors. This work was supported by the U. S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division, under Award No. DE-SC0020045.

References

  • Valatin and Butler [1958] J. Valatin and D. Butler, On the collective properties of a boson system, Nuovo Cimento 10, 37 (1958).
  • Evans and Imry [1969] W. Evans and Y. Imry, On the pairing theory of the bose superfluid, Nuovo Cimento B 63, 155 (1969).
  • Lampert [1958] M. A. Lampert, Mobile and immobile effective-mass-particle complexes in nonmetallic solids, Phys. Rev. Lett. 1, 450 (1958).
  • Moskalenko [1958] S. A. Moskalenko, Opt. Spektrosk 5, 147 (1958).
  • Wang et al. [2017] Z. Wang, A. E. Feiguin, W. Zhu, O. A. Starykh, A. V. Chubukov, and C. D. Batista, Chiral liquid phase of simple quantum magnets, Phys. Rev. B 96, 184409 (2017).
  • Strockoz et al. [2022] J. Strockoz, D. S. Antonenko, D. LaBelle, and J. W. Venderbos, Excitonic instability towards a Potts-nematic quantum paramagnet, arXiv:2211.11739  (2022).
  • Korshunov [1985] S. Korshunov, Two-dimensional superfluid Fermi liquid with pp-pairing, Zh. Eksp. Teor. Fiz 89, 539 (1985).
  • Volovik [1992] G. E. Volovik, Exotic properties of superfluid 3He, Vol. 1 (World Scientific, 1992).
  • Aligia et al. [2005] A. A. Aligia, A. P. Kampf, and J. Mannhart, Quartet formation at (100)/(110)(100)/(110) interfaces of dd-wave superconductors, Phys. Rev. Lett. 94, 247004 (2005).
  • Berg et al. [2009] E. Berg, E. Fradkin, and S. A. Kivelson, Charge-4e4e superconductivity from pair-density-wave order in certain high-temperature superconductors, Nature Physics 5, 830 (2009).
  • Moon [2012] E.-G. Moon, Skyrmions with quadratic band touching fermions: A way to achieve charge 4e4e superconductivity, Phys. Rev. B 85, 245123 (2012).
  • Khalaf et al. [2022] E. Khalaf, P. Ledwith, and A. Vishwanath, Symmetry constraints on superconductivity in twisted bilayer graphene: Fractional vortices, 4e4e condensates, or nonunitary pairing, Phys. Rev. B 105, 224508 (2022).
  • Jiang et al. [2017] Y.-F. Jiang, Z.-X. Li, S. A. Kivelson, and H. Yao, Charge-4e4e superconductors: A Majorana quantum Monte Carlo study, Phys. Rev. B 95, 241103 (2017).
  • Gnezdilov and Wang [2022] N. V. Gnezdilov and Y. Wang, Solvable model for a charge-4e4e superconductor, Phys. Rev. B 106, 094508 (2022).
  • Ge et al. [2022] J. Ge, P. Wang, Y. Xing, Q. Yin, H. Lei, Z. Wang, and J. Wang, Discovery of charge-4e4e and charge-6e6e superconductivity in kagome superconductor CsV3Sb5CsV_{3}Sb_{5}, arXiv:2201.10352  (2022).
  • Varma and Wang [2023] C. M. Varma and Z. Wang, Extended superconducting fluctuation region and 6e and 4e flux-quantization in a kagome compound with a normal state of 3q-order, arXiv:2307.00448  (2023).
  • Noziéres and Saint James [1982] P. Noziéres and D. Saint James, Particle vs. pair condensation in attractive Bose liquids, Journal de Physique 43, 1133 (1982).
  • Wu [2005] C. Wu, Competing orders in one-dimensional spin-3/23/2 fermionic systems, Phys. Rev. Lett. 95, 266404 (2005).
  • Mukerjee et al. [2006] S. Mukerjee, C. Xu, and J. E. Moore, Topological defects and the superfluid transition of the s=1s=1 spinor condensate in two dimensions, Phys. Rev. Lett. 97, 120406 (2006).
  • Herland et al. [2010] E. V. Herland, E. Babaev, and A. Sudbø, Phase transitions in a three dimensional U(1)×U(1)U(1)\times U(1) lattice London superconductor: Metallic superfluid and charge-4e4e superconducting states, Phys. Rev. B 82, 134511 (2010).
  • Radzihovsky and Vishwanath [2009] L. Radzihovsky and A. Vishwanath, Quantum liquid crystals in an imbalanced Fermi gas: Fluctuations and fractional vortices in Larkin-Ovchinnikov states, Phys. Rev. Lett. 103, 010404 (2009).
  • Radzihovsky [2011] L. Radzihovsky, Fluctuations and phase transitions in Larkin-Ovchinnikov liquid-crystal states of a population-imbalanced resonant Fermi gas, Phys. Rev. A 84, 023611 (2011).
  • Agterberg et al. [2011] D. F. Agterberg, M. Geracie, and H. Tsunetsugu, Conventional and charge-six superfluids from melting hexagonal Fulde-Ferrell-Larkin-Ovchinnikov phases in two dimensions, Phys. Rev. B 84, 014513 (2011).
  • Agterberg et al. [2020] D. F. Agterberg, J. C. S. Davis, S. D. Edkins, E. Fradkin, D. J. V. Harlingen, S. A. Kivelson, P. A. Lee, L. Radzihovsky, J. M. Tranquada, and Y. Wang, The physics of pair-density waves: Cuprate superconductors and beyond, Annual Review of Condensed Matter Physics 11, 231 (2020).
  • Shaffer et al. [2023] D. Shaffer, F. J. Burnell, and R. M. Fernandes, Weak-coupling theory of pair density wave instabilities in transition metal dichalcogenides, Phys. Rev. B 107, 224516 (2023).
  • Yu [2023] Y. Yu, Nondegenerate surface pair density wave in the kagome superconductor csv3sb5{\mathrm{csv}}_{3}{\mathrm{sb}}_{5}: Application to vestigial orders, Phys. Rev. B 108, 054517 (2023).
  • Wu and Wang [2023] Y.-M. Wu and Y. Wang, dd-wave charge-4e4e superconductivity from fluctuating pair density waves, arXiv:2303.17631  (2023).
  • Fernandes and Fu [2021] R. M. Fernandes and L. Fu, Charge-4e4e superconductivity from multicomponent nematic pairing: Application to twisted bilayer graphene, Phys. Rev. Lett. 127, 047001 (2021).
  • Jian et al. [2021] S.-K. Jian, Y. Huang, and H. Yao, Charge-4e4e superconductivity from nematic superconductors in two and three dimensions., Phys. Rev. Lett. 127, 227001 (2021).
  • Zeng et al. [2021] M. Zeng, L.-H. Hu, H.-Y. Hu, Y.-Z. You, and C. Wu, Phase-fluctuation induced time-reversal symmetry breaking normal state, arXiv:2102.06158  (2021).
  • Rampp et al. [2022] M. A. Rampp, E. J. König, and J. Schmalian, Topologically enabled superconductivity, Phys. Rev. Lett. 129, 077001 (2022).
  • Curtis et al. [2023] J. B. Curtis, N. R. Poniatowski, Y. Xie, A. Yacoby, E. Demler, and P. Narang, Stabilizing fluctuating spin-triplet superconductivity in graphene via induced spin-orbit coupling, Phys. Rev. Lett. 130, 196001 (2023).
  • Poduval and Scheurer [2023] P. P. Poduval and M. S. Scheurer, Vestigial singlet pairing in a fluctuating magnetic triplet superconductor: Applications to graphene moiré systems, arXiv:2301.01344  (2023).
  • Annett [1990] J. F. Annett, Symmetry of the order parameter for high-temperature superconductivity, Advances in Physics 39, 83 (1990).
  • Sigrist and Ueda [1991] M. Sigrist and K. Ueda, Phenomenological theory of unconventional superconductivity, Rev. Mod. Phys. 63, 239 (1991).
  • Sauls [1994] J. Sauls, The order parameter for the superconducting phases of UPt3UPt_{3}Advances in Physics 43, 113 (1994).
  • Schemm et al. [2014] E. Schemm, W. Gannon, C. Wishne, W. Halperin, and A. Kapitulnik, Observation of broken time-reversal symmetry in the heavy-fermion superconductor UPt3UPt_{3}Science 345, 190 (2014).
  • Avers et al. [2020] K. E. Avers, W. J. Gannon, S. J. Kuhn, W. P. Halperin, J. Sauls, L. DeBeer-Schmitt, C. Dewhurst, J. Gavilano, G. Nagy, U. Gasser, et al., Broken time-reversal symmetry in the topological superconductor UPt3UPt_{3}Nature Physics 16, 531 (2020).
  • Cao et al. [2021] Y. Cao, D. Rodan-Legrain, J. M. Park, N. F. Yuan, K. Watanabe, T. Taniguchi, R. M. Fernandes, L. Fu, and P. Jarillo-Herrero, Nematicity and competing orders in superconducting magic-angle graphene, Science 372, 264 (2021).
  • Matano et al. [2016] K. Matano, M. Kriener, K. Segawa, Y. Ando, and G.-q. Zheng, Spin-rotation symmetry breaking in the superconducting state of CuxBi2Se3Cu_{x}Bi_{2}Se_{3}Nature Physics 12, 852 (2016).
  • Pan et al. [2016] Y. Pan, A. Nikitin, G. Araizi, Y. Huang, Y. Matsushita, T. Naka, and A. De Visser, Rotational symmetry breaking in the topological superconductor SrxBi2Se3Sr_{x}Bi_{2}Se_{3} probed by upper-critical field experiments, Scientific reports 6, 28632 (2016).
  • Asaba et al. [2017] T. Asaba, B. J. Lawson, C. Tinsman, L. Chen, P. Corbae, G. Li, Y. Qiu, Y. S. Hor, L. Fu, and L. Li, Rotational symmetry breaking in a trigonal superconductor NbNb-doped Bi2Se3Bi_{2}Se_{3}Phys. Rev. X 7, 011009 (2017).
  • Hecker et al. [2023] M. Hecker, R. Willa, J. Schmalian, and R. M. Fernandes, Cascade of vestigial orders in two-component superconductors: Nematic, ferromagnetic, ss-wave charge-4e4e, and dd-wave charge-4e4e states, Phys. Rev. B 107, 224503 (2023).
  • Fradkin et al. [2015] E. Fradkin, S. A. Kivelson, and J. M. Tranquada, Colloquium: Theory of intertwined orders in high temperature superconductors, Rev. Mod. Phys. 87, 457 (2015).
  • Fernandes et al. [2019] R. M. Fernandes, P. P. Orth, and J. Schmalian, Intertwined vestigial order in quantum materials: Nematicity and beyond, Annual Review of Condensed Matter Physics 10, 133 (2019).
  • Gali and Fernandes [2022] V. Gali and R. M. Fernandes, Role of electromagnetic gauge-field fluctuations in the selection between chiral and nematic superconductivity, Phys. Rev. B 106, 094509 (2022).
  • Hecker and Schmalian [2018] M. Hecker and J. Schmalian, Vestigial nematic order and superconductivity in the doped topological insulator CuxBi2Se3Cu_{x}Bi_{2}Se_{3}npj Quantum Materials 3, 1 (2018).
  • Willa [2020] R. Willa, Symmetry-mixed bound-state order, Phys. Rev. B 102, 180503 (2020).
  • Venderbos and Fernandes [2018] J. W. F. Venderbos and R. M. Fernandes, Correlations and electronic order in a two-orbital honeycomb lattice model for twisted bilayer graphene, Phys. Rev. B 98, 245103 (2018).
  • Scheurer and Samajdar [2020] M. S. Scheurer and R. Samajdar, Pairing in graphene-based moiré superlattices, Phys. Rev. Research 2, 033062 (2020).
  • Wang et al. [2021] Y. Wang, J. Kang, and R. M. Fernandes, Topological and nematic superconductivity mediated by ferro-su(4) fluctuations in twisted bilayer graphene, Phys. Rev. B 103, 024506 (2021).
  • Lake et al. [2022] E. Lake, A. S. Patri, and T. Senthil, Pairing symmetry of twisted bilayer graphene: A phenomenological synthesis, Phys. Rev. B 106, 104506 (2022).
  • Dong et al. [2023] Z. Dong, A. V. Chubukov, and L. Levitov, Transformer spin-triplet superconductivity at the onset of isospin order in bilayer graphene, Phys. Rev. B 107, 174512 (2023).
  • Volovik and Gor’kov [1985] G. Volovik and L. Gor’kov, Superconducting classes in heavy-fermion systems, Zh. Eksp. Teor. Fiz 88, 1412 (1985).
  • Blagoeva et al. [1990] E. J. Blagoeva, G. Busiello, L. De Cesare, Y. T. Millev, I. Rabuffo, and D. I. Uzunov, Fluctuation-induced first-order transitions in unconventional superconductors, Phys. Rev. B 42, 6124 (1990).
  • Blume and Hsieh [1969] M. Blume and Y. Hsieh, Biquadratic exchange and quadrupolar ordering, J. Appl. Phys. 40, 1249 (1969).
  • Andreev and Grishchuk [1984] A. Andreev and I. Grishchuk, Spin nematics, Sov. Phys. JETP 60, 267 (1984).
  • Haenel et al. [2022] R. Haenel, T. Tummuru, and M. Franz, Incoherent tunneling and topological superconductivity in twisted cuprate bilayers, Phys. Rev. B 106, 104505 (2022).
  • Moshe and Zinn-Justin [2003] M. Moshe and J. Zinn-Justin, Quantum field theory in the large N limit: a review, Physics Reports 385, 69 (2003).
  • Fischer and Berg [2016] M. H. Fischer and E. Berg, Fluctuation and strain effects in a chiral pp-wave superconductor, Phys. Rev. B 93, 054501 (2016).
  • Nie et al. [2017] L. Nie, A. V. Maharaj, E. Fradkin, and S. A. Kivelson, Vestigial nematicity from spin and/or charge order in the cuprates, Phys. Rev. B 96, 085142 (2017).
  • Fu [2014] L. Fu, Odd-parity topological superconductor with nematic order: Application to CuxBi2Se3Cu_{x}Bi_{2}Se_{3}Phys. Rev. B 90, 100509 (2014).
  • Kozii et al. [2019] V. Kozii, H. Isobe, J. W. F. Venderbos, and L. Fu, Nematic superconductivity stabilized by density wave fluctuations: Possible application to twisted bilayer graphene, Phys. Rev. B 99, 144507 (2019).
  • Chichinadze et al. [2020] D. V. Chichinadze, L. Classen, and A. V. Chubukov, Nematic superconductivity in twisted bilayer graphene, Phys. Rev. B 101, 224513 (2020).
  • Sun et al. [2019] Y. Sun, S. Kittaka, T. Sakakibara, K. Machida, J. Wang, J. Wen, X. Xing, Z. Shi, and T. Tamegai, Quasiparticle evidence for the nematic state above Tc{T}_{\mathrm{c}} in SrxBi2Se3Sr_{x}Bi_{2}Se_{3}Phys. Rev. Lett. 123, 027002 (2019).
  • Cho et al. [2020] C.-w. Cho, J. Shen, J. Lyu, O. Atanov, Q. Chen, S. H. Lee, Y. San Hor, D. J. Gawryluk, E. Pomjakushina, M. Bartkowiak, M. Hecker, J. Schmalian, and R. Lortz, Z3Z_{3}-vestigial nematic order due to superconducting fluctuations in the doped topological insulator NbxBi2Se3Nb_{x}Bi_{2}Se_{3} and CuxBi2Se3Cu_{x}Bi_{2}Se_{3}Nature communications 11, 1 (2020).
  • Ghosh et al. [2020] S. K. Ghosh, M. Smidman, T. Shang, J. F. Annett, A. D. Hillier, J. Quintanilla, and H. Yuan, Recent progress on superconductors with time-reversal symmetry breaking, Journal of Physics: Condensed Matter 33, 033001 (2020).
  • Naaman et al. [2001] O. Naaman, W. Teizer, and R. C. Dynes, Fluctuation dominated josephson tunneling with a scanning tunneling microscope, Phys. Rev. Lett. 87, 097004 (2001).
  • Bastiaans et al. [2019] K. M. Bastiaans, D. Cho, D. Chatzopoulos, M. Leeuwenhoek, C. Koks, and M. P. Allan, Imaging doubled shot noise in a josephson scanning tunneling microscope, Phys. Rev. B 100, 104506 (2019).
  • Yang et al. [2013] H. U. Yang, E. Hebestreit, E. E. Josberger, and M. B. Raschke, A cryogenic scattering-type scanning near-field optical microscope, Review of Scientific Instruments 84, 023701 (2013).
  • McLeod et al. [2017] A. S. McLeod, E. van Heumen, J. G. Ramirez, S. Wang, T. Saerbeck, S. Guenon, M. Goldflam, L. Anderegg, P. Kelly, A. Mueller, M. K. Liu, I. K. Schuller, and D. N. Basov, Nanotextured phase coexistence in the correlated insulator V2O3V_{2}O_{3}Nature Physics 13, 80 (2017).

Supplementary Material: Local condensation of charge-4e4e superconductivity at a nematic domain wall Matthias Hecker Rafael M. Fernandes September 7, 2025

Supplementary Material: Local condensation of charge-4e4e superconductivity at a nematic domain wall


SI Derivation of the variational free energy

Here, we derive the variational free energy associated with the ansatz 𝒮0\mathcal{S}_{0} of the main text, i.e. we evaluate equation (3) of the main text. For convenience, we first repeat the setup of the problem and the notations introduced in the main text. Expressing the two-component superconducting order parameter 𝚫=(Δ1,Δ2)\boldsymbol{\Delta}=(\Delta_{1},\Delta_{2}) in terms of the four-component Nambu basis 𝚫^=(𝚫,𝚫¯)\hat{\boldsymbol{\Delta}}=\left(\boldsymbol{\Delta},\bar{\boldsymbol{\Delta}}\right), we can rewrite the real-valued and complex-valued bilinear combinations as

Ψn\displaystyle\Psi^{n} =𝚫^Mn𝚫^,\displaystyle=\hat{\boldsymbol{\Delta}}^{\dagger}M^{n}\hat{\boldsymbol{\Delta}}, ψn\displaystyle\psi^{n} =𝚫^mn𝚫^.\displaystyle=\hat{\boldsymbol{\Delta}}^{\dagger}m^{n}\hat{\boldsymbol{\Delta}}. (S1)

Here, we defined the matrices

MA1g\displaystyle M^{A_{1g}} =τ0σ0/2,\displaystyle=\tau^{0}\sigma^{0}/2, mA1g\displaystyle m^{A_{1g}} =τ0σ,\displaystyle=\tau^{0}\sigma^{-}, MA2g\displaystyle M^{A_{2g}} =τyσz/2,\displaystyle=\tau^{y}\sigma^{z}/2,
MB1g\displaystyle M^{B_{1g}} =τzσ0/2,\displaystyle=\tau^{z}\sigma^{0}/2, mB1g\displaystyle m^{B_{1g}} =τzσ,\displaystyle=\tau^{z}\sigma^{-},
MB2g\displaystyle M^{B_{2g}} =τxσ0/2,\displaystyle=\tau^{x}\sigma^{0}/2, mB2g\displaystyle m^{B_{2g}} =τxσ,\displaystyle=\tau^{x}\sigma^{-}, (S2)

with σ±=(σx±𝗂σy)/2\sigma^{\pm}=(\sigma^{x}\pm\mathsf{i}\sigma^{y})/2 and τi\tau^{i}, σi\sigma^{i} acting respectively on the internal superconducting subspace and on the Nambu space. For our specific setting of a one-dimensional grid of length LL with lattice sites labeled i,j=1,,Ni,j=1,\dots,N, the superconducting action becomes

𝒮\displaystyle\mathcal{S} =LTi,j𝚫i[r0δij+12fij0]τ0𝚫j+𝒮int,\displaystyle=\frac{L}{T}\sum_{i,j}\boldsymbol{\Delta}_{i}^{\dagger}\left[r_{0}\delta_{ij}+\frac{1}{2}f_{ij}^{0}\right]\tau^{0}\boldsymbol{\Delta}_{j}\;+\;\mathcal{S}^{\mathrm{int}}\,,

where we introduced the gradient term

𝒮grad\displaystyle\mathcal{S}^{\mathrm{grad}} =L2Ti,j𝚫iτ0fij0𝚫j,\displaystyle=\frac{L}{2T}\sum_{i,j}\boldsymbol{\Delta}_{i}^{\dagger}\tau^{0}f_{ij}^{0}\boldsymbol{\Delta}_{j}, (S3)

described in terms of the hopping function fij0=t02(2δijδi,j+1δi,j1)f_{ij}^{0}=\frac{t_{0}}{2}\left(2\delta_{ij}-\delta_{i,j+1}-\delta_{i,j-1}\right) and the stiffness parameter t0>0t_{0}>0. Recall that r0=a0(TT0)r_{0}=a_{0}(T-T_{0}) denotes the bare superconducting transition temperature with a0,T0>0a_{0},T_{0}>0. The interaction part is given by [43] (see also Eq. (4) in main text)

𝒮int\displaystyle\mathcal{S}^{\mathrm{int}} =LTi[u(ΨiA1g)2+v(ΨiA2g)2+w(ΨiB1g)2],\displaystyle=\frac{L}{T}\sum_{i}\Big{[}u\,(\Psi_{i}^{A_{1g}})^{2}+v\,(\Psi_{i}^{A_{2g}})^{2}+w\,(\Psi_{i}^{B_{1g}})^{2}\Big{]}, (S4)

where the Landau parameters satisfy the conditions u>0u>0 and v,w>uv,w>-u, in order for the action to be bounded.

As discussed in the main text, within the Gaussian variational approach, we choose a trial action

𝒮0\displaystyle\mathcal{S}_{0} =12LTi,j𝚫^i𝒢i,j1𝚫^j,\displaystyle=\frac{1}{2}\frac{L}{T}\sum_{i,j}\hat{\boldsymbol{\Delta}}_{i}^{\dagger}\,\mathcal{G}_{i,j}^{-1}\,\hat{\boldsymbol{\Delta}}_{j}, (S5)

that is characterized by the inverse Green’s function

𝒢ij1\displaystyle\mathcal{G}_{ij}^{-1} =Gi1δij+fij0MA1g,\displaystyle=G_{i}^{-1}\,\delta_{ij}+f_{ij}^{0}M^{A_{1g}}, (S6)
Gi1\displaystyle G_{i}^{-1} =2RiMA1g+2n𝔾ΦinMn+n𝔾(ϕ¯inmn+h.c.),\displaystyle=2R_{i}M^{A_{1g}}+2\sum_{n\in\mathbb{G}_{\mathbb{R}}}\Phi_{i}^{n}M^{n}+\sum_{n\in\mathbb{G}_{\mathbb{C}}}\left(\bar{\phi}_{i}^{n}m^{n}+\mathrm{h.c.}\right), (S7)

which contains all variational parameters. Recall that we use Φin\Phi_{i}^{n} for real-valued variational composite order parameters and (ϕin,ϕ¯in)(\phi_{i}^{n},\bar{\phi}_{i}^{n}) for the complex-valued ones, where nn denotes the irreducible representation (IR) according to which the composite transforms. Moreover, we also define the mass renormalization parameter Ri=r0+ΦiA1gR_{i}=r_{0}+\Phi_{i}^{A_{1g}}. For convenience of notation, we introduce the IR sets 𝔾={A2g,B1g,B2g}\mathbb{G}_{\mathbb{R}}=\{A_{2g},B_{1g},B_{2g}\} and 𝔾={A1g,B1g,B2g}\mathbb{G}_{\mathbb{C}}=\{A_{1g},B_{1g},B_{2g}\} in Eq. (S7). Note that the local inverse Green’s function (S7) is identical to Eq. (5) in the main text.

Since the trial action (S5) is Gaussian, it is straightforward to evaluate the variational free energy [Eq. (6) in the main text]

Fv\displaystyle F_{v} =TlogZ0+T𝒮𝒮00,\displaystyle=-T\log Z_{0}+T\langle\mathcal{S}-\mathcal{S}_{0}\rangle_{0}, (S8)

with Z0=D(𝚫,𝚫¯)e𝒮0Z_{0}=\int D(\boldsymbol{\Delta},\bar{\boldsymbol{\Delta}})\,e^{-\mathcal{S}_{0}} being the partition function of the trial action, see for example Ref. [43] for technical details. In the real-space representation (S5), it is convenient to promote the 44-component field 𝚫^i\hat{\boldsymbol{\Delta}}_{i} to a (4N)(4N)-component field via 𝚫^^=iP^i𝚫^i\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\boldsymbol{\Delta}}}}=\sum_{i}\hat{P}_{i}\hat{\boldsymbol{\Delta}}_{i}, where the projector P^i\hat{P}_{i} is a (4N)×4(4N)\times 4 dimensional matrix whose elements are either 0 or 11, and P^iTP^j=δij𝟙4\hat{P}_{i}^{T}\hat{P}_{j}=\delta_{ij}\mathbbm{1}_{4}. The resulting variational free energy FvF_{v}, up to an unimportant constant, is given by

Fv\displaystyle F_{v} =T2logdet(𝒢^^1)+Ti{2[r0Ri+U~A1gGiiA1g]GiiA1g\displaystyle=\frac{T}{2}\log\det\Big{(}\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}}^{-1}\Big{)}+T\!\sum_{i}\!\Big{\{}2\big{[}r_{0}-R_{i}\!+\tilde{U}_{A_{1g}}G_{ii}^{A_{1g}}\big{]}G_{ii}^{A_{1g}}
2n𝔾(ΦinU~nGiin)Giinn𝔾[(ϕinu~ngiin)g¯iin+c.c.]}.\displaystyle-2\sum_{n\in\mathbb{G}_{\mathbb{R}}}\left(\Phi_{i}^{n}\!-\tilde{U}_{n}G_{ii}^{n}\right)G_{ii}^{n}-\sum_{n\in\mathbb{G}_{\mathbb{C}}}\big{[}\left(\phi_{i}^{n}-\tilde{u}_{n}g_{ii}^{n}\right)\bar{g}_{ii}^{n}+\mathrm{c.c.}\big{]}\!\Big{\}}. (S9)

Here, 𝒢^^\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}} is the inverse of 𝒢^^1=i,jP^i𝒢ij1P^jT\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}}^{-1}=\sum_{i,j}\hat{P}_{i}\mathcal{G}_{ij}^{-1}\hat{P}_{j}^{T} and GijnG_{ij}^{n} is given by the decomposition of 𝒢ij=P^iT𝒢^^P^j\mathcal{G}_{ij}=\hat{P}_{i}^{T}\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}}\hat{P}_{j} onto the symmetry channels according to

𝒢ij\displaystyle\mathcal{G}_{ij} =2n{A1g,𝔾}GijnMn+n𝔾[gijn(mn)+g¯ijnmn].\displaystyle=2\!\!\!\sum_{n\in\{A_{1g},\mathbb{G}_{\mathbb{R}}\}}\!\!G_{ij}^{n}M^{n}+\!\!\sum_{n\in\mathbb{G}_{\mathbb{C}}}\!\!\Big{[}g_{ij}^{n}(m^{n})^{\dagger}\!+\bar{g}_{ij}^{n}m^{n}\Big{]}. (S10)

Finally, in Eq. (S9), the effective interaction parameters {U~n,u~n}=TL{Un,un}\left\{\tilde{U}_{n},\tilde{u}_{n}\right\}=\frac{T}{L}\left\{U_{n},u_{n}\right\} in each symmetry channel are given by

UA1g\displaystyle U_{A_{1g}} =3u+v+w,\displaystyle=3u+v+w, uA1g\displaystyle u_{A_{1g}} =uv+w,\displaystyle=u-v+w, UA2g\displaystyle U_{A_{2g}} =u+3vw,\displaystyle=u+3v-w,
UB1g\displaystyle U_{B_{1g}} =uv+3w,\displaystyle=u-v+3w, uB1g\displaystyle u_{B_{1g}} =u+v+w,\displaystyle=u+v+w,
UB2g\displaystyle U_{B_{2g}} =uvw,\displaystyle=u-v-w, uB2g\displaystyle u_{B_{2g}} =u+vw,\displaystyle=u+v-w, (S11)

which agree with the expressions found in the bulk case [43].

SII details of the free energy minimization

In this section, we present a few technical details related to the minimization of the free energy (S9). As emphasized in the main text, we model the domain wall through the boundary conditions Φ1B1g=ΦNB1g=Φ0B1g\Phi_{1}^{B_{1g}}=-\Phi_{N}^{B_{1g}}=\Phi_{0}^{B_{1g}}, R1=RN=R0R_{1}=R_{N}=R_{0}, and all other fields Φ1,Nn=ϕ1,Nn=0\Phi_{1,N}^{n}=\phi_{1,N}^{n}=0, where Φ0B1g\Phi_{0}^{B_{1g}} and R0R_{0} are the corresponding bulk values.

The free energy (S9) effectively depends only on the parameters TT, t^0t0/uT/L\hat{t}_{0}\equiv t_{0}/\sqrt{uT/L}, {v,w}/u\{v,w\}/u, and {r0,Ri,Φin,ϕin}/t0\{r_{0},R_{i},\Phi_{i}^{n},\phi_{i}^{n}\}/t_{0}. In the spirit of the Ginzburg-Landau approach, we assume the most important temperature dependence to be in r0=a0(TT0)r_{0}=a_{0}(T-T_{0}), and we set T=T0T=T_{0} elsewhere, with T0T_{0} denoting the bare SC transition temperature. To facilitate the minimization of the free energy (S9), we supply the minimizer with the gradient expressions given by

FvXi\displaystyle\frac{\partial F_{v}}{\partial X_{i}} =Tj{VjA1gGjjA1gXi+n𝔾VjnGjjnXi\displaystyle=T\!\sum_{j}\!\Bigg{\{}V_{j}^{A_{1g}}\frac{\partial G_{jj}^{A_{1g}}}{\partial X_{i}}+\sum_{n\in\mathbb{G}_{\mathbb{R}}}V_{j}^{n}\frac{\partial G_{jj}^{n}}{\partial X_{i}}
+n𝔾(vjngjjnXi+c.c.)},\displaystyle\quad+\sum_{n\in\mathbb{G}_{\mathbb{C}}}\left(v_{j}^{n}\frac{\partial g_{jj}^{n}}{\partial X_{i}}+\mathrm{c.c.}\right)\Bigg{\}}, (S12)

where Xi{Ri,Φin,ϕin,ϕ¯in}X_{i}\in\{R_{i},\Phi_{i}^{n},\phi_{i}^{n},\bar{\phi}_{i}^{n}\} can be any of the variational parameters. In writing Eq. (S12), we exploited the fact that the partial derivatives FvXi|Gn,gn=0\frac{\partial F_{v}}{\partial X_{i}}\big{|}_{G^{n},g^{n}}=0 vanish [43]. Moreover, we defined

VjA1g\displaystyle V_{j}^{A_{1g}} =2[r0Rj+2U~A1gGjjA1g],\displaystyle=2\big{[}r_{0}-R_{j}\!+2\tilde{U}_{A_{1g}}G_{jj}^{A_{1g}}\big{]}, vjn\displaystyle v_{j}^{n} =(ϕ¯jn2u~ng¯jjn),\displaystyle=-\left(\bar{\phi}_{j}^{n}-2\tilde{u}_{n}\bar{g}_{jj}^{n}\right),
Vjn\displaystyle V_{j}^{n} =2(Φjn2U~nGjjn).\displaystyle=-2\left(\Phi_{j}^{n}\!-2\tilde{U}_{n}G_{jj}^{n}\right). (S13)

To determine the remaining derivatives in Eq. (S12), we use the Green’s function (S10) to identify

Gjjn\displaystyle G_{jj}^{n} =12tr[𝒢jjMn],\displaystyle=\frac{1}{2}\mathrm{tr}\left[\mathcal{G}_{jj}M^{n}\right], gjjn\displaystyle g_{jj}^{n} =12tr[𝒢jjmn].\displaystyle=\frac{1}{2}\mathrm{tr}\left[\mathcal{G}_{jj}m^{n}\right]. (S14)

Then, using the introduced projector matrix P^i\hat{P}_{i}, we obtain the relationship

𝒢jjXi\displaystyle\frac{\partial\mathcal{G}_{jj}}{\partial X_{i}} =P^jT𝒢^^𝒢^^1Xi𝒢^^P^j=i1i2𝒢ji1𝒢i1i21Xi𝒢i2j,\displaystyle=-\hat{P}_{j}^{T}\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}}\frac{\partial\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}}^{-1}}{\partial X_{i}}\hat{\vphantom{\rule{1.0pt}{6.57643pt}}\smash{\hat{\mathcal{G}}}}\hat{P}_{j}=-\sum_{i_{1}i_{2}}\mathcal{G}_{ji_{1}}\frac{\partial\mathcal{G}_{i_{1}i_{2}}^{-1}}{\partial X_{i}}\mathcal{G}_{i_{2}j}, (S15)

which leads to the expressions

𝒢jjRi\displaystyle\frac{\partial\mathcal{G}_{jj}}{\partial R_{i}} =2𝒢jiMA1g𝒢ij,\displaystyle=-2\mathcal{G}_{ji}M^{A_{1g}}\mathcal{G}_{ij}, 𝒢jjΦin\displaystyle\frac{\partial\mathcal{G}_{jj}}{\partial\Phi_{i}^{n}} =2𝒢jiMn𝒢ij,\displaystyle=-2\mathcal{G}_{ji}M^{n}\mathcal{G}_{ij},
𝒢jjϕ¯in\displaystyle\frac{\partial\mathcal{G}_{jj}}{\partial\bar{\phi}_{i}^{n}} =𝒢jimn𝒢ij.\displaystyle=-\mathcal{G}_{ji}m^{n}\mathcal{G}_{ij}. (S16)

It is now straightforward to compute the derivatives in Eq. (S14) by using Eq. (S16); we find, for instance

GjjnRi\displaystyle\frac{\partial G_{jj}^{n}}{\partial R_{i}} =tr[𝒢jiMA1g𝒢ijMn],\displaystyle=-\mathrm{tr}\left[\mathcal{G}_{ji}M^{A_{1g}}\mathcal{G}_{ij}M^{n}\right],
gjjnRi\displaystyle\frac{\partial g_{jj}^{n}}{\partial R_{i}} =tr[𝒢jiMA1g𝒢ijmn],\displaystyle=-\mathrm{tr}\left[\mathcal{G}_{ji}M^{A_{1g}}\mathcal{G}_{ij}m^{n}\right], (S17)

and similar expressions for the other variational parameters.