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

Interfacial Magnon-Mediated Superconductivity in Twisted Bilayer Graphene

Bjørnulf Brekke, Asle Sudbø and Arne Brataas Center for Quantum Spintronics, Department of Physics, Norwegian University of Science and Technology, NO-7491 Trondheim, Norway
Abstract

The interfacial coupling between electrons and magnons in adjacent layers can mediate an attractive electron-electron interaction and induce superconductivity. We consider magic-angle twisted bilayer graphene sandwiched between two ferromagnetic insulators to optimize this effect. As a result, magnons induce an interlayer superconducting state characterized by pp-wave symmetry. We investigate two candidate ferromagnets. The van der Waals ferromagnet CrI3 stands out because it allows compression to tune the superconducting state with an exponential sensitivity. This control adds a new dimension to the tunability of twisted bilayer graphene. Our results open a new path for exploring magnon-induced superconductivity.

Heterostructures of ferromagnets (FM) and conductors are currently attracting considerable attention in spintronics. The interfacial coupling between the localized spins and itinerant electrons gives rise to intriguing phenomena such as RKKY interactions, spin-transfer, and spin-pumping [1, 2]. Furthermore, the coupling between electrons and magnons can mediate an attractive electron-electron interaction [3, 4, 5]. This effect is analogous to the electron-phonon coupling in conventional Bardeen-Cooper-Schrieffer (BCS) superconductivity [6]. Superconductivity mediated by magnons has been studied experimentally in different materials [7, 8, 9]. Furthermore, superconductivity mediated by (antiferromagnetic) magnons might also appear in certain high-TcT_{c} superconductors [10, 11].

Superconductivity induced by interfacial coupling to magnons could exist in various material combinations. Examples are normal metals coupled to ferro- and antiferromagnets [3, 5], as well as ferromagnets and antiferromagnets coupled to the surface of topological insulators [12, 4, 13, 14]. The ferromagnetic case has also been experimentally studied [15, 16], showing a superconducting state with a critical temperature TcT_{c} significantly higher than the intrinsic superconductivity of two materials. These studies consider either surface effects or monolayer conductors.

Systems designed for interfacial magnon-mediated superconductivity require specific properties. In general, superconducting critical temperatures TcT_{c} are exponentially sensitive to interaction strength. In the present case, this is the magnon-mediated electron-electron interaction. Therefore, the electron states should be localized at the interface. Thus, the conducting layer should be as thin as possible yet stable. Furthermore, the electron density of states (DOS) should be large at the Fermi level. In fulfilling these conditions, twisted bilayer graphene (TBG) stands out as an ideal candidate [17].

Twisted bilayer graphene is a two-dimensional material. This renders the electron-magnon-induced effects in TBG more robust than in 3D normal metals, where interactions are constrained to the surface. The relative twist angle of the two graphene layers creates a long-period moiré pattern that, in turn, gives rise to flat electronic bands at certain magic angles [18, 19, 20]. Flat bands at the magic angles greatly enhance the electron DOS. TBG is, therefore, a laboratory for studying the transition from weak- to strong coupling superconductivity by tuning the twist angle. Graphene can couple to conventional ferromagnets [21, 22, 23]. TBG is also an interesting component of van der Waals heterostructures such as spin valves [24, 25]. Moreover, it is an intrinsic superconductor with a critical temperature of about Tc=1.7T_{c}=1.7 K at half filling [26]. The underlying mechanism is still under debate, and the explanations range from phonon-mediated superconductivity [27, 28, 29, 30] to non-BCS type mechanisms [31, 32, 33].

In this Letter, we consider another path to superconductivity in TBG via magnons in adjacent layers. Fig. 1 shows twisted bilayer graphene sandwiched between two identical ferromagnets with oppositely aligned magnetization. The interfacial coupling to magnons gives rise to an effective electron-electron interaction. In combination, TBG´s valley degree of freedom causes a multicomponent superconductor. We find that two of the interlayer coupling channels are suitable for Cooper pair formation. Using a BCS model, we find a pp-wave superconducting state with a critical temperature of the same order of magnitude as that of the intrinsic mechanism.

Refer to caption
Figure 1: We consider a heterostructure consisting of twisted bilayer graphene (TBG) sandwiched between two oppositely aligned ferromagnets (FM).

To describe the heterostructure, we use a Hamiltonian H=HTBG+HFM+HintH=H_{\mathrm{TBG}}+H_{\mathrm{FM}}+H_{\mathrm{int}}, where the first term describes the electrons in the twisted bilayer of graphene sheets, and the second term describes the magnons in the top and bottom layer ferromagnets. The last term describes the interfacial coupling between the ferromagnets and the graphene layers.

We consider a continuum model for the ferromagnets given by

HFM=d2𝒓[J(𝒎)2Kzmz2)],\displaystyle H_{\mathrm{FM}}=\int\mathrm{d}^{2}\bm{r}\left[J(\nabla\bm{m})^{2}-K_{z}m_{z}^{2}\right)]\,, (1)

where 𝐦\mathbf{m} is the magnetization, J>0J>0 is the exchange coupling, and the easy-axis anisotropy is parametrized by Kz>0K_{z}>0. A Holstein-Primakoff transformation to second order in magnon operators and a subsequent Fourier transform yields the magnon Hamiltonian

HFM=𝒒ω𝒒(a𝒒a𝒒+b𝒒b𝒒),\displaystyle\begin{split}H_{\mathrm{FM}}=&\sum_{\bm{q}}\omega_{\bm{q}}\left(a^{\dagger}_{\bm{q}}a_{\bm{q}}+b_{\bm{q}}^{\dagger}b_{\bm{q}}\right)\,,\end{split} (2)

where a𝒒()a_{\bm{q}}^{(\dagger)} and b𝒒()b_{\bm{q}}^{(\dagger)} are the magnon annihilation (creation) operators of momentum 𝒒\bm{q} in the top and bottom layer, respectively. The magnon dispersion is ω𝒒=2JMq2+2MKz\omega_{\bm{q}}=2JMq^{2}+2MK_{z}, where MM is the ground state magnetization.

TBG consists of two graphene monolayers with a relative twist angle θ\theta, as shown in Fig. 2. In the decoupled limit, the top layer has two Dirac cones K1K_{1} and K1K^{\prime}_{1}. Similarly, the bottom layer has two Dirac cones K2K_{2} and K2K^{\prime}_{2}. Because of the relative twist, the Dirac points are related by the three vectors 𝒒1=kθ(0,1)T\bm{q}_{1}=k_{\theta}(0,-1)^{T}, 𝒒2=kθ(3/2,1/2)T\bm{q}_{2}=k_{\theta}\left(\sqrt{3}/2,1/2\right)^{T} and 𝒒3=kθ(3/2,1/2)T\bm{q}_{3}=k_{\theta}\left(-\sqrt{3}/2,1/2\right)^{T}. Here, kθ=2kDsin(θ/2)k_{\theta}=2k_{D}\sin{\theta/2} is the magnitude of 𝒒j\bm{q}_{j}. The coefficient kDk_{D} is the magnitude of the Dirac cone momenta in the monolayers. For decoupled layers, the electrons at crystal momentum 𝒌\bm{k} near K1K_{1} or K2K_{2} are governed by the single layer Dirac Hamiltonian hK(𝒌)=v𝝈𝒌h^{K}(\bm{k})=v\bm{\sigma}^{*}\cdot\bm{k}, whereas the electrons near the cones K1K^{\prime}_{1} and K2K^{\prime}_{2} are governed by hK(𝒌)=v𝝈𝒌h^{K^{\prime}}(\bm{k})=-v\bm{\sigma}\cdot\bm{k}. Here, 𝝈\bm{\sigma} is the Pauli matrix vector, and vv is the graphene Fermi velocity.

To model the low-energy electron bands of TBG, we use the Bistritzer-MacDonald Hamiltonian [20]. In the following, we present the main features required to apply the model for magnon-induced superconductivity. We start with an effective spin-degenerate 4×44\times 4 Hamiltonian describing the electrons at cone K1K_{1} and interlayer hopping in momentum space to cone K2K_{2}. It is given by

𝒌K1K2=[hK(𝒌)T1T2T3T1hK(𝒌1)00T20hK(𝒌2)0T300hK(𝒌3)].\displaystyle\mathcal{H}_{\bm{k}}^{K_{1}K_{2}}=\begin{bmatrix}h^{K}(\bm{k})&T_{1}&T_{2}&T_{3}\\ T_{1}^{\dagger}&h^{K}(\bm{k}_{1})&0&0\\ T_{2}^{\dagger}&0&h^{K}(\bm{k}_{2})&0\\ T_{3}^{\dagger}&0&0&h^{K}(\bm{k}_{3})\end{bmatrix}\,. (3)

The three other Hamiltonians describing electrons similarly at the cones (K1,K2,K2)(K_{1}^{\prime},K_{2},K_{2}^{\prime}) must also be considered (see Supplemental Material). In Eq. (3), each element is a 2×22\times 2 matrix in sublattice space. The Hamiltonian acts on four two-component spinors (ψK1(𝒌),ψK2(𝒌1),ψK2(𝒌2),ψK2(𝒌3))T\left(\psi_{K_{1}}(\bm{k}),\psi_{K_{2}}(\bm{k}_{1}),\psi_{K_{2}}(\bm{k}_{2}),\psi_{K_{2}}(\bm{k}_{3})\right)^{T}. Here, the first spinor component describes electrons near K1K_{1}, whereas the other three components describe the electrons near K2K_{2}. The crystal momentum 𝒌j𝒌𝒒j\bm{k}_{j}\equiv\bm{k}-\bm{q}_{j}, where 𝒌\bm{k} is measured relative to the top layer‘s Dirac cone K1\mathrm{K}_{1}.

Refer to caption
Figure 2: The Brillouin zones of two graphene layers twisted by an angle θ\theta. These vectors constitute the moirè Brillouin zone with Dirac cones KM and KM{}^{\prime}_{\mathrm{M}}.

The Hamiltonian at K1K_{1} couples to the Dirac cone K2K_{2} through three hopping processes. The interlayer hopping is captured by the hopping matrices T1=w(σ0+σx)T_{1}=w(\sigma_{0}+\sigma_{x}), T2=w(σ01/2σx3/2σy)T_{2}=w\left(\sigma_{0}-1/2\sigma_{x}-\sqrt{3}/2\sigma_{y}\right) and T3=w(σ01/2σx3/2σy)T_{3}=w\left(\sigma_{0}-1/2\sigma_{x}-\sqrt{3}/2\sigma_{y}\right). The hopping strength w113w\approx 113 meV [34].

The Hamiltonian in Eq. (3) exhibits low-energy eigenstates

ΨK1K2=(1,h11T1,h21T2,h31T3)T,\displaystyle\Psi^{K_{1}K_{2}}=\begin{pmatrix}1,&-h_{1}^{-1}T_{1}^{\dagger},&-h_{2}^{-1}T_{2}^{\dagger},&-h_{3}^{-1}T_{3}^{\dagger}\end{pmatrix}^{T}, (4)

where we used the shorthand notation hj=hK(qj)h_{j}=h^{K}(-q_{j}). Projecting the Hamiltonian on these low-energy states, we get an effective 2×22\times 2 sublattice space Hamiltonian on the form

ΨK1K2|𝒌K1K2|ΨK1K2=v𝝈𝒌.\displaystyle\begin{split}\bra{\Psi^{K_{1}K_{2}}}\mathcal{H}_{\bm{k}}^{K_{1}K_{2}}\ket{\Psi^{K_{1}K_{2}}}=v^{\star}\bm{\sigma}^{*}\cdot\bm{k}.\end{split} (5)

This effective Hamiltonian has the form of a single-layer Dirac Hamiltonian with a renormalized velocity

vv=13α21+6α2,\displaystyle\frac{v^{\star}}{v}=\frac{1-3\alpha^{2}}{1+6\alpha^{2}}\,, (6)

where α=w/vkθ\alpha=w/vk_{\theta}. Note how α2=1/3\alpha^{2}=1/3 yields a vanishing Fermi velocity and flat bands. This value corresponds to the largest ”magic angle.”

The Hamiltonian in Eq. (3) models electrons at cone K1K_{1} in the top layer and three allowed interlayer hoppings to K2K_{2}. As noted above, the related Hamiltonians for the K1K_{1}^{\prime} point in the top layer and K2K_{2} and K2K_{2}^{\prime} in the bottom layer must also be considered (see Supplemental Material). Diagonalizing all four effective 2×22\times 2 Dirac Hamiltonians, we find the resulting Hamiltonian

H~nI,s(𝒌)=εn𝒌c𝒌nIsc𝒌nIs,\displaystyle\tilde{H}_{n}^{I,s}(\bm{k})=\varepsilon_{n\bm{k}}c^{\dagger}_{\bm{k}nIs}c_{\bm{k}nIs}, (7)

with a linear electron dispersion εn𝒌=nv|𝒌|\varepsilon_{n\bm{k}}=nv^{\star}\absolutevalue{\bm{k}}. The cc-operators are creation- and annihilation operators for the upper n=1n=1 and lower n=1n=-1 bands. The eigenstates are superpositions of states at Dirac cones in both layers. Each state is ”based” at one of the four cones I{K1,K1,K2,K2I\in\{K_{1},K_{1}^{\prime},K_{2},K^{\prime}_{2}} with a threefold contribution from the opposing layer. The relative weight of the contributions depends on the twist angle θ\theta and interlayer hopping strength ww. We index the eigenstates according to the cone at which it is based. Furthermore, ss is the spin index, and the crystal momentum 𝒌\bm{k} is taken with respect to cone II.

Refer to caption
Figure 3: An intersection of the electronic bands at constant energy in the moirè Brillouin zone, with moirè cones KMK_{M} and KMK^{\prime}_{M} relative to the center γ\gamma. The bands are split due to the interfacial s-d coupling giving rise to Eq. (9). The bands are doubly degenerate with states labeled according to the cone at which it is based and spin.

We model the interfacial coupling with a conventional s-d Hamiltonian

Hint=Jsdd2𝒓𝒎(𝒓)𝒔(𝒓),\displaystyle H_{\mathrm{int}}=J_{\mathrm{s-d}}\int\mathrm{d}^{2}\bm{r}\bm{m}(\bm{r})\cdot\bm{s}(\bm{r}), (8)

where 𝒔\bm{s} is the spin operator of the itinerant electrons and JsdJ_{\mathrm{s-d}} is the coupling strength.

As explained above, we take into account spin fluctuations by means of a Holstein-Primakoff transformation. We only consider the interfacial coupling between the ferromagnets and their nearest graphene layer. For instance, the first component of the electron state in Eq. (4) couples to the top ferromagnetic layer, whereas the three next components couple to the bottom layer. The interfacial coupling gives rise to a layer-dependent spin splitting

Hss=Δss𝒌,I,sls(c𝒌Isc𝒌Is),\displaystyle H_{\mathrm{ss}}=\Delta_{ss}\sum_{\bm{k},I,s}ls\left(c_{\bm{k}Is}^{\dagger}c_{\bm{k}Is}\right), (9)

where the spin splitting is given by

Δss=JsdM(16α2)1+6α2.\displaystyle\Delta_{ss}=\frac{J_{\mathrm{s-d}}M(1-6\alpha^{2})}{1+6\alpha^{2}}. (10)

The layer index ll can be extracted directly from II. It takes the value l=1l=1 for I{K1,K1}I\in\{K_{1},K^{\prime}_{1}\} and l=1l=-1 for I{K2,K2}I\in\{K_{2},K^{\prime}_{2}\}. The spin index s=1s=1 for spin up and s=1s=-1 for spin down. The spin-split electron dispersion in the moirè Brillouin zone is shown in Fig. 3. In the case of a monolayer conductor sandwiched between two oppositely aligned ferromagnets, the spin splitting vanishes exactly. This is not necessarily the case for a bilayer, as seen from Eq. (9). Each of the electron eigenstates is asymmetrically delocalized in the two layers. Hence, the net magnetic field shifts the energy of the electron states. The sign of the shift depends on spin and the layer in which the state is ”based.” At the special twist angle α2=1/6\alpha^{2}=1/6, the states are symmetrically distributed in the two layers such that the spin shift cancels.

We next consider the electron-magnon coupling

Hem=𝒒,𝒌,I[gI,𝒌,𝒒a(c𝒌+𝒒Ic𝒌I)a𝒒+h.c.+gI,𝒌,𝒒b(c𝒌𝒒Ic𝒌I)b𝒒+h.c.].\displaystyle\begin{split}H_{\mathrm{e-m}}=&\sum_{\bm{q},\bm{k},I}\left[g_{I,\bm{k},\bm{q}}^{a}\left(c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)a_{\bm{q}}+h.c.\right.\\ &+\left.g_{I,\bm{k},-\bm{q}}^{b}\left(c_{\bm{k}-\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)b^{\dagger}_{\bm{q}}+h.c.\right]\,.\end{split} (11)

The coupling parameters gI,𝒌,𝒒ag_{I,\bm{k},\bm{q}}^{a} and gI,𝒌,𝒒bg_{I,\bm{k},-\bm{q}}^{b} depend on the index II, in addition to the ferromagnetic layers aa and bb to which it couples. Their full form is given in the Supplemental material.

We derive an effective electron-electron interaction between electrons in state II and II^{\prime} via a Schrieffer-Wolff transformation [35], obtaining

Heff=𝒌,𝒌,I,IV𝒌𝒌IIc𝒌Ic𝒌Ic𝒌Ic𝒌I,\displaystyle H_{\mathrm{eff}}=\sum_{\bm{k},\bm{k}^{\prime},I,I^{\prime}}V_{\bm{k}\bm{k}^{\prime}II^{\prime}}c_{\bm{k}^{\prime}I\uparrow}^{\dagger}c_{-\bm{k}^{\prime}I^{\prime}\downarrow}^{\dagger}c_{-\bm{k}I\downarrow}c_{\bm{k}I^{\prime}\uparrow}, (12)

where the interaction strength V𝒌𝒌IIV_{\bm{k}\bm{k}^{\prime}II^{\prime}} is given by

V𝒌𝒌II=2ω𝒌+𝒌(j{a,b}gI,𝒌,𝒌+𝒌,jgI,𝒌,𝒌+𝒌j)ω𝒌+𝒌2(ε𝒌ε𝒌)2.\displaystyle\begin{split}V_{\bm{k}\bm{k}^{\prime}II^{\prime}}=\frac{2\omega_{\bm{k^{\prime}+k}}\left(\sum_{j\in\{a,b\}}g_{I,-\bm{k},\bm{k}^{\prime}+\bm{k},}^{j}g_{I^{\prime},-\bm{k},\bm{k}^{\prime}+\bm{k}}^{j}\right)}{\omega^{2}_{\bm{k}^{\prime}+\bm{k}}-(\varepsilon_{\bm{k}^{\prime}}-\varepsilon_{\bm{k}})^{2}}.\end{split} (13)

The rich I,II,I^{\prime} structure of Eq. (13) yields, in principle, a large number of possible pairing channels. However, not all of them are suitable for the formation of Cooper pairs. For instance, states based in the same layer with opposite spins are spin split and thus not suitable for spin-unpolarized Cooper-pairing either in the spin-singlet or spin-triplet channels. Spin-unpolarized Cooper-pairing is the only possibility when the electron-magnon coupling originates from collinear spin ground states in the magnetic insulator [36]. Hence, we focus on the coupling between electrons in different layers. This excludes half of the pairing channel candidates. Furthermore, requiring the Cooper pairs to have a zero net momentum with respect to the moiré Brillouin zone leaves two possible pairing channels. These are inter-layer intra-valley pairs denoted by

{I=K1,I¯=K2},\displaystyle\left\{I=K_{1},\bar{I}=K_{2}\right\}, {I=K1,I¯=K2}.\displaystyle\left\{I=K^{\prime}_{1},\bar{I}=K^{\prime}_{2}\right\}. (14)

Within these pairing channels, the interaction can be further decomposed in terms of pairing symmetry. To that end, we approximate the magnon frequency to be constant ω𝒌+𝒌=ωM\omega_{\bm{k}^{\prime}+\bm{k}}=\omega_{M}. In this approximation, the full angular dependence originates from the coupling constants gg in Eq. (13). The effective interaction decomposes into an ss-, a pp-, and a dd-wave component. In the low-frequency limit (ε𝒌ε𝒌)20(\varepsilon_{\bm{k}^{\prime}}-\varepsilon_{\bm{k}})^{2}\rightarrow 0, the ss- and dd-wave components are repulsive. In contrast, the pp-wave symmetry component is attractive and enables Cooper pair formation. Hence, we expect TBG to exhibit an interlayer magnon-mediated superconducting state with pp-wave symmetry.

We now give an estimate of the critical temperature from the conventional BCS expression

kBTc1.13ωMe1λ,\displaystyle k_{B}T_{c}\approx 1.13\omega_{M}e^{-\frac{1}{\lambda}}, (15)

where ωM\omega_{M} is the characteristic magnon frequency. The coupling λ=VND\lambda=VN^{\prime}_{D} depends on the effective interaction VV and the DOS NDN^{\prime}_{D} per valley per layer per spin. The electron DOS is enhanced near the magic angle due to the flat energy bands. Ref. [37] reports a total DOS ND100eV1nm2N_{D}\gtrapprox 100\;\mathrm{eV}^{-1}\mathrm{nm}^{-2} close to the magic angle. This suggests a DOS ND=12.5eV1nm2N_{D}^{\prime}=12.5\;\mathrm{eV}^{-1}\mathrm{nm}^{-2}. Although the BCS theory does not predict the critical temperature, it may be used to obtain estimates of TcT_{c}. An important feature of Eq. (15) is the non-perturbative renormalization of the magnon energy scale, in that TcT_{c} depends exponentially on the inverse of the DOS and the interfacial s-d coupling.

The repulsive Coulomb interaction can be detrimental to the superconducting state. TBG´s Coulomb interaction is largely screened at the magic angle for long wavelengths due to a large twist-angle dependent dielectric constant ϵ>250\epsilon>250 [38]. The Coulomb coupling strength is μ=VC(kθ)ND1\mu=V_{C}(k_{\theta})N_{D}\approx 1. It slightly exceeds the attractive magnon-mediated interaction. However, for two reasons, the superconducting state is robust in the presence of this Coulomb repulsion. First, the superconducting gap function is of interlayer pp-wave symmetry. Hence, the Cooper pairs circumvent the significant on-site ss-wave contribution of the Coulomb interaction. However, we will not consider the decomposition of the Coulomb interaction as the approximation is already at a crude level. Second, the Coulomb interaction is frequency independent at the scale of the magnon cut-off frequencies. To account for this, we adopt the Morel-Anderson model [39] to find an effective coupling strength

μ=μ1+μln(ωpωM)0.13.\displaystyle\mu^{*}=\frac{\mu}{1+\mu\ln{\frac{\omega_{p}}{\omega_{M}}}}\approx 0.13. (16)

Here, we used the observed interband plasmon frequency ωp200meV\omega_{p}\approx 200\;\mathrm{meV} as the Coulomb interaction cut off [40]. The effective interaction strength takes the value λ=λμ\lambda^{*}=\lambda-\mu^{*}.

Interfacial coupling between graphene and ferromagnets has been studied theoretically and experimentally for numerous materials [41, 42, 22, 43, 44, 45, 46]. Hence, there are several candidates. Here, we consider two specific ferromagnets.

EuO is a ferromagnetic semiconductor with Curie temperature Tc=69T_{c}=69 K. It has an fcc unit cell with lattice constant 5.15.1 Å. Hence, two magnetic Eu2+ ions per unit cell, each with spin S=7/2S=7/2, are located at the interface and thus accessible for interfacial s-d coupling [47]. EuO thin films can be deposited on graphene epitaxially [41, 42]. The induced exchange splitting is found to be Δ=36\Delta=36 meV [48]. At the wave number kθk_{\theta}, ωM=0.15\omega_{M}=0.15 meV is an appropriate frequency cut-off. These parameters suggest an effective coupling strength λ0.8\lambda^{*}\approx 0.8 and a critical temperature TC0.5T_{C}\approx 0.5 K.

CrI3 is a van der Waals ferromagnet down to the monolayer limit [49]. The crystal has two magnetic ions per unit cell. Each magnetic ion carries a magnetic moment S=3μBS=3\mu_{B} [50]. CrI3 hosts two magnonic modes accessible for electron-magnon coupling. Their respective energies at momentum 𝒒=0\bm{q}=0 are 0.30.3 meV and 1717 meV [51]. In a graphene-CrI3 heterostructure, CrI3 is theoretically found to induce an exchange splitting of 2020 meV [46]. Considering CrI3 as the ferromagnet, we find a coupling constant λ0.4\lambda^{*}\approx 0.4 and critical temperature Tc0.3T_{c}\approx 0.3 K.

Van der Waals ferromagnets, such as CrI3, are particularly interesting candidates. This is because the interfacial exchange splitting increases significantly under compression. For CrI3, a slight decrease in the interlayer gap can enhance the exchange splitting. A moderate reduction of the interlayer distance Δd=0.5\Delta d=-0.5 Å leads to an exchange splitting of 80 meV. The splitting can reach values up to 150 meV [45]. The enhanced interfacial interaction renders the higher energy magnon branch of CrI3 accessible for electron-magnon coupling. We expect this to increase the critical temperature significantly. In this way, the van der Waals spin valve exhibits a tunable compression-controlled superconducting state that connects the weak- and strong-coupling regimes. Due to the limited validity of the weak-coupling BCS model, we do not estimate the critical temperature for the compressed heterostructure.

In conclusion, we have demonstrated that interfacial magnons can induce superconductivity in twisted bilayer graphene. The magnons yield a multicomponent superconducting state due to the valley structure of TBG. We find an attractive interlayer channel suitable for Cooper pairs with pp-wave symmetry. Moreover, we have considered two promising candidate ferromagnets, EuO and CrI3. Both exhibit a critical temperature of the same order of magnitude as the intrinsic superconducting mechanism. Using van der Waals magnets is particularly interesting because the interlayer interaction strength is tunable through compression. For this reason, the superconducting state is tunable both via the twist angle and external compression. Twisted bilayer graphene sandwiched between ferromagnets is, therefore, a promising platform in which to explore magnon-mediated superconductivity.

The Research Council of Norway (RCN) supported this work through its Centres of Excellence funding scheme, project number 262633, ”QuSpin”, as well as RCN project number 323766.

References

  • Bruno [1995] P. Bruno, Theory of interlayer magnetic coupling, Physical Review B 52, 411 (1995).
  • Brataas et al. [2012] A. Brataas, A. D. Kent, and H. Ohno, Current-induced torques in magnetic materials, Nat Mater 11, 372 (2012).
  • Rohling et al. [2018] N. Rohling, E. L. Fjærbu, and A. Brataas, Superconductivity induced by interfacial coupling to magnons, Physical Review B 97, 115401 (2018).
  • Hugdal et al. [2018] H. G. Hugdal, S. Rex, F. S. Nogueira, and A. Sudbø, Magnon-induced superconductivity in a topological insulator coupled to ferromagnetic and antiferromagnetic insulators, Phys. Rev. B 97, 195438 (2018).
  • Fjærbu et al. [2019] E. L. Fjærbu, N. Rohling, and A. Brataas, Superconductivity at metal-antiferromagnetic insulator interfaces, Physical Review B 100, 125432 (2019).
  • Bardeen et al. [1957] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Theory of superconductivity, Physical review 108, 1175 (1957).
  • Saxena et al. [2000] S. Saxena, P. Agarwal, K. Ahilan, F. Grosche, R. Haselwimmer, M. Steiner, E. Pugh, I. Walker, S. Julian, P. Monthoux, et al., Superconductivity on the border of itinerant-electron ferromagnetism in UGe2, Nature 406, 587 (2000).
  • Aoki et al. [2001] D. Aoki, A. Huxley, E. Ressouche, D. Braithwaite, J. Flouquet, J.-P. Brison, E. Lhotel, and C. Paulsen, Coexistence of superconductivity and ferromagnetism in URhGe, Nature 413, 613 (2001).
  • Pfleiderer et al. [2001] C. Pfleiderer, M. Uhlarz, S. Hayden, R. Vollmer, H. v. Löhneysen, N. Bernhoeft, and G. Lonzarich, Coexistence of superconductivity and ferromagnetism in the d-band metal ZrZn2, Nature 412, 58 (2001).
  • Si et al. [2016] Q. Si, R. Yu, and E. Abrahams, High-temperature superconductivity in iron pnictides and chalcogenides, Nature Reviews Materials 1, 1 (2016).
  • Lee et al. [2006] P. A. Lee, N. Nagaosa, and X.-G. Wen, Doping a Mott insulator: Physics of high-temperature superconductivity, Reviews of modern physics 78, 17 (2006).
  • Kargarian et al. [2016] M. Kargarian, D. K. Efimkin, and V. Galitski, Amperean pairing at the surface of topological insulators, Physical Review Letters 117, 076806 (2016).
  • Erlandsen et al. [2020] E. Erlandsen, A. Brataas, and A. Sudbø, Magnon-mediated superconductivity on the surface of a topological insulator, Physical Review B 101, 094503 (2020).
  • Thingstad et al. [2021] E. Thingstad, E. Erlandsen, and A. Sudbø, Eliashberg study of superconductivity induced by interfacial coupling to antiferromagnets, Phys. Rev. B 104, 014508 (2021).
  • Gong et al. [2015] X.-X. Gong, H.-X. Zhou, P.-C. Xu, D. Yue, K. Zhu, X.-F. Jin, H. Tian, G.-J. Zhao, and T.-Y. Chen, Possible p-wave superconductivity in epitaxial Bi/Ni bilayers, Chinese Physics Letters 32, 067402 (2015).
  • Gong et al. [2017] X. Gong, M. Kargarian, A. Stern, D. Yue, H. Zhou, X. Jin, V. M. Galitski, V. M. Yakovenko, and J. Xia, Time-reversal symmetry-breaking superconductivity in epitaxial bismuth/nickel bilayers, Science advances 3, e1602579 (2017).
  • Andrei and MacDonald [2020] E. Y. Andrei and A. H. MacDonald, Graphene bilayers with a twist, Nature materials 19, 1265 (2020).
  • Li et al. [2010] G. Li, A. Luican, J. Lopes dos Santos, A. Castro Neto, A. Reina, J. Kong, and E. Andrei, Observation of van hove singularities in twisted graphene layers, Nature physics 6, 109 (2010).
  • Suárez Morell et al. [2010] E. Suárez Morell, J. D. Correa, P. Vargas, M. Pacheco, and Z. Barticevic, Flat bands in slightly twisted bilayer graphene: Tight-binding calculations, Phys. Rev. B 82, 121407(R) (2010).
  • Bistritzer and MacDonald [2011] R. Bistritzer and A. H. MacDonald, Moiré bands in twisted double-layer graphene, Proceedings of the National Academy of Sciences 108, 12233 (2011).
  • Haugen et al. [2008] H. Haugen, D. Huertas-Hernando, and A. Brataas, Spin transport in proximity-induced ferromagnetic graphene, Phys. Rev. B 77, 115406 (2008).
  • Wei et al. [2016] P. Wei, S. Lee, F. Lemaitre, L. Pinel, D. Cutaia, W. Cha, F. Katmis, Y. Zhu, D. Heiman, J. Hone, et al., Strong interfacial exchange field in the graphene/EuS heterostructure, Nature materials 15, 711 (2016).
  • Wu et al. [2017] Y.-F. Wu, H.-D. Song, L. Zhang, X. Yang, Z. Ren, D. Liu, H.-C. Wu, J. Wu, J.-G. Li, Z. Jia, et al., Magnetic proximity effect in graphene coupled to a BiFeO3 nanoplate, Physical Review B 95, 195426 (2017).
  • Geim and Grigorieva [2013] A. K. Geim and I. V. Grigorieva, Van der Waals heterostructures, Nature 499, 419 (2013).
  • Cardoso et al. [2018] C. Cardoso, D. Soriano, N. A. García-Martínez, and J. Fernández-Rossier, Van der waals spin valves, Phys. Rev. Lett. 121, 067701 (2018).
  • Cao et al. [2018] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Unconventional superconductivity in magic-angle graphene superlattices, Nature 556, 43 (2018).
  • Lian et al. [2019] B. Lian, Z. Wang, and B. A. Bernevig, Twisted bilayer graphene: A phonon-driven superconductor, Phys. Rev. Lett. 122, 257002 (2019).
  • Wu et al. [2018] F. Wu, A. H. MacDonald, and I. Martin, Theory of phonon-mediated superconductivity in twisted bilayer graphene, Phys. Rev. Lett. 121, 257001 (2018).
  • Peltonen et al. [2018] T. J. Peltonen, R. Ojajärvi, and T. T. Heikkilä, Mean-field theory for superconductivity in twisted bilayer graphene, Physical Review B 98, 220504(R) (2018).
  • Choi and Choi [2018] Y. W. Choi and H. J. Choi, Strong electron-phonon coupling, electron-hole asymmetry, and nonadiabaticity in magic-angle twisted bilayer graphene, Physical Review B 98, 241412(R) (2018).
  • Oh et al. [2021] M. Oh, K. P. Nuckolls, D. Wong, R. L. Lee, X. Liu, K. Watanabe, T. Taniguchi, and A. Yazdani, Evidence for unconventional superconductivity in twisted bilayer graphene, Nature 600, 240 (2021).
  • Yankowitz et al. [2019] M. Yankowitz, S. Chen, H. Polshyn, Y. Zhang, K. Watanabe, T. Taniguchi, D. Graf, A. F. Young, and C. R. Dean, Tuning superconductivity in twisted bilayer graphene, Science 363, 1059 (2019).
  • Lu et al. [2019] X. Lu, P. Stepanov, W. Yang, M. Xie, M. A. Aamir, I. Das, C. Urgell, K. Watanabe, T. Taniguchi, G. Zhang, et al., Superconductors, orbital magnets and correlated states in magic-angle bilayer graphene, Nature 574, 653 (2019).
  • Jung et al. [2014] J. Jung, A. Raoux, Z. Qiao, and A. H. MacDonald, Ab initio theory of moiré superlattice bands in layered two-dimensional materials, Physical Review B 89, 205414 (2014).
  • Schrieffer and Wolff [1966] J. R. Schrieffer and P. A. Wolff, Relation between the Anderson and Kondo Hamiltonians, Physical Review 149, 491 (1966).
  • Mæland and Sudbø [2022] K. Mæland and A. Sudbø, Topological superconductivity mediated by skyrmionic magnons, arxiv:2211.05129  (2022).
  • Carr et al. [2019] S. Carr, S. Fang, Z. Zhu, and E. Kaxiras, Exact continuum model for low-energy electronic states of twisted bilayer graphene, Phys. Rev. Res. 1, 013001 (2019).
  • Goodwin et al. [2019] Z. A. H. Goodwin, F. Corsetti, A. A. Mostofi, and J. Lischner, Attractive electron-electron interactions from internal screening in magic-angle twisted bilayer graphene, Phys. Rev. B 100, 235424 (2019).
  • Morel and Anderson [1962a] P. Morel and P. W. Anderson, Calculation of the Superconducting State Parameters with Retarded Electron-Phonon Interaction, Phys. Rev. 125, 1263 (1962a).
  • Hesp et al. [2021] N. C. Hesp, I. Torre, D. Rodan-Legrain, P. Novelli, Y. Cao, S. Carr, S. Fang, P. Stepanov, D. Barcons-Ruiz, H. Herzig Sheinfux, et al., Observation of interband collective excitations in twisted bilayer graphene, Nature Physics 17, 1162 (2021).
  • Swartz et al. [2012] A. G. Swartz, P. M. Odenthal, Y. Hao, R. S. Ruoff, and R. K. Kawakami, Integration of the ferromagnetic insulator EuO onto graphene, ACS nano 6, 10063 (2012).
  • Averyanov et al. [2018] D. V. Averyanov, I. S. Sokolov, A. M. Tokmachev, O. E. Parfenov, I. A. Karateev, A. N. Taldenkov, and V. G. Storchak, High-temperature magnetism in graphene induced by proximity to EuO, ACS applied materials & interfaces 10, 20767 (2018).
  • Wang et al. [2015] Z. Wang, C. Tang, R. Sachs, Y. Barlas, and J. Shi, Proximity-induced ferromagnetism in graphene revealed by the anomalous Hall effect, Physical review letters 114, 016603 (2015).
  • Farooq and Hong [2019] M. U. Farooq and J. Hong, Switchable valley splitting by external electric field effect in graphene/CrI3 heterostructures, npj 2D Materials and Applications 3, 1 (2019).
  • Zhang et al. [2018] J. Zhang, B. Zhao, T. Zhou, Y. Xue, C. Ma, and Z. Yang, Strong magnetization and chern insulators in compressed graphene/cri3{\mathrm{graphene}\text{/}\mathrm{cri}}_{3} van der waals heterostructures, Phys. Rev. B 97, 085401 (2018).
  • Holmes et al. [2020] A. M. Holmes, S. Pakniyat, S. A. H. Gangaraj, F. Monticone, M. Weinert, and G. W. Hanson, Exchange splitting and exchange-induced nonreciprocal photonic behavior of graphene in CrI3-graphene van der Waals heterostructures, Physical Review B 102, 075435 (2020).
  • Mauger and Godart [1986] A. Mauger and C. Godart, The magnetic, optical, and transport properties of representatives of a class of magnetic semiconductors: The europium chalcogenides, Physics Reports 141, 51 (1986).
  • Yang et al. [2013] H. X. Yang, A. Hallal, D. Terrade, X. Waintal, S. Roche, and M. Chshiev, Proximity effects induced in graphene by magnetic insulators: First-principles calculations on spin filtering and exchange-splitting gaps, Phys. Rev. Lett. 110, 046603 (2013).
  • Huang et al. [2017] B. Huang, G. Clark, E. Navarro-Moratalla, D. R. Klein, R. Cheng, K. L. Seyler, D. Zhong, E. Schmidgall, M. A. McGuire, D. H. Cobden, et al., Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit, Nature 546, 270 (2017).
  • McGuire et al. [2015] M. A. McGuire, H. Dixit, V. R. Cooper, and B. C. Sales, Coupling of crystal structure and magnetism in the layered, ferromagnetic insulator CrI3, Chemistry of Materials 27, 612 (2015).
  • Cenker et al. [2021] J. Cenker, B. Huang, N. Suri, P. Thijssen, A. Miller, T. Song, T. Taniguchi, K. Watanabe, M. A. McGuire, D. Xiao, et al., Direct observation of two-dimensional magnons in atomically thin CrI3, Nature Physics 17, 20 (2021).
  • Morel and Anderson [1962b] P. Morel and P. W. Anderson, Calculation of the Superconducting State Parameters with Retarded Electron-Phonon Interaction, Phys. Rev. 125, 1263 (1962b).
  • Dietrich et al. [1975] O. W. Dietrich, A. J. Henderson, and H. Meyer, Spin-wave analysis of specific heat and magnetization in EuO and EuS, Phys. Rev. B 1210.1103/PhysRevB.12.2844 (1975).

Supplemental material to ”Interfacial Magnon-Mediated Superconductivity in Twisted Bilayer Graphene”

I Ferromagnet

We consider a continuum model of magnetization dynamics in a ferromagnetic layer. The Hamiltonian is

Hm=d2𝒓[J(𝒎)2Kzmz2],\displaystyle H_{m}=\int\mathrm{d}^{2}\bm{r}\left[J(\nabla\bm{m})^{2}-K_{z}m_{z}^{2}\right]\,, (S1)

where the exchange constant J>0J>0, the easy-axis anisotropy Kz>0K_{z}>0 and 𝒎\bm{m} is the magnetization. We use this model for both the top and bottom ferromagnetic layers introduced in the main text. We employ the Holstein-Primakoff transformation for the variables

mx=12(m++m),\displaystyle m_{x}=\frac{1}{2}(m^{+}+m^{-})\,, my=12i(m+m),\displaystyle m_{y}=\frac{1}{2i}(m^{+}-m^{-})\,, mz=mz.\displaystyle m_{z}=m_{z}\,. (S2)

We quantize the magnetization to quadratic order in the bosonic operators for the top layer using

m+=2Ma𝒓,\displaystyle m^{+}=\sqrt{2M}a_{\bm{r}}\,, m=2Ma𝒓,\displaystyle m^{-}=\sqrt{2M}a_{\bm{r}}^{\dagger}\,, mz=(Ma𝒓a𝒓).\displaystyle m^{z}=(M-a_{\bm{r}}^{\dagger}a_{\bm{r}})\,. (S3)

The ground state magnetization MM of the bottom layer is oppositely aligned. Hence, we quantize the operators as

m+=2Mb𝒓,\displaystyle m^{+}=\sqrt{2M}b^{\dagger}_{\bm{r}}\,, m=2Mb𝒓,\displaystyle m^{-}=\sqrt{2M}b_{\bm{r}}\,, mz=(Mb𝒓b𝒓).\displaystyle m^{z}=-(M-b_{\bm{r}}^{\dagger}b_{\bm{r}})\,. (S4)

We consider the Hamiltonian in momentum space by using the Fourier transforms

a𝒓=1𝒱𝒌ei𝒌𝒓a𝒌,\displaystyle a_{\bm{r}}=\frac{1}{\sqrt{\mathcal{V}}}\sum_{\bm{k}}e^{i\bm{k}\cdot\bm{r}}a_{\bm{k}}, b𝒓=1𝒱𝒌ei𝒌𝒓b𝒌,\displaystyle b_{\bm{r}}=\frac{1}{\sqrt{\mathcal{V}}}\sum_{\bm{k}}e^{i\bm{k}\cdot\bm{r}}b_{\bm{k}}, (S5)

where 𝒱\mathcal{V} is the area of the ferromagnet. The transformations yield the magnon Hamiltonian

Hm=𝒒ω𝒒(a𝒒a𝒒+b𝒒b𝒒),\displaystyle\begin{split}H_{m}=&\sum_{\bm{q}}\omega_{\bm{q}}\left(a^{\dagger}_{\bm{q}}a_{\bm{q}}+b_{\bm{q}}^{\dagger}b_{\bm{q}}\right)\,,\end{split} (S6)

with ω𝒒=2JMq2+2MKz\omega_{\bm{q}}=2JMq^{2}+2MK_{z} as stated in the main text. Here, and throughout the text, we set =1\hbar=1.

II Bistritzer-MacDonald model for TBG

A single sheet of graphene has two Dirac cones in the first Brillouin zone located at the high-symmetry points KK and KK^{\prime}. These cones are related by inversion symmetry and are described by Dirac Hamiltonians of opposite chirality

hK(𝒌)=v𝝈𝒌,\displaystyle h^{K}(\bm{k})=v\bm{\sigma}^{*}\cdot\bm{k}\,, hK(𝒌)=v𝝈𝒌.\displaystyle h^{K^{\prime}}(\bm{k})=-v\bm{\sigma}\cdot\bm{k}\,. (S7)

Here, 𝝈\bm{\sigma} is the vector of the Pauli matrices acting on the sublattice space. In the bilayer case, we use K1K_{1} and K1K_{1}^{\prime} for the top layer and K2K_{2} and K2K_{2}^{\prime} for the bottom layer. The Bistritzer-MacDonald (BM) model describes interlayer hopping between these Dirac cones in twisted bilayer graphene [20]. The hopping from momentum 𝒌\bm{k} relative to the cone K1K_{1} in the top layer to the bottom layer cone K2K_{2} is described by the Hamiltonian

𝒌K1K2=[hK(𝒌)T1T2T3T1hK(𝒌𝒒1)00T20hK(𝒌𝒒2)0T300hK(𝒌𝒒3)].\displaystyle\mathcal{H}_{\bm{k}}^{K_{1}K_{2}}=\begin{bmatrix}h^{K}(\bm{k})&T_{1}&T_{2}&T_{3}\\ T_{1}^{\dagger}&h^{K}(\bm{k}-\bm{q}_{1})&0&0\\ T_{2}^{\dagger}&0&h^{K}(\bm{k}-\bm{q}_{2})&0\\ T_{3}^{\dagger}&0&0&h^{K}(\bm{k}-\bm{q}_{3})\end{bmatrix}. (S8)

Here, each element is a 2×22\times 2 matrix with respect to sublattice space. The 𝒒\bm{q} vectors are defined as

𝒒1=kθ(0,1)T,\displaystyle\bm{q}_{1}=k_{\theta}(0,-1)^{T}, 𝒒2=kθ(32,12)T\displaystyle\bm{q}_{2}=k_{\theta}\left(\frac{\sqrt{3}}{2},\frac{1}{2}\right)^{T} 𝒒3=kθ(32,12)T,\displaystyle\bm{q}_{3}=k_{\theta}\left(-\frac{\sqrt{3}}{2},\frac{1}{2}\right)^{T}, (S9)

where the 𝒒\bm{q}-vector magnitude is kθ=2kDsin(θ/2)k_{\theta}=2k_{D}\sin{\theta/2}. The factor kDk_{D} is the magnitude of the Dirac cone momenta in the single layers. The interlayer hopping matrices are T1=w(σ0+σx)T_{1}=w(\sigma_{0}+\sigma_{x}), T2=w(σ01/2σx3/2σy)T_{2}=w\left(\sigma_{0}-1/2\sigma_{x}-\sqrt{3}/2\sigma_{y}\right) and T3=w(σ01/2σx3/2σy)T_{3}=w\left(\sigma_{0}-1/2\sigma_{x}-\sqrt{3}/2\sigma_{y}\right), where ww is the interlayer hopping strength.

The Hamiltonian acts on the four two-component spinors (ψK1(𝒌),ψK2(𝒌1),ψK2(𝒌2),ψK2(𝒌3))T\left(\psi_{K_{1}}(\bm{k}),\psi_{K_{2}}(\bm{k}_{1}),\psi_{K_{2}}(\bm{k}_{2}),\psi_{K_{2}}(\bm{k}_{3})\right)^{T}. Each component is a spinor in the sublattice space. The first component ψK1(𝒌)\psi_{K_{1}}(\bm{k}) is located near K1K_{1} whereas the three components ψK2(𝒌1)\psi_{K_{2}}(\bm{k}_{1}), ψK2(𝒌2)\psi_{K_{2}}(\bm{k}_{2}) and ψK2(𝒌3)\psi_{K_{2}}(\bm{k}_{3}) are located in the bottom layer near K2K_{2} at the three distinct momenta 𝒌j=𝒌𝒒j\bm{k}_{j}=\bm{k}-\bm{q}_{j}. The crystal momentum 𝒌\bm{k} is relative to the valley K1K_{1}. In this basis, we project the Hamiltonian on the two-fold low-energy eigenstate

𝚿K1K2=(1,h11T1,h21T2,h31T3)T,\displaystyle\bm{\Psi}^{K_{1}K_{2}}=\begin{pmatrix}1,&-h_{1}^{-1}T_{1}^{\dagger},&-h_{2}^{-1}T_{2}^{\dagger},&-h_{3}^{-1}T_{3}^{\dagger}\end{pmatrix}^{T}, (S10)

where hj=hK(𝒒j)h_{j}=h^{K}(-\bm{q}_{j}). The state is normalized to

|𝚿K1K2|2=1+6α2.\displaystyle\absolutevalue{\bm{\Psi}^{K_{1}K_{2}}}^{2}=1+6\alpha^{2}. (S11)

This is evident from the identity

h11T1T1h11+h21T2T2h21+h31T3T3h31=6α2σ0.\displaystyle h_{1}^{-1}T_{1}^{\dagger}T_{1}h_{1}^{-1}+h_{2}^{-1}T_{2}^{\dagger}T_{2}h_{2}^{-1}+h_{3}^{-1}T_{3}^{\dagger}T_{3}h_{3}^{-1}=6\alpha^{2}\sigma_{0}. (S12)

Here, σ0\sigma_{0} denotes the 2×22\times 2 identity matrix. The parameter α=w/vkθ\alpha=w/vk_{\theta}, where vv is the single layer graphene Fermi velocity. Direct application of the Hamiltonian on the state ΨK1K2\Psi^{K_{1}K_{2}} in Eq. (S8) shows that the state is a zero eigenstate at 𝒌=0\bm{k}=0 and has a linear dispersion with respect to 𝒌\bm{k}. More precisely, the projected low-energy effective Hamiltonian is

𝚿K1K2|𝒌K1K2|𝚿K1K2=v1+6α2[𝝈𝒌+j=13Tjhj1𝝈𝒌hj1Tj]=v𝝈𝒌,\displaystyle\bra{\bm{\Psi}^{K_{1}K_{2}}}\mathcal{H}_{\bm{k}}^{K_{1}K_{2}}\ket{\bm{\Psi}^{K_{1}K_{2}}}=\frac{v}{1+6\alpha^{2}}\left[\bm{\sigma}^{*}\cdot\bm{k}+\sum_{j=1}^{3}T_{j}h_{j}^{-1\dagger}\bm{\sigma}^{*}\cdot\bm{k}h_{j}^{-1}T_{j}^{\dagger}\right]=v^{\star}\bm{\sigma}^{*}\cdot\bm{k}, (S13)

where

vv=13α21+6α2\displaystyle\frac{v^{\star}}{v}=\frac{1-3\alpha^{2}}{1+6\alpha^{2}} (S14)

shows the twist-angle dependence of the renormalized Fermi velocity vv^{\star}.

II.1 The full Hamiltonian

So far, we have modeled the interlayer hopping from cone K1K_{1} to cone K2K_{2}. In total, there are four inequivalent Dirac cones, two in each layer. In this section, we argue by symmetry to find the Hamiltonians for the three other interlayer hopping processes. We omit the spin quantum number in this section because the interlayer hopping and the Dirac Hamiltonian are spin independent.

In general, the total Hamiltonian

H𝒌=𝒌K1K2𝒌K1K2𝒌K2K1𝒌K2K1,\displaystyle H_{\bm{k}}=\mathcal{H}_{\bm{k}}^{K_{1}K_{2}}\oplus\mathcal{H}_{\bm{k}}^{K^{\prime}_{1}K^{\prime}_{2}}\oplus\mathcal{H}_{\bm{k}}^{K_{2}K_{1}}\oplus\mathcal{H}_{\bm{k}}^{K^{\prime}_{2}K^{\prime}_{1}}, (S15)

must be invariant under the symmetries of the twisted bilayer. We relate the terms of the Hamiltonian by considering symmetry transformations that relate the cones. Ref. [27] outlines a similar procedure.

II.2 Time-reversal symmetry

We start by considering time-reversal symmetry 𝒯\mathcal{T}. It relates primed and unprimed Dirac cones, such that KKK\leftrightarrow K^{\prime}. Furthermore, it acts as complex conjugation on the sublattice Pauli matrices. The hopping matrices transform in the following way,

T2()T3(),\displaystyle T_{2}^{(\dagger)}\leftrightarrow T_{3}^{(\dagger)}, (S16)

whereas T1()T_{1}^{(\dagger)} is left invariant. The crystal momentum 𝒌𝒌\bm{k}\rightarrow-\bm{k}. In total, we have that

𝒌K1K2=[hK(𝒌)T1T3T2T1hK(𝒌+𝒒1)00T30hK(𝒌+𝒒2)0T200hK(𝒌+𝒒3)].\displaystyle\mathcal{H}_{\bm{k}}^{K^{\prime}_{1}K^{\prime}_{2}}=\begin{bmatrix}h^{K^{\prime}}(\bm{k})&T_{1}&T_{3}&T_{2}\\ T_{1}^{\dagger}&h^{K^{\prime}}(\bm{k}+\bm{q}_{1})&0&0\\ T_{3}^{\dagger}&0&h^{K^{\prime}}(\bm{k}+\bm{q}_{2})&0\\ T_{2}^{\dagger}&0&0&h^{K^{\prime}}(\bm{k}+\bm{q}_{3})\end{bmatrix}. (S17)

II.3 C2xC_{2x} rotation symmetry

We now use the twofold C2xC_{2x} symmetry to find HK2K1H^{K_{2}K_{1}} and HK2K1H^{K_{2}^{\prime}K^{\prime}_{1}}. This symmetry exchanges the sublattices and flips the sign of the kyk_{y} component. Hence,

(σx,σy)(σx,σy),\displaystyle\left(\sigma_{x},\sigma_{y}\right)\rightarrow\left(\sigma_{x},-\sigma_{y}\right), (kx,ky)(kx,ky).\displaystyle\left(k_{x},k_{y}\right)\rightarrow\left(k_{x},-k_{y}\right). (S18)

Additionally, C2xC_{2x} exchanges the top and bottom layers. In total,

𝒌K2K1=[hK(𝒌)T1T3T2T1hK(𝒌+𝒒1)00T30hK(𝒌+𝒒3)0T200hK(𝒌+𝒒2)].\displaystyle\mathcal{H}_{\bm{k}}^{K_{2}K_{1}}=\begin{bmatrix}h^{K}(\bm{k})&T_{1}&T_{3}&T_{2}\\ T_{1}^{\dagger}&h^{K}(\bm{k}+\bm{q}_{1})&0&0\\ T_{3}^{\dagger}&0&h^{K}(\bm{k}+\bm{q}_{3})&0\\ T_{2}^{\dagger}&0&0&h^{K}(\bm{k}+\bm{q}_{2})\end{bmatrix}. (S19)

II.4 C2x𝒯C_{2x}\mathcal{T} symmetry

The combination of the symmetries yields

𝒌K2K1=[hK(𝒌)T1T2T3T1hK(𝒌𝒒1)00T20hK(𝒌𝒒3)0T300hK(𝒌𝒒2)].\displaystyle\mathcal{H}_{\bm{k}}^{K_{2}^{\prime}K^{\prime}_{1}}=\begin{bmatrix}h^{K^{\prime}}(\bm{k})&T_{1}&T_{2}&T_{3}\\ T_{1}^{\dagger}&h^{K^{\prime}}(\bm{k}-\bm{q}_{1})&0&0\\ T_{2}^{\dagger}&0&h^{K^{\prime}}(\bm{k}-\bm{q}_{3})&0\\ T_{3}^{\dagger}&0&0&h^{K^{\prime}}(\bm{k}-\bm{q}_{2})\end{bmatrix}. (S20)

II.5 Summary of full Hamiltonian

To summarize this section, we list the four distinct Hamiltonians

𝒌K1K2=[hK(𝒌)T1T2T3T1hK(𝒌𝒒1)00T20hK(𝒌𝒒2)0T300hK(𝒌𝒒3)],\displaystyle\mathcal{H}_{\bm{k}}^{K_{1}K_{2}}=\begin{bmatrix}h^{K}(\bm{k})&T_{1}&T_{2}&T_{3}\\ T_{1}^{\dagger}&h^{K}(\bm{k}-\bm{q}_{1})&0&0\\ T_{2}^{\dagger}&0&h^{K}(\bm{k}-\bm{q}_{2})&0\\ T_{3}^{\dagger}&0&0&h^{K}(\bm{k}-\bm{q}_{3})\end{bmatrix}, (S21a)
𝒌K1K2=[hK(𝒌)T1T3T2T1hK(𝒌+𝒒1)00T30hK(𝒌+𝒒2)0T200hK(𝒌+𝒒3)],\displaystyle\mathcal{H}_{\bm{k}}^{K_{1}^{\prime}K_{2}^{\prime}}=\begin{bmatrix}h^{K^{\prime}}(\bm{k})&T_{1}&T_{3}&T_{2}\\ T_{1}^{\dagger}&h^{K^{\prime}}(\bm{k}+\bm{q}_{1})&0&0\\ T_{3}^{\dagger}&0&h^{K^{\prime}}(\bm{k}+\bm{q}_{2})&0\\ T_{2}^{\dagger}&0&0&h^{K^{\prime}}(\bm{k}+\bm{q}_{3})\end{bmatrix}, (S21b)
𝒌K2K1=[hK(𝒌)T1T3T2T1hK(𝒌+𝒒1)00T30hK(𝒌+𝒒3)0T200hK(𝒌+𝒒2)],\displaystyle\mathcal{H}_{\bm{k}}^{K_{2}K_{1}}=\begin{bmatrix}h^{K}(\bm{k})&T_{1}&T_{3}&T_{2}\\ T_{1}^{\dagger}&h^{K}(\bm{k}+\bm{q}_{1})&0&0\\ T_{3}^{\dagger}&0&h^{K}(\bm{k}+\bm{q}_{3})&0\\ T_{2}^{\dagger}&0&0&h^{K}(\bm{k}+\bm{q}_{2})\end{bmatrix}, (S21c)
𝒌K2K1=[hK(𝒌)T1T2T3T1hK(𝒌𝒒1)00T20hK(𝒌𝒒3)0T300hK(𝒌𝒒2)].\displaystyle\mathcal{H}_{\bm{k}}^{K_{2}^{\prime}K_{1}^{\prime}}=\begin{bmatrix}h^{K^{\prime}}(\bm{k})&T_{1}&T_{2}&T_{3}\\ T_{1}^{\dagger}&h^{K^{\prime}}(\bm{k}-\bm{q}_{1})&0&0\\ T_{2}^{\dagger}&0&h^{K^{\prime}}(\bm{k}-\bm{q}_{3})&0\\ T_{3}^{\dagger}&0&0&h^{K^{\prime}}(\bm{k}-\bm{q}_{2})\end{bmatrix}. (S21d)

We can now find the low-energy eigenstates that diagonalize each of the Hamiltonians. For a compact notation, we use

±hj=hK(qj),\displaystyle\pm h_{j}=h^{K}(\mp{q}_{j}), ±hj=hK(qj).\displaystyle\pm h^{\prime}_{j}=h^{K^{\prime}}(\mp{q}_{j}). (S22)

Inspired by Eq. (S10), we read off the normalized eigenstates directly as

𝚿K1K2\displaystyle\bm{\Psi}^{K_{1}K_{2}} =11+6α2(1,h11T1,h21T2,h31T3)T,\displaystyle=\frac{1}{\sqrt{1+6\alpha^{2}}}\left(1,-h_{1}^{-1}T_{1}^{\dagger},-h_{2}^{-1}T_{2}^{\dagger},-h_{3}^{-1}T_{3}^{\dagger}\right)^{T}, (S23a)
𝚿K1K2\displaystyle\bm{\Psi}^{K_{1}^{\prime}K_{2}^{\prime}} =11+6α2(1,h11T1,h21T3,h31T2)T,\displaystyle=\frac{1}{\sqrt{1+6\alpha^{2}}}\left(1,{h^{\prime}_{1}}^{-1}T_{1}^{\dagger},{h^{\prime}_{2}}^{-1}T_{3}^{\dagger},{h^{\prime}_{3}}^{-1}T_{2}^{\dagger}\right)^{T}, (S23b)
𝚿K2K1\displaystyle\bm{\Psi}^{K_{2}K_{1}} =11+6α2(1,h11T1,h31T3,h21T2)T,\displaystyle=\frac{1}{\sqrt{1+6\alpha^{2}}}\left(1,h_{1}^{-1}T_{1}^{\dagger},h_{3}^{-1}T_{3}^{\dagger},h_{2}^{-1}T_{2}^{\dagger}\right)^{T}, (S23c)
𝚿K2K1\displaystyle\bm{\Psi}^{K_{2}^{\prime}K^{\prime}_{1}} =11+6α2(1,h11T1,h31T2,h21T3)T.\displaystyle=\frac{1}{\sqrt{1+6\alpha^{2}}}\left(1,-{h^{\prime}_{1}}^{-1}T_{1}^{\dagger},-{h^{\prime}_{3}}^{-1}T_{2}^{\dagger},-{h^{\prime}_{2}}^{-1}T_{3}^{\dagger}\right)^{T}. (S23d)

Each eigenstate is taken with respect to a specific basis, i.e., the basis of the Hamiltonian, to which the eigenstate belong. For completeness, we present the bases as

K1K2\displaystyle\mathcal{B}^{K_{1}K_{2}} =(ψK1(𝒌),ψK2(𝒌𝒒1),ψK2(𝒌𝒒2),ψK2(𝒌𝒒3))T,\displaystyle=\left(\psi_{K_{1}}(\bm{k}),\psi_{K_{2}}(\bm{k}-\bm{q}_{1}),\psi_{K_{2}}(\bm{k}-\bm{q}_{2}),\psi_{K_{2}}(\bm{k}-\bm{q}_{3})\right)^{T}, (S24a)
K1K2\displaystyle\mathcal{B}^{K_{1}^{\prime}K_{2}^{\prime}} =(ψK1(𝒌),ψK2(𝒌+𝒒1),ψK2(𝒌+𝒒2),ψK2(𝒌+𝒒3))T,\displaystyle=\left(\psi_{K^{\prime}_{1}}(\bm{k}),\psi_{K^{\prime}_{2}}(\bm{k}+\bm{q}_{1}),\psi_{K^{\prime}_{2}}(\bm{k}+\bm{q}_{2}),\psi_{K^{\prime}_{2}}(\bm{k}+\bm{q}_{3})\right)^{T}, (S24b)
K2K1\displaystyle\mathcal{B}^{K_{2}K_{1}} =(ψK2(𝒌),ψK1(𝒌+𝒒1),ψK1(𝒌+𝒒3),ψK1(𝒌+𝒒2))T,\displaystyle=\left(\psi_{K_{2}}(\bm{k}),\psi_{K_{1}}(\bm{k}+\bm{q}_{1}),\psi_{K_{1}}(\bm{k}+\bm{q}_{3}),\psi_{K_{1}}(\bm{k}+\bm{q}_{2})\right)^{T}, (S24c)
K2K1\displaystyle\mathcal{B}^{K_{2}^{\prime}K^{\prime}_{1}} =(ψK2(𝒌),ψK1(𝒌𝒒1),ψK1(𝒌𝒒3),ψK1(𝒌𝒒2))T.\displaystyle=\left(\psi_{K^{\prime}_{2}}(\bm{k}),\psi_{{K^{\prime}_{1}}}(\bm{k}-\bm{q}_{1}),\psi_{K^{\prime}_{1}}(\bm{k}-\bm{q}_{3}),\psi_{K^{\prime}_{1}}(\bm{k}-\bm{q}_{2})\right)^{T}. (S24d)

II.6 The effective 2×22\times 2 electron Hamiltonian

In this section, we consider the effective Hamiltonians in Eq. (II.5) with respect to their respective low-energy eigenstates given in Eq. (S23). As seen from Eq. (S13), this yields an effective 2×22\times 2 model. To simplify notation, we write only the first of the two superscripts used in Eqs. (II.5) and (S23). Nevertheless, we emphasize that the eigenstates are still superpositions of electron states at cones in distinct layers. The set of effective Hamiltonians is given by

𝚿Klη|𝒌Klη|𝚿Klη=v(ησxkxσyky).\displaystyle\bra{\bm{\Psi}^{K^{\eta}_{l}}}\mathcal{H}^{K^{\eta}_{l}}_{\bm{k}}\ket{\bm{\Psi}^{K^{\eta}_{l}}}=v^{\star}(\eta\sigma_{x}k_{x}-\sigma_{y}k_{y}). (S25)

Here, we introduced η\eta to account for the chirality of the Dirac cones. That is, η=1\eta=1 and η=1\eta=-1 corresponds to the Dirac cones KK and KK^{\prime}, respectively. The subscript ll denotes the layer of the cone. The eigenvalues of the effective Hamiltonians are

ε±=±v|𝒌|,\displaystyle\varepsilon_{\pm}=\pm v^{\star}\absolutevalue{\bm{k}}, (S26)

and the corresponding eigenvectors are

ϕ𝒌η=12(ηeηφ𝒌,1)T,\displaystyle\phi_{\bm{k}-}^{\eta}=\frac{1}{\sqrt{2}}(-\eta e^{\eta\varphi_{\bm{k}}},1)^{T}, ϕ𝒌+η=12(ηeηφ𝒌,1)T.\displaystyle\phi_{\bm{k}+}^{\eta}=\frac{1}{\sqrt{2}}(\eta e^{\eta\varphi_{\bm{k}}},1)^{T}. (S27)

Here, φ𝒌\varphi_{\bm{k}} is the angle between 𝒌\bm{k} and the xx-axis. The eigenvectors span a unitary transformation U𝒌ηU_{\bm{k}\eta} such that

vU𝒌η(ησxkxσyky)U𝒌η=(ε00ε+),\displaystyle v^{\star}U_{\bm{k}\eta}^{\dagger}(\eta\sigma_{x}k_{x}-\sigma_{y}k_{y})U_{\bm{k}\eta}=\begin{pmatrix}\varepsilon_{-}&0\\ 0&\varepsilon_{+}\end{pmatrix}, (S28)

where

U𝒌η=12(ηeiηφ𝒌ηeiηφ𝒌11)\displaystyle U_{\bm{k}\eta}=\frac{1}{\sqrt{2}}\begin{pmatrix}-\eta e^{i\eta\varphi_{\bm{k}}}&&\eta e^{i\eta\varphi_{\bm{k}}}\\ 1&&1\end{pmatrix} U𝒌η=12(ηeiηφ𝒌1ηeiηφ𝒌1)\displaystyle U^{\dagger}_{\bm{k}\eta}=\frac{1}{\sqrt{2}}\begin{pmatrix}-\eta e^{-i\eta\varphi_{\bm{k}}}&&1\\ \eta e^{-i\eta\varphi_{\bm{k}}}&&1\end{pmatrix} (S29)

Accordingly, the eigenstates in Eq. (S23) are related to band electron operators by the transformations

c𝒌I=U𝒌η𝚿I\displaystyle c_{\bm{k}I}=U^{\dagger}_{\bm{k}\eta}\bm{\Psi}^{I} c𝒌I=𝚿IU𝒌η\displaystyle c^{\dagger}_{\bm{k}I}=\bm{\Psi}^{I\dagger}U_{\bm{k}\eta} (S30a)
𝚿I=U𝒌ηc𝒌I\displaystyle\bm{\Psi}^{I}=U_{\bm{k}\eta}c_{\bm{k}I} 𝚿I=c𝒌IU𝒌η\displaystyle\bm{\Psi}^{I\dagger}=c^{\dagger}_{\bm{k}I}U^{\dagger}_{\bm{k}\eta} (S30b)

where

c𝒌I=(c𝒌Ic+𝒌I),\displaystyle c_{\bm{k}I}=\begin{pmatrix}c_{-\bm{k}I}\\ c_{+\bm{k}I}\end{pmatrix}, c𝒌I=(c𝒌Ic+𝒌I).\displaystyle c^{\dagger}_{\bm{k}I}=\begin{pmatrix}c^{\dagger}_{-\bm{k}I}&c^{\dagger}_{+\bm{k}I}\end{pmatrix}. (S31)

Here, we introduced the index I{K1,K1,K2,K2}I\in\{K_{1},K_{1}^{\prime},K_{2},K_{2}^{\prime}\}. It denotes at which Dirac cone each state is ”based.” Note that η\eta can be extracted directly from II. The effective electron Hamiltonian for each of the cones is

H~nI,s(𝒌)=εn𝒌cn𝒌Iscn𝒌Is.\displaystyle\tilde{H}_{n}^{I,s}(\bm{k})=\varepsilon_{n\bm{k}}c^{\dagger}_{n\bm{k}Is}c_{n\bm{k}Is}. (S32)

where nn denotes the energy levels n=±n=\pm and we introduced the electron spin ss.

III Interfacial electron-magnon coupling

In this section, we consider the s-d coupling between the ferromagnetic layers and the electrons in the graphene layers. We consider a local coupling

Hint=Jsdd2𝒓𝒎(𝒓)𝒔(𝒓),\displaystyle H_{\mathrm{int}}=J_{\mathrm{s-d}}\int\mathrm{d}^{2}\bm{r}\bm{m}(\bm{r})\cdot\bm{s}(\bm{r}), (S33)

where 𝒎\bm{m} is the magnetization in the ferromagnet and 𝒔\bm{s} is the spin of the itinerant electrons. We rewrite the spin operators in terms of raising and lowering operators

sx=12(s++s)\displaystyle s_{x}=\frac{1}{2}(s^{+}+s^{-}) sx=12i(s+s).\displaystyle s_{x}=\frac{1}{2i}(s^{+}-s^{-}). (S34)

As a result, we find

Hint=Jsdd2𝒓[12(m+s+ms+)+mzsz].\displaystyle H_{\mathrm{int}}=J_{\mathrm{s-d}}\int\mathrm{d}^{2}\bm{r}\left[\frac{1}{2}\left(m^{+}s^{-}+m^{-}s^{+}\right)+m_{z}s_{z}\right]. (S35)

We use the same Holstein-Primakoff transformation as in Eqs. (S3) and (S4) with a subsequent Fourier transform to find

HintT=𝒒,𝒌JsdM2𝒱(a𝒒ψa,𝒌+𝒒,ψa,𝒌,+a𝒒ψa,𝒌𝒒,ψa,𝒌,)+𝒌JsdM(ψa,𝒌,ψa,𝒌,ψa,𝒌,ψa,𝒌,),\displaystyle H^{T}_{\mathrm{int}}=\sum_{\bm{q},\bm{k}}J_{\mathrm{s-d}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\left(a_{\bm{q}}\psi^{\dagger}_{a,\bm{k}+\bm{q},\uparrow}\psi_{a,\bm{k},\downarrow}+a^{\dagger}_{\bm{q}}\psi^{\dagger}_{a,\bm{k}-\bm{q},\downarrow}\psi_{a,\bm{k},\uparrow}\right)+\sum_{\bm{k}}J_{\mathrm{s-d}}M\left(\psi^{\dagger}_{a,\bm{k},\uparrow}\psi_{a,\bm{k},\uparrow}-\psi^{\dagger}_{a,\bm{k},\downarrow}\psi_{a,\bm{k},\downarrow}\right), (S36a)
and
HintB=𝒒,𝒌JsdM2𝒱(b𝒒ψb,𝒌𝒒,ψb,𝒌,+b𝒒ψb,𝒌+𝒒,ψb,𝒌,)𝒌JsdM(ψb,𝒌,ψb,𝒌,ψb,𝒌,ψb,𝒌,),\displaystyle H^{B}_{\mathrm{int}}=\sum_{\bm{q},\bm{k}}J_{\mathrm{s-d}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\left(b^{\dagger}_{\bm{q}}\psi^{\dagger}_{b,\bm{k}-\bm{q},\uparrow}\psi_{b,\bm{k},\downarrow}+b_{\bm{q}}\psi^{\dagger}_{b,\bm{k}+\bm{q},\downarrow}\psi_{b,\bm{k},\uparrow}\right)-\sum_{\bm{k}}J_{\mathrm{s-d}}M\left(\psi^{\dagger}_{b,\bm{k},\uparrow}\psi_{b,\bm{k},\uparrow}-\psi^{\dagger}_{b,\bm{k},\downarrow}\psi_{b,\bm{k},\downarrow}\right), (S36b)

for the top (T) and bottom layer (B). Here, ψa,𝒌,s\psi^{\dagger}_{a,\bm{k},s} and ψb,𝒌,s\psi^{\dagger}_{b,\bm{k},s} are operators of electron states located in the top and bottom layer, respectively, with momentum 𝒌\bm{k} and spin ss. We use the layer indices aa and bb to emphasize that the states are not associated with particular points in the Brillouin zone. The low-energy electron states, however, are located close to the Dirac cones. We now project the Hamiltonians in Eq. (S36a) and (S36b) on the electron states given in Eq. (S24). Furthermore, we restrict our discussion to intravalley scattering. Thus, the top layer Hamiltonian can be written as

HintT=JsdM2𝒱𝒒,𝒌a𝒒((ψ,𝒌+𝒒K1ψ,𝒌K1+ψ,𝒌+𝒒K1ψ,𝒌K1)+j=13(ψ,𝒌+𝒒j+𝒒K2ψ,𝒌+𝒒jK2+ψ,𝒌𝒒j+𝒒K2ψ,𝒌𝒒jK2))+h.c.+JsdMs,𝒌s((ψs,𝒌K1ψs,𝒌K1+ψs,𝒌K1ψs,𝒌K1)+j=13(ψs,𝒌+𝒒jK2ψs,𝒌+𝒒jK2+ψs,𝒌𝒒jK2ψs,𝒌𝒒jK2)).\displaystyle\begin{split}H^{T}_{\mathrm{int}}=J_{\mathrm{s-d}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\sum_{\bm{q},\bm{k}}a_{\bm{q}}\left(\left(\psi^{{K}_{1}\dagger}_{\uparrow,\bm{k}+\bm{q}}\psi^{{K}_{1}}_{\downarrow,\bm{k}}+\psi^{{K^{\prime}}_{1}\dagger}_{\uparrow,\bm{k}+\bm{q}}\psi^{{K^{\prime}}_{1}}_{\downarrow,\bm{k}}\right)+\sum^{3}_{j=1}\left(\psi^{K_{2}\dagger}_{\uparrow,\bm{k}+\bm{q}_{j}+\bm{q}}\psi^{K_{2}}_{\downarrow,\bm{k}+\bm{q}_{j}}+\psi^{{K^{\prime}}_{2}\dagger}_{\uparrow,\bm{k}-\bm{q}_{j}+\bm{q}}\psi^{{K^{\prime}}_{2}}_{\downarrow,\bm{k}-\bm{q}_{j}}\right)\right)+h.c.\\ +J_{\mathrm{s-d}}M\sum_{s,\bm{k}}s\biggl{(}\left(\psi^{{K}_{1}\dagger}_{s,\bm{k}}\psi^{{K}_{1}}_{s,\bm{k}}+\psi^{{K^{\prime}}_{1}\dagger}_{s,\bm{k}}\psi^{{K^{\prime}}_{1}}_{s,\bm{k}}\right)+\sum^{3}_{j=1}\left(\psi^{K_{2}\dagger}_{s,\bm{k}+\bm{q}_{j}}\psi^{K_{2}}_{s,\bm{k}+\bm{q}_{j}}+\psi^{{K^{\prime}}_{2}\dagger}_{s,\bm{k}-\bm{q}_{j}}\psi^{{K^{\prime}}_{2}}_{s,\bm{k}-\bm{q}_{j}}\right)\biggr{)}.\end{split} (S37)

Here, the electron spin index s=±1s=\pm 1 is positive for spin up \uparrow and negative for spin down \downarrow. We now transform the coupling and project the Hamiltonian onto the low-energy states given in Eq. (S23). In this case, the coupling simplifies to

HintT=Jsd1+6α2M2𝒱𝒒,𝒌a𝒒(Ψ,𝒌+𝒒K1Ψ,𝒌K1+Ψ,𝒌+𝒒K1Ψ,𝒌K1+6α2(Ψ,𝒌+𝒒K2Ψ,𝒌K2+Ψ,𝒌+𝒒K2Ψ,𝒌K2))+h.c+JsdM1+6α2s,𝒌s(Ψs,𝒌K1Ψs,𝒌K1+Ψs,𝒌K1Ψs,𝒌K1+6α2(Ψs,𝒌K2Ψs,𝒌K2+Ψs,𝒌K2Ψs,𝒌K2)).\displaystyle\begin{split}H^{T}_{\mathrm{int}}=\frac{J_{\mathrm{s-d}}}{1+6\alpha^{2}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\sum_{\bm{q},\bm{k}}a_{\bm{q}}\biggl{(}\Psi^{K_{1}\dagger}_{\uparrow,\bm{k}+\bm{q}}\Psi^{K_{1}}_{\downarrow,\bm{k}}+\Psi^{K^{\prime}_{1}\dagger}_{\uparrow,\bm{k}+\bm{q}}\Psi^{K^{\prime}_{1}}_{\downarrow,\bm{k}}+6\alpha^{2}\left(\Psi^{K_{2}\dagger}_{\uparrow,\bm{k}+\bm{q}}\Psi^{K_{2}}_{\downarrow,\bm{k}}+\Psi^{K^{\prime}_{2}\dagger}_{\uparrow,\bm{k}+\bm{q}}\Psi^{K^{\prime}_{2}}_{\downarrow,\bm{k}}\biggr{)}\right)+h.c\\ +\frac{J_{\mathrm{s-d}}M}{1+6\alpha^{2}}\sum_{s,\bm{k}}s\biggl{(}\Psi^{K_{1}\dagger}_{s,\bm{k}}\Psi^{K_{1}}_{s,\bm{k}}+\Psi^{K_{1}^{\prime}\dagger}_{s,\bm{k}}\Psi^{K_{1}^{\prime}}_{s,\bm{k}}+6\alpha^{2}\left(\Psi^{K_{2}\dagger}_{s,\bm{k}}\Psi^{K_{2}}_{s,\bm{k}}+\Psi^{K^{\prime}_{2}\dagger}_{s,\bm{k}}\Psi^{K^{\prime}_{2}}_{s,\bm{k}}\right)\biggr{)}.\end{split} (S38)

The bottom ferromagnet couples to the bottom graphene layer in a similar way. The explicit form is

HintB=Jsd1+6α2M2𝒱𝒒𝒌b𝒒(Ψ𝒌𝒒K2Ψ𝒌K2+Ψ𝒌𝒒K2Ψ𝒌K2+6α2(Ψ𝒌𝒒K1Ψ𝒌K1+Ψ𝒌𝒒K1Ψ𝒌K1))+h.c+JsdM1+6α2s,𝒌s((Ψs,𝒌K2Ψs,𝒌K2+Ψs,𝒌K2Ψs,𝒌K2)+6α2(Ψs,𝒌K1Ψs,𝒌K1+Ψs,𝒌K1Ψs,𝒌K1))\displaystyle\begin{split}H^{B}_{\mathrm{int}}=\frac{J_{\mathrm{s-d}}}{1+6\alpha^{2}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\sum_{\bm{q}\bm{k}}b^{\dagger}_{\bm{q}}\biggl{(}\Psi^{K_{2}\dagger}_{\uparrow\bm{k}-\bm{q}}\Psi^{K_{2}}_{\downarrow\bm{k}}+\Psi^{K^{\prime}_{2}\dagger}_{\uparrow\bm{k}-\bm{q}}\Psi^{K^{\prime}_{2}}_{\downarrow\bm{k}}+6\alpha^{2}\left(\Psi^{K_{1}^{\prime}\dagger}_{\uparrow\bm{k}-\bm{q}}\Psi^{K^{\prime}_{1}}_{\downarrow\bm{k}}+\Psi^{K_{1}\dagger}_{\uparrow\bm{k}-\bm{q}}\Psi^{K_{1}}_{\downarrow\bm{k}}\right)\biggr{)}+h.c\\ +\frac{J_{\mathrm{s-d}}M}{1+6\alpha^{2}}\sum_{s,\bm{k}}-s\biggl{(}\left(\Psi^{K_{2}\dagger}_{s,\bm{k}}\Psi^{K_{2}}_{s,\bm{k}}+\Psi^{K^{\prime}_{2}\dagger}_{s,\bm{k}}\Psi^{K^{\prime}_{2}}_{s,\bm{k}}\right)+6\alpha^{2}\left(\Psi^{K^{\prime}_{1}\dagger}_{s,\bm{k}}\Psi^{K^{\prime}_{1}}_{s,\bm{k}\uparrow}+\Psi^{K_{1}\dagger}_{s,\bm{k}}\Psi^{K_{1}}_{s,\bm{k}}\right)\biggr{)}\end{split} (S39)

The total s-d coupling is the sum of Eqs. (S38) and (S39). Note how the exchange splitting does not cancel out except at the special angle α2=1/6\alpha^{2}=1/6. However, this is not the magic angle α2=1/3\alpha^{2}=1/3. The coupling is expressed in terms of the low-energy eigenstates of the BM model. We transform the coupling using Eq. (S30), to express the coupling in terms of band electrons. To simplify the model, we consider coupling to the lower band only. In terms of lower band electron operators, the interfacial coupling is

HsdT=Jsd1+6α2M2𝒱𝒒,𝒌a𝒒(c𝒌+𝒒K2ϕ𝒌+𝒒Kϕ𝒌Kc𝒌K2+c𝒌+𝒒K1ϕ𝒌+𝒒Kϕ𝒌Kc𝒌K1+6α2(c𝒌+𝒒K2ϕ𝒌+𝒒Kϕ𝒌Kc𝒌K2c𝒌+𝒒K1ϕ𝒌+𝒒Kϕ𝒌Kc𝒌K1))+h.c.+JsdM1+6α2s,𝒌s((c𝒌K1sc𝒌K1+c𝒌K1sc𝒌K1s)+6α2(c𝒌K2sc𝒌K2s+c𝒌K2sc𝒌K2s)).\displaystyle\begin{split}H^{T}_{\mathrm{s-d}}=\frac{J_{\mathrm{s-d}}}{1+6\alpha^{2}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\sum_{\bm{q},\bm{k}}a_{\bm{q}}\biggl{(}c_{\bm{k}+\bm{q}K_{2}^{\prime}\uparrow}^{\dagger}\phi^{K\dagger}_{\bm{k}+\bm{q}}\phi^{K}_{\bm{k}}c_{\bm{k}K_{2}^{\prime}\downarrow}+c_{\bm{k}+\bm{q}K_{1}^{\prime}\uparrow}^{\dagger}\phi^{K^{\prime}\dagger}_{\bm{k}+\bm{q}}\phi^{K^{\prime}}_{\bm{k}}c_{\bm{k}K_{1}^{\prime}\downarrow}\\ +6\alpha^{2}\left(c_{\bm{k}+\bm{q}K_{2}\uparrow}^{\dagger}\phi^{K\dagger}_{\bm{k}+\bm{q}}\phi^{K}_{\bm{k}}c_{\bm{k}K_{2}\downarrow}c_{\bm{k}+\bm{q}K_{1}^{\prime}\uparrow}^{\dagger}\phi^{K^{\prime}\dagger}_{\bm{k}+\bm{q}}\phi^{K^{\prime}}_{\bm{k}}c_{\bm{k}K_{1}^{\prime}\downarrow}\right)\biggr{)}+h.c.\\ +\frac{J_{\mathrm{s-d}}M}{1+6\alpha^{2}}\sum_{s,\bm{k}}s\biggl{(}\left(c_{\bm{k}K_{1}s}^{\dagger}c_{\bm{k}K_{1}\uparrow}+c_{\bm{k}K_{1}^{\prime}s}^{\dagger}c_{\bm{k}K_{1}^{\prime}s}\right)+6\alpha^{2}\left(c_{\bm{k}K_{2}s}^{\dagger}c_{\bm{k}K_{2}s}+c_{\bm{k}K_{2}^{\prime}s}^{\dagger}c_{\bm{k}K^{\prime}_{2}s}\right)\biggr{)}.\end{split} (S40)

The ϕ\phi-factors can be written as

ϕ𝒌+𝒒ηϕ𝒌η=12(eiη(φ𝒌φ𝒌+𝒒)+1),\displaystyle\phi_{\bm{k}+\bm{q}}^{\eta\dagger}\phi^{\eta}_{\bm{k}}=\frac{1}{2}\left(e^{i\eta(\varphi_{\bm{k}}-\varphi_{\bm{k}+\bm{q}})}+1\right), ϕ𝒌𝒒ηϕ𝒌η=12(eiη(φ𝒌φ𝒌𝒒)+1).\displaystyle\phi_{\bm{k}-\bm{q}}^{\eta\dagger}\phi^{\eta}_{\bm{k}}=\frac{1}{2}\left(e^{i\eta(\varphi_{\bm{k}}-\varphi_{\bm{k}-\bm{q}})}+1\right). (S41)

The analogous coupling for the bottom layer coupling is

HsdB=Jsd1+6α2M2𝒱𝒒𝒌b𝒒(c𝒌𝒒K2ϕ𝒌𝒒Kϕ𝒌Kc𝒌K2+c𝒌𝒒K2ϕ𝒌𝒒Kϕ𝒌Kc𝒌K2+6α2(c𝒌𝒒K1ϕ𝒌𝒒Kϕ𝒌Kc𝒌K1+c𝒌𝒒K1ϕ𝒌𝒒Kϕ𝒌Kc𝒌K1))+h.c.+JsdM1+6α2s,𝒌(s)((c𝒌K2sc𝒌K2s+c𝒌K2sc𝒌K2s)+6α2(c𝒌K1sc𝒌K1s+c𝒌K1sc𝒌K1s)).\displaystyle\begin{split}H^{B}_{\mathrm{s-d}}=\frac{J_{\mathrm{s-d}}}{1+6\alpha^{2}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}\sum_{\bm{q}\bm{k}}b^{\dagger}_{\bm{q}}\biggl{(}c_{\bm{k}-\bm{q}K_{2}\uparrow}^{\dagger}\phi^{K\dagger}_{\bm{k}-\bm{q}}\phi^{K}_{\bm{k}}c_{\bm{k}K_{2}\downarrow}+c_{\bm{k}-\bm{q}K^{\prime}_{2}\uparrow}^{\dagger}\phi^{K^{\prime}\dagger}_{\bm{k}-\bm{q}}\phi^{K^{\prime}}_{\bm{k}}c_{\bm{k}K^{\prime}_{2}\downarrow}\\ +6\alpha^{2}\left(c_{\bm{k}-\bm{q}K_{1}\uparrow}^{\dagger}\phi^{K\dagger}_{\bm{k}-\bm{q}}\phi^{K}_{\bm{k}}c_{\bm{k}K_{1}\downarrow}+c_{\bm{k}-\bm{q}K_{1}^{\prime}\uparrow}^{\dagger}\phi^{K^{\prime}\dagger}_{\bm{k}-\bm{q}}\phi^{K^{\prime}}_{\bm{k}}c_{\bm{k}K_{1}^{\prime}\downarrow}\right)\biggr{)}+h.c.\\ +\frac{J_{\mathrm{s-d}}M}{1+6\alpha^{2}}\sum_{s,\bm{k}}(-s)\biggl{(}\left(c_{\bm{k}K_{2}s}^{\dagger}c_{\bm{k}K_{2}s}+c_{\bm{k}K_{2}^{\prime}s}^{\dagger}c_{\bm{k}K_{2}^{\prime}s}\right)+6\alpha^{2}\left(c_{\bm{k}K^{\prime}_{1}s}^{\dagger}c_{\bm{k}K^{\prime}_{1}s}+c_{\bm{k}K_{1}s}^{\dagger}c_{\bm{k}K_{1}s}\right)\biggr{)}.\end{split} (S42)

From Eq. (S40) and (S42), we can extract the exchange splitting Δss\Delta_{ss} and electron-magnon coupling constant gg introduced in the main text.

IV Effective interaction

The electron-magnon coupling can give rise to an effective electron-electron interaction. We consider the system Hamiltonian

H=He+Hm+Hem.\displaystyle H=H_{\mathrm{e}}+H_{\mathrm{m}}+H_{\mathrm{em}}. (S43)

The electron Hamiltonian is

He=\displaystyle H_{\mathrm{e}}= 𝒌,I,sε𝒌c𝒌Isc𝒌Is+Δss𝒌,I,sls(c𝒌Isc𝒌Is),\displaystyle\sum_{\bm{k},I,s}\varepsilon_{\bm{k}}c_{\bm{k}Is}^{\dagger}c_{\bm{k}Is}+\Delta_{ss}\sum_{\bm{k},I,s}ls\left(c_{\bm{k}Is}^{\dagger}c_{\bm{k}Is}\right), (S44a)
where we introduced the layer index ll. It can be extracted directly from the index II, and takes the value l=1l=1 for I{K1,K1}I\in\{K_{1},K^{\prime}_{1}\} and l=1l=-1 for I{K2,K2}I\in\{K_{2},K^{\prime}_{2}\}. Hence, top-layer spin-up electrons experience the same exchange spin shift as a bottom-layer spin-down electron due to the identical but anti-parallel ferromagnets. The magnitude of the spin splitting is
Δss=JsdM(16α2)1+6α2.\displaystyle\Delta_{ss}=\frac{J_{\mathrm{s-d}}M(1-6\alpha^{2})}{1+6\alpha^{2}}. (S44b)
The electron dispersion ε𝒌\varepsilon_{\bm{k}} is that of the lower band. The magnon Hamiltonian is
Hm=\displaystyle H_{\mathrm{m}}= 𝒒ω𝒒(a𝒒a𝒒+b𝒒b𝒒),\displaystyle\sum_{\bm{q}}\omega_{\bm{q}}\left(a_{\bm{q}}^{\dagger}a_{\bm{q}}+b_{\bm{q}}^{\dagger}b_{\bm{q}}\right), (S44c)
and the electron-magnon coupling is
Hem=𝒒𝒌IgI,𝒌,𝒒a(c𝒌+𝒒Ic𝒌I)a𝒒+(gI,𝒌,𝒒a)(c𝒌Ic𝒌+𝒒I)a𝒒+gI,𝒌,𝒒b(c𝒌𝒒Ic𝒌I)b𝒒+(gI,𝒌,𝒒b)(c𝒌Ic𝒌𝒒I)b𝒒.\displaystyle\begin{split}H_{\mathrm{em}}=&\sum_{\bm{q}\bm{k}I}g_{I,\bm{k},\bm{q}}^{a}\left(c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)a_{\bm{q}}+(g_{I,\bm{k},\bm{q}}^{a})^{*}\left(c_{\bm{k}I\downarrow}^{\dagger}c_{\bm{k}+\bm{q}I\uparrow}\right)a_{\bm{q}}^{\dagger}\\ &+g_{I,\bm{k},-\bm{q}}^{b}\left(c_{\bm{k}-\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)b^{\dagger}_{\bm{q}}+(g_{I,\bm{k},-\bm{q}}^{b})^{*}\left(c_{\bm{k}I\downarrow}^{\dagger}c_{\bm{k}-\bm{q}I\uparrow}\right)b_{\bm{q}}.\end{split} (S44d)

The coupling coefficients depend on the layer at which the electron state is based and the chirality of the Dirac cone II. To account for this, we consider two disjoint subsets of II. That is, we denote {K1,K1}\{K_{1},K^{\prime}_{1}\} by I1I_{1} and {K2,K2}\{K_{2},K^{\prime}_{2}\} by I2I_{2}. The coupling constants are then

gI1,𝒌,𝒒a=\displaystyle g_{I_{1},\bm{k},\bm{q}}^{a}= gI2,𝒌,𝒒b=g0ϕ𝒌+𝒒ηϕ𝒌η,\displaystyle g_{I_{2},\bm{k},\bm{q}}^{b}=g_{0}\phi^{\eta\dagger}_{\bm{k}+\bm{q}}\phi^{\eta}_{\bm{k}}, (S45a)
gI2,𝒌,𝒒a=\displaystyle g_{I_{2},\bm{k},\bm{q}}^{a}= gI1,𝒌,𝒒b=6α2g0ϕ𝒌+𝒒ηϕ𝒌η,\displaystyle g_{I_{1},\bm{k},\bm{q}}^{b}=6\alpha^{2}g_{0}\phi^{\eta\dagger}_{\bm{k}+\bm{q}}\phi^{\eta}_{\bm{k}}, (S45b)

where

g0=Jsd1+6α2M2𝒱.\displaystyle g_{0}=\frac{J_{\mathrm{s-d}}}{1+6\alpha^{2}}\frac{\sqrt{M}}{\sqrt{2\mathcal{V}}}. (S46)

The aa and bb superscripts denote the top and bottom layers, respectively. The ϕ\phi-factors are defined in Eq. (S27).

We relabel the indices of HemH_{\mathrm{em}} and use (g𝒌+𝒒,𝒒I)=g𝒌,𝒒I(g_{\bm{k}+\bm{q},-\bm{q}}^{I})^{*}=g_{\bm{k},\bm{q}}^{I}. The relabeling simplifies the electron-magnon coupling to

Hem=𝒒,𝒌,IgI,𝒌,𝒒a(c𝒌+𝒒Ic𝒌I)a𝒒+gI,𝒌,𝒒a(c𝒌+𝒒Ic𝒌I)a𝒒+gI,𝒌,𝒒b(c𝒌+𝒒Ic𝒌I)b𝒒+gI,𝒌,𝒒b(c𝒌+𝒒Ic𝒌I)b𝒒.\displaystyle\begin{split}H_{\mathrm{em}}=&\sum_{\bm{q},\bm{k},I}g_{I,\bm{k},\bm{q}}^{a}\left(c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)a_{\bm{q}}+g_{I,\bm{k},\bm{q}}^{a}\left(c_{\bm{k}+\bm{q}I\downarrow}^{\dagger}c_{\bm{k}I\uparrow}\right)a_{-\bm{q}}^{\dagger}\\ &+g_{I,\bm{k},\bm{q}}^{b}\left(c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)b^{\dagger}_{-\bm{q}}+g_{I,\bm{k},\bm{q}}^{b}\left(c_{\bm{k}+\bm{q}I\downarrow}^{\dagger}c_{\bm{k}I\uparrow}\right)b_{\bm{q}}.\end{split} (S47)

We split the Hamiltonian into two parts to perform the Schrieffer-Wolff transformation [35]. Let H0=He+HmH_{0}=H_{\mathrm{e}}+H_{\mathrm{m}} and H1=HemH_{1}=H_{\mathrm{em}} such that we have H=H0+ηH1H=H_{0}+\eta H_{1}. The small expansion parameter η\eta should not be confused with the valley index. We consider the canonical transformation

H=eηSHeηS\displaystyle H^{\prime}=e^{-\eta S}He^{\eta S} (S48)

and expand to find

H=H0+η(H1+[H0,S])+η2[H1,S]+η22[[H0,S],S]+𝒪(η3).\displaystyle\begin{split}H^{\prime}=H_{0}+\eta(H_{1}+\left[H_{0},S\right])+\eta^{2}\left[H_{1},S\right]+\frac{\eta^{2}}{2}\left[\left[H_{0},S\right],S\right]+\mathcal{O}(\eta^{3}).\end{split} (S49)

To eliminate the linear term, we require

H1+[H0,S]=0.\displaystyle H_{1}+\left[H_{0},S\right]=0. (S50)

To that end, we make the ansatz

S=𝒒,𝒌,IgI,𝒌,𝒒ax𝒌,𝒒,Iaa𝒒(c𝒌+𝒒Ic𝒌I)+gI,𝒌,𝒒ay𝒌,𝒒,Iaa𝒒(c𝒌+𝒒Ic𝒌I)+gI,𝒌,𝒒by𝒌,𝒒,Ibb𝒒(c𝒌+𝒒Ic𝒌I)+gI,𝒌,𝒒bx𝒌,𝒒,Ibb𝒒(c𝒌+𝒒Ic𝒌I).\displaystyle\begin{split}S=&\sum_{\bm{q},\bm{k},I}g_{I,\bm{k},\bm{q}}^{a}x^{a}_{\bm{k},\bm{q},I}a_{\bm{q}}\left(c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)+g_{I,\bm{k},\bm{q}}^{a}y^{a}_{\bm{k},\bm{q},I}a_{-\bm{q}}^{\dagger}\left(c_{\bm{k}+\bm{q}I\downarrow}^{\dagger}c_{\bm{k}I\uparrow}\right)\\ &+g_{I,\bm{k},\bm{q}}^{b}y^{b}_{\bm{k},\bm{q},I}b^{\dagger}_{-\bm{q}}\left(c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}\right)+g_{I,\bm{k},\bm{q}}^{b}x^{b}_{\bm{k},\bm{q},I}b_{\bm{q}}\left(c_{\bm{k}+\bm{q}I\downarrow}^{\dagger}c_{\bm{k}I\uparrow}\right).\end{split} (S51)

We evaluate Eq. (S50) and find

x𝒌,𝒒,Ia=x𝒌,𝒒,Ib=1ε𝒌ε𝒌+𝒒+ω𝒒y𝒌,𝒒,Ia=y𝒌,𝒒,Ib=1ε𝒌ε𝒌+𝒒ω𝒒,\displaystyle\begin{split}x^{a}_{\bm{k},\bm{q},I}=x^{b}_{\bm{k},\bm{q},I}=\frac{1}{\varepsilon_{\bm{k}}-\varepsilon_{\bm{k}+\bm{q}}+\omega_{\bm{q}}}\\ y^{a}_{\bm{k},\bm{q},I}=y^{b}_{\bm{k},\bm{q},I}=\frac{1}{\varepsilon_{\bm{k}}-\varepsilon_{\bm{k}+\bm{q}}-\omega_{\bm{q}}},\end{split} (S52)

where we used that ω𝒒=ω𝒒\omega_{-\bm{q}}=\omega_{\bm{q}}. The parameters xx and yy are independent of the index II to second order in the small coupling parameter JsdJ_{\mathrm{s-d}}. The effective interaction is

Heff=η22[H1,S].\displaystyle H_{\mathrm{eff}}=\frac{\eta^{2}}{2}\left[H_{1},S\right]. (S53)

The commutator [H1,S]\left[H_{1},S\right] has the form [Aa,Bb]\left[Aa,Bb\right] with A,BccA,B\propto c^{\dagger}c containing products of fermion operators and a,bxa+yaa,b\propto xa+ya^{\dagger} containing sums of boson operators. The general relationship

[Aa,Bb]=AB[a,b]+[A,B]ab[A,B][a,b]\displaystyle\left[Aa,Bb\right]=AB[a,b]+\left[A,B\right]ab-\left[A,B\right][a,b] (S54)

shows that there are three types of contributions. The last term represents a one-body electron operator, which can be shown to vanish [35]. The second term describes an effective coupling of an electron to two magnons. We are interested in the effective electron-electron interaction described by the first term. The magnon operators of distinct layers commute

[a𝒒+b𝒒,a𝒒+b𝒒]=[a𝒒,a𝒒]+[b𝒒,b𝒒],\displaystyle\left[a_{\bm{q}}+b^{\dagger}_{\bm{q}},a^{\dagger}_{\bm{q}}+b_{\bm{q}}\right]=\left[a_{\bm{q}},a^{\dagger}_{\bm{q}}\right]+\left[b^{\dagger}_{\bm{q}},b_{\bm{q}}\right], (S55)

such that we can treat each layer independently to get

Heff=𝒒,𝒌,𝒌,I,I(gI,𝒌,𝒒agI,𝒌,𝒒a+gI,𝒌,𝒒bgI,𝒌,𝒒b)(y𝒌,𝒒x𝒌,𝒒)c𝒌+𝒒Ic𝒌Ic𝒌𝒒Ic𝒌I.\displaystyle\begin{split}H_{\mathrm{eff}}=\sum_{\bm{q},\bm{k},\bm{k}^{\prime},I,I^{\prime}}\left(g_{I,\bm{k},\bm{q}}^{a}g_{I^{\prime},\bm{k}^{\prime},-\bm{q}}^{a}+g_{I,\bm{k},\bm{q}}^{b}g_{I^{\prime},\bm{k}^{\prime},-\bm{q}}^{b}\right)(y_{\bm{k}^{\prime},-\bm{q}}-x_{\bm{k},\bm{q}})c_{\bm{k}+\bm{q}I\uparrow}^{\dagger}c_{\bm{k}I\downarrow}c_{\bm{k}^{\prime}-\bm{q}I^{\prime}\downarrow}^{\dagger}c_{\bm{k}^{\prime}I^{\prime}\uparrow}.\end{split} (S56)

IV.1 Effective BCS type pairing

We consider a BCS-type pairing. Hence, we let 𝒌=𝒌\bm{k}^{\prime}=-\bm{k}. In this case,

y𝒌,𝒒x𝒌,𝒒=2ω𝒒(ε𝒌ε𝒌+𝒒)2ω𝒒2,\displaystyle y_{-\bm{k},-\bm{q}}-x_{\bm{k},\bm{q}}=\frac{2\omega_{\bm{q}}}{(\varepsilon_{\bm{k}}-\varepsilon_{\bm{k}+\bm{q}})^{2}-\omega_{\bm{q}}^{2}}, (S57)

where we used that ε𝒌𝒒=ε𝒌+𝒒\varepsilon_{-\bm{k}-\bm{q}}=\varepsilon_{\bm{k}+\bm{q}}. Now, we relabel 𝒌+𝒒𝒌\bm{k}+\bm{q}\rightarrow\bm{k}^{\prime} and subsequently 𝒌𝒌\bm{k}\rightarrow-\bm{k} to get

Heff=𝒌𝒌IIVeff(𝒌,𝒌,I,I)(c𝒌Ic𝒌Ic𝒌Ic𝒌I).\displaystyle H_{\mathrm{eff}}=\sum_{\bm{k}\bm{k}^{\prime}II^{\prime}}V_{\mathrm{eff}}({\bm{k},\bm{k}^{\prime}},I,I^{\prime})\left(c_{\bm{k}^{\prime}I\uparrow}^{\dagger}c_{-\bm{k}^{\prime}I^{\prime}\downarrow}^{\dagger}c_{-\bm{k}I\downarrow}c_{\bm{k}I^{\prime}\uparrow}\right). (S58)

The effective interaction is

Veff(𝒌,𝒌,I,I)=2ω𝒌+𝒌(j{a,b}gI,𝒌,𝒌+𝒌,jgI,𝒌,𝒌+𝒌j)ω𝒌+𝒌2(ε𝒌ε𝒌)2.\displaystyle\begin{split}V_{\mathrm{eff}}({\bm{k},\bm{k}^{\prime}},I,I^{\prime})=\frac{2\omega_{\bm{k^{\prime}+k}}\left(\sum_{j\in\{a,b\}}g_{I,-\bm{k},\bm{k}^{\prime}+\bm{k},}^{j}g_{I^{\prime},-\bm{k},\bm{k}^{\prime}+\bm{k}}^{j}\right)}{\omega^{2}_{\bm{k}^{\prime}+\bm{k}}-(\varepsilon_{\bm{k}^{\prime}}-\varepsilon_{\bm{k}})^{2}}.\end{split} (S59)

Eq. (S58) governs the effective interaction between electrons based at Dirac cones II and II^{\prime}. Of the multiple possible channels, only a few are suitable for Cooper pair formation. We denote the partner of II by I¯\bar{I} and consider the pairs

{I=K1,I¯=K2},\displaystyle\left\{I=K_{1},\bar{I}=K_{2}\right\}, {I=K1,I¯=K2}.\displaystyle\left\{I=K^{\prime}_{1},\bar{I}=K^{\prime}_{2}\right\}. (S60)

Note how the electron states are based in distinct layers. For this reason, they are not spin split.

For this particular choice of electron pairs, the effective interaction simplifies to

Veff(𝒌,𝒌,I,I¯)=ω𝒌+𝒌ω𝒌+𝒌2(ε𝒌ε𝒌)212α2Jsd2(1+6α2)2M𝒱w𝒌𝒌ηη,\displaystyle\begin{split}V_{\mathrm{eff}}({\bm{k},\bm{k}^{\prime}},I,\bar{I})=\frac{\omega_{\bm{k^{\prime}+k}}}{\omega^{2}_{\bm{k}^{\prime}+\bm{k}}-(\varepsilon_{\bm{k}^{\prime}}-\varepsilon_{\bm{k}})^{2}}\frac{12\alpha^{2}J^{2}_{\mathrm{s-d}}}{(1+6\alpha^{2})^{2}}\frac{M}{\mathcal{V}}w^{\eta\eta^{\prime}}_{\bm{k}\bm{k}^{\prime}},\end{split} (S61)

where

w𝒌𝒌ηη=ϕ𝒌ηϕ𝒌ηϕ𝒌ηϕ𝒌η.\displaystyle w^{\eta\eta^{\prime}}_{\bm{k}\bm{k}^{\prime}}=\phi_{\bm{k}^{\prime}}^{\eta\dagger}\phi_{-\bm{k}}^{\eta}\phi_{\bm{k}^{\prime}}^{\eta^{\prime}\dagger}\phi_{-\bm{k}}^{\eta^{\prime}}. (S62)

The angular dependence of the effective interaction is determined by the magnon-dispersion relation ω𝒌+𝒌\omega_{\bm{k}+\bm{k}^{\prime}} and w𝒌𝒌ηηw^{\eta\eta^{\prime}}_{\bm{k}\bm{k}^{\prime}}. To simplify the angular dependence, we approximate the magnon dispersion to be constant ω𝒌+𝒌=ωM\omega_{\bm{k}+\bm{k}^{\prime}}=\omega_{M} and expand w𝒌𝒌ηηw^{\eta\eta^{\prime}}_{\bm{k}\bm{k}^{\prime}}

w𝒌𝒌ηη=ϕ𝒌ηϕ𝒌ηϕ𝒌ηϕ𝒌η=eiη+η2(ϕ𝒌ϕ𝒌)(14(ei(ϕ𝒌ϕ𝒌)+ei(ϕ𝒌ϕ𝒌))12)w𝒌𝒌ηη=eiη(ϕ𝒌ϕ𝒌)(14(ei(ϕ𝒌ϕ𝒌)+ei(ϕ𝒌ϕ𝒌))12)\displaystyle\begin{split}w^{\eta\eta^{\prime}}_{\bm{k}\bm{k}^{\prime}}=&\phi_{\bm{k}^{\prime}}^{\eta\dagger}\phi_{-\bm{k}}^{\eta}\phi_{\bm{k}^{\prime}}^{\eta^{\prime}\dagger}\phi_{-\bm{k}}^{\eta^{\prime}}\\ =&e^{i\frac{\eta+\eta^{\prime}}{2}(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}\left(\frac{1}{4}\left(e^{i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}+e^{-i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}\right)-\frac{1}{2}\right)\\ w^{\eta\eta}_{\bm{k}\bm{k}^{\prime}}=&e^{i\eta(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}\left(\frac{1}{4}\left(e^{i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}+e^{-i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}\right)-\frac{1}{2}\right)\end{split} (S63)

In the last line, we used that II and I¯\bar{I} have the same chirality. Eq. (S63) suggests that the effective interaction is separable with respect to angular character.

First, let (I,I¯)=(K1,K2)(I,\bar{I})=(K_{1},K_{2}) such that η=1\eta=1. For this pair,

w𝒌𝒌ηη=\displaystyle w^{\eta\eta}_{\bm{k}\bm{k}^{\prime}}= (14(e2i(ϕ𝒌ϕ𝒌)+1)12ei(ϕ𝒌ϕ𝒌)).\displaystyle\left(\frac{1}{4}\left(e^{2i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}+1\right)-\frac{1}{2}e^{i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}\right). (S64a)
The two first terms correspond to dd- and ss-wave characters with respect to rotations in the plane. In the low-frequency limit (ε𝒌ε𝒌)20(\varepsilon_{\bm{k}^{\prime}}-\varepsilon_{\bm{k}})^{2}\rightarrow 0, these channels render a repulsive interaction. Conversely, the last term yields an attractive interaction with a pp-wave character. Second, we consider (I,I¯)=(K1,K2)(I,\bar{I})=(K^{\prime}_{1},K^{\prime}_{2}) such that η=1\eta=-1. This pair yields the coupling
w𝒌𝒌ηη=\displaystyle w^{\eta\eta}_{\bm{k}\bm{k}^{\prime}}= (14(e2i(ϕ𝒌ϕ𝒌)+1)12ei(ϕ𝒌ϕ𝒌)).\displaystyle\left(\frac{1}{4}\left(e^{-2i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}+1\right)-\frac{1}{2}e^{-i(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})}\right). (S64b)

with an attractive pp-wave interaction. The pairing channels and effective interactions are related through time-reversal symmetry.

V BCS gap equation

In this section, we consider the effective electron Hamiltonian.

H=𝒌Isε𝒌Isc𝒌Isc𝒌Is+𝒌𝒌IVeff(𝒌,𝒌,I,I¯)c𝒌Ic𝒌I¯c𝒌Ic𝒌I¯.\displaystyle H=\sum_{\bm{k}Is}\varepsilon_{\bm{k}Is}c_{\bm{k}Is}^{\dagger}c_{\bm{k}Is}+\sum_{\bm{k}\bm{k}^{\prime}I}V_{\mathrm{eff}}({\bm{k},\bm{k}^{\prime}},I,\bar{I})c_{\bm{k}I\uparrow}^{\dagger}c_{-\bm{k}\bar{I}\downarrow}^{\dagger}c_{-\bm{k}^{\prime}I\downarrow}c_{\bm{k}^{\prime}\bar{I}\uparrow}. (S65)

Here, the first term corresponds to the electron dispersion Eq. (S44a), including the exchange splitting effect. The second term corresponds to the effective interaction in Eq. (S58) for the pairs defined in Eq. (S60).

Following the conventional BCS approach, we introduce two hermitian conjugate gap parameters. We define

ΔI𝒌=𝒌V𝒌𝒌II¯c𝒌Ic𝒌I¯,\displaystyle\Delta_{I\bm{k}}=-\sum_{\bm{k^{\prime}}}V_{\bm{k}\bm{k}^{\prime}I\bar{I}}\left<c_{-\bm{k}^{\prime}I\downarrow}c_{\bm{k}^{\prime}\bar{I}\uparrow}\right>, (S66a)
ΔI𝒌=𝒌V𝒌𝒌I¯Ic𝒌I¯c𝒌I.\displaystyle\Delta_{I\bm{k}}^{\dagger}=-\sum_{\bm{k}^{\prime}}V_{\bm{k}^{\prime}\bm{k}\bar{I}I}\left<c_{\bm{k}^{\prime}\bar{I}\uparrow}^{\dagger}c_{-\bm{k}^{\prime}I\downarrow}^{\dagger}\right>. (S66b)

Here, V𝒌𝒌II¯V_{\bm{k}\bm{k}^{\prime}I\bar{I}} is shorthand notation for Veff(𝒌,𝒌,I,I¯)V_{\mathrm{eff}}({\bm{k},\bm{k}^{\prime}},I,\bar{I}). The resulting mean-field Hamiltonian is

HBCS=𝒌Isε𝒌Isc𝒌Isc𝒌Is𝒌IΔI𝒌c𝒌I¯c𝒌I𝒌IΔI𝒌c𝒌Ic𝒌I¯const.\displaystyle\begin{split}H_{\mathrm{BCS}}=\sum_{\bm{k}Is}\varepsilon_{\bm{k}Is}c_{\bm{k}Is}^{\dagger}c_{\bm{k}Is}-\sum_{\bm{k}I}\Delta_{I\bm{k}}^{\dagger}c_{-\bm{k}\bar{I}\downarrow}c_{\bm{k}I\uparrow}-\sum_{\bm{k}I}\Delta_{I\bm{k}}c_{\bm{k}I\uparrow}^{\dagger}c_{-\bm{k}\bar{I}\downarrow}^{\dagger}-\mathrm{const}.\end{split} (S67)

To diagonalize this Hamiltonian, we introduce new fermionic operators through the Bogoliubov transformation

(γ𝒌Iγ𝒌I¯)=(uI𝒌vI𝒌vI𝒌uI𝒌)(c𝒌Ic𝒌I¯).\displaystyle\begin{pmatrix}\gamma_{\bm{k}I\uparrow}\\ \gamma_{-\bm{k}\bar{I}\downarrow}^{\dagger}\end{pmatrix}=\begin{pmatrix}u_{I\bm{k}}^{*}&-v_{I\bm{k}}\\ v_{I\bm{k}}^{*}&u_{I\bm{k}}\end{pmatrix}\begin{pmatrix}c_{\bm{k}I\uparrow}\\ c_{-\bm{k}\bar{I}\downarrow}^{\dagger}\end{pmatrix}. (S68)

We require the γ\gamma-operators to satisfy fermionic anticommutation relations

{γ𝒌I,γ𝒌I}=γ𝒌Iγ𝒌I+γ𝒌Iγ𝒌I\displaystyle\left\{\gamma_{\bm{k}I\uparrow},\gamma_{\bm{k}I\uparrow}^{\dagger}\right\}=\gamma_{\bm{k}I\uparrow}\gamma_{\bm{k}I\uparrow}^{\dagger}+\gamma_{\bm{k}I\uparrow}^{\dagger}\gamma_{\bm{k}I\uparrow} (S69)

which is equivalent to

|uI𝒌|2+|vI𝒌|2=1.\displaystyle\absolutevalue{u_{I\bm{k}}}^{2}+\absolutevalue{v_{I\bm{k}}}^{2}=1. (S70)

Furthermore, we choose uI𝒌u_{I\bm{k}} and vI𝒌v_{I\bm{k}} to diagonalize the Hamiltonian with respect to the quasiparticle γ\gamma operators. We find a special solution by choosing

ΔI𝒌=|ΔI𝒌|eiϕ𝒌,\displaystyle\Delta_{I\bm{k}}=\absolutevalue{\Delta_{I\bm{k}}}e^{i\phi_{\bm{k}}}, uI𝒌=|uI𝒌|,\displaystyle u_{I\bm{k}}=\absolutevalue{u_{I\bm{k}}}, vI𝒌=|vI𝒌|eiϕ𝒌.\displaystyle v_{I\bm{k}}=\absolutevalue{v_{I\bm{k}}}e^{i\phi_{\bm{k}}}. (S71)

For this choice, the equation simplifies to

2ε𝒌I|uI𝒌||vI𝒌|+|ΔI𝒌|(|vI𝒌|2|uI𝒌|2)=0.\displaystyle 2\varepsilon_{\bm{k}I\uparrow}\absolutevalue{u_{I\bm{k}}}\absolutevalue{v_{I\bm{k}}}+\absolutevalue{\Delta_{I\bm{k}}}\left(\absolutevalue{v_{I\bm{k}}}^{2}-\absolutevalue{u_{I\bm{k}}}^{2}\right)=0. (S72)

Here, we used the identity ε𝒌Is=ε𝒌I¯s¯\varepsilon_{-\bm{k}Is}=\varepsilon_{-\bm{k}\bar{I}\bar{s}}, with s¯\bar{s} being the opposite spin of ss. By squaring the equation and employing Eq. (S70), we find the relations

|uI𝒌|2=12(1+ε𝒌Iε𝒌I2+|ΔI𝒌|2),\displaystyle\absolutevalue{u_{I\bm{k}}}^{2}=\frac{1}{2}\left(1+\frac{\varepsilon_{\bm{k}I\uparrow}}{\sqrt{\varepsilon_{\bm{k}I\uparrow}^{2}+\absolutevalue{\Delta_{I\bm{k}}}^{2}}}\right), |vI𝒌|2=12(1ε𝒌Iε𝒌I2+|ΔI𝒌|2).\displaystyle\absolutevalue{v_{I\bm{k}}}^{2}=\frac{1}{2}\left(1-\frac{\varepsilon_{\bm{k}I\uparrow}}{\sqrt{\varepsilon_{\bm{k}I\uparrow}^{2}+\absolutevalue{\Delta_{I\bm{k}}}^{2}}}\right). (S73a)
and
uI𝒌vI𝒌=ΔI𝒌2ε𝒌I2+|ΔI𝒌|2.\displaystyle u_{I\bm{k}}v_{I\bm{k}}=\frac{\Delta_{I\bm{k}}}{2\sqrt{\varepsilon_{\bm{k}I\uparrow}^{2}+\absolutevalue{\Delta_{I\bm{k}}}^{2}}}. (S73b)

The diagonal form of the Hamiltonian in Eq. (S67) is

HBCS=𝒌I(ε𝒌I2+|ΔI𝒌|2)γ𝒌Iγ𝒌I+(ε𝒌I2+|ΔI𝒌|2)γ𝒌I¯γ𝒌I¯.\displaystyle\begin{split}H_{\mathrm{BCS}}=\sum_{\bm{k}I}\left(\sqrt{\varepsilon_{\bm{k}I\uparrow}^{2}+\absolutevalue{\Delta_{I\bm{k}}}^{2}}\right)\gamma_{\bm{k}I\uparrow}^{\dagger}\gamma_{\bm{k}I\uparrow}+\left(\sqrt{\varepsilon_{\bm{k}I\uparrow}^{2}+\absolutevalue{\Delta_{I\bm{k}}}^{2}}\right)\gamma_{-\bm{k}\bar{I}\downarrow}^{\dagger}\gamma_{-\bm{k}\bar{I}\downarrow}.\end{split} (S74)

The quasiparticle dispersion is spin-degenerate for this specific choice of electron pairs.

To establish the superconducting gap equation, we evaluate the expectation values in Eq. (S66b) in terms of the quasiparticle operators. We find the gap equation to be

ΔI𝒌=𝒌V𝒌𝒌II¯ΔI𝒌2E𝒌Itanh((βE𝒌I2)).\displaystyle\begin{split}\Delta_{I\bm{k}}=-\sum_{\bm{k}^{\prime}}V_{\bm{k}\bm{k}^{\prime}I\bar{I}}\frac{\Delta_{I\bm{k}^{\prime}}}{2E_{\bm{k}^{\prime}I}}\tanh{\left(\frac{\beta E_{\bm{k}^{\prime}I}}{2}\right)}.\end{split} (S75)

The quasiparticle dispersion is given by

E𝒌I=(ε𝒌+ΔssI)2+|ΔI𝒌|2\displaystyle E_{\bm{k}I}=\sqrt{\left(\varepsilon_{\bm{k}}+\Delta^{I}_{ss}\right)^{2}+\absolutevalue{\Delta_{I\bm{k}}}^{2}} (S76)

where the layer dependent exchange splitting is ΔssI=Δss\Delta^{I}_{ss}=\Delta_{ss} for I{K1,K1}I\in\{K_{1},K^{\prime}_{1}\} and ΔssI=Δss\Delta^{I}_{ss}=-\Delta_{ss} for I{K2,K2}I\in\{K_{2},K^{\prime}_{2}\}. We consider the gap equation for the attractive pp-wave channel specifically. To that end, we let Veff(𝒌,𝒌,I,I¯)=Veiη(ϕ𝒌ϕ𝒌)V_{\mathrm{eff}}({\bm{k},\bm{k}^{\prime}},I,\bar{I})=Ve^{i\eta(\phi_{\bm{k}}-\phi_{\bm{k}^{\prime}})} with

V=6α2Jsd2ωM(1+6α2)2M𝒱\displaystyle V=\frac{-6\alpha^{2}J^{2}_{\mathrm{s-d}}}{\omega_{M}(1+6\alpha^{2})^{2}}\frac{M}{\mathcal{V}} (S77)

and ΔI𝒌=ΔIeiηϕ𝒌\Delta_{I\bm{k}}=\Delta_{I}e^{i\eta\phi_{\bm{k}}}. To estimate the order of magnitude of the critical temperature, we consider the gap equation to the second order in the small coupling parameter JsdJ_{\mathrm{s-d}}. In this case, the critical temperature TcT_{c} is given by the standard expression

Tc1.13ωMkBe1λ.\displaystyle T_{c}\approx\frac{1.13\omega_{M}}{k_{B}}e^{-\frac{1}{\lambda}}. (S78)

Here, kBk_{B} is the Boltzmann constant and λ=VND\lambda=VN_{D}^{\prime} where NDN_{D}^{\prime} is the electron density of states.

VI Coulomb screening and candidate materials

Twisted bilayer graphene serves as an excellent candidate for exhibiting magnon-mediated superconductivity. One of the reasons is the greatly enhanced electron density of states (DOS) close to the magic angles. Ref. [37] reports a DOS over ND=100eV1nm2N_{D}=100\;\mathrm{eV}^{-1}\mathrm{nm}^{-2} close to the magic angle. This suggests a DOS per valley per spin and per layer of ND=12.5eV1nm2N_{D}^{\prime}=12.5\;\mathrm{eV}^{-1}\mathrm{nm}^{-2}.

The Coulomb interaction in momentum space is given by

VC(q)=2πϵe2q,\displaystyle V_{C}(q)=\frac{2\pi}{\epsilon}\frac{e^{2}}{q}, (S79)

where ee is the electron charge, ϵ\epsilon is the dielectric constant and qq is the wave vector. The repulsive interaction can be detrimental to the superconducting state. However, it is largely screened by the high DOS. Ref. [38] reports a twist-angle-dependent dielectric constant ϵ\epsilon. At the magic angle, they find ϵ>250\epsilon>250 for long wavelengths. This gives rise to a significant screening effect. The Coulomb coupling strength for the relevant wavelengths is then

μ=VC(kθ)ND1.\displaystyle\mu=V_{C}(k_{\theta})N_{D}^{\prime}\approx 1. (S80)

The Coulomb coupling strength exceeds the attractive magnon-mediated coupling strength. However, it is effectively frequency independent at the magnon frequency cut-off. To account for this, we employ the Morel-Anderson model [52]. It effectively renormalizes the Coulomb coupling strength as

μ=μ1+μln(ωpωM)0.13.\displaystyle\mu^{*}=\frac{\mu}{1+\mu\ln{\frac{\omega_{p}}{\omega_{M}}}}\approx 0.13. (S81)

Here, we used the observed interband plasmon frequency ωp200meV\omega_{p}\approx 200\;\mathrm{meV} as the Coulomb interaction cut off [40]. We used ωM=0.3\omega_{M}=0.3 meV for the magnon frequency, justified by the ferromagnets used for critical temperature estimates. The effective interaction strength takes the value λ=λμ\lambda^{*}=\lambda-\mu^{*}. This shows that the Coulomb repulsion only weakly affects the superconducting state due to the TBG’s enhanced screening properties.

The on-site interaction is a substantial part of the repulsive Coulomb interaction. This part has an ss-wave symmetry. The pp-wave symmetry superconducting state circumvents the on-site repulsion completely, such that the Coulomb repulsion has an even weaker effect. We will not consider the decomposition of the Coulomb effect in further detail because the treatment is already at a crude level.

Interfacial coupling between graphene and ferromagnets has been studied both theoretically and experimentally for numerous materials [41, 42, 22, 43, 44, 45, 46]. Here, we consider two specific materials to give an order of magnitude estimate for the critical temperature.

EuO is a ferromagnetic semiconductor with Curie temperature Tc=69T_{c}=69 K. It has an fcc unit cell with lattice constant 5.15.1 Å. Hence, two magnetic Eu2+ ions per unit cell, each with spin S=7μBS=7\mu_{B}, are located at the interface and thus accessible for interfacial s-d coupling [47]. EuO thin films can be deposited on graphene epitaxially [41, 42]. The induced exchange coupling is found to be Δ=36\Delta=36 meV [48]. Due to the long periodicity of the moirè structure, we consider coupling to long-wavelength magnons. Hence, we consider magnons with a wavenumber qkθq\lesssim k_{\theta}. At this momentum, the magnon dispersion is cut off at ωM0.15\omega_{M}\approx 0.15 meV [53]. These parameters yield a coupling constant λ0.8\lambda^{*}\approx 0.8 and a critical temperature of Tc0.5T_{c}\approx 0.5 K when taking Coulomb interaction into account.

CrI3 is a van der Waals ferromagnet down to the monolayer limit [49]. The crystal has two magnetic ions per unit cell. Each magnetic ion carries a magnetic moment S=3μBS=3\mu_{B} [50]. CrI3 hosts two magnonic modes accessible for electron-magnon coupling. Their respective energies at momentum 𝒒=0\bm{q}=0 are 0.30.3 meV and 1717 meV [51]. The magnon dispersion with respect to the momentum kθk_{\theta} is negligible compared to the magnon gap. Hence, we use ωM=0.3\omega_{M}=0.3 meV. In a graphene-CrI3 heterostructure, CrI3 is theoretically found to induce an interfacial exchange splitting of 2020 meV [46]. Considering CrI3 as the ferromagnet, we find a coupling constant λ0.4\lambda^{*}\approx 0.4 and critical temperature Tc0.3T_{c}\approx 0.3 K.