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

Spin-caloric resistance of Dirac plasma in a graphene Corbino device

Alex Levchenko Department of Physics, University of Wisconsin-Madison, Madison, Wisconsin 53706, USA
(December 9, 2024)
Abstract

The thermal resistance of a spin-polarized hydrodynamic Dirac plasma in graphene is considered. A mechanism for the coupling of heat and spin flows is discussed, demonstrating that spin diffusion and spin thermocurrent modify viscous dissipation, leading to a significant enhancement of thermal resistance. Practical calculations are then presented for graphene devices in the Corbino geometry.

I Introduction

Thermoelectric (TE) and thermomagnetic (TM) transport effects in metals reflect particle-hole asymmetry and arise from the coupling of charge and heat currents Abrikosov (1988). A different class of thermomagnetoelectric (TME) effects emerges when one considers coupling of charge and heat flow with the spin current Aronov (1976); Johnson and Silsbee (1987). The distinction between TM and TME phenomena lies in the former being driven by the Lorentz force of the applied external magnetic field exerted on the charge flow, while the latter arises from the flow of spins driven by a magnetic field.

In metallic magnetic heterostructures, such as spin valves, spin-dependent thermoelectric effects can be understood through two main models. The first model involves two parallel spin-transport channels with distinct thermoelectric properties. Alternatively, these effects can be attributed to collective phenomena, influenced by spin waves, which are also present in insulating ferromagnets. For further references, refer to the collection of papers and reviews on spin caloritronics Bauer et al. (2010, 2012). The two-channel model, incorporating spin-dependent conductivities, can be derived from the kinetic equation. This model assumes different drift velocities for two spin subsystems and is particularly suitable in the weak interaction limit. Essentially, it extends the single-particle Drude-Sommerfeld theory to account for spin-polarized transport. Consequently, it adheres to Mott’s law for the Seebeck coefficient and the Wiedemann-Franz law for spin-dependent thermal conductivity. In contrast, in the collision-dominated regime, the drift velocity is the same for all the spin components. The reason is that frequent collisions between electrons with different spins form a common drift of the electron system. This scenario can be realized in high-mobility and low-density semiconductor and semimetal devices, where strong correlations lead to the electron liquid reaching a hydrodynamic limit Gurzhi et al. (2006); Andreev et al. (2011). In this situation, the coupling of spin and heat currents exhibits a distinct character, leading to the specific mechanism of thermal resistance explored in this paper.

The hydrodynamic behavior of the electron liquid has been demonstrated in a number of recent experiments, see reviews Lucas and Fong (2018); Levchenko and Schmalian (2020); Narozhny (2022); Fritz and Scaffidi (2024) on this topic and references therein. Of particular relevance to this work are strong thermoelectric anomalies observed in graphene devices tuned to the proximity of charge neutrality Crossno et al. (2016); Ghahari et al. (2016). Physically, these profound features, namely giant enhancement of the Lorenz ratio and violation of the Mott relation by inelastic scattering, stem from the decoupling of charge and heat flows.

The hydrodynamic equations in the presence of finite spin-polarization of an electron liquid induced by an in-plane magnetic field should be supplemented by the conservation law for spin. This is particularly relevant to graphene multilayers, as the spin relaxation time is known to be long, even at room temperature, in the range of nanoseconds, with corresponding spin diffusion lengths reaching 10 μ\mum Drögeler et al. (2014, 2016). In the linear response, the spin current consists of a spin diffusion current and a spin thermocurrent induced by temperature gradients. According to Onsager’s reciprocity principle Onsager (1931), the heat flux acquires an additional spin-dependent contribution related to the Peltier effect. These processes modify the viscous flow of the electron fluid, thus influencing dissipative friction and consequently strongly increasing thermal resistance.

Refer to caption
Figure 1: A schematic for the graphene Corbino disk. Inner electrode has radius r1r_{1} and is kept at temperature T1T_{1}. The outer electrode has radius r2r_{2} and temperature T2<T1T_{2}<T_{1}. The arrows pointing outwards from the center mark heat fluxes carried by electron fluid in graphene layer which is subjected to the in-plane magnetic field HH.

To explore this physics, a particularly useful device geometry is a graphene Corbino membrane, as illustrated in Fig. 1. In this setup, the continuity of currents enforces the electronic flow profile to be spatially nonuniform, even at uniform particle and entropy density. This amplifies the interplay of dissipative viscous, spin-diffusion, and spin-thermal effects. It should be noted that charge and thermal transport measurements in gated graphene Corbino devices have been reported, including observations of viscous and magneto-thermal effects Yan and Fuhrer (2010); Faugeras et al. (2010); Waissman et al. (2024); Talanov et al. (2024); Zeng et al. (2024). Additionally, powerful thermal and electrochemical potential imaging techniques can be implemented Waissman et al. (2022); Kumar et al. (2022); Krebs et al. (2023); Völkl et al. (2024). Thus far, theoretical works have focused on the orbital effects of the magnetic field in the magnetotransport Shavit et al. (2019); Kavokin et al. (2020); Levchenko et al. (2022); Gall et al. (2023), while hydrodynamic effects dominated by spin thermotransport remain unexplored, which motivates this work.

II Hydrodynamic theory

To this end, consider a two-dimensional electron system subjected to an in-plane magnetic field. In the regime where the rate of momentum-conserving electron-electron collisions exceeds the rates of momentum, energy, and spin relaxation, the corresponding macroscopic hydrodynamic equations of the electron fluid can be written as follows:

tn+𝒋=0,\partial_{t}n+\bm{\nabla}\cdot\bm{j}=0, (1a)
ts+𝒋s=s˙,\partial_{t}s+\bm{\nabla}\cdot\bm{j}_{s}=\dot{s}, (1b)
tσ+𝒋σ=σ˙,\partial_{t}\sigma+\bm{\nabla}\cdot\bm{j}_{\sigma}=\dot{\sigma}, (1c)
t𝒑+k𝒖+P+enϕ=𝚺^.\partial_{t}\bm{p}+k\bm{u}+\bm{\nabla}P+en\bm{\nabla}\phi=\bm{\nabla}\cdot\hat{\bm{\Sigma}}. (1d)

Here PP is the pressure, 𝒖\bm{u} is the hydrodynamic velocity, kk denotes the disorder-induced friction coefficient, and 𝚺^Σij\hat{\bm{\Sigma}}\equiv\Sigma_{ij} denotes the viscous stress tensor Landau and Lifshitz (1987)

Σij=η(iuj+jui)+(ζη)δijkuk,\Sigma_{ij}=\eta(\partial_{i}u_{j}+\partial_{j}u_{i})+(\zeta-\eta)\delta_{ij}\partial_{k}u_{k}, (2)

with η\eta and ζ\zeta being, respectively, shear (first) and bulk (second) viscosity of the electron liquid. The electric potential ϕ\phi is related to the electron charge density enen by the Poisson equation. The densities of particles, entropy, momentum, and spin are denoted by nn, ss, 𝒑\bm{p}, and σ\sigma respectively. The corresponding current densities of particles, entropy, and spin are labeled respectively as 𝒋\bm{j}, 𝒋s\bm{j}_{s}, and 𝒋σ\bm{j}_{\sigma}. The local rate of entropy production due to electron-electron collisions is denoted by s˙\dot{s} in Eq. (1b) Likewise, σ˙\dot{\sigma} in Eq. (1c) corresponds to the spin relaxation. The continuity equation for the spin current is written in analogy to the hydrodynamic theory of spin waves Halperin and Hohenberg (1969). It should be noted that only spin component along the field appears in this framework as the other spin components are not conserved due to spin precession.

We proceed to determine the entropy production rate. For this purpose, we recall that the entropy density ss is a function of the conserved quantities, and its differential is given by the general thermodynamic relation Landau and Lifshitz (1980)

ds=dεT(μ+eϕ)dnTμσdσT𝒖d𝒑T,ds=\frac{d\varepsilon}{T}-\frac{(\mu+e\phi)dn}{T}-\frac{\mu_{\sigma}d\sigma}{T}-\frac{\bm{u}\cdot d\bm{p}}{T}, (3)

where μ\mu and μσ\mu_{\sigma} are chemical potentials for particle and spin. Noting that all these thermodynamic variables correspond to conserved quantities, we express their time derivatives in terms of the divergences of the corresponding currents from Eq. (1). This yields the following expression for S˙=s˙𝑑𝒓\dot{S}=\int\dot{s}d\bm{r},

S˙=1T[𝒋ε+(μ+eϕ)𝒋+μσ𝒋σ\displaystyle\dot{S}=\int\frac{1}{T}\left[-\bm{\nabla}\cdot\bm{j}_{\varepsilon}+(\mu+e\phi)\bm{\nabla}\cdot\bm{j}+\mu_{\sigma}\bm{\nabla}\cdot\bm{j}_{\sigma}\right.
+k𝒖2+n𝒖eϕ+𝒖𝚷^]d𝒓,\displaystyle\left.+k\bm{u}^{2}+n\bm{u}\bm{\nabla}e\phi+\bm{u}\cdot\bm{\nabla}\cdot\hat{\bm{\Pi}}\right]d\bm{r}, (4)

where we used tε=𝒋ε\partial_{t}\varepsilon=-\bm{\nabla}\cdot\bm{j}_{\varepsilon} for the energy current density 𝒋ε\bm{j}_{\varepsilon}, and similarly for tn\partial_{t}n, tσ\partial_{t}\sigma, and t𝒑\partial_{t}\bm{p} from Eqs. (1a)-(1c). The tensor 𝚷^Πij=PδijΣij\hat{\bm{\Pi}}\equiv\Pi_{ij}=P\delta_{ij}-\Sigma_{ij} denotes the local part of the momentum flux. Next, using the thermodynamic identity P=nμ+sT+σμσ\bm{\nabla}P=n\bm{\nabla}\mu+s\bm{\nabla}T+\sigma\bm{\nabla}\mu_{\sigma} and integrating by parts we get

TS˙=[𝒋sT𝒋(μ+eϕ)𝒋σμσ\displaystyle T\dot{S}=\int\left[-\bm{j}^{\prime}_{s}\bm{\nabla}T-\bm{j}^{\prime}\bm{\nabla}(\mu+e\phi)-\bm{j}^{\prime}_{\sigma}\bm{\nabla}\mu_{\sigma}\right.
+k𝒖2+Σijjui]d𝒓\displaystyle\left.+k\bm{u}^{2}+\Sigma_{ij}\partial_{j}u_{i}\right]d\bm{r} (5)

where we introduced dissipative fluxes 𝒋s=𝒋ss𝒖\bm{j}^{\prime}_{s}=\bm{j}_{s}-s\bm{u}, 𝒋=𝒋n𝒖\bm{j}^{\prime}=\bm{j}-n\bm{u}, and 𝒋σ=𝒋σσ𝒖\bm{j}^{\prime}_{\sigma}=\bm{j}_{\sigma}-\sigma\bm{u} that are proportional to gradients of the equilibrium parameters. The entropy flux is related to the energy flux as follows 𝒋s=1T[𝒋ε(μ+eϕ)𝒋μσ𝒋σ]\bm{j}_{s}=\frac{1}{T}[\bm{j}_{\varepsilon}-(\mu+e\phi)\bm{j}-\mu_{\sigma}\bm{j}_{\sigma}].

In a situation with a vanishing electromotive force, e𝓔=(μ+eϕ)e\bm{\mathcal{E}}=-\bm{\nabla}(\mu+e\phi), the second term in the above expression is absent since 𝒋=0\bm{j}^{\prime}=0. Recalling the definition S˙=iXix˙i\dot{S}=-\sum_{i}X_{i}\dot{x}_{i} from Ref. Landau and Lifshitz (1980) and identifying x˙i\dot{x}_{i} with the entropy current 𝒋s\bm{j}^{\prime}_{s} and the spin diffusion current 𝒋σ\bm{j}^{\prime}_{\sigma} we conclude that the role of the corresponding thermodynamic forces is played by T/T\bm{\nabla}T/T and μσ/T\bm{\nabla}\mu_{\sigma}/T respectively. Introducing the kinetic coefficients following the standard convention x˙i=γijXj\dot{x}_{i}=-\gamma_{ij}X_{j} we write

(𝒋s𝒋σ)=1T(γ11γ12γ21γ22)(Tμσ).\left(\begin{array}[]{c}\bm{j}^{\prime}_{s}\\ \bm{j}^{\prime}_{\sigma}\end{array}\right)=-\frac{1}{T}\left(\begin{array}[]{cc}\gamma_{11}&\gamma_{12}\\ \gamma_{21}&\gamma_{22}\end{array}\right)\left(\begin{array}[]{c}\bm{\nabla}T\\ \bm{\nabla}\mu_{\sigma}\end{array}\right). (6)

We can express γ11=κ\gamma_{11}=\kappa in terms of the thermal conductivity κ\kappa, and γ22\gamma_{22} in terms the spin diffusion DσD_{\sigma} as γ22=TDσ\gamma_{22}=TD_{\sigma}. Since the spin density changes sign with respect to time reversal symmetry and energy does not, the Onsager symmetry principle gives γ12(H)=γ21(H)\gamma_{12}(H)=-\gamma_{21}(-H). On the other hand both γ12(H)\gamma_{12}(H) and γ21(H)\gamma_{21}(H) are odd functions of the magnetic field HH. Therefore, we conclude that γ12(H)=γ21(H)\gamma_{12}(H)=\gamma_{21}(H). Thus the matrix of kinetic coefficients is symmetric. In the following we denote γ12γσ\gamma_{12}\equiv\gamma_{\sigma} and introduce

Ξ^γ^/T=(κ/Tγσ/Tγσ/TDσ).\hat{\Xi}\equiv\hat{\gamma}/T=\left(\begin{array}[]{cc}\kappa/T&\gamma_{\sigma}/T\\ \gamma_{\sigma}/T&D_{\sigma}\end{array}\right). (7)

As a reminder, for temperatures T<EFT<E_{F} in the Fermi liquid theory κEF2/T\kappa\sim E^{2}_{F}/T Abrikosov and Khalatnikov (1959) and DσA/T2D_{\sigma}\sim A/T^{2} Hone (1961), modulo logarithmic renormalizations specific to the 2D case, see e.g. Ref. Lyakhov and Mishchenko (2003). A detailed analytical calculations of these coefficients in the limit of weak interactions, rs=V/EF1r_{s}=V/E_{F}\lesssim 1, where VV is the characteristic energy of Coulomb interaction and EFE_{F} is the bare Fermi energy, can be found in Ref. Sykes and Brooker (1970).

For compactness of further expressions, it is convenient to introduce the column-vector notations

x=(sσ),𝑿=(Tμσ).\vec{x}=\left(\begin{array}[]{c}s\\ \sigma\end{array}\right),\quad\vec{\bm{X}}=\left(\begin{array}[]{c}\bm{\nabla}T\\ \bm{\nabla}\mu_{\sigma}\end{array}\right). (8)

Then it can be readily checked that the sum of the first and third term in the right hand side of Eq. (II) can be rewritten as 𝑿𝕋Ξ^𝑿\vec{\bm{X}}^{\mathbb{T}}\hat{\Xi}\vec{\bm{X}}, with the superscript 𝕋\mathbb{T} denoting the transposition. As a result,

TS˙=[Σijjui+𝑿𝕋Ξ^𝑿+k𝒖2]𝑑𝒓,T\dot{S}=\int\left[\Sigma_{ij}\partial_{j}u_{i}+\vec{\bm{X}}^{\mathbb{T}}\hat{\Xi}\vec{\bm{X}}+k\bm{u}^{2}\right]d\bm{r}, (9)

where summation over the repeated indices is implicit. The current densities take the final form in explicit notations

𝒋s=s𝒖κTTγσTμσ,\bm{j}_{s}=s\bm{u}-\frac{\kappa}{T}\bm{\nabla}T-\frac{\gamma_{\sigma}}{T}\bm{\nabla}\mu_{\sigma}, (10a)
𝒋σ=σ𝒖γσTTDσμσ,\bm{j}_{\sigma}=\sigma\bm{u}-\frac{\gamma_{\sigma}}{T}\bm{\nabla}T-D_{\sigma}\bm{\nabla}\mu_{\sigma}, (10b)

where the first term in the right hand side represents the equilibrium components of the currents, while the remaining two terms in each expression represent the dissipative components. With the appropriate boundary conditions for a given device geometry, the system of equations Eqs. (1), (9), and (10) enable determination of the linear response transport coefficients (see Appendix A for additional discussion).

III Graphene Corbino device

We apply the above formalism to consider heat transport in a graphene flake shaped like a Corbino disk, see Fig. 1. The aspect ratio of the disk is defined by the ratio of its radii, p=r2/r1>1p=r_{2}/r_{1}>1. In the linear response regime, a small heat current is induced in the device by a temperature difference ΔT=T1T2\Delta T=T_{1}-T_{2} between the electrodes. The calculations presented below will focus on the regime of charge neutrality, where n0n\to 0. For simplicity, we also ignore the effect of bulk viscosity, which is known to vanish for both parabolic and linear spectra of quasiparticle excitations Pitaevskii and Lifshitz (1981). The latter corresponds to the case of graphene monolayer.

In a steady state, the continuity equation of entropy and spin currents can be presented in the form specific to the cylindrical geometry Li et al. (2022); Levchenko et al. (2022)

xu(r)Ξ^X(r)=I2πr,I=(IsIσ),\vec{x}u(r)-\hat{\Xi}\vec{X}(r)=\frac{\vec{I}}{2\pi r},\quad\vec{I}=\left(\begin{array}[]{c}I_{s}\\ I_{\sigma}\end{array}\right), (11)

where u(r)u(r) is the radial hydrodynamic velocity and X(r)\vec{X}(r) is the column-vector of thermodynamically conjugate forces, with r[r1,r2]r\in[r_{1},r_{2}] being the radial coordinate. The force balance condition exerted on an element of the fluid takes the form of the Navier-Stokes equation, which can be found with the help of Eqs. (1d) and (2):

ηΔ^u(r)ku(r)=x𝕋X,Δ^=1rddr(rddr)1r2.\eta\hat{\Delta}u(r)-ku(r)=\vec{x}^{\mathbb{T}}\vec{X},\quad\hat{\Delta}=\frac{1}{r}\frac{d}{dr}\left(r\frac{d}{dr}\right)-\frac{1}{r^{2}}. (12)

Equations (11) and (12) determine the spatial dependence of the flow velocity, temperature gradients, and gradients of the spin chemical potential in the interior of the Corbino disk in terms of the entropy and spin currents I\vec{I}. The crucial point to realize here is that even though thermal resistance is defined under the global condition of vanishing spin current, locally dissipative parts of these currents are nonzero due to finite X(r)\vec{X}(r) in the bulk of the flow. This can be seen explicitly.

For this purpose, we use Eq. (11) to exclude X(r)\vec{X}(r) and arrive at the equation for u(r)u(r) in the form

Δ^ukuul2=12πrη(x𝕋Ξ^1I).\hat{\Delta}u-ku-\frac{u}{l^{2}}=-\frac{1}{2\pi r\eta}(\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{I}). (13)

Here we introduced the characteristic length scale ll defined by

l2=x𝕋Ξ^1xη=s2[Dσ2σγσsT+σ2κs2T]η[κDσTγσ2T2].l^{-2}=\frac{\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{x}}{\eta}=\frac{s^{2}\left[D_{\sigma}-2\frac{\sigma\gamma_{\sigma}}{sT}+\frac{\sigma^{2}\kappa}{s^{2}T}\right]}{\eta\left[\frac{\kappa D_{\sigma}}{T}-\frac{\gamma^{2}_{\sigma}}{T^{2}}\right]}. (14)

The general solution of Eq. (13) consists of the linear superposition of solutions to the homogeneous equation and a particular solution due to the right hand side. The former is given by a linear combination of modified Bessel functions of the first and second kinds. These functions describe deviations of the hydrodynamic flow from that in the bulk of the disk. Due to properties of Bessel functions they are exponentially localized near the inner and outer boundaries on the length scale of the order of ll. For the graphene monolayer this length scale is of the order of thermal de Broglie length llT=v/Tl\sim l_{T}=v/T, which in itself is of the order of electron mean free path leel_{\text{ee}}. Therefore, these solutions contribute to the resistance of the contacts. Consideration of these effects goes beyond the limit of hydrodynamic theory. We are interested in the solution that describes the hydrodynamic mode in the bulk of the device. This solution can be easily found as a particular solution of the inhomogeneous equation, yielding the result

u(r)=f2πrx𝕋Ξ^1Ix𝕋Ξ^1x,f=11+kl2/η.u(r)=\frac{f}{2\pi r}\frac{\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{I}}{\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{x}},\quad f=\frac{1}{1+kl^{2}/\eta}. (15)

The factor defined by ff can be safely set to unity. Indeed, viscosity of the Dirac fluid is η(T/v)2lT2\eta\sim(T/v)^{2}\sim l^{-2}_{T} Müller et al. (2009). In addition, friction coefficient can be related to the spatial average of the local density fluctuations δn(𝒓)\delta n(\bm{r}) of the globally charge neutral system, namely ke2ςδn2k\sim\frac{e^{2}}{\varsigma}\langle\delta n^{2}\rangle Li et al. (2020), where ς\varsigma is the intrinsic conductivity of graphene Mishchenko (2008); Fritz et al. (2008); Kashuba (2008). Therefore, kl2/ηδn2lT41kl^{2}/\eta\sim\langle\delta n^{2}\rangle l^{4}_{T}\ll 1. Based on this argument, we set f1f\to 1 in what follows. Knowledge of u(r)u(r) enables us to compute dissipative friction term and also the viscous term in Eq. (9). For a purely radial flow, there are only two nonvanishing components of the stress tensor that contribute to the dissipated power. In cylindrical coordinates, these are

Σrr=2ηur,Σϕϕ=2ηur.\Sigma_{rr}=2\eta\frac{\partial u}{\partial r},\quad\Sigma_{\phi\phi}=2\eta\frac{u}{r}. (16)

As the next step, we need to determine X(r)\vec{X}(r), which governs contribution of the relative transport mode to the dissipation rate in Eq. (9). The required solution can be obtained by multiplying Eq. (11) by a row vector ϰ𝕋=(σ,s)\vec{\varkappa}^{\mathbb{T}}=(\sigma,-s) from the left. The result is

X(r)=ϰ2πrϰ𝕋Iϰ𝕋Ξ^ϰ,\vec{X}(r)=-\frac{\vec{\varkappa}}{2\pi r}\frac{\vec{\varkappa}^{\mathbb{T}}\vec{I}}{\vec{\varkappa}^{\mathbb{T}}\hat{\Xi}\vec{\varkappa}}, (17)

which completes solution of the hydrodynamic equations.

IV Spin-caloric resistance

We introduce spin caloric resistance matrix ^\hat{\mathcal{R}} that can be determined by equating the Joule heat,

𝒫=𝕋^,𝕋=(Iq,Iσ),\mathcal{P}=\vec{\mathcal{I}}^{\mathbb{T}}\hat{\mathcal{R}}\vec{\mathcal{I}},\quad\vec{\mathcal{I}}^{\mathbb{T}}=(I_{q},I_{\sigma}), (18)

to the rate of energy dissipation in the bulk flow as defined by Eq. (9). Here Iq=TIsI_{q}=TI_{s} is the heat current. The thermal resistance RthR_{\text{th}} can be found by setting the spin current to zero Iσ0I_{\sigma}\to 0. The computed dissipation in Eq. (18) is then simply 𝒫=11Iq2\mathcal{P}=\mathcal{R}_{11}I^{2}_{q}. On the other hand, employing general thermodynamic relations, one finds the entropy production rate S˙=IqΔT/T2=RthIq2/T2\dot{S}=I_{q}\Delta T/T^{2}=R_{\text{th}}I^{2}_{q}/T^{2}. As a consequence, from the condition 𝒫=TS˙\mathcal{P}=T\dot{S} we determine that

Rth=T11=1TIs2(TS˙)Iσ0.R_{\text{th}}=T\mathcal{R}_{11}=\frac{1}{TI^{2}_{s}}(T\dot{S})_{I_{\sigma}\to 0}. (19)

We insert Eqs. (15), (29), and (17) into Eq. (9), set Iσ0I_{\sigma}\to 0, compute each of the radial integrals individually, and bring the resulting expression into Eq. (19) that will cancel Is2I^{2}_{s} in the denominator. The resulting formula can be written as a sum of respective contributions (see Appendix B for technical details)

Rth=RΣ+RΞ+Rk.R_{\text{th}}=R_{\Sigma}+R_{\Xi}+R_{k}. (20)

The first term describes viscous dissipation modified by the spin diffusion and spin thermocurrent, which reads

RΣ=η(p21)πTs2r22(DσσγσsT)2(Dσ+σsβσ)2,R_{\Sigma}=\frac{\eta(p^{2}-1)}{\pi Ts^{2}r^{2}_{2}}\frac{\left(D_{\sigma}-\frac{\sigma\gamma_{\sigma}}{sT}\right)^{2}}{\left(D_{\sigma}+\frac{\sigma}{s}\beta_{\sigma}\right)^{2}}, (21a)
with the notation βσ=(σκ/sT)(2γσ/T)\beta_{\sigma}=(\sigma\kappa/sT)-(2\gamma_{\sigma}/T). The second term captures contribution of the relative mode and is expressed solely in terms of the spin-dependent intrinsic dissipative coefficients of matrix Ξ^\hat{\Xi}
RΞ=12πT(σ/s)2lnpDσ+σsβσ.R_{\Xi}=\frac{1}{2\pi T}\frac{(\sigma/s)^{2}\ln p}{D_{\sigma}+\frac{\sigma}{s}\beta_{\sigma}}. (21b)
The third term describes an extra contribution due to disorder-induced friction
Rk=klnp2πTs2(DσσγσsT)2(Dσ+σsβσ)2.R_{k}=\frac{k\ln p}{2\pi Ts^{2}}\frac{\left(D_{\sigma}-\frac{\sigma\gamma_{\sigma}}{sT}\right)^{2}}{\left(D_{\sigma}+\frac{\sigma}{s}\beta_{\sigma}\right)^{2}}. (21c)

Eqs. (46) and (21) represent the main results of this work. Next we analyze several limiting cases.

In pristine systems, k0k\to 0, and without magnetic field, σ0\sigma\to 0, the result simplifies to Rth=η(p21)/πTs2r22R_{\text{th}}=\eta(p^{2}-1)/\pi Ts^{2}r^{2}_{2} (see Appendix C for an alternative derivation). For a graphene monolayer ηs(T/v)2\eta\sim s\sim(T/v)^{2}, so that Rth1/T3R_{\text{th}}\propto 1/T^{3}. In contrast, in bilayer graphene (BLG) the entropy density has different temperature dependence, smTs\sim m^{*}T, where mm^{*} is the effective mass of the band structure. From the bound on the viscosity to entropy ratio, known from the strongly interacting quantum field theories Kovtun et al. (2005), one can deduce that ηT\eta\propto T, therefore Rth1/T2R_{\text{th}}\propto 1/T^{2} in BLG. These predictions can be tested in experiments.

The leading field dependent correction to RthR_{\text{th}} comes from RΞR_{\Xi}. At low magnetic fields, we write the spin density in the form σ=χH\sigma=\chi H, where χ\chi denotes the spin susceptibility. As a result, for the relative thermal magnetoresistance, one obtains

ΔRth(H)Rth(H)Rth(0)Rth(0)p2lnp2(p21)r12(χH)2ηDσ,\Delta R_{\text{th}}(H)\equiv\frac{R_{\text{th}}(H)-R_{\text{th}}(0)}{R_{\text{th}}(0)}\approx\frac{p^{2}\ln p}{2(p^{2}-1)}\frac{r^{2}_{1}(\chi H)^{2}}{\eta D_{\sigma}}, (22)

which is controlled by the spin diffusion coefficient. We conclude that the spin caloric effect enables the extraction of DσD_{\sigma} and, thus, the spin conductance via the corresponding Einstein relation, in the transport regime dominated by electron collisions. This approach could offer valuable complementary experimental tools in addition to earlier methods based on spin Coulomb drag measurements Weber et al. (2005).

These results can be readily generalized to the magnetotransport away from charge neutrality. There are three key modifications that need to be incorporated. (i) At finite particle density nn, the thermoelectric effects couple charge and heat flows. Therefore, we have to augment Eq. (10) by the respective expression for the electric current density

𝒋e=en𝒖+ς𝑬γTT,\bm{j}_{e}=en\bm{u}+\varsigma\bm{E}-\frac{\gamma}{T}\bm{\nabla}T, (23)

where γ\gamma is the intrinsic thermoelectric coefficient and the electromotive force (EMF) e𝑬e\bm{E} is the gradient of the local electrochemical potential. It is important to realize that finite 𝑬\bm{E} is generated even in a purely thermal bias setup. Since thermal resistance is defined at vanishing spin and electric currents, vanishing of 𝒋𝒆\bm{j_{e}} implies a local relationship 𝑬=(en/ς)𝒖\bm{E}=-(en/\varsigma)\bm{u}. Note that at nsn\ll s the thermoelectric contribution to 𝒋e\bm{j}_{e} may be neglected because γ/Tn/s1\gamma/T\sim n/s\ll 1. (ii) The Coulomb force en𝑬-en\bm{E} must be added to the right-hand-side of the force balance condition, Eq. (12). Inserting 𝑬\bm{E} into the force-balance condition gives a force term (en)2𝒖/ς(en)^{2}\bm{u}/\varsigma, which amounts to the following redefinition of the effective friction coefficient in Eq. (12), kk(n)=k+e2ςn2k\to k(n)=k+\frac{e^{2}}{\varsigma}n^{2}. The form of the equation remains intact thus its solution for 𝒖\bm{u}, which is still given by Eq. (15). (iii) An additional term must be added to the entropy production rate in Eq. (9), which takes a simple form

TS˙=ς𝑬2𝑑𝒓T\dot{S}=\varsigma\int\bm{E}^{2}d\bm{r} (24)

with the neglect of thermoelectric contribution γ/T1\propto\gamma/T\ll 1. With these ingredients, we use Eq. (15) to calculate Eq. (24) and thus determine additional contribution to thermal resistance due to intrinsic electric conductivity

Rς=e2ςn2s2f2lnp2πT(DσσγσsT)2(Dσ+σsβσ)2.R_{\varsigma}=\frac{e^{2}}{\varsigma}\frac{n^{2}}{s^{2}}\frac{f^{2}\ln p}{2\pi T}\frac{\left(D_{\sigma}-\frac{\sigma\gamma_{\sigma}}{sT}\right)^{2}}{\left(D_{\sigma}+\frac{\sigma}{s}\beta_{\sigma}\right)^{2}}. (25)

In this limit ff defined in Eq. (15) has noticeable density dependence. Therefore RΣR_{\Sigma} and RkR_{k} should be also multiplied by a factor f2f^{2}, since ufu\propto f, but not RΞR_{\Xi}, which remains the same. The limit of charge neutrality corresponds to nδn2n\ll\sqrt{\langle\delta n^{2}\rangle}. In modern encapsulated devices density inhomogeneity of monolayer graphene is in the range δn2(5÷10)×109\sqrt{\langle\delta n^{2}\rangle}\sim(5\div 10)\times 10^{9} cm-2. This is still well within the domain of Dirac plasma {n,δn2}<s\{n,\sqrt{\langle\delta n^{2}\rangle}\}<s at T100T\sim 100 K. In this limit one may approximate fs2/(s2+e2ςn2)f\approx s^{2}/(s^{2}+\frac{e^{2}}{\varsigma}n^{2}). Therefore, RΞR_{\Xi} remains the leading spin-dependent contribution to thermal resistance at all relevant densities n<sn<s.

V Estimates

It is instructive to estimate the order of magnitude of computed thermal resistances. For this purpose consider Eq. (46) in the limit σ0\sigma\to 0 as an example. Recall that for a monolayer graphene ηs(kBT/v)2\eta\sim s\sim(k_{B}T/\hbar v)^{2}, where Boltzmann and Planck constants were restored to have proper units in SI system. Let us take T100T\sim 100 K and v106v\sim 10^{6} m/s. The typical radius of the Corbino disk in experiments is r1,21μr_{1,2}\gtrsim 1\mum. For this values, let us convert the conventional K/W units of RthR_{\text{th}} into Ω\Omega units of electrical resistance by using the Lorenz number LL as a conversion factor through the Wiedemann-Franz law L=Rel/TRthL=R_{\text{el}}/TR_{\text{th}}. This gives Rth(/e2)1/(sr2)100ΩR_{\text{th}}\sim(\hbar/e^{2})1/(sr^{2})\sim 100\,\Omega. The measured values of thermal resistance reported in Ref. Talanov et al. (2024) are in the range 50÷600Ω\sim 50\div 600\,\Omega for temperatures in the range T25÷200T\sim 25\div 200 K. The estimate matches well.

The kinematic viscosity measured in Corbino geometry as reported in Refs. Talanov et al. (2024); Zeng et al. (2024) is in the range ν0.1÷0.5\nu\sim 0.1\div 0.5 m2/s. The kinematic viscosity is given by the shear viscosity per mass density ν=η/ρ\nu=\eta/\rho. The mass density of the Dirac thermal cloud can be estimated as ρT3/v4\rho\sim T^{3}/v^{4}. Putting this together and restoring proper conversion units one finds νv2/kBT\nu\sim\hbar v^{2}/k_{B}T, which gives ν0.1\nu\sim 0.1 m2/s for T=100T=100 K. Again this gives good agreement.

It is harder to estimate the magnitude of the relative MR in Eq. (22) because of unknown values of spin diffusion and spin susceptibility in the hydrodynamic regime, but form the structure of the expression it is expected that the result is significant even at relatively weak field.

Acknowledgments

I am grateful to Anton Andreev and Philip Kim for insightful discussions. This work was supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences (BES) under Award No. DE-SC0020313 and H. I. Romnes Faculty Fellowship provided by the University of Wisconsin-Madison Office of the Vice Chancellor for Research and Graduate Education with funding from the Wisconsin Alumni Research Foundation.

Appendix A Assumptions and approximations

In the main part of this paper, the calculation of the spin thermoresistance is described under the assumption of the hydrodynamic regime of electron flow, characterized by frequent collisions and strong correlations. The key physical insight developed in this work relates to the pronounced spin-heat transport coupling near charge neutrality, where thermal anomalies are expected to be strong. The coupling equations for the spin and entropy currents in the presence of a temperature gradient and a spin chemical potential were derived starting from the generic conservation equations for entropy, charge, spin, and the force-balance condition. These equations are solved in a Corbino disk geometry.

The applicability of the hydrodynamic limit assumes that the electron mean free path leel_{\text{ee}}, due to momentum-conserving collisions, is the shortest length scale in the problem. For monolayer graphene (MLG), this scale is of the order of the thermal de Broglie length leelT=v/Tl_{\text{ee}}\sim l_{T}=v/T. Momentum-relaxing collisions on disorder induce friction on the liquid flow. For smooth disorder potential, the friction coefficient kk can be expressed in terms of the local variations of the particle density δn(𝒓)\delta n(\bm{r}) and entropy density δs(𝒓)\delta s(\bm{r}) as follows Li et al. (2020)

k=(sδnnδs)22(n2κT2nsγT+s2ςe2).k=\frac{\langle(s\delta n-n\delta s)^{2}\rangle}{2\left(\frac{n^{2}\kappa}{T}-\frac{2ns\gamma}{T}+\frac{s^{2}\varsigma}{e^{2}}\right)}. (26)

At charge neutrality, n0n\to 0, intrinsic thermoelectric coefficients vanishes, γ0\gamma\to 0, and kk simplifies to k=e22ςδn2k=\frac{e^{2}}{2\varsigma}\langle\delta n^{2}\rangle, which is determined by the intrinsic conductivity. For MLG devices the intrinsic conductivity is known to be of the order of conductance quantum e2/2π\sim e^{2}/2\pi (in units of =1\hbar=1) modulo logarithmic renormalizations in the weak-coupling theory Mishchenko (2008); Fritz et al. (2008); Kashuba (2008)

ς=e22παT2,αT=αg1+αg4lnΛT,\varsigma=\frac{e^{2}}{2\pi\alpha^{2}_{T}},\qquad\alpha_{T}=\frac{\alpha_{g}}{1+\frac{\alpha_{g}}{4}\ln\frac{\Lambda}{T}}, (27)

where αg=e2/εv\alpha_{g}=e^{2}/\varepsilon v is the temperature-independent dimensionless interaction constant (ε\varepsilon is the dielectric constant of the surrounding medium), which determines the bare strength of the electron-electron interactions, and Λ\Lambda is the cutoff in the scheme of the renormalization group.

In the model explored in this paper, a long-range disorder potential is assumed, characterized by a correlation radius ξ\xi longer than the equilibration length but shorter than the distance between the electrodes, namely lT<ξ<r2r1l_{T}<\xi<r_{2}-r_{1}. These conditions can be securely met in modern devices. Indeed, for boron nitride encapsulated graphene devices, scanning probes reveal that the correlation radius of these fluctuations is somewhere in the range ξ100\xi\sim 100 nm and local strength is in the range of 5\sim 5 meV. Corbino devices can be designed with the radial difference 5μ\sim 5\mum.

In the model long spin relaxation time is assumed. This is particularly relevant to graphene multilayers, as the spin relaxation, even at room temperature, in the range of nanoseconds, with corresponding spin diffusion lengths reaching lσ10μl_{\sigma}\sim 10\mum.

Appendix B Spin-caloric magnetoresistance

In this section we provide additional details for the computation of the thermal resistances RΣR_{\Sigma}, RΞR_{\Xi}, and RkR_{k}, which enter Eq. (20) in the main text.

B.1 Viscous contribution

A particular solution of the hydrodynamic equations for the radial flow is just the consequence of the continuity of entropy current. As shown in the main text [Eq. (15)], it takes the form

u(r)=A2πr,A=(x𝕋Ξ^1I)(x𝕋Ξ^1x).u(r)=\frac{A}{2\pi r},\quad A=\frac{(\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{I})}{(\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{x})}. (28)

This expression gives us two nonvanishing components of the stress tensor

Σrr=2ηur=2ηA2πr2,Σϕϕ=2ηur=2ηA2πr2.\Sigma_{rr}=2\eta\frac{\partial u}{\partial r}=-2\eta\frac{A}{2\pi r^{2}},\quad\Sigma_{\phi\phi}=2\eta\frac{u}{r}=2\eta\frac{A}{2\pi r^{2}}. (29)

Using the definition of the stress tensor Σij\Sigma_{ij} Landau and Lifshitz (1987), the entropy production can be equivalently rewritten as follows

TS˙=12ηijΣij2d2r=12η(Σrr2+Σϕϕ2)d2rT\dot{S}=\frac{1}{2\eta}\sum_{ij}\int\Sigma^{2}_{ij}d^{2}r=\frac{1}{2\eta}\int(\Sigma^{2}_{rr}+\Sigma^{2}_{\phi\phi})d^{2}r (30)

where integration goes over the two-dimensional area of the Corbino disk. Since Σrr2=Σϕϕ2\Sigma^{2}_{rr}=\Sigma^{2}_{\phi\phi} for u1/ru\propto 1/r flow, we thus have

TS˙=12η×2×r1r2(ηAπr2)22πr𝑑r=ηA2π(1r121r22).T\dot{S}=\frac{1}{2\eta}\times 2\times\int^{r_{2}}_{r_{1}}\left(\frac{\eta A}{\pi r^{2}}\right)^{2}2\pi rdr=\frac{\eta A^{2}}{\pi}\left(\frac{1}{r^{2}_{1}}-\frac{1}{r^{2}_{2}}\right). (31)

As the next step, we calculate the coefficient AA, which consists of two terms:

x𝕋Ξ^1I=1DetΞ^[(sDσσγσT)Is+(σκTsγσT)Iσ],\displaystyle\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{I}=\frac{1}{\mathrm{Det}\,\hat{\Xi}}\left[\left(sD_{\sigma}-\sigma\frac{\gamma_{\sigma}}{T}\right)I_{s}+\left(\sigma\frac{\kappa}{T}-s\frac{\gamma_{\sigma}}{T}\right)I_{\sigma}\right], (32a)
x𝕋Ξ^1x=1DetΞ^[s2Dσ2sσγσT+σ2κT].\displaystyle\vec{x}^{\mathbb{T}}\hat{\Xi}^{-1}\vec{x}=\frac{1}{\mathrm{Det}\,\hat{\Xi}}\left[s^{2}D_{\sigma}-2s\sigma\frac{\gamma_{\sigma}}{T}+\sigma^{2}\frac{\kappa}{T}\right]. (32b)

Therefore

(A)Iσ0=IssDσσγσsTDσ+σ2κs2T2σγσsT.(A)_{I_{\sigma}\to 0}=\frac{I_{s}}{s}\frac{D_{\sigma}-\frac{\sigma\gamma_{\sigma}}{sT}}{D_{\sigma}+\frac{\sigma^{2}\kappa}{s^{2}T}-2\frac{\sigma\gamma_{\sigma}}{sT}}. (33)

The corresponding resistance is given by

RΣ\displaystyle R_{\Sigma} =1TIs2(TS˙)Iσ0\displaystyle=\frac{1}{TI^{2}_{s}}(T\dot{S})_{I_{\sigma}\to 0}
=ηπTs2(1r121r22)(DσσγσsTDσ+σ2κs2T2σγσsT)2\displaystyle=\frac{\eta}{\pi Ts^{2}}\left(\frac{1}{r^{2}_{1}}-\frac{1}{r^{2}_{2}}\right)\left(\frac{D_{\sigma}-\frac{\sigma\gamma_{\sigma}}{sT}}{D_{\sigma}+\frac{\sigma^{2}\kappa}{s^{2}T}-2\frac{\sigma\gamma_{\sigma}}{sT}}\right)^{2} (34)

After a trivial algebraic step, this expression can be reduced to the form of Eq. (21a) from the main text.

B.2 Spin diffusion contribution

The solution for the spatial distribution of the thermodynamic forces in the bulk of the Corbino device is given by [see Eq. (17) in the main text]:

X=ϰ2πrϰ𝕋Iϰ𝕋Ξ^ϰ.\vec{X}=-\frac{\vec{\varkappa}}{2\pi r}\frac{\vec{\varkappa}^{\mathbb{T}}\vec{I}}{\vec{\varkappa}^{\mathbb{T}}\hat{\Xi}\vec{\varkappa}}. (35)

In the entropy product rate we need an expression

X𝕋Ξ^X=B(2πr)2,B=(ϰ𝕋I)2ϰ𝕋Ξ^ϰ,\vec{X}^{\mathbb{T}}\hat{\Xi}\vec{X}=\frac{B}{(2\pi r)^{2}},\quad B=\frac{(\vec{\varkappa}^{\mathbb{T}}\vec{I})^{2}}{\vec{\varkappa}^{\mathbb{T}}\hat{\Xi}\vec{\varkappa}}, (36)

which gives us

TS˙=X𝕋Ξ^Xd2r=r1r2B(2πr)22πr𝑑r=B2πlnr2r1.T\dot{S}=\int\vec{X}^{\mathbb{T}}\hat{\Xi}\vec{X}d^{2}r=\int^{r_{2}}_{r_{1}}\frac{B}{(2\pi r)^{2}}2\pi rdr=\frac{B}{2\pi}\ln\frac{r_{2}}{r_{1}}. (37)

To obtain BB we calculate vector products

ϰ𝕋I=σIssIσ,\displaystyle\vec{\varkappa}^{\mathbb{T}}\vec{I}=\sigma I_{s}-sI_{\sigma}, (38a)
ϰ𝕋Ξ^ϰ=s2Dσ2sσγσT+σ2κT.\displaystyle\vec{\varkappa}^{\mathbb{T}}\hat{\Xi}\vec{\varkappa}=s^{2}D_{\sigma}-2s\sigma\frac{\gamma_{\sigma}}{T}+\sigma^{2}\frac{\kappa}{T}. (38b)

As a result, for the corresponding resistance one obtains

RΞ=(σ/s)2ln(r2/r1)2πT(1Dσ+σ2κs2T2σγσsT),R_{\Xi}=\frac{(\sigma/s)^{2}\ln(r_{2}/r_{1})}{2\pi T}\left(\frac{1}{D_{\sigma}+\frac{\sigma^{2}\kappa}{s^{2}T}-2\frac{\sigma\gamma_{\sigma}}{sT}}\right), (39)

where we used (B)Iσ0(B)_{I_{\sigma}\to 0}, that reduces to Eq. (21b) from the main text of the paper.

B.3 Disorder friction contribution

The dissipative power associated with the disorder-induced friction takes the form

TS˙=ku2d2r=kr1r2(A2πr)22πr𝑑r=kA22πlnr2r1T\dot{S}=k\int u^{2}d^{2}r=k\int^{r_{2}}_{r_{1}}\left(\frac{A}{2\pi r}\right)^{2}2\pi rdr=\frac{kA^{2}}{2\pi}\ln\frac{r_{2}}{r_{1}} (40)

The corresponding resistance is given by

Rk=1TIs2(TS˙)Iσ0=kln(r2/r1)2πTs2(DσσγσsTDσ+σ2κs2T2σγσsT)2,R_{k}=\frac{1}{TI^{2}_{s}}(T\dot{S})_{I_{\sigma}\to 0}=\frac{k\ln(r_{2}/r_{1})}{2\pi Ts^{2}}\left(\frac{D_{\sigma}-\frac{\sigma\gamma_{\sigma}}{sT}}{D_{\sigma}+\frac{\sigma^{2}\kappa}{s^{2}T}-2\frac{\sigma\gamma_{\sigma}}{sT}}\right)^{2}, (41)

where we used (A)Iσ0(A)_{I_{\sigma}\to 0} from the previous section, Eq. (33). The above expression can be easily reduced to the form of Eq. (21c).

Appendix C Thermal resistance

To elucidate the physical origin of the viscous contribution to the thermal resistance, it is instructive to derive it from an alternative consideration. To this end, observe one crucial point, that the obtained solution for the thermodynamic force X(r)\vec{X}(r), obeys the following relation x𝕋X=0\vec{x}^{\mathbb{T}}\vec{X}=0. In explicit notations it reads

sT+σμσ=0.s\bm{\nabla}T+\sigma\bm{\nabla}\mu_{\sigma}=0. (42)

It can be understood as the force balance condition exerted on the element of the fluid where local forces due to the temperature gradients are compensated by forces on electron spins arising from the local gradients of the spin chemical potential. In total, this relation reveals the global expulsion of the force density from the bulk of the device. Consider now the flow without spin polarization σ0\sigma\to 0. This means that T=0\bm{\nabla}T=0 so that the temperature of electron liquid is constant T(r)=TlT(r)=T_{l}. Yet, to induce the heat flow, we must apply the temperature bias ΔT=T1T2\Delta T=T_{1}-T_{2} across the entire device between inner and outer electrodes, which is not consistent with the constant temperature profile. This apparent paradox is resolved by the fact that there are temperature drops at the contacts. These drops create an excess pressure difference that is compensated by the radial viscous stresses arising in the liquid that exert additional radial force on the contacts. Recall the thermodynamic identity for the pressure differential, dP=ndμ+sdTdP=nd\mu+sdT Landau and Lifshitz (1980), where μ\mu is the chemical potential. At charge neutrality, n0n\to 0, the extra pressure is only due to temperature drops. Therefore, the pressure jumps at the boundary with the contacts are given by

s(TiTl)=Σrr(ri),s(T_{i}-T_{l})=\Sigma_{rr}(r_{i}), (43)

where TiT_{i} is the temperature of the ii-th contact. Note that the temperature of the liquid, TlT_{l} is either higher or lower than the temperatures of both leads, depending on the direction of the heat flow. This is a consequence of the bulk character of entropy production. It should be noted that a similar consequence of force expulsion occurs in charge transport away from charge neutrality. The voltage drop between the contacts and the electron liquid has the same sign at both boundaries.

Taking the difference in Eq. (43) we obtain

s(T1T2)=Σrr(r1)Σrr(r2)=ηAπ(1r121r22).s(T_{1}-T_{2})=\Sigma_{rr}(r_{1})-\Sigma_{rr}(r_{2})=-\frac{\eta A}{\pi}\left(\frac{1}{r^{2}_{1}}-\frac{1}{r^{2}_{2}}\right). (44)

where we used Eq. (29). Without spin polarization A=Is/sA=I_{s}/s, see Eq. (33), therefore

sΔT=ηIsπs(1r121r22).s\Delta T=-\frac{\eta I_{s}}{\pi s}\left(\frac{1}{r^{2}_{1}}-\frac{1}{r^{2}_{2}}\right). (45)

In the linear response to ΔT\Delta T, the heat current is given by Iq=GthΔTI_{q}=-G_{\text{th}}\Delta T, where GthG_{\text{th}} is thermal conductance. Using Iq=TIsI_{q}=TI_{s} one finds the corresponding resistance

RthGth1=ΔTTIs=ηπTs2(1r121r22).R_{\text{th}}\equiv G^{-1}_{\text{th}}=-\frac{\Delta T}{TI_{s}}=\frac{\eta}{\pi Ts^{2}}\left(\frac{1}{r^{2}_{1}}-\frac{1}{r^{2}_{2}}\right). (46)

It coincides with Eq. (21a) at σ0\sigma\to 0, which was obtained by matching dissipated power to Joule heating. This argument can be generalized to the case of finite spin polarization.

References

  • Abrikosov (1988) A. A. Abrikosov, Fundamentals of the Theory of Metals (Dover Publications, Inc., 1988).
  • Aronov (1976) A. G. Aronov, “Spin injection in metals and polarization of nuclei,” JETP Lett 24, 32 (1976).
  • Johnson and Silsbee (1987) Mark Johnson and R. H. Silsbee, “Thermodynamic analysis of interfacial transport and of the thermomagnetoelectric system,” Phys. Rev. B 35, 4959–4972 (1987).
  • Bauer et al. (2010) Gerrit E.W. Bauer, Allan H. MacDonald,  and Sadamichi Maekawa, “Spin caloritronics,” Solid State Communications 150, 459–460 (2010).
  • Bauer et al. (2012) Gerrit E. W. Bauer, Eiji Saitoh,  and Bart J. van Wees, “Spin caloritronics,” Nature Materials 11, 391–399 (2012).
  • Gurzhi et al. (2006) R. N. Gurzhi, A. N. Kalinenko, A. I. Kopeliovich, P. V. Pyshkin,  and A. V. Yanovsky, “Dynamics of a spin-polarized electron liquid: Spin oscillations with a low decay,” Phys. Rev. B 73, 153204 (2006).
  • Andreev et al. (2011) A. V. Andreev, Steven A. Kivelson,  and B. Spivak, “Hydrodynamic description of transport in strongly correlated electron systems,” Phys. Rev. Lett. 106, 256804 (2011).
  • Lucas and Fong (2018) Andrew Lucas and Kin Chung Fong, “Hydrodynamics of electrons in graphene,” Journal of Physics: Condensed Matter 30, 053001 (2018).
  • Levchenko and Schmalian (2020) Alex Levchenko and Jörg Schmalian, “Transport properties of strongly coupled electron–phonon liquids,” Annals of Physics 419, 168218 (2020).
  • Narozhny (2022) Boris N. Narozhny, “Hydrodynamic approach to two-dimensional electron systems,” La Rivista del Nuovo Cimento 45, 661–736 (2022).
  • Fritz and Scaffidi (2024) L. Fritz and T. Scaffidi, “Hydrodynamic electronic transport,” Annual Review of Condensed Matter Physics 15, 17–44 (2024).
  • Crossno et al. (2016) Jesse Crossno, Jing K. Shi, Ke Wang, Xiaomeng Liu, Achim Harzheim, Andrew Lucas, Subir Sachdev, Philip Kim, Takashi Taniguchi, Kenji Watanabe, Thomas A. Ohki,  and Kin Chung Fong, “Observation of the Dirac fluid and the breakdown of the Wiedemann-Franz law in graphene,” Science 351, 1058–1061 (2016).
  • Ghahari et al. (2016) Fereshte Ghahari, Hong-Yi Xie, Takashi Taniguchi, Kenji Watanabe, Matthew S. Foster,  and Philip Kim, “Enhanced thermoelectric power in graphene: Violation of the Mott relation by inelastic scattering,” Phys. Rev. Lett. 116, 136802 (2016).
  • Drögeler et al. (2014) Marc Drögeler, Frank Volmer, Maik Wolter, Bernat Terrés, Kenji Watanabe, Takashi Taniguchi, Gernot Güntherodt, Christoph Stampfer,  and Bernd Beschoten, “Nanosecond spin lifetimes in single- and few-layer graphene–hbn heterostructures at room temperature,” Nano LettersNano Letters 14, 6050–6055 (2014).
  • Drögeler et al. (2016) Marc Drögeler, Christopher Franzen, Frank Volmer, Tobias Pohlmann, Luca Banszerus, Maik Wolter, Kenji Watanabe, Takashi Taniguchi, Christoph Stampfer,  and Bernd Beschoten, “Spin lifetimes exceeding 12 ns in graphene nonlocal spin valve devices,” Nano LettersNano Letters 16, 3533–3539 (2016).
  • Onsager (1931) Lars Onsager, “Reciprocal relations in irreversible processes. I.” Phys. Rev. 37, 405–426 (1931).
  • Yan and Fuhrer (2010) Jun Yan and Michael S. Fuhrer, “Charge transport in dual gated bilayer graphene with Corbino geometry,” Nano LettersNano Letters 10, 4521–4525 (2010).
  • Faugeras et al. (2010) Clement Faugeras, Blaise Faugeras, Milan Orlita, M. Potemski, Rahul R. Nair,  and A. K. Geim, “Thermal conductivity of graphene in Corbino membrane geometry,” ACS NanoACS Nano 4, 1889–1892 (2010).
  • Waissman et al. (2024) Jonah Waissman, Artem V. Talanov, Aaron Hui, Zhongying Yan, Terry Phang, Kenji Watanabe, Takash Taniguchi, Brian J. Skinner,  and Philip Kim, “Observation of electronic viscous dissipation in graphene magneto-thermal transport,” in L100H (APS March Meeting, https://march.aps.org/sessions/M07/8, 2024).
  • Talanov et al. (2024) Artem Talanov, Jonah Waissman, Aaron Hui, Brian Skinner, Kenji Watanabe, Takashi Taniguchi,  and Philip Kim, “Observation of electronic viscous dissipation in graphene magneto-thermal transport,”  (2024), arXiv:2406.13799 [cond-mat.mes-hall] .
  • Zeng et al. (2024) Yihang Zeng, Haoyu Guo, Olivia M. Ghosh, Kenji Watanabe, Takashi Taniguchi, Leonid S. Levitov,  and Cory R. Dean, “Quantitative measurement of viscosity in two-dimensional electron fluids,”  (2024), arXiv:2407.05026 [cond-mat.mes-hall] .
  • Waissman et al. (2022) Jonah Waissman, Laurel E. Anderson, Artem V. Talanov, Zhongying Yan, Young J. Shin, Danial H. Najafabadi, Mehdi Rezaee, Xiaowen Feng, Daniel G. Nocera, Takashi Taniguchi, Kenji Watanabe, Brian Skinner, Konstantin A. Matveev,  and Philip Kim, “Electronic thermal transport measurement in low-dimensional materials with graphene non-local noise thermometry,” Nature Nanotechnology 17, 166–173 (2022).
  • Kumar et al. (2022) C. Kumar, J. Birkbeck, J. A. Sulpizio, D. Perello, T. Taniguchi, K. Watanabe, O. Reuven, T. Scaffidi, Ady Stern, A. K. Geim,  and S. Ilani, “Imaging hydrodynamic electrons flowing without Landauer–Sharvin resistance,” Nature 609, 276–281 (2022).
  • Krebs et al. (2023) Zachary J. Krebs, Wyatt A. Behn, Songci Li, Keenan J. Smith, Kenji Watanabe, Takashi Taniguchi, Alex Levchenko,  and Victor W. Brar, “Imaging the breaking of electrostatic dams in graphene for ballistic and viscous fluids,” Science 379, 671–676 (2023).
  • Völkl et al. (2024) T. Völkl, A. Aharon-Steinberg, T. Holder, E. Alpern, N. Banu, A. K. Pariari, Y. Myasoedov, M. E. Huber, M. Hücker,  and E. Zeldov, “Demonstration and imaging of cryogenic magneto-thermoelectric cooling in a van der Waals semimetal,” Nature Physics  (2024), 10.1038/s41567-024-02417-z.
  • Shavit et al. (2019) Michal Shavit, Andrey Shytov,  and Gregory Falkovich, “Freely flowing currents and electric field expulsion in viscous electronics,” Phys. Rev. Lett. 123, 026801 (2019).
  • Kavokin et al. (2020) A. V. Kavokin, B. L. Altshuler, S. G. Sharapov, P. S. Grigoryev,  and A. A. Varlamov, “The Nernst effect in Corbino geometry,” Proceedings of the National Academy of Sciences 117, 2846–2851 (2020).
  • Levchenko et al. (2022) Alex Levchenko, Songci Li,  and A. V. Andreev, “Hydrodynamic magnetoresistance in graphene Corbino devices,” Phys. Rev. B 106, L201306 (2022).
  • Gall et al. (2023) Vanessa Gall, Boris N. Narozhny,  and Igor V. Gornyi, “Corbino magnetoresistance in neutral graphene,” Phys. Rev. B 107, 235401 (2023).
  • Landau and Lifshitz (1987) L. D. Landau and E. M. Lifshitz, Fluid Mechanics, 2nd ed., Course of Theoretical Physics Series, Vol. 6 (Butterworth-Heinemann, Oxford, 1987).
  • Halperin and Hohenberg (1969) B. I. Halperin and P. C. Hohenberg, “Hydrodynamic theory of spin waves,” Phys. Rev. 188, 898–918 (1969).
  • Landau and Lifshitz (1980) L. D. Landau and E. M. Lifshitz, Statistical Physics, Part 1, 3rd ed., Course of Theoretical Physics Series, Vol. 5 (Butterworth-Heinemann, Oxford, 1980).
  • Abrikosov and Khalatnikov (1959) A. A. Abrikosov and I. M. Khalatnikov, “The theory of a fermi liquid (the properties of liquid 3He at low temperatures),” Rep. Prog. Phys. 22, 329 (1959).
  • Hone (1961) Daniel Hone, “Self-diffusion in liquid He3,” Phys. Rev. 121, 669–673 (1961).
  • Lyakhov and Mishchenko (2003) A. O. Lyakhov and E. G. Mishchenko, “Thermal conductivity of a two-dimensional electron gas with Coulomb interaction,” Phys. Rev. B 67, 041304 (2003).
  • Sykes and Brooker (1970) J. Sykes and G. A. Brooker, “The transport coefficients of a fermi liquid,” Annals of Physics 56, 1–39 (1970).
  • Pitaevskii and Lifshitz (1981) L. P. Pitaevskii and E. M. Lifshitz, Physical Kinetics, 1st ed., Course of Theoretical Physics Series, Vol. 10 (Butterworth-Heinemann, 1981).
  • Li et al. (2022) Songci Li, Alex Levchenko,  and A. V. Andreev, “Hydrodynamic thermoelectric transport in Corbino geometry,” Phys. Rev. B 105, 125302 (2022).
  • Müller et al. (2009) Markus Müller, Jörg Schmalian,  and Lars Fritz, “Graphene: A nearly perfect fluid,” Phys. Rev. Lett. 103, 025301 (2009).
  • Li et al. (2020) Songci Li, Alex Levchenko,  and A. V. Andreev, “Hydrodynamic electron transport near charge neutrality,” Phys. Rev. B 102, 075305 (2020).
  • Mishchenko (2008) E. G. Mishchenko, “Minimal conductivity in graphene: Interaction corrections and ultraviolet anomaly,” Europhysics Letters 83, 17005 (2008).
  • Fritz et al. (2008) Lars Fritz, Jörg Schmalian, Markus Müller,  and Subir Sachdev, “Quantum critical transport in clean graphene,” Phys. Rev. B 78, 085416 (2008).
  • Kashuba (2008) Alexander B. Kashuba, “Conductivity of defectless graphene,” Phys. Rev. B 78, 085415 (2008).
  • Kovtun et al. (2005) P. K. Kovtun, D. T. Son,  and A. O. Starinets, “Viscosity in strongly interacting quantum field theories from black hole physics,” Phys. Rev. Lett. 94, 111601 (2005).
  • Weber et al. (2005) C. P. Weber, N. Gedik, J. E. Moore, J. Orenstein, J. Stephens,  and D. D. Awschalom, “Observation of spin Coulomb drag in a two-dimensional electron gas,” Nature 437, 1330–1333 (2005).