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

institutetext: Instituto de Física La Plata IFLP-CONICET & Departamento de Física, FCE-UNLP
C.C. 67, 1900 La Plata, Argentina

Pomeranchuk instabilities in holographic metals

Gastón Giordano    Nicolás Grandi    Adrián Lugo gaston.giordano@fisica.unlp.edu.ar grandi@fisica.unlp.edu.ar lugo@fisica.unlp.edu.ar
Abstract

We develop a method to detect instabilities leading to nematic phases in strongly coupled metallic systems. We do so by adapting the well-known Pomeranchuk technique to a weakly coupled system of fermions in a curved asymptotically AdS bulk. The resulting unstable modes are interpreted as corresponding to instabilities on the dual strongly coupled holographic metal. We apply our technique to a relativistic 3+13+1-dimensional bulk with generic quartic fermionic couplings, and explore the phase diagram at zero temperature for finite values of the fermion mass and chemical potential, varying the couplings. We find a wide region of parameters where the system is stable, which is simply connected and localized around the origin of coupling space.

1 Introduction

Strongly correlated electron systems have been at the heart of most recent research in condensed matter theory sachdev_2000 . The underlying strongly coupled dynamics is thought to be responsible of the richness of the phase diagram of high TcT_{c} superconductors. It contains normal metallic regions, as well as an exotic phase known as “strange metal” or “non-Fermi liquid” RevModPhys.73.797 which is believed to be described by strongly coupled fermionic degrees of freedom. There are also regions in which rotational symmetry is broken, the so-called “nematic” phases, as well as inhomogeneous “smectic” and “chessboard” phases.

The transition from an isotropic Fermi liquid to a nematic phase is believed to be driven by a Pomeranchuk instability pomeranchuk . Such instability arises when an excitation of the ground state of the Fermi liquid results in a net decrease of the total energy. In Landau theory, such perturbation is represented by a deformation of the Fermi surface. By decomposing the deformation onto an orthonormal basis, Pomeranchuk obtained a set of conditions under which the Fermi liquid is stable. This method can be generalized to lattice systems or anisotropic Fermi surfaces PhysRevB.78.115104 ; PhysRevB.80.075108 and to finite temperature and magnetic field uno ; dos ; RodriguezPonte2013 , and it relies on the weakness of the quasi-particle coupling, or in other words on the validity of the Landau formula. This implies that from the strange metal perspective, since the dynamics is strongly coupled, the detection of fermionic instabilities becomes more difficult.

The holographic description AHARONY2000183 of a homogeneous fermionic phase has been shown to account for some of the interesting properties of the strange metal Lee:2008xf ; Liu:2009dm ; Faulkner:2009wj ; Faulkner:2010da ; Hartnoll:2010gu ; Hartnoll:2011dm . In particular, the resulting spectral function is compatible with a Fermi surface with or without long-lived quasi-particles. Based on that, in this paper, we analyze Pomeranchuk instabilities of the strange metal phase from the holographic perspective. We use a holographic background in which we propagate a Dirac spinor. Being weakly coupled, such spinor can be described by Landau’s theory, and its stability under Fermi surface deformations can be studied. This accounts for a description of the anisotropic instabilities of the dual strange metal.

The generalization of Pomeranchuk method to arbitrary curved spaces is not possible, since it relies on a momentum space representation of the fermionic system. Nevertheless, the special kind of “planar” bulk spacetimes used in holography have the additional feature of a translational symmetry in the spatial directions spanning the boundary. This allows for a labeling of the bulk states by a momentum index k\vec{k}, complemented by an additional index mm labeling the oscillation mode in the holographic direction. Then the D+1D+1 dimensional curved space fermionic system can be interpreted as a (D1)+1(D-1)+1 dimensional flat space system with multiple fermion species labeled by mm. Pomeranchuk method can then be straightforwardly applied to such multi-fermion flat space weakly coupled system.

In what follows, we sketch the necessary steps needed to go from a 3+13+1 dimensional action for spinors in AdS spacetime to a 2+12+1 dimensional Hamiltonian for a multi-fermion system in flat space. Then we construct the corresponding Landau theory, and apply the Pomeranchuk method to it. To improve readability, only the relevant steps of the calculations are shown in the bulk of the paper, leaving the details to the appendices 111Notice that Pomeranchuk instabilities were studied in the holographic context in Edalati:2012eh from a different perspective, focusing in the distortions of spectral functions. See also Liu:2014mva for a related issue..

2 Holographic setup

We consider a holographic background consisting of a metric and an electromagnetic field with the generic “planar” form

G\displaystyle G =\displaystyle= L2(fdt2+gdz2+dx2+dy2z2),A=hdt.\displaystyle L^{2}\left(-f\,dt^{2}+g\,dz^{2}+\frac{dx^{2}+dy^{2}}{z^{2}}\right)\,,\qquad\qquad A=h\,dt\,. (1)

This is a solution of the Einstein-Maxwell equations with a negative cosmological constant, as well as possible additional matter contributions to the energy-momentum tensor. The metric components f(z)f(z) and g(z)g(z) are functions of the holographic coordinate zz, and the geometry asymptotes AdS spacetime as zz goes to zero, provided f(z)g(z)1/z2f(z)\sim g(z)\sim 1/z^{2}. The electric potential h(z)h(z) is also a function of zz, and approaches a constant h(z)μh(z)\sim\mu at the boundary, which is identified with the chemical potential on the holographic theory.

Specific backgrounds with the form (1) are the Reisner-Nordstrom AdS black-hole Lee:2008xf ; Liu:2009dm ; Faulkner:2010da ; Faulkner:2009wj , the holographic superconductor Hartnoll_2008 , and the electron star at zero Hartnoll:2010gu ; Hartnoll:2011dm ; Cubrovic:2011xm and finite Puletti:2010de ; Hartnoll:2010ik temperatures. Although we keep our formalism as general as possible, in order to get concrete results in Section 4 we apply it to the background of reference Hartnoll:2010gu .


In the background defined above we propagate a spinorial perturbation Ψ\Psi, whose dynamics is dictated by the free Dirac action

S𝖿𝗋𝖾𝖾=d4x|G|Ψ¯(/𝒟𝗆)Ψ,S_{\sf free}=-\int d^{4}x\,\sqrt{|G|}\;\bar{\Psi}\left(/\!\!\!\!{\cal D}-{\sf m}\right)\Psi\,, (2)

where /𝒟/\!\!\!\!{\cal D} stands for the covariant derivative containing both curved spacetime and gauge contributions, contracted with the curved space Dirac matrices. Since the energy-momentum tensor and the electric current are quadratic in the spinor perturbation, to linear order in Ψ\Psi we can work in the probe limit in which the background (1) is not perturbed.

The general solution to (2) can be decomposed as

Ψ(t,x,z)\displaystyle\Psi(t,\vec{x},z) =\displaystyle= zf(z)14αmk1𝒩αmkcαmk(t)eikxψαmk(z),\displaystyle\frac{z}{f(z)^{\frac{1}{4}}}\,\sum_{\alpha m\vec{k}}\frac{1}{{\cal N}_{\alpha m\vec{k}}}\;\,c_{\alpha m\vec{k}}(t)\;\,e^{i\vec{k}\cdot\vec{x}}\;\psi_{\alpha m\vec{k}}(z)\,, (3)

in terms of time dependent coefficients cαmk(t)c_{\alpha m\vec{k}}(t) and zz-dependent spinors ψαmk(z)\psi_{\alpha m\vec{k}}(z). Here the label α=1,2\alpha=1,2 is a spin index, and k2\vec{k}\in\mathbb{R}^{2} represents the momentum in the xyxy plane, while mm\in\mathbb{N} characterizes the number of oscillations in the zz direction. Finally 𝒩αmk{\cal N}_{\alpha m\vec{k}} is a normalization constant and the factor z/f1/4(z)z/f^{1/4}(z) was introduced for calculational convenience. The spinors ψαmk(z)\psi_{\alpha m\vec{k}}(z) are written in terms of the solution to an ordinary differential equation in the variable zz, with a Schrödinger-like form. When physically meaningful boundary conditions are imposed at the extremes of the zz range, a (possibly complex) quantized dispersion relation is obtained ωm(k)\omega_{m}(k), with which we can write cαmk(t)exp(iωm(k)t)c_{\alpha m\vec{k}}(t)\propto\exp(-i\omega_{m}(k)t).

Expression (3) allows us to quantize the system by promoting the coefficients cαmk(t)c_{\alpha m\vec{k}}(t) to operators cαmkc_{\alpha m\vec{k}} and cαmkc^{\dagger}_{\alpha m\vec{k}} satisfying fermionic anticommutation relations. Then cαmkc_{\alpha m\vec{k}}^{\dagger} creates a fermionic perturbation with momentum k\vec{k}, spin α\alpha and mode index mm, and cαmkc_{\alpha m\vec{k}} annihilates it.


The steps of the derivation leading to decomposition (3) as well as the explicit form of the components of the spinor ψαmk(z)\psi_{\alpha mk}(z) are reviewed in detail in Appendix A.

3 Interactions

Now we introduce interactions among the spinorial perturbations, which represent 1/N1/N corrections to the holographic description. We do so by supplementing the action with the additional term

S𝗂𝗇𝗍=d4x|G|Tσσσ¯σ¯Ψ¯σ¯Ψ¯σ¯ΨσΨσ,S_{\sf int}=\int d^{4}x\,\sqrt{|G|}\;T^{\bar{\sigma}\bar{\sigma}^{\prime}}_{\sigma\sigma^{\prime}}\;\bar{\Psi}_{\bar{\sigma}}\;\bar{\Psi}_{\bar{\sigma}^{\prime}}\;\Psi^{\sigma}\;\Psi^{\sigma^{\prime}}\,, (4)

where we made explicit the spinor indices σ{1,2,3,4}\sigma\in\{1,2,3,4\}, and the Lorentz invariant tensor Tσσσ¯σ¯T^{\bar{\sigma}\bar{\sigma}^{\prime}}_{\sigma\sigma^{\prime}} represents the most general four fermion covariant interaction in curved space PhysRevB.59.7140 .

Tσ3σ4σ1σ2\displaystyle T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}} =\displaystyle= g1δσ1δσ2σ3+σ4g2(Γ5)σ1(Γ5)σ2σ3+σ4g3(Γa)σ1(Γa)σ2σ3σ4\displaystyle g_{1}\;\delta^{\sigma_{1}}{}_{\sigma_{3}}\;\delta^{\sigma_{2}}{}_{\sigma_{4}}+g_{2}\;(\Gamma^{5})^{\sigma_{1}}{}_{\sigma_{3}}\;(\Gamma^{5})^{\sigma_{2}}{}_{\sigma_{4}}+g_{3}\;(\Gamma^{a})^{\sigma_{1}}{}_{\sigma_{3}}\;(\Gamma_{a})^{\sigma_{2}}{}_{\sigma_{4}} (5)
+\displaystyle+ g4(ΓaΓ5)σ1(ΓaΓ5)σ2σ3σ4g54([Γa,Γb]Γ5)σ1([Γa,Γb]Γ5)σ2σ3.σ4\displaystyle g_{4}\;(\Gamma^{a}\,\Gamma^{5})^{\sigma_{1}}{}_{\sigma_{3}}\;(\Gamma_{a}\,\Gamma^{5})^{\sigma_{2}}{}_{\sigma_{4}}-\frac{g_{5}}{4}\;\left([\Gamma^{a},\Gamma^{b}]\,\Gamma^{5}\right)^{\sigma_{1}}{}_{\sigma_{3}}\;\left([\Gamma_{a},\Gamma_{b}]\,\Gamma^{5}\right)^{\sigma_{2}}{}_{\sigma_{4}}\,. (6)

We see that it is written completely in terms of Dirac matrices, and depends linearly on a set of five coupling constants g1,,g5g_{1},\dots,g_{5}, which we assume are small.

We can now obtain the Hamiltonian resulting from (2) and (4), it takes the form

H\displaystyle H =\displaystyle= αmd2kωm(k)cαmkcαmk\displaystyle\sum_{\alpha m}\!\int\!d^{2}\!k\;\omega_{m}(k)\,c_{\alpha m\vec{k}}^{\dagger}c_{\alpha m\vec{k}}
+\displaystyle+ α1α2α3α4m1m2m3m4d2kd2kd2qtα3m3k;α4m4kα1m1(k+q);α2m2(kq)cα1m1(k+q)cα2m2(kq)cα3m3kcα4m4k.\displaystyle\sum_{\underset{m_{1}m_{2}m_{3}m_{4}}{\mbox{\tiny$\alpha_{1}\alpha_{2}\alpha_{3}\alpha_{4}$\normalsize}}}\!\!\!\int d^{2}\!k\,d^{2}\!k^{\prime}d^{2}\!q\;\,t^{\alpha_{1}m_{1}({\vec{k}}\!+\!\vec{q});\alpha_{2}m_{2}({\vec{k}}^{\prime}\!-\!\vec{q})}_{\alpha_{3}m_{3}{\vec{k}};\alpha_{4}m_{4}{\vec{k}}^{\prime}}\;c_{\alpha_{1}m_{1}(\vec{k}+\vec{q})}^{\dagger}\;c_{\alpha_{2}m_{2}(\vec{k}^{\prime}-\vec{q})}^{\dagger}c_{\alpha_{3}m_{3}{\vec{k}}}\;c_{\alpha_{4}m_{4}{\vec{k}}^{\prime}}\,.

Here the frequencies ωm(k)\omega_{m}(k) play the role of dispersion relations for each fermionic mode. On the other hand the new tensor tα3m3k3;α4m4k4α1m1k1;α2m2k2t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}} contains the information about the interaction strengths among different modes, and it results from contracting the position space interaction tensor Tσσσ¯σ¯T^{\bar{\sigma}\bar{\sigma}^{\prime}}_{\sigma\sigma^{\prime}} with integrals in the zz direction of quartic products of the free fermionic eigenstates ψαmk(z)\psi_{\alpha m\vec{k}}(z).


The explicit forms of the interaction tensors Tσσσ¯σ¯T^{\bar{\sigma}\bar{\sigma}^{\prime}}_{\sigma\sigma^{\prime}} and tα3m3k3;α4m4k4α1m1k1;α2m2k2t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}}. as well as that of the aforementioned quartic integrals, is not relevant for the moment. For the details of the derivation of (3) we refer the reader to Appendix B.


With equation (3), we have succeded in re-writing the bulk dynamics as that of a second quantized Hamiltonian for a two dimensional multi-fermion system in which the index mm denotes fermion species. Since the coupling in the bulk is assumed to be weak, we can safely rely on the Landau description of the Fermi liquid.

Assuming that the ground state of the Hamiltonian (3) is characterized by a certain set of occupation numbers NαmkN_{\alpha m\vec{k}}, then the excitations can be described by their variations δNαmk\delta N_{\alpha m\vec{k}}. The grand canonical energy of an excitation is then written as the Landau formula

δΩ(T,μ)=d2kαmϵm(k)δNαmk+12d2kd2kαmαmfαmαm(k,k)δNαmkδNαmk,\delta\Omega(T,\mu)\!=\!\int d^{2}\vec{k}\,\sum_{\alpha m}\,\epsilon_{m}(k)\delta N_{\alpha m\vec{k}}+\frac{1}{2}\int\!d^{2}\vec{k}\,d^{2}\vec{k}^{\prime}\!\!\sum_{\alpha m\alpha^{\prime}m^{\prime}}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\,\delta N_{\alpha m\vec{k}}\,\delta N_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\,, (8)

where fαmα¯m¯(k,k¯)f_{\alpha m\bar{\alpha}\bar{m}}(\vec{k},\vec{\bar{k}}) is the so-called “interaction function” which is obtained from the tensor tα3m3k3;α4m4k4α1m1k1;α2m2k2t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}}, and the quasiparticle dispersion relation ϵm(k)\epsilon_{m}(k) takes the form

ϵm(k)=ωm(k)+αmd2kfαmαm(k,k)Nαmk.\epsilon_{m}(k)=\omega_{m}(k)+\sum_{\alpha^{\prime}m^{\prime}}\int\!d^{2}\vec{k}^{\prime}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\,N_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\,. (9)

For future use, notice that the quasiparticle dispersion relation equals the frequency plus corrections of first order in the coupling constants.

For the most general covariant quartic perturbation (4), the interaction function takes a particularly simple angular dependence in the momentum plane. Indeed, if we write k=k(cosθ,sinθ)\vec{k}=k\,(\cos\theta,\sin\theta) then it can be decomposed in only three Fourier modes, as

fαm;αm(k,k)=fαmk;αmk0+cos(θθ)fαmk;αmkc+sin(θθ)fαmk;αmks,f_{\alpha m;\alpha^{\prime}m^{\prime}}(\vec{k},{\vec{k}}^{\prime})=f_{\alpha mk;\alpha^{\prime}m^{\prime}k^{\prime}}^{0}+\cos(\theta-\theta^{\prime})\;f_{\alpha mk;\alpha^{\prime}m^{\prime}k^{\prime}}^{c}+\sin(\theta-\theta^{\prime})\;f_{\alpha mk;\alpha^{\prime}m^{\prime}{k}^{\prime}}^{s}\,, (10)

where the constant, sine and cosine coefficients depend on a reduced subset of the components of the tensor tα3m3k3;α4m4k4α1m1k1;α2m2k2t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}}, or in other words on the position space interaction tensor Tσσσ¯σ¯T^{\bar{\sigma}\bar{\sigma}^{\prime}}_{\sigma\sigma^{\prime}} contracted with integrals of the fermionic states ψαmk(z)\psi_{\alpha m\vec{k}}(z).


The details of the construction of the Landau description starting with the underlying Hamiltonian (3), as well as the explicit form of the interaction function (10) in terms of the interaction tensor and wavefuntion integrals, are presented in Appendix C.


We can now make use of the Pomeranchuk technique for the above defined two dimensional multicomponent Landau Fermi liquid, in order to diagnose instabilities arising from an anisotropic deformation of its Fermi surface.

We start by decomposing the deformation on the occupation numbers, that charactherize the excitations of the Fermi liquid, in the form

δNαmk=𝖧(ϵm(k)+δgαm(k))𝖧(ϵm(k))=δ(ϵm(k))δgαm(k)+12δ(ϵm(k))δgαm(k)2+\delta N_{\alpha m\vec{k}}={\sf H}\big{(}\!-\!\epsilon_{m}(k)+\delta g_{\alpha m}(\vec{k})\big{)}-{\sf H}\left(\!-\epsilon_{m}(k)\right)=\delta\big{(}\!-\!\epsilon_{m}(k)\big{)}\;\delta g_{\alpha m}(\vec{k})+\frac{1}{2}\;\delta^{\prime}\big{(}\!-\epsilon_{m}(k)\big{)}\;\delta g_{\alpha m}(\vec{k})^{2}+\dots (11)

where 𝖧(){\sf H}(\cdot) is the Heavyside unit-step function, and δgαm(k)\delta g_{\alpha m}(\vec{k}) are arbitrary functions of the momentum characterizing the excitation. When plugged back into (8) the delta functions force the evaluation of the expression on the Fermi momentum kFmk_{F}^{m} defined by ϵm(kFm)=0\epsilon_{m}(k_{F}^{m})=0. The remaining angular integrals result on a grand canonical energy that is a quadratic form in the Fourier components δgαmn(c,s)\delta g_{\alpha m}{}_{n}^{(c,s)} of the parameters δgαm(k)|k=kFm=n=0(δgαmcosnc(nθ)+δgαmsinns(nθ))\delta g_{\alpha m}(\vec{k})|_{k=k_{F}^{m}}=\sum_{n=0}^{\infty}\left(\delta g_{\alpha m}{}_{n}^{c}\;\cos(n\theta)+\delta g_{\alpha m}{}_{n}^{s}\;\sin(n\theta)\right). Notice that, since the interaction function (10) has only three Fourier components, the only modes that depend on the couplings (and hence could lead to negative minors in the quadratic form) are δgαm0c\delta g_{\alpha m}{}_{0}^{c} and δgαm1(c,s)\delta g_{\alpha m}{}_{1}^{(c,s)}. The first one is isotropic, while the remaining two completely break the rotational symmetry.

If the grand canonical energy of an excitation is negative, the system decreases its energy by creating more such excitations. To avoid such instability, we need to impose that the aforementioned quadratic form is positive definite. This is guaranteed provided all the minors of the quadratic kernel are positive, or in other words

|kFmvFm(δαmαm+2πkFmvFmfαmαm0)|M×M\displaystyle\left|\frac{k_{F}^{m}}{v_{F}^{m}}\left(\delta_{\alpha m\alpha^{\prime}m^{\prime}}+2\,\pi\,\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}\;f^{0}_{\alpha m\alpha^{\prime}m^{\prime}}\right)\right|_{M\times M} >\displaystyle> 0,M\displaystyle 0\quad,\quad\forall M\in\mathbb{N} (12)
|kFmvFm(δαmαm+πkFmvFm(fαmαmc+ifαmαms))|M×M\displaystyle\left|\frac{k_{F}^{m}}{v_{F}^{m}}\left(\delta_{\alpha m\alpha^{\prime}m^{\prime}}+\pi\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}\,\left({f^{c}_{\alpha m\alpha^{\prime}m^{\prime}}}+i{f^{s}_{\alpha m\alpha^{\prime}m^{\prime}}}\right)\right)\,\right|_{M\times M} >\displaystyle> 0,M\displaystyle 0\quad,\quad\forall M\in\mathbb{N} (13)

where vFm=dϵm(k)/dk|k=kFmv_{F}^{m}=d\epsilon_{m}(k)/dk|_{k=k_{F}^{m}} are the Fermi velocities and fαmαm(0,c,s)f^{(0,c,s)}_{\alpha m\alpha^{\prime}m^{\prime}} are the Landau parameters, given by the Fourier components in (10) evaluated at the corresponding Fermi momenta. Notice that, provided the Fermi velocities are positive, the quotient kFm/vFm{k_{F}^{m}}/{v_{F}^{m}} can be removed from the prefactors. Moreover, we can perturbatively replace kFm/vFm{k_{F}^{m}}/{v_{F}^{m}} in the parenthesis by its free version, defined by ωm(kfm𝖿𝗋𝖾𝖾)=0\omega_{m}(k_{f}^{m\,{\sf free}})=0 and vFm𝖿𝗋𝖾𝖾=dωm(k)/dk|k=kFm𝖿𝗋𝖾𝖾v_{F}^{m\,{\sf free}}=d\omega_{m}(k)/dk|_{k=k_{F}^{m\,{\sf free}}}.

When some of the minors in (12) is negative, a linear combination of the modes δgαm0c\delta g_{\alpha m}{}_{0}^{c} causes a net decrease in the energy, resulting in an instability. As mentioned above, such modes are isotropic and therefore the corresponding instability would be a Stoner one, i.e. an instability with respect to adding or removing particles from the ground state. On the other hand, if the negative minor shows up in (13), the unstable direction correspond to a linear combination of the modes δgαm1(s,c)\delta g_{\alpha m}{}_{1}^{(s,c)}, and the resulting deformation completely breaks rotational invariance. It is tempting to imply that such symmetry breaking pattern is inherited by the phase that stabilizes the system after the phase transition.

A review of Pomeranchuk method, including the details of the calculations resulting in formulas (12)-(13), can be seen in Appendix D.

4 Summary of the method

In summary, in order to test the anisotropic Pomeranchuk instabilities of an holographic Fermi liquid, we need to

  1. 1.

    Obtain the fermionic modes ψαmk(z)\psi_{\alpha mk}(z) in the expansion (3), that we read from (58) and (64)), and their frequencies ωm(k)\omega_{m}(k). This is done by solving a Schrödinger-like equation with potential (44) in the holographic direction for a function ϕmk(z)\phi_{mk}(z). Notice that we only need the modes with very small frequency, since we will evaluate them at the free Fermi momentum kFm𝖿𝗋𝖾𝖾k_{F}^{m\,{\sf free}} where the frequency vanishes.

  2. 2.

    Calculate the free Fermi velocities as vFm𝖿𝗋𝖾𝖾=dωFm/dk|k=kFm𝖿𝗋𝖾𝖾v_{F}^{m\,{\sf free}}=d\omega_{F}^{m}/dk|_{k=k_{F}^{m\,{\sf free}}}. Incidentally, this is why we need modes with non-vanishing small frequency, even if just those with vanishing frequency contribute to the quantities evaluated at the free Fermi momenta.

  3. 3.

    Integrate a particular set of quartic products of fermionic modes in the holographic direction given in (LABEL:eq:fermion.integrals.all), in order to go from the coupling tensor in spin indices Tσσσ¯σ¯T^{\bar{\sigma}\bar{\sigma}^{\prime}}_{\sigma\sigma^{\prime}} given by to tα3m3k3;α4m4k4α1m1k1;α2m2k2t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}} and then to the Landau parameters fαmαm(c,s)f^{(c,s)}_{\alpha m\alpha^{\prime}m^{\prime}}.

  4. 4.

    Check whether all the minors in (12)-(13) are positive. Notice that each minor of M×MM\times M has degree MM on the interaction function, which is linear in the coupling constants. This implies that each condition imposes a polynomial restriction of order MM in coupling space.

At any point in coupling space at which condition 4 is not satisfied, the quadratic form has a negative eigenmode, implying that the grand canonical energy is decreased by the corresponding excitation. This results in an instability of the system. If the negative minor appears in (12), the resulting instability is isotropic, while if it shows up in (13), it breaks completely the rotational symmetry.

5 Application to the electron star

We applied the above steps to the specific example of the zero temperature electron star background of Hartnoll:2010gu . The electron star represents the holographic dual to the ground state of a highly degenerate system of fermions at zero temperature. It is constructed by coupling the gravitational and electromagnetic degrees of freedom in the bulk to a perfect fluid representing the fermions, whose equation of state is obtained in the Thomas-Fermi approximation. A summary of the electron star solution is given in Appendix E.

In order to obtain the fermionic modes and calculate the Fermi momenta and velocities as required by points 1 and 2, we used a Wentzel-Kramers-Brillouin (WKB) approximation on the spinorial perturbation. This is consistent with the use of the Thomas-Fermi approximation for the bulk fluid. The relevant details of the WKB method are presented in Appendix F.

We explored the phase diagram by turning on the coupling constants g1,,g5g_{1},\dots,g_{5} by pairs. Sitting at different values of the non-vanishing couplings on every coupling plane, we evaluated the minors in equations (12) and (13) order by order. For example, on the plane g2g_{2}-g3g_{3}, we obtained the plots at the top of Fig.1. The regions are painted with a color that depends on the order MM of the first unstable minor. The plot on the top left correspond to instabilities that preserve rotational invariance, while that on the top right denotes instabilities that break it completely. These plots were combined on the one at the bottom. The resulting unstable regions of the phase diagram can be seen in Fig. 2, in which all different coupling planes are presented. We see that as MM grows (and so does the degree of the polynomial in point 4 above) the region in coupling space where the system is stable gets smaller, converging to an isolated wide island around the origin, which is simply connected.


In order to interpret our results from the boundary perspective, we recall that the electron star describes a holographic strongly coupled theory whose ground state has a fermionic condensate Hartnoll:2010gu . Classical spinor perturbations on the bulk then correspond to fermionic excitations on the boundary theory Hartnoll:2011dm .

The interesting point is that, as we decompose the bulk spinor into zz-modes, the set of Fermi momenta we obtained for them corresponds exactly to those defining the “holographically smeared Fermi surface” in reference Hartnoll:2011dm . It is natural then to interpret collective bulk modes, made by deforming the Fermi surface of each bulk zz mode, as corresponding to collective excited states on the boundary theory. These “deformations of the smeared Fermi surface” become unstable when a bulk Pomeranchuk instability shows up, leading to an instability of the boundary state that could easily be interpreted as a boundary Pomeranchuk instability.

Refer to caption
Refer to caption

​​​​​​​​​​​​​​​​​​​​​​​ ​​​​​​​​​​​​​​​​​​​​​​​ ​​​​​​​​​​​​​​​​​​​​​​​ Refer to caption   Refer to caption

Figure 1: Construction of the unstable regions in the coupling plane g2g_{2}-g3g_{3}. The plots on the top correspond to the unstable regions in equation (12) (left) and (13) (right). The plot on the bottom combines both equations. The colors denote the order MM of the first negative minor.
Refer to caption

​​​​ Refer to captionRefer to caption
 
Refer to caption ​​​​ Refer to captionRefer to caption
 
Refer to caption ​​​​​ Refer to caption ​​​ Refer to caption
 
Refer to caption  Refer to caption

Figure 2: Phase diagrams showing the unstable regions in the five-dimensional coupling space g1g_{1}, g2,g3,g4,g5g_{2},g_{3},g_{4},g_{5}. Each plot corresponds to one of the ten two-dimensional coupling planes. We see that there is an island of stability around the origin. The lighter regions become unstable at higher MM. The plots correspond to m^=0.4\hat{m}=0.4, λ=2\lambda=2 and γ=100\gamma=100 (see Appendix E for details on these parameters).

6 Conclusions and outlook

We developed a general method to study anisotropic fermionic instabilities on a strongly coupled Fermi liquid via the AdS/CFT duality. It entails to perform a Pomeranchuk analysis on the dual bulk fermion, adapting the formalism to a curved space-time with a planar slice. The key step is to rewrite the single fermion in the bulk in terms of its modes in the holographic direction, resulting on a multi-component systems of lower dimensional fermions which move in the planar slice. By analyzing the minors of a quadratic form, we are able to detect when the fermionic modes become unstable under anisotropic deformations of their Fermi surface. The resulting phase transition breaks the rotational symmetry in the planar slice, and consequently in the boundary theory, potentially leading to a nematic phase.

We kept the formalism as general as possible, being suitable to be applied to many Euclidean-invariant boundary theories. This includes the zero temperature electron star Hartnoll:2010gu ; Hartnoll:2011dm , the finite temperature electron star Puletti:2010de ; Hartnoll:2010ik , as well as other examples222Since the method relays on the existence of a bulk Fermi surface, to adapt it to backgrounds such like the holographic metal Lee:2008xf ; Liu:2009dm ; Faulkner:2010da ; Faulkner:2009wj , the holographic superconductor Hartnoll_2008 , the magnetically charged black-hole Albash_2008 , or the Lifshitz geometry Kachru:2008yh , one would need to work in the limit in which a bulk fermion condensate would not backreact. This limit can be subtle, and it is hard to reconcile with the validity conditions for the WKB method, so the fermion wave functions should be solved using a different approximation.. We applied the method to the zero-temperature electron star background, being able to identify the unstable region on a five dimensional coupling space.

The method can be extended to more general situations relevant to condensed matter systems. Some examples of possible future directions are:

  • Finite doping: the inclusion of a doping axis should be straitforward, following the lines described in Kiritsis2016 .

  • Finite magnetic field: the inclusion of a magnetic field would require a magnetically charged background Albash_2008 333See footnote 2., and it would modify the bulk fermion coupling. Also, the Pomeranchuk analysis needs to be adapted dos .

  • Finite temperature fluid: as it stands, the method can be applied to the background of Puletti:2010de ; Hartnoll:2010ik in which the temperature is included via the presence of a horizon, while the bulk fluid is approximated as a finite temperature Fermi liquid. Including the effect of temperature in the bulk matter would modify the background, as well as the Pomeranchuk analysis uno .

  • Lifshitz scaling: considering Lifshitz geometries Kachru:2008yh would require a rewriting of the asymptotic conditions on the near boundary fermion, but the rest of the analysis remains mostly unchanged3{}^{\ref{foot2}}.

  • Lattice fermions: describing an underlying lattice in the holographic setup would imply the inclusion of a lattice-like metric Horowitz2012 or momentum relaxation Andrade2014 , and a modification of the Pomeranchuk analysis according to the lines of PhysRevB.78.115104 .

These are only some of the situations to which the method can be extended, we think that they deserve further investigation.

As a last remark, one could wonder whether Pomeranchuk instabilities are evident in physical quantities as we approach the edge of the unstable region444We would like to thank the referee for rising this point.. In a weakly coupled Fermi liquid, standard Landau theory arguments show that Landau parameters appear into some response functions. This happens for example for the compressibility or the spin susceptibility. The expressions are such that these functions become infinite when the instability is triggered by the correct Landau parameter.

Recalling that, according to Kubo formulae, response functions are written in terms of two-point correlators of local quantities, we can calculate them in holography. This is done by perturbing the background fields with the right IR boundary conditions and then evaluating the quotient of the leading and subleading parts of the perturbations at the UV. However, it is evident that in such calculation the fermionic couplings g1,,g5g_{1},\dots,g_{5} would only enter at higher loops. Since there is no reason to assume that bulk perturbation theory would break down close to the edge of the unstable region, the natural conclusion is that, in the strongly coupled boundary theory, the response functions are not sensitive to the instability.

Aknowledgements

The authors thank Pablo Rodríguez Ponte and Ignacio Salazar Landea for relevant contributions during the early stages of this work. We also thank Carlos Lamas for helpful comments. This work has been funded by the CONICET grants PIP-2017-1109, PIP-2015-0688 and PUE 084 “Búsqueda de Nueva Física”, and UNLP grants PID-X791, PID-11/X910.

Appendix A Fermionic states in a holographic background

In this Appendix, we work out a basis of free fermionic modes in a charged planar holographic background. We do so by writing the background in terms of vielbein and spin connection (section A.1), deriving the form of the corresponding Dirac equation (section A.2), separating it into spin components and Fourier modes in the planar and time directions (section A.2.1), and obtaining an effective Schrödinger equation that describes the fermion profile in the holographic direction (section A.2.2). We write the solution (section A.3) and notice that boundary conditions would generically imply a quantization of the frequencies as functions of the momenta (section A.3.1), resulting in a bunch of 2+12+1 dimensional dispersion relations. We finally obtain the normalization constants, and check that with the so defined basis the resulting second quantized theory satisfy correctly normalized commutation relations (section A.3.1).

A.1 The background

We work with a generic planar asymptotically AdS charged background, with the form

G\displaystyle G =\displaystyle= L2(fdt2+gdz2+dx2z2)=ηABωAωB,\displaystyle L^{2}\left(-f\,dt^{2}+g\,dz^{2}+\frac{d\vec{x}^{2}}{z^{2}}\right)=\eta_{AB}\,\omega^{A}\,\omega^{B}\,, (14)
A\displaystyle A =\displaystyle= hdt=AAωA.\displaystyle h\,dt=A_{A}\;\omega^{A}\,. (15)

Here ff, gg and hh, and consequently ωA\omega^{A} and AAA^{A}, are functions of zz. Since we want to couple fermions to the present background, we need the explicit form of the vielbein and dual vector basis {ωA,eA}\{\omega^{A},e_{A}\} that we wrote in the second equality. In terms of the planar indices AA =t¯,x¯,y¯,z¯=\underline{t},\underline{x},\underline{y},\underline{z}, they read

ωt¯Lfdt,et¯1Lft,ωx¯Lzdx,ex¯zLx,ωy¯Lzdy,ey¯zLy,ωz¯Lgdz,ez¯1Lgz.\begin{array}[]{lcl}\omega^{\underline{t}}\equiv L\,\sqrt{f}\;dt&,&\qquad e_{\underline{t}}\equiv\frac{1}{L\,\sqrt{f}}\;\partial_{t}\,,\cr\omega^{\underline{x}}\equiv\frac{L}{z\,}\;dx&,&\qquad e_{\underline{x}}\equiv\frac{z\,}{L}\;\partial_{x}\,,\cr\omega^{\underline{y}}\equiv\frac{L\,}{z}\;dy&,&\qquad e_{\underline{y}}\equiv\frac{z}{L\,}\;\partial_{y}\,,\cr\omega^{\underline{z}}\equiv L\,\sqrt{g}\;dz&,&\qquad e_{\underline{z}}\equiv\frac{1}{L\,\sqrt{g}}\;\partial_{z}\,.\end{array} (16)

This allows us to calculate the non-zero components of the spin connection {ωA}B\{\omega^{A}{}_{B}\} by imposing metricity ωAB=ωBA\omega_{AB}=-\omega_{BA}, and torsionless dωA+ωABωB=0\;d\omega^{A}+\omega^{A}{}_{B}\wedge\omega^{B}=0\; conditions. They take the form

ωt¯z¯\displaystyle\omega^{\underline{t}}{}_{\underline{z}} =\displaystyle= ωt¯z¯=+ωz¯t¯=(lnf)2gωt¯,\displaystyle-\omega_{{\underline{t}}{\underline{z}}}=+\omega_{{\underline{z}}{\underline{t}}}=\frac{(\ln f)^{\prime}}{2\,\sqrt{g}}\;\omega^{\underline{t}}\,, (17)
ωx¯z¯\displaystyle\omega^{\underline{x}}{}_{\underline{z}} =\displaystyle= +ωx¯z¯=ωz¯x¯=1zgωx¯,\displaystyle+\omega_{{\underline{x}}{\underline{z}}}=-\omega_{{\underline{z}}{\underline{x}}}=-\frac{1}{z\sqrt{g}}\;\omega^{\underline{x}}\,, (18)
ωy¯z¯\displaystyle\omega^{\underline{y}}{}_{\underline{z}} =\displaystyle= +ωy¯z¯=ωz¯y¯=1zgωy¯.\displaystyle+\omega_{{\underline{y}}{\underline{z}}}=-\omega_{{\underline{z}}{\underline{y}}}=-\frac{1}{z\sqrt{g}}\;\omega^{\underline{y}}\,. (19)

This allows us to write the Dirac equation for a charged fermion in the present background, as we do in the next subsection.

A.2 The Dirac equation

A.2.1 Separation of variables and spin components

We consider a four-component Dirac spinor Ψ\Psi with charge qq under the U(1)U(1) gauge field coupled to gravity through the covariant derivative

ΨΓA𝒟AΨ=ΓA(eA(Ψ)+i2ωBCΣBCAΨiqAAΨ),\not{\cal D}\,\Psi\equiv\Gamma^{A}\,{\cal D}_{A}\,\Psi=\Gamma^{A}\left(e_{A}(\Psi)+\frac{i}{2}\,\omega^{BC}{}_{A}\,\Sigma_{BC}\,\Psi-i\,q\,A_{A}\,\Psi\right)\,, (20)

where the Dirac gamma-matrices obey {ΓA,ΓB}=2ηAB\{\Gamma^{A},\Gamma^{B}\}=2\,\eta^{AB}, and ΣAB[ΓA,ΓB]/4i\Sigma_{AB}\equiv[\Gamma_{A},\Gamma_{B}]/{4\,i} are the generators in the spinorial representation of the local Lorentz group in 3+13+1 dimensions. We use along the paper the following representation for the gamma matrices

Γt¯(iσ100iσ1),Γx¯(σ200σ2),Γy¯(0σ2σ20),Γz¯(σ300σ3),\Gamma^{\underline{t}}\equiv\left(\begin{array}[]{cc}i\,\sigma_{1}&0\\ 0&i\,\sigma_{1}\end{array}\right)\,,\quad\quad\Gamma^{\underline{x}}\equiv\left(\begin{array}[]{cc}-\sigma_{2}&0\\ 0&\sigma_{2}\end{array}\right)\,,\quad\quad\Gamma^{\underline{y}}\equiv\left(\begin{array}[]{cc}0&\sigma_{2}\\ \sigma_{2}&0\end{array}\right)\,,\quad\quad\Gamma^{\underline{z}}\equiv\left(\begin{array}[]{cc}\sigma_{3}&0\\ 0&\sigma_{3}\end{array}\right)\,, (21)

where {σ1,σ2,σ3}\{\sigma_{1},\sigma_{2},\sigma_{3}\} are the Pauli matrices.

The free fermionic modes satisfy the Dirac equation

(m)Ψ=0.(\not{\cal D}-m)\,\Psi=0\,. (22)

To solve it, we find convenient to work in momentum space

Ψωk(t,x,z)=Ψωk(x,z)eiωt=1𝒩ωkzf(z)14ei(kxωt)ψωk(z),\Psi_{\omega\vec{k}}(t,\vec{x},z)=\Psi_{\omega\vec{k}}(\vec{x},z)e^{-i\omega t}=\frac{1}{{\cal N}_{\omega\vec{k}}}\;\frac{z}{f(z)^{\frac{1}{4}}}\,e^{i(\vec{k}\cdot\vec{x}-\omega t)}\,\psi_{\omega\vec{k}}(z)\,, (23)

where ω\omega, k\vec{k} are the energy and momentum along the xyxy plane, 𝒩ωk{\cal N}_{\omega\vec{k}} is a normalization constant, and the factor z/f(z)1/4z/f(z)^{1/4} has been included for later convenience. We can use rotational invariance to refer the momentum to the xx-axis, as

R[θ](kxky)=(k0),withR[θ]=eiθσ2=(cosθsinθsinθcosθ).\displaystyle R[\theta]\left(\begin{array}[]{c}k_{x}\\ k_{y}\end{array}\right)=\left(\begin{array}[]{c}k\\ 0\end{array}\right)\,,\qquad{\rm with}\qquad R[\theta]=e^{i\,\theta\,\sigma_{2}}=\left(\begin{array}[]{rr}\cos\theta&\sin\theta\\ -\sin\theta&\cos\theta\end{array}\right)\,. (30)

This allows us to write ψωk(z)=S[θ]ψωk(z)\psi_{\omega\vec{k}}(z)=S[\theta]\,\psi_{\omega k}(z) where

S[θ]=eiθΣx¯y¯=cosθ2+Γx¯y¯sinθ2=(cosθ2 12×2sinθ2 12×2sinθ2 12×2cosθ2 12×2),S[\theta]=e^{i\,\theta\,\Sigma_{\underline{x}\underline{y}}}=\cos\frac{\theta}{2}+\Gamma_{\underline{x}\underline{y}}\,\sin\frac{\theta}{2}=\left(\begin{array}[]{rr}\cos\frac{\theta}{2}\,1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}&-\sin\frac{\theta}{2}\,1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}\\ \sin\frac{\theta}{2}\,1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}&\cos\frac{\theta}{2}\,1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}\end{array}\right)\,, (31)

is the spinor representation of the rotation matrix (30), where 12×21_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}} is the 2×22\times 2 identity matrix, and we used the explicit form of the generator Σx¯y¯14i[Γx¯;Γy¯]=12iΓx¯y¯\Sigma_{{\underline{x}}{\underline{y}}}\equiv\frac{1}{4\,i}\,[\Gamma_{\underline{x}}\,;\Gamma_{\underline{y}}]=\frac{1}{2\,i}\,\Gamma_{\underline{x}\underline{y}}. The resulting form of the equation (22) for ψωk(z)\psi_{\omega k}(z) now reads

(1g(z)Γz¯zif(z)Γt¯(ω+qh(z))+ikzΓx¯mL)ψωk(z)=0.\displaystyle\left(\frac{1}{\sqrt{g(z)}}\,\Gamma_{\underline{z}}\,\partial_{z}-\frac{i}{\sqrt{f(z)}}\,\Gamma_{\underline{t}}\,\left(\omega+q\,{h(z)}\right)+i\,k\,z\,\Gamma_{\underline{x}}-m\,L\right)\;\psi_{\omega k}(z)=0\,. (32)

Notice that this equation is real due to the choice of gamma matrices (21).

Let us now introduce the projectors

Πα12(1+(1)α+1Γz¯Γt¯Γx¯)={(12×2000),α=1(00012×2),α=2\Pi_{\alpha}\equiv\frac{1}{2}\left(1+(-1)^{\alpha+1}\;\Gamma^{\underline{z}}\,\Gamma^{\underline{t}}\,\Gamma^{\underline{x}}\right)=\left\{\begin{array}[]{l}\left(\begin{array}[]{cc}1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}&0\\ 0&0\end{array}\right)\qquad,\qquad\alpha=1\\ \\ \left(\begin{array}[]{cc}0&0\\ 0&1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}\end{array}\right)\qquad,\qquad\alpha=2\end{array}\right. (33)

With their help we can decompose the Dirac field into “spin” states as ψωk(z)=(Π1+Π2)ψωk(z)=α=12ψαωk(z)\;\psi_{\omega k}(z)=(\Pi_{1}+\Pi_{2})\,\psi_{\omega k}(z)=\sum_{\alpha=1}^{2}\,\psi_{\alpha\omega k}(z), where the projected fields are

ψαωk(z)Παψωk(z)={(ψωk(1)(z)0),α=1(0ψωk(2)(z)),α=2\psi_{\alpha\omega k}(z)\equiv\Pi_{\alpha}\psi_{\omega k}(z)=\left\{\begin{array}[]{l}\left(\begin{array}[]{c}\psi^{(1)}_{\omega k}(z)\\ 0\end{array}\right)\qquad,\qquad\alpha=1\\ \\ \left(\begin{array}[]{c}0\\ \psi^{(2)}_{\omega k}(z)\end{array}\right)\qquad,\qquad\alpha=2\end{array}\right. (34)

where each of the ψωk(α)\psi_{\omega k}^{(\alpha)} have two components. Inserting the resulting decomposition in (23), we obtain a solution with energy ω\omega momentum k\vec{k} and “spin” α\alpha of the form

Ψαωk(t,x,z)=Ψαωk(x,z)eiωt=1𝒩αωkzf(z)14ei(kxωt)S[θ]ψαωk(z).\Psi_{\alpha\omega\vec{k}}(t,\vec{x},z)=\Psi_{\alpha\omega\vec{k}}(\vec{x},z)e^{-i\omega t}=\frac{1}{{\cal N}_{\alpha\omega\vec{k}}}\;\frac{z}{f(z)^{\frac{1}{4}}}\,e^{i(\vec{k}\cdot\vec{x}-\omega t)}\,S[\theta]\;\psi_{\alpha\omega k}(z)\,. (35)

Plugging this back into the field equation (32) we get a decoupled system for the bi-spinors ψωk(α)(z)\psi^{(\alpha)}_{\omega k}(z), as

ψωk(α)(z)+g(z)(ω+qh(z)f(z)iσ2+()αkzσ1mLσ3)ψωk(α)(z)=0,α=1,2\psi^{(\alpha)^{\prime}}_{\omega k}(z)+\sqrt{g(z)}\,\left(\frac{\omega+q\,{h(z)}}{\sqrt{f(z)}}\,i\,\sigma_{2}+(-)^{\alpha}\,k\,z\,\sigma_{1}-m\,L\,\sigma_{3}\right)\,\psi^{(\alpha)}_{\omega k}(z)=0\qquad,\qquad\,\alpha=1,2 (36)

This system can now be turned into a second order, Schrödinger-like equation for a single function ϕωk(z)\phi_{\omega k}(z), that is completely determined (up to normalization) by imposing smoothness at the boundary and smoolthness/ingoing boundary conditions at the horizon. We do this in the next subsection 555 It is worth to notice that from the representation (21) the generators of the Lorentz subgroup in 2+12+1 dimensions are Σt¯x¯=12i(σ300σ3),Σt¯y¯=12i(0σ3σ30),Σx¯y¯=12i(012×212×20),\Sigma_{{\underline{t}}{\underline{x}}}=\frac{1}{2\,i}\,\left(\begin{array}[]{cc}-\sigma_{3}&0\\ 0&\sigma_{3}\end{array}\right)\quad,\quad\Sigma_{\underline{t}\underline{y}}=\frac{1}{2\,i}\,\left(\begin{array}[]{cc}0&\sigma_{3}\\ \sigma_{3}&0\end{array}\right)\quad,\quad\Sigma_{\underline{x}\underline{y}}=\frac{1}{2\,i}\,\left(\begin{array}[]{cc}0&-1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}\\ 1_{\mbox{\tiny$2\!\!\times\!\!2$\normalsize}}&0\end{array}\right)\,, (37) From here it should be clear that ψωk(α)(z)\psi^{(\alpha)}_{\omega k}(z) are not Dirac spinors in 2+12+1 dimensions, since they mix under Lorentz transformations. .

A.2.2 The effective Schrödinger equation.

We want to transform (36) into a second order Schrödinger equation to which we can apply our intuitions regarding wave functions. In order to do that, let us introduce the functions

fk±(z)g(z)(zk±ω+qh(z)f(z)).f_{k}^{\pm}(z)\equiv\sqrt{g(z)}\;\left(z\,k\pm\frac{\omega+q\,{h(z)}}{\sqrt{f(z)}}\right)\,. (38)

Now, if we consider first α=2\alpha=2, we can parameterize the bi-spinor ψωk(2)(z)\psi^{(2)}_{\omega k}(z) in terms of two functions ψωk(2)(z)\psi^{(2)-}_{\omega k}(z) and ϕωk(2)(z)\phi^{(2)}_{\omega k}(z), in the form

ψωk(2)(z)(fk+(z)ϕω,k(2)(z)ψωk(2)(z)),\psi^{(2)}_{\omega k}(z)\equiv\left(\begin{array}[]{c}\sqrt{f_{k}^{+}(z)}\,\phi^{(2)}_{\omega,k}(z)\\ \psi^{(2)-}_{\omega k}(z)\end{array}\right)\,, (39)

then substituting in (36) we get from its components the pair of coupled first order equations

ϕωk(2)(z)+(fk+(z)2fk+(z)mLg(z))ϕωk(2)(z)+fk+(z)ψωk(2)(z)\displaystyle\phi^{(2)}_{\omega k}{}^{\prime}(z)+\left(\frac{f_{k}^{+}{}^{\prime}(z)}{2\,f_{k}^{+}(z)}-m\,L\;\sqrt{g(z)}\right)\;\phi^{(2)}_{\omega k}(z)+\sqrt{f^{+}_{k}(z)}\;\psi^{(2)-}_{\omega k}(z) =\displaystyle= 0,\displaystyle 0\,, (40)
ψωk(2)(z)+mLg(z)ψωk(2)(z)+fk(z)fk+(z)ϕωk(2)(z)\displaystyle\psi^{(2)-}_{\omega k}{}^{\prime}(z)+m\,L\;\sqrt{g(z)}\;\psi^{(2)-}_{\omega k}(z)+f^{-}_{k}(z)\sqrt{f^{+}_{k}(z)}\;\phi^{(2)}_{\omega k}(z) =\displaystyle= 0,\displaystyle 0\,, (41)

from the first of which we can rewrite

ψωk(2)(z)=1fk+(z)(ϕωk(2)(z)+(fk+(z)2fk+(z)mLg(z))ϕωk(2)(z)),\psi^{(2)-}_{\omega k}(z)=-\frac{1}{\sqrt{f^{+}_{k}(z)}}\;\left(\phi^{(2)}_{\omega k}{}^{\prime}(z)+\left(\frac{f_{k}^{+}{}^{\prime}(z)}{2\,f_{k}^{+}(z)}-m\,L\;\sqrt{g(z)}\right)\;\phi^{(2)}_{\omega k}(z)\right)\,, (42)

and now plugging this into the second equation, we get a second order Schrödinger-like equation for ϕωk(2)\phi^{(2)}_{\omega k}

ϕωk(2)(z)′′+U(z)ϕωk(2)(z)=0,-\phi^{(2)}_{\omega k}{}^{\prime\prime}(z)+U(z)\;\phi^{(2)}_{\omega k}(z)=0\,, (43)

where the potential is given by

U(z)\displaystyle U(z)\! =\displaystyle= g(z)(k2z2(ω+qh(z))2f(z)+m2L2)fk+(z)′′2fk+(z)+34(fk+(z)fk+(z))2mLg(z)(ln(fk+(z)g(z))),\displaystyle\!g(z)\!\left(k^{2}\!z^{2}\!-\!\frac{(\omega+q{h(z)})^{2}}{f(z)}\!+\!m^{2}\!L^{2}\right)\!-\!\frac{f_{k}^{+}{}^{\prime\prime}(z)}{2f_{k}^{+}(z)}+\frac{3}{4}\left(\frac{f_{k}^{+}{}^{\prime}(z)}{f_{k}^{+}(z)}\right)^{2}\!\!-mL\sqrt{g(z)}\left(\ln\left(\frac{f_{k}^{+}(z)}{\sqrt{g(z)}}\right)\right)^{\prime}\,, (44)

If instead we consider α=1\alpha=1, we can parametrize the bi-spinor ψωk(1)(z)\psi^{(1)}_{\omega k}(z) in terms of the functions ψωk(1)+(z)\psi^{(1)+}_{\omega k}(z) and ϕωk(1)(z)\phi^{(1)}_{\omega k}(z), as

ψωk(1)(z)(ψωk(1)+(z)fk+(z)ϕω,k(1)(z)).\psi^{(1)}_{\omega k}(z)\equiv\left(\begin{array}[]{c}\psi^{(1)+}_{\omega k}(z)\\ \sqrt{f_{k}^{+}(z)}\,\phi^{(1)}_{\omega,k}(z)\end{array}\right)\,. (46)

After introducing it in (36) we get as before a pair of coupled first order equations

ϕωk(1)(z)+(fk+(z)2fk+(z)+mLg(z))ϕωk(1)(z)fk+(z)ψωk(1)+(z)\displaystyle\phi^{(1)}_{\omega k}{}^{\prime}(z)+\left(\frac{f_{k}^{+}{}^{\prime}(z)}{2\,f_{k}^{+}(z)}+m\,L\;\sqrt{g(z)}\right)\;\phi^{(1)}_{\omega k}(z)-\sqrt{f^{+}_{k}(z)}\;\psi^{(1)+}_{\omega k}(z) =\displaystyle= 0,\displaystyle 0\,, (47)
ψωk(1)+(z)mLg(z)ψωk(1)+(z)fk(z)fk+(z)ϕωk(1)(z)\displaystyle\psi^{(1)+}_{\omega k}{}^{\prime}(z)-m\,L\;\sqrt{g(z)}\;\psi^{(1)+}_{\omega k}(z)-f^{-}_{k}(z)\sqrt{f^{+}_{k}(z)}\;\phi^{(1)}_{\omega k}(z) =\displaystyle= 0.\displaystyle 0\,. (48)

From the first equation we can obtain

ψωk(1)+(z)=1fk+(z)(ϕωk(1)(z)+(fk+(z)2fk+(z)+mLg(z))ϕωk(1)(z)),\psi^{(1)+}_{\omega k}(z)=\frac{1}{\sqrt{f^{+}_{k}(z)}}\;\left(\phi^{(1)}_{\omega k}{}^{\prime}(z)+\left(\frac{f_{k}^{+}{}^{\prime}(z)}{2\,f_{k}^{+}(z)}+m\,L\;\sqrt{g(z)}\right)\;\phi^{(1)}_{\omega k}(z)\right)\,, (49)

and by plugging it into the second equation of (47) we get a second order Schrödinger-like equation for the function ϕωk(1)\phi^{(1)}_{\omega k}, with the form

ϕωk(1)(z)′′+U(z)ϕωk(1)(z)=0,-\phi^{(1)}_{\omega k}{}^{\prime\prime}(z)+U(z)\;\phi^{(1)}_{\omega k}(z)=0\,, (50)

with exactly the same potential (44) as before.

We look for solutions of (43) or equivalently (50) which do no diverge neither at the boundary nor at the horizon. In other words, we need normalizable solutions to the Schrödinger-like problem.

A last remark: we could try a solution of equations (36) such that: ψωk(1)(z)=ψω(k)(2)(z)\;\psi^{(1)}_{\omega k}(z)=\psi^{(2)}_{\omega(-k)}(z) in (39) or (46). However, that would not be a smart choice since it would introduce a vanishing denominator at (42) or (49) whenever fk+(z)f^{+}_{k}(z) vanishes at negative value of the momenta. A better definition is to write (39) and (46) with ϕωk(1)(z)=ϕωk(2)(z)ϕωk(z)\phi^{(1)}_{\omega k}(z)=\phi^{(2)}_{\omega k}(z)\equiv\phi_{\omega k}(z) the unique solution of (50), as we will show nextly.

A.3 General form of the free fermionic bases

A.3.1 Quantization of the frequencies

We must now impose boundary conditions on our solutions, in order to ensure smoothness everywhere. In particular, we impose regular boundary conditions in the ultraviolet, and we can chose either regular or in-going boundary conditions in the infrared.

The boundary conditions have an important consequence: the frequency ω\omega is not free but is related to the modulus kk of the two-momentum through a dispersion relation that is in general labelled by some integer index mm, resulting in ω=ωm(k)\omega=\omega_{m}(k) (see section F.2 for explicit forms of dispersion relations). This allows us to replace the indices kωk\omega in our functions of the previous sections by kmkm.

Taking into account these facts, and collecting the results given in (36)-(50) we can write for the fermionic mode in (35), the general form

Ψαmk(x,z)\displaystyle\Psi_{\alpha m\vec{k}}(\vec{x},z) =\displaystyle= 1𝒩αmkzf(z)14eikxeiθΣx¯y¯(1(1)α2ψmk(1)+(z)1(1)α2fk+(z)ϕmk(z)1+(1)α2fk+(z)ϕmk(z)1+(1)α2ψmk(2)(z)),\displaystyle\frac{1}{{\cal N}_{\alpha mk}}\;\frac{z}{f(z)^{\frac{1}{4}}}\,e^{i\vec{k}\cdot\vec{x}}\,e^{i\,\theta\,\Sigma_{\underline{x}\underline{y}}}\;\left(\begin{array}[]{l}\frac{1-(-1)^{\alpha}}{2}\psi^{(1)+}_{mk}(z)\\ \\ \frac{1-(-1)^{\alpha}}{2}\sqrt{f_{k}^{+}(z)}\;\phi_{mk}(z)\\ \\ \frac{1+(-1)^{\alpha}}{2}\sqrt{f_{k}^{+}(z)}\;\phi_{mk}(z)\\ \\ \frac{1+(-1)^{\alpha}}{2}\psi^{(2)-}_{mk}(z)\end{array}\right)\,,\qquad (58)

where ψmk(2)(z)\psi^{(2)-}_{mk}(z) and ψmk(1)+(z)\psi^{(1)+}_{mk}(z) are given in (42) and (49) respectively, in terms of the solution ϕmk(z)\phi_{mk}(z) of (43) (or equivalently of (50)). Notice that the 44-tuple in parenthesis on the right corresponds to what was called ψαωk(z)\psi_{\alpha\omega k}(z) in equation (35), and that will be referred bellow as ψαmk(z)\psi_{\alpha mk}(z) in attention to the quantization of the frequencies.

A.3.2 Normalization

To completely define the modes (58) we need to fix their normalization. To this end we introduce in the space of spinors the following scalar product Birrell:1982ix

(Ψ1;Ψ2)(t)Σt𝑑zd2x|H|Ψ1(t,x,z)Ψ2(t,x,z),(\Psi_{1};\Psi_{2})(t)\equiv\int_{\Sigma_{t}}dz\,d^{2}x\;\sqrt{|H|}\;\Psi_{1}(t,\vec{x},z)^{\dagger}\;\Psi_{2}(t,\vec{x},z)\,, (59)

where Σt\Sigma_{t} is the space of constant tt and

H=L2(g(z)dz2+dx2z2);|H|=L3g(z)z2,H=L^{2}\,\left(g(z)\;dz^{2}+\frac{d\vec{x}^{2}}{z^{2}}\right)\qquad;\qquad\sqrt{|H|}=L^{3}\,\frac{\sqrt{g(z)}}{z^{2}}\,, (60)

is the induced metric on it. It is not difficult to prove that in the space of solutions to the Dirac equation (22) the operator iti\partial_{t} is hermitian, i.e.

(Ψ1;itΨ2)(t)=(itΨ1;Ψ2)(t).(\Psi_{1};i\partial_{t}\Psi_{2})(t)=(i\partial_{t}\Psi_{1};\Psi_{2})(t)\,. (61)

Standard arguments then show that eigenspinors of iti\partial_{t} with different eigenvalues are orthogonal with respect to (59). By applying this result to the spinors (58) we know that they result orthogonal for different mm’s. By using this fact we straightforwardly get the orthonormality relation

(Ψαmk;Ψαmk)Σt𝑑zd2xHΨαmk(x,z)Ψαmk(x,z)=δααδmmδ2(kk),\left(\Psi_{\alpha m\vec{k}};\Psi_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\right)\equiv\int_{\Sigma_{t}}dz\,d^{2}\vec{x}\;\sqrt{H}\;\Psi^{\dagger}_{\alpha m\vec{k}}(\vec{x},z)\;\Psi_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}(\vec{x},z)=\delta_{\alpha\alpha^{\prime}}\;\delta_{mm^{\prime}}\;\delta^{2}(\vec{k}-\vec{k}^{\prime})\,, (62)

if the normalization constant is fixed such that

|𝒩αmk|2=(2π)2L30𝑑zg(z)f(z)ψαmk(z)ψαmk(z).|{\cal N}_{\alpha mk}|^{2}=(2\,\pi)^{2}\,L^{3}\,\int_{0}^{\infty}\,dz\,\sqrt{\frac{g(z)}{f(z)}}\;\psi_{\alpha mk}(z)^{\dagger}\;\psi_{\alpha mk}(z)\,. (63)

Finally, after the calculations above, the general solution of (22) can be written in coordinates space as

Ψ(t,x,z)=d2kαmcαmk(t)Ψαmk(x,z).\Psi(t,\vec{x},z)=\int\,d^{2}\vec{k}\;\sum_{\alpha m}\,c_{\alpha m\vec{k}}(t)\;\Psi_{\alpha m\vec{k}}(\vec{x},z)\,. (64)

For later use, it is worth to remind that (62) yields for the above solutions the completeness relation

d2kαmΨαmk(x,z)Ψαmk(x,z)=1|H|δ(zz)δ2(xx).\int\,d^{2}\vec{k}\;\sum_{\alpha m}\,\Psi_{\alpha m\vec{k}}(\vec{x},z)\;\Psi^{\dagger}_{\alpha m\vec{k}}(\vec{x}^{\prime},z^{\prime})=\frac{1}{\sqrt{|H|}}\,\delta(z-z^{\prime})\;\delta^{2}(\vec{x}-\vec{x}^{\prime})\,. (65)

As we will see in the next section, the above defined normalization provides canonical anti-commutation relations for the corresponding Hamiltonian variables. This allows us to define canonical creation and annihilation operators for the fermionic modes in the bulk. With this, we will be able to establish a standard Landau description for the bulk Fermi liquid.

Appendix B Hamiltonian theory

In this Appendix we develop the Hamiltonian theory for the fermions in the bulk, using the orthonormal basis obtained in the previous section. In section B.1, we define our dynamics for the fermionic degrees of freedom, and write the generic form of the corresponding Hamiltonian. Then in section B.2 we deal with its free part, while the interacting part is worked out in section B.3.

B.1 Dynamical setup

The action for our spinor field that propagates in the asymptotically AdS bulk reads

SΨ=d4x|G|Ψ=d4x|G|(Ψ¯(m)Ψ+Tσ3σ4σ1σ2Ψ¯σ1Ψ¯σ2Ψσ3Ψσ4),S^{\Psi}=\int d^{4}x\sqrt{-|G|}\;{\cal L}^{\Psi}=-\int d^{4}x\sqrt{-|G|}\;\left(\bar{\Psi}\,(\not{\cal D}-m)\Psi+T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}}\;\bar{\Psi}_{\sigma_{1}}\;\bar{\Psi}_{\sigma_{2}}\;\Psi^{\sigma_{3}}\;\Psi^{\sigma_{4}}\right)\,, (66)

where the σ\sigma’s are spin indices running from 11 to 44, and we have defined the conjugate spinor Ψ¯ΨiΓt¯\bar{\Psi}\equiv\Psi^{\dagger}\,i\,\Gamma^{\underline{t}}. We included a four-fermion interaction term, which takes the most general form that respects covariance in four-dimensional curved space and fermion number conservation. It is written in terms of an invariant tensor Tσ3σ4σ1σ2T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}} given by

Tσ3σ4σ1σ2\displaystyle T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}} =\displaystyle= g1δσ1δσ2σ3+σ4g2(Γ5)σ1(Γ5)σ2σ3+σ4g3(Γa)σ1(Γa)σ2σ3σ4\displaystyle g_{1}\;\delta^{\sigma_{1}}{}_{\sigma_{3}}\;\delta^{\sigma_{2}}{}_{\sigma_{4}}+g_{2}\;(\Gamma^{5})^{\sigma_{1}}{}_{\sigma_{3}}\;(\Gamma^{5})^{\sigma_{2}}{}_{\sigma_{4}}+g_{3}\;(\Gamma^{a})^{\sigma_{1}}{}_{\sigma_{3}}\;(\Gamma_{a})^{\sigma_{2}}{}_{\sigma_{4}} (67)
+\displaystyle+ g4(ΓaΓ5)σ1(ΓaΓ5)σ2σ3σ4g54([Γa,Γb]Γ5)σ1([Γa,Γb]Γ5)σ2σ3,σ4\displaystyle g_{4}\;(\Gamma^{a}\,\Gamma^{5})^{\sigma_{1}}{}_{\sigma_{3}}\;(\Gamma_{a}\,\Gamma^{5})^{\sigma_{2}}{}_{\sigma_{4}}-\frac{g_{5}}{4}\;\left([\Gamma^{a},\Gamma^{b}]\,\Gamma^{5}\right)^{\sigma_{1}}{}_{\sigma_{3}}\;\left([\Gamma_{a},\Gamma_{b}]\,\Gamma^{5}\right)^{\sigma_{2}}{}_{\sigma_{4}}\,, (68)

where we defined Γ5iΓt¯Γx¯Γy¯Γz¯\;\Gamma^{5}\equiv-i\,\Gamma^{\underline{t}}\,\Gamma^{\underline{x}}\,\Gamma^{\underline{y}}\,\Gamma^{\underline{z}}\,\,. This tensor satisfies the condition Tσ3σ4σ1σ2=Tσ4σ3σ2σ1T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}}=T^{\sigma_{2}\sigma_{1}}_{\sigma_{4}\sigma_{3}}, as well as the Lorentz invariance property

Tσ3σ4σ1σ2=S[Λ]σ1Sδ1[Λ]σ2Tδ3δ4δ1δ2δ2S[Λ]1δ3Sσ3[Λ]1δ4,σ4\displaystyle T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}}=S[\Lambda]^{\sigma_{1}}{}_{\delta_{1}}\;S[\Lambda]^{\sigma_{2}}{}_{\delta_{2}}\;T^{\delta_{1}\delta_{2}}_{\delta_{3}\delta_{4}}\;{S[\Lambda]^{-1}}^{\delta_{3}}{}_{\sigma_{3}}\;{S[\Lambda]^{-1}}^{\delta_{4}}{}_{\sigma_{4}}\,, (69)

for any Lorentz transformation Λ\Lambda, where S[Λ]S[\Lambda] denotes its spinorial representation. This property will be useful in what follows, in particular when we apply it to the rotation on the x¯y¯\underline{x}\underline{y} plane represented by the unitary matrix S[θ]S[\theta] defined in (31).

From the action (66) the momentum conjugate to the spinor field is

π=i|H|Ψ,\pi=i\,\sqrt{|H|}\,\Psi^{\dagger}\,, (70)

allowing us to write the Hamiltonian as

H=d2x𝑑z(πtΨ|G|Ψ),H=\int d^{2}\vec{x}\,dz\,\left(\pi\,\partial_{t}\Psi-\sqrt{-|G|}\,{\cal L}^{\Psi}\right)\,, (71)

where Ψ{\cal L}^{\Psi} can be read from (66).

Inserting the momentum (70) into the canonical anti-commutation relations for coordinates and momenta, we get that the spinor field satisfies

{Ψ(t,x,z),Ψ(t,x,z)}=1|H|δ(zz)δ2(xx).\{\Psi(t,\vec{x},z),\Psi^{\dagger}(t,\vec{x}^{\prime},z^{\prime})\}=\frac{1}{\sqrt{|H|}}\,\delta(z-z^{\prime})\;\delta^{2}(\vec{x}-\vec{x}^{\prime})\,. (72)

Then from the decomposition (64) we get the standard anti-commutation relations

{cαmk(t),cαmk(t)}=δααδmmδ2(kk),\{c_{\alpha m\vec{k}}(t),c^{\dagger}_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}(t)\}=\delta_{\alpha\alpha^{\prime}}\;\delta_{mm^{\prime}}\;\delta^{2}(\vec{k}-\vec{k}^{\prime})\,, (73)

identifying cαmkc^{\dagger}_{\alpha m\vec{k}} and cαmkc_{\alpha m\vec{k}} as the creation and annihilation operators respectively for the fermionic modes in the bulk. In the rest of this section we rewrite the bulk dynamics given by (71) in terms of them.

B.2 The free Hamiltonian

Let us first take the free part of the Hamiltonian (71), which reads

H𝖿𝗋𝖾𝖾\displaystyle H_{\sf free} =\displaystyle= d2x𝑑z|H|(iΨ(i2ωBCΣBCtiqh(z))Ψ|Gtt|Ψ¯(Γz¯𝒟z¯+Γi¯𝒟i¯m)Ψ)\displaystyle-\!\int d^{2}\vec{x}\int dz\sqrt{|H|}\left(i\Psi^{\dagger}\left(\frac{i}{2}\omega^{BC}{}_{t}\Sigma_{BC}-i\,q\,h(z)\,\right)\Psi\right.-\left.\sqrt{|G_{tt}|}\bar{\Psi}\,(\Gamma^{\underline{z}}{\cal D}_{\underline{z}}+\Gamma^{\underline{i}}{\cal D}_{\underline{i}}-m)\Psi\right) (74)

now using the Dirac equation we get

H𝖿𝗋𝖾𝖾=12d2x𝑑z|H|iΨtΨ+h.c..H_{\sf free}=\frac{1}{2}\int d^{2}\vec{x}\,\int dz\,\sqrt{|H|}\;i\Psi^{\dagger}\partial_{t}\Psi+h.c.\,. (76)

Then inserting the decomposition (64) and using the orthogonality relation (65) we get the free Hamiltonian in terms of creation and annihilation operators, as

H𝖿𝗋𝖾𝖾=αmd2kωm(k)cαmkcαmk.H_{\sf free}=\sum_{\alpha m}\int d^{2}\vec{k}\;\omega_{m}(k)\;c^{\dagger}_{\alpha m\vec{k}}\;c_{\alpha m\vec{k}}\,. (77)

In other words, the bulk degrees of freedom are described by a set of independent fermionic species labelled with an integer index mm corresponding to the mode on the zz direction. Each species has a dispersion relation given by ωm(k)\omega_{m}(k), with a two-fold degeneracy given by the spin α\alpha.

B.3 The interaction Hamiltonian

Regarding the interacting part of the Hamiltonian, we have

H𝗂𝗇𝗍=𝑑zd2x|H|Tσ3σ4σ1σ2Ψ¯σ1Ψ¯σ2Ψσ3Ψσ4.\displaystyle H_{\sf int}=\int dz\,d^{2}{\vec{x}}\,\sqrt{|H|}\,\;T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}}\;\bar{\Psi}_{\sigma_{1}}\;\bar{\Psi}_{\sigma_{2}}\;\Psi^{\sigma_{3}}\;\Psi^{\sigma_{4}}\,. (78)

Plugging the decomposition (64) into (78) we get

H𝗂𝗇𝗍\displaystyle H_{\sf int}\! =\displaystyle= α1α2α3α4m1m2m3m4d2k1d2k4δ(2)(k1+k2k3k4)tα3m3k3;α4m4k4α1m1k1;α2m2k2cα1m1k1cα2m2k2cα3m3k3cα4m4k4,\displaystyle\!\!\!\!\sum_{\underset{m_{1}m_{2}m_{3}m_{4}}{\mbox{\tiny$\alpha_{1}\alpha_{2}\alpha_{3}\alpha_{4}$\normalsize}}}\!\!\!\int d^{2}\!k_{1}\dots d^{2}\!k_{4}\,\delta^{(2)}\!(\vec{k}_{1}\!+\!\vec{k}_{2}\!-\!\vec{k}_{3}\!-\!\vec{k}_{4})\,\;t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}}\;c_{\alpha_{1}m_{1}{\vec{k}}_{1}}^{\dagger}\;c_{\alpha_{2}m_{2}{\vec{k}}_{2}}^{\dagger}c_{\alpha_{3}m_{3}{\vec{k}}_{3}}\;c_{\alpha_{4}m_{4}{\vec{k}}_{4}}\,,

where the x\vec{x} integral has been performed explicitly giving origin to the momentum conservation δ\delta-function, while the zz integral is contained in the definition of the momentum-space interaction tensor

tα3m3k3;α4m4k4α1m1k1;α2m2k2\displaystyle t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}} =\displaystyle= L3(2π)2S[θ1]Sδ1σ1[θ2]Tσ3σ4σ1σ2δ2σ2S[θ3]σ3Sδ3[θ4]σ4Iα3m3k3;α4m4k4;δ3δ4α1m1k1;α2m2k2;δ1δ2δ4,\displaystyle L^{3}(2\pi)^{2}\,S[\theta_{1}]^{\dagger}{}^{\delta^{1}}{}_{\!\!\sigma^{1}}\,S[\theta_{2}]^{\dagger}{}^{\delta^{2}}{}_{\!\!\sigma^{2}}\,T^{\sigma_{1}\sigma_{2}}_{\sigma_{3}\sigma_{4}}\,S[\theta_{3}]^{\sigma^{3}}\!{}_{\!\delta^{3}}\,S[\theta_{4}]^{\sigma^{4}}\!{}_{\!\delta^{4}}\,I^{\alpha_{1}m_{1}{k}_{1};\alpha_{2}m_{2}{k}_{2};\delta_{1}\delta_{2}}_{\alpha_{3}m_{3}{k}_{3};\alpha_{4}m_{4}{k}_{4};\delta_{3}\delta_{4}}\,,

in terms of the integrals

Iα3m3k3;α4m4k4;δ3δ4α1m1k1;α2m2k2;δ1δ2\displaystyle I^{\alpha_{1}m_{1}{k}_{1};\alpha_{2}m_{2}{k}_{2};\delta_{1}\delta_{2}}_{\alpha_{3}m_{3}{k}_{3};\alpha_{4}m_{4}{k}_{4};\delta_{3}\delta_{4}} =\displaystyle= 𝑑zz2g(z)f(z)ψ¯α1m1k1;δ1(z)𝒩α1m1k1ψ¯α2m2k2;δ2(z)𝒩α2m2k2ψα3m3k3δ3(z)𝒩α3m3k3ψα4m4k4δ4(z)𝒩α4m4k4,\displaystyle\int dz\;z^{2}\;\frac{\sqrt{g(z)}}{f(z)}\;\frac{\bar{\psi}_{\alpha_{1}m_{1}k_{1};\delta_{1}}(z)}{{\cal N}_{\alpha_{1}m_{1}k_{1}}}\;\frac{\bar{\psi}_{\alpha_{2}m_{2}k_{2};\delta_{2}}(z)}{{\cal N}_{\alpha_{2}m_{2}k_{2}}}\;\frac{\psi_{\alpha_{3}m_{3}k_{3}}^{\delta^{3}}(z)}{{\cal N}_{\alpha_{3}m_{3}k_{3}}}\;\frac{\psi_{\alpha_{4}m_{4}k_{4}}^{\delta^{4}}(z)}{{\cal N}_{\alpha_{4}m_{4}k_{4}}}\,, (81)

where functions ψαωk(z)\psi_{\alpha\omega k}(z) are defined as in equation (35) and correspond to the 44-tuples in parenthesis on the right of equation (58). As we show below, only a subset of the integrals (LABEL:eq:fermion.integrals.all) is needed for a Landau description of the bulk fluid.

Appendix C Landau description of the bulk fermions

In the previous section we rewrite the dynamics of the Dirac spinor in the bulk as that of a bunch of fermionic species in one less dimension, indexed by the energy mode mm and the spin α\alpha. In the present appendix, we develop a description in terms of the Landau theory for a multi-component Fermi liquid. We do that by using a perturbative approach.

C.1 Perturbative derivation

We work in the grand canonical ensemble at chemical potential μ\mu and temperature T=1/βT=1/\beta (that we will take to zero at the end of the calculations). We define the expectation value of an operator 𝒪{\cal O} by

𝒪=1Ztr(eβ(HμN)𝒪),whereZ=tr(eβ(HμN))=eβΩ(T,μ),\langle{\cal O}\rangle=\frac{1}{Z}\,{\rm tr}\left(e^{-\beta\left(H-\mu\,N\right)}{\cal O}\right)\,,\qquad\quad{\rm where}\qquad\quad Z={\rm tr}\left(e^{-\beta\left(H-\mu\,N\right)}\right)=e^{-\beta\,\Omega(T,\mu)}\,, (83)

where NN is the number of particles and Ω(T,μ)\Omega(T,\mu) is the grand canonical potential. The last formula can be inverted according to

Ω(T,μ)=1βlogZ.\Omega(T,\mu)=-\frac{1}{\beta}\log Z\,. (84)

To perform a perturbative approximation, we write the Hamiltonian as

H=H𝖿𝗋𝖾𝖾+H𝗂𝗇𝗍,H=H_{\sf free}+H_{\sf int}\,, (85)

and assume that the constants gig_{i} in (67) are small. With the help of the free Hamiltonian H𝖿𝗋𝖾𝖾H_{\sf free} we define the zeroth order quantities

𝒪𝖿𝗋𝖾𝖾=1Z𝖿𝗋𝖾𝖾tr(eβ(H𝖿𝗋𝖾𝖾μN)𝒪)whereZ𝖿𝗋𝖾𝖾=tr(eβ(H𝖿𝗋𝖾𝖾μN)).\langle{\cal O}\rangle_{\sf free}=\frac{1}{Z_{\sf free}}\,{\rm tr}\left(e^{-\beta\left(H_{\sf free}-\mu\,N\right)}{\cal O}\right)\qquad\quad{\rm where}\qquad\quad Z_{\sf free}={\rm tr}\left(e^{-\beta\left(H_{\sf free}-\mu\,N\right)}\right)\,. (86)

Then expanding (83) to first order in gig_{i}, we obtain

ZZ𝖿𝗋𝖾𝖾(1βH𝗂𝗇𝗍𝖿𝗋𝖾𝖾),Z\approx Z_{\sf free}\left(1-\beta\langle H_{\sf int}\rangle_{\sf free}\right)\,, (87)

which implies for the grand canonical potential

Ω(T,μ)1βlogZ𝖿𝗋𝖾𝖾+H𝗂𝗇𝗍𝖿𝗋𝖾𝖾.\Omega(T,\mu)\approx-\frac{1}{\beta}\log Z_{\sf free}+\langle H_{\sf int}\rangle_{\sf free}\,. (88)

For the first term in (88) we can write

Z𝖿𝗋𝖾𝖾=eβd2kαmωm(k)Nαmk,Z_{\sf free}=e^{-\beta\int d^{2}\vec{k}\,\sum_{\alpha m}\;\omega_{m}(k)N_{\alpha m\vec{k}}}\,, (89)

where NαmkN_{\alpha m\vec{k}} are the occupation numbers of the one-particle states Nαmk=cαmkcαmkN_{\alpha m\vec{k}}=\langle c^{\dagger}_{\alpha m\vec{k}}c_{\alpha m\vec{k}}\rangle. This implies for the free part of the grand canonical potential

1βlogZ𝖿𝗋𝖾𝖾=d2kαmωm(k)Nαmk.-\frac{1}{\beta}\log Z_{\sf free}=\int d^{2}\vec{k}\,\sum_{\alpha m}\;\omega_{m}(k)N_{\alpha m\vec{k}}\,. (90)

On the other hand, to write the second term in (88), we use (B.3) to have

H𝗂𝗇𝗍𝖿𝗋𝖾𝖾=α1α2α3α4m1m2m3m4d2k1d2k4δ(2)(k1+k2k3k4)tα3m3k3;α4m4k4α1m1k1;α2m2k2cα1m1k1cα2m2k2cα3m3k3cα4m4k4𝖿𝗋𝖾𝖾\langle H_{\sf int}\rangle_{\sf free}=\!\!\!\!\sum_{\underset{m_{1}m_{2}m_{3}m_{4}}{\mbox{\tiny$\alpha_{1}\alpha_{2}\alpha_{3}\alpha_{4}$\normalsize}}}\!\!\!\int d^{2}\!k_{1}\dots d^{2}\!k_{4}\,\delta^{(2)}\!(\vec{k}_{1}\!+\!\vec{k}_{2}\!-\!\vec{k}_{3}\!-\!\vec{k}_{4})\,\;t^{\alpha_{1}m_{1}{\vec{k}}_{1};\alpha_{2}m_{2}{\vec{k}}_{2}}_{\alpha_{3}m_{3}{\vec{k}}_{3};\alpha_{4}m_{4}{\vec{k}}_{4}}\,\left\langle\!c_{\alpha_{1}m_{1}{\vec{k}}_{1}}^{\dagger}\,c_{\alpha_{2}m_{2}{\vec{k}}_{2}}^{\dagger}c_{\alpha_{3}m_{3}{\vec{k}}_{3}}\,c_{\alpha_{4}m_{4}{\vec{k}}_{4}}\!\right\rangle_{\sf free} (91)

Calculating the expectation value in the right hand side

cα1m1k1cα2m2k2cα3m3k3cα4m4k4𝖿𝗋𝖾𝖾=Nα3m3k3Nα4m4k4×\displaystyle\left\langle c_{\alpha_{1}m_{1}\vec{k}_{1}}^{\dagger}c_{\alpha_{2}m_{2}\vec{k}_{2}}^{\dagger}c_{\alpha_{3}m_{3}\vec{k}_{3}}c_{\alpha_{4}m_{4}\vec{k}_{4}}\right\rangle_{\sf free}=N_{\alpha_{3}m_{3}\vec{k}_{3}}\,N_{\alpha_{4}m_{4}\vec{k}_{4}}\times (92)
×(δα1α4δα2α3δm1m4δm2m3δ2(k1k)δ2(k2k3)δα1α3δα2α4δm1m3δm2m4δ2(k1k3)δ2(k2k4)),\displaystyle\times\left(\delta_{\alpha_{1}\alpha_{4}}\,\delta_{\alpha_{2}\alpha_{3}}\,\delta_{m_{1}m_{4}}\,\delta_{m_{2}m_{3}}\,\delta^{2}(\vec{k}_{1}-\vec{k}^{\prime})\,\delta^{2}(\vec{k}_{2}-\vec{k}_{3})-\delta_{\alpha_{1}\alpha_{3}}\,\delta_{\alpha_{2}\alpha_{4}}\,\delta_{m_{1}m_{3}}\,\delta_{m_{2}m_{4}}\,\delta^{2}(\vec{k}_{1}-\vec{k}_{3})\,\delta^{2}(\vec{k}_{2}-\vec{k}_{4})\right)\,,

we replace back in (91) to get

H𝗂𝗇𝗍𝖿𝗋𝖾𝖾\displaystyle\left\langle H_{\sf int}\right\rangle_{\sf free} =\displaystyle= 12d2kd2kαmαmfαmαm(k,k)NαmkNαmk,\displaystyle\frac{1}{2}\int\!d^{2}\vec{k}\,d^{2}\vec{k}^{\prime}\,\sum_{\alpha m\alpha^{\prime}m^{\prime}}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\,N_{\alpha m\vec{k}}\,N_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\,, (93)

where the “Landau interaction functions” fαmαm(k,k)f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime}) are defined according to

fαmαm(k,k)tαmk;αmkαmk;αmktαmk;αmkαmk;αmk+tαmk;αmkαmk;αmktαmk;αmkαmk;αmk.f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\equiv t^{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}}_{\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\alpha m{\vec{k}}}-t^{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}}_{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}}+t^{\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\alpha m{\vec{k}}}_{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}}-t^{\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\alpha m{\vec{k}}}_{\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\alpha m{\vec{k}}}\,. (94)

Here we used the definition (B.3), and re-absorbed into the couplings an infinite constant δ2(0)\delta^{2}(\vec{0}) coming from collapsing the momentum integrals666Formally this is accomplished by regularizing through the introduction of a finite volume in coordinate space, and then taking the infinite volume limit at the end of the calculations.. The Landau interaction functions contain all the information about the interactions on the multi-component Fermi liquid. They verify the following relations

fαmαm(k,k)=fαmαm(k,k);fαmαm(k,k)=0.f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})=f_{\alpha^{\prime}m^{\prime}\alpha m}(\vec{k}^{\prime},\vec{k})\qquad;\qquad f_{\alpha m\alpha m}(\vec{k},\vec{k})=0\,. (95)

At this point of the discussion, a remark is in order. It is usual for condense matter theorists to think about the Landau interaction function as a correlation function. More precisely, let us introduce the index A(αmk)A\equiv(\alpha m\vec{k}) and denote by

ΓA1A2;A3A4(1PI)cA1cA2cA3cA41PI\Gamma^{(1PI)}_{A_{1}A_{2};A_{3}A_{4}}\equiv\left\langle c_{A_{1}}\,c_{A_{2}}\,c^{\dagger}_{A_{3}}\,c^{\dagger}_{A_{4}}\right\rangle_{1PI} (96)

the one-particle irreducible vertex that describes the interacction of two elementary excitations. Then it can be shown that 777 As they are not essential for the argument, we do not care here with numerical factors and propagators residues, that in any case are irrelevant in the order at which we work, see nozieresdupuis for details.

fAAΓAA;AA(1PI)f_{AA^{\prime}}\sim\Gamma^{(1PI)}_{AA^{\prime};A^{\prime}A} (97)

At leading order in perturbation theory the right hand side of (96) reads peskin

ΓA1A2;A3A4(1PI)|𝗅𝖾𝖺𝖽𝗂𝗇𝗀𝗈𝗋𝖽𝖾𝗋\displaystyle\left.\Gamma^{(1PI)}_{A_{1}A_{2};A_{3}A_{4}}\right|_{{\sf leading~order}} =\displaystyle= i𝑑tH𝗂𝗇𝗍cA1cA2cA3cA4\displaystyle-i\,\left\langle\int dt\,H_{\sf int}\;c_{A_{1}}\,c_{A_{2}}\,c^{\dagger}_{A_{3}}\,c^{\dagger}_{A_{4}}\right\rangle (98)
\displaystyle\sim tA4A3A1A2tA3A4A1A2+tA3A4A2A1tA4A3A2A1\displaystyle t^{A_{1}A_{2}}_{A_{4}A_{3}}-t^{A_{1}A_{2}}_{A_{3}A_{4}}+t^{A_{2}A_{1}}_{A_{3}A_{4}}-t^{A_{2}A_{1}}_{A_{4}A_{3}} (99)

where in the first line all the operators are in the interaction picture, with the interacting Hamiltonian given in (B.3), and in the second line we omitted a momentum conservation delta-function. By plugging (98) into (97) we get our result (94).

Now we can use (93) and (94) to write the grand canonical potential as

Ω(T,μ)=d2kαmωm(k)Nαmk+12d2kd2kαmαmfαmαm(k,k)NαmkNαmk.\Omega(T,\mu)=\int d^{2}\vec{k}\,\sum_{\alpha m}\;\omega_{m}(k)N_{\alpha m\vec{k}}+\frac{1}{2}\int\!d^{2}\vec{k}\,d^{2}\vec{k}^{\prime}\,\sum_{\alpha m\alpha^{\prime}m^{\prime}}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\,N_{\alpha m\vec{k}}\,N_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\,. (100)

A perturbation δNαmk\delta N_{\alpha m\vec{k}} of the ground state occupation numbers NαmkN_{\alpha m\vec{k}} gives us a variation of the grand canonical potential with the form

δΩ(T,μ)=d2kαmϵm(k)δNαmk+12d2kd2kαmαmfαmαm(k,k)δNαmkδNαmk,\delta\Omega(T,\mu)=\int d^{2}\vec{k}\,\sum_{\alpha m}\;\epsilon_{m}(k)\delta N_{\alpha m\vec{k}}+\frac{1}{2}\int\!d^{2}\vec{k}\,d^{2}\vec{k}^{\prime}\,\sum_{\alpha m\alpha^{\prime}m^{\prime}}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\,\delta N_{\alpha m\vec{k}}\,\delta N_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\,, (101)

where the quasiparticle dispersion relation ϵm(k)\epsilon_{m}(k) has been defined as

ϵm(k)=ωm(k)+αmd2kfαmαm(k,k)Nαmk.\epsilon_{m}(k)=\omega_{m}(k)+\sum_{\alpha^{\prime}m^{\prime}}\int\!d^{2}\vec{k}^{\prime}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\,N_{\alpha^{\prime}m^{\prime}\vec{k}^{\prime}}\,. (102)

For any perturbation δNαmk\delta N_{\alpha m\vec{k}} the quantity δΩ(T,μ)\delta\Omega(T,\mu) must be positive to have a stable ground state. Whenever it becomes negative, an instability is triggered.

C.2 Explicit form of the Landau interaction functions

We want to obtain a more explicit form of the Landau interaction functions fαmαm(k,k)f_{\alpha m\,\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\; suitable of being implemented into a numerical code. The integrals (LABEL:eq:fermion.integrals.all) we need are

Iαmk;δ1δ2αmk;δ3δ4\displaystyle I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}} =\displaystyle= Iαmk;αmk;δ3δ4αmk;αmk;δ1δ2,\displaystyle I^{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\delta_{1}\delta_{2}}_{\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\alpha m{\vec{k}};\delta_{3}\delta_{4}}\,, (103)

and they satify Iαmk;δ1δ2αmk;δ3δ4=Iαmk;δ2δ1αmk;δ4δ3I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}}=I^{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{4}\delta_{3}}_{\alpha m{k};\delta_{2}\delta_{1}} what can be used to rewrite (B.3) as

tαmk;αmkαmk;αmk\displaystyle t^{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}}_{\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime};\alpha m{\vec{k}}} =\displaystyle= S[θθ]Tδ3σ2σ1δ2δ1σ1S[θθ]σ2Iαmk;δ1δ2αmk;δ3δ4δ4,\displaystyle S[\theta-\theta^{\prime}]^{\dagger}{}^{\delta^{1}}{}_{\sigma_{1}}\;T^{\sigma_{1}\delta_{2}}_{\delta_{3}\sigma_{2}}\;S[\theta-\theta^{\prime}]^{\sigma^{2}}{}_{\delta^{4}}\;I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}}\,, (104)
tαmk;αmkαmk;αmk\displaystyle t^{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}}_{\alpha m{\vec{k}};\alpha^{\prime}m^{\prime}{\vec{k}}^{\prime}} =\displaystyle= S[θθ]Tσ2δ3σ1δ2δ1σ1S[θθ]σ2Iαmk;δ1δ2αmk;δ3δ4δ4.\displaystyle S[\theta-\theta^{\prime}]^{\dagger}{}^{\delta^{1}}{}_{\sigma_{1}}\;T^{\sigma_{1}\delta_{2}}_{\sigma_{2}\delta_{3}}\;S[\theta-\theta^{\prime}]^{\sigma^{2}}{}_{\delta^{4}}\;I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}}\,. (105)

To further disentangle the angle dependence, we find convenient to use the explicit form of the spinorial rotations matrices (31) to write

S[θθ]Sδ1σ1[θθ]σ2=δ4P0+σ1δ4δ1σ2cos(θθ)Pc+σ1δ4δ1σ2sin(θθ)Ps,σ1δ4δ1σ2S[\theta-\theta^{\prime}]^{\dagger}{}^{\delta_{1}}{}_{\sigma_{1}}\;S[\theta-\theta^{\prime}]^{\sigma^{2}}{}_{\delta^{4}}=P_{0}{}^{\delta_{1}\sigma_{2}}_{\sigma_{1}\delta_{4}}+\cos(\theta-\theta^{\prime})\;P_{c}{}^{\delta_{1}\sigma_{2}}_{\sigma_{1}\delta_{4}}+\sin(\theta-\theta^{\prime})\;P_{s}{}^{\delta_{1}\sigma_{2}}_{\sigma_{1}\delta_{4}}\,, (106)

where the constant, cosine and sine auxiliary tensors are given respectively as

P0σ1δ4δ1σ2\displaystyle P_{0}{}^{\delta_{1}\sigma_{2}}_{\sigma_{1}\delta_{4}} \displaystyle\equiv 12(δδ1δσ2σ1δ4Γx¯y¯Γx¯y¯δ1σ1)σ2δ4=+P0,δ4σ1σ2δ1\displaystyle\frac{1}{2}\;\left(\delta^{\delta_{1}}{}_{\sigma_{1}}\;\delta^{\sigma_{2}}{}_{\delta_{4}}-\Gamma_{\underline{x}\underline{y}}{}^{\delta_{1}}{}_{\sigma_{1}}\;\Gamma_{\underline{x}\underline{y}}{}^{\sigma_{2}}{}_{\delta_{4}}\right)=+P_{0}{}^{\sigma_{2}\delta_{1}}_{\delta_{4}\sigma_{1}}\,, (107)
Pcσ1δ4δ1σ2\displaystyle P_{c}{}^{\delta_{1}\sigma_{2}}_{\sigma_{1}\delta_{4}} \displaystyle\equiv 12(δδ1δσ2σ1+δ4Γx¯y¯Γx¯y¯δ1σ1)σ2δ4=+Pc,δ4σ1σ2δ1\displaystyle\frac{1}{2}\;\left(\delta^{\delta_{1}}{}_{\sigma_{1}}\;\delta^{\sigma_{2}}{}_{\delta_{4}}+\Gamma_{\underline{x}\underline{y}}{}^{\delta_{1}}{}_{\sigma_{1}}\;\Gamma_{\underline{x}\underline{y}}{}^{\sigma_{2}}{}_{\delta_{4}}\right)=+P_{c}{}^{\sigma_{2}\delta_{1}}_{\delta_{4}\sigma_{1}}\,, (108)
Psσ1δ4δ1σ2\displaystyle P_{s}{}^{\delta_{1}\sigma_{2}}_{\sigma_{1}\delta_{4}} \displaystyle\equiv 12(δδ1Γx¯y¯σ1σ2δ4Γx¯y¯δσ2δ1σ1)δ4=Ps.δ4σ1σ2δ1\displaystyle\frac{1}{2}\;\left(\delta^{\delta_{1}}{}_{\sigma_{1}}\;\Gamma_{\underline{x}\underline{y}}{}^{\sigma_{2}}{}_{\delta_{4}}-\Gamma_{\underline{x}\underline{y}}{}^{\delta_{1}}{}_{\sigma_{1}}\;\delta^{\sigma_{2}}{}_{\delta_{4}}\right)=-P_{s}{}^{\sigma_{2}\delta_{1}}_{\delta_{4}\sigma_{1}}\,. (109)

By inserting this expressions into (104)-(105) we finally obtain from (94) a completely factorized form for the Landau interaction functions, as

fαm;αm(k,k)=fαmk;αmk0+cos(θθ)fαmk;αmkc+sin(θθ)fαmk;αmks,f_{\alpha m;\alpha^{\prime}m^{\prime}}(\vec{k},{\vec{k}}^{\prime})=f_{\alpha mk;\alpha^{\prime}m^{\prime}k^{\prime}}^{0}+\cos(\theta-\theta^{\prime})\;f_{\alpha mk;\alpha^{\prime}m^{\prime}k^{\prime}}^{c}+\sin(\theta-\theta^{\prime})\;f_{\alpha mk;\alpha^{\prime}m^{\prime}{k}^{\prime}}^{s}\,, (110)

written in terms of three independent constant, cosine and sine components, which are given according to

fαmk;αmk0\displaystyle f_{\alpha m{k};\alpha^{\prime}m^{\prime}{k}^{\prime}}^{0} =\displaystyle= Iαmk;δ1δ2αmk;δ3δ4(Tδ3δ4δ1δ2Tδ4δ3δ1δ2+Γx¯y¯(Tσ4δ3σ1δ2Tδ3σ4σ1δ2)δ1σ1Γx¯y¯)σ4δ4\displaystyle I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}}\;\left(T^{\delta_{1}\delta_{2}}_{\delta_{3}\delta_{4}}-T^{\delta_{1}\delta_{2}}_{\delta_{4}\delta_{3}}+\Gamma_{\underline{x}\underline{y}}{}^{\delta_{1}}{}_{\sigma_{1}}\,\left(T^{\sigma_{1}\delta_{2}}_{\sigma_{4}\delta_{3}}-T^{\sigma_{1}\delta_{2}}_{\delta_{3}\sigma_{4}}\right)\,\Gamma_{\underline{x}\underline{y}}{}^{\sigma_{4}}{}_{\delta_{4}}\right) (111)
=\displaystyle= +fαmk;αmk0,\displaystyle+f_{\alpha^{\prime}m^{\prime}{k}^{\prime};\alpha m{k}}^{0}\,, (113)
fαmk;αmkc\displaystyle f_{\alpha m{k};\alpha^{\prime}m^{\prime}{k}^{\prime}}^{c} =\displaystyle= Iαmk;δ1δ2αmk;δ3δ4(Tδ3δ4δ1δ2Tδ4δ3δ1δ2Γx¯y¯(Tσ4δ3σ1δ2Tδ3σ4σ1δ2)δ1σ1Γx¯y¯)σ4δ4\displaystyle I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}}\;\left(T^{\delta_{1}\delta_{2}}_{\delta_{3}\delta_{4}}-T^{\delta_{1}\delta_{2}}_{\delta_{4}\delta_{3}}-\Gamma_{\underline{x}\underline{y}}{}^{\delta_{1}}{}_{\sigma_{1}}\,\left(T^{\sigma_{1}\delta_{2}}_{\sigma_{4}\delta_{3}}-T^{\sigma_{1}\delta_{2}}_{\delta_{3}\sigma_{4}}\right)\,\Gamma_{\underline{x}\underline{y}}{}^{\sigma_{4}}{}_{\delta_{4}}\right) (115)
=\displaystyle= +fαmk;αmkc,\displaystyle+f^{c}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\alpha m{k}}\,, (117)
fαmk;αmks\displaystyle f_{\alpha m{k};\alpha^{\prime}m^{\prime}{k}^{\prime}}^{s} =\displaystyle= Iαmk;δ1δ2αmk;δ3δ4((Tδ3σδ1δ2Tσδ3δ1δ2)Γx¯y¯σδ4Γx¯y¯(Tδ3δ4σδ2Tδ4δ3σδ2)δ1σ)\displaystyle I^{\alpha m{k};\delta_{3}\delta_{4}}_{\alpha^{\prime}m^{\prime}{k}^{\prime};\delta_{1}\delta_{2}}\;\left(\left(T^{\delta_{1}\delta_{2}}_{\delta_{3}\sigma}-T^{\delta_{1}\delta_{2}}_{\sigma\delta_{3}}\right)\,\Gamma_{\underline{x}\underline{y}}{}^{\sigma}{}_{\delta_{4}}\,-\Gamma_{\underline{x}\underline{y}}{}^{\delta_{1}}{}_{\sigma}\,\left(T^{\sigma\delta_{2}}_{\delta_{3}\delta_{4}}-T^{\sigma\delta_{2}}_{\delta_{4}\delta_{3}}\right)\right) (119)
=\displaystyle= fαmk;αmks.\displaystyle-f_{\alpha^{\prime}m^{\prime}{k}^{\prime};\alpha m{k}}^{s}\,. (121)

Notice that we dropped any auxiliary quantity, having obtained a formula for the Landau interaction functions which is completely written in terms of the interaction tensor (67) and the integrals (103) of the form (LABEL:eq:fermion.integrals.all), where the radial functions were defined in equations (35)-(58) in terms of the solutions of equation (50).

Appendix D Pomeranchuk method

We give in this section a brief introduction to Pomeranchuk’s method to detect instabilities in a Landau fermi liquid. We focus on the isotropic, two dimensional case, since it is what concern us in this paper. We begin by explaining the method for a single spinless fermion (section D.1), the generalization for multiple species of spinful fermions is given later (section D.2).

D.1 Single spinless fermion

Given a two-dimensional system of spinless fermion with quasiparticle dispersion relation ϵ(k)\epsilon(k), its Fermi surface is defined in momentum space by the relation ϵ(kF)=0\epsilon(k_{F})=0. In the ground state all the single quasiparticle states with momenta kkFk\leq k_{F} are occupied, while those with momenta k>kFk>k_{F} are empty. In other words, we can write the occupation number as Nk=𝖧(ϵ(k))N_{\vec{k}}={\sf H}(-\epsilon(k)), where 𝖧{\sf H} is the Heavyside step function.

A fermionic excitation can then be represented by a small deformation of the Fermi surface, charactherized by a variation of the occupation numbers δNk\delta N_{\vec{k}}. The excitation energy at weak coupling is then given by Landau’s formula (101)

δΩ=d2kϵ(k)δNk+12d2kd2kf(k,k)δNkδNk,\delta\Omega=\int\;d^{2}\vec{k}\;\epsilon(k)\;\delta N_{\vec{k}}+\frac{1}{2}\int\!\!d^{2}\vec{k}\;d^{2}\vec{k}^{\prime}\;f(\vec{k},\vec{k}^{\prime})\;\delta N_{\vec{k}}\;\delta N_{\vec{k}^{\prime}}\,, (122)

where f(k,k)f(\vec{k},\vec{k}^{\prime}) is the Landau interaction function. The variation δNk\delta N_{\vec{k}} on the occupation numbers of the state take the values δNk=0,±1\delta N_{\vec{k}}=0,\pm 1 at any point with k\vec{k} of momentum space. They can be parametrized as

δNk=𝖧(ϵ(k)+δg(k))𝖧(ϵ(k)),\delta N_{\vec{k}}={\sf H}\big{(}\!-\!\epsilon(k)+\delta g(\vec{k})\big{)}-{\sf H}\left(\!-\epsilon(k)\right)\,, (123)

where δg(k)\delta g(\vec{k}) is an auxiliary function that characterizes the deformation of the Fermi surface. By using the expansion of the Heavyside function in terms of Dirac delta function and its derivatives, we get888This is a shorthand calculation, if the reader is uncomfortable with it, (124) can be smoothed by making the replacement 𝖧(x)1/(1+exp(b2x)){\sf H}(x)\rightarrow 1/(1+\exp(-b^{2}\,x)) and then taking the b2b^{2}\to\infty limit.

δNk=δ(ϵ(k))δg(k)+12δ(ϵ(k))δg(k)2+.\delta N_{\vec{k}}=\delta\big{(}\!-\!\epsilon(k)\big{)}\;\delta g(\vec{k})+\frac{1}{2}\;\delta^{\prime}\big{(}\!-\epsilon(k)\big{)}\;\delta g(\vec{k})^{2}+\dots\,. (124)

When replacing (124) into (122), the Dirac delta functions enforce the integrands to be evaluated at ϵ(k)=0\epsilon(k)=0, namely at the Fermi surface. Using polar coordinates in momentum space, this implies that the integrand must be evaluated at k=kFk=k_{F} and then the momentum integrals reduce to angular ones

δΩ=kFvF𝑑θ12δg(θ)2+kF2vF2𝑑θ𝑑θ12f(θθ)δg(θ)δg(θ),\delta\Omega\!=\frac{k_{F}}{v_{F}}\;\int\!d\theta\,\frac{1}{2}\,\delta g(\theta)^{2}+\frac{k_{F}{}^{2}}{v_{F}{}^{2}}\;\int\!d\theta\;d\theta^{\prime}\;\frac{1}{2}\,f(\theta-\theta^{\prime})\;\delta g(\theta)\;\delta g(\theta^{\prime})\,, (125)

where by using rotational invariance we wrote f(θθ)=f(kF,θ;kF,θ)\;f(\theta-\theta^{\prime})=f(k_{F},\theta;k_{F},\theta^{\prime})\; and δg(θ)=δg(kF,θ)\;\delta g(\theta)=\delta g(k_{F},\theta). Here the Fermi velocity is defined as vF=dϵ(k)/dk|k=kFv_{F}=d\epsilon(k)/dk|_{k=k_{F}}.

A Fourier expansion now allows us to write

f(θθ)=n=0fnccos(n(θθ))+n=1fnssin(n(θθ)),f(\theta-\theta^{\prime})=\sum_{n=0}^{\infty}f_{n}^{c}\;\cos\left(n\,(\theta-\theta^{\prime})\right)+\sum_{n=1}^{\infty}f_{n}^{s}\;\sin\left(n\,(\theta-\theta^{\prime})\right)\,, (126)

and similarly, to parameterize the deformations by the amplitudes {δgnc;δgns}\{\delta g_{n}^{c};\delta g_{n}^{s}\} in the decomposition

δg(θ)=n=0δgnccos(nθ)+n=1δgnssin(nθ).\delta g(\theta)=\sum_{n=0}^{\infty}\delta g_{n}^{c}\;\cos\left(n\,\theta\right)+\sum_{n=1}^{\infty}\delta g_{n}^{s}\;\sin\left(n\,\theta\right)\,. (127)

Replacing (126) and (127) in (125) we obtain:

δΩ=πkFvF(1+2πkFvFf0c)δg0c2+π2kFvFn=1(1+πkFvFfnc)(δgnc+2δgns)2.\delta\Omega=\pi\,\frac{k_{F}}{v_{F}}\;\left(1+2\,\pi\,\frac{k_{F}}{v_{F}}\;f^{c}_{0}\right)\;\delta{g^{c}_{0}}^{2}+\frac{\pi}{2}\,\frac{k_{F}}{v_{F}}\;\sum_{n=1}^{\infty}\,\left(1+\pi\,\frac{k_{F}}{v_{F}}\;f_{n}^{c}\right)\;\left(\delta g_{n}^{c}{}^{2}+\delta g_{n}^{s}{}^{2}\right)\,. (128)

We see that the excitation energy ends up written as a quadratic form in the deformation amplitudes {δgnc;δgns}\{\delta g_{n}^{c};\delta g_{n}^{s}\}. In order to have a stable system, δΩ\delta\Omega has to be positive for any possible excitation, or in other words for any possible amplitudes. This implies that the above defined quadratic form must be positive definite, leading to the stability conditions:

1+2πkFvFf0c\displaystyle 1+2\,\pi\,\frac{k_{F}}{v_{F}}\;f^{c}_{0} >\displaystyle> 0,\displaystyle 0\,, (129)
1+πkFvFfnc\displaystyle 1+\pi\,\frac{k_{F}}{v_{F}}\;f_{n}^{c} >\displaystyle> 0,n.\displaystyle 0\,,\qquad\qquad\forall\,n\in\mathbb{N}\,. (130)

If any of these conditions is violated, the system becomes unstable. Notice that the parameters fnsf_{n}^{s} completely disappear from the calculation. The parameters fncf_{n}^{c} are called the “Landau parameters” of the Fermi liquid.

It is important to stress that, when working in perturbation theory, the dispersion relation is defined by (102) as

ϵ(k)=ω(k)+d2kf(k,k)Nk.\epsilon(k)=\omega(k)+\int\!d^{2}\vec{k}^{\prime}\,f(\vec{k},\vec{k}^{\prime})\,N_{\vec{k}^{\prime}}\,. (131)

This implies that only the zeroth order forms of the Fermi momentum and Fermi velocity have to be kept in equations (129). Indeed, since the second term in those inequalities contains a Landau parameter fncf^{c}_{n}, it is already first order in the coupling constants. Thus we can replace (129) by

1+2πkF𝖿𝗋𝖾𝖾vF𝖿𝗋𝖾𝖾f0c\displaystyle 1+2\,\pi\,\frac{k_{F}^{\sf free}}{v_{F}^{\sf free}}\;f^{c}_{0} >\displaystyle> 0,\displaystyle 0\,, (132)
1+πkF𝖿𝗋𝖾𝖾vF𝖿𝗋𝖾𝖾fnc\displaystyle 1+\pi\,\frac{k_{F}^{\sf free}}{v_{F}^{\sf free}}\;f_{n}^{c} >\displaystyle> 0,n\displaystyle 0\,,\qquad\qquad\forall\,n\in\mathbb{N} (133)

where kF𝖿𝗋𝖾𝖾k_{F}^{\sf free} is obtained from ω(kF𝖿𝗋𝖾𝖾)=0\omega(k_{F}^{\sf free})=0 and we defined vF𝖿𝗋𝖾𝖾=dω(k)/dk|k=kF𝖿𝗋𝖾𝖾v_{F}^{\sf free}=d\omega(k)/dk|_{k=k_{F}^{\sf free}}.

D.2 Multiple fermionic species with spin

The above procedure can be easily generalized to many species of spinful fermions, the resulting quadratic form being in general non-diagonal in the indices denoting spin α\alpha and species mm. As we know, the excitation energy of such fermionic system is given by (101)

δΩ=αmd2kϵm(k)δNαm(k)+12αmαmd2kd2kfαmαm(k,k)δNαm(k)δNαm(k).\delta\Omega=\sum_{\alpha m}\int\!\!d^{2}k\,\epsilon_{m}(k)\;\delta N_{\alpha m}(\vec{k})+\frac{1}{2}\sum_{\alpha m\alpha^{\prime}m^{\prime}}\int\!\!d^{2}k\;d^{2}k^{\prime}\;f_{\alpha m\alpha^{\prime}m^{\prime}}(\vec{k},\vec{k}^{\prime})\;\delta N_{\alpha m}(\vec{k})\;\delta N_{\alpha^{\prime}m^{\prime}}(\vec{k}^{\prime})\,. (134)

When the deformations of the occupation numbers get spin and species indices δNαm(k)\delta N_{\alpha m}(\vec{k}), so do the functions δgαm(k)\delta g_{\alpha m}(\vec{k}) parameterizing the deformations

δNαm(k)=𝖧(ϵm(k)+δgαm(k))𝖧(ϵm(k)).\delta N_{\alpha m}(\vec{k})={\sf H}\big{(}\!-\!\epsilon_{m}(k)+\delta g_{\alpha m}(\vec{k})\big{)}-{\sf H}\left(\!-\epsilon_{m}(k)\right)\,. (135)

Notice that each species and spin component has its own Fermi momentum kFmk_{F}^{m} at which it dispersion relation vanishes ϵm(kFm)=0\epsilon_{m}(k_{F}^{m})=0. Going through the same steps as before we get for the energy fluctuation

δΩ=αmkFmvFm𝑑θ12δgαm(θ)2+12αmαmkFmvFmkFmvFm𝑑θ𝑑θfαmαm(θθ)δgαm(θ)δgαm(θ),\delta\Omega\!=\sum_{\alpha m}\,\frac{k_{F}^{m}}{v_{F}^{m}}\int\!d\theta\;\frac{1}{2}\,\delta g_{\alpha m}(\theta)^{2}+\frac{1}{2}\!\!\sum_{\alpha m^{\prime}\alpha m^{\prime}}\frac{k_{F}^{m}}{v_{F}^{m}}\,\frac{k_{F}^{{}^{\prime}m^{\prime}}}{v_{F}^{{}^{\prime}m^{\prime}}}\!\int\!d\theta\,d\theta^{\prime}\,f_{\alpha m\alpha^{\prime}m^{\prime}}(\theta-\theta^{\prime})\,\delta g_{\alpha m}(\theta)\;\delta g_{\alpha^{\prime}m^{\prime}}(\theta^{\prime}), (136)

where fαmαm(θθ)=fαmαm(kFαm,θ;kFαm,θ)\;f_{\alpha m\alpha^{\prime}m^{\prime}}(\theta\!-\!\theta^{\prime})=f_{\alpha m\alpha^{\prime}m^{\prime}}(k_{F}^{\alpha m},\theta;\,k_{F}^{\alpha^{\prime}m^{\prime}},\theta^{\prime}) and vFm=dϵm(k)/dk|k=kFmv_{F}^{m}=d\epsilon_{m}(k)/dk|_{k=k_{F}^{m}}.

Now we must decompose the interaction function as in (126)

fαmαm(θθ)=n=0fαmαmnccos(n(θθ))+n=1fαmαmnssin(n(θθ)).f_{\alpha m\alpha^{\prime}m^{\prime}}(\theta-\theta^{\prime})=\sum_{n=0}^{\infty}{f_{\alpha m\alpha^{\prime}m^{\prime}}}_{n}^{c}\;\cos\left(n\,(\theta-\theta^{\prime})\right)+\sum_{n=1}^{\infty}{f_{\alpha m\alpha^{\prime}m^{\prime}}}_{n}^{s}\;\sin\left(n\,(\theta-\theta^{\prime})\right)\,. (137)

Although from (110) we can already see that in our case only the modes n=0,1n=0,1 are present in the decomposition (137), we mantain the general analysis for now. The deformation parameters can be decomposed in parallel with (127) to get

δgαm(θ)=n=0δgαmnccos(nθ)+n=1δgαmnssin(nθ),\delta g_{\alpha m}(\theta)=\sum_{n=0}^{\infty}{\delta g_{\alpha m}}_{n}^{c}\;\cos\left(n\,\theta\right)+\sum_{n=1}^{\infty}{\delta g_{\alpha m}}_{n}^{s}\;\sin\left(n\,\theta\right)\,, (138)

in terms of the deformation amplitudes {δgαmnc,δgαmns}\{{\delta g_{\alpha m}}_{n}^{c},{\delta g_{\alpha m}}_{n}^{s}\}.

Going ahead to plug the decompositions (137) and (138) into (136), we get the generalization of (128) to be

δΩ\displaystyle\delta\Omega\! =\displaystyle= αmαmπkFmvFm(δαmαm+2πkFmvFmfαmαm0c)δgαm0cδgαm0c+\displaystyle\!\sum_{\alpha m\alpha^{\prime}m^{\prime}}\,\pi\,\frac{k_{F}^{m}}{v_{F}^{m}}\;\left(\delta_{\alpha m\alpha^{\prime}m^{\prime}}+2\,\pi\,\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}\;{f_{\alpha m\alpha^{\prime}m^{\prime}}}^{c}_{0}\right)\;\delta{g_{\alpha m}}^{c}_{0}\;{\delta g_{\alpha^{\prime}m^{\prime}}}^{c}_{0}+ (141)
+n=1αmαmπ2kFmvFm(δαmαm+πkFmvFmfαmαmnc)(δgαmncδgαmnc+δgαmnsδgαmns)+\displaystyle+\sum_{n=1}^{\infty}\,\sum_{\alpha m\alpha^{\prime}m^{\prime}}\frac{\pi}{2}\,\frac{k_{F}^{m}}{v_{F}^{m}}\;\left(\delta_{\alpha m\alpha^{\prime}m^{\prime}}+\pi\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}{f_{\alpha m\alpha^{\prime}m^{\prime}}}_{n}^{c}\right)\left({\delta g_{\alpha m}}_{n}^{c}\;{\delta g_{\alpha^{\prime}m^{\prime}}}_{n}^{c}+{\delta g_{\alpha m}}_{n}^{s}\;{\delta g_{\alpha^{\prime}m^{\prime}}}_{n}^{s}\right)+
+n=1αmαmπ22kFmvFmkFmvFmfαmαmns(δgαmnsδgαmncδgαmncδgαmns).\displaystyle+\sum_{n=1}^{\infty}\,\sum_{\alpha m\alpha^{\prime}m^{\prime}}\;\frac{\pi^{2}}{2}\,\frac{k_{F}^{m}}{v_{F}^{m}}\;\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}\;{f_{\alpha m\alpha^{\prime}m^{\prime}}}_{n}^{s}\;\left({\delta g_{\alpha m}}_{n}^{s}\;{\delta g_{\alpha^{\prime}m^{\prime}}}_{n}^{c}-{\delta g_{\alpha m}}_{n}^{c}\;{\delta g_{\alpha^{\prime}m^{\prime}}}_{n}^{s}\right)\,.

On a stable state, this quadratic form has to be positive definite. Neccesary and sufficient conditions for stability can be obtained by considering the following two disantangled cases.

  • Putting δgαm0c0{\delta g_{\alpha m}}^{c}_{0}\neq 0 and δgαmnc=δgαmns=0\;{\delta g_{\alpha m}}_{n}^{c}=\delta{g_{\alpha m}}_{n}^{s}=0 for n1n\geq 1, we have that the first line in (141) must be positive definite, that applying Sylvester’s criterion is equivalent to,

    |kFmvFm(δαmα¯m¯+2πkFmvFmfαmαm0c)|M×M>0\displaystyle\left|\frac{k_{F}^{m}}{v_{F}^{m}}\;\left(\delta_{\alpha m\bar{\alpha}\bar{m}}+2\,\pi\,\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}\;{f_{\alpha m\alpha^{\prime}m^{\prime}}}^{c}_{0}\right)\right|_{M\times M}>0\; , M,\displaystyle\;\forall M\in\mathbb{N}\,, (142)

    where ||M×M|\cdots|_{M\times M} stands for the MM-th minor, that is the determinant of the M×MM\times M upper-left submatrix.

  • Considering now the opposite case: δgαm0c=0{\delta g_{\alpha m}}^{c}_{0}=0 and δgαmnc0,δgαmns0{\delta g_{\alpha m}}_{n}^{c}\neq 0,\delta{g_{\alpha m}}_{n}^{s}\neq 0 for n1n\geq 1, we have to impose the positivity of the quadratic form: δΩ=n=1δΩn\delta\Omega=\sum_{n=1}^{\infty}\delta\Omega_{n} defined by the second and third lines in (141) for each nn separetely, since they do not couple. To simplify the analysis, we find convenient to introduce the perturbation vectors in indices (αm)(\alpha m): un(δgαm)nc\vec{u}_{n}\equiv(\delta g_{\alpha m}{}_{n}^{c}) and vn(δgαm)ns\vec{v}_{n}\equiv(\delta g_{\alpha m}{}_{n}^{s}). Then we can write,

    δΩn=un𝐒𝐧tun+vn𝐒𝐧tvnun𝐀𝐧tvn+vn𝐀𝐧tun,\delta\Omega_{n}=\vec{u}_{n}{}^{t}\;{\bf S_{n}}\;\vec{u}_{n}+\vec{v}_{n}{}^{t}\;{\bf S_{n}}\;\vec{v}_{n}-\vec{u}_{n}{}^{t}\;{\bf A_{n}}\;\vec{v}_{n}+\vec{v}_{n}{}^{t}\;{\bf A_{n}}\;\vec{u}_{n}\,, (143)

    where the matrices 𝐒𝐧=𝐒𝐧t{\bf S_{n}}={\bf S_{n}}{}^{t} and 𝐀𝐧=𝐀𝐧t{\bf A_{n}}=-{\bf A_{n}}{}^{t} are read from (141). By further defining the complex vectors znun+ivn\vec{z}_{n}\equiv\vec{u}_{n}+i\;\vec{v}_{n}\;, (143) takes the simple form

    δΩn=zn𝐇𝐧zn;𝐇𝐧=𝐒𝐧+i𝐀𝐧.\delta\Omega_{n}=\vec{z}_{n}{}^{\dagger}\;{\bf H_{n}}\;\vec{z}_{n}\qquad;\qquad{\bf H_{n}}={\bf S_{n}}+i\;{\bf A_{n}}\,. (144)

    From here it is clear that the positivity condition of δΩ\delta\Omega is just that the hermitian matrices 𝐇𝐧=𝐇𝐧{\bf H_{n}}={\bf H_{n}}{}^{\dagger} must be positive, for any nn\in\mathbb{N}. By using the explicit expressions of 𝐒𝐧{\bf S_{n}} and 𝐀𝐧{\bf A_{n}}, this is equivalent to ask that

    |kFmvFm(δαmαm+πkFmvFm(fαmαmnc+ifαmαmns))|M×M>0,M.\left|\frac{k_{F}^{m}}{v_{F}^{m}}\;\left(\delta_{\alpha m\alpha^{\prime}m^{\prime}}+\pi\frac{k_{F}^{m^{\prime}}}{v_{F}^{m^{\prime}}}\,\left({f_{\alpha m\alpha^{\prime}m^{\prime}}}_{n}^{c}+i{f_{\alpha m\alpha^{\prime}m^{\prime}}}_{n}^{s}\right)\right)\,\right|_{M\times M}>0\quad,\quad\forall M\in\mathbb{N}\,. (145)

In conclusion, if any of the minors in (142) and (145) is negative, the quadratic form has a negative mode that would lead to an excitation with negative energy: δΩ<0\,\delta\Omega<0, thus triggering an instability on the fermionic system.

D.3 Summary and application to the holographic setup

In summary, in order to perform the stability analysis, all we need are

  • The value of the free Fermi momenta kFm𝖿𝗋𝖾𝖾k_{F}^{m\;{\sf free}}, obtained from ωm(kFm𝖿𝗋𝖾𝖾)=0\omega_{m}(k_{F}^{m\;{\sf free}})=0. Indeed, notice that, since kFmk_{F}^{m} is always positive, we can remove it from the pre-factor in (142) and (145). Regarding the second term, since kFmk_{F}^{m} is multiplied by the interaction function, we can replace it by kFm𝖿𝗋𝖾𝖾k_{F}^{m\;{\sf free}}.

  • The value of the free Fermi velocities vFm𝖿𝗋𝖾𝖾=dωm(k)/dk|k=kFm𝖿𝗋𝖾𝖾v_{F}^{m\;{\sf free}}=\left.d\omega_{m}(k)/dk\right|_{k=k_{F}^{m\;{\sf free}}}, as long as we assume that vFmv_{F}^{m} is always positive in other to remove it from the overall pre-factor in (142) and (145).

  • The Fourier components of the Landau parameters, evaluated at the free Fermi momenta kFm𝖿𝗋𝖾𝖾k_{F}^{m\;{\sf free}}. Notice that equation (110) implies that only the modes with n=0,1n=0,1 enter into the stability analysis in the holographic setup. The Landau parameters are obtained from (111) in terms of the integrals (103) evaluated at the free Fermi momenta. Since ωm(kFm𝖿𝗋𝖾𝖾)=0\omega_{m}(k_{F}^{m\;{\sf free}})=0, we need only static wave functions in (103).

Appendix E The electron star background

In this appendix, we summarize the construction of the electron star solution. We do so in order to fix notation and to have a fully self-contained discussion. However, all the results presented in this section were obtained in the original reference Hartnoll:2010gu to which the interest reader is referred.


We work with the previously presented Ansatz (14)

ds2\displaystyle ds^{2} =\displaystyle= L2(f(z)dt2+g(z)dz2+dx2z2),\displaystyle L^{2}\left(-f(z)\,dt^{2}+g(z)\,dz^{2}+\frac{d\vec{x}^{2}}{z^{2}}\right)\,, (146)
A\displaystyle A =\displaystyle= h(z)dt.\displaystyle h(z)\,dt\,. (147)

The field equations of (3+1)(3+1)-dimensional Einstein-Maxwell theory in the presence of a negative cosmological constant 3/L2-3/L^{2} plus matter in the form of a charged perfect fluid are

RMN12gMNR3L2gMN\displaystyle R_{MN}-\frac{1}{2}\,g_{MN}\,R-\frac{3}{L^{2}}\,g_{MN} =\displaystyle= κ2(TMN𝖬𝖺𝗑𝗐𝖾𝗅𝗅+TMN𝖥𝗅𝗎𝗂𝖽),\displaystyle\kappa^{2}\,\left(T^{\sf Maxwell}_{MN}+T^{\sf Fluid}_{MN}\right)\,, (148)
NFMN\displaystyle\nabla_{N}F^{MN} =\displaystyle= e2JM,\displaystyle e^{2}\,J^{M}\,, (149)

with κ\kappa and e2e^{2} the gravitational and electromagnetic couplings respectively. Here the contributions to the energy-momentum tensor and current density read

TMN𝖬𝖺𝗑𝗐𝖾𝗅𝗅\displaystyle T^{\sf Maxwell}_{MN} =\displaystyle= 1e2(FMPFNP14gMNFPQFPQ),\displaystyle\frac{1}{e^{2}}\,\left(F_{MP}\,F_{N}^{P}-\frac{1}{4}\,g_{MN}\,F_{PQ}\,F^{PQ}\right)\,, (150)
TMN𝖥𝗅𝗎𝗂𝖽\displaystyle T^{\sf Fluid}_{MN} =\displaystyle= (ρ+p)uMuN+pgMN,\displaystyle(\rho+p)\,u_{M}\,u_{N}+p\,g_{MN}\,, (151)
JM\displaystyle J_{M} =\displaystyle= σuM,\displaystyle\sigma\,u_{M}\,, (152)

The functions pp, ρ\rho and σ\sigma are the pressure, energy density and charge density of the fluid, respectively, and uMu_{M} its four-velocity. Furthermore, the conservation equations must hold

NTMN=0;MJM=0.\nabla^{N}T_{MN}=0\qquad;\qquad\nabla_{M}J^{M}=0\,. (153)

Since we aim to describe a relativistic system of fermions of mass mm, we can try a description of the energy and charge densities in terms of the density of states in flat space of a free fermion gas g(E)=π2E(E2m2)1/2\;g(E)=\pi^{-2}\,E\,(E^{2}-m^{2})^{1/2}. This description would be accurate as long as the wavelength of the fermions is much shorter than the local curvature radius. Furthermore, we must consider the equation of state of the system in terms of a grand canonical potential Ω=pV\Omega=-p\,V. At zero temperature the fluid functions read

σ\displaystyle\sigma =\displaystyle= mεF𝑑Eg(E)=1π2mϵF𝑑EE(E2m2)1/2,\displaystyle\int_{m}^{\varepsilon_{F}}dE\,g(E)=\frac{1}{\pi^{2}}\,\int_{m}^{\epsilon_{F}}dE\,E\,(E^{2}-m^{2})^{1/2}\,, (154)
ρ\displaystyle\rho =\displaystyle= mεF𝑑Eg(E)E=1π2mϵF𝑑EE2(E2m2)1/2,\displaystyle\int_{m}^{\varepsilon_{F}}dE\,g(E)\,E=\frac{1}{\pi^{2}}\,\int_{m}^{\epsilon_{F}}dE\,E^{2}\,(E^{2}-m^{2})^{1/2}\,, (155)
p\displaystyle p =\displaystyle= ρ+εFσ.\displaystyle-\rho+\varepsilon_{F}\,\sigma\,. (156)

where εF\varepsilon_{F} is the bulk Fermi energy. Consistency with Ansatz (146) imposes that all the functions be dependent only on zz. Moreover, the velocity of the fluid must be unitary, which identifies it with the vierbein temporal vector, i.e. uMMet¯=1/(Lf)u^{M}\,\partial_{M}\equiv e_{\underline{t}}=1/(L\,\sqrt{f}), see (16). This last fact implies that the bulk Fermi energy coincides with the local (i.e. measured by comoving observers) chemical potential εFh/Lf\;\varepsilon_{F}\equiv{h}/L{\sqrt{f}}.

It is convenient to work with the dimensionless quantities

p^=κ2L2p,ρ^=κ2L2ρ,σ^=eκL2σ,h^=h/γ.\hat{p}=\kappa^{2}\,L^{2}\,p\qquad,\qquad\hat{\rho}=\kappa^{2}\,L^{2}\,\rho\qquad,\qquad\hat{\sigma}=e\,\kappa\,L^{2}\,\sigma\qquad,\qquad\hat{h}=h/\gamma\,. (157)

Here we introduced γeL/κ\gamma\equiv{e\,L}/{\kappa}. In terms of these scaled variables (154) can be integrated to obtain the explicit expressions

σ^\displaystyle\hat{\sigma} =\displaystyle= β2^3(h^2fm^2)32,\displaystyle\frac{\hat{\beta^{2}}}{3}\,\left(\frac{\hat{h}^{2}}{f}-\hat{m}^{2}\right)^{\frac{3}{2}}\,, (158)
ρ^\displaystyle\hat{\rho} =\displaystyle= β^28[h^2fm^2(m^2h^f+2h^3f32)m^4ln(h^f+h^2fm^2)+m^4lnm^],\displaystyle\frac{\hat{\beta}^{2}}{8}\,\left[\sqrt{\frac{\hat{h}^{2}}{f}-\hat{m}^{2}}\;\left(-\hat{m}^{2}\,\frac{\hat{h}}{\sqrt{f}}+2\,\frac{\hat{h}^{3}}{f^{\frac{3}{2}}}\right)-\hat{m}^{4}\,\ln\left(\frac{\hat{h}}{\sqrt{f}}+\sqrt{\frac{\hat{h}^{2}}{f}-\hat{m}^{2}}\right)+\hat{m}^{4}\,\ln\hat{m}\right]\,, (159)
p^\displaystyle\hat{p} =\displaystyle= ρ^+h^fσ^,\displaystyle-\hat{\rho}+\frac{\hat{h}}{\sqrt{f}}\,\hat{\sigma}\,, (160)

where we introduced the two independent parameters β^\hat{\beta} and m^\hat{m} that can be used to define the electron star

β^eγπ,m^mLγ.\hat{\beta}\equiv\frac{e\,\gamma}{\pi}\qquad,\qquad\hat{m}\equiv\frac{m\,L}{\gamma}\,. (161)

By plugging (146) and (158) in (148) we get the field equations

zff+zgg+σ^z2gh^f+4\displaystyle\frac{z\,f^{\prime}}{f}+\frac{z\,g^{\prime}}{g}+\hat{\sigma}\,z^{2}\,g\,\frac{\hat{h}}{\sqrt{f}}+4 =\displaystyle= 0,\displaystyle 0\,, (162)
zffz2h^22f+(3+p^)z2g1\displaystyle\frac{z\,f^{\prime}}{f}-\frac{z^{2}\,\hat{h}^{\prime 2}}{2\,f}+(3+\hat{p})\,z^{2}\,g-1 =\displaystyle= 0,\displaystyle 0\,, (163)
z2h^′′h+σ^z2gh^f(zh^2h^fh^2)\displaystyle\frac{z^{2}\,\hat{h}^{\prime\prime}}{h}+\hat{\sigma}\,z^{2}\,g\,\frac{\hat{h}}{\sqrt{f}}\,\left(\frac{z\,\hat{h}^{\prime}}{2\,\hat{h}}-\frac{f}{\hat{h}^{2}}\right) =\displaystyle= 0.\displaystyle 0\,. (164)

These equations have to be solved numerically for each value of m^\hat{m} and β^\hat{\beta} to obtain the metric and gauge functions (f,g,h^)(f,g,\hat{h}).

Regarding the boundary conditions in the IR, we observe from (158) that the electron star extends from infinity to the radius zsz_{s} such that

p^(zs)=ρ^(zs)=σ^(zs)=0,TMN𝖥𝗅𝗎𝗂𝖽(zs)=JM(zs)=0,\hat{p}(z_{s})=\hat{\rho}(z_{s})=\hat{\sigma}(z_{s})=0\,,\qquad\Longleftrightarrow\qquad T^{\sf Fluid}_{MN}(z_{s})=J_{M}(z_{s})=0\,, (165)

whose position is determined by the equation

h^(zs)2m^2f(zs)=0.\hat{h}(z_{s})^{2}-\hat{m}^{2}\;f(z_{s})=0\,. (166)

Outside the electron star, i.e. 0<z<zs0<z<z_{s}, we have that σ=p=ρ=0\sigma=p=\rho=0, implying that the equations of motion (162) reduce to

ff+gg+4z=0;1zffh^22f+3g1z2=0;h^′′=0,\frac{f^{\prime}}{f}+\frac{g^{\prime}}{g}+\frac{4}{z}=0\qquad;\qquad\frac{1}{z}\,\frac{f^{\prime}}{f}-\frac{\hat{h}^{\prime 2}}{2\,f}+3\,g-\frac{1}{z^{2}}=0\qquad;\qquad\hat{h}^{\prime\prime}=0\,, (167)

whose general solution is the AdS Reissner-Nördstrom black hole

f(z)=c2z2M^z+Q^22z2,g(z)=c2z4f(z),h^(z)=μ^Q^z.f(z)=\frac{c^{2}}{z^{2}}-\hat{M}\,z+\frac{\hat{Q}^{2}}{2}\,z^{2}\qquad,\qquad g(z)=\frac{c^{2}}{z^{4}\,f(z)}\qquad,\qquad\hat{h}(z)=\hat{\mu}-\hat{Q}\,z\,. (168)

To specify this solution we have to give the four integration constants. (c,M^,μ^,Q^)(c,\hat{M},\hat{\mu},\hat{Q}). In the electron star context they must be obtained by matching (f,g,h^,h^)(f,g,\hat{h},\hat{h}^{\prime}) at the radius z=zsz=z_{s} with the solution inside the star.


On the other hand, in the IR limit zz\rightarrow\infty the solution acquires the form of a Lifshitz metric. In fact, an Ansatz for large zz of the form:

f=1z2λ(1+f1zα+),g=gz2(1+g1zα+),h^=hzλ(1+h1zα+),f=\frac{1}{z^{2\lambda}}\,\left(1+f_{1}\,z^{\alpha}+\dots\right)\quad,\quad g=\frac{g_{\infty}}{z^{2}}\,\left(1+g_{1}\,z^{\alpha}+\dots\right)\quad,\quad\hat{h}=\frac{h_{\infty}}{z^{\lambda}}\,\left(1+h_{1}\,z^{\alpha}+\dots\right)\,, (169)

solves (162) if α\alpha takes some of the following three values

α0=2+λ,α±=1+2λ±129λ321λ2+40λ28m^2λ(43λ)2(1m^2)λ1.\alpha_{0}=2+\lambda\qquad,\qquad\alpha_{\pm}=1+\frac{2}{\lambda}\pm\frac{1}{2}\,\sqrt{\frac{9\lambda^{3}-21\lambda^{2}+40\lambda-28-\hat{m}^{2}\lambda(4-3\lambda)^{2}}{(1-\hat{m}^{2})\lambda-1}\,.} (170)

The parameter λ\lambda is called the dynamical critical exponent. One can show that an IR non-singular, asymptotically exact Lifshitz solution exists only if

111m^2λ,1\leq\frac{1}{1-\hat{m}^{2}}\leq\lambda\,, (171)

and the root selected is the negative one: α=α\alpha=\alpha_{-}, see Hartnoll:2010gu for details. From the leading order of (162) we get,

g2\displaystyle g^{2}_{\infty} =\displaystyle= 36λ4(λ1)β^4(λ(1m^2)1)3,\displaystyle\frac{36\,\lambda^{4}\,(\lambda-1)}{\hat{\beta}^{4}\,\left(\lambda\,(1-\hat{m}^{2})-1\right)^{3}}\,, (172)
h2\displaystyle h^{2}_{\infty} =\displaystyle= 11λ,\displaystyle 1-\frac{1}{\lambda}\,, (173)
g(3+p^())\displaystyle g_{\infty}\,(3+\hat{p}(\infty)) =\displaystyle= 1+2λ+h22λ2.\displaystyle 1+2\,\lambda+\frac{h_{\infty}{}^{2}}{2}\,\lambda^{2}\,. (174)

The first two equations determine the coefficients that define the Lifshitz metric in terms of λ\lambda and m^\hat{m}, while that the last equation relates implicitly β^\hat{\beta} in terms of λ\lambda and m^\hat{m} also. So it is possible, and convenient to make numerics, to think the Lifshitz parameter and the mass as the free parameters of the electron star. The expansions (169) result completely determined by (λ,m^\lambda,\hat{m}), and by f1f_{1} that remains free. However, the scalings (t,x,y,z)(bt,ax,ay,az)(t,x,y,z)\rightarrow(b\,t,a\,x,a\,y,a\,z) leave the Ansatz (146) invariant and induce the invariance of the system (162) under

zaz\displaystyle z\rightarrow a\,z\quad , ga2g,\displaystyle\quad g\rightarrow a^{-2}\,g\,, (175)
fb2f\displaystyle f\rightarrow b^{-2}\,f\quad , h^b1h^.\displaystyle\quad\hat{h}\rightarrow b^{-1}\,\hat{h}\,. (176)

These scalings allow to fix the leading order term of ff in (169) together with the absolute value of f1f_{1}, while the sign to get the solution with the right behaviour in the UV results to be the negative one. In the paper we work out the electron star solution by fixing f1=1f_{1}=-1.

In figure 3 we present the electron star background solution obtained after solving numerically the system (162). We can see how the electron star solution matches the AdS-RN and Lifshitz solutions in the UV (z/zs1z/z_{s}\ll 1) and IR (z/zs1z/z_{s}\gg 1) respectively.

Refer to caption
Refer to caption
Refer to caption
Figure 3: We plot the electron star background functions ff, gg and hh vs. z/zsz/z_{s} (light blue line), compared to the Lifshitz solution (orange) and AdS Reissner-Nördstrom black hole (lihgt green) respectively. The parameters are chosen as m^=0.4\hat{m}=0.4 and λ=2\lambda=2.

An important point to be stressed is that, in the weak gravity regime both κL\kappa\ll L and κ1/m\kappa\ll 1/m conditions should hold, or what is the same, e2πβ^e^{2}\ll\pi\,\hat{\beta} and e1/m^e\ll 1/{\hat{m}}. On the other hand, in Hartnoll:2010gu was shown that the flat treatment (158) is consistent only if we restrict the electron star parameters to the region β^1,m^1\,\hat{\beta}\sim 1,\,\hat{m}\sim 1. Together with the weak gravity conditions, they imply that

γ1emL1.\gamma\sim\frac{1}{e}\sim m\,L\gg 1\,. (177)

We will see in appendix F that these conditions are closely linked to the validity of the WKB approximation for the fermionic perturbations.

Appendix F The WKB solution

In the electron star background, we are assuming that we have a large number of particles inside one AdS radius. This is equivalent to the statement that the particle wavelength is much shorter than the characteristic length of the background in which it is moving. This implies that we can solve the effective Schrödinger equation (43) (or equivalently (50)) in the Wentzel-Kramers-Brillouin (WKB) approximation, as we do in the rest of this section.

F.1 Basics of the WKB approximation

Let us remember basic facts about the WKB approximation (see for example merzbacher1998quantum ). Let us consider the one-dimensional Schrödinger equation written in the form

ϕ′′(z)+U(z)ϕ(z)=0.-\phi^{\prime\prime}(z)+U(z)\;\phi(z)=0\,. (178)

We will assume that the potential is repulsive and positive at the boundary, implying that at zz close enough to z=0z=0 we have U(z)>0U(z)>0 and U(z)<0U^{\prime}(z)<0. As we move into positive values of zz we have a first turning point z0z_{0} at which U(z0)=0U(z_{0})=0 and U(z0)<0U^{\prime}(z_{0})<0. At larger values of zz additional turning points zrz_{r} appear at which U(zr)=0U(z_{r})=0.

Away from any given turning point, we introduce the functions

u±(z;zr)=|U(z)|14exp(±zrz𝑑zU(z)).u_{\pm}(z;z_{r})=|U(z)|^{-\frac{1}{4}}\;\exp{\left(\pm\int_{z_{r}}^{z}dz\,\sqrt{U(z)}\right)}\,. (179)

As can be checked by direct substitution, they are good approximate solutions of (178) as long as U(z)U(z) is large enough, which means in particular that we can use them away from the turning points. On the other hand, close to any turning point we can linearize the potential, and the solution is then written in terms of Airy functions AiA\mbox{\scriptsize$i$\normalsize} and BiB\mbox{\scriptsize$i$\normalsize}. Then the WKB approximate solution around zrz_{r} takes the form

ϕ𝖶𝖪𝖡(z)={L+(r)u+(z;zr)+L(r)u(z;zr),zrz|U(zr)|13A+(r)Ai(w)+A(r)Bi(w)|w=U(zr)13(zzr),|zzr||2U(zr)U′′(zr)|R+(r)u+(z;zr)+R(r)u(z;zr),zzr|U(zr)|13\displaystyle\phi^{\sf WKB}(z)=\left\{\begin{array}[]{lcrcl}L^{(r)}_{+}\;u_{+}(z;z_{r})+L^{(r)}_{-}\;u_{-}(z;z_{r})&,&\qquad z_{r}-z&\gg&|U^{\prime}(z_{r})|^{-\frac{1}{3}}\\ \\ A^{(r)}_{+}\;A\mbox{\scriptsize$i$\normalsize}(w)+A^{(r)}_{-}\;B\mbox{\scriptsize$i$\normalsize}(w)\big{|}_{w=U^{\prime}(z_{r})^{-\frac{1}{3}}\,(z-z_{r})}&,&|z-z_{r}|&\ll&\left|\frac{2\,U^{\prime}(z_{r})}{U^{\prime\prime}(z_{r})}\right|\\ \\ R^{(r)}_{+}\;u_{+}(z;z_{r})+R^{(r)}_{-}\;u_{-}(z;z_{r})&,&\qquad z-z_{r}&\gg&|U^{\prime}(z_{r})|^{-\frac{1}{3}}\end{array}\right.\qquad (185)

where L±(r),R±(r)L_{\pm}^{(r)},R_{\pm}^{(r)} and A±(r)A_{\pm}^{(r)} are numerical constants to be determined in order to ensure the boundary conditions and the continuity around each turning point.

The coefficients of the functions u±(x,xr)u_{\pm}(x,x_{r}) at left and right of a given turning point are linearly related, as in

(R+(r)R(r))=𝐌(zr)(L+(r)L(r)),\left(\begin{array}[]{c}R^{(r)}_{+}\\ R^{(r)}_{-}\end{array}\right)={\bf M}(z_{r})\;\left(\begin{array}[]{c}L^{(r)}_{+}\\ L^{(r)}_{-}\end{array}\right)\,, (186)

where the explicit form of the matrix 𝐌(zr){\bf M}(z_{r}) is obtained by matching the functions u±(z;zr)u_{\pm}(z;z_{r}) across the turning point using the intermediate Airy form of the solution. It has the expression 𝐌(zr)=𝐌{\bf M}(z_{r})={\bf M} when U(zr)>0U^{\prime}(z_{r})>0, and 𝐌(zr)=𝐌{\bf M}(z_{r})={\bf M^{\dagger}} when U(zr)<0U^{\prime}(z_{r})<0, with

𝐌e+iπ4(1ii212).{\bf M}\equiv e^{+i\frac{\pi}{4}}\;\left(\begin{array}[]{cc}1&-i\\ -\frac{i}{2}&\frac{1}{2}\end{array}\right)\,. (187)

The relation between the approximate WKB solutions around any pair of successive turning points can be found by shifting the limits of integration in (179) from zrz_{r} to rr+1r_{r+1}

u±(z;zr)\displaystyle u_{\pm}(z;z_{r}) =\displaystyle= φr±u±(z;zr+1),\displaystyle\varphi_{r}^{\pm}\;u_{\pm}(z;z_{r+1})\,, (188)

in terms of the connection coefficients φr±\varphi_{r}^{\pm}, which read

φr±\displaystyle\varphi_{r}^{\pm} =\displaystyle= exp(±zrzr+1𝑑zU(z)).\displaystyle\exp{\left(\pm\int_{z_{r}}^{z_{r+1}}dz\,\sqrt{U(z)}\right)}\,. (189)

Compatibility with the form (185) for the solution around each turning point thus implies the additional linear relation

(L+(r+1)L(r+1))=𝐖(zr)(R+(r)R(r)),\left(\begin{array}[]{c}L^{(r+1)}_{+}\\ L^{(r+1)}_{-}\end{array}\right)={\bf W}(z_{r})\;\left(\begin{array}[]{c}R^{(r)}_{+}\\ R^{(r)}_{-}\end{array}\right)\,, (190)

in terms of the connection matrix

𝐖(zr)=(φr+00φr).{\bf W}(z_{r})=\left(\begin{array}[]{cc}\varphi_{r}^{+}&0\\ 0&\varphi_{r}^{-}\end{array}\right)\,. (191)

With all the above, we can express the whole set of coefficients of the solution around each of the turning points, in terms of L+(0)L^{(0)}_{+} and L(0)L^{(0)}_{-}, as follows

(L+(r)L(r))\displaystyle\left(\begin{array}[]{c}L^{(r)}_{+}\\ L^{(r)}_{-}\end{array}\right) =\displaystyle= 𝐖(zr1)𝐌(zr1)𝐖(z0)𝐌(z0)(L+(0)L(0)),\displaystyle{\bf W}(z_{r-1})\;{\bf M}(z_{r-1})\;\dots\bf W(z_{0})\;{\bf M}(z_{0})\;\left(\begin{array}[]{c}L^{(0)}_{+}\\ L^{(0)}_{-}\end{array}\right)\,, (196)
(R+(r1)R(r1))\displaystyle\left(\begin{array}[]{c}R^{(r-1)}_{+}\\ R^{(r-1)}_{-}\end{array}\right) =\displaystyle= 𝐌(zr1)𝐖(zr2)𝐌(zr2)𝐖(z0)𝐌(z0)(L+(0)L(0)).\displaystyle{\bf M}(z_{r-1})\;{\bf W}(z_{r-2})\;{\bf M}(z_{r-2})\;\dots\bf W(z_{0})\;{\bf M}(z_{0})\;\left(\begin{array}[]{c}L^{(0)}_{+}\\ L^{(0)}_{-}\end{array}\right)\,. (201)

From this, we can write the relation between the initial (leftmost) and final (rightmost) coefficients for a system with a total of RR turning points. It takes the form

(R+(R1)R(R1))=𝐕(L+(0)L(0)),\left(\begin{array}[]{c}R^{(R-1)}_{+}\\ R^{(R-1)}_{-}\end{array}\right)={\bf V}\;\left(\begin{array}[]{c}L^{(0)}_{+}\\ L^{(0)}_{-}\end{array}\right)\,, (202)

where

𝐕=(V11V12V21V22)=𝐌(zR1)𝐖(zR2)𝐌(zR2)𝐖(z0)𝐌(z0).{\bf V}=\left(\begin{array}[]{cc}V_{11}&V_{12}\\ V_{21}&V_{22}\end{array}\right)={\bf M}(z_{R-1})\;{\bf W}(z_{R-2})\;{\bf M}(z_{R-2})\;\dots\bf W(z_{0})\;{\bf M}(z_{0})\,. (203)

In the present context, Bohr-Sommerfeld quantization relations arise when we impose boundary conditions at both extremes of the zz axis. This leads to constraints on the matrix elements of 𝐕{\bf V} and consequently on the parameters of the theory.

F.2 Application to the electron star background

After the re-scalings ω=γω^,k=γk^\omega=\gamma\,\hat{\omega},\;k=\gamma\,\hat{k}, together with (157) and (161), the potential (44) entering into the effective Schrödinger equation acquires the following form

U(z)\displaystyle U(z) =\displaystyle= γ2g(z)(k^2z2(ω^+qh^(z))2f(z)+m^2)+γm^g(z)(ln(fk+(z)g(z)))+(fk+(z))′′fk+(z).\displaystyle\gamma^{2}\,g(z)\,\left(\hat{k}^{2}\,z^{2}-\frac{(\hat{\omega}+q\,\hat{h}(z))^{2}}{f(z)}+\hat{m}^{2}\right)+\gamma\,\hat{m}\,\sqrt{g(z)}\;\left(\ln\left(f_{k}^{+}(z)\,\sqrt{g(z)}\right)\right)^{\prime}+\frac{\left(\sqrt{f_{k}^{+}(z)}\right)^{\prime\prime}}{\sqrt{f_{k}^{+}(z)}}. (204)

As we will be working in an electron star background, we have to consider the large γ\gamma limit. From (204) we then see that this regime coincides with the semiclassical limit where the WKB approximation is reliable. In such limit we can neglect the last two terms in (204) and consider

U(z)γ2g(z)(k^2z2(ω^+qh^(z))2f(z)+m^2).U(z)\approx\gamma^{2}\,g(z)\,\left(\hat{k}^{2}\,z^{2}-\frac{(\hat{\omega}+q\,\hat{h}(z))^{2}}{f(z)}+\hat{m}^{2}\right)\,. (206)

Close to the AdS boundary z=0z=0, we can replace the functions f,gf,g and h^\hat{h} by their corresponding UV expansions (168), obtaining U(z)γ2m^2/z2U(z)\sim\gamma^{2}\,\hat{m}^{2}/z^{2}. This implies that the potential diverges at the boundary. Since the resulting u(z,z0)u_{-}(z,z_{0}) function in (179) diverges as we move into the boundary z=0z=0, in order to have a normalizable solution, we need to impose L(0)=0L_{-}^{(0)}=0.

In the deep IR, the functions take their Lifshitz form (169), and we get U(z)γ2g(k^2ω^2z2(λ1))U(z)\sim\gamma^{2}{g_{\infty}}\,(\hat{k}^{2}-\hat{\omega}^{2}z^{2(\lambda-1)}). In consequence, for zero frequency the potential goes to a positive constant at infinity. The corresponding function u+(z,zR1)u_{+}(z,z_{R-1}) diverging, we need to impose R+(R1)=0R^{(R-1)}_{+}=0. Combined with the conditions at the boundary we obtain R+(R1)=V11L+(0)=0R^{(R-1)}_{+}=V_{11}L^{(0)}_{+}=0 or in other words

V11=0,V_{11}=0\,, (207)

where the matrix element V11V_{11} was defined in (203) in terms of integrals involving the potential U(z)U(z). For finite frequency on the other hand, the potential diverges into negative values. This results in an u(z,zR1)u_{-}(z,z_{R-1}) with the form of a wave moving into smaller zz values, i.e. entering the bulk from infinity, which implies that we must impose R(R1)=0R^{(R-1)}_{-}=0 in order to avoid causality issues. Since R(R1)=V21L+(0)=0R^{(R-1)}_{-}=V_{21}L^{(0)}_{+}=0 this implies

V21=0.V_{21}=0\,. (208)

Equations (207) and (208) impose a constraint between the parameters entering into U(z)U(z). It often has a discrete set of solutions, being then understood as a quantization condition (see below).

According to the summary on section D.3, using this information we need to obtain the free Fermi momentum kFm𝖿𝗋𝖾𝖾k_{F}^{m\;{\sf free}} and Fermi velocity vFm𝖿𝗋𝖾𝖾=dωm(k)/dk|k=kFm𝖿𝗋𝖾𝖾v_{F}^{m\;{\sf free}}=\left.d\omega_{m}(k)/dk\right|_{k=k_{F}^{m\;{\sf free}}} for each mode mm in the zz direction, as well as its static wave function ψαmkFm𝖿𝗋𝖾𝖾δ(z)\psi_{\alpha m{k_{F}^{m\,{\sf free}}}}^{\delta}(z) or in other words the solution ϕkω(2)\phi_{k\omega}^{(2)} of the effective Schrödinger equation (43) with k=kFk=k_{F} and ω=0\omega=0.

F.2.1 Fermi momenta and static wave functions

Of particular interest to our problem are the Fermi momenta, i.e. the values kFm𝖿𝗋𝖾𝖾{k_{F}^{m}}^{\sf free} of the momentum kk satisfying ω^m(kFm𝖿𝗋𝖾𝖾)=0\hat{\omega}_{m}({k_{F}^{m}}^{\sf free})=0. This implies that the corresponding potential in (206) has a vanishing frequency, and then goes to a positive constant at infinity. In consequence it has none or an even number of turning points.

The two last terms in the parenthesis in (206) are positive outside the star, and negative inside it, according to the rule (166). This implies that for kk large enough, the parenthesis remains always positive and there are no turning points. On the other hand, for kk smaller than some critical value kk^{*}, the potential becomes negative somewhere inside the star, giving rise to a pair of turning points z0,z1z_{0},z_{1}. Only in this last case we are able to find values of the free Fermi momenta, all of them inside a “Fermi ball” Lee:2008xf of radius kk^{*}.

The connection matrix (203) has only three factors 𝐕=𝐌𝐖(z0)𝐌{\bf V}={\bf M}\;{\bf W}(z_{0})\;{\bf M^{\dagger}} and the relevant matrix element in equation (207) takes the form

V11\displaystyle V_{11} =\displaystyle= 2cos(z0z1𝑑zU(z))=0,\displaystyle 2\,\cos{\left(\int_{z_{0}}^{z_{1}}dz\,\sqrt{-U(z)}\right)}=0\,, (209)

implying that the Bohr-Sommerfeld condition (207) reads

γz0z1dzg(z)(q2h^(z)2f(z)(kFm𝖿𝗋𝖾𝖾)2z2m^2)=(m+12)π,m=0,1,.\gamma\int_{z_{0}}^{z_{1}}dz\,\sqrt{g(z)\,\left(\frac{q^{2}\hat{h}(z)^{2}}{f(z)}-({k_{F}^{m}}^{\sf free})^{2}\,z^{2}-\hat{m}^{2}\right)}=\left(m+\frac{1}{2}\right)\,\pi\quad,\quad m=0,1,\dots\,. (210)

In terms of an integer mm, this equation determines the free Fermi momentum kFm𝖿𝗋𝖾𝖾{k_{F}^{m}}^{\sf free} of the mm-th fermionic mode. The results of these calculations are shown in Fig. 4

Refer to caption
Figure 4: The Fermi momenta are plotted in terms of mm for different values of γ=\gamma= 100,100, 150,150, 200,200, 250250 from bottom to top. For each value of γ\gamma there exists a maximum value of the Fermi momentum.

Next, we can replace the obtained values of the free Fermi momenta kFm𝖿𝗋𝖾𝖾{k_{F}^{m}}^{\sf free} into equations (179)-(191) to obtain the corresponding solutions of (43) ϕkω(2)=ϕm𝖶𝖪𝖡(z)\phi_{k\omega}^{(2)}=\phi^{\sf WKB}_{m}(z) needed to build the static wave-functions. In our calculations we found useful to replace the Airy functions interpolating around the turning points in (185) by a quartic polynomial. Results are shown in Fig. 5.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Profiles of the function ϕm𝖶𝖪𝖡(z)\phi_{m}^{\sf WKB}(z) are shown for the values m{0,,7}m\in\{0,\dots,7\} respectively. We see the smoothly interpolated solutions (light blue curves), and purely the WKB solutions (orange curves) that in the vicinity of the turning points were replaced by a quartic polynomial.

F.2.2 Fermi velocities

To calculate the Fermi velocities vFm𝖿𝗋𝖾𝖾=dωm(k)/dk|k=kFm𝖿𝗋𝖾𝖾v_{F}^{m\;{\sf free}}=\left.d\omega_{m}(k)/dk\right|_{k=k_{F}^{m\;{\sf free}}}, we need to concentrate in low energy excitations, with momenta very near to the Fermi momentum k=kF𝖿𝗋𝖾𝖾+dkk=k_{F}^{\sf free}+dk. This would result into small frequencies ω^=0+dω^\hat{\omega}=0+d\hat{\omega}. As discussed previously, for any non-zero frequency the potential has an odd number of turning points. As our frequency is very small, along with the two turning points of the zero frequency case z0,z1z_{0},z_{1}, only a third one z2z_{2} appears. As dω^d\hat{\omega} goes to zero, the third turning point z2z_{2} moves to infinity.

In this case the resulting connection matrix 𝐕\bf V in (203) has the five factors 𝐕=𝐌𝐖(z1)𝐌𝐖(z0)𝐌\quad{\bf V}={\bf M^{\dagger}}\;{\bf W}(z_{1})\;{\bf M}\;{\bf W}(z_{0})\;{\bf M^{\dagger}}, and (208) reads

tan(z0z1𝑑zU(z)π2)=i4e2z1z2(ω^)𝑑zU(z).\tan\left(\,\int_{z_{0}}^{z_{1}}dz\,\sqrt{-U(z)}-\frac{\pi}{2}\right)=-\frac{i}{4}\,e^{-2\,\int_{z_{1}}^{z_{2}(\hat{\omega})}dz\,\sqrt{U(z)}}\,. (211)

The solution for the dispersion relation is necessarily complex, defining a problem of quasi-normal modes. Since the integral in the exponent of the right hand side of (211) diverges as dω^d\hat{\omega} goes to zero, the right hand side is very small. This implies that we can write (211) as

z0z1𝑑zU(z)=(m+12)π.\int_{z_{0}}^{z_{1}}dz\,\sqrt{-U(z)}=\left(m+\frac{1}{2}\right)\,\pi\,. (212)

To the first order in dk^d\hat{k} and dω^d\hat{\omega}, we get

k^F𝖿𝗋𝖾𝖾dk^z0z1𝑑zg(z)z2U(z)dω^z0z1𝑑zqh(z)g(z)f(z)U(z)=0,\hat{k}_{F}^{\sf free}\,d\hat{k}\,\int_{z_{0}}^{z_{1}}dz\,\frac{g(z)z^{2}}{\sqrt{-U(z)}}-d\hat{\omega}\,\int_{z_{0}}^{z_{1}}dz\,\frac{qh(z)g(z)}{f(z)\sqrt{-U(z)}}=0\,, (213)

implying for the Fermi velocity

vF𝖿𝗋𝖾𝖾=k^F𝖿𝗋𝖾𝖾z0z1𝑑zz2g(z)U(z)z0z1𝑑zg(z)h(z)f(z)U(z),v_{F}^{\sf free}=\hat{k}_{F}^{\sf free}\,\frac{\int_{z_{0}}^{z_{1}}dz\,\frac{z^{2}\,g(z)}{\sqrt{-U(z)}}}{\int_{z_{0}}^{z_{1}}dz\,\frac{g(z)\,h(z)}{f(z)\,\sqrt{-U(z)}}}\,, (214)

where the potential and the turning points are evaluated at zero frequency and k^=k^F\hat{k}=\hat{k}_{F}.


In is worth to mention that a possible obstacle to get straight the dispersion relation at leading order from (212) arises from the non-analyticity of the complex right hand side of (211). However, as shown in Hartnoll:2011dm , it only affects (in fact, determines!) the imaginary part of the dispersion relation, while the real part is given as usual by the analytical left hand side.

References