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

Non-linear electrodynamics non-minimally coupled to gravity:
symmetric-hyperbolicity and causal structure.

Érico Goulart Federal University of São João d’El Rei, C.A.P. Rod.: MG 443, KM 7, CEP-36420-000, Ouro Branco, MG, Brazil    Santiago Esteban Perez Bergliaffa Departamento de Física Teórica, Instituto de Física, Universidade do Estado de Rio de Janeiro, 20550-013, Rio de Janeiro, Brazil
Abstract

It is shown here that symmetric hyperbolicity, which guarantees well-posedness, leads to a set of two inequalities for matrices whose elements are determined by a given theory. As a part of the calculation, carried out in a mostly-covariant formalism, the general form for the symmetrizer, valid for a general Lagrangian theory, was obtained. When applied to nonlinear electromagnetism linearly coupled to curvature, the inequalities lead to strong constraints on the relevant quantities, which were illustrated with applications to particular cases. The examples show that non-linearity leads to constraints on the field intensities, and non-minimal coupling imposes restrictions on quantities associated to curvature.

I Introduction

Well-posedness of the initial value problem stands out as a basic requirement to be satisfied by any relativistic field theory. Broadly speaking, it implies that solutions for a given problem exist, are unique, and depend continuously on the initial data. Hence, well-posedness is at the roots of physics, for it amounts to the predictability power of a given theory. While it is difficult to decide whether a given non-linear theory has a well-posed initial value problem, a necessary and sufficient condition for well-posedness around a given solution is that all the linearised problems obtained by linearising near such a solution are well-posed, see for instance Kreiss and Lorenz (1989); Papallo and Reall (2017).

In the linearized regime, well-possedness means that the equation of motion is hyperbolic. At least three notions of hyperbolicity can be distinguished Kreiss and Lorenz (1989); Beig (2006): (a) weak hyperbolicity, in which all the roots of the characteristic equation are real, (b) strong hyperbolicity implies that there is an energy estimate that sets a bound for the energy of a solution at a given time in terms of the initial energy 111In fact, strong hyperbolicity is a necessary and sufficient condition for the initial-value problem to be well-posed., and (c) symmetric hyperbolicity, which is a sufficient condition for well-posedness Sarbach and Tiglio (2012). It follows that symmetric hyperbolicity implies strong hyperbolicity, which in turn implies weak hyperbolicity.

There are many examples in which the requirement of some kind of hyperbolicity imposes severe restrictions on the Lagrangian of a given theory. For instance, as shown in Papallo and Reall (2017), the equations of motion of Lovelock’ss theory are always weakly hyperbolic for weak fields but not strongly hyperbolic in a generic weak-field background. The well-posedness of Horneski theory has been analyzed in Papallo and Reall (2017); Papallo (2017). As shown in the latter reference, the most general Horndeski theory which is strongly hyperbolic for weak fields in a generalized harmonic gauge is simply k-essence coupled to Einstein’s gravity (see also Bernard et al. (2019)). The well-posedness of scalar-tensor effective field theory was studied in Kovács and Reall (2020), where it was shown that the equations of motion are strongly hyperbolic at weak coupling.

We would like to explore here the constraints imposed by the requirement of well-posedness in the case of nonlinear electromagnetic theories coupled to gravity. Nonlinear electromagnetism has been widely studied in several contexts. A non-exhaustive list of applications and references includes black holes Moreno and Sarbach (2003); Bretón and Perez Bergliaffa (2015); Toshmatov et al. (2018); Magos and Bretón (2020); Falciano et al. (2021), astrophysics Heyl and Hernquist (2004); Harding and Lai (2006); Turolla et al. (2015) and cosmology Novello et al. (2004, 2007); Campanelli et al. (2008); Vollick (2008); Kruglov (2015). There are also several articles devoted to different aspects of the propagation of perturbations in nonlinear electromagnetic theories, such as Gutierrez et al. (1981); Deser et al. (1999); Goulart de Oliveira Costa and Esteban Perez Bergliaffa (2009); Perlick et al. (2018). The matter of (symmetric) hyperbolicity for nonlinear electromagnetism minimally coupled to gravity was analyzed in a flat spacetime background in Abalos et al. (2015), while the hyperbolicity of Maxwell’s equations with a local (and possibly nonlinear) constitutive law in flat spacetime was considered in Perlick (2011). Our aim here is to determine the restrictions that follow from the requirement of symmetric hyperbolicity on nonlinear electromagnetism non-minimally coupled to gravity, several aspects of which have been studied in detail in Balakin and Lemos (2005); Balakin et al. (2008, 2010); Dereli and Sert (2011). We shall restrict to couplings linear in the curvature, since higher-order couplings produce higher-than-second order equations for the gravitational field Balakin and Lemos (2005).

As mentioned above, a sufficient condition for well-posedness to hold is that the system under study admits a symmetric-hyperbolic representation. The theory of first-order symmetric hyperbolic systems, originally due to Friedrichs Friedrichs (1954), has been extensively developed (see for instance Beig (2006) and references therein). To study the evolution of a system we shall adopt here the modern geometric approach to the subject outlined by Geroch in Hall et al. (1996), in which covariance is kept during most of the calculation, instead of using a 3+1 decomposition of spacetime (as for instance in Perlick (2011)) 222For a list of other approaches to the analysis of the evolution of minimally coupled electromagnetism, see Abalos et al. (2015).. We shall see that such approach leads to a general form for the symmetrizer, and to conditions for symmetric hyperbolicity that are easier to evaluate than those for other types of hyperbolicity.

The structure of the paper is as follows. In Section II the basic notation used in the equations of motion is presented. Symmetric hyperbolicity, the related concept of symmetrizer, and the role of the constraints are analyzed in Section III. The equations defining the characteristic cones will be deduced in Section IV. Our main result, namely the explicit form of the matrices that are needed to investigate the symmetric hyperbolicity of any nonlinear electromagnetic theory non-minimally coupled to gravity can be found in Section V. In Section VI, some examples of the restrictions imposed by symmetric hyperbolicity are presented for different theories. Our closing remarks are presented in Section VII.

II Lagrangians and equations of motion

To begin with, let \mathcal{M} denote a smooth four-dimensional spacetime with a Lorentzian metric 𝒈\boldsymbol{g} of signature (+,,,)(+,-,-,-). For the sake of concreteness we assume \mathcal{M} to be also oriented and globally hyperbolic i.e. M×ΣM\cong\mathbb{R}\times\Sigma, with Σ\Sigma a codimension-11 hypersurface. Sticking to the conventions RabRacbcR_{ab}\equiv R^{c}_{\phantom{a}acb} and [ab]abba[ab]\equiv ab-ba, the canonical decomposition of the Riemman tensor into its irreducible parts reads

Rcdab=Wcdab+12δ[c[aSd]b]+R12gcdab,R^{ab}_{\phantom{a}\phantom{a}cd}=W^{ab}_{\phantom{a}\phantom{a}cd}+\frac{1}{2}\delta^{[a}_{\phantom{a}[c}S^{b]}_{\phantom{a}d]}+\frac{R}{12}g^{ab}_{\phantom{a}\phantom{a}cd}, (1)

where WabcdW_{abcd} is the Weyl conformal tensor, SabS_{ab} is the traceless part of the Ricci tensor, RR is the scalar curvature and gabcd=ga[cgd]bg_{abcd}=g_{a[c}g_{d]b} is the Kulkarni-Nomizu product of the metric with itself. Clearly, each factor in the decomposition has the same algebraic symmetries as the full Riemann tensor, that is

Rabcd=Rbacd=Rabdc,Rabcd=Rcdab,R_{abcd}=-R_{bacd}=-R_{abdc},\quad\quad\quad R_{abcd}=R_{cdab}, (2)
Rabcd+Racdb+Radbc=0.R_{abcd}+R_{acdb}+R_{adbc}=0. (3)

An arbitrary rank four covariant tensor satisfying Eqs. (2) has 21 independent components and is often referred to as a double symmetric (2,2)(2,2)-form (the skew pairs can be interchanged). If, in addition, the tensor satisfies Eq. (3), it is called an algebraic curvature tensor (see for instance Besse (1987)).

If we are given a double symmetric (2,2)(2,2)-form on \mathcal{M}, say χ\chi, we may obtain its left-dual χ\star\chi and right-dual χ\chi\star by

χabcd=12εabpqχpqcd,χabcd=12εcdpqχabpq,\star\chi_{abcd}=\frac{1}{2}\varepsilon_{ab}^{\phantom{a}\phantom{a}pq}\chi_{pqcd},\quad\quad\quad\chi_{abcd}\star=\frac{1}{2}\varepsilon^{pq}_{\phantom{a}\phantom{a}cd}\chi_{abpq}, (4)

where εabcd\varepsilon_{abcd} is the Levi-Civita tensor with ε0123=g\varepsilon_{0123}=\sqrt{-g}. Notice that the place of the \star indicates the pair of skew indices which are Hodge-dualized. A direct consequence is that

χabcd=χabcd,χabcd=χabcd.\star\star\chi_{abcd}=-\chi_{abcd},\quad\quad\quad\chi_{abcd}\star\star=-\chi_{abcd}. (5)

We aim at constructing Lagrangians describing the coupling of gravity with the electromagnetic field using invariants that are at most linear in the curvature 333As discussed in Balakin and Lemos (2005), higher-order couplings lead to equations of motion with derivatives of the metric higher than two., and respecting the U(1)U(1) gauge invariance of electromagnetism. Recalling that every algebraic curvature tensor satisfy the Ruse-Lanczos identity, we may construct the following independent rank four tensors (see the Appendix for details)

χcdab(1)gcdab,(2)χcdabεcdab,{}^{(1)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv g^{ab}_{\phantom{a}\phantom{a}cd},\quad\quad\quad^{(2)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv\varepsilon^{ab}_{\phantom{a}\phantom{a}cd}, (6)
χcdab(3)Rgcdab,(4)χcdabRεcdab,{}^{(3)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv Rg^{ab}_{\phantom{a}\phantom{a}cd},\quad\quad\quad^{(4)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv R\varepsilon^{ab}_{\phantom{a}\phantom{a}cd}, (7)
χcdab(5)Wcdab,(6)χcdabWcdab,{}^{(5)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv W^{ab}_{\phantom{a}\phantom{a}cd},\quad\quad\quad^{(6)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv\star W^{ab}_{\phantom{a}\phantom{a}cd}, (8)
χcdab(7)δ[c[aSd]b].{}^{(7)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\equiv\delta^{[a}_{\phantom{a}[c}S^{b]}_{\phantom{a}d]}. (9)

Notice that (Γ)χabcd{\phantom{a}}^{(\Gamma)}\chi_{abcd} (Γ=1,2,,7)(\Gamma=1,2,...,7) are double symmetric (2,2)(2,2)-forms by construction and it can be checked that they exhaust the interesting possibilities: the first pair, Eqs. (6), involves only algebraic terms in the metric, while the remaining Eqs. (7)-(9) contain at most linear terms in the curvature.

Let Fab=[aAb]F_{ab}=\partial_{[a}A_{b]} denote a test electromagnetic field propagating on \mathcal{M}. We shall focus on the propagation properties of such a field on a fixed gravitational background. The EM field will be described by the gauge-invariant action functional of the form

S=14d4xg(I1,I2,,I7),S=\frac{1}{4}\int d^{4}x\sqrt{-g}\ \mathcal{L}(I^{1},I^{2},...,I^{7}), (10)

where the factor 1/41/4 is introduced for future convenience and the Lagrangian density is taken as an arbitrary smooth function of the following scalars

IΓ(Γ)HabFab,where(Γ)Hab12(Γ)χcdabFcd.I^{\Gamma}\equiv{\phantom{a}}^{(\Gamma)}H^{ab}F_{ab},\quad{\rm where}\quad\quad{\phantom{a}}^{(\Gamma)}H^{ab}\equiv\frac{1}{2}{\phantom{a}}^{(\Gamma)}\chi^{ab}_{\phantom{a}\phantom{a}cd}F^{cd}. (11)

In analogy with electrodynamics in material media, we call (Γ)Hab{\phantom{a}}^{(\Gamma)}H^{ab} the Γ\Gamma-th induction tensor and (Γ)χcdab{\phantom{a}}^{(\Gamma)}\chi^{ab}_{\phantom{a}\phantom{a}cd} the Γ\Gamma-th constitutive tensor. Particular instances described by Eq. (10) include Maxwell’s theory, minimally-coupled nonlinear electrodynamics, and the three-parameter non-minimal Einstein–Maxwell model originating from QED vacuum polarization in a background gravitational field (see for instance Drummond and Hathrell (1980)), among others. With the conventions presented above, the variation of the action with respect to the 4-potential yields a coupled system of first-order quasi-linear PDEs for the fields, given by

bHab=0,bFab=0.\nabla_{b}H^{ab}=0,\quad\quad\quad\nabla_{b}\star F^{ab}=0. (12)

Here, a\nabla_{a} is the covariant derivative compatible with the metric and the full induction tensor is defined by

HabΓ(Γ)Hab,H^{ab}\equiv\displaystyle\sum\mathcal{L}_{\Gamma}{\phantom{a}}^{(\Gamma)}H^{ab}, (13)

with Γ/IΓ\mathcal{L}_{\Gamma}\equiv\partial\mathcal{L}/\partial I^{\Gamma}, for conciseness. Notice that the system defined by Eqs. (12) is composed of eight equations for only six unknowns. In relevant physical situations, the equations must include six dynamical equations (defined with respect to some time coordinate, to be identified later) and two constraints, which are to be imposed on initial data.

III Symmetric-hyperbolicity

In order to study whether the equations of motion admit a symmetric-hyperbolic representation, it is convenient to recast the system of equations Eqs. (12) in a unified manner as

KAβm(x,Φ)mΦβ+JA(x,Φ)=0,K_{A\phantom{a}\beta}^{\phantom{a}m}(x,\Phi)\partial_{m}\Phi^{\beta}+J_{A}(x,\Phi)=0, (14)

where xx\in{\mathcal{M}}, KAβmK_{A\phantom{a}\beta}^{\phantom{a}m} is the principal part of the PDE and JA(x,Φ)J_{A}(x,\Phi) stands for semi-linear contributions (whose explicit form is unnecessary for our discussion). Here capital Latin indices (A=1,,8)(A=1,...,8) stand for the space of multi-tensorial equations, lowercase Latin indices (m=0,1,2,3)(m=0,1,2,3) stand for space-time indices, and greek indices (β=1,..,6)(\beta=1,..,6) for tensorial unknowns. To start with, we introduce an ordering of the antisymmetric indices to obtain the six possible collective quantities

1(01)2(02)3(03)4(32)5(13)6(21).1\rightarrow(01)\quad 2\rightarrow(02)\quad 3\rightarrow(03)\quad 4\rightarrow(32)\quad 5\rightarrow(13)\quad 6\rightarrow(21). (15)

Making the identification ΦβFbc\Phi^{\beta}\rightarrow F^{bc} and performing simple manipulations in Eqs. (12), the principal part is then written as

KAαm=12(Xabcm,εabcm),K_{A\phantom{a}\alpha}^{\phantom{a}m}=\frac{1}{2}\left(X_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m},\ \varepsilon_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m}\right), (16)

with

XabcdΓΓ(Γ)χabcd+4ΓΛΓΛ(Γ)Hab(Λ)Hcd.X_{abcd}\equiv\sum_{\Gamma}\mathcal{L}_{\Gamma}{\phantom{a}}^{(\Gamma)}\chi_{abcd}+4\sum_{\Gamma}\sum_{\Lambda}\mathcal{L}_{\Gamma\Lambda}{\phantom{a}}^{(\Gamma)}H_{ab}{\phantom{a}}^{(\Lambda)}H_{cd}. (17)

XabcdX_{abcd} consists of a main term involving only first partial derivatives of the Lagrangian density and a nonlinear term including the Hessian matrix of the latter. Clearly, if the Lagrangian is a linear combination of the invariants IΓI^{\Gamma}, the last term vanishes and linear equations of motion follow. More importantly, XabcdX_{abcd} is always a symmetric double (2,2)-form independently on the specific form of the Lagrangian. This is a direct consequence of the symmetries of the constitutive tensors together with the symmetry of the Hessian matrix, i.e. ΓΛ=ΛΓ\mathcal{L}_{\Gamma\Lambda}=\mathcal{L}_{\Lambda\Gamma}.

In what follows we shall use the covariant approach for first-order symmetric-hyperbolic systems outlined in Hall et al. (1996): a symmetric hyperbolization of Eqs. (14) means that there exist a smooth symmetrizer hαAh^{A}_{\phantom{a}\alpha} and a covector field nmn_{m}, such that:

  1. 1.

    K^αβmhαAKAβm\hat{K}_{\alpha\phantom{a}\beta}^{\phantom{a}m}\equiv h^{A}_{\phantom{a}\alpha}K_{A\phantom{a}\beta}^{\phantom{a}m} is symmetric in the indices α,β\alpha,\ \beta.

  2. 2.

    The matrix K^αβ(n)K^αβmnm\hat{K}_{\alpha\beta}(n)\equiv\hat{K}_{\alpha\phantom{a}\beta}^{\phantom{a}m}n_{m} is positive-definite.

Roughly speaking, the first statement means that it should be possible to construct from Eqs. (14) a new subsystem of first-order quasi-linear PDEs given by

K^αβmmΦβ+J^α=0\hat{K}_{\alpha\phantom{a}\beta}^{\phantom{a}m}\partial_{m}\Phi^{\beta}+\hat{J}_{\alpha}=0 (18)

with J^αhαAJA\hat{J}_{\alpha}\equiv h^{A}_{\phantom{a}\alpha}J_{A}. Clearly, the new system contains only evolution equations and its dimension is equivalent to the number of unknown fields. The second statement means that the new system can be solved uniquely for any given set of initial conditions on a hypersurface Σ\Sigma with normal covector nmn_{m}. What are the remaining equations the symmetrizer does not capture? For the system to be consistent they should not be of the evolution type. In other words, they must be satisfied automatically once they are satisfied initially i.e., they should be what we normally call the constraints444See references Hall et al. (1996); Beig (2006); Kato (1975) for additional details..

The geometrical meaning of the covectors introduced in the previous paragraph is as follows. At a spacetime point pp\in\mathcal{M}, the collection of all covectors satisfying condition (2) is denoted by SpS_{p}. This set defines a nonempty open, convex cone at pp and the tangent vectors paTpp^{a}\in T_{p}\mathcal{M} such that pana>0p^{a}n_{a}>0 for all naSpn_{a}\in S_{p} determine the cone of influence of the physical field, i.e., the maximal speed of propagation in any given direction. It turns out that the latter is also a nonempty open, convex cone at pp.

In the next subsection, we start by showing that a family of symmetrizers parametrized by a vector field always exists for the system of first order PDEs given by Eqs. (12). It is important to point out that our result is general in the sense that it does not depend on the specific form of the Lagrangian density. We then obtain the conditions that a theory must satisfy for positive-definiteness (see condition (2) above) to hold. This is first obtained for a particular covector and henceforth generalized by inspecting the characteristic varieties (dispersion relations), which are necessarily well-behaved for symmetric hyperbolic systems.

III.1 Symmetrizer

In order to find a symmetrizer, it is convenient to work with projections. In other words, we seek a multi-tensorial field hαAh^{A}_{\phantom{a}\alpha} such that the quantity δϕα(hαAKAβm)δψβ\delta\phi^{\alpha}(h^{A}_{\phantom{a}\alpha}K_{A\phantom{a}\beta}^{\phantom{a}m})\delta{\psi}^{\beta} is symmetric in δϕ\delta\phi and δψ\delta{\psi}. Making the identifications

δϕαAab,δψαBab,\delta\phi^{\alpha}\rightarrow A^{ab},\quad\quad\quad\delta{\psi}^{\alpha}\rightarrow B^{ab}, (19)

where AA and BB are generic bivectors, we obtain from Eq. (16), the relation

KAβmδψβ=(12XabcmBbc,Bam).K_{A\phantom{a}\beta}^{\phantom{a}m}\delta\psi^{\beta}=\left(\frac{1}{2}X_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m}B^{bc},\ \star B_{a}^{\phantom{a}m}\right). (20)

Since there is no known practical procedure to obtain a symmetrizer for an arbitrary system of PDE’s, hyperbolizations are found, for a sufficiently low number of dimensions, by solving explicitly the algebraic equations inherent to the system defined by Eq. (14) and, in higher dimensions, by guessing. Let us show that the symmetrizer is given by the projection

δϕαhαA=(Aqa,12XqrsaArs)tq\delta\phi^{\alpha}h^{A}_{\phantom{a}\alpha}=\left(A^{a}_{\phantom{a}q}\ ,\ \frac{1}{2}\star X^{a}_{\phantom{a}qrs}A^{rs}\right)t^{q} (21)

where tqt^{q} is an auxiliary vector field and Xqrsa\star X^{a}_{\phantom{a}qrs} is the left Hodge dual as defined before. Indeed, multiplying (21) by (20) one obtains

δϕα(hαAKAβm)δψβ=12(XabcmAqaBbc+XqrsaArsBam)tq\delta\phi^{\alpha}(h^{A}_{\phantom{a}\alpha}K_{A\phantom{a}\beta}^{\phantom{a}m})\delta{\psi}^{\beta}=\frac{1}{2}\left(X_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m}A^{a}_{\phantom{a}q}B^{bc}+\star X^{a}_{\phantom{a}qrs}A^{rs}\star B_{a}^{\phantom{a}m}\right)t^{q} (22)

Now, defining

YqaXqrsaArs,YqaXqrsaArs,Y^{a}_{\phantom{a}q}\equiv X^{a}_{\phantom{a}qrs}A^{rs},\quad\quad\quad\star Y^{a}_{\phantom{a}q}\equiv\star X^{a}_{\phantom{a}qrs}A^{rs}, (23)

and using the well known identity valid for anti-symmetric tensors

(Yaq)(Bam)=12(YlnBln)δmq+YamBaq,(\star Y^{aq})(\star B_{am})=-\frac{1}{2}(Y_{ln}B^{ln})\delta_{m}^{\phantom{a}q}+Y_{am}B^{aq}, (24)

it follows that

δϕα(hαAKAβm)δψβ=12[Xabcm(AqaBbc+AbcBqa)12(XabcdBabAcd)δqm]tq,\delta\phi^{\alpha}(h^{A}_{\phantom{a}\alpha}K_{A\phantom{a}\beta}^{\phantom{a}m})\delta{\psi}^{\beta}=\frac{1}{2}\left[X_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m}(A^{a}_{\phantom{a}q}B^{bc}+A^{bc}B^{a}_{\phantom{a}q})-\frac{1}{2}(X_{abcd}B^{ab}A^{cd})\delta^{m}_{\phantom{a}q}\right]t^{q}, (25)

which is obviously symmetric in AA and BB since Xabcd=XcdabX_{abcd}=X_{cdab}. Therefore, the symmetrizer itself is given by the simple expression

hαA=12(gqrza,Xqrza)tq,h^{A}_{\phantom{a}\alpha}=\frac{1}{2}(g^{a}_{\phantom{a}qrz},\star X^{a}_{\phantom{a}qrz})t^{q}, (26)

It is remarkable that Eq. (26) is valid independently of the specific content of the tensor field XabcdX_{abcd}: in particular it does not matter whether the equations of motion contain quasi-linear terms or not.

The application of Eq. (26) to Eq. (16), yields the object

K^αβm=14(gqrzaXabcm+Xqrzaεabcm)tq,\hat{K}_{\alpha\phantom{a}\phantom{a}\beta}^{\phantom{a}m}=\frac{1}{4}(g^{a}_{\phantom{a}qrz}X_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m}+\star X^{a}_{\phantom{a}qrz}\varepsilon_{a\phantom{a}\phantom{a}bc}^{\phantom{a}m})t^{q}, (27)

which, after straightforward algebraic manipulations, becomes

K^αβm=14(gq[aXb]cdm+gq[cXd]abm+δqmXabcd)tq.\hat{K}_{\alpha\phantom{a}\phantom{a}\beta}^{\phantom{a}m}=-\frac{1}{4}\left(g_{q[a}X_{b]\phantom{a}\phantom{a}cd}^{\phantom{a}m}+g_{q[c}X_{d]\phantom{a}\phantom{a}ab}^{\phantom{a}m}+\delta^{m}_{\phantom{a}q}X_{abcd}\right)t^{q}. (28)

Notice that this equation is indeed symmetric in the exchange of antisymmetric indices abcdab\Leftrightarrow cd and that the auxiliary vector field tqt^{q} remains (up to now) arbitrary, so we can use it at our disposal. This concludes the first step of our task.

III.2 Positive definitness

Let us next investigate whether the above symmetrizer constitutes a true hyperbolization of the equations of motion. This will be achieved if we manage to find a covector field nmn_{m} such that the characteristic matrix

K^αβ(t,n)=14[t[aXb]cdm+t[cXd]abm+tmXabcd]nm\hat{K}_{\alpha\beta}(t,n)=-\frac{1}{4}[t_{[a}X_{b]\phantom{a}\phantom{a}cd}^{\phantom{a}m}+t_{[c}X_{d]\phantom{a}\phantom{a}ab}^{\phantom{a}m}+t^{m}X_{abcd}]n_{m} (29)

is positive definite, i.e.

δϕαK^αβδϕβ>0,\delta\phi^{\alpha}\hat{K}_{\alpha\beta}\delta\phi^{\beta}>0, (30)

for all nonzero vectors δϕα6\delta\phi^{\alpha}\in\mathbb{R}^{6}. A direct calculation gives the equivalent inequality

[taXbcdmAabAcd+14tmXabcdAabAcd]nm>0,-\left[t_{a}X_{b\phantom{a}\phantom{a}cd}^{\phantom{a}m}A^{ab}A^{cd}+\frac{1}{4}t^{m}X_{abcd}A^{ab}A^{cd}\right]n_{m}>0, (31)

where AabA^{ab} is an arbitrary, but nonzero bivector. Notice that the latter is a fully covariant expression in the sense that it does not depend on any particular choice of coordinates. It will restrict, however, the choices of admissible covector and auxiliary vector fields.

In order to proceed, we shall decompose AabA^{ab} and XabcdX_{abcd} with respect to the auxiliary vector field tqt^{q}. To do so, assuming that tqt^{q} is timelike, future-directd and normalized, we write the bivector as

Aab=𝔞[atb]+εabcdtc𝔟d,A_{ab}=\mathfrak{a}_{[a}t_{b]}+\varepsilon_{ab}^{\phantom{a}\phantom{a}cd}t_{c}\mathfrak{b}_{d}, (32)

with

𝔞aAabtb,𝔟aAabtb,𝔞ata=0𝔟ata=0.\mathfrak{a}_{a}\equiv A_{ab}t^{b},\quad\quad\mathfrak{b}_{a}\equiv\star A_{ab}t^{b},\quad\quad\mathfrak{a}_{a}t^{a}=0\quad\quad\mathfrak{b}_{a}t^{a}=0. (33)

Similarly, for any double (2,2)(2,2)-form, we have555See for instance Senovilla (2000) for similar decompositions.

Xabcd={gabpq(gcdrspr+εcdrspr)+εabpq(gcdrspr+εcdrs𝕊pr)}tqts,X_{abcd}=\left\{g_{abpq}\left(g_{cdrs}\mathbb{P}^{pr}+\varepsilon_{cdrs}\mathbb{Q}^{pr}\right)+\varepsilon_{abpq}\left(g_{cdrs}\mathbb{R}^{pr}+\varepsilon_{cdrs}\mathbb{S}^{pr}\right)\right\}t^{q}t^{s}, (34)

with the 2-index tensors given by

abXacbdtctd,abXacbdtctd,abXacbdtctd,𝕊abXacbdtctd,\mathbb{P}_{ab}\equiv X_{acbd}t^{c}t^{d},\quad\mathbb{Q}_{ab}\equiv-X_{acbd}\star t^{c}t^{d},\quad\mathbb{R}_{ab}\equiv-\star X_{acbd}t^{c}t^{d},\quad\mathbb{S}_{ab}\equiv\star X_{acbd}\star t^{c}t^{d}, (35)

and orthogonal to tqt^{q} by construction. In general, each of the latter has 9 independent components since the number of independent components of XabcdX_{abcd} is 36. However, in the particular case in which Xabcd=XcdabX_{abcd}=X_{cdab}, the following simplified relations are valid:

ab=ba,𝕊ab=𝕊ba,ab=ba,\mathbb{P}_{ab}=\mathbb{P}_{ba},\quad\quad\mathbb{S}_{ab}=\mathbb{S}_{ba},\quad\quad\mathbb{Q}_{ab}=\mathbb{R}_{ba}, (36)

so that ab\mathbb{P}_{ab} and 𝕊ab\mathbb{S}_{ab} have 6 independent components each, while ab\mathbb{Q}_{ab} (or, equivalently, ab\mathbb{R}_{ab}) have the remaining 9 components. Notice that, in the context of electrodynamics in material media, these tensors would be related to the permittivity, permeability and magneto-electric cross terms of the medium Balakin and Zayats (2018).

Using the above decompositions in Eq. (31), the new covariant inequality

(nmtm)[pr𝔞p𝔞r𝕊pr𝔟p𝔟r]>2(nmtn)εnpmq[qr𝔞p𝔞r+𝕊qr𝔞p𝔟r](n_{m}t^{m})[\mathbb{P}_{pr}\mathfrak{a}^{p}\mathfrak{a}^{r}-\mathbb{S}_{pr}\mathfrak{b}^{p}\mathfrak{b}^{r}]>2(n_{m}t^{n})\varepsilon^{mq}_{\phantom{a}\phantom{a}np}[\mathbb{R}_{qr}\mathfrak{a}^{p}\mathfrak{a}^{r}+\mathbb{S}_{qr}\mathfrak{a}^{p}\mathfrak{b}^{r}] (37)

follows, which is to be satisfied for any 33-vectors 𝔞p\mathfrak{a}^{p} and 𝔟q\mathfrak{b}^{q}, not vanishing simultaneously. Furthermore, since our considerations are essentially algebraic we may restrict to an arbitrary point pp\in\mathcal{M}. In order to complete our symmetric hyperbolization, it suffices to find a specific covector naTpn_{a}\in T^{*}_{p}\mathcal{M} satisfying inequality (37) since, if this is the case, there will be a unique connected, open and convex cone, SpS_{p}, containing the initial covector, and this cone will exhaust all possibilities. The natural choice here is the covector na=gabtbn_{a}=g_{ab}t^{b}, since the right-hand-side of Eq. (37) will vanish. Now, since nmtm>0n_{m}t^{m}>0 for the latter, we obtain the equivalent inequalities

pr𝔞p𝔞r>0,𝕊pr𝔟p𝔟r<0,\mathbb{P}_{pr}\mathfrak{a}^{p}\mathfrak{a}^{r}>0,\quad\quad\quad\mathbb{S}_{pr}\mathfrak{b}^{p}\mathfrak{b}^{r}<0,

for all nonzero vectors 𝔞p,𝔟q\mathfrak{a}^{p},\mathfrak{b}^{q}. In words, a symmetric hyperbolization is achieved for a covector nan_{a} if the 2-index tensors ab\mathbb{P}_{ab} and 𝕊ab\mathbb{S}_{ab}, obtained from XabcdX_{abcd} and Xabcd\star X_{abcd}\star via contractions with nan^{a}, satisfy the above inequalities.

For the sake of completness, let us a repeat the calculations in an adapted frame at pp, such that

gab(p)=ηab,tq=δ0q.g_{ab}(p)=\eta_{ab},\quad\quad\quad t^{q}=\delta^{q}_{\phantom{a}0}. (38)

Using Eq. (34), it can be shown that XabcdX_{abcd} has the following block matrix display:

Xαβ=(T𝕊),X_{\alpha\beta}=\left(\begin{array}[]{cc}\mathbb{P}&\mathbb{Q}\\ \\ \mathbb{Q}^{T}&\mathbb{S}\end{array}\right), (39)

where \mathbb{P}, \mathbb{Q} and 𝕊\mathbb{S} denote the covariant 3×33\times 3 matrices constructed with the corresponding tensors in the obvious way. In order to compute Eq. (29) in matrix form it is convenient to define the auxiliary 3×33\times 3 matrices (a similar notation was used in Perlick (2011))

A1=(000001010),A2=(001000100),A3=(010100000).\textbf{A}^{1}=\left(\begin{array}[]{ccc}0&0&0\\ 0&0&1\\ 0&-1&0\end{array}\right),\quad\quad\textbf{A}^{2}=\left(\begin{array}[]{ccc}0&0&-1\\ 0&0&0\\ 1&0&0\end{array}\right),\quad\quad\textbf{A}^{3}=\left(\begin{array}[]{ccc}0&1&0\\ -1&0&0\\ 0&0&0\end{array}\right).

Taking into account that the transposition relation (Ak)T=Ak(\textbf{A}^{k})^{T}=-\textbf{A}^{k} holds, a direct calculation gives:

K^αβ(n)=(n0nk(AkAkT)nkAk𝕊nk𝕊Akn0𝕊).\hat{K}_{\alpha\beta}(n)=\left(\begin{array}[]{cc}n_{0}\mathbb{P}-n_{k}(\mathbb{Q}\textbf{A}^{k}-\textbf{A}^{k}\mathbb{Q}^{T})&n_{k}\textbf{A}^{k}\mathbb{S}\\ \\ -n_{k}\mathbb{S}\textbf{A}^{k}&-n_{0}\mathbb{S}\end{array}\right). (40)

The above relation may be thought as a linear map from TxT^{*}_{x}\mathcal{M} to the 21-dimensional space of symmetric 6×66\times 6 matrices Sym6\mbox{Sym}_{6} . The set of all symmetric positive definite matrices forms an open convex cone in Sym6\mbox{Sym}_{6} with apex on the origin. It turns out that the image of the particular covector nm=ηmntnn_{m}=\eta_{mn}t^{n} will lie inside this cone whenever

0,𝕊0.\mathbb{P}\succ 0,\quad\quad\quad\quad\mathbb{S}\prec 0. (41)

since nm=(1,0,0,0)n_{m}=(1,0,0,0) in our frame. Therefore, if Eqs. (41) are satisfied pointwisely, symmetric hyperbolicity is guaranteed for the corresponding covector field. In other words, if the auxiliary vector field tat^{a} is vorticity-free, then initial data given on a hypersurface Σ\Sigma with normal covectors tat_{a} are uniquely evolved away from the hypersurface. As will be shown in Sect. VI, the requirements set by Eqs.(41) may yield severe constraints on the Lagrangian density, the intensity of the curvature tensor or the electromagnetic field.

III.3 Constraints

We now discuss the remaining equations, which are left aside by the symmetrizer. To do so, we recall that, acording to Geroch’s formalism, a constraint is a tensor cAnc^{An} such that

cAnKAαm+cAmKAαn=0.c^{An}K_{A\phantom{a}\alpha}^{\phantom{a}m}+c^{Am}K_{A\phantom{a}\alpha}^{\phantom{a}n}=0. (42)

It is straightforward to show that, in our case, this tensor is given by

cAn=(xgan,ygan)c^{An}=(xg^{an},yg^{an}) (43)

where x,yx,y\in\mathbb{R}. Indeed, multiplying Eq. (43) by (16) one obtains

cAnKAαm=12(xXbcnm+yεbcnm)c^{An}K_{A\phantom{a}\alpha}^{\phantom{a}m}=\frac{1}{2}(xX^{nm}_{\phantom{a}\phantom{a}\phantom{a}bc}+y\varepsilon^{nm}_{\phantom{a}\phantom{a}\phantom{a}bc}) (44)

which is obviously antisymmetric in nn and mm. Here, the emergence of two real numbers reveals that the vector space of constraints is actually 2-dimensional, as expected.

From the above calculation one concludes that the constraints are complete, in the sense that the number of constraint equations plus the number of evolution equations equals the number of initial equations. The constraints are integrable if the equation

cAnn(KAβmmΦβ+JA)=0,c^{An}\nabla_{n}(K_{A\phantom{a}\beta}^{\phantom{a}m}\nabla_{m}\Phi^{\beta}+J_{A})=0, (45)

is identically satisfied solely due the algebraic structure of the principal symbol, independently of the equations (14) of the original system. We leave for the reader to verify that this is indeed the case. This means that the original system of equations is equivalent to a symmetric-hyperbolic one with two additional integrable constraints.

IV Characteristic cones

Suppose that we manage to find a symmetric hyperbolization with constraints, as described in the previous sections. We have seen that this choice, however, is far from unique. Indeed, by the continuity of Eq.(40), any small deformation of tat_{a} will be such that condition (2) of symmetric hyperbolization is satisfied. It turns out that the set of all admissible covectors SpS_{p} in TpT_{p}^{*}\mathcal{M} is determined by the unique connected, open, convex, positive cone containing the initial covector Beig (2006) . Its existence is related to the hyperbolicity of the characteristic polynomial, defined by

p(n)det(K^αβ(n))p(n)\equiv\mbox{det}\left(\hat{K}_{\alpha\beta}(n)\right) (46)

which is a homogeneous multivariate polynomial of degree 66 in our case. Recall that, at a spacetime point, such a polynomial is called hyperbolic in a direction tat_{a} if p(ta)>0p(t_{a})>0 and the univariate polynomial p(ua+λta)p(u_{a}+\lambda t_{a}) only has real roots for all covectors uatau_{a}\neq t_{a}. Now, the vanishing set of the characteristic polynomial will define an algebraic variety: the cone of characteristic conormals (or characteristic cone, for brevity). In general, it will consist of different codimension 1 sheets which may be nested, intersect along lines, or even coincide. Geometrically, hyperbolicity in the direction of tat_{a} is the requirement that every line parallel to tat_{a} intersects this algebraic variety at exactly ``6"``6" points (counting multiplicities). Clearly, this condition severely constrains the topology of the characteristic cones, thus guaranteeing well-behaved propagation for small wavy excitations. In particular, it was shown by Garding Gårding (1959) that the closure of the connected component of tat_{a} in the set {na|p(na)0}\{n_{a}|p(n_{a})\neq 0\} is necessarily convex: the hyperbolicity cone of the polynomial. That symmetric hyperbolicity implies the hyperbolicity of the characteristic polynomial is direct. To see this one simply observes that the equation

det(K^αβ(a+λt))=0\mbox{det}(\hat{K}_{\alpha\beta}(a+\lambda t))=0 (47)

characterizes the eigenvalues of the quadratic form K^αβ(am)\hat{K}_{\alpha\beta}(a_{m}) relative to the metric K^αβ(tm)\hat{K}_{\alpha\beta}(t_{m}) - and these eigenvalues have to be real (see Beig (2006) for further details).

In order to compute the characteristic polynomial explicitly, we substitute Eq. (40) into Eq. (46) and use Schur’s determinant identity to obtain the product of determinants

p(n)=det(𝕊)q(n),p(n)=-\mbox{det}(\mathbb{S})q(n), (48)

where

q(n)det(n02n0nk(AkAkT)nknlAk𝕊Al).q(n)\equiv\mbox{det}\left(n_{0}^{2}\mathbb{P}-n_{0}n_{k}(\mathbb{Q}\textbf{A}^{k}-\textbf{A}^{k}\mathbb{Q}^{T})-n_{k}n_{l}\textbf{A}^{k}\mathbb{S}\textbf{A}^{l}\right). (49)

Notice that q(t)q(t) is necessarily positive definite, since >0\mathbb{P}>0 and 𝕊<0\mathbb{S}<0 for this particular covector. Interestingly, the determinant in Eq. (49) has been calculated several times in literature, see e.g. Perlick (2011),Rivera and Schuller (2011). In particular, it can be shown that it factorizes as

q(n)=(nmtm)2P(n),q(n)=(n_{m}t^{m})^{2}P(n), (50)

where

P(n)=124εa1a2a3a4εb1b2b3b4Xa1a2b1c1Xc2a3b2c3Xc4a4b3b4nc1nc2nc3nc4.P(n)=\frac{1}{24}\varepsilon_{a_{1}a_{2}a_{3}a_{4}}\varepsilon_{b_{1}b_{2}b_{3}b_{4}}X^{a_{1}a_{2}b_{1}c_{1}}X^{c_{2}a_{3}b_{2}c_{3}}X^{c_{4}a_{4}b_{3}b_{4}}n_{c_{1}}n_{c_{2}}n_{c_{3}}n_{c_{4}}. (51)

In other words, the sixth order polynomial always reduces to a product of a quadratic polynomial and a quartic polynomial. Since the vanishing set of the quadratic polynomial gives a non-compact variety (plane) inconsistent with the constraint equations, the characteristic cone is given by the covectors kak_{a} that satisfy the fourth-order Fresnel equation

P(k)(XpqarXbpcsXrsdq)kakbkckd=0,P(k)\sim\left(\star X_{pq}^{\phantom{a}\phantom{a}ar}X^{bpcs}X^{dq}_{\phantom{a}\phantom{a}rs}\star\right)k_{a}k_{b}k_{c}k_{d}=0, (52)

The fourth rank tensor defined by the terms between parentheses is called the Kummer tensor, whereas its totally symmetrized version is usually called the Tamm-Rubilar tensor. If symmetric hyperbolicity holds, the above polynomial is necessarily hyperbolic in the direction of tat_{a}, its vanishing set determining the causal structure of the theory up to a conformal factor. In specific situations, where XabcdX_{abcd} is sufficiently simple, Eq. (52) will reduce to the more familiar product of quadratic polynomials 666A general Ansatz leading to bi-metricity in nonlinear electromagnetism was presented in Visser et al. (2003).

P(k)(g(1)abkakb)(g(2)abkckd)=0.P(k)\sim\left(g^{ab}_{(1)}k_{a}k_{b}\right)\left(g^{ab}_{(2)}k_{c}k_{d}\right)=0. (53)

In these cases, symmetric hyperbolicity guarantees that the rank-2 contravariant tensors g(1)abg^{ab}_{(1)} and g(2)abg^{ab}_{(2)} are nondegenerate, necessarily of Lorentzian type and with the same signature. Furthermore, if these tensors coincide 777Covariant conditions on the Fresnel surface for birefringence to be absent were derived in Itin (2005). , then only one of them must be considered in the dispersion relation (reduced polynomial) , i.e.,

Pred(k)g(1)abkakb=0.P_{red}(k)\sim g^{ab}_{(1)}k_{a}k_{b}=0. (54)

It is important to point out that the abovementioned result stating that symmetric hyperbolicity leads to a cone strongly constraints the possible shape of the Fresnel surfaces, defined in a convenient 3-space: they must be topologically equivalent to those obtained from the 4-d hyperbolicity cone. Hence, open surfaces are not allowed by symmetric-hyperbolic propagation 888For examples of Fresnel surfaces in Maxwell´s theory non-minimally coupled to gravity, see Balakin and Zayats (2018). .

V General result

We have seen that symmetric hyperbolicity imposes restrictions on admissible physical theories. Importantly, they are always expressed in terms of matrix inequalities which are much easier to guarantee than to check the hyperbolicity of the corresponding characteristic polynomial. In order to obtain explicit expressions for the matrix inequalities, it is convenient to introduce the projection tensor hab=gabtatbh_{ab}=g_{ab}-t_{a}t_{b}, which projects arbitrary tensors onto the “rest spaces” orthogonal to tqt^{q}, i.e,

hab=hba,hachbc=hab,habtb=0.h_{ab}=h_{ba},\quad\quad\quad h_{ac}h^{c}_{\phantom{a}b}=h_{ab},\quad\quad\quad h_{ab}t^{b}=0. (55)

Similarly, we decompose the electromagnetic 2-form, the Weyl tensor and the traceless part of the Ricci tensor as

Fab=E[atb]+εabcdtcBd,F_{ab}=E_{[a}t_{b]}+\varepsilon_{abcd}t^{c}B^{d}, (56)
Wabcd={gabpq(gcdrsprεcdrspr)εabpq(gcdrspr+εcdrspr)}tqts,W_{abcd}=\left\{g_{abpq}\left(g_{cdrs}\mathcal{E}^{pr}-\varepsilon_{cdrs}\mathcal{B}^{pr}\right)-\varepsilon_{abpq}\left(g_{cdrs}\mathcal{B}^{pr}+\varepsilon_{cdrs}\mathcal{E}^{pr}\right)\right\}t^{q}t^{s}, (57)
Sab=Statb+Q(atb)+Nab,S_{ab}=St_{a}t_{b}+Q_{(a}t_{b)}+N_{ab}, (58)

with the following definitions

EaFabtb,BaFabtb,E_{a}\equiv F_{ab}t^{b},\quad\quad\quad B_{a}\equiv\star F_{ab}t^{b}, (59)
abWacbdtctd,abWacbdtctd,\mathcal{E}_{ab}\equiv W_{acbd}t^{c}t^{d},\quad\quad\quad\mathcal{B}_{ab}\equiv\star W_{acbd}t^{c}t^{d}, (60)
SSabtatb,QahabSbctc,NabhachbdScd.S\equiv S_{ab}t^{a}t^{b},\quad\quad Q_{a}\equiv h_{a}^{\phantom{a}b}S_{bc}t^{c},\quad\quad N_{ab}\equiv h_{a}^{\phantom{a}c}h_{b}^{\phantom{a}d}S_{cd}. (61)

According to the latter, the tensor fields {Ea,Ba,Qa,ab,ab,Nab}\{E_{a},B_{a},Q_{a},\mathcal{E}_{ab},\mathcal{B}_{ab},N_{ab}\} are automatically orthogonal to the auxiliary vector field.

In order to compute ab\mathbb{P}_{ab} and 𝕊ab\mathbb{S}_{ab} from Eq. (17), the following relations involving the constitutive tensors are useful

(1)χabcdtbtd=hac,(2)χabcdtbtd=0,{\phantom{a}}^{(1)}\chi_{abcd}t^{b}t^{d}=h_{ac},\quad\quad{\phantom{a}}^{(2)}\chi_{abcd}t^{b}t^{d}=0, (62)
(3)χabcdtbtd=Rhac,(4)χabcdtbtd=0,{\phantom{a}}^{(3)}\chi_{abcd}t^{b}t^{d}=Rh_{ac},\quad\quad{\phantom{a}}^{(4)}\chi_{abcd}t^{b}t^{d}=0, (63)
(5)χabcdtbtd=ac,(6)χabcdtbtd=ac,{\phantom{a}}^{(5)}\chi_{abcd}t^{b}t^{d}=\mathcal{E}_{ac},\quad\quad{\phantom{a}}^{(6)}\chi_{abcd}t^{b}t^{d}=\mathcal{B}_{ac}, (64)
(7)χabcdtbtd=Nac.{\phantom{a}}^{(7)}\chi_{abcd}t^{b}t^{d}=N_{ac}. (65)

Similarly, for the induction tensors, Eq. (11), one obtains

(1)Habtb=Ea,(2)Habtb=Ba,{\phantom{a}}^{(1)}H_{ab}t^{b}=E_{a},\quad\quad{\phantom{a}}^{(2)}H_{ab}t^{b}=B_{a}, (66)
(3)Habtb=REa,(4)Habtb=RBa,{\phantom{a}}^{(3)}H_{ab}t^{b}=RE_{a},\quad\quad{\phantom{a}}^{(4)}H_{ab}t^{b}=RB_{a}, (67)
(5)Habtb=abEbabBb,(6)Habtb=abBb+abEb,{\phantom{a}}^{(5)}H_{ab}t^{b}=\mathcal{E}_{ab}E^{b}-\mathcal{B}_{ab}B^{b},\quad\quad{\phantom{a}}^{(6)}H_{ab}t^{b}=\mathcal{E}_{ab}B^{b}+\mathcal{B}_{ab}E^{b}, (68)
(7)Habtb=SEa+NabEbεabcdtbQcEd,{\phantom{a}}^{(7)}H_{ab}t^{b}=SE_{a}+N_{ab}E^{b}-\varepsilon_{abcd}t^{b}Q^{c}E^{d}, (69)

Using the above relations and the Ruse-Lanczos identities (see the Appendix A), it can be checked that the tensors ab\mathbb{P}_{ab} and 𝕊ab\mathbb{S}_{ab} are given by:

ab\displaystyle\mathbb{P}_{ab} =\displaystyle= (1+3R)hab+5ab+6ab+7Nab\displaystyle(\mathcal{L}_{1}+\mathcal{L}_{3}R)h_{ab}+\mathcal{L}_{5}\mathcal{E}_{ab}+\mathcal{L}_{6}\mathcal{B}_{ab}+\mathcal{L}_{7}N_{ab}
+4EaEb(11+2R13)\displaystyle+4E_{a}E_{b}(\mathcal{L}_{11}+2R\mathcal{L}_{13})
+4E(aBb)[12+R(14+23)]\displaystyle+4E_{(a}B_{b)}\left[\mathcal{L}_{12}+R(\mathcal{L}_{14}+\mathcal{L}_{23})\right]
+4BaBb(22+2R24)\displaystyle+4B_{a}B_{b}(\mathcal{L}_{22}+2R\mathcal{L}_{24})
+4[15E(a+25B(a][b)cEcb)cBc]\displaystyle+4[\mathcal{L}_{15}E_{(a}+\mathcal{L}_{25}B_{(a}][\mathcal{E}_{b)c}E^{c}-\mathcal{B}_{b)c}B^{c}]
+4[16E(a+26B(a][b)cBc+b)cEc]\displaystyle+4[\mathcal{L}_{16}E_{(a}+\mathcal{L}_{26}B_{(a}][\mathcal{E}_{b)c}B^{c}+\mathcal{B}_{b)c}E^{c}]
+4[17E(a+27B(a][SEb)+Nb)cEcεb)pqrQptqEr]+\displaystyle+4[\mathcal{L}_{17}E_{(a}+\mathcal{L}_{27}B_{(a}][SE_{b)}+N_{b)c}E^{c}-\varepsilon_{b)pqr}Q^{p}t^{q}E^{r}]+...
𝕊ab\displaystyle\mathbb{S}_{ab} =\displaystyle= (1+3R)hab5ab6ab+7Nab\displaystyle-(\mathcal{L}_{1}+\mathcal{L}_{3}R)h_{ab}-\mathcal{L}_{5}\mathcal{E}_{ab}-\mathcal{L}_{6}\mathcal{B}_{ab}+\mathcal{L}_{7}N_{ab}
+4BaBb(11+2R13)\displaystyle+4B_{a}B_{b}(\mathcal{L}_{11}+2R\mathcal{L}_{13})
4E(aBb)[12+R(14+23)]\displaystyle-4E_{(a}B_{b)}[\mathcal{L}_{12}+R(\mathcal{L}_{14}+\mathcal{L}_{23})]
+4EaEb(22+2R24)\displaystyle+4E_{a}E_{b}(\mathcal{L}_{22}+2R\mathcal{L}_{24})
+4[15B(a25E(a][b)cBc+b)cEc]\displaystyle+4[\mathcal{L}_{15}B_{(a}-\mathcal{L}_{25}E_{(a}][\mathcal{E}_{b)c}B^{c}+\mathcal{B}_{b)c}E^{c}]
4[16B(a26E(a][b)cEcb)cBc]\displaystyle-4[\mathcal{L}_{16}B_{(a}-\mathcal{L}_{26}E_{(a}][\mathcal{E}_{b)c}E^{c}-\mathcal{B}_{b)c}B^{c}]
4[17B(a27E(a][SBb)+Nb)cBcεb)pqrQptqBr]+\displaystyle-4[\mathcal{L}_{17}B_{(a}-\mathcal{L}_{27}E_{(a}][SB_{b)}+N_{b)c}B^{c}-\varepsilon_{b)pqr}Q^{p}t^{q}B^{r}]+...

where the dots stand for possible nonlinear terms in the irreducible parts of the curvature tensor, which we shall not discuss in this work. The tensors ab\mathbb{P}_{ab} and 𝕊ab\mathbb{S}_{ab} given above are the most general ones when dealing with symmetric hyperbolicity in models of nonlinear electromagnetism coupled linearly to curvature. The inequalities given in Eqs. (41), defined in terms of such tensors, lead to drastic restrictions on admissible theories: there must be a strong compromise between the electromagnetic quantities {Ea,Ba}\{E^{a},B^{a}\}, the spacetime curvature expressed via {ab,ab,Qa,Nab}\{\mathcal{E}_{ab},\mathcal{B}_{ab},Q_{a},N_{ab}\} and the partial derivatives of the Lagrangian density.

VI Applications

Let us present next several examples that illustrate the restrictions obtained so far. For the sake of completeness, we display, in each case, the tensor XabcdX_{abcd}, the relevant inequalities and the corresponding characteristic polynomial. A generic feature of the latter is non-factorization: in general, the characteristic varieties are described by vanishing sets of fourth-order polynomials which do not split into products of second-order polynomials. Needless to say, this fact makes it considerably difficult to guarantee hyperbolicity straight from the characteristic polynomial. In constrast with this fact, symmetric-hyperbolicity automatically implies that the characteristic varieties are necessarily well-behaved.

VI.1 Maxwell electrodynamics =I1\mathcal{L}=-I^{1}

Our first example is Maxwell’s linear theory, which is the simplest possible case in our setting. Using Eq.(17), it follows that

Xabcd=gabcd.X_{abcd}=-g_{abcd}. (72)

Taking the auxiliary vector field tqt^{q} as timelike, future-directed and normalized, Eqs.(V) and (V) give the 2-index tensors as

ab=hab,𝕊ab=hab.\mathbb{P}_{ab}=-h_{ab},\quad\quad\quad\mathbb{S}_{ab}=h_{ab}. (73)

In a frame such that gab=ηabg_{ab}=\eta_{ab} and tq=δ0qt^{q}=\delta^{q}_{\phantom{a}0}, the corresponding 3×33\times 3 matrices are

ij=δij,𝕊ij=δij,i,j=1,2,3.\mathbb{P}_{ij}=\delta_{ij},\quad\quad\quad\mathbb{S}_{ij}=-\delta_{ij},\quad\quad\quad i,j=1,2,3. (74)

Since they satisfy Eqs.(41) identically, symmetric-hyperbolicity is guaranteed for all timelike covector fields nm=ηmntnn_{m}=\eta_{mn}t^{n}. A direct calculation using Eq. (52) leads to

P(k)\displaystyle P(k) \displaystyle\sim (gabkakb)2.\displaystyle(g^{ab}k_{a}k_{b})^{2}.

Only one of the quadratic polinomials is needed to obtain the usual dispersion relation for linear electromagnetic waves in vacuum:

Pred(k)gabkakb=0.P_{red}(k)\sim g^{ab}k_{a}k_{b}=0. (75)

Clearly, the cone of hyperbolicity at pp, SpS_{p}, is the connected component of tat_{a} i.e., it coincides with the set of future-directed timelike covectors, as expected.

VI.2 Minimally coupled nonlinear electrodynamics

VI.2.1 Single invariant: =(I1)\mathcal{L}=\mathcal{L}(I^{1})

If the Lagrangian density is an arbitrary function of the invariant I1I^{1}, a direct inspection of Eq. (17) gives

Xabcd=1gabcd+411FabFcd.X_{abcd}=\mathcal{L}_{1}g_{abcd}+4\mathcal{L}_{11}F_{ab}F_{cd}. (76)

From Eqs.(V) and (V), we obtain the projections

ab=1hab+411EaEb,𝕊ab=1hab+411BaBb,\displaystyle\mathbb{P}_{ab}=\mathcal{L}_{1}h_{ab}+4\mathcal{L}_{11}E_{a}E_{b},\quad\quad\quad\mathbb{S}_{ab}=-\mathcal{L}_{1}h_{ab}+4\mathcal{L}_{11}B_{a}B_{b}, (77)

Simple manipulations using Eq.(41) then yield the matrix inequalities

1δij411EiEj0,\mathcal{L}_{1}\delta_{ij}-4\mathcal{L}_{11}E_{i}E_{j}\prec 0, (78)
1δij+411BiBj0.\mathcal{L}_{1}\delta_{ij}+4\mathcal{L}_{11}B_{i}B_{j}\prec 0. (79)

In particular, they imply the condition 1<0\mathcal{L}_{1}<0, since the theory should be well-behaved when either the electric or magnetic field vanish. However, notice that for field intensities violating the inequalities, the propagation of small wavy excitations about the background field will be, in general, ill-posed.

A straightforward calculation using Eq. (52) gives the well-known bi-metric dispersion relation

P(k)(g(1)abkakb)(g(2)abkckd),P(k)\sim\left(g^{ab}_{(1)}k_{a}k_{b}\right)\left(g^{ab}_{(2)}k_{c}k_{d}\right), (80)

with the effective metrics given by

g(1)ab12gab,g(2)ab(1gab+411τab),τabFacFcb,g^{ab}_{(1)}\equiv\mathcal{L}_{1}^{2}g^{ab},\quad\quad\quad g^{ab}_{(2)}\equiv-(\mathcal{L}_{1}g^{ab}+4\mathcal{L}_{11}\tau^{ab}),\quad\quad\quad\tau^{ab}\equiv F^{ac}F^{b}_{\phantom{a}c}, (81)

as was shown for instance in Novello et al. (2000). Needless to say, if Eqs. (78) and (79) hold, the effective metrics are well-behaved, and both possess the same Lorentzian signature. However, it should be clear from our discussion that symmetric-hyperbolicity requires more than the simple Lorentzian nature of the effective metrics.

VI.2.2 Two invariants: (I1,I2)\mathcal{L}(I^{1},I^{2})

The case of two invariants is naturally more involved. The symmetric double (2,2)(2,2) form becomes

Xabcd=1gabcd+4{11FabFcd+12(FabFcd+FabFcd)+22FabFcd}.X_{abcd}=\mathcal{L}_{1}g_{abcd}+4\{\mathcal{L}_{11}F_{ab}F_{cd}+\mathcal{L}_{12}(F_{ab}\star F_{cd}+\star F_{ab}F_{cd})+\mathcal{L}_{22}\star F_{ab}\star F_{cd}\}. (82)

Apart from minor details on definitions, this tensor coincides with the so-called jump tensor obtained by Obukhov and Rubilar in Obukhov and Rubilar (2002). Notice also that the term involving 2\mathcal{L}_{2} in Eq. (17) was discarded, since it is proportional to the Bianchi identity. We then obtain the tensors (see Eqs.(V) and (V))

ab=1hab+4{11EaEb+12(EaBb+BaEb)+22BaBb},\mathbb{P}_{ab}=\mathcal{L}_{1}h_{ab}+4\{\mathcal{L}_{11}E_{a}E_{b}+\mathcal{L}_{12}(E_{a}B_{b}+B_{a}E_{b})+\mathcal{L}_{22}B_{a}B_{b}\}, (83)
𝕊ab=1hab+4{11BaBb12(EaBb+BaEb)+22EaEb},\mathbb{S}_{ab}=-\mathcal{L}_{1}h_{ab}+4\{\mathcal{L}_{11}B_{a}B_{b}-\mathcal{L}_{12}(E_{a}B_{b}+B_{a}E_{b})+\mathcal{L}_{22}E_{a}E_{b}\}, (84)

which, in the frame described above, lead to the inequalities

1δij4{11EiEj+12(EiBj+BiEj)+22BiBj}0,\mathcal{L}_{1}\delta_{ij}-4\{\mathcal{L}_{11}E_{i}E_{j}+\mathcal{L}_{12}(E_{i}B_{j}+B_{i}E_{j})+\mathcal{L}_{22}B_{i}B_{j}\}\prec 0, (85)
1δij+4{11BiBj12(EiBj+BiEj)+22EiEj}0,\mathcal{L}_{1}\delta_{ij}+4\{\mathcal{L}_{11}B_{i}B_{j}-\mathcal{L}_{12}(E_{i}B_{j}+B_{i}E_{j})+\mathcal{L}_{22}E_{i}E_{j}\}\prec 0, (86)

A tedious calculation using Eq. (52) yields the effective metrics

g(1)ab=𝒳gab+(𝒴+𝒴2𝒳𝒵)tab,\displaystyle g^{ab}_{(1)}=\mathcal{X}g^{ab}+(\mathcal{Y}+\sqrt{\mathcal{Y}^{2}-\mathcal{X}\mathcal{Z}})t^{ab}, (87)
g(2)ab=𝒳gab+(𝒴𝒴2𝒳𝒵)tab,\displaystyle g^{ab}_{(2)}=\mathcal{X}g^{ab}+(\mathcal{Y}-\sqrt{\mathcal{Y}^{2}-\mathcal{X}\mathcal{Z}})t^{ab}, (88)

with

𝒳=12+21(G22F)+(12121122)G2,\displaystyle\mathcal{X}=\mathcal{L}_{1}^{2}+2\mathcal{L}_{1}(G-\mathcal{L}_{22}F)+(\mathcal{L}_{12}\mathcal{L}_{12}-\mathcal{L}_{11}\mathcal{L}_{22})G^{2},
𝒴=21(11+22)+4(12121122)F,\displaystyle\mathcal{Y}=2\mathcal{L}_{1}(\mathcal{L}_{11}+\mathcal{L}_{22})+4(\mathcal{L}_{12}\mathcal{L}_{12}-\mathcal{L}_{11}\mathcal{L}_{22})F,
𝒵=(11221212),\displaystyle\mathcal{Z}=(\mathcal{L}_{11}\mathcal{L}_{22}-\mathcal{L}_{12}\mathcal{L}_{12}),

and we have used the notation I1FI^{1}\equiv F, I2GI^{2}\equiv G, tabFacFcbt^{ab}\equiv F^{ac}F^{b}_{\phantom{a}c} for conciseness. This result coincides with Obukhov and Rubilar (2002); De Lorenci et al. (2000), with minor modifications of notation.

VI.3 Nonminimally coupled nonlinear electrodynamics.

Since the equations of motion that follow from Eq. (10) are very general, it is convenient to consider some particular cases. Hence, in what follows we assume that the Lagrangian density takes the form

(I1,I2)+αI3+βI5+γI7,\mathcal{L}(I^{1},I^{2})+\alpha I^{3}+\beta I^{5}+\gamma I^{7}, (89)

where (I1,I2)\mathcal{L}(I^{1},I^{2}) is an arbitrary function of the usual electromagnetic invariants, and α,β\alpha,\beta and γ\gamma are phenomenological parameters to be determined in principle from experiments. Such a density is still sufficiently general to include most of the relevant important models present in literature.

Let us examine next some particular cases that follow from Eq.(89).

Linear non-minimal model

The corresponding Lagrangian density is given by

=I1+αI3+βI5+γI7,\mathcal{L}=-I^{1}+\alpha I^{3}+\beta I^{5}+\gamma I^{7}, (90)

since we expect to recover Maxwell’s theory in the flat spacetime regime. This model describes several possible non-minimal couplings of linear electrodynamics with gravity, and encompasses for instance the modifications of Maxwell´s theory due to one-loop vacuum polarization contributions Drummond and Hathrell (1980). Clearly, the corresponding equations of motion are linear and, using Eq. (17), one obtains the double symmetric (2,2)(2,2)-form as

Xabcd=(αR1)gabcd+βWabcd+γ(gacSbdgadSbc+gbdSacgbcSad).X_{abcd}=(\alpha R-1)g_{abcd}+\beta W_{abcd}+\gamma(g_{ac}S_{bd}-g_{ad}S_{bc}+g_{bd}S_{ac}-g_{bc}S_{ad}). (91)

Notice that the irreducible parts of the Riemann tensor contribute in different ways to the equations of motion. Using the above prescription in Eqs.(V) and (V), we obtain the following 22-index tensors

ab=(αR1)hab+βab+γNab\mathbb{P}_{ab}=(\alpha R-1)h_{ab}+\beta\mathcal{E}_{ab}+\gamma N_{ab} (92)
𝕊ab=(1αR)habβab+γNab\mathbb{S}_{ab}=(1-\alpha R)h_{ab}-\beta\mathcal{E}_{ab}+\gamma N_{ab} (93)

Let us consider next several sub-cases:

Scalar curvature coupling (α0,β=0,γ=0)(\alpha\neq 0,\beta=0,\gamma=0).

This is by far the simplest type of non-minimal coupling. Indeed, since the double symmetric (2,2)(2,2)-form in Eq.(91) reduces to

Xabcd=(αR1)gabcd,X_{abcd}=(\alpha R-1)g_{abcd}, (94)

the relevant 22-index tensors read

ab=(αR1)hab,𝕊ab=(1αR)hab.\mathbb{P}_{ab}=(\alpha R-1)h_{ab},\quad\quad\quad\mathbb{S}_{ab}=(1-\alpha R)h_{ab}. (95)

Hence, the simple matrix inequality

(1αR)δij0(1-\alpha R)\delta_{ij}\succ 0 (96)

follows. Clearly, symmetric-hyperbolicity requires that α<1/R\alpha<1/R, which may forbid good propagation for sufficiently high curvature, for a given α\alpha. Regarding the characteristic cone, Eq. (52) gives

P(k)(gabkakb)2,P(k)\sim(g^{ab}k_{a}k_{b})^{2}, (97)

which shows that in this case the dispersion relation is governed by the background metric i.e. the causal structure is not changed by the coupling.

Weyl coupling (α=0,β0,γ=0)(\alpha=0,\beta\neq 0,\gamma=0).

It follows from Eqs. (92) and (93) that

ab=𝕊ab=hab+βab.\mathbb{P}_{ab}=-\mathbb{S}_{ab}=-h_{ab}+\beta\mathcal{E}_{ab}.

Although the ensuing inequalities need to be examined on a case by case basis, they will lead to limits on the components of the electric part of the Weyl tensor 999Conversely, for a given geometry, the inequalities may furnish β\beta-dependent limits on the region of space-time where the propagation is symmetric-hyperbolic..

In order to compute the dispersion relation explicitly, we first recall that the Weyl conformal tensor has only one independent Hodge dual i.e. Wabcd=Wabcd\star W_{abcd}=W_{abcd}\star. A direct calculation using Eq. (52) gives a quartic equation of the type

P(k)Gabcdkakbkckd,P(k)\sim G^{abcd}k_{a}k_{b}k_{c}k_{d}, (98)

with the Kummer tensor given by

Gabcdgabgcdβ23(WapbqWpqcd+14gabWcpqrWpqrd)+β36WpqarWbpcsWrsdq.G^{abcd}\equiv g^{ab}g^{cd}-\frac{\beta^{2}}{3}\left(W^{apbq}W^{c\phantom{a}d}_{\phantom{a}p\phantom{a}q}+\frac{1}{4}g^{ab}W^{cpqr}W^{d}_{\phantom{a}pqr}\right)+\frac{\beta^{3}}{6}\star W_{pq}^{\phantom{a}\phantom{a}ar}W^{bpcs}W^{dq}_{\phantom{a}\phantom{a}rs}\star. (99)

Using the following identites due to Debever and Lanczos

WapcqWbdpqWapcqWbdpq=AgabgcdW_{apcq}W_{b\phantom{a}d}^{\phantom{a}p\phantom{a}q}-\star W_{apcq}\star W_{b\phantom{a}d}^{\phantom{a}p\phantom{a}q}=Ag_{ab}g_{cd} (100)
WapqrWbpqr=2AgabW_{apqr}W_{b}^{\phantom{a}pqr}=2Ag_{ab} (101)

with A18WpqrsWpqrsA\equiv\frac{1}{8}W_{pqrs}W^{pqrs}, we obtain the simplified expression

Gabcd=(1β26A)gabgcdβ23WapbqWpqcd+β33WpqarWbpcsWrsdq.G^{abcd}=\left(1-\frac{\beta^{2}}{6}A\right)g^{ab}g^{cd}-\frac{\beta^{2}}{3}W^{apbq}W^{c\phantom{a}d}_{\phantom{a}p\phantom{a}q}+\frac{\beta^{3}}{3}W_{pq}^{\phantom{a}\phantom{a}ar}W^{bpcs}W^{dq}_{\phantom{a}\phantom{a}rs}. (102)
Traceless Ricci couplings (α=0,β=0,γ0)(\alpha=0,\beta=0,\gamma\neq 0).

The relevant matrices take the form

ab=hab+γNab,\mathbb{P}_{ab}=-h_{ab}+\gamma N_{ab}, (103)
𝕊ab=hab+γNab.\mathbb{S}_{ab}=h_{ab}+\gamma N_{ab}. (104)

As in the previous case, the corresponding inequalities will relate the coupling constant γ\gamma to the curvature quantities described by NabN_{ab}. The Kummer tensor is given by

Gabcd\displaystyle G^{abcd} =\displaystyle= gabgcdγgabχpcpd(7)+γ22(χpapb(7)χqcqd(7)(7)χapbqχpqcd(7))\displaystyle g^{ab}g^{cd}-\gamma g^{ab}\ {}^{(7)}\chi^{cpd}_{\phantom{a}\phantom{a}\phantom{a}p}+\frac{\gamma^{2}}{2}\left({}^{(7)}\chi^{apb}_{\phantom{a}\phantom{a}\phantom{a}p}\ {}^{(7)}\chi^{cqd}_{\phantom{a}\phantom{a}\phantom{a}q}-^{(7)}\chi^{apbq}\ {}^{(7)}\chi^{c\phantom{a}d}_{\phantom{a}p\phantom{a}q}\right)
+γ36(7)χapbqχcrds(7)χprqs(7).\displaystyle\quad\quad\quad\quad\quad+\frac{\gamma^{3}}{6}\ ^{(7)}\chi^{apbq}\ {}^{(7)}\chi^{crds}\ {}^{(7)}\chi_{prqs}.
Non-minimally coupled EM in a cosmological background.

Another relevant example is that of the propagation in a cosmological background described by the flat Friedman-Lemâitre-Robertson-Walker metric, for which the Weyl tensor is null. In such a case, only the terms associated to 1{\cal L}_{1}, 3{\cal L}_{3}, and 7{\cal L}_{7} survive in Eqs. (V) and (V). Using Einstein´s equations, the traceless part of the Ricci tensor is given in terms of the matter by

Sab=TabT4gab,S_{ab}=T_{ab}-\frac{T}{4}g_{ab},

where TabT_{ab} is the energy-momentum tensor and TT, its trace. In a convenient tetrad basis, in which Tab=diag(ρ,p,p,p)T_{ab}={\rm diag}\;(\rho,p,p,p), it follows that

ij=δij[1+β(ρ3p)12γ(ρ+p)],\mathbb{P}_{ij}=-\delta_{ij}\left[1+\beta(\rho-3p)-\frac{1}{2}\gamma(\rho+p)\right], (105)
𝕊ij=δij[1+β(ρ3p)+12γ(ρ+p)],\mathbb{S}_{ij}=\delta_{ij}\left[1+\beta(\rho-3p)+\frac{1}{2}\gamma(\rho+p)\right], (106)

Hence, the inequalities (41) lead to

1+β(ρ3p)>12|γ|(ρ+p),1+\beta(\rho-3p)>\frac{1}{2}|\gamma|(\rho+p),

which is trivially satisfied in the case of linear EM. The propagation will be symmetric hyperbolic if this inequaly is satisfied at all times for which the model is valid.

VII Conclusions

Well-posedness is a basic requirement for any field theory, and is guaranteed by symmetric hyperbolicity. We have obtained a general form for the symmetrizer, given in Eq. (26), valid for a general Lagrangian theory. We have shown that symmetric hyperbolicity leads to a set of two inequalities for the matrices \mathbb{P} and 𝕊\mathbb{S}, whose elements are determined by a given theory. Regarding the constraints, we have verified that they are integrable.

When applied to nonlinear electromagnetism linearly coupled to curvature, the matrices \mathbb{P} and 𝕊\mathbb{S} are expressed in terms the fields, the Lagrangian, and its derivatives, and also of the different quantities associated to curvature. They lead to strong constraints on the relevant quantities, which were illustrated with applications to several particular cases. The examples show that while in the linear theory, no constraint arises from symmetric hyperbolic propagation, non-linearity leads to constraints on the field intensities, and non-minimal coupling imposes restrictions on quantities associated to curvature. In the general case, symmetric hyperbolicity relates the electromagnetic quantities {Ea,Ba}\{E^{a},B^{a}\}, the spacetime curvature expressed via {ab,ab,Qa,Nab}\{\mathcal{E}_{ab},\mathcal{B}_{ab},Q_{a},N_{ab}\} and the partial derivatives of the Lagrangian density.

The ideas presented here can be applied in other settings, such as electromagnetism in material media. We plan to return to this problem in a future publication.

Appendix A Ruse-Lanczos identity

Let χabcd\chi_{abcd} denote an arbitrary double symmetric (2,2) form at a point pp\in\mathcal{M}. Its double Hodge dual is defined as

χabcd=14εabpqεcdrsχpqrs.\star\chi_{abcd}\star=\frac{1}{4}\varepsilon_{ab}^{\phantom{a}\phantom{a}pq}\varepsilon_{cd}^{\phantom{a}\phantom{a}rs}\chi_{pqrs}. (107)

Writing the traces of χabcd\chi_{abcd} as

χabχacbc,χχaa,\chi_{ab}\equiv\chi^{c}_{\phantom{a}acb},\quad\quad\quad\chi\equiv\chi^{a}_{\phantom{a}a}, (108)

we may construct the symmetric trace-free tensor

ψabχab14gabχ.\psi_{ab}\equiv\chi_{ab}-\frac{1}{4}g_{ab}\chi. (109)

Using elementary algebraic manipulations, it can be shown that

χabcd+χabcd=ga[cψd]b+gb[dψc]a,\star\chi_{abcd}\star+\chi_{abcd}=g_{a[c}\psi_{d]b}+g_{b[d}\psi_{c]a}, (110)

which is called the Ruse-Lanczos identity Hall (2004); de Felice and Clarke (1992). Aplying this identity to the constitutive tensors results in the following useful relations

(1)χcdab+(1)χcdab=0,\displaystyle\star^{(1)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star+^{(1)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0,
(2)χcdab+(2)χcdab=0,\displaystyle\star^{(2)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star+^{(2)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0,
(3)χcdab+(3)χcdab=0,\displaystyle\star^{(3)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star+^{(3)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0,
(4)χcdab+(4)χcdab=0,\displaystyle\star^{(4)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star+^{(4)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0,
(5)χcdab+(5)χcdab=0,\displaystyle\star^{(5)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star+^{(5)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0,
(6)χcdab+(6)χcdab=0,\displaystyle\star^{(6)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star+^{(6)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0,
(7)χcdab(7)χcdab=0.\displaystyle\star^{(7)}\chi^{ab}_{\phantom{a}\phantom{a}cd}\star-^{(7)}\chi^{ab}_{\phantom{a}\phantom{a}cd}=0.

References