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

Magnetic equilibria of relativistic axisymmetric stars: The impact of flow constants

Arthur G. Suvorov [email protected] Manly Astrophysics, 15/41-42 East Esplanade, Manly, New South Wales 2095, Australia Theoretical Astrophysics, Eberhard Karls University of Tübingen, Tübingen, D-72076, Germany    Kostas Glampedakis [email protected] Departamento de Física, Universidad de Murcia, Murcia, E-30100, Spain Theoretical Astrophysics, Eberhard Karls University of Tübingen, Tübingen, D-72076, Germany
Abstract

Symmetries and conservation laws associated with the ideal Einstein-Euler system, for stationary and axisymmetric stars, can be utilized to define a set of flow constants. These quantities are conserved along flow lines in the sense that their gradients are orthogonal to the four-velocity. They are also conserved along surfaces of constant magnetic flux, making them powerful tools to identify general features of neutron star equilibria. One important corollary of their existence is that mixed poloidal-toroidal fields are inconsistent with the absence of meridional flows except in some singular sense, a surprising but powerful result first proven by Bekenstein and Oron. In this work, we revisit the flow constant formalism to rederive this result together with several new ones concerning both nonlinear and perturbative magnetic equilibria. Our investigation is supplemented by some numerical solutions for multipolar magnetic fields on top of a Tolman-VII background, where strict power-counting of the flow constants is used to ensure a self-consistent treatment.

I Introduction

The intricate magnetic fields of neutron stars govern most, if not all, of their persistent and transient activity. Outbursts from the magnetar subclass are likely powered by an ultra-strong internal field Duncan and Thompson (1992); Thompson and Duncan (1993), for instance, while it is the magnetospheric field that largely controls the morphology of radio pulsars Beskin et al. (1993); Melrose and Yuen (2016) and X-ray binaries Patruno and Watts (2021); Glampedakis and Suvorov (2021). Magnetohydrodynamic (MHD) modelling of neutron stars, particularly within the realm of general relativity (GR), is thus crucial to interpret their emissions.

Though theorised earlier from stability arguments Prendergast (1956); Wright (1973); Tayler (1973), the pioneering simulations by Braithwaite and collaborators Braithwaite and Spruit (2004); Braithwaite and Nordlund (2006); Braithwaite and Spruit (2006) led to the popularity of the ‘twisted torus’ model: an almost axisymmetric field with a toroidal component that is confined to an equatorial torus. These configurations are stable over many dynamical (Alfvén) timescales (see also Refs. Braithwaite (2009); Akgün et al. (2013); Mitchell et al. (2015)), and thus are a leading candidate for the field structure of mature neutron stars (though it has been argued that non-axisymmetric end states are more likely Braithwaite (2008); Becerra et al. (2022)). MHD evolutions of pulsar magnetospheres suggest that the inclination angle made between the magnetic and rotation axes tends to zero at late times Philippov et al. (2014), further promoting the degree of axisymmetry. There is thus reasonable motivation for exploring the set of stationary and axisymmetric (GR-)MHD equilibria, even if a neutron star never truly attains such a state (because of processes related to finite thermal Pons et al. (2009) and electrical Gourgouliatos et al. (2022) conductivity, for example).

It is generally advantageous to exploit symmetries when approaching a theoretical problem. In cases of stationary and axisymmetric MHD equilibria, there exists a number of flow constants (also known as streamline invariants Markakis et al. (2017)) which reduce the complexity of the system Chandrasekhar (1956); Woltjer (1959); Mestel (1999). More precisely, these ‘flow constants’ are functions which are conserved along flow lines in the sense that their gradient is orthogonal to the (four-)velocity. Equivalently, these are functions which are conserved along surfaces of constant magnetic flux. Their introduction allows for the basic (GR)MHD equations to be rewritten in a form that is more conducive to analytic and numerical investigation, like the familiar example of the Grad-Shafranov (GS) equation which may contain some general toroidal function of the magnetic flux (i.e. a flow constant) Lovelace et al. (1986); Ioka and Sasaki (2003).

In the GR context, these flow constants were derived and detailed by Bekenstein and Oron (hereafter BO; Bekenstein and Oron (1978, 1979)) and others since, including Ioka and Sasaki (IS; Ioka and Sasaki (2003, 2004)) and Gourgoulhon and collaborators Gourgoulhon (2006); Gourgoulhon et al. (2011); Markakis et al. (2017); Uryū et al. (2019); Tsokaros et al. (2022). Using this formalism, BO were able to prove a surprising result about GRMHD equilibria: configurations with mixed poloidal-toroidal fields are singular in the limit of a vanishing meridional flow. That is, the limit of vanishing meridional flow also corresponds to the limit of either a purely poloidal or purely toroidal magnetic field and a circular spacetime Oron (2002). The former is at least intuitive on a dynamical level: inducing a toroidal field requires a poloidal current, and hence meridional flow (Frieben and Rezzolla, 2012). It is remarkable though that one cannot completely suppress the meridional flow if considering a sequence of mixed-field equilibria. As purely poloidal or toroidal fields are unstable Lasky et al. (2011); Ciolfi et al. (2011); Kiuchi et al. (2011); Ciolfi and Rezzolla (2012), the BO theorem suggests that meridional flows are a necessary ingredient in realistic neutron star models (see also Ref. Ofengeim and Gusakov (2018)). This result seems to have gone largely unnoticed in the literature.

The goal of this work is to revisit the flow constant formalism and its corollaries, both at the non-linear and perturbative levels. Indeed, owing to the complexity of non-linear GRMHD, most models of neutron star equilibria, such as those of Colaiuda et al. (C08; Colaiuda et al. (2008)) and Ciolfi et al. (C09; Ciolfi et al. (2009)), invoke a perturbative scheme using an expansion in the magnetic field. We discuss the various scenarios in the context of the BO results, aiming to provide a fresh look at the theory landscape and its implications for neutron star structure. The special roles of the induction and GS equations are discussed in detail. Our results are complemented by some numerical solutions for perturbative equilibria, which differ from existing ones in the literature: we argue in Sec. IV.5 that previous choices may not be physically admissible.

This work is split into two parts. A reader whose interests are more focussed towards non-linear and general corollaries of the flow constant formalism can contain their reading to Secs. II and III, while one who is primarily interested in perturbative equilibria may skip ahead to Sec. IV. Numerical constructions of multipolar ‘twisted-torus’ solutions are presented in Sec. V, with discussion and conclusions given in Sec. VI.

Natural units c=G=1c=G=1 are assumed throughout, unless explicitly stated otherwise. Greek indices run over spacetime coordinates, while Latin indices are reserved for purely meridional components (rr,θ\theta). As a quick guide, some previous literature that is frequently referred to is abbreviated as follows: BO78 = Ref. Bekenstein and Oron (1978), BO79 = Ref. Bekenstein and Oron (1979), IS03 = Ref. Ioka and Sasaki (2003), IS04 = Ref. Ioka and Sasaki (2004), C08 = Ref. Colaiuda et al. (2008), and C09 = Ref. Ciolfi et al. (2009).

II General-relativistic MHD equilibria with flow constants

This section presents certain key relationships between flow constants and ‘primitive’ quantities, such as the poloidal magnetic field and the fluid velocity. Although the results presented in this section can largely be found in previous literature (most notably Refs. Bekenstein and Oron (1978, 1979); Ioka and Sasaki (2003, 2004); Gourgoulhon (2006); Gourgoulhon et al. (2011); Markakis et al. (2017); Uryū et al. (2019); Tsokaros et al. (2022); Oron (2002)), we provide a summary here, with the eventual aim of pointing out important corollaries and various limits in Section III. A self-contained derivation of the flow constants (following the steps of the original derivation of BO78 and BO79) can be found in Appendix A. For completeness and to highlight the role of GR, an analogous derivation in the Newtonian limit is provided in Appendix B.

II.1 Metric tensor and conventions

At the full non-linear level, we work with a general stationary and axisymmetric metric gμνg_{\mu\nu} in spherical-like coordinates {t,r,θ,φ}\{t,r,\theta,\varphi\} with signature {,+,+,+}\{-,+,+,+\}. Later on it will be convenient to introduce a static and spherically symmetric ‘background’, where the magnetic field is treated perturbatively (see Sec. IV). Here and in Sec. III, MHD dynamics are treated non-linearly, though the fluid is assumed to be perfect, infinitely conducting, and barotropic.

The stress-energy tensor appearing in the Einstein equations contains both fluid and electromagnetic components. The fluid piece is written as

Tfluidμν=(ϵ+p)uμuν+pgμν,T^{\mu\nu}_{\rm fluid}=\left(\epsilon+p\right)u^{\mu}u^{\nu}+pg^{\mu\nu}, (1)

where ϵ,p,uμ\epsilon,p,u^{\mu} denote the fluid energy density, pressure, and four-velocity, respectively. The former parameter can be expressed in terms of the rest-mass density and specific internal energy ϵ~\tilde{\epsilon} as

ϵ=ρ(1+ϵ~).\epsilon=\rho(1+\tilde{\epsilon}). (2)

With regard to these densities, there is a certain degree of notational confusion in the literature. In Refs. Ciolfi et al. (2009, 2010), for instance, ρ\rho is used to denote the energy density, which makes it appear at first glance that the resulting GS equation disagrees with others in the literature Ioka and Sasaki (2003, 2004), though this is not the case (cf. Sec. IV.5). The notation used here is the typical one when working with the Tolman-Oppenheimer-Volkoff (TOV) equations, for example. For a barotropic fluid, the energy density and rest-mass density are related through the first law of thermodynamics, d(ϵ/ρ)=pd(1/ρ)d(\epsilon/\rho)=-pd(1/\rho).

Expression (1) corresponds to the stress-energy of a single fluid, where we have ρ=mn\rho=mn for baryon mass, mm, and number density, nn. In reality, neutron star matter comprises multiple particle species (e,p,n,e^{-},p,n,\ldots) and a multi-fluid description is needed (see, for instance, Refs. Prix (2005); Glampedakis et al. (2012)). The distinction between particle constituents is important when considering the dynamics of magnetised stars, as the four-current is entirely sourced by the flow of the electrons relative to the other charged particles in the limit of charge neutrality. The single fluid description is based on the additional assumptions of a negligible electron inertia and a common velocity for the rest of the particles (charged and uncharged).

The flow constants we introduce in the next section that arise from the induction equation should thus be strictly associated with the electron number density and not the global one. This point is discussed further in Sec. III.3. For a single fluid in the ideal MHD limit, the electromagnetic stress-energy tensor is given by

TEMαβ=gαβB28π+14π(B2uαuβBαBβ),T^{\alpha\beta}_{\rm EM}=g^{\alpha\beta}\frac{B^{2}}{8\pi}+\frac{1}{4\pi}\left(B^{2}u^{\alpha}u^{\beta}-B^{\alpha}B^{\beta}\right), (3)

where BμB^{\mu} is the magnetic field four-vector. The electric and magnetic components are generally given as Eμ=FμνuνE_{\mu}=F_{\mu\nu}u^{\nu} and Bα=12gϵαβγδuβFγδB_{\alpha}=\tfrac{1}{2}\sqrt{-g}\epsilon_{\alpha\beta\gamma\delta}u^{\beta}F^{\gamma\delta} for Levi-Civita symbol ϵαβγδ\epsilon_{\alpha\beta\gamma\delta} and Faraday tensor Fμν=Aν,μAμ,νF_{\mu\nu}=A_{\nu,\mu}-A_{\mu,\nu}. Here AμA_{\mu} is the vector potential, whose existence is implied by Maxwell’s equations, [αFμν]=0\nabla_{[\alpha}F_{\mu\nu]}=0, and Eμ=0E_{\mu}=0 defines the ideal MHD condition of infinite conductivity.

A result of immediate interest concerns a theorem by Carter Carter (1969, 1973). While a strictly azimuthal flow (ui=0)(u^{i}=0) implies that one can consider a circular spacetime, gti=gφi=0g_{ti}=g_{\varphi i}=0, inclusion of meridional flows or mixed poloidal-toroidal fields generally forces these extra off-diagonal terms to be non-zero and means one cannot work with the Papapetrou metric Papapetrou (1966). Intuitively, if one considers a radial flow (ur0u^{r}\neq 0), for instance, it is clear that the physical dynamics are not symmetric under time reversal if the system is also rotating. Similarly, for a mixed magnetic field there will generally be a poloidal current and the same logic applies (Frieben and Rezzolla, 2012). Spacetimes with purely toroidal fields are still circular however, since one can then exploit the azimuthal coordinate/gauge freedom Gourgoulhon et al. (2011). Additionally, one can always set grθ=0g_{r\theta}=0 regardless.

The necessity of non-circular components is one key reason why constructing mixed poloidal-toroidal equilibria at a non-linear level is so difficult in GR Bocquet et al. (1995); it is only recently that such equilibria have been numerically constructed Uryū et al. (2019) and evolved Tsokaros et al. (2022) self-consistently111It has been argued that even for mixed, virial-strength fields, the degree of non-circularity is of order 103\lesssim 10^{-3} Oron (2002); Pili et al. (2017). Whether this remains true for all physical setups is unclear owing to the numerical nature of the statement, but for computational purposes a circular approximation may be reasonable. (see also Refs. Lovelace et al. (1986); Fujisawa et al. (2013); Ofengeim and Gusakov (2018) for some Newtonian solutions). Aside from mathematical complexity, these extra pieces enrich the physical system as discussed in Sec. VI.

II.2 The BO flow constants

In this section we list the formulae for the flow constants, as derived in the BO papers Bekenstein and Oron (1978, 1979) and the aforementioned references. A proper derivation is relegated to Appendix A owing to the length of the calculations involved. Each of the constants here are derived from the conservation law μTμν=μ(Tfluidμν+TEMμν)=0\nabla_{\mu}T^{\mu\nu}=\nabla_{\mu}\left(T^{\mu\nu}_{\rm fluid}+T^{\mu\nu}_{\rm EM}\right)=0 together with the Einstein equations Rμν12Rgμν=8πTμνR_{\mu\nu}-\tfrac{1}{2}Rg_{\mu\nu}=8\pi T_{\mu\nu} of a general stationary and axisymmetric spacetime.

Although we use the term ‘flow’, Eq. (107) from Appendix A shows that for any conserved QQ (i.e. uαQ,α=0u^{\alpha}Q_{,\alpha}=0) one necessarily has that BαQ,α=0B^{\alpha}Q_{,\alpha}=0. Introducing the magnetic potential ψ=Aφ\psi=A_{\varphi}, which can be shown to satisfy Bαψ,α=0B^{\alpha}\psi_{,\alpha}=0, we express the flow constants in a more familiar form using ψ\psi. For circular spacetimes, it is always possible to introduce a zero-angular-momentum observer such that the poloidal field admits a Chandrasekhar-like description with Brψ,θB^{r}\propto\psi_{,\theta} and Bθψ,rB^{\theta}\propto\psi_{,r} up to metric factors Bocquet et al. (1995). Although the decomposition is not as simple in general, we still have that BiAφB^{i}\propto A_{\varphi} by definition and thus Bi=𝒪(ψ)B^{i}=\mathcal{O}(\psi) to leading order since uiAt,i=0u^{i}A_{t,i}=0 from the ideal MHD condition uνFtν=0u^{\nu}F_{t\nu}=0 (see also Sec. 2.1.2 in C08).

Adopting the notation of IS03 and IS04, we have the following relations amongst the five flow constants 𝒞(ψ),Ω(ψ),E(ψ),L(ψ),{\cal C}(\psi),\Omega(\psi),E(\psi),L(\psi), and D(ψ)D(\psi):

E=(μ+B24πn)ut𝒞4π(ut+Ωuφ)Bt,E=-\left(\mu+\frac{B^{2}}{4\pi n}\right)u_{t}-\frac{{\cal C}}{4\pi}(u_{t}+\Omega u_{\varphi})B_{t}, (4)
L=(μ+B24πn)uφ+𝒞4π(ut+Ωuφ)Bφ,L=\left(\mu+\frac{B^{2}}{4\pi n}\right)u_{\varphi}+\frac{{\cal C}}{4\pi}(u_{t}+\Omega u_{\varphi})B_{\varphi}, (5)
D=EΩL,D=E-\Omega L, (6)

where μ\mu is the specific enthalpy (this is defined in Appendix A and should not be confused with the chemical potential unless the temperature is everywhere zero). Some of these constants can be used to express the magnetic field BμB^{\mu} and meridional flow uiu^{i} (i={r,θ}i=\{r,\theta\}) via

Bμ=𝒞n[(ut+Ωuφ)uμ+δtμ+Ωδφμ],B^{\mu}=-{\cal C}n\left[\,(u_{t}+\Omega u_{\varphi})u^{\mu}+\delta_{t}^{\mu}+\Omega\delta^{\mu}_{\varphi}\,\right], (7)

and

B2=𝒞n(Bt+ΩBφ).B^{2}=-{\cal C}n(B_{t}+\Omega B_{\varphi}). (8)

Using (8) in Eqs. (4) and (5) one finds the alternative expressions

E\displaystyle E =μut+𝒞Ω4π(utBφuφBt),\displaystyle=-\mu u_{t}+\frac{{\cal C}\Omega}{4\pi}(u_{t}B_{\varphi}-u_{\varphi}B_{t}), (9)
L\displaystyle L =μuφ+𝒞4π(utBφuφBt).\displaystyle=\mu u_{\varphi}+\frac{{\cal C}}{4\pi}(u_{t}B_{\varphi}-u_{\varphi}B_{t}). (10)

It is illuminating to decompose (7) into components, viz.

Bi\displaystyle B^{i} =𝒞n(ut+Ωuφ)ui,\displaystyle=-{\cal C}n(u_{t}+\Omega u_{\varphi})u^{i}, (11)
Bφ\displaystyle B^{\varphi} =𝒞n[Ω+(ut+Ωuφ)uφ],\displaystyle=-{\cal C}n\left[\Omega+(u_{t}+\Omega u_{\varphi})u^{\varphi}\right], (12)
Bt\displaystyle B^{t} =𝒞n[1+(ut+Ωuφ)ut].\displaystyle=-{\cal C}n\left[1+(u_{t}+\Omega u_{\varphi})u^{t}\right]. (13)

Following IS04 and BO79, these constants can be interpreted as follows. E(ψ)E(\psi) and L(ψ)L(\psi) represent the total enthalpy of the plasma and specific enthalpic angular momentum of the magnetic field, respectively; D(ψ)D(\psi) symbolises the specific plasma energy according to a corotating observer and Ω(ψ)\Omega(\psi) relates to the angular velocity222Note that what is called Ω\Omega here generally differs from the actual angular velocity uφ/utu^{\varphi}/u^{t}, with agreement only in the limit that the toroidal field vanishes; see Eq. (4.42) of Ref. Gourgoulhon et al. (2011), noting that what we call Ω\Omega is denoted there by ω\omega and A-A by BO. Because of its hybrid nature, Ω\Omega could also be considered ‘induction-like’. of the fluid. These constants can be thought of as ‘Bernoulli-like’, in the sense that there is a simple but non-trivial hydrodynamic limit; the constant DD in particular can be directly identified with the Bernoulli integral (see Appendices). The remaining constant 𝒞(ψ){\cal C}(\psi) – representing the magnetic field strength relative to the magnitude of meridional flow – is instead ‘induction-like’. The existence of this constant is a direct consequence of the ideal assumption Eμ=0E_{\mu}=0 and becomes trivial in the non-magnetic limit. Moreover, expression (11) shows that in the limit of vanishing meridional flow, ui0u^{i}\rightarrow 0, the poloidal field generally vanishes unless 𝒞{\cal C}\rightarrow\infty.

II.3 The Newtonian flow constants

Flow constants analogous to the ones found by BO78 in the context of GRMHD equilibria are known to exist in Newtonian axisymmetric-stationary systems. These are discussed, for instance, in Mestel’s textbook Mestel (1999) but their origins can be traced back to older work Prendergast (1956); Chandrasekhar (1956); Woltjer (1959). Here we summarise the expressions relating these constants with the various hydromagnetic parameters (adopting standard vectorial notation); a detailed derivation of these expressions can be found in Appendix B.

The defining property of a Newtonian flow constant QQ is 𝐯Q=0\mathbf{v}\cdot\bm{\nabla}Q=0. It is easy to show (see Appendix B) that the same flow constant is a function Q(ψ)Q(\psi), where the magnetic scalar potential ψ\psi is defined via the poloidal field component

𝐁p=ψ×φ.\mathbf{B}_{\rm p}=\bm{\nabla}\psi\times\bm{\nabla}\varphi. (14)

As in GR, the Newtonian constants can be grouped according to their ‘equation of origin’, namely the ideal induction [Eq. (137)] or the Euler equation [Eq. (136)]. In the first group we find the flow constants 𝒞N{\cal C}_{\rm N} and ΩN\Omega_{\rm N} which, as their notation suggests, are analogous to the GR ones 𝒞{\cal C} and Ω\Omega. These appear in the following relation between the magnetic field and fluid velocity

𝐁=𝒞Nρ(𝐯ϖΩN𝝋^),\mathbf{B}={\cal C}_{\rm N}\rho\left(\mathbf{v}-\varpi\Omega_{\rm N}\bm{\hat{\varphi}}\right), (15)

where ϖ=rsinθ\varpi=r\sin\theta. In particular, the poloidal/meridional component of this expression is

𝐁p=𝒞Nρ𝐯p.\mathbf{B}_{\rm p}={\cal C}_{\rm N}\rho\mathbf{v}_{\rm p}. (16)

In the second group we find the following three Bernoulli-like flow constants which result by projecting the Euler equation along the 𝐯\mathbf{v} and 𝐁\mathbf{B} vectors

EN\displaystyle E_{\rm N} =12v2+Φ+hΩN𝒞N4πϖBφ,\displaystyle=\frac{1}{2}v^{2}+\Phi+h-\frac{\Omega_{\rm N}{\cal C}_{\rm N}}{4\pi}\varpi B_{\varphi}, (17)
LN\displaystyle L_{\rm N} =ϖ(vφ𝒞N4πBφ),\displaystyle=\varpi\left(v_{\varphi}-\frac{{\cal C}_{\rm N}}{4\pi}B_{\varphi}\right), (18)
DN\displaystyle D_{\rm N} =12v2+Φ+hϖΩNvφ,\displaystyle=\frac{1}{2}v^{2}+\Phi+h-\varpi\Omega_{\rm N}v_{\varphi}, (19)

where hh is the fluid enthalpy and Φ\Phi is the gravitational potential.

III Corollaries of the flow constant formalism

In this section, we make use of the flow constant formalism to establish some corollaries regarding the possible set of magnetic equilibria.

III.1 The ‘no toroidal field theorem’ and necessity of meridional flow

One of the most surprising consequences of the flow constant MHD formalism is the theorem derived in BO79 under a circular spacetime assumption (see their Sec. VI for more context): “the absence of meridional circulation implies the vanishing of the toriodal [sic] magnetic field.”

The validity of this theorem hinges on the simultaneous presence of the meridional flow and the flow constant 𝒞{\cal C} in the ‘induction-originated’ equation (7). Indeed, setting ui=0u^{i}=0 in (11), and in combination with a non-vanishing poloidal field Bi0B^{i}\neq 0, leads to the requirement 𝒞{\cal C}\to\infty. This divergence would cause the hydrodynamical constants E,LE,L to diverge too, unless [see Eqs. (4), (5)]

Bt\displaystyle B_{t} =gttBt+gtφBφ+gtiBi=0,\displaystyle=g_{tt}B^{t}+g_{t\varphi}B^{\varphi}+g_{ti}B^{i}=0, (20)
Bφ\displaystyle B_{\varphi} =gφφBφ+gtφBt+gφiBi=0.\displaystyle=g_{\varphi\varphi}B^{\varphi}+g_{t\varphi}B^{t}+g_{\varphi i}B^{i}=0. (21)

The determinant of this linear system is non-vanishing in the absence of horizons, and as a consequence the unique solution is the trivial one, Bt=Bφ=0B^{t}=B^{\varphi}=0.

As an additional intuitive argument supporting the theorem’s conclusion, consider how the toroidal field appears to a locally inertial observer. The relevant tetrad component reads B(φ)L(ψ)/𝒞(ψ)B_{(\varphi)}\propto L(\psi)/{\cal C}(\psi) up to metric factors (e.g. Ioka and Sasaki (2004); Ciolfi et al. (2009); see Sec. V). Since 𝒞{\cal C}\to\infty in the ui0u^{i}\to 0 limit, keeping the above non-zero requires that LL\to\infty at an equally fast rate if the observer is to witness a toroidal field. This is clearly incompatible with the hydrodynamic limit, and thus we are forced to conclude that the toroidal field must vanish to maintain consistency.

Therefore a stationary-axisymmetric MHD equilibrium without meridional flow must be purely poloidal as well as purely spacelike. As far as astrophysically relevant MHD equilibria are concerned this is not an acceptable situation as it is well known that purely poloidal fields are generically unstable (e.g. Lasky et al. (2011); Ciolfi and Rezzolla (2012)). This means that a stable mixed poloidal-toroidal configuration should be accompanied by some amount of meridional flow.

As an aside we note that a vanishing meridional flow is still compatible with a purely toroidal field (which is also generically unstable Frieben and Rezzolla (2012); Kiuchi et al. (2011)). Under these circumstances, setting ui=Bi=0u^{i}=B^{i}=0 in (11) does not imply a divergent flow constant 𝒞{\cal C}. Moreover, the vanishing poloidal field implies Frφ=Ftr=0F_{r\varphi}=F_{tr}=0 and again Ωuφ/ut\Omega\neq u^{\varphi}/u^{t} (see Appendix A). The two non-vanishing field components then read Bt=𝒞n(uφΩut)uφB^{t}={\cal C}n(u^{\varphi}-\Omega u^{t})u_{\varphi} and Bφ=𝒞n(uφΩut)utB^{\varphi}=-{\cal C}n(u^{\varphi}-\Omega u^{t})u_{t}; these expressions can be compared with those found in Ref. Kiuchi and Yoshida (2008), which studied nonlinear toroidal equilibria.

III.2 The Grad-Shafranov equation

Without presenting details, which are carefully laid out by IS03, the Euler equation can be shown to lead to the GS equation333The following conversion rule should be applied when adapting formulae from IS to our notation: μIS=μ/m\mu_{\rm IS}=\mu/m, QIS=Q/mQ_{\rm IS}=Q/m where Q={𝒞,L,E,D}Q=\{{\cal C},L,E,D\}.

0=\displaystyle 0= JφΩJt+1𝒞g[(μur),θ(μuθ),r]\displaystyle J^{\varphi}-\Omega J^{t}+\frac{1}{{\cal C}\sqrt{-g}}\left[\,(\mu u_{r})_{,\theta}-(\mu u_{\theta})_{,r}\,\right] (22)
+nuφ(LΛ𝒞)nut[EΛ(𝒞Ω)],\displaystyle+nu^{\varphi}\left(L^{\prime}-\Lambda{\cal C}^{\prime}\right)-nu^{t}\left[\,E^{\prime}-\Lambda({\cal C}\Omega)^{\prime}\,\right],

where

Λ=14π(utBφuφBt)\Lambda=\frac{1}{4\pi}(u_{t}B_{\varphi}-u_{\varphi}B_{t}) (23)

and Jμ=νFμν/4πJ^{\mu}=\nabla_{\nu}F^{\mu\nu}/4\pi is the four-current. Many studies of equilibria concern themselves only with Eq. (22), discarding other information under the assumption that the other ideal MHD equations are trivial. This is not really the case, however, as the number of free functions in (22) make conceptually obvious. As pointed out in Sec. II.2, the flow constant 𝒞{\cal C} arises directly from the induction equation and thus key physical points are missed if one considers (22) in isolation (see also Sec. III.3). One such point was described in Sec. III.1, namely that setting the meridional flow to zero but keeping a mixed poloidal-toroidal field is inconsistent, even though such a choice leaves (22) well behaved. Another concerns the nature of Alfvén points (i.e. where the magnetic energy density matches the kinetic energy density of the flow) and the stellar surface, which we discuss in Sec. III.4.

III.3 The role of the induction equation

Although we work within a relativistic framework in this paper, many of the issues we discuss carry over at a Newtonian level. This is exemplified by the presence of BO-like flow constants in Newtonian systems (Sec. II.3). Owing to the complexity of multifluid dynamics in GR, especially when convective motions and mixed magnetic fields are permitted, we opt to describe some physical features related to induction here in a Newtonian language. In particular, the fact that ‘induction-like’ flow constants are only associated with the electron fluid while the others are associated with the bulk applies to both frameworks. This bares on the physical interpretation of various limits, as discussed throughout, some of which have been addressed in detail by Prix Prix (2005).

Our objective in this section, therefore, is to provide a careful rederivation of the induction-originated 𝒞N{\cal C}_{\rm N} and ΩN\Omega_{\rm N} flow constants by accounting for the distinct proton and electron fluid velocities. These are denoted, respectively, as 𝐯c\mathbf{v}_{\rm c} and 𝐯e\mathbf{v}_{\rm e}. The induction equation itself can be derived from Faraday’s law, ×𝐄=t𝐁/c\bm{\nabla}\times\mathbf{E}=-\partial_{t}\mathbf{B}/c, and the Euler equation for the electron fluid with the inertial terms omitted, i.e.

𝐄=1c𝐯e×𝐁+(μe+meΦ).\mathbf{E}=-\frac{1}{c}\mathbf{v}_{\rm e}\times\mathbf{B}+\bm{\nabla}(\mu_{\rm e}+m_{\rm e}\Phi). (24)

The resulting equation is

t𝐁=×(𝐯e×𝐁).\partial_{t}\mathbf{B}=\bm{\nabla}\times(\mathbf{v}_{\rm e}\times\mathbf{B}). (25)

The manipulations of the ‘single-fluid’ induction equation described in Appendix B can be exactly repeated if we change 𝐯𝐯e\mathbf{v}\to\mathbf{v}_{\rm e} and ρρe\rho\to\rho_{\rm e}. In terms of the poloidal/meridional and toroidal/azimuthal components of the vectors 𝐁,𝐯\mathbf{B},\mathbf{v},

𝐁=𝐁p+Bφ𝝋^,𝐯e=𝐯ep+veφ𝝋^,\mathbf{B}=\mathbf{B}_{\rm p}+B_{\varphi}\bm{\hat{\varphi}},\qquad\mathbf{v}_{\rm e}=\mathbf{v}_{\rm ep}+v_{e\varphi}\bm{\hat{\varphi}}, (26)

we find

𝐁p=𝒞Nρe𝐯ep,\mathbf{B}_{\rm p}={\cal C}_{\rm N}\rho_{\rm e}\mathbf{v}_{\rm ep}, (27)

where 𝒞N{\cal C}_{\rm N} is constant along poloidal field lines, i.e.

𝐁p𝒞N=0.\mathbf{B}_{\rm p}\cdot\bm{\nabla}{\cal C}_{\rm N}=0. (28)

In axisymmetry this is equivalent to

𝐯e𝒞N=0,\mathbf{v}_{\rm e}\cdot\bm{\nabla}{\cal C}_{\rm N}=0, (29)

which shows that 𝒞N{\cal C}_{\rm N} is also conserved with respect to the electron flow.

The poloidal field can be written in a divergence-free form in terms of the scalar potential ψ(r,θ)\psi(r,\theta),

𝐁p=ψ×φ.\mathbf{B}_{\rm p}=\bm{\nabla}\psi\times\bm{\nabla}\varphi. (30)

In this parametrisation the poloidal field lines lie in constant ψ\psi surfaces and therefore the same must be true for the flow lines of 𝐯ep\mathbf{v}_{\rm ep}. As a consequence, 𝒞N=𝒞N(ψ){\cal C}_{\rm N}={\cal C}_{\rm N}(\psi).

Further manipulation of Eq. (25) leads to the flow constant ΩN(ψ)\Omega_{\rm N}(\psi) as part of an expression for the toroidal field

Bφ=𝒞Nρe(vφeϖΩN).B_{\varphi}={\cal C}_{\rm N}\rho_{\rm e}\left(v_{\varphi}^{\rm e}-\varpi\Omega_{\rm N}\right). (31)

In combination with (27), this allows us to write the Newtonian counterpart of Eq. (7) as

𝐁=𝒞Nρe(𝐯eϖΩN𝝋^).\mathbf{B}={\cal C}_{\rm N}\rho_{\rm e}\left(\mathbf{v}_{\rm e}-\varpi\Omega_{\rm N}\bm{\hat{\varphi}}\right). (32)

The final step consists in expressing 𝐯e\mathbf{v}_{\rm e} in terms of the total current 𝐉\mathbf{J} and the proton velocity (which in non-superfluid matter can be taken to coincide with the neutron fluid velocity). Making use of the assumed charge neutrality of the system, ne=ncn_{\rm e}=n_{\rm c}, we have

𝐯e=𝐯c𝐉/ne,\mathbf{v}_{\rm e}=\mathbf{v}_{\rm c}-\mathbf{J}/n_{\rm e}, (33)

which then implies that

𝐁=𝒞Nρe(𝐯cϖΩN𝝋^)𝒞Nme𝐉.\mathbf{B}={\cal C}_{\rm N}\rho_{\rm e}\left(\mathbf{v}_{\rm c}-\varpi\Omega_{\rm N}\bm{\hat{\varphi}}\right)-{\cal C}_{\rm N}m_{\rm e}\mathbf{J}. (34)

Assuming ne=0n_{\rm e}=0 at the surface, this expression can easily accommodate a non-vanishing magnetic field as long as 𝐉0\mathbf{J}\neq 0, thus alleviating any pathological behaviour at the surface (see related discussion in the following section). In practice, this may not be an issue since astrophysical neutron stars are endowed with a non-vacuum magnetosphere in which nen_{\rm e} is non-vanishing Goldreich and Julian (1969). Eq. (34) may also invalidate the BO result applying in the limit 𝐯cp0\mathbf{v}_{\rm cp}\to 0 (see the final paragraph of Appendix B), as 𝒞N{\cal C}_{{\rm N}}\to\infty is no longer necessary to maintain 𝐁p0\mathbf{B}_{p}\neq 0 if the poloidal current 𝐉p\mathbf{J}_{\rm p} does not vanish.

As a final point, although entropy gradients have been ignored here, one generally has 𝐯s=0\mathbf{v}\cdot\nabla s=0 in the single fluid case, meaning that s=s(ψ)s=s(\psi). Yoshida et al. (2012) argue that this forces a vanishing meridional flow else convectively unstable regions with ps<0\nabla p\cdot\nabla s<0 exist in a realistic, stratified star. This is clearly in conflict with the BO conclusion, though the above considerations, which naturally carry over to GR, may remedy the situation.

III.4 The stellar surface and other singularities

In the context of twisted-torus models where one considers the GS equation in isolation, the vanishing density at the stellar surface generally poses no problem to regularity. In fact, in the exterior of a static star where n=0n=0 and there is a purely poloidal ‘test’ field, Eq. (22) decouples with respect to the angular coordinates and admits an analytical solution in terms of hypergeometric functions Konno et al. (1999). Although rotating stars will not be surrounded by vacuum in reality Goldreich and Julian (1969), the vacuum GS equation necessarily describes a force-free magnetosphere (see Ref. Pétri (2016) for a discussion of ‘force-free’ in the GR context).

When considering non-zero velocity fields, the limit n0n\to 0 needs special attention as a result of equations (7) and (8). The two subequations which are most relevant here are written again for convenience, viz.

Bi\displaystyle B^{i} =𝒞n(ut+Ωuφ)ui,\displaystyle=-{\cal C}n\left(u_{t}+\Omega u_{\varphi}\right)u^{i}, (35)
Bφ\displaystyle B^{\varphi} =𝒞n[Ω+(ut+Ωuφ)uφ].\displaystyle=-{\cal C}n\left[\Omega+(u_{t}+\Omega u_{\varphi})u^{\varphi}\right]. (36)

The toroidal field can be assumed to be confined in the stellar interior, rendering Eq. (36) trivial at the surface and beyond, without placing any restriction on 𝒞{\cal C}. The poloidal equations are clearly problematic though at n=0n=0, unless 𝒞{\cal C}\to\infty for every ψ\psi curve that crosses the surface or one carefully treats the multi-fluid dynamics, as in Sec. III.3. This and a related problem can be described in terms of the poloidal Alfvén Mach number444It is somewhat unintuitive that MAM_{\rm A} does not depend explicitly on the magnetic field, though this is because the poloidal flow scales with the poloidal magnetic field, as evident from expression (35).,

MA2=4πμ𝒞2n(gtt+2Ωgtϕ+Ω2gϕϕ).M^{2}_{\rm A}=-\frac{4\pi\mu}{{\cal C}^{2}n\left(g_{tt}+2\Omega g_{t\phi}+\Omega^{2}g_{\phi\phi}\right)}. (37)

From Eqs. (5) and (8), it can be shown that uφu_{\varphi} tends to diverge as MA1M_{\rm A}\to 1 (i.e. at the Alfvén surface) unless LL and Ω\Omega are proportional to each other there (Ioka and Sasaki, 2004). This problem is discussed in the Newtonian context by Mestel Mestel (1968) and Ogilvie Ogilvie (2016) (see also section VI C in Ref. Gourgoulhon et al. (2011)). Unfortunately, this does not really help reduce the pool of flow constants since proportionality is only required in a particular limit. The more obvious problem is that for a finite value of MAM_{\rm A} to be maintained one requires 𝒞2n>0{\cal C}^{2}n>0 everywhere, which would seem to demand 𝒞ψ1{\cal C}\sim\psi^{-1} with a field confined to the interior such that ψ0\psi\to 0 towards the surface (see also Sec. IV). Although this choice was the one adopted in IS04, it is clearly unrealistic.

There are several possibilities worth considering in order to avoid this conundrum:

  • (i)

    The ideal MHD approximation (upon which the BO relations are based) breaks down in the limit n0n\to 0. This is somewhat expected as, in reality, it is meaningless to discuss a current or velocity gradient without matter (e.g., manipulations of the ideal MHD assumption Eμ=Fμνuν=0E_{\mu}=F_{\mu\nu}u^{\nu}=0 become dubious; see Appendix A).

  • (ii)

    A non-relaxed crust alters (or invalidates) some of the BO relations. That is, the term TshearμνT^{\mu\nu}_{\rm shear}, defined through the Carter-Quintana equations Carter and Quintana (1972), may also enter into the total stress-energy tensor. This term will naturally adjust the boundary conditions and flow constant structure (see Ref. Kojima et al. (2022) for some results in the Newtonian context). The presence of an ocean further complicates the dynamics.

  • (iii)

    The function 𝒞(ψ){\cal C}(\psi) behaves in such a way that it tends to infinity everywhere except in an equatorial torus. Suppose that ψ\psi attains a maximum ψc\psi_{c} at the stellar surface, and we write 𝒞=𝒞~/[ψ(ψψc)Θ(ψψc)]{\cal C}={\tilde{\mathcal{C}}}/\left[\psi(\psi-\psi_{c})\Theta(\psi-\psi_{c})\right] for constant 𝒞~{\tilde{\mathcal{C}}} with the meridional flow confined to the region ψ>ψc\psi>\psi_{c}. Outside of this torus (e.g. at the surface) we have 𝒞{\cal C}\to\infty but ui0u^{i}\to 0 and so BiB^{i} from Eq. (35) can be non-zero even if n0n\to 0 there, while inside both uiu^{i} and 𝒞{\cal C} are finite with n>0n>0 and so BiB^{i} is well behaved there also. Although there are mathematical issues with this approach related to the undefined limit of the Heaviside function as ψψc\psi\to\psi_{c}, one has to bear in mind that the use of discontinuous functions is a convenient way of modelling sharp yet continuous gradients of the real physical system.

  • (iv)

    The neutron star is unlikely to be surrounded by vacuum in reality as mentioned previously, arguably rendering this issue moot. Care must be taken in this case to choose appropriate boundary conditions, for instance using the Goldreich-Julian number density nGJBμuμn_{\rm GJ}\sim B_{\mu}u^{\mu} Goldreich and Julian (1969). If n0n\to 0 anywhere in the exterior universe however, this problem reappears at that interface. Note this is related to point (i), as the ideal MHD approximation itself may not be valid in a tenuous magnetosphere.

IV Perturbative equilibria

This section considers perturbative MHD equilibria, where ‘weak’ (see below) magnetic fields are superimposed as perturbations on a non-magnetic background. For most of what follows, the background star is assumed to be static; the rotating case is discussed separately in Sec. IV.3. The hydrostatic metric we use takes the form

ds2\displaystyle ds^{2} =gμνTOVdxμdxν\displaystyle=g_{\mu\nu}^{\rm TOV}dx^{\mu}dx^{\nu}
=e2νdt2+e2λdr2+r2(dθ2+sin2θdφ2),\displaystyle=-e^{2\nu}dt^{2}+e^{2\lambda}dr^{2}+r^{2}\left(d\theta^{2}+\sin^{2}\theta d\varphi^{2}\right), (38)

where we use the superscript ‘TOV’ to indicate that, at a background level, we work with a static and spherically symmetric star whose geometry is described by the TOV equations.

IV.1 General remarks

The following sections consider the leading-order behaviour of the flow constants, essentially through a power-counting analysis. A similar procedure was undertaken by IS04, however they assumed that rotation is always sub-leading and that ψ0\psi\to 0 at the stellar surface (cf. Sec. III.4). One goal of this section is to reexamine their results when these conditions are relaxed.

Firstly, it is instructive to define perturbative when considering an expansion in terms of magnetic parameters. If, as is typical in numerical GR investigations, we further consider ‘units’ of length such that the star has unit mass, M=1M_{\star}=1, the basic measure of magnetic flux becomes the dimensionless number [G cm2]=1.39×1030(1.4M/M~)[\text{G cm}^{2}]=1.39\times 10^{-30}(1.4M_{\odot}/\tilde{M}) for physical-units mass M~\tilde{M}. Given that the magnetic flux is related to the local field strength through ψBR2\psi\approx BR_{\star}^{2}, where RR_{\star} denotes the stellar radius, we see that expansions in powers of ψ\psi are well behaved (ψ1\psi\ll 1) provided that (B/G)(R/cm)2×1.39×1030(1.4M/M~)1(B/\text{G})(R_{\star}/\text{cm})^{2}\times 1.39\times 10^{-30}(1.4M_{\odot}/\tilde{M})\ll 1, which requires B7.2×1017B\ll 7.2\times 10^{17}\,G for R=106R_{\star}=10^{6}\,cm and M~=1.4M\tilde{M}=1.4M_{\odot}. This same limit can be independently derived by demanding that the magnetic pressure at the center of the star is much less than the minimum central hydrostatic pressure. This limit is physically equivalent to a magnetic-to-gravitational-binding energy ratio much below unity.

Although we consider expansions in ψ\psi, it is not obvious that this is the most natural approach when considering a rotating star. In reality, most if not all neutron stars are both ‘slow’ (rotating much slower than the break-up limit) and ‘weakly magnetised’ (as defined above), and it may be more appropriate to consider a double expansion in both uφu^{\varphi} and ψ\psi simultaneously in the form of a magnetic-Hartle-Thorne scheme (see Ref. Steil (2018) for such an approach). We ignore such complications here (though see Footnote 5 below).

IV.2 The magnetic field as a perturbation of a non-rotating star

Here we treat the magnetic field as a perturbation on a static star, with the magnetic stream function ψ\psi used as a perturbative bookkeeping parameter. We assume the following scaling for the magnetic field components

Bi,Bφ=𝒪(ψ),Bt=𝒪(ψγ),B^{i},B^{\varphi}={\cal O}(\psi),\quad B^{t}={\cal O}(\psi^{\gamma}), (39)

for some γ1\gamma\geq 1. To be more precise, the scaling Bi=𝒪(ψ)B^{i}={\cal O}(\psi) is a direct consequence of the definition of ψ=Aφ\psi=A_{\varphi} (see Sec. II.2) and we impose a comparable toroidal field to maintain linearity (cf. Sec. IV.5). This ‘3+1’ scaling accommodates the presence of a mixed poloidal-toroidal field to leading order, while also taking into account the post-Newtonian character of BtB^{t}.

Metric perturbations appear first at 𝒪(ψ2){\cal O}(\psi^{2}) since TEMαβB2T^{\alpha\beta}_{\rm EM}\sim B^{2} Ioka and Sasaki (2003), i.e.

gμν=gμνTOV+𝒪(ψ2).g_{\mu\nu}=g_{\mu\nu}^{\rm TOV}+{\cal O}(\psi^{2}). (40)

As discussed earlier, a magnetic field may induce fluid motion in an otherwise static configuration. We suppose that azimuthal flow appears at some sub-leading order,

uφ=𝒪(ψη),Ω=𝒪(ψη),u^{\varphi}={\cal O}(\psi^{\eta}),\qquad\Omega={\cal O}(\psi^{\eta}), (41)

for η1\eta\geq 1, together with ut=𝒪(1)u^{t}={\cal O}(1) set by the wind equation utut=1u^{t}u_{t}=-1 at leading order. The meridional flow scaling is taken to be

ui=𝒪(ψβ),u^{i}={\cal O}(\psi^{\beta}), (42)

for some βη\beta\geq\eta. The flow constants E,LE,L, and DD are hydrodynamical. That is, they are present even as ψ0\psi\to 0. We Taylor-expand these through

E\displaystyle E =E0+E1ψ+E2ψ2+𝒪(ψ3),\displaystyle=E_{0}+E_{1}\psi+E_{2}\psi^{2}+{\cal O}(\psi^{3}), (43)
L\displaystyle L =L0+L1ψ+L2ψ2+𝒪(ψ3),\displaystyle=L_{0}+L_{1}\psi+L_{2}\psi^{2}+{\cal O}(\psi^{3}), (44)

where E0E_{0} and L0L_{0} are non-magnetic parameters, and the expansion for DD follows from Eq. (6). Last but not least, the flow constant 𝒞{\cal C} is assumed to scale as

𝒞=𝒪(ψα),{\cal C}={\cal O}(\psi^{\alpha}), (45)

where α0\alpha\neq 0, since induction-like constants should not enter at background order. The above scalings are not independent; inserting them in Eqs. (11), (12), (13) and (8), we obtain

ψnutψα+β,\psi\sim nu_{t}\psi^{\alpha+\beta}, (46)
ψnutuφψαnψα+η,\psi\sim nu_{t}u^{\varphi}\psi^{\alpha}\sim n\,\psi^{\alpha+\eta}, (47)
ψγn(Ωutuφ)ψα+η,\psi^{\gamma}\sim n(\Omega u^{t}-u^{\varphi})\psi^{\alpha+\eta}, (48)

and

ψ2n(Bt+ΩBφ)ψαnψα+k,k=min[γ,η+1],\psi^{2}\sim n(B_{t}+\Omega B_{\varphi})\psi^{\alpha}\sim n\psi^{\alpha+k},\,k=\mbox{min}[\gamma,\eta+1], (49)

respectively, where the symbol \sim is used to relate parameters with a non-trivial ψ\psi-scaling. The balance of ψ\psi-powers in the first two expressions leads to

α=1η,β=η.\alpha=1-\eta,\qquad\beta=\eta. (50)

Subsequently, Eqs. (48) and (49) lead to

γ1+η,Ωutuφ=𝒪(ψγ1)𝒪(ψη).\gamma\geq 1+\eta,\qquad\Omega u^{t}-u^{\varphi}={\cal O}(\psi^{\gamma-1})\lesssim{\cal O}(\psi^{\eta}). (51)

The resulting scalings 𝒞ψ1{\cal C}\sim\psi^{-1} and ui,uφψ2u^{i},u^{\varphi}\sim\psi^{2} are consistent with the analysis of IS04 if we set η=2\eta=2. More generally, however, any η>1\eta>1 with α=1η<0\alpha=1-\eta<0 is viable, though demanding integer solutions to permit (Laurent/Taylor) expansions forces η=2\eta=2 and α=1\alpha=-1. The remaining ‘hydrodynamical’ Eqs. (9), (10) are compatible with them and offer no new information about the scalings. It is also easy to verify that

gttTOV(ut)2=1+𝒪(ψ2).g_{tt}^{\rm TOV}(u^{t})^{2}=-1+{\cal O}(\psi^{2}). (52)

One last remark concerns the non-magnetic (ψ0\psi\to 0) limit of the MHD equations. The balance of ψ\psi-powers implemented here means that all of them should automatically behave smoothly in that limit (for example, the above purely magnetic equations reduce to a trivial 0=00=0).

For a static star, the above analysis thus determines the leading-order behaviour of the flow constants irrespective of the true system’s initial conditions.

IV.3 The magnetic field as a perturbation of a rotating star

Much of the previous section’s results depend on the crucial assumption of a non-rotating star in the limit ψ0\psi\to 0. More astrophysically relevant is the scenario where the star is rigidly rotating in the non-magnetic limit and gtφ0g_{t\varphi}\neq 0 (though cf. Footnote 5 below). In such a case,

uφ=𝒪(1),Ω=𝒪(1).u^{\varphi}={\cal O}(1),\qquad\Omega={\cal O}(1). (53)

As rigid rotation may preclude comparable energy partitions, we instead allow for a more general scaling of the toroidal field through

Bφ=𝒪(ψλ),B^{\varphi}={\cal O}(\psi^{\lambda}), (54)

together with555Note that if AtuiA_{t}\sim u^{i} and uφ=𝒪(1)u^{\varphi}=\mathcal{O}(1), the arguments outlined at the beginning of Sec. II.2 would appear to imply Bi=𝒪(ψχ)B^{i}={\cal O}(\psi^{\chi}) with χ=min[β,1]\chi=\text{min}[\beta,1]. This, however, forces λ=0\lambda=0 if β1\beta\leq 1 which leads to a contradiction. Physically speaking though, this simply echoes the sentiment that it does not really make sense to talk about a non-magnetic star with circulating charged particles (i.e. the ‘background’ does not truly exist). On the other hand, if the fluid is composed only of uncharged matter, then the flow constant 𝒞{\cal C} does not exist; cf. Sec. III.3. Bi=𝒪(ψ)B^{i}={\cal O}(\psi). As before, we assume

ui=𝒪(ψβ),𝒞=𝒪(ψα).u^{i}={\cal O}(\psi^{\beta}),\qquad{\cal C}={\cal O}(\psi^{\alpha}). (55)

Upon inserting these in Eqs. (11), (12), and (13) we find

ψ\displaystyle\psi nutψα+β,\displaystyle\sim nu_{t}\psi^{\alpha+\beta}, (56)
ψλ\displaystyle\psi^{\lambda} nutuφψαnψα,\displaystyle\sim nu_{t}u^{\varphi}\psi^{\alpha}\sim n\psi^{\alpha}, (57)
ψγ\displaystyle\psi^{\gamma} n(Ωutuφ)ψα,\displaystyle\sim n(\Omega u^{t}-u^{\varphi})\psi^{\alpha}, (58)

from which we easily deduce that

α=λ,α+β=1,γ=α.\alpha=\lambda,\qquad\alpha+\beta=1,\qquad\gamma=\alpha. (59)

If we have comparable poloidal and toroidal field strengths, then we find that α=1,β=0\alpha=1,~{}\beta=0. That is,

ui=𝒪(1),𝒞=𝒪(ψ),u^{i}={\cal O}(1),\qquad{\cal C}={\cal O}(\psi), (60)

for λ=1\lambda=1. These scalings could be preempted from Eq. (50) with η=0\eta=0, though are markedly different to those obtained for a non-rotating star in Sec. IV.2. In particular, a leading-order meridional flow is an uncomfortable result as it implies a strong deviation from rigid rotation in the non-magnetic limit. This impasse prompts us to explore non-standard scalings for the toroidal field component, in combination with a sub-leading and non-integer-scaling meridional flow, β>0\beta>0. This requirement gives λ=1β<1\lambda=1-\beta<1, which entails a toroidally-dominated MHD equilibrium unless ψ\psi itself is non-perturbative (i.e. ψλ>ψ\psi^{\lambda}>\psi for 0<ψ,λ<10<\psi,\lambda<1). Some discussion on this point is given in Sec. VI.

IV.4 Linearised Grad-Shafranov equation

Here we consider a perturbative and integer expansion of the GS equation (22) without 𝒪(1)\mathcal{O}(1) rotation. Based on the previously obtained scalings and writing

𝒞=𝒞~/ψ,{\cal C}={{\tilde{\mathcal{C}}}}/{\psi}, (61)

where 𝒞~{\tilde{\mathcal{C}}} is a constant parameter, we have

Jφ𝒪(ψ),ΩJt𝒪(ψ2),Λ=14πutBφ+𝒪(ψ5),J^{\varphi}\sim{\cal O}(\psi),\,\,\Omega J^{t}\lesssim{\cal O}(\psi^{2}),\,\,\Lambda=\frac{1}{4\pi}u_{t}B_{\varphi}+{\cal O}(\psi^{5}), (62)
1𝒞g[(μur),θ(μuθ),r]ψβα𝒞~𝒪(ψ3),\frac{1}{{\cal C}\sqrt{-g}}\left[\,(\mu u_{r})_{,\theta}-(\mu u_{\theta})_{,r}\,\right]\sim\frac{\psi^{\beta-\alpha}}{\tilde{{\cal C}}}\sim{\cal O}(\psi^{3}), (63)
nuφ(LΛ𝒞)\displaystyle nu^{\varphi}\left(L^{\prime}-\Lambda{\cal C}^{\prime}\right) =nuφ(L1+2L2ψαΛ𝒞~ψα1)\displaystyle=nu^{\varphi}\left(L_{1}+2L_{2}\psi-\alpha\Lambda\tilde{{\cal C}}\psi^{\alpha-1}\right) (64)
=n𝒞~uφΛψ2+𝒪(ψ2),\displaystyle=n\tilde{{\cal C}}u^{\varphi}\frac{\Lambda}{\psi^{2}}+{\cal O}(\psi^{2}),

and

nut[EΛ(𝒞Ω)]\displaystyle-nu^{t}\left[E^{\prime}-\Lambda({\cal C}\Omega)^{\prime}\right] =nut[E1+2E2ψ\displaystyle=-nu^{t}\Big{[}E_{1}+2E_{2}\psi (65)
Λ𝒞~ψ(ΩΩψ1)]+𝒪(ψ2).\displaystyle-\frac{\Lambda\tilde{{\cal C}}}{\psi}\left(\Omega^{\prime}-\Omega\psi^{-1}\right)\Big{]}+{\cal O}(\psi^{2}).

The above imply that (22) takes the form

𝒪(ψ2)=Jφ+n𝒞~uφΛψ2nut[E1+2E2ψΩ2Λ𝒞~],\mathcal{O}(\psi^{2})=J^{\varphi}+n\tilde{{\cal C}}\frac{u^{\varphi}\Lambda}{\psi^{2}}-nu^{t}\left[E_{1}+2E_{2}\psi-\Omega_{2}\Lambda\tilde{{\cal C}}\right], (66)

with Ω(ψ)=Ω2ψ2\Omega(\psi)=\Omega_{2}\psi^{2}. Upon inserting the approximate Λ\Lambda from Eq. (62), one finally obtains the 𝒪(ψ)\mathcal{O}(\psi) GS equation

0=\displaystyle 0= Jφ+n4π{𝒞~uφutBφψ2\displaystyle J^{\varphi}+\frac{n}{4\pi}\Big{\{}\tilde{{\cal C}}\frac{u^{\varphi}u_{t}B_{\varphi}}{\psi^{2}} (67)
ut[4π(E1+2E2ψ)Ω2utBφ𝒞~]}.\displaystyle-u^{t}\left[4\pi\left(E_{1}+2E_{2}\psi\right)-\Omega_{2}u_{t}B_{\varphi}\tilde{{\cal C}}\right]\Big{\}}.

By using expression (10), one can also rewrite this equation using the flow constant LL (modulo some factors) in exchange for utBφu_{t}B_{\varphi}; see Sec. V.

The first interesting aspect of Eq. (67) we point out is that the term nutE1\sim nu^{t}E_{1} is 𝒪(1){\cal O}(1), and therefore we should expect E1=0E_{1}=0. That is to say, a strict power-counting argument appears to preclude the possibility of E10E_{1}\neq 0, else in the non-magnetic limit (67) unphysically implies that 0=nut0=nu^{t}. The choice E10E_{1}\neq 0 is however made in the literature (where Ei+1E_{i+1} are usually rebranded as cic_{i}), as discussed in the next section.

A second point concerns the explicit appearance of Ω\Omega, which is absent in the linearised equation presented by IS04 [see their Eq. (64)]. We believe this is because these authors implicitly assume η>2\eta>2 rather than η=2\eta=2 as written in their Eq. (58), meaning that all Ω\Omega terms are ignored as being higher order. [Note also the typo in their Eq. (60) in the expansion of D(ψ)D(\psi).]

IV.5 Some remarks on previous literature

In this section, we discuss the perturbative GS equation and compare various approaches towards it found in the literature. Ignoring Ω\Omega corrections at all orders, writing E(ψ)F(ψ)E^{\prime}(\psi)\propto F(\psi) and L0/𝒞~ζL_{0}/{\tilde{\mathcal{C}}}\propto\zeta, one can eventually write the linear GS equation in coordinate form as

0=Jφζ2e2νψr2sin2θ(ϵ+p)F(ψ),0=J_{\varphi}-{\zeta^{2}}e^{-2\nu}\psi-r^{2}\sin^{2}\theta(\epsilon+p)F(\psi), (68)

with F′′(ψ)=0F^{\prime\prime}(\psi)=0. We write Eq. (68) in this way so as to match the 𝒪(ψ)\mathcal{O}(\psi) equation presented by C08 [i.e. their Eq. (57)]. Sometimes the parameter ζ\zeta is upgraded to a general function of ψ\psi, making (68) non-linear (see below).

Strict linearisation of the GS equation limits the possible set of equilibria matching to an exterior poloidal electrovacuum, with the only options being (i) the entire field is confined within the star, (ii) BφB^{\varphi} is discontinuous and there is a non-zero surface current, or (iii) ζ=0\zeta=0 and Bφ=0B^{\varphi}=0 identically. To see this, note that if one matches to an exterior where ψ0\psi\neq 0 then ζ2ψ0\zeta^{2}\psi\neq 0 at the surface unless ζ=0\zeta=0 everywhere or ζ(R)=0\zeta(R_{\star})=0 imposed by hand discontinuously. Option (i) is that selected by IS04 and others (e.g. Haskell et al. (2008)), while (ii) is chosen by C08.

While neither of these options characterise a realistic system, they are the only choices available for a mixed poloidal-toroidal field at strictly linear order. An alternative is to consider non-linear choices for ζ2ψ\zeta^{2}\psi and/or F(ψ)F(\psi), allowing a toroidal field that decays regularly towards the boundary (à la twisted torus), such as in Refs. Ciolfi et al. (2009, 2010); Ciolfi and Rezzolla (2013). However, if one were to take ζψ\zeta\propto\psi (for instance), quantities explicitly related to uiu^{i} should appear in the GS equation since meridional flow enters at 𝒪(ψ3)\mathcal{O}(\psi^{3}). Omitting these terms thus implies that power counting is not handled self-consistently, presenting a conceptual problem. At quadratic order one could avoid the appearance of meridional terms (i.e. taking ζψ\zeta\propto\sqrt{\psi}), though this again likely rules out a regular toroidal component. Moreover, metric perturbations should be accounted for as these appear at 𝒪(ψ2)\mathcal{O}(\psi^{2}), though this is often circumvented with the Cowling approximation.

A related issue surrounds F(ψ)F(\psi); C08 and others (e.g. Ciolfi et al. (2009, 2010); Glampedakis et al. (2012); Ciolfi and Rezzolla (2013)) Taylor-expand this function as

F(ψ)=c0+c1ψ+𝒪(ψ2),F(\psi)=c_{0}+c_{1}\psi+{\cal O}(\psi^{2}), (69)

though declare c0=𝒪(B)c_{0}={\cal O}(B). In C08, the authors study the simpler system (68) with c1=0c_{1}=0, which admits decoupled multipoles. In C09, by contrast, both c0c_{0} and c1c_{1} are kept and the multipole expansion becomes non-trivial (cf. Section V). In either case, however, it seems inconsistent to keep c00c_{0}\neq 0 since this would imply one does not recover 0=00=0 in the non-magnetic limit, ψ0\psi\rightarrow 0.

The above two points show that there are subtle issues involved in the construction of magnetic equilibria, aside from those related to induction explored previously. In the next section we resolve the second of these points, meaning that we numerically build equilibria where c10c_{1}\neq 0 but c0=0c_{0}=0, doing so in the flow-constant language. Solving the non-linear system self-consistently with a twisted-torus configuration, meridional flow, and metric backreactions is left to future work.

V Worked examples of mixed, multipolar MHD equilibria

To start our numerical investigation of the impact of discarding the E1E_{1} coefficient while keeping E2E_{2}, it is convenient to express (67) in component form. While we ignore rotational corrections at background order, it turns out that introducing Ω=Ω2ψ2\Omega=\Omega_{2}\psi^{2} simply leads to a rescaling of the E2E_{2} constant due to Eq. (6), and thus setting Ω2=0\Omega_{2}=0 leads to no loss of generality in our approach. We also opt to replace the constant E2E_{2} by D2D_{2} through Eq. (6) to avoid confusion, as EE_{\ell} is often used to represent the energy stored in an \ell-pole. One eventually finds

0=eν[r(eνλrψ)+eν+λr2sinθθ(θψsinθ)]\displaystyle 0=e^{-\nu}\left[\partial_{r}\left(e^{\nu-\lambda}\partial_{r}\psi\right)+\frac{e^{\nu+\lambda}}{r^{2}}\sin\theta\partial_{\theta}\left(\frac{\partial_{\theta}\psi}{\sin\theta}\right)\right] (70)
+(4πL0𝒞~)2eλ2νψ+8πD2r2sin2θeλ(p+ϵ)ψ,\displaystyle+\left(\frac{4\pi L_{0}}{\tilde{\mathcal{C}}}\right)^{2}e^{\lambda-2\nu}\psi+8\pi D_{2}r^{2}\sin^{2}\theta e^{\lambda}\left(p+\epsilon\right)\psi,

which agrees with previous literature once conventions have been accounted for (e.g., the tttt-component of the TOV metric in C09 is eνe^{\nu} rather than e2νe^{2\nu}).

In order to ensure that the toroidal field is confined to the stellar interior, while maintaining a self-consistently linearised666As discussed in Sec. IV.5, one cannot guarantee that BφB^{\varphi} decays smoothly as rRr\to R_{\star} in a strictly 𝒪(ψ)\mathcal{O}(\psi) setup. It is necessary that nonlinearities enter into Eq. (LABEL:eq:gscomp) via LL or 𝒞{\cal C} to avoid this; the choice L0L0(ψψc)Θ(|ψ/ψc|1)L_{0}\to L_{0}(\psi-\psi_{c})\Theta(|\psi/\psi_{c}|-1), similar to that made by C09 and Ref. Glampedakis et al. (2012), is one possibility. Strict linearisation thus leads to some unrealistic features, whether it be ψ0\psi\rightarrow 0 towards the boundary, like IS04 impose, or a jump discontinuity. equation that is not plagued by singularities at the stellar surface, we impose a linear variant on choice (iii) described in Sec. III.4 in the form

𝒞~𝒞~/Θ(|ψ/ψc|1).{\tilde{\mathcal{C}}}\to{\tilde{\mathcal{C}}}/\Theta(|\psi/\psi_{c}|-1). (71)

Independently of the choice of 𝒞{\cal C} or LL, we note that D20D_{2}\neq 0 prevents a decoupling of the multipolar components in Eq. (LABEL:eq:gscomp). Indeed, the penultimate term behaves like ψ\sim\psi while the last behaves as D2ψsin2θ\sim D_{2}\psi\sin^{2}\theta, breaking the angular symmetry. The obvious choice D2=0D_{2}=0 is not permitted since the boundary conditions cannot be satisfied (see below). Therefore, even the linear GS problem requires a multipolar solution. To make progress, we follow the method of C09 by projecting the equation onto a Legendre basis. We introduce a multipolar expansion

ψ(r,θ)==1maxf(r)sinθdP(cosθ)dθ,\psi(r,\theta)=\sum_{\ell=1}^{\ell_{\rm max}}f_{\ell}(r)\sin\theta\frac{dP_{\ell}(\cos\theta)}{d\theta}, (72)

where a finite max\ell_{\rm max} is necessary because of numerical limitations. Making use of the orthogonality relations

0π𝑑θdP(cosθ)dθdP(cosθ)dθsinθδ,\int^{\pi}_{0}d\theta\frac{dP_{\ell}(\cos\theta)}{d\theta}\frac{dP_{\ell^{\prime}}(\cos\theta)}{d\theta}\sin\theta\propto\delta_{\ell\ell^{\prime}}, (73)

we project Eq. (LABEL:eq:gscomp) into harmonics by multiplying by Legendre polynomials and integrating. This produces a coupled set of max\ell_{\rm max} ordinary differential equations for the functions f(r)f_{\ell}(r) (see Sec. V.1 for numerical details).

The system is closed by setting appropriate boundary conditions and choosing a hydrostatic equation of state (EOS). For the former, we make the standard choices that (i) the field behaves as regularly as possible towards the origin, f(r0)αr+1f_{\ell}(r\rightarrow 0)\rightarrow\alpha_{\ell}r^{\ell+1} for some constants α\alpha_{\ell} [see, e.g., Eq. (24) of C09], and (ii) the field matches smoothly to electrovacuum. The latter condition can be straightforwardly expressed since Eq. (LABEL:eq:gscomp) can be solved analytically when p=ϵ=L0=0p=\epsilon=L_{0}=0 via hypergeometric functions that decay as f,extrf_{\ell,\rm ext}\sim r^{-\ell} with amplitudes representing the electromagnetic multipole moments.

We work with the analytic Tolman-VII solution for simplicity, the details of which are reproduced here for the reader’s convenience. The energy density takes the form (e.g. Jiang and Yagi (2019))

ϵ(r)=ϵc(1x2),\epsilon(r)=\epsilon_{c}\left(1-x^{2}\right), (74)

for central value ϵc=15M/8πR3\epsilon_{c}=15M_{\star}/8\pi R_{\star}^{3} and dimensionless radius x=r/Rx=r/R_{\star}, which implies that grrg_{rr} is found through

e2λ=18π15R2ϵcx2(53x2).e^{-2\lambda}=1-\frac{8\pi}{15}R_{\star}^{2}\epsilon_{c}x^{2}(5-3x^{2}). (75)

The pressure and gttg_{tt} component of the metric are rather more involved, and are given by

p(r)=14πR2[3Ce2λtanϕTC2(53x2)],p(r)=\frac{1}{4\pi R_{\star}^{2}}\left[\sqrt{3Ce^{-2\lambda}}\tan\phi_{\rm T}-\frac{C}{2}\left(5-3x^{2}\right)\right], (76)

and

e2ν=(15C/3)cos2ϕT,e^{2\nu}=\left(1-5C/3\right)\cos^{2}\phi_{\rm T}, (77)

where C=M/RC=M_{\star}/R_{\star} is the compactness and

ϕT=\displaystyle\phi_{\rm T}= arctanC3(12C)+log16+12C3C\displaystyle\arctan\sqrt{\frac{C}{3(1-2C)}}+\log\sqrt{\frac{1}{6}+\sqrt{\frac{1-2C}{3C}}} (78)
12log(x256+e2λ3C).\displaystyle-\frac{1}{2}\log\left(x^{2}-\frac{5}{6}+\sqrt{\frac{e^{-2\lambda}}{3C}}\right).

At the stellar surface (r=R)(r=R_{\star}) we have p=ϵ=0p=\epsilon=0 and the metric functions match to the Schwarzschild exterior. We now have a well-defined boundary problem.

V.1 Numerical methods

A few numerical remarks are in order before we can present a solution. Firstly, integrals involving the Heaviside function, i.e. over the toroidal terms that are radially-rescaled functions of the form

I(r)=0π𝑑θΘ[|ψ(r,θ)ψc|1]ψ(r,θ)dP(cosθ)dθ,I_{\ell}(r)=\int^{\pi}_{0}d\theta\Theta\left[\left|\frac{\psi(r,\theta)}{\psi_{c}}\right|-1\right]\psi(r,\theta)\frac{dP_{\ell}(\cos\theta)}{d\theta}, (79)

cannot be evaluated trivially like the poloidal components (though see Ref. Glampedakis et al. (2012) for an analytic approach). One method is to use a smooth, sigmoid-like approximation for the Heaviside function, which allows II_{\ell} to be evaluated analytically for low values of \ell. In this work, the integrals (79) are approximated with a sum via Simpson’s method. That is, we approximate angular integrals over a small subregion θ(a,b)\theta\in(a,b) via

ab𝑑θX(r,θ)baX(r,a)6+2X(r,a+b2)3+X(r,b)6,\frac{\int^{b}_{a}d\theta X(r,\theta)}{b-a}\approx\frac{X\left(r,a\right)}{6}+\frac{2X\left(r,\frac{a+b}{2}\right)}{3}+\frac{X\left(r,b\right)}{6}, (80)

where the full range 0θ<π0\leq\theta<\pi is subdivided into a uniform grid of nϕn_{\phi} points. Experimentation reveals that a resolution of nϕ=256n_{\phi}=256 produces solutions that are not visibly different from those with nϕ>256n_{\phi}>256 for most cases. However, for fields with L0/𝒞~0.1L_{0}/{\tilde{\mathcal{C}}}\gtrsim 0.1 the toroidal field occupies only a small volume (see Sec. V.2), and so higher resolution is needed to accurately evaluate II_{\ell}.

Solutions to the projected GS equations are achieved via a fourth-order Runge-Kutta method applied to the vector (f1,f2,,fmax)(f_{1},f_{2},\ldots,f_{\ell_{\rm max}}), with step size chosen such that the global error is controlled to be at most one part in 106\gtrsim 10^{6} (see Sec. V.3). The surface boundary condition can be written as (see, e.g., C09)

f(R)=f,ext(R)f,ext(R)f(R),f_{\ell}^{\prime}(R_{\star})=\frac{f_{\ell,\rm ext}^{\prime}(R_{\star})}{f_{\ell,\rm ext}(R_{\star})}f_{\ell}(R_{\star}), (81)

which guarantees we match to some multipolar exterior.

Numerical implementation of the condition f(r0)αr+1f_{\ell}(r\rightarrow 0)\rightarrow\alpha_{\ell}r^{\ell+1} is trickier, and must be achieved by shooting for solutions with the most regularity (i.e. with the shallowest gradients around r=0r=0). We perform a sequence of refining scans over values of D2D_{2} and multipolar coefficients f(R)/f(R)f_{\ell}^{\prime}(R_{\star})/f_{\ell}(R_{\star}), for any given values of the toroidal amplitude L0L_{0} and compactness M/RM_{\star}/R_{\star}, until we converge to a solution that is regular within a specified tolerance. Such a procedure is computationally expensive for max>3\ell_{\rm max}>3 because we must scan over a set of dimension max\ell_{\rm max} (noting that we can fix the dipole moment without loss of generality, though cf. Sec. V.3), and typically we need a precision of order 104\lesssim 10^{-4} in the eigenvalues to avoid divergences. Greater precision is required for strong toroidal fields, else the solutions display a numerical twist and makes it appear as if ψ\psi has additional nodes. Curiously, C08 found that disconnected solutions with multiple nodes are genuine features of dipolar, c00c_{0}\neq 0 equilibria (see their Fig. 2d), while in our case they only appear if the eigenvalues are approximated with insufficient precision. Regardless, keeping terms up to the octupole gives a representation of the true solution to (LABEL:eq:gscomp) at the 10%\lesssim 10\% level, even for fields with strong toroidal fields, as found by C09 and us here.

V.2 Results

Implementing the approach described above, we present mixed poloidal-toroidal and multipolar solutions to (LABEL:eq:gscomp) with a Tolman-VII background (74)–(78). In general, we can plot the field line structure, as measured by a locally inertial observer, using the tetrad components of the field given by Ioka and Sasaki (2004) (up to sign convention; cf. C09)

B(r)=ψ,θr2sinθ,B_{(r)}=-\frac{\psi_{,\theta}}{r^{2}\sin\theta}, (82)
B(θ)=eλrsinθψ,r,B_{(\theta)}=\frac{e^{-\lambda}}{r\sin\theta}\psi_{,r}, (83)

and

B(φ)=4πL0𝒞~rsinθΘ(|ψ/ψc|1)eνψ.B_{(\varphi)}=-\frac{4\pi L_{0}}{{\tilde{\mathcal{C}}}r\sin\theta}\Theta(|\psi/\psi_{c}|-1)e^{-\nu}\psi. (84)

For concreteness, we set the stellar compactness equal to 0.20.2, i.e. R=5MR_{\star}=5M_{\star} in geometric units.

Refer to caption
Refer to caption
Figure 1: Field line structure of the solution to the dipole projection of the linearised GS equation (LABEL:eq:gscomp), for a Tolman-VII background of compactness R=5MR_{\star}=5M_{\star}, where we set L0=0L_{0}=0 (left panel) and L0=0.04𝒞~L_{0}=0.04{\tilde{\mathcal{C}}} (right panel). The colour scale shows the relative magnitude of the poloidal field relative to the (arbitrary) polar field value, with redder shades indicating a stronger field. The shaded ellipsoids at the equator delimit the surface |ψ|ψc|\psi|\geq\psi_{c}, where the toroidal field resides, although B(φ)B_{(\varphi)} vanishes identically in the poloidal case (left). The red annulus depicts the ‘crustal’ region 0.9RrR0.9R_{\star}\leq r\leq R_{\star}.

As a first (relatively crude) approximation, consider the pure dipole case, f2(r)=0f_{\ell\geq 2}(r)=0. Fixing the dipole moment to some arbitrary (though sub-virial) value, we find the purely poloidal (L0=0L_{0}=0) solution with D20.69214D_{2}\approx 0.69214. The field line structure is shown in the left panel of Fig. 1, with the colour scale indicating the poloidal strength relative to the polar value, Bp=|B(r)(R,0)|B_{p}=|B_{(r)}(R_{\star},0)|. Although there is no toroidal field here by construction, we still show the region |ψ|>ψc|\psi|>\psi_{c} for comparison with mixed-field cases. The internal field attains a maximum of 3.2\approx 3.2 times the polar value BpB_{p}, as indicated by the colour scale. Instead setting L0=0.04𝒞~L_{0}=0.04\mathcal{\tilde{C}} to include a modest toroidal field, we find D20.40386D_{2}\approx 0.40386. The associated field line structure is shown in the right panel of Fig. 1.

The strength of the poloidal field is reduced for increasing L0L_{0}. For L0=0.04𝒞~L_{0}=0.04{\tilde{\mathcal{C}}}, for instance, the field achieves a maximum Bmax/Bp2.3B_{\rm max}/B_{p}\approx 2.3 (right panel). Moreover, as found in previous literature though in a different notation, increasing the ratio L0/𝒞~L_{0}/{\tilde{\mathcal{C}}} shrinks the toroidal volume (as is evident from the size of the equatorial rings in Figure 1). The eigenvalue D2D_{2} decreases monotonically as we increase L0L_{0}, though the gradient softens after L0/𝒞~0.1L_{0}/{\tilde{\mathcal{C}}}\gtrsim 0.1. Eventually, if one sets a very large L0L_{0} then BφB_{\varphi} occupies such a small volume that the solution in most of the star is unaffected, and the eigenvalue D2D_{2} asymptotes towards a (max\ell_{\rm max}-dependent) floor value of D2,min0.236D_{2,\rm min}\approx 0.236. This is simply a consequence of the linearisation (cf. Refs. Glampedakis et al. (2012); Ciolfi and Rezzolla (2013)); choosing more complicated flow constants allows one to control the toroidal volume and strength simultaneously (see Sec. V.3).

We next construct a purely poloidal but multipolar solution with max=3\ell_{\rm max}=3. Although we could include a quadrupolar term, likely important for astrophysical systems as evidence from NICER and hotspot modelling suggests that hemispherically asymmetric magnetic geometries are prevalent in Nature (such as in PSR J0030+0451 Bilous et al. (2019)), the trivial solution f2(r)=0f_{2}(r)=0 is permitted because the even and odd parity sectors decouple, as also noted by C09. Although equilibria with f2(r)0f_{2}(r)\neq 0 exist777The situation is subtle. Mixed polarity solutions exist only for sufficiently large values of L0L_{0} given a compactness. To see this, note that in the purely poloidal case the equations fully decouple and unless the eigenvalue D2D_{2} happens to match in both equations there will be no solution. Although we do not show the field structure, an example of a nodeless, dipole-dominated solution is L0=0.06𝒞~L_{0}=0.06{\tilde{\mathcal{C}}}, D20.2936D_{2}\approx 0.2936, and f1(R)/f2(R)90.11f_{1}(R_{\star})/f_{2}(R_{\star})\approx 90.11., we focus on odd-parity multipoles for simplicity.

Including an octupole moment, the numerical scan reveals D20.6750D_{2}\approx 0.6750 and f1(R)/f3(R)22.7795f_{1}(R_{\star})/f_{3}(R_{\star})\approx 22.7795 for L0=0L_{0}=0. The field structure is shown in the left panel of Fig. 2. The largeness of f1(R)/f3(R)f_{1}(R_{\star})/f_{3}(R_{\star}) indicates that the octupole contribution is relatively weak, though its impact in the core is evident because the field maximum increases to B/Bp3.7B/B_{p}\approx 3.7 (15%\approx 15\% larger than the dipole case; Fig. 1). A mixed field solution with L0=0.04𝒞~L_{0}=0.04{\tilde{\mathcal{C}}}, requiring D20.3905D_{2}\approx 0.3905 and f1(R)/f3(R)10.320f_{1}(R_{\star})/f_{3}(R_{\star})\approx 10.320, is depicted in the right panel. As before, the volume of the toroidal region decreases as we increase L0L_{0} and the poloidal field decreases in strength (Bmax/Bp2.6B_{\rm max}/B_{p}\approx 2.6). Additionally, including a toroidal component also requires that the octupole contribution becomes stronger, with the ‘kink’ in the field lines near polar latitudes around r0.9Rr\sim 0.9R_{\star} becoming more prominent.

Refer to caption
Refer to caption
Figure 2: Similar to Fig. 1, though including octupole contributions, for L0=0L_{0}=0 (left) and L0=0.04𝒞~L_{0}=0.04{\tilde{\mathcal{C}}} (right).

Larger values of L0L_{0} demand smaller ratios of f1(R)/f3(R)f_{1}(R_{\star})/f_{3}(R_{\star}); for example, for L0=0.1𝒞~L_{0}=0.1\tilde{\mathcal{C}} we find D20.234492D_{2}\approx 0.234492 and f1(R)/f3(R)6.0101f_{1}(R_{\star})/f_{3}(R_{\star})\approx 6.0101. This solution, depicted in Fig. 3, is interesting because the toroidal center migrates towards the surface and shrinks in volume enough that BφB_{\varphi} becomes confined to the region 0.9x10.9\lesssim x\leq 1 representing a (hypothetical) crust. Such equilibria may be more realistic than ones with core-dominated fields owing to the torque instability described in Ref. Glampedakis and Lasky (2015). Similar to Fig. 2, the solution we find is nodeless and the toroidal field is confined strictly by the neutral curves, unlike some of the peculiar solutions found by C08 and C09. Although not shown, a range of equilibria with varying L0L_{0} were constructed and we found that the nodeless property persisted. In C08, by contrast, there are discrete ranges (of ζ\zeta) such that magnetically-disconnected solutions exist (see their Fig. 2). In cases with mixed polarity, we find solutions with some minimum number of nodes only.

We conclude by noting that the important distinction with cases where D10D_{1}\neq 0 (i.e. c00c_{0}\neq 0) is that solutions constructed for D1=0D_{1}=0 are unique, for a given EOS, because there are no free parameters. In C09, for instance, they were able to find solutions with an integer number of nodes in ψ\psi because of the c00c_{0}\neq 0 freedom (see their Fig. 6). One might argue that solutions where ψ0\psi\rightarrow 0 internally, at least for odd-multipoles, are physically undesirable as they indicate a singular limit of the flow constant formalism (cf. IS04 and Sec. III.4). In our case, it seems that only the minimum node solution can be found, indicating that that these extra solutions are artifacts of the additional freedom. Although this is difficult to prove rigorously, we have performed fine scans over the parameter space and can only consistently converge to the eigenvalues given above; no other choices give a regular solution. Overall though, the main quantitative features of our solutions are similar to those found in the aforementioned references.

Refer to caption
Figure 3: Similar to Fig. 2, though with a stronger toroidal field, viz. L0=0.1𝒞~L_{0}=0.1\tilde{\mathcal{C}}, that is confined to the ‘crust’.

Although we could continue this sequence of solutions by increasing max\ell_{\rm max}, the field line geometry is unlikely to change dramatically because the ratio f(R)/f1(R)f_{\ell}(R_{\star})/f_{1}(R_{\star}) decreases as a function of \ell (as found by C09).

V.2.1 Meridional flow

Even though the meridional flow does not enter into Eq. (LABEL:eq:gscomp) at linear order, its presence is guaranteed by the existence of a toroidal field (see Sec. III.1). The tetrad components of the velocity vector can be found through [see, e.g., Eqns. (135–137) in IS04]

v(r)=Θ(|ψ/ψc|1)𝒞~(ϵ+p)r2e2νψψθ,v_{(r)}=\frac{\Theta(|\psi/\psi_{c}|-1)}{\tilde{\mathcal{C}}\left(\epsilon+p\right)r^{2}}e^{-2\nu}\psi\frac{\partial\psi}{\partial\theta}, (85)
v(θ)=Θ(|ψ/ψc|1)𝒞~(ϵ+p)rsinθeλ2νψψr,v_{(\theta)}=-\frac{\Theta(|\psi/\psi_{c}|-1)}{{\tilde{\mathcal{C}}\left(\epsilon+p\right)r\sin\theta}}e^{-\lambda-2\nu}\psi\frac{\partial\psi}{\partial r}, (86)

and

v(φ)=eν[ΩrsinθL0Θ(|ψ/ψc|1)e2ν4π𝒞~2(ϵ+p)rsinθψ2].v_{(\varphi)}=e^{-\nu}\left[\Omega r\sin\theta-\frac{L_{0}\Theta(|\psi/\psi_{c}|-1)e^{-2\nu}}{4\pi\tilde{\mathcal{C}}^{2}\left(\epsilon+p\right)r\sin\theta}\psi^{2}\right]. (87)

It is interesting to remark that B(φ)B_{(\varphi)} depends on the combination L0/𝒞~L_{0}/\tilde{\mathcal{C}}, while the meridional components (85) and (86) depend only on 𝒞~\tilde{\mathcal{C}}. This is again an artifact of the linearisation procedure, and shows that a family of meridional flows of essentially arbitrary amplitude exist for a given ψ\psi (provided the perturbative criteria is satisfied).

Fig. 4 shows the meridional flow, in the case where ψ\psi corresponds to the dipole shown in the right panel of Fig. 1. Even though the denominator in the flow components (85) and (86) diverges at the stellar surface, the overall velocity vanishes there because of Eq. (71).

Refer to caption
Figure 4: Meridional flow, restricted to the toroidal region, for the solution shown in the right panel of Fig. 1. For improved visibility, only the eastern hemisphere of the star is shown and the ‘crustal’ region is not depicted. The colour scale shows the relative amplitude of the (clockwise) flow, normalised by an arbitrary constant as described in text.

V.3 Regular toroidally-dominated configurations free of surface currents

As described in Sec. V.2, linearisation imposes a limit to the toroidal energy stored in the field: increasing L0L_{0} in an attempt to boost B(φ)B_{(\varphi)} shrinks the toroidal volume. Moreover, the strict linearisation of the GS equation implies that it is not possible to have a regular toroidal field which is non-zero internally and vanishes externally if ψ0\psi\neq 0 at the boundary. These restrictions can be lifted simultaneously if we allow for a non-linear E(ψ)E(\psi) and L(ψ)L(\psi), though the price to pay is self-consistency with power-counting, as detailed in Sec. IV.5. Nevertheless, we can compute solutions with non-linear flow constants. Motivated by the choices made in Ref. Ciolfi and Rezzolla (2013), though again setting E1=0E_{1}=0, we consider the functions

E(ψ)=\displaystyle E(\psi)= E2ψ[k1(|ψ/ψc|1)κΘ(|ψ/ψc|1)\displaystyle E_{2}\psi\Big{[}k_{1}\left(\left|{\psi}/{\psi_{c}}\right|-1\right)^{\kappa}\Theta\left(\left|{\psi}/{\psi_{c}}\right|-1\right) (88)
k2L(ψ)k3],\displaystyle-k_{2}L(\psi)-k_{3}\Big{]},
L(ψ)=L0ψ(|ψ/ψc|1)Θ(|ψ/ψc|1),L(\psi)=L_{0}\psi\left(\left|{\psi}/{\psi_{c}}\right|-1\right)\Theta\left(\left|{\psi}/{\psi_{c}}\right|-1\right), (89)

for parameters L0,κ,k1,k2L_{0},\kappa,k_{1},k_{2}, and k3k_{3}. The choice k1=k2=0k_{1}=k_{2}=0 returns (43) with a rescaled value of E2E_{2}. The motivation for the above inclusions are that one wishes to enhance the region of closed field lines – achieved through k1k_{1} and κ\kappa – while allowing the toroidal energy density to grow by minimising the azimuthal current JφJ_{\varphi} – achieved through k2k_{2} (see Ref. Ciolfi and Rezzolla (2013) for a discussion). Furthermore, the non-linear choice (89) (also made in other works, e.g. Glampedakis et al. (2012)) ensures that the toroidal field decays smoothly towards the stellar surface for suitable κ\kappa, thus preventing surface currents. A dipolar solution to Eq. (22), though again at the expense of ignoring meridional components and metric corrections, is found using the method described in Sec. V.1 for the choices κ=4\kappa=4, k1=1k_{1}=1, k2=10k_{2}=10, k3=0.1k_{3}=-0.1, and L0=1.1𝒞~L_{0}=1.1{\tilde{\mathcal{C}}}. The eigenvalue we converge to is D29.49276D_{2}\approx 9.49276. The field components, exhibiting a strongly dominant B(φ)B_{(\varphi)}, are shown in Fig. 5. Note that because we consider a non-linear GS equation here, the exact normalisation of the dipole moment affects the solution888It is clear from these solutions that non-linear dynamics are important as concerns the magnetic field itself even for amplitudes well below the virial limit. The impact on the stellar structure is tiny however, even though metric corrections appear at 𝒪(ψ2)\mathcal{O}(\psi^{2}).. The figures here correspond to a physical value of Bp5×1014B_{p}\gtrsim 5\times 10^{14}\,G.

The combined effect of these choices is that the toroidal field is both strong and vast. Furthermore, B(φ)B_{(\varphi)} decays smoothly to zero as ψψc\psi\to\psi_{c}, indicating an absence of surface currents and discontinuities. Direct integration reveals that the toroidal-to-total energy ratio is

EtorEtot=𝑑VBφ2/8π𝑑VB2/8π=0.956,\frac{E_{\rm tor}}{E_{\rm tot}}=\frac{\int dVB_{\varphi}^{2}/8\pi}{\int dVB^{2}/8\pi}=0.956, (90)

meaning that the toroidal field houses 95.6%\approx 95.6\% of the total magnetic energy. Such a configuration may be relevant for magnetars. For instance, 36\approx 36\,ks phase-modulations in the X-ray pulses of 1E 1547.0–5408 have been consistently observed Makishima et al. (2021), possibly indicating a freely precessing source. Since the spin period of this object is P=2.087P=2.087\,s, the modulation periodicity suggests a quadrupolar ellipticity of order 6×105\sim 6\times 10^{-5}, and thus a toroidal field of strength B(φ)1016B_{(\varphi)}\sim 10^{16}\,G since the polar field strength is only Bp1014B_{p}\sim 10^{14}\,G (see also Ref. Suvorov (2023)).

Though more formal checks can be made with the GR virial relations (e.g. Kiuchi and Yoshida (2008)), the extent of numerical errors, quantified by the left-right mismatch in the radial projection(s) of Eq. (LABEL:eq:gscomp) in dimensionless units (i.e. the residual), is shown in Fig. 6. The local error reaches 109\sim 10^{-9} at worst, indicating a reliable solution.

Refer to caption
Figure 5: Tetrad components of the magnetic field – B(r)B_{(r)} (dotted), B(θ)B_{(\theta)} (dashed), and B(φ)B_{(\varphi)} (solid) evaluated at either the equator or pole – for a solution to the non-linear GS equation with the flow constant choices (88) and (89), with κ=4\kappa=4, k1=1k_{1}=1, k2=10k_{2}=10, k3=0.1k_{3}=-0.1, and L0=1.1𝒞~L_{0}=1.1{\tilde{\mathcal{C}}}.
Refer to caption
Figure 6: Residual of the numerical solution shown in Fig. 5. This plot is typical for solutions shown in this work.

VI Discussion

In this paper we revisit the impact of ‘flow constants’, introduced by BO in the GR context Bekenstein and Oron (1978, 1979), on the construction of stationary and axisymmetric neutron star equilibria. The main results of this work are two-fold. The first part presents a modern review of the general formalism in both GR and Newtonian gravity (Sec. II), spelling out important corollaries at the general, non-linear level (Sec. III). One key argument we present is that consideration of the GS equation alone gives the misleading impression that terms related to meridional flows can be discarded: such a choice is inconsistent with a mixed poloidal-toroidal field according to the BO theorem (Sec. III.1), required in canonical twisted-torus models and from a stability standpoint. The second part explores perturbative constructions, where an expansion in the sub-virial flux ψ\psi is taken (Sec. IV), with numerical equilibria built in Sec. V. We find that power-counting arguments suggest that some choices for flow constants made in previous literature are not, strictly speaking, self-consistent. Fortunately though, our numerical equilibria for static stars, where strict order-counting is enforced, are both qualitatively and quantitatively similar to existing models (see, e.g., Refs. Glampedakis et al. (2012); Colaiuda et al. (2008); Ciolfi et al. (2009); Ciolfi and Rezzolla (2013)).

In considering the perturbative equilibria of stars that are rotating at background order, we find that power-balancing relations for the flow constants predict that the toroidal field and meridional flow scale as Bφ=𝒪(ψλ)B^{\varphi}=\mathcal{O}(\psi^{\lambda}) and ui=𝒪(ψ1λ)u^{i}=\mathcal{O}(\psi^{1-\lambda}), respectively, with 0<λ10<\lambda\leq 1 (Sec. IV.3). Since |ψ|1|\psi|\ll 1 for the perturbative scheme to be valid, this implies that the toroidal field is larger than the poloidal one, though not necessarily in the global sense of total energies: the toroidal field may be locally stronger but confined to a small volume (cf. Sec. V.2). It is generally thought that strong toroidal fields are necessary to explain frequent magnetar outbursts Pons and Perna (2011), a conclusion which is supported by stability theory Akgün et al. (2013) together with observations of large pulse fractions Igoshev et al. (2021) and 10\gtrsim 10\,ks phase modulations in X-ray lightcurves Makishima et al. (2021). Based on this, one may conjecture that magnetar birth conditions preferentially select values of λ1\lambda\ll 1 while ‘ordinary’ neutron stars have λ1\lambda\lesssim 1. However, in modelling rotating stars one must be careful when taking expansions for the flow constants [such as in Eq. (43)] as simple Taylor series are precluded by the necessity of non-integer powers (i.e. one should use Puiseux series instead). For future work, it would be worthwhile considering self-consistent GS equilibria at successive (fractional) orders in ψ\psi, including meridional flow, metric corrections, and multi-fluids (see the discussion in Sec. III.3).

Given that the equilibria constructed here are moulded around twisted tori, we expect that the stability results regarding such configurations carry over Braithwaite and Spruit (2006). Whether mixed poloidal-toroidal fields are long-term stable is not a fully settled matter though (especially in GR), as non-axisymmetric instabilities may be endemic to barotropic stars Lander and Jones (2012) while convective instabilities can operate in non-barotropic systems with meridional flows Yoshida et al. (2012).

In a full, time-dependent simulation, one should be able to specify initial data which eventually translate into flow constants when equilibrium is reached, if indeed it ever is. While in considering the time-independent GRMHD equations one is free to choose the flow constants arbitrarily, it is far from obvious whether a star set up with some initial data would actually reach such a state. This is highlighted by recent ‘long-term’ GRMHD simulations, where late-time behaviour depends sensitively on the initial energy partition Sur et al. (2022). Even if one ignores meridional flows and rotation, the toroidal function is totally free and rich families of GS-equilibria exist Glampedakis and Lasky (2016). Providing a definitive answer to how these constants behave given some initial data, or which combinations of them are stable, is an extremely difficult problem which we do not solve here. A future approach to this kind of ‘inverse problem’ would be to try and fit the flow constants to (stable) numerical solutions. To this end, a crucial ingredient, not present in our investigation, is the magnetic helicity: it has been argued that magnetic fields evolve so as to minimise changes in the global helicity (i.e. to be as conserved as possible) Spruit (2008); Ciolfi et al. (2010).

The existence of meridional flow has important physical consequences. One effect was considered by IS04, who noted that non-circularity induces additional types of frame dragging into spacetime which violate reflection symmetry about the equator. Such effects could, in principle, be tied to equatorially-asymmetric hotspot formation Bilous et al. (2019) or natal kicks Ioka and Sasaki (2004).

The meridional flow could lead to an additional channel of magnetic diffusion via the viscous damping of the internal circulation. The associated dissipation timescale is identical to the standard viscous timescale, tviscR2ρ/ηt_{\rm visc}\sim R^{2}_{\star}\rho/\eta, where η\eta is the shear viscosity coefficient. As an example, we consider mature neutron stars with a neutron superfluid, in which case viscosity is dominated by electron-electron scattering. The corresponding viscosity coefficient is ηee6×1020ρ152T82g/cm s\eta_{\rm ee}\approx 6\times 10^{20}\rho_{15}^{2}T_{8}^{-2}\,\mbox{g}/\mbox{cm s} Cutler and Lindblom (1987), where T8=T/108KT_{8}=T/10^{8}\,\mbox{K} and ρ15=ρ/1015gcm3\rho_{15}=\rho/10^{15}\,\mbox{g}\,\mbox{cm}^{-3}, and we find tvisc0.05ρ151T82yrt_{\rm visc}\sim 0.05\,\rho_{15}^{-1}T_{8}^{2}\,\mbox{yr} (we note that the corresponding timescale for non-superfluid matter is of the same order of magnitude with a modified density scaling ρ1ρ5/4\rho^{-1}\to\rho^{-5/4}). This timescale is much shorter than typical neutron star ages. Since the poloidal field and meridional flow are proportional to each other [Eq. (11)], this naïvely suggests that poloidal fields should be comparatively weak in mature stars, unless somehow regenerated by the toroidal field through MHD exchanges or if the proportionality factor 𝒞{\cal C} adjusts in tandem. Either way, time-dependent effects of a meridional flow could have a profound influence on the magnetic evolution and observational appearance of neutron stars. Future observations of gravitational waves may elucidate the situation.

Acknowledgements.
We thank Luciano Rezzolla, Riccardo Ciolfi, and the anonymous referee for comments. K. G. acknowledges support from research grant PID2020-1149GB-I00 of the Spanish Ministerio de Ciencia e Innovación. A. G. S. was supported by the Alexander von Humboldt Foundation and the European Union’s Horizon 2020 Programme under the AHEAD2020 project (Grant No. 871158) during the early stages of this work.

Appendix A Flow constant derivation (GR)

This Appendix provides a self-contained derivation of the flow constants originally discussed in BO78 & BO79. Under the assumption of a stationary-axisymmetric system we have Q,t=Q,φ=0Q_{,t}=Q_{,\varphi}=0 where QQ stands for any geometric or electromagnetic function. Note that nothing else is assumed about the metric or the fluid flow.

The first group of flow constants does not depend on the Euler equation and therefore is a good starting point for our analysis. By definition, we have

Ftφ=0,\displaystyle F_{t\varphi}=0, (91)
Frφ,θ=Fθφ,r,\displaystyle F_{r\varphi,\theta}=F_{\theta\varphi,r}, (92)
Ftr,θ=Ftθ,r.\displaystyle F_{tr,\theta}=F_{t\theta,r}. (93)

The t,r,φt,r,\varphi components of the ideal MHD condition Eμ=Fμνuν=0E_{\mu}=F_{\mu\nu}u^{\nu}=0 lead to

Ftθuθ+Ftrur=0,\displaystyle F_{t\theta}u^{\theta}+F_{tr}u^{r}=0, (94)
Frφur+Fθφuθ=0,\displaystyle F_{r\varphi}u^{r}+F_{\theta\varphi}u^{\theta}=0, (95)
Frtut+Frθuθ+Frφuφ=0.\displaystyle F_{rt}u^{t}+F_{r\theta}u^{\theta}+F_{r\varphi}u^{\varphi}=0. (96)

Ignoring the singular possibility ui=0u^{i}=0 for now, we may solve (94) and (95) for Ftθ,FθφF_{t\theta},F_{\rm\theta\varphi} and insert the results in (93) and (92). We find, respectively

ddτlogFtr\displaystyle\frac{d}{d\tau}\log F_{tr} =urFtr,rFtr+uθFtr,θFtr=uθ(uruθ),r\displaystyle=u^{r}\frac{F_{tr,r}}{F_{tr}}+u^{\theta}\frac{F_{tr,\theta}}{F_{tr}}=-u^{\theta}\left(\frac{u^{r}}{u^{\theta}}\right)_{,r} (97)
ddτlogFrφ\displaystyle\frac{d}{d\tau}\log F_{r\varphi} =urFrφ,rFrφ+uθFrφ,θFrφ=uθ(uruθ),r\displaystyle=u^{r}\frac{F_{r\varphi,r}}{F_{r\varphi}}+u^{\theta}\frac{F_{r\varphi,\theta}}{F_{r\varphi}}=-u^{\theta}\left(\frac{u^{r}}{u^{\theta}}\right)_{,r} (98)

It follows that

ddτlog(FrφFtr)=0Ftr/Frφ=Ω,\frac{d}{d\tau}\log\left(\frac{F_{r\varphi}}{F_{tr}}\right)=0~{}\Rightarrow~{}F_{tr}/F_{r\varphi}=\Omega, (99)

where Ω\Omega is a flow constant. Using this in (94) and (95)

Ftθ/Fθφ=Ω.F_{t\theta}/F_{\theta\varphi}=\Omega. (100)

Meanwhile, we can rewrite (98) as

ddτlogFrφ=u,μμ+uμu,μθuθ.\frac{d}{d\tau}\log F_{r\varphi}=-u^{\mu}_{~{},\mu}+u^{\mu}\frac{u^{\theta}_{~{},\mu}}{u^{\theta}}. (101)

The baryon conservation equation μ(nuμ)=0\nabla_{\mu}(nu^{\mu})=0 can be written in the equivalent form u,μμ=d[log(gn)]/dτu^{\mu}_{~{},\mu}=-d[\log(\sqrt{-g}n)]/d\tau and the previous expression becomes

ddτlog(Frφgnuθ)=0Fφrgnuθ=𝒞,\frac{d}{d\tau}\log\left(\frac{F_{r\varphi}}{\sqrt{-g}nu^{\theta}}\right)=0~{}\Rightarrow~{}\frac{F_{\varphi r}}{\sqrt{-g}nu^{\theta}}={\cal C}, (102)

where 𝒞{\cal C} is another flow constant. This result can be combined with (95) to similarly show that

Fθφgnur=𝒞.\frac{F_{\theta\varphi}}{\sqrt{-g}nu^{r}}={\cal C}. (103)

Finally, (96) in combination with (99), (102) leads to

Frθ=g𝒞n(uφΩut).F_{r\theta}=\sqrt{-g}{\cal C}n(u^{\varphi}-\Omega u^{t}). (104)

Using the above expressions in the definition of the magnetic field we find,

Bi=𝒞n(ut+Ωuφ)ui.B^{i}=-{\cal C}n(u_{t}+\Omega u_{\varphi})u^{i}. (105)

In fact, using the definition of BμB^{\mu}, we can produce the more general four-vectorial expression

Bα=𝒞n[(ut+Ωuφ)uα+δtα+Ωδφα].B^{\alpha}=-{\cal C}n\left[\,(u_{t}+\Omega u_{\varphi})u^{\alpha}+\delta_{t}^{\alpha}+\Omega\delta_{\varphi}^{\alpha}\,\right]. (106)

This is a key relation between the magnetic field, the flow constants, and the fluid parameters. With the help of this relation we can prove the following interesting property for any constant dQ/dτ=uαQ,α=0dQ/d\tau=u^{\alpha}Q_{,\alpha}=0:

BαQ,α=𝒞n[(ut+Ωuφ)uαQ,α+Q,t+ΩQ,φ]=0,B^{\alpha}Q_{,\alpha}=-{\cal C}n\left[(u_{t}+\Omega u_{\varphi})u^{\alpha}Q_{,\alpha}+Q_{,t}+\Omega Q_{,\varphi}\right]=0, (107)

meaning that BαB^{\alpha} and QQ share the same level surfaces.

The final conservation law of this first group relates to the magnetic energy. It originates from the orthogonality between the Lorentz force and the four-velocity, i.e.

uαβTEMαβ=uαFαβJβ=0.u_{\alpha}\nabla_{\beta}T^{\alpha\beta}_{\rm EM}=-u_{\alpha}F^{\alpha\beta}J_{\beta}=0. (108)

Use of (3) for TEMμνT_{\rm EM}^{\mu\nu} together with the orthogonality property uαBα=0u_{\alpha}B^{\alpha}=0 allows us to rewrite this condition as

ddτB2=2B2αuα2uαBββBα.\frac{d}{d\tau}B^{2}=-2B^{2}\nabla_{\alpha}u^{\alpha}-2u_{\alpha}B^{\beta}\nabla_{\beta}B^{\alpha}. (109)

With the help of αuα=uαn,α/n\nabla_{\alpha}u^{\alpha}=-u^{\alpha}n_{,\alpha}/n and uαBββBα=BαBββuαu^{\alpha}B^{\beta}\nabla_{\beta}B_{\alpha}=-B^{\alpha}B^{\beta}\nabla_{\beta}u_{\alpha} this expression becomes

ddτB2\displaystyle\frac{d}{d\tau}B^{2} =2B2ddτlogn+2BαBββuα.\displaystyle=2B^{2}\frac{d}{d\tau}\log n+2B^{\alpha}B^{\beta}\nabla_{\beta}u_{\alpha}. (110)

Expanding the covariant derivative in the first and third term,

0=B2ddτlogn+BαBβu,βαBαB,βαuβ.0=B^{2}\frac{d}{d\tau}\log n+B_{\alpha}B^{\beta}u^{\alpha}_{~{},\beta}-B_{\alpha}B^{\alpha}_{~{},\beta}u^{\beta}. (111)

Moreover, we can exploit another equivalent expression for dB2/dτdB^{2}/d\tau in the form

uβBαB,βα=ddτB2uβBαBα,β,u^{\beta}B_{\alpha}B^{\alpha}_{~{},\beta}=\frac{d}{d\tau}B^{2}-u^{\beta}B^{\alpha}B_{\alpha,\beta}, (112)

together with BβBαu,βα=uαBβBα,βB^{\beta}B_{\alpha}u^{\alpha}_{,~{}\beta}=-u^{\alpha}B^{\beta}B_{\alpha,\beta}. Inserting these in (111) and after some rearrangement we obtain

ddτlog(B2n)=1B2(BαdBαdτuβBiBβ,i).\frac{d}{d\tau}\log\left(\frac{B^{2}}{n}\right)=\frac{1}{B^{2}}\left(B^{\alpha}\frac{dB_{\alpha}}{d\tau}-u^{\beta}B^{i}B_{\beta,i}\right). (113)

After expanding the right-hand-side terms and using (106) we can finally arrive to a ‘magnetic energy’ conservation law

ddτ[B2n+𝒞(Bt+ΩBφ)]=0.\frac{d}{d\tau}\left[\frac{B^{2}}{n}+{\cal C}(B_{t}+\Omega B_{\varphi})\right]=0. (114)

This implies the existence of a new flow constant {\cal F},

B2n+𝒞(Bt+ΩBφ)=.\frac{B^{2}}{n}+{\cal C}(B_{t}+\Omega B_{\varphi})={\cal F}. (115)

In the second group of conservation laws we find expressions derivable from the Euler equation. Projecting that equation along BμB^{\mu}

(ϵ+p+B24π)Bαaα=Bαp,α+B24παBα.\left(\epsilon+p+\frac{B^{2}}{4\pi}\right)B_{\alpha}a^{\alpha}=-B^{\alpha}p_{,\alpha}+\frac{B^{2}}{4\pi}\nabla_{\alpha}B^{\alpha}. (116)

For the divergence of the magnetic field we can write

αBα=12gϵαβγδαuβFγδ.\nabla_{\alpha}B^{\alpha}=\frac{1}{2\sqrt{-g}}\epsilon^{\alpha\beta\gamma\delta}\nabla_{\alpha}u_{\beta}F_{\gamma\delta}. (117)

This can be inverted

12gϵαβγδFγδ=uαBβuβBα,\frac{1}{2\sqrt{-g}}\epsilon^{\alpha\beta\gamma\delta}F_{\gamma\delta}=u^{\alpha}B^{\beta}-u^{\beta}B^{\alpha}, (118)

and we can rewrite (117) in the far more elegant form,

αBα=Bαuββuα.\nabla_{\alpha}B^{\alpha}=B^{\alpha}u^{\beta}\nabla_{\beta}u_{\alpha}. (119)

The same is true for Eq. (116), which becomes

(ϵ+p)αBα=Bαp,α.(\epsilon+p)\nabla_{\alpha}B^{\alpha}=-B^{\alpha}p_{,\alpha}. (120)

Introducing the chemical potential, which is just the specific enthalpy in the limit of zero temperature as considered here,

μ=dϵ/dn=(ϵ+p)/n,\mu=d\epsilon/dn=(\epsilon+p)/n, (121)

we can easily show that p,α=nμ,α=(ϵ+p)μ,α/μp_{,\alpha}=n\mu_{,\alpha}=(\epsilon+p)\mu_{,\alpha}/\mu, and as a result (120) reduces to a total divergence

α(μBα)=0.\nabla_{\alpha}(\mu B^{\alpha})=0. (122)

An equivalent form for this expression is

μ(gBi),i=gμ,iBi.\mu\left(\sqrt{-g}B^{i}\right)_{,i}=-\sqrt{-g}{\mu_{,i}}B^{i}. (123)

Combining this with (105) and the baryon conservation equation (gnui),i=0(\sqrt{-g}nu^{i})_{,i}=0 leads to to the conservation law

ddτlog[μ(ut+Ωuφ)]=0.\frac{d}{d\tau}\log\left[\mu(u_{t}+\Omega u_{\varphi})\right]=0. (124)

The corresponding flow constant DD is defined as

D=μ(ut+Ωuφ)=μuα(δtα+Ωδφα).D=\mu(u_{t}+\Omega u_{\varphi})=\mu u_{\alpha}(\delta_{t}^{\alpha}+\Omega\delta_{\varphi}^{\alpha}). (125)

The remaining conservation laws represent a generalisation of the hydrodynamical Bernoulli theorem (e.g. Gourgoulhon (2006)). If ξα\xi^{\alpha} is a Killing vector, it is easy to show that α(Tαβξβ)=0\nabla_{\alpha}(T^{\alpha\beta}\xi_{\beta})=0. For Tαβ=TEMαβ+TfluidαβT^{\alpha\beta}=T^{\alpha\beta}_{\rm EM}+T^{\alpha\beta}_{\rm fluid} this identity leads to

nuα(χuβξβ),α14π[ξβBβαBα+Bα(ξβBβ),α]\displaystyle nu^{\alpha}\left(\chi u^{\beta}\xi_{\beta}\right)_{,\alpha}-\frac{1}{4\pi}\left[\xi_{\beta}B^{\beta}\nabla_{\alpha}B^{\alpha}+B^{\alpha}\left(\xi_{\beta}B^{\beta}\right)_{,\alpha}\right]
+ξβgiβ(p+B28π),i=0,\displaystyle+\xi_{\beta}g^{i\beta}\left(p+\frac{B^{2}}{8\pi}\right)_{,i}=0, (126)

where we defined the generalised chemical potential χ=μ+B2/4πn\chi=\mu+B^{2}/4\pi n. Using (122)

nuα(χuβξβ),α+Bα4π[ξβBβμ,αμ(ξβBβ),α]\displaystyle nu^{\alpha}\left(\chi u^{\beta}\xi_{\beta}\right)_{,\alpha}+\frac{B^{\alpha}}{4\pi}\left[\xi_{\beta}B^{\beta}\frac{\mu_{,\alpha}}{\mu}-\left(\xi_{\beta}B^{\beta}\right)_{,\alpha}\right]
+ξβgiβ(p+B28π),i=0.\displaystyle+\xi_{\beta}g^{i\beta}\left(p+\frac{B^{2}}{8\pi}\right)_{,i}=0. (127)

Furthermore, using (106) for BαB^{\alpha} as well as (124) allow us to rewrite this expression as

ddτ[χuβξβ+𝒞4π(ut+Ωuφ)ξβBβ]=ξin(p+B28π),i.\frac{d}{d\tau}\left[\,\chi u^{\beta}\xi_{\beta}+\frac{{\cal C}}{4\pi}(u_{t}+\Omega u_{\varphi})\xi_{\beta}B^{\beta}\,\right]=-\frac{\xi^{i}}{n}\left(p+\frac{B^{2}}{8\pi}\right)_{,i}. (128)

Insofar the spacetime is endowed with the two standard Killing vectors, ξα={ξtα,ξφα}={δtα,δφα}\xi^{\alpha}=\{\xi^{\alpha}_{t},\xi^{\alpha}_{\varphi}\}=\{\delta_{t}^{\alpha},\delta_{\varphi}^{\alpha}\}, the right hand side term in the preceding expression vanishes and we arrive at the two conserved quantities

E\displaystyle E =(χuα+𝒞D4πμBα)ξtα=χut𝒞D4πμBt,\displaystyle=-\left(\chi u_{\alpha}+\frac{{\cal C}D}{4\pi\mu}B_{\alpha}\right)\xi_{t}^{\alpha}=-\chi u_{t}-\frac{{\cal C}D}{4\pi\mu}B_{t}, (129)
L\displaystyle L =(χuα+𝒞D4πμBα)ξφα=χuφ+𝒞D4πμBφ.\displaystyle=\left(\chi u_{\alpha}+\frac{{\cal C}D}{4\pi\mu}B_{\alpha}\right)\xi_{\varphi}^{\alpha}=\chi u_{\varphi}+\frac{{\cal C}D}{4\pi\mu}B_{\varphi}. (130)

These expressions represent MHD generalisations of the conserved specific energy and angular momentum of a non-magnetic fluid. In particular, the bracketed terms can be identified as the system’s canonical four-momentum.

A final non-trivial result follows from the combination of E,LE,L and Eq. (115). We have

EΩL+D=D/4πμ,E-\Omega L+D={{\cal F}D}/4\pi\mu, (131)

and given that μ\mu is not constant along flux surfaces, we should have D=0{\cal F}D=0. Noticing that D=0D=0 implies a vanishing poloidal field according to (105), we find that the only acceptable solution is =0{\cal F}=0. Then, D=(EΩL)D=-(E-\Omega L) and the conservation law (115) reduces to

B2=𝒞n(Bt+ΩBφ).B^{2}=-{\cal C}n(B_{t}+\Omega B_{\varphi}). (132)

It is clear now that the constants EE and DD are degenerate in the non-magnetic limit and thus one may say that DD has ‘no perfect fluid analogue’ as commented by BO78.

For the remainder of this Appendix we assume, in addition to stationarity and axisymmetry, a strictly azimuthal fluid flow i.e., uμ=(ut,0,0,uφ)u^{\mu}=(u^{t},0,0,u^{\varphi}) [as discussed elsewhere in the paper, this type of flow entails a circular metric and as a consequence uμ=(ut,0,0,uφ)u_{\mu}=(u_{t},0,0,u_{\varphi})]. From the ideal MHD equation (96) we have utFαt+uφFαφ=0u^{t}F_{\alpha t}+u^{\varphi}F_{\alpha\varphi}=0 and therefore, in the absence of meridional flow (and a non-vanishing poloidal field), the flow constant Ω\Omega can be identified with the angular frequency

Ω=uφ/ut.\Omega={u^{\varphi}}/{u^{t}}. (133)

In terms of the scalar magnetic potentials Φ=At,ψ=Aφ\Phi=A_{t},\psi=A_{\varphi} we can write Fαt=Φ,α,Fαφ=ψ,αF_{\alpha t}=\Phi_{,\alpha},F_{\alpha\varphi}=\psi_{,\alpha} and from the previous expressions we find Φ=Φ(ψ)\Phi=\Phi(\psi) and Ω=dΦ/dψ\Omega=-d\Phi/d\psi. The Ω=Ω(ψ)\Omega=\Omega(\psi) result represents the relativistic version of Ferraro’s theorem for differentially rotating MHD systems (e.g. Gourgoulhon et al. (2011)).

Finally, from the definition of BαB^{\alpha} we find the following expressions for the poloidal field components

Br=ψ,θgut,Bθ=ψ,rgut.B^{r}=\frac{\psi_{,\theta}}{\sqrt{-g}\,u^{t}},\qquad B^{\theta}=-\frac{\psi_{,r}}{\sqrt{-g}\,u^{t}}. (134)

These can be combined to give Biψ,i=Bαψ,α=0B^{i}\psi_{,i}=B^{\alpha}\psi_{,\alpha}=0. This result represents a well known geometric property, namely, that the poloidal field lines lie in ψ\psi level surfaces.

Appendix B Flow constant derivation (Newtonian)

Here we present a self-contained derivation of the flow constants in the Newtonian context; unlike their GR counterparts these are well known, for example, a concise introduction can be found in Mestel’s textbook Mestel (1999), though original derivations can be traced back to Prendergast Prendergast (1956), Chandrasekhar Chandrasekhar (1956), and Woltjer Woltjer (1959). Some important exceptions arise between the GR and Newtonian cases, which we point out here.

The ideal MHD equations for a ‘cold’ Newtonian system are

tρ+(ρ𝐯)=0,\displaystyle\partial_{t}\rho+\bm{\nabla}\cdot(\rho\mathbf{v})=0, (135)
ρ(t+𝐯)𝐯=ρΦρh+14π(×𝐁)×𝐁\displaystyle\rho(\partial_{t}+\mathbf{v}\cdot\bm{\nabla})\mathbf{v}=-\rho\bm{\nabla}\Phi-\rho\bm{\nabla}h+\frac{1}{4\pi}(\bm{\nabla}\times\mathbf{B})\times\mathbf{B}
=ρ(Φ+h)(B28π)+14π(𝐁)𝐁,\displaystyle=-\rho\bm{\nabla}\left(\Phi+h\right)-\bm{\nabla}\left(\frac{B^{2}}{8\pi}\right)+\frac{1}{4\pi}(\mathbf{B}\cdot\bm{\nabla})\mathbf{B}, (136)
𝐁=0,t𝐁=×(𝐯×𝐁),\displaystyle\bm{\nabla}\cdot\mathbf{B}=0,\qquad\partial_{t}\mathbf{B}=\bm{\nabla}\times(\mathbf{v}\times\mathbf{B}), (137)

together with an EOS and the Poisson equation 2Φ=4πGρ\nabla^{2}\Phi=4\pi G\rho for the gravitational potential. Note that we have opted for working with enthalpy instead of pressure,

dp=ρdhp=ρh.dp=\rho dh~{}\Rightarrow~{}\bm{\nabla}p=\rho\bm{\nabla}h. (138)

A general strategy for deriving conservation laws for scalar quantities is to take the scalar product of the above equations with the two available vectors, {𝐯,𝐁}\{\mathbf{v},\mathbf{B}\}, and subsequently impose axisymmetry and stationarity. In the present context a conservation law means ‘conservation along flow lines’, in other words

𝐯Q=0,\mathbf{v}\cdot\bm{\nabla}Q=0, (139)

for any scalar function QQ. Since by assumption tQ=0\partial_{t}Q=0 this condition amounts to Lie transportation, vQ=0{\cal L}_{v}Q=0.

The ideal MHD condition is encapsulated in the induction equation and, as in the GR analysis, we start our discussion from there. The assumed symmetries of the system imply

𝐯×𝐁=Φe,\mathbf{v}\times\mathbf{B}=\bm{\nabla}\Phi_{e}, (140)

from which it follows that the electrostatic potential Φe\Phi_{e} is conserved, 𝐯Φe=0\mathbf{v}\cdot\bm{\nabla}\Phi_{e}=0.

The decomposition of the vectors 𝐯,𝐁\mathbf{v},\mathbf{B} into poloidal and toroidal components,

𝐯=𝐯p+vφ𝝋^,𝐁=𝐁p+Bφ𝝋^.\mathbf{v}=\mathbf{v}_{\rm p}+v_{\varphi}\bm{\hat{\varphi}},\qquad\mathbf{B}=\mathbf{B}_{\rm p}+B_{\varphi}\bm{\hat{\varphi}}. (141)

allows us to write (140) as

𝐯p×𝐁p+Bφ(𝐯p×𝝋^)vφ(𝐁p×𝝋^)=Φe.\mathbf{v}_{\rm p}\times\mathbf{B}_{\rm p}+B_{\varphi}(\mathbf{v}_{\rm p}\times\bm{\hat{\varphi}})-v_{\varphi}(\mathbf{B}_{\rm p}\times\bm{\hat{\varphi}})=\bm{\nabla}\Phi_{e}. (142)

The projection of this equation along 𝝋^\bm{\hat{\varphi}} yields,

𝐯p=k(r,θ)𝐁p.\mathbf{v}_{\rm p}=k(r,\theta)\mathbf{B}_{\rm p}. (143)

This expression should represent the Newtonian analogue of Eq. (105), so we expect kk to be inversely proportional to ρ\rho. This is easy to show with the help of 𝐁p=0,(ρ𝐯p)=0\bm{\nabla}\cdot\mathbf{B}_{\rm p}=0,\bm{\nabla}\cdot(\rho\mathbf{v}_{\rm p})=0. From these we have 𝐁p(ρk)=0\mathbf{B}_{\rm p}\cdot\bm{\nabla}(\rho k)=0 which is equivalent to the conservation law

𝐯𝒞N=0,\mathbf{v}\cdot\bm{\nabla}{\cal C}_{\rm N}=0, (144)

with 𝒞N=1/ρk{\cal C}_{\rm N}=1/\rho k. Thus we have found

𝐁p=𝒞Nρ𝐯p,\mathbf{B}_{\rm p}={\cal C}_{\rm N}\rho\mathbf{v}_{\rm p}, (145)

which represents the Newtonian analogue of (105).

The poloidal field can be expressed in terms of the magnetic potential function ψ(r,θ)\psi(r,\theta) in the usual way,

𝐁p=ψ×φ.\mathbf{B}_{\rm p}=\bm{\nabla}\psi\times\bm{\nabla}\varphi. (146)

This parametrisation guarantees that poloidal field lines as well as the meridional flow lines lie in constant ψ\psi surfaces, 𝐁pψ=0,𝐯pψ=0\mathbf{B}_{\rm p}\cdot\bm{\nabla}\psi=0,\mathbf{v}_{\rm p}\cdot\bm{\nabla}\psi=0. We can conclude that

𝐯ψ=0,\mathbf{v}\cdot\bm{\nabla}\psi=0, (147)

which means that conserved quantities are functions Q=Q(ψ)Q=Q(\psi), see (139). Using the above expressions in (142) gives

ϖ1(vφkBφ)ψ=Φe,\displaystyle\varpi^{-1}\left(v_{\varphi}-kB_{\varphi}\right)\bm{\nabla}\psi=\bm{\nabla}\Phi_{e}, (148)

where ϖ=rsinθ\varpi=r\sin\theta. The curl of this expression leads to the conserved quantity

ϖ1(vφkBφ)=ΩN(ψ)\varpi^{-1}\left(v_{\varphi}-kB_{\varphi}\right)=\Omega_{\rm N}(\psi) (149)

This result allows us to write the Newtonian counterpart of (106)

𝐁=𝒞Nρ(𝐯ϖΩN𝝋^),\mathbf{B}={\cal C}_{\rm N}\rho\left(\mathbf{v}-\varpi\Omega_{\rm N}\bm{\hat{\varphi}}\right), (150)

and identify ΩN\Omega_{\rm N} as the Newtonian analogue of Ω\Omega (notice that in the absence of a toroidal field this constant reduces to the angular frequency parameter, vφ=ϖΩNv_{\varphi}=\varpi\Omega_{\rm N}).

The remaining ‘hydrodynamical’ constants DND_{\rm N}, ENE_{\rm N}, LNL_{\rm N} should emerge from projections of the Euler equation. The azimuthal component of that equation gives

𝐯p(ϖvφ)=14πρ𝐁p(ϖBφ),\mathbf{v}_{\rm p}\cdot\bm{\nabla}(\varpi v_{\varphi})=\frac{1}{4\pi\rho}\mathbf{B}_{\rm p}\cdot\bm{\nabla}(\varpi B_{\varphi}), (151)

Using (150) we can write this expression as a conservation law, 𝐯LN=0\mathbf{v}\cdot\bm{\nabla}L_{\rm N}=0, where

LN=ϖ(vφ𝒞N4πBφ).L_{\rm N}=\varpi\left(v_{\varphi}-\frac{{\cal C}_{\rm N}}{4\pi}B_{\varphi}\right). (152)

The last two constants ENE_{\rm N} and DND_{\rm N} are found by dotting the Euler equation with 𝐯\mathbf{v} and 𝐁\mathbf{B}. We find,

12ρtv2+ρ𝐯[(𝐯)𝐯]+ρ𝐯(Φ+h)\displaystyle\frac{1}{2}\rho\partial_{t}v^{2}+\rho\mathbf{v}\cdot[(\mathbf{v}\cdot\bm{\nabla})\mathbf{v}]+\rho\mathbf{v}\cdot\bm{\nabla}\left(\Phi+h\right)
+𝐯(B28π)14π𝐯[(𝐁)𝐁]=0,\displaystyle+\mathbf{v}\cdot\bm{\nabla}\left(\frac{B^{2}}{8\pi}\right)-\frac{1}{4\pi}\mathbf{v}\cdot[(\mathbf{B}\cdot\bm{\nabla})\mathbf{B}]=0, (153)
ρ𝐁t𝐯+ρ𝐁[(𝐯)𝐯]+ρ𝐁(Φ+h)=0.\rho\mathbf{B}\cdot\partial_{t}\mathbf{v}+\rho\mathbf{B}\cdot[(\mathbf{v}\cdot\bm{\nabla})\mathbf{v}]+\rho\mathbf{B}\cdot\bm{\nabla}\left(\Phi+h\right)=0. (154)

We can manipulate some of the terms appearing in these expressions, viz.

ρ𝐯[(𝐯)𝐯]\displaystyle\rho\mathbf{v}\cdot[(\mathbf{v}\cdot\bm{\nabla})\mathbf{v}] =(12ρv2𝐯)+12v2tρ,\displaystyle=\bm{\nabla}\cdot\left(\frac{1}{2}\rho v^{2}\mathbf{v}\right)+\frac{1}{2}v^{2}\partial_{t}\rho,
ρ𝐯(Φ+h)\displaystyle\rho\mathbf{v}\cdot\bm{\nabla}\left(\Phi+h\right) =[ρ𝐯(Φ+h)]+(Φ+h)tρ.\displaystyle=\bm{\nabla}\cdot\left[\rho\mathbf{v}\left(\Phi+h\right)\right]+\left(\Phi+h\right)\partial_{t}\rho. (155)

For our stationary system, Eq. (B) becomes

0=\displaystyle 0= [ρ𝐯(12v2+Φ+h)]\displaystyle\bm{\nabla}\cdot\left[\rho\mathbf{v}\left(\frac{1}{2}v^{2}+\Phi+h\right)\right] (156)
+𝐯(B28π)14π𝐯[(𝐁)𝐁].\displaystyle+\mathbf{v}\cdot\bm{\nabla}\left(\frac{B^{2}}{8\pi}\right)-\frac{1}{4\pi}\mathbf{v}\cdot[(\mathbf{B}\cdot\bm{\nabla})\mathbf{B}]. (157)

In the non-magnetic case this equation would have led to the familiar Bernoulli flow-line constant =v2/2+Φ+h{\cal B}=v^{2}/2+\Phi+h Chandrasekhar (1956). The absence of ρ\rho in the magnetic terms in (157) does not allow the same kind of manipulation (unless ρ\rho is constant). However, the fact that 𝐁p=0\bm{\nabla}\cdot\mathbf{B}_{\rm p}=0 and 𝐯p=𝐁p/𝒞Nρ\mathbf{v}_{\rm p}=\mathbf{B}_{\rm p}/{\cal C}_{\rm N}\rho may allow for a 𝐁pQ=0\mathbf{B}_{\rm p}\cdot\bm{\nabla}Q=0 conservation law. Using index notation (i,j={r,θ,φ})(i,j=\{r,\theta,\varphi\}) for clarity, the magnetic tension term can be written as

viBjjBi=Bjj(viBi)BiBjjvi.v_{i}B^{j}\nabla_{j}B^{i}=B^{j}\nabla_{j}(v_{i}B^{i})-B^{i}B^{j}\nabla_{j}v_{i}. (158)

At the same time, from the induction equation we have

Bjjvij(vjBi)=0,B^{j}\nabla_{j}v^{i}-\nabla_{j}(v^{j}B^{i})=0, (159)

and the magnetic tension term becomes

viBjjBi=[(𝐯×𝐁)×𝐁]+12𝐯B2.v_{i}B^{j}\nabla_{j}B^{i}=\bm{\nabla}\cdot[(\mathbf{v}\times\mathbf{B})\times\mathbf{B}]+\frac{1}{2}\mathbf{v}\cdot\bm{\nabla}B^{2}. (160)

The resulting Euler equation is

0=[ρ𝐯p(12v2+Φ+h)14π(𝐯×𝐁)×𝐁],0=\bm{\nabla}\cdot\left[\rho\mathbf{v}_{\rm p}\left(\frac{1}{2}v^{2}+\Phi+h\right)-\frac{1}{4\pi}(\mathbf{v}\times\mathbf{B})\times\mathbf{B}\right], (161)

where we have exploited axisymmetry to set 𝐯𝐯p\mathbf{v}\to\mathbf{v}_{\rm p} in the first term. The last term can be written as

(𝐯×𝐁)×𝐁=ϖΩN(Bp2𝝋^𝐁pBφ).(\mathbf{v}\times\mathbf{B})\times\mathbf{B}=-\varpi\Omega_{\rm N}(B^{2}_{\rm p}\bm{\hat{\varphi}}-\mathbf{B}_{\rm p}B_{\varphi}). (162)

In axisymmetry the first right-hand-side term is divergence free and the Euler equation reduces to

𝐯[(12v2+Φ+h)14πϖΩN𝒞NBφ]=0.\mathbf{v}\cdot\bm{\nabla}\left[\left(\frac{1}{2}v^{2}+\Phi+h\right)-\frac{1}{4\pi}\varpi\Omega_{\rm N}{\cal C}_{\rm N}B_{\varphi}\right]=0. (163)

We have thus arrived to the MHD generalisation of the Bernoulli constant

EN=12v2+Φ+hΩN𝒞N4πϖBφ.E_{\rm N}=\frac{1}{2}v^{2}+\Phi+h-\frac{\Omega_{\rm N}{\cal C}_{\rm N}}{4\pi}\varpi B_{\varphi}. (164)

Unlike the analogous GR result (129), (164) does not depend on the chemical potential.

Our last order of business for this section is the manipulation of Eq. (154). Removing the time-derivative

𝐁[(𝐯)𝐯]+𝐁(Φ+h)=0.\mathbf{B}\cdot\left[(\mathbf{v}\cdot\bm{\nabla})\mathbf{v}\right]+\mathbf{B}\cdot\bm{\nabla}\left(\Phi+h\right)=0. (165)

The first term can be manipulated with the help of the induction equation to give

Bivjjvi=j(vjviBi)12Bjjv2.B_{i}v^{j}\nabla_{j}v^{i}=\nabla_{j}(v^{j}v^{i}B_{i})-\frac{1}{2}B^{j}\nabla_{j}v^{2}. (166)

The axisymmetric Euler equation is equivalent to

𝐯[(𝐁𝐯)𝒞Nρ+Φ+h12v2]=0,\mathbf{v}\cdot\bm{\nabla}\left[\frac{(\mathbf{B}\cdot\mathbf{v})}{{\cal C}_{\rm N}\rho}+\Phi+h-\frac{1}{2}v^{2}\right]=0, (167)

and the corresponding flow constant is

DN=𝐁𝐯𝒞Nρ+Φ+h12v2.D_{\rm N}=\frac{\mathbf{B}\cdot\mathbf{v}}{{\cal C}_{\rm N}\rho}+\Phi+h-\frac{1}{2}v^{2}. (168)

With the help of (150) we can rewrite this as

DN=12v2+Φ+hϖΩNvφ.D_{\rm N}=\frac{1}{2}v^{2}+\Phi+h-\varpi\Omega_{\rm N}v_{\varphi}. (169)

As in GR, it can be noticed that the constants ENE_{\rm N} and DND_{\rm N} are degenerate in the non-magnetic limit (both being equal to the Bernoulli constant {\cal B} in different frames).

We remark that we have been unable to obtain an analogous \mathcal{F} flow constant in the Newtonian case, though the relation DN=ENΩNLND_{\rm N}=E_{\rm N}-\Omega_{\rm N}L_{\rm N} essentially corresponds to setting =0\mathcal{F}=0 in GRMHD, see Eq. (131).

The ‘no toroidal field theorem’ discussed in Sec. III.1 persists in the present Newtonian context. Indeed, considering the limit of vanishing meridional flow, 𝐯p0\mathbf{v}_{\rm p}\to 0, we find that (145) predicts 𝒞N{\cal C}_{\rm N}\to\infty if we require a poloidal magnetic field to be present. However, we can see that the divergence of 𝒞N{\cal C}_{\rm N} makes both the hydrodynamic constants ENE_{\rm N} and LNL_{\rm N} divergent unless Bφ0B_{\varphi}\to 0 since there are no other terms available for counterbalancing. As in GR, we are thus forced to conclude that the field must be purely poloidal.

References