Issue |
A&A
Volume 550, February 2013
|
|
---|---|---|
Article Number | A43 | |
Number of page(s) | 16 | |
Section | Astrophysical processes | |
DOI | https://doi.org/10.1051/0004-6361/201220082 | |
Published online | 23 January 2013 |
Equation of state for magnetized Coulomb plasmas⋆
1
CRAL (UMR CNRS 5574), École Normale Supérieure de Lyon,
69364
Lyon Cedex 07,
France
e-mail: palex@astro.ioffe.ru, chabrier@ens-lyon.fr
2
Ioffe Physical-Technical Institute, Politekhnicheskaya 26, 194021
St. Petersburg,
Russia
3
Isaac Newton Institute of Chile, St. Petersburg Branch,
Russia
4
School of Physics, University of Exeter,
Exeter,
EX4 4QL,
UK
Received:
23
July
2012
Accepted:
5
December
2012
We have developed an analytical equation of state (EOS) for magnetized fully-ionized plasmas that cover a wide range of temperatures and densities, from low-density classical plasmas to relativistic, quantum plasma conditions. This EOS directly applies to calculations of structure and evolution of strongly magnetized white dwarfs and neutron stars. We review available analytical and numerical results for thermodynamic functions of the nonmagnetized and magnetized Coulomb gases, liquids, and solids. We propose a new analytical expression for the free energy of solid Coulomb mixtures. Based on recent numerical results, we have constructed analytical approximations for the thermodynamic functions of harmonic Coulomb crystals in quantizing magnetic fields. The analytical description ensures a consistent evaluation of all astrophysically important thermodynamic functions based on the first, second, and mixed derivatives of the free energy. Our numerical code for calculation of thermodynamic functions based on these approximations has been made publicly available. Using this code, we calculate and discuss the effects of electron screening and magnetic quantization on the position of the melting point in a range of densities and magnetic fields relevant to white dwarfs and outer envelopes of neutron stars. We consider also the thermal and mechanical structure of a magnetar envelope and argue that it can have a frozen surface which covers the liquid ocean above the solid crust.
Key words: dense matter / equation of state / magnetic fields / stars: neutron / white dwarfs
The Fortran code that realizes the analytical approximations described in this paper is available at http://www.ioffe.ru/astro/EIP/ and at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsarc.u-strasbg.fr/viz-bin/qcat?J/A+A/550/A43
© ESO, 2013
1. Introduction
Coulomb plasmas, i.e., the fully-ionized plasmas whose thermodynamics is strongly affected by electrostatic interactions, are encountered in many physical and astrophysical situations (e.g., Fortov 2009). Full ionization is reached either at high temperatures T and low densities ρ (thermal ionization) or at high densities ρ (pressure ionization). The latter case is typical of the interior conditions of low-mass stars, brown dwarfs, or giant planets (Chabrier & Baraffe 2000) as well as the interior and envelope conditions of white dwarfs and neutron stars. Coulomb interactions are crucial for the equation of state (EOS) under such conditions. In the interior or the envelope of compact objects such as white dwarfs and neutron stars, the electrons can be weakly or strongly degenerate, the plasma can be in the liquid or solid state, the electrons can have various degrees of degeneracy and relativism, and the quantum effects on ion motion can be substantial. Therefore, a wide-range EOS is needed for calculations of the structure and evolution of such stars.
In a previous work (Potekhin & Chabrier 2000, 2010, hereafter Papers I and II, respectively) we proposed a set of analytical expressions for the calculations of the EOSs of the Coulomb plasmas without magnetic fields and presented a code for thermodynamic functions based on the first, second, and mixed derivatives of the analytical Helmholtz free energy F with respect to density ρ and temperature T. This code has been employed in astrophysical modeling and adapted for the use in the Modules for Experiments in Stellar Astrophysics (mesa; Paxton et al. 2011).
The Bohr – van Leeuwen theorem states that an EOS of charged pointlike classical particles is not affected by a magnetic field (van Leeuwen 1921). However, a magnetic field can affect thermodynamic functions through intrinsic magnetic moments of particles and by the quantization of the motion of charged particles in Landau orbitals (Landau 1930; Landau & Lifshitz 1977). These effects can be important, for example, in magnetic white dwarfs whose magnetic fields B can reach 107−109 G (e.g., Wickramasinghe & Ferrario 2000, and references therein) and neutron stars with typical B ~ 108−1014 G (e.g., Haensel et al. 2007, and references therein).
In this paper, we systematically consider analytical expressions for thermodynamic functions of magnetized Coulomb plasmas, discuss their validity range, and introduce some practical modifications. We take account of analytical and numerical results, currently available for various contributions to the Helmholtz free energy in quantizing magnetic fields. Taking advantage of recently published numerical results (Baiko 2009), we construct an analytical description of the thermodynamic functions of harmonic Coulomb crystals in quantizing magnetic fields.
In Sect. 2 we give definitions and simple estimates for the plasma parameters that determine different thermodynamic regimes. In Sect. 3 we outline the EOS of a nonmagnetized Coulomb plasma as the reference case. In Sect. 4 we consider the Boltzmann and Fermi gases in quantizing magnetic fields, present a general analytical description of their EOS, and simplify them for several limiting cases. In Sect. 5 we review nonideal contributions to the EOS of a Coulomb liquid in a strong magnetic field. In Sect. 6 we derive an analytical approximation for the EOS of a strongly magnetized Coulomb crystal. In Sect. 7 we present and discuss examples of thermodynamic functions for conditions typical of white dwarfs and neutron-star envelopes. The summary is given in Sect. 8. In the Appendices we give the explicit expressions for the thermodynamic functions used in Sect. 3.
2. Basic definitions
2.1. General parameters
Let ne and ni be the electron and ion number densities, A and Z the ion mass and charge numbers, respectively. In this paper we consider the neutral plasmas (therefore ne = Zni) that contain a single type of ion and include neither positrons (they can be described using the same analytical functions as for the electrons; see, e.g., Blinnikov et al. 1996; Timmes & Arnett 1999), nor free neutrons (see Haensel et al. 2007 for a review).
The state of a free-electron gas is determined by the electron number density
ne and temperature T. Instead of
ne it is convenient to introduce the dimensionless density
parameter
rs = ae/a0,
where a0 is the Bohr radius and
. The
parameter rs can be quickly evaluated from the relations
where
n24 ≡ ne/1024 cm-3
and
ρ0 = 2.6752 (A/Z)
g cm-3. The analogous density parameter for the ions is
RS = aimi(Ze)2/ħ2 = 1822.89AZ7/3 rs,
where mi is the ion mass and
is the
ion sphere radius.
At stellar densities it is convenient to use, instead of rs,
the nonmagnetic relativity parameter (1)where
pF = ħ (3π2ne)1/3
is the electron Fermi momentum in the absence of a magnetic field, and
ρ6 ≡ ρ/106
g cm-3. The Fermi energy (without the rest energy) for the electron gas is
and the Fermi temperature
TF ≡ ϵF/kB = Tr (γr − 1),
where
Tr ≡ mec2/kB = 5.93 × 109 K,
,
and kB is the Boltzmann constant. A useful measure of electron
degeneracy is
θ = T/TF.
In the nonrelativistic limit (xr ≪ 1),
K,
and
(2)where
(3)In the opposite
ultrarelativistic limit (xr ≫ 1),
θ ≈ (263 Γe)-1. The strength of the Coulomb
interaction of nonrelativistic ions is characterized by the Coulomb coupling parameter
(4)where
T6 ≡ T/106 K.
Thermal de Broglie wavelengths of free ions and electrons are usually defined as
(5)although in some
publications these definitions differ by a numerical factor. The quantum effects on ion
motion are important either at
λi ≳ ai or at
T ≪ Tp, where
Tp ≡ ħωp/kB
is the ion plasma temperature and
ωp = (4πe2 niZ2/mi(1/2
is the ion plasma frequency. Since
,
the importance of the quantum effects in strongly coupled plasmas (i.e., at Γ ≫ 1) is
determined by parameter
(6)
2.2. Magnetic-field parameters
In the nonrelativistic theory (Landau & Lifshitz
1977), the energy of an electron in magnetic field
B equals , where
pz is the momentum component along
B,
ωc = eB/mec
is the electron cyclotron frequency,
characterizes a Landau level, σ = ±1 determines the spin projection on
the field, and nL is the non-negative integer Landau number
related to the quantization of the kinetic motion transverse to the field. In the
relativistic theory (Johnson & Lippmann
1949; Berestetskiĭ et al. 1982), the
kinetic energy ϵ of an electron at the Landau level n
and its longitudinal momentum pz are
inter-related as
The
levels
are double-degenerate with respect to σ. Their splitting due to the
anomalous magnetic moment of the electron is
≈ (mec2αf/2π) b
at b ≪ 1 and
~(mec2αf/2π) [lnb − 1.584] 2
at b ≫ 1 (see Schwinger 1988;
Suh & Mathews 2001), which is always much
smaller than ħωc and is negligible in the compact stars.
Convenient dimensionless parameters that characterize the magnetic field in a plasma are
the ratios of the electron cyclotron energy ħωc to the Hartree
unit of energy, to the electron rest energy, and to
kBT:
(9)where
B0 = 2.3505 × 109 G,
(10)where
αf = e2/ħc
is the fine-structure constant, and
(11)where
B12 ≡ B/1012 G.
The magnetic length
gives a characteristic transverse scale of the electron wave function.
For the ions, the cyclotron energy is
ħωci = Z (me/mi) ħωc,
and the parameter analogous to ζ is (12)Another important
parameter is the ratio of the ion cyclotron frequency to the plasma frequency,
(13)
2.3. Free energy and thermodynamic functions
The Helmholtz free energy F of a plasma can be conveniently written as
(14)where
and
denote the ideal free energy of the ions and the electrons, and the last three terms
represent an excess free energy arising from the electron-electron, ion-ion, and
ion-electron interactions, respectively. In the nonideal plasmas, correlations between any
plasma particles depend on all interactions, therefore the separation in Eq. (14) is just a question of convenience.
An important reference case is the model of one-component plasma (OCP). In this model,
the electrons are replaced by a rigid (nonpolarizable) background of the uniform charge
distribution. It is convenient to define Fii as the difference
between F and
in the OCP model. Still stronger simplification is the ion-sphere model, in which the
interaction energy in the OCP is evaluated as the electrostatic energy of a positive ion
in the negatively charged sphere of radius ai (Salpeter 1961). The electron exchange-correlation term
is defined as
in the model of an electron gas without consideration of the ions, which are replaced by
an uniform positive background to ensure the global charge neutrality. The ion-electron
(electron polarization) contribution Fie, then, is the
difference between F and the other terms, when interactions between all
types of particles are taken into account.
The pressure P, the internal energy U, and the entropy S of an ensemble of particles in volume V can be found from the thermodynamic relations P = −(∂F/∂V)T, S = −(∂F/∂T)V, and U = F + TS. The second-order thermodynamic functions are derived by differentiating these first-order functions. The decomposition (Eq. (14)) induces analogous decompositions of P, U, S, the heat capacity CV = (∂S/∂lnT)V, and the logarithmic derivatives χT = (∂lnP/∂lnT)V and χρ = −(∂lnP/∂lnV)T. Other second-order functions can be expressed through these functions by Maxwell relations (e.g., Landau & Lifshitz 1980).
3. EOS of nonmagnetized Coulomb plasmas
3.1. Ideal part of the free energy
The free energy of a gas of
Ni = niV
nonrelativistic classical ions is (15)where
Mspin is the spin multiplicity. Accordingly,
,
,
, and
. In the OCP, Eq. (15) can be written in terms of the
dimensionless plasma parameters (Sect. 2) as
(16)The free energy of the
electron gas is given by
(17)where
Ne = neV is the
number of electrons and μe is the electron chemical potential
without the rest energy mec2. The
pressure and the number density are functions of μe and
T:
where
χe ≡ μe/kBT,
τ ≡ T/Tr,
and
(20)is the Fermi-Dirac
integral. The internal energy is
(21)Since we use
V and T as independent variables, we need to find
μe(V,T). This can be done either by
inverting Eq. (19) numerically, or from
the analytical approximation given in Chabrier &
Potekhin (1998). Then the second-order thermodynamic functions are obtained using
relations of the type
We
use analytical approximations for
Iν(χe,τ)
based on the fits of Blinnikov et al. (1996) and
accurate typically to a few parts in 104, with maximum error
<0.2% at τ ≤ 100 (Chabrier & Potekhin 1998). These approximations are given by
different expressions in three ranges of χe: below, within,
and above the interval 0.6 ≤ χe < 14.
In particular, at large χe the Sommerfeld expansion (e.g.,
Chandrasekhar 1957; Girifalco 1973) yields1
(24)where
is the electron
chemical potential (without the rest energy) in the relativistic units,
Here,
we have introduced notations
and
. At
strong degeneracy,
,
, and
. In Paper II we also
described an alternative expansion in powers of τ, which allows one to
avoid numerical cancellations of close terms at small
(we switch to this alternative expansion at
).
The discontinuities of the Blinnikov et al. (1996) approximations for Iν(χe,τ) at χe = 0.6 and χe = 14 are typically a few parts in 104 at τ ≲ 102, but they may reach ≈ 1% for the second derivatives. This accuracy is sufficient for many applications. Nevertheless, the jumps may produce problems, e.g., when higher derivatives are evaluated numerically in a stellar evolution code. In our calculations of white-dwarf evolution (to be published elsewhere), we smoothly interpolate between the two analytical approximations for the adjacent intervals near the boundary at the cost of a slight violation of the thermodynamic consistency in the interpolation regions (this version of the EOS code is now also available at our web site).
If a higher accuracy is needed, one can numerically calculate tables of Iν(χe,τ) (e.g., Timmes & Arnett 1999) and interpolate in them with an algorithm that preserves thermodynamic consistency (Timmes & Swesty 2000) and is available at mesa (Paxton et al. 2011).
3.2. Nonideal contributions
3.2.1. Electron and ion liquids
The contribution to the free energy due to the electron-electron interactions has been studied by many authors. For the reasons explained in Paper II, we adopt the fit to Fee derived by Ichimaru et al. (1987A.1) (see Appendix A).
The ion-ion interactions are described using the OCP model. In the liquid regime, the numerical results obtained for the OCP of nonrelativistic pointlike charged particles in different intervals of the Coulomb coupling parameter from Γ = 0 to Γ ~ 200 by different numerical and analytical methods are reproduced by a simple expression given in Appendix B.1. The accurate fit for classical OCP is supplemented by the Wigner-Kirkwood correction, which extends the applicability range of our approximations to lower temperatures T ~ Tp. In spite of the significant progress in numerical ab initio modeling of quantum ion liquids, available results do not currently allow us to establish an analytical extension to still lower temperatures T ≪ Tp (see Chabrier et al. 2002, for references and discussion).
3.2.2. Coulomb crystal
At T < Tm, where
Tm is the melting temperature, the ions in thermodynamic
equilibrium are arranged in the body-centered cubic (bcc) Coulomb lattice. In the
harmonic approximation (e.g., Kittel 1963), the
free energy of the lattice is (29)where
U0 = NiC0(Ze)2/ai
is the classical static-lattice energy, C0 ≈ −0.9 is the
Madelung constant,
(30)accounts
for zero-point quantum vibrations,
u1 = ⟨ ωkα ⟩ ph/ωp ≈ 0.5
is the reduced first moment of phonon frequencies,
(31)is
the thermal contribution,
ωkα are
phonon frequencies, and ⟨ ... ⟩ ph denotes the averaging
over phonon polarizations α and wave vectors
k in the first Brillouin zone. Here we do not separate
the classical-gas free energy, therefore Flat replaces
in Eq. (14).
Beyond the harmonic-lattice approximation, the total reduced free energy
flat ≡ Flat/NikBT
can be written as (32)Here, the first
three terms correspond to the three terms in Eq. (29), and fah is the anharmonic
correction. The most accurate values of the constants C0 and
u1 were calculated by Baiko
(2000) (see Appendix B.2). For
fth = Fth/NikBT,
we use the highly precise fit of Baiko et al.
(2001) (Appendix B.2). In the classical
limit η ≪ 1, it reduces to
fth ≃ 3lnη − 2.49389−1.5u1η + η2/24,
where the term with u1 cancels that in Eq. (32) and the last term represents the
Wigner-Kirkwood quantum correction
(Eq. (B.3)), which is the same in the
liquid and solid phases (Pollock & Hansen
1973). In the opposite limit
T ≪ Tp, we have
fth ≃ −209.3323 η-3 (here
the constant is given for the bcc crystal; for other lattice types, see Baiko et al. 2001).
Anharmonic corrections for Coulomb lattices were studied by many authors in the limits η → 0 and η → ∞, but only a few numerical results of low precision are available at finite η values (see Paper II for references and discussion). In Paper II we constructed an analytical interpolation between these limits, which is applicable at arbitrary η and is consistent with the available numerical results within accuracy of the latter ones (Appendix B.2). It should be replaced by a more accurate function in the future when accurate finite-temperature anharmonic quantum corrections become available.
3.2.3. Electron polarization
Electron polarization in Coulomb liquids was studied by perturbation (Galam & Hansen 1976; Yakovlev & Shalybkov 1989) and hypernetted-chain (HNC) techniques (Chabrier & Ashcroft 1990; Chabrier & Potekhin 1998; Paper I). The results have been reproduced by an analytical expression (Appendix C.1), which exactly recovers the Debye-Hückel limit for the weakly coupled (Γ ≪ 1) electron-ion plasmas and the Thomas-Fermi limit for the strongly coupled (Γ ≫ 1) Coulomb liquids at Z ≫ 1.
For classical ions, the simplest screening model consists in replacing the Coulomb potential by the Yukawa potential. Molecular-dynamics and path-integral Monte Carlo simulations of classical liquid and solid Yukawa systems were performed in several works (e.g., Hamaguchi et al. 1997; Militzer & Graham 2006). However, the Yukawa interaction reflects only the small-wavenumber asymptote of the electron dielectric function (Jancovici 1962; Galam & Hansen 1976). The first-order perturbation approximation for the dynamical matrix of a classical Coulomb solid with the polarization corrections was developed by Pollock & Hansen (1973). The phonon spectrum in such a quantum crystal has been calculated only in the harmonic approximation (Baiko 2002), which has a restricted applicability to this problem (for example, it is obviously incapable of reproducing the polarization contribution to the heat capacity in the classical limit η → 0, where it gives CV = 3NikB independent of the polarization).
In Paper I we calculated Fie using the semiclassical perturbation theory of Galam & Hansen (1976) with a model structure factor, and fit the results by an analytical function of xr and η. In Paper II we improved the η-dependence of this function to completely eliminate the screening contribution in the strong quantum limit η ≪ 1, because the employed model of the structure factor failed at η ≲ 1. The latter approximation is reproduced in Appendix C.2. It can be improved in the future, when the polarization corrections for the quantum Coulomb crystal at η ≲ 1 have been accurately evaluated.
3.2.4. Ion mixtures
In Sects. 3.2.1–3.2.3 we have considered plasmas containing identical ions. In the case where
several ion types are present in a strongly coupled Coulomb plasma, a common
approximation is the linear mixing rule (LMR), (33)where
xj are the number fractions of ions with
charge numbers Zj and
.
In Eq. (33),
fex ≡ Fex/NikBT
is the reduced nonideal part of the free energy, Fex is the
excess free energy, which is equal to Fii in the case of the
rigid charge-neutralizing electron background and to
Fii + Fie + Fee
in the case of the polarizable background. The high accuracy of Eq. (33) for binary ionic mixtures in the rigid
background was first demonstrated by calculations in the HNC approximation (Hansen et al. 1977; Brami et al. 1979) and confirmed later by Monte Carlo simulations (DeWitt et al. 1996; Rosenfeld 1996; DeWitt & Slattery
2003). The validity of the LMR in the case of an ionic mixture immersed in a
polarizable finite-temperature electron background has been examined by Hansen et al. (1977) in the first-order thermodynamic
perturbation approximation and by Chabrier &
Ashcroft (1990) by solving the HNC equations with effective screened
potentials. These authors found that the LMR remains accurate when the electron response
is taken into account in the inter-ionic potential, as long as the Coulomb coupling is
strong (Γj > 1,
∀j).
However, the LMR is not exact, and Eq. (33) should be replaced by the Debye – Hückel formula in the limit of weak coupling (Γj ≪ 1, ∀j). Even in the strong-coupling regime, the small deviations from the LMR are important for establishing phase equilibria (see Medin & Cumming 2010). The deviations from the LMR were studied by Brami et al. (1979); Chabrier & Ashcroft (1990); DeWitt et al. (1996); DeWitt & Slattery (2003); Potekhin et al. (2009a,b) for strongly coupled Coulomb liquids and by Ogata et al. (1993) and DeWitt & Slattery (2003) for Coulomb solids.
The analytical expression that describes deviations from the LMR,
Δf ≡ f − fLM, in Coulomb
liquids for arbitrary coupling parameters Γj reads (Potekhin et al. 2009b)
(34)where
,
⟨ Γ ⟩ = Γe ⟨ Z5/3 ⟩ ,
δ is defined either as
(35)for
rigid electron background model, or as
(36)for
polarizable background, and parameters a, b,
α, and β depend on the plasma composition as
follows:
For
Coulomb solids, one should distinguish regular crystals containing different ion types
and disordered solid mixtures, where different ions are randomly distributed in regular
lattice sites (Ogata et al. 1993). Each regular
lattice type corresponds to a fixed composition, whereas random lattices allow variable
fractions of different ion types. The free energy correction Δf mainly
arises from the difference in the Madelung energies. It is generally larger for regular
crystals than for “random” crystals with the same composition. Ogata et al. (1993) performed Monte Carlo simulations of solid ionic
mixtures and fitted the calculated deviation, Δfsol, from
linear-mixing prediction for the reduced free energy in a random binary ion crystal.
Medin & Cumming (2010) and Hughto et al. (2012) used this fit to study the phase
separation and solidification of ion mixtures in the interiors of white dwarfs. We note,
however, that the fit of Ogata et al. (1993)
exhibits nonphysical features: for example, it is nonmonotonic as a function of
RZ = Z2/Z1
at a fixed number fraction
x2 = 1 − x1 for a binary ion
mixture with
Z2/Z1 > 2
and x2 < 0.5. A much simpler fit,
which does not exhibit unphysical behavior, was suggested by DeWitt & Slattery (2003). It can be written as
However, the latter fit is
valid only for relatively small charge ratios
RZ ≲ 3/2. We replace
it by the expression
(39)where
(40)and
The approximation in
Eq. (40) reproduces reasonably well
the results of both Ogata et al. (1993) and DeWitt & Slattery (2003) for random
two-component ionic bcc lattices. For a multicomponent ion crystal, Medin & Cumming (2010) proposed the
extrapolation from the two-component plasma case
(41)where the indices are
arranged so that
Zj < Zj + 1.
4. EOS of a fully ionized magnetized gas
4.1. Ions
We consider only nondegenerate and nonrelativistic ions (for a discussion of the EOS of
degenerate nuclear matter in strong magnetic fields see, e.g., Broderick et al. 2000; Suh &
Mathews 2001). In this case (cf. Potekhin et al.
1999) (42)The last term arises from
the energy of the magnetic moments of the ions in the magnetic field,
(43)where
Mspin is the ion spin multiplicity, and
gi is the g-factor
(gi = 5.5857 for protons). For ions with spin one-half
(Mspin = 2), the expression in the square brackets in
Eq. (43) simplifies to
[2cosh(gi ζi/4)] .
For zero-spin ions, such as 4He, 56Fe, and other even-even nuclei in
the ground state, Fspin = 0.
The ion pressure obeys the nonmagnetic ideal-gas relation
, but expressions for the
internal energy and heat capacity are different:
Here,
the terms uspin and cspin arise
from Fspin,
They
simplify at Mspin = 2:
(48)
4.2. Electrons
4.2.1. General case
Thermodynamic functions of the electron gas in a magnetic field are easily derived from
first principles (Landau & Lifshitz
1980). The number of quantum states per longitudinal momentum interval
Δpz for an electron with given
B-projections of the spin and the orbital moment and
with a fixed Landau number n in a volume V equals
(Landau & Lifshitz 1977). Thus one can
express the electron number density ne and the grand
thermodynamic potential
as
where
ϵn(pz)
is given by Eq. (7) and
∑ σ denotes the sum over spin projections, which
amounts to the factor 2 for n ≥ 1 since we neglect the anomalous
magnetic moment of electrons. This derivation equally holds in the relativistic and
nonrelativistic theories. Equations (49)
and (50) can be rewritten, using
integration by parts, as
where
and
The
free energy
is
given by Eqs. (17), (51), and (52).
The calculation of ne,
, and
their derivatives at given χe and τ can be
performed using Eqs. (51, (52) and the same analytical approximations
to the Fermi-Dirac integrals as for the nonmagnetized electron gas. The reduced electron
chemical potential χe at constant
ne and T is found by numerical inversion
of Eq. (51). Then the derivatives over
T at constant V and over V at
constant T are given by Eqs. (22) and (23). We use this
approach in the current research, but we should note that for quantizing magnetic fields
it is less precise than at B = 0. As mentioned in Sect. 3.1, the inaccuracy of the employed approximations for
Iν(χe,τ)
is within a fraction of percent, but it grows for the derivatives. Since the first
derivatives are already employed in Eq. (51), evaluation of the second-order thermodynamic functions such as
χT or
CV involves third derivatives.
Therefore, the error in the evaluation of these functions may rise to several percent.
This level of accuracy may be sufficient for astrophysical applications, but otherwise
one should resort to a thermodynamically consistent interpolation in numerical tables of
the Fermi-Dirac integrals (Timmes & Arnett
1999; Timmes & Swesty 2000).
Equations (51) and (52) can be simplified in several limiting
cases considered below.
4.2.2. Strongly quantizing and nonquantizing magnetic fields
The field is called strongly quantizing if most of the electrons reside on the ground
Landau level. The electron Fermi momentum, then, equals (53)where
is
the zero-field Fermi momentum at the given density. Equation (53) can be written as
pF = mecxB,
where
(54)is
the relativity parameter modified by the field and
(Sect. 2.1). With increasing
ne at constant B and zero temperature,
the electrons start to populate the first excited Landau level when
ne reaches
.
Therefore, the field is strongly quantizing at
T ≪ Tcycl and
ρ < ρB,
where
Tcycl = ħωc/kB ≈ 1.3434 × 108B12 K
and
(55)The condition
ne < nB
can be written as
. Then
Eq. (53) shows that in a strongly
quantizing field
,
except for densities ne close to the threshold
nB. Thus TF
is reduced, compared to its nonmagnetic value
, by
factor
. In the nonrelativistic
limit,
, and the
parameter
θ = T/TF
becomes
(56)where
θ0 is the nonmagnetic value given by Eq. (2).
The opposite case of a nonquantizing magnetic field occurs at T ≫ TB, where TB is the temperature at which the thermal kinetic energy of the electrons becomes sufficient to smear their distribution over many Landau levels. It can be estimated as TB = Tcycl, if ρ ≤ ρB and TB = Tcycl/γr, if ρ > ρB (a more sophisticated but qualitatively similar definition of TB was introduced by Lai 2001). Then we can approximately replace the sum over Landau level numbers n by the integral over a continuous variable n. Integrating over n by parts, we can reduce Eq. (52) to Eq. (18) and Eq. (51) to Eq. (19), i.e., to recover the zero-field thermodynamics. At ρ ≫ ρB, the electrons also fill many Landau levels and the magnetic field can be approximately treated as nonquantizing.
In the intermediate region, where the magnetic field is neither strongly quantizing nor nonquantizing, the summation over n manifests itself in quantum oscillations of the thermodynamic functions with changing B and/or ρ, similar to the de Haas – van Alphen oscillations of magnetic susceptibility (e.g., Landau & Lifshitz 1980). The oscillations are smoothed by the thermal broadening of the Fermi distribution function and by the quantum broadening of the Landau levels (particularly, owing to electron collisions; see Yakovlev & Kaminker 1994, for references). Some examples of such oscillations will be given in Sect. 7.
Figure 1 presents the ρ – T diagram of outer neutron-star envelopes at two magnetic field strengths, B = 1012 G and 1015 G, assuming fully-ionized iron (this assumption may be crude in the lower left part of the diagram). In the strongly-quantizing magnetic-field domain, bounded by ρB and Tcycl, the dependence TF(ρ) is steeper than at B = 0, in agreement with Eq. (56). The line Tm(ρ) separates Coulomb liquid from Coulomb crystal. Near the lower right corner of the figure, where T ≪ Tp, the quantum effects on the ions become important (i.e., the ions cannot be treated as classical pointlike particles). In the lower left corner, at ρ < ρs, the plasma is unstable to the phase separation into the gaseous and condensed phases (this phase transition will be discussed in Sect. 7.3).
![]() |
Fig. 1 Characteristic density-temperature domains at B = 1012 G (blue online) and 1015 G (red online) for fully-ionized iron. Solid lines indicate the Fermi temperature as function of density, the dotted line shows the plasma temperature, the dot-dashed line shows the melting temperature as function of density, short and long dashes delimit the domains of strongly quantizing magnetic field and of magnetic condensation, respectively, and the heavy dots mark the critical point for the condensation (Sect. 7.3). |
4.2.3. Strongly degenerate electrons
If the electrons are strongly degenerate, then one can apply the Sommerfeld expansion
(Sect. 3.1) and obtain
,
where
is
the zero-temperature value and ΔF is a thermal correction. According to
Eqs. (24) and (52), the zero-temperature pressure
P0 is
(57)where
is the relativistic unit of pressure, nmax is the maximum
integer n for which
, and
.
According to Eqs. (24) and (51), the Fermi energy
ϵF is determined by the condition
(58)In order to obtain
the chemical potential
μe = ϵF + Δϵ
with fractional accuracy ~
,
we retain two terms in Eq. (24), insert
it into Eq. (51), approximate
in the vicinity of
by
(59)where
and
, and drop the
higher-order terms containing
. Then
(60)The thermal correction
to the pressure equals
(61)and the thermal
correction to the free energy and internal energy
(62)The leading
contribution to the heat capacity is
. As in the nonmagnetic case,
is
proportional to T at T → 0, but with a different
proportionality coefficient.
4.2.4. Strongly degenerate electrons in a strongly quantizing magnetic field
If the magnetic field is strongly quantizing and the electrons are strongly degenerate
(which corresponds to the triangular domains in Fig. 1 defined by conditions
ρ < ρB
and T < TF), then
In
the nonrelativistic (xB ≪ 1) and
ultrarelativistic (xB ≫ 1) limits, we have
and
,
respectively. Compared with the nonmagnetic case (Papers I and II), the dependence
is
steeper, but P is lower everywhere except for
ne ≈ nB.
Thus, a strongly quantizing magnetic field softens the EOS of degenerate electrons.
The thermal corrections (Eqs. (60)–(62)) simplify to
The
last equation differs from the nonmagnetic case (Paper II) in that
xB replaces
xr and π2/3
replaces π2.
4.2.5. Nonrelativistic limit
In the nonrelativistic limit
(pF ≪ mec and
T ≪ Tr), Eqs. (51) and (52) simplify to (69)In the
nondegenerate regime (T ≫ TF), one has
Iν(χ) ≈ eχ Γ(ν + 1),
where Γ(ν + 1) is the gamma-function. Then Eq. (69) yields
and
(70)which provides the free
energy
.
As follows from Eq. (70), the reduced
internal energy and heat capacity
of the
Boltzmann gas decrease with increasing ζ. In a strongly quantizing
magnetic field (ζ ≫ 1), they tend to 1/2 instead of
3/2 because the gas becomes effectively one-dimensional. The only
kinetic degree is along the magnetic field.
In the nonquantizing field (ζ ≪ 1), the two last terms in Eq. (70) cancel out, so that the standard
expression is recovered. In the strongly
quantizing, nondegenerate regime
(ρ < ρB
and
TF ≪ T ≪ Tcycl),
the last term of Eq. (70) vanishes,
which yields
(71)
4.3. Thermodynamic and kinetic pressures
The above expressions for pressure of a magnetized gas of charged particles are based on
the principles of thermodynamics (Landau &
Lifshitz 1980), according to which
P = −(∂F/∂V)T,B.
Alternatively, the pressure can be calculated from the microscopic dynamics as the sum of
the changes of kinetic momenta of all particles reflected off a unit surface per unit
time. The result of the latter calculation, the kinetic pressure
Pkin, depends on the orientation of the surface relative to
B (Canuto & Chiu
1968). If the surface is perpendicular to B, then
one gets the kinetic pressure , which
acts along the field lines, the longitudinal pressure. If the surface is parallel to the
field, one gets a different (transverse) kinetic pressure, which can be expressed (Blandford & Hernquist 1982) as
(72)where
M = −∂Ω/∂B is
the magnetization.
In order to resolve the apparent paradox, one should take the magnetization current density jm = c ∇ × M into account; when boundaries are present, this volume current should be supplemented by the surface current cM × B/B (see, e.g., Griffith 1999). As argued by Blandford & Hernquist (1982), if we compress the electron gas perpendicular to B then we must do work against the Lorentz force density jm × B/c, which gives an additional contribution to the total transverse pressure and makes it equal to P. Because this point still causes confusion in some publications, let us illustrate it with a simple example.
If the pressure were anisotropic, then one might expect an anisotropic density gradient in a strongly magnetized star. Let us consider a small volume element in the star, assuming that we can treat B, T, and gravitational acceleration g as constants within this volume, and z axis is directed along g. Hydrostatic equilibrium implies that the density of gravitational force, ρg, is balanced by the density of forces created by plasma particles. The crucial point is that the magnetization contributes to this balance.
Let us compare the cases where B is parallel and
perpendicular to g. In the first case, the
z-component of the Lorentz force is absent, and we get the standard
equation of hydrostatic equilibrium: . In the second case, the
gradients of the kinetic pressure
and of the Lorentz force
density B dM/dz
act in parallel. In the constant and uniform magnetic field,
dM/dz is not zero, but is related
to the density gradient:
(73)Then the
equilibrium condition takes the form
which
is the same as in the first case. Furthermore, one can express P through
ρ using some EOS. In the considered example,
dP/dz = [∂P(ρ,T,B)/∂ρ] dρ/dz,
so that dρ/dz is also the same in
both cases. Thus, the stellar hydrostatic profile is determined by the isotropic
thermodynamic pressure P, which automatically includes magnetization.
5. Magnetic effects on the EOS of a Coulomb liquid
5.1. Electron exchange and correlation
The effects of a magnetic field on the contribution to the free energy due to electron
exchange and correlation were studied either in the regime of strong degeneracy and
strongly quantizing magnetic fields (Danz &
Glasser 1971; Banerjee et al. 1974; Fushiki et al. 1989; see also Morbec & Capelle 2008 for an instructive discussion of the
previous results and the inclusion of the second Landau level contribution), or at low
densities (Alastuey & Jancovici 1980; Cornu 1998; Steinberg
et al. 1998, 2000). In a previous work
(Potekhin et al. 1999) we suggested a
modification of the field-free expression for Fee, which
matches available exact limiting expressions, including the cases of nonquantizing,
strongly quantizing degenerate, and strongly quantizing nondegenerate plasmas. The
modification consists in replacing
Fee(θ,Γe) by
Fee(θ ∗ ,Γe),
where (74)ξ = [1−(4/ζ)tanh(ζ/4)] 1/2,
and θ0 and θB are
given by Eqs. (2) and (56), respectively, at fixed
ne and T.
5.2. Wigner-Kirkwood term
For the same reason as in Sect. 3.2.1, the treatment
of the quantum effects in the ion liquid is restricted by the Wigner-Kirkwood term. Its
expression in an arbitrary magnetic field was obtained by Alastuey & Jancovici (1980): (75)The function in the
square brackets monotonously varies from 1 at ζi → 0 to
1/3 at ζi → ∞, reflecting the effective
reduction of the degrees of freedom of the classical ion motion from
d = 3 at B = 0 to d = 1 for a strongly
quantizing field. At small ζi,
.
5.3. Electron-ion correlations
Using the linear response theory in the Thomas-Fermi limit, Fushiki et al. (1989) evaluated the electron polarization energy for a
dense plasma in a strongly quantizing magnetic field at zero temperature, assuming that
the ions remain classical (unaffected by the field). A comparison with the analogous
zero-field result shows that the strongly quantizing magnetic field
() increases the polarization
energy at high densities (rs ≪ 1) by a factor of
(Potekhin et al. 1999).
Recently, Sharma & Reddy (2011) calculated the screening of the ion-ion potential due to electrons in a large magnetic field B at T = 0, using the one-loop representation of the polarization function. Their results for the strongly quantizing magnetic field show that the screening is anisotropic, and the screened ion potential exhibits Friedel oscillations with period πħ/pF in a cylinder of a radius ~πħ/pF along the magnetic field line that passes through the Coulomb center. Sharma & Reddy suggest that this long-range oscillatory behavior can affect the ion lattice structure. However, finite temperature should damp these oscillations, so that they are pronounced only at T ≪ TF, i.e., deep within the triangular domains formed by the lines TF and ρB in Fig. 1. At the typical pulsar magnetic fields B ~ 1012 G, this requires an unusually low temperature of the neutron-star crust. On the other hand, the conditions T ≪ TF and ρ < ρB can be easily fulfilled in the outer crust of magnetars at B ~ 1015 G (cf. Fig. 1 and the top panel of Fig. 8), but in this case the Friedel oscillations are strongly suppressed because the electrons are ultrarelativistic.
To the best of our knowledge, the magnetic effects on the electron polarization energy have not been calculated at finite temperatures or in the case where the field is not strongly quantizing. In view of the limited scope and limited applicability of the available results on the magnetic effects, we use the nonmagnetic expression for Fie in our code (Appendix C).
6. Harmonic Coulomb crystals in the magnetic field
The magnetic effects on Coulomb crystals have been studied only in the harmonic approximation. Nagai & Fukuyama (1982, 1983) calculated phonon spectra of body-centered cubic (bcc), face-centered cubic (fcc), and hexagonal closely-packed (hcp) OCP lattices. They compared the energies of zero-point vibrations at different values of parameters β and RS and found conditions of stability of every lattice type. However, Baiko (2000, 2009) noticed that their choice of the magnetic-field direction did not provide the minimum of the total energy.
Usov et al. (1980) obtained the equations for oscillation modes of a harmonic OCP crystal and studied its phonon spectrum in a quantizing magnetic field in several limiting cases. These authors discovered a “soft” phonon mode with dispersion relation ωkα ∝ k2 near the center of the Brillouin zone, which leads to the unusual dependence of the heat capacity of the lattice CV,lat ∝ T3/2 at T → 0 instead of the Debye law CV,lat ∝ T3. Usov et al. (1980) argued that a strong magnetic field should increase stability of the crystal.
Baiko (2000, 2009) studied the magnetic effects on the phonon spectrum of the harmonic Coulomb crystals and calculated its energy, entropy, and heat capacity. We have found that his results can be approximately reproduced by the analytical expressions presented below.
6.1. Thermal phonon contributions
Without a magnetic field, the thermal phonon contribution fth to the reduced free energy of a Coulomb crystal F/NikBT is a function of a single argument η, described by a simple analytical expression (Baiko et al. 2001). The magnetic field introduces the second independent dimensionless argument β. The functional dependence of thermodynamic functions on η and β is not simple. Baiko (2009) identified five characteristic sectors of the η – β plane:
-
1.
η < 1 and β < η-1 – weakly magnetized classical crystal,
-
2.
η > 1 and β < η-1 – weakly magnetized quantum crystal,
-
3.
η < 1 and β > η-1 – strongly magnetized classical crystal,
-
4.
η > β > η-1 – strongly magnetized quantum crystal,
-
5.
β > η > 1 – very strongly magnetized quantum crystal.
For astrophysical applications, we have constructed an analytical representation of the EOS of the magnetized Coulomb crystal, which is asymptotically exact in each of the five sectors far from their boundaries, exactly recovers the nonmagnetic fit of Baiko et al. (2001) in the limit β → 0, and reaches a reasonable compromise between simplicity and accuracy.
The term fth in Eq. (32) can be rewritten as
fth = uth − sth,
where
uth = Uth/NikBT
and
sth = Sth/NikB
are the thermal contributions to the reduced internal energy and entropy. We approximately
represent uth by the function (76)where
ψ = 12.5 (β/η)3/2 + 119 (β/η)2
and ζi = βη, and we represent
sth by the function
(77)In these equations,
and
are the values of uth and sth at
β = 0. Equations (76)
and (77) exactly reproduce the known
asymptotic limits: uth = 3 in the classical nonmagnetic limit
(η ≪ β ≪ 1), uth = 2 in
the classical magnetic limit (η ≪ β-1 ≪ 1),
uth = 1 in the case where
β ≫ η ≫ 1, and
uth = 0.6sth ∝ η−3/2,
if η → ∞ at β = constant.
The functions and
are displayed in
Figs. 2 and 3.
Their accuracy is seen from a comparison with the numerical results (Baiko 2009), also shown in the figures. However, if the complete
consistency of different thermodynamic functions is required, Eqs. (76) and (77) should not be used directly, but should be first combined into
Then
one can calculate thermodynamic functions by differentiating the function
fth(η,β). In this way we obtain, for
example,
Note
that the relation between the phonon contributions to pressure and internal energy,
2PthV = Uth,
which is standard for a nonmagnetized harmonic crystal, is invalid in the strongly
magnetized crystal because both dimensionless arguments η and
β depend on density.
Approximations in Eqs. (78)–(80) are shown in Figs. 2–4. Their reasonable behavior beyond the range of available numerical data is demonstrated by plotting them also at larger β = 103 and 104.
![]() |
Fig. 2 Thermal phonon contribution to the reduced internal energy uth = Uth/NikBT as a function of log (T/Tp) = −log η at β = ħωci/kBTp = 0, 1, 10, 100, and 103 (numbers near the lines). The analytical approximation in Eq. (76) (dotted lines) and in Eq. (78) (short-dashed lines) are compared with the numerical results of Baiko (2009) (solid lines for β = 1, 10, and 100). |
![]() |
Fig. 3 Thermal phonon contribution to the reduced entropy
sth = Sth/NikBT
as a function of
log (T/Tp) at
β = ħωci/kBTp = 0,
0.1, 1, 10, 100, and 104 (numbers near the lines). The analytical
approximations in Eq. (77) (dotted
lines) and in Eq. (79) (dashed
lines) are compared with the numerical results of Baiko (2009) (solid lines for |
![]() |
Fig. 4 Thermal phonon contribution to the reduced heat capacity CV,lat/NikBT as a function of log (T/Tp) at β = ħωci/kBTp = 0, 1, 10, 100, and 103 (numbers near the lines). The analytical approximation in Eq. (80) (short-dashed lines) is compared with the numerical results of Baiko (2009) (solid lines for β = 0, 1, 10, and 100). The dotted lines correspond to the first term on the r.h.s. of Eq. (80). |
6.2. Zero-point vibrations
Because motions of the ions are confined by the magnetic field in the transverse direction, they exhibit quantum oscillations in the ground state (Landau 1930; Landau & Lifshitz 1977). The energy of these oscillations is ħωci/2 for every ion, which gives the term ζi/2 in Eq. (42). In a Coulomb crystal, the motion of an ion is confined in an effective potential well, centered at its equilibrium lattice site. The total energy of the zero-point quantum lattice oscillations Uq is given by Eq. (30).
In the case where the crystal is placed in a magnetic field,
Uq includes contributions due to both magnetic and lattice
confinements of the ion motion. However, since the magnetic contribution
Niħωci/2 is
common in all phase states, we take it as the zero energy point in our code and separate
it from the lattice contribution that is specific to the solid phase. Then Eq. (30) becomes
(82)and in Eq. (32) we have
, where we have defined
(83)The reduced frequency
moment
still depends
on β, because the character of ion vibrations is affected by the magnetic
field (they become essentially one-dimensional if B is extremely large),
but the latter dependence is relatively weak. Having extracted
from the
available numerical results for u1 (Baiko 2009; Baiko & Yakovlev 2012), we can represent it by the
simple interpolation
(84)where
is the
zero-field value and
is the limit
of
at β → ∞. Only
one of the three phonon branches contributes to
in the latter
limit, therefore
. For
the bcc crystal,
, whereas
varies
between 0.18 and 0.19 depending on the orientation of the lattice in the magnetic field.
In Fig. 5 we show
and the
logarithm (base 10) of u1 versus β.
![]() |
Fig. 5 Reduced first moment of phonon frequencies |
Usov et al. (1980) noticed that the energy of a
crystal depends on its orientation in a strong magnetic field. However, numerical
calculations (Baiko 2000, 2009) show that this dependence is very weak. For example, the
difference Δu1 between the values of
u1 for two orientations, where the field lines connect an
ion with its nearest neighbor in the first case and with a next-order nearest neighbor in
the second case, is approximately (85)with
saturation level (Δu1)max = 0.0064 for the bcc
lattice.
6.3. Comparison with the Baiko-Yakovlev fit
After the present work was completed, we became aware of an independent study by Baiko & Yakovlev (2012, priv. comm.), who developed another set of approximations for the free energy of the harmonic Coulomb crystal in a magnetic field. They presented the free energy as a sum of three terms corresponding to the contributions from each of the three phonon modes in the bcc crystal. Thus each of these terms has a clear physics meaning, while our fitting expressions give only the total contribution, which cannot be easily decomposed to three parts corresponding to the separate phonon modes.
Unlike our fit, the fit of Baiko & Yakovlev (2012) does not exactly reproduce the very accurate results of Baiko et al. (2001) in the limit β → 0. At finite β, both sets of fitting expressions accurately reproduce the asymptotes at T ≪ Tp and T ≫ Tp and have similar accuracies within several percent points in the intermediate range 0.1Tp/β ≲ T ≲ 10Tp. Meanwhile, our approximation is simpler: the Baiko-Yakovlev approximation contains 27 independent numerical fitting parameters, whereas our fits (76) and (77) contain together only 9 such parameters.
7. Examples and discussion
7.1. Thermodynamic functions
Characteristic features of the EOS can be seen in Fig. 6. Here, we have chosen the plasma parameters that are typical for outer
envelopes of isolated neutron stars: we consider fully-ionized iron
(Z = 26, A = 56) at
T = 107 K and B = 1012 G (for
illustration, the density range is extended to ρ ≲ 105
neglecting the bound states that can be important in this
ρ – T domain). We plot the normalized pressure
p = P/nikBT,
entropy
S/NikB,
heat capacity
cV = CV/NikB,
and logarithmic derivatives of pressure χρ
and χT as functions of density. Dashed lines
show these functions in the absence of quantizing magnetic field. The vertical dotted
lines marked by numbers separate different characteristic domains, consecutively entered
with increasing density: onset of electron degeneracy at B = 0
() and at
B = 1012 G
(TF = T), population of excited Landau
levels (ρ = ρB), melting
point with formation of a classical Coulomb crystal
(Tm = T), and quantum effects in the
crystal (Tp = T).
At low densities, the ideal-gas values are approached: p = 1 + Z, χρ = χT = 1, cV = (3 + 3Z)/2 at B = 0, and cV = (3 + Z)/2 at B = 1012 G. The latter difference is because at ρ < ρB the electron gas is effectively one-dimensional due to the strong magnetic quantization.
With increasing density, the reduced pressure p first decreases below its ideal-gas value due to the Coulomb nonideality and then increases due to the electron degeneracy. The increase occurs earlier at zero field than in the strong magnetic field, because of the delayed onset of the degeneracy (Sect. 4). When ρ > ρB ≈ 1.5 × 104 g cm-3, the thermodynamic functions approach their zero-field values. The gradually decreasing oscillations correspond to consecutive filling of the electron Landau levels. The magnetic field B = 1012 G does not affect the ion contributions in Fig. 6, because it is nonquantizing for the iron nuclei at T = 107 K (ζi = 0.00342).
The liquid-solid phase transition occurs in Fig. 6 at ρ ≈ 8.25 × 104 g cm-3, where we adopt the classical OCP melting condition Γ = 175.2 (Paper I). With further increase in density (ρ ≳ 106) the degeneracy becomes so strong that the energy and pressure are nearly independent of T and χT strongly decreases. The normalized heat capacity gradually tends to its value cV = 3 characteristic of the classical simple crystal. At still higher density the ion motions become quantized (Tp ≫ T) which leads to the further decrease in the heat capacity and the entropy.
![]() |
Fig. 6 Reduced thermodynamic functions
P/nikBT,
S/NikB,
CV/NikB,
χρ, and
χT for a fully-ionized nonmagnetic
(dashed lines) and magnetized (B = 1012 K, solid lines)
iron plasma at T = 107 K. The vertical dotted lines mark
the densities at which (1) |
![]() |
Fig. 7 Characteristics of the melting transition of nonideal carbon and iron plasmas at different field strengths B (marked near the curves). Lower panel: the value Γm of the Coulomb coupling parameter Γ at the melting point as function of mass density ρ. Upper panel: normalized latent heat per ion at the melting transition. The dot-dashed and dashed segments of the curves correspond to the domains of nonperturbative quantum effects (T < 0.5 Tp) and electron response (Z2 Ry > 0.1 ϵF), respectively. The dotted horizontal lines mark the OCP values. The filled and open circles mark the positions of the real and virtual condensed surfaces (see text in Sect. 7.3). |
7.2. Melting
The electron polarization, ion quantum effects, and quantizing magnetic field can shift the melting temperature. The lower panel of Fig. 7 shows the Coulomb coupling parameter Γ at the melting (that is, the value Γm at which the free energies of the two phases are equal to each other); the upper panel displays the difference between the internal energies in the liquid and solid phases at the melting point (the latent heat Qm = Uliq − Usol), divided by NikBT. We plot the data for fully ionized 12C and 56Fe at B = 0 and 1013 G for carbon, B = 0, 1014 G, and 1015 G for iron. The density range shown in the figure is typical for the outer envelopes of a neutron star and is also relevant for white dwarfs.
The position of the melting point is very sensitive to the accuracy of the free energies of the Coulomb liquid and crystal (see, e.g., Paper I). The polarization and quantum corrections to the classical OCP free energy are not known sufficiently well for finding the position of the melting point in the whole interval of densities shown in Fig. 7. The dot-dashed curves in this figure correspond to the domain where T < 0.5 Tp. Here, the perturbation theory for the quantum effects in the liquid phase becomes progressively inaccurate. The dashed curves correspond to the domain where the binding energy of the hydrogen-like ion exceeds 10% of the Fermi energy, Z2 Ry > 0.1 ϵF, which corresponds to ae ≳ 0.6a0/Z. Here, the perturbation treatment of the electron polarization in the crystal starts to be inaccurate. In addition, the position of the melting point cannot be traced to Γ ≲ 100–120, because the available results for the anharmonic corrections to the free energy of a Coulomb crystal (Appendix B.2) are accurate only at larger Γ. Nevertheless, we can evaluate Γm in a certain interval of densities for each type of ions (the solid segments of the curves in the figure).
The values of Qm in Fig. 7 roughly (within a factor of two) agree with the OCP value
at Γm = 175.2 and
with the values used in theoretical models of white dwarf cooling (e.g., Hansen 2004 and references therein). Most of the
neutron star cooling models currently ignore the release of the latent heat at the
crystallization of the neutron star envelopes (e.g., Yakovlev & Pethick 2004, and references therein).
Figure 7 shows that the strong magnetic fields tend to decrease Γm and thus stabilize the Coulomb crystal, in qualitative agreement with previous conjectures (Ruderman 1971; Kaplan & Glasser 1972; Usov et al. 1980; Lai 2001). At densities ρ ~ 107–108 g cm-3 corresponding to the “sensitivity strip” in the neutron-star cooling theory (Yakovlev & Pethick 2004), the stabilization proves to be significant at the magnetar field strengths B = 1014–1015 G. The results for 56Fe in this B interval, shown in the lower panel of Fig. 7, can be roughly (within 10%) described by the formula Γm(B) ≈ Γm(0)/(1 + 0.2 β). At the typical pulsar field strengths, B = 1012–1013 G, the effect is noticeable at lower densities. However, these conclusions remain preliminary in the absence of an evaluation of the magnetic-field effects on the anharmonic corrections. In view of the limited applicability and incompleteness of the evaluation of Γm with account of the quantum, polarization, and magnetic effects, in applications we use the classical OCP value Γm = 175.2 as the fiducial melting criterion.
7.3. Magnetic condensation
Ruderman (1971) suggested that the strong magnetic
field may stabilize molecular chains (polymers) aligned with the magnetic field and
eventually turn the surface of a neutron star into the metallic solid state. Later studies
have provided support for this conjecture, although the critical temperature
Tcrit, below which this condensation occurs, remains very
uncertain. Condensed surface density ρs is usually estimated
as (86)where
ξ ~ 1 is an unknown numerical factor, which absorbs the theoretical
uncertainty (Lai 2001; Medin & Lai 2006). The value ξ = 1 corresponds
to the EOS provided by the ion-sphere model (Salpeter
1961), which is close to the uniform model of Fushiki et al. (1989). For comparison, the results of the zero-temperature
Thomas-Fermi model for 56Fe at
G
(Rögnvaldsson et al. 1993) can be approximated
(within 4%) by ρs,ξ with
, whereas
the finite-temperature Thomas-Fermi model of Thorolfsson
et al. (1998) does not predict magnetic condensation at all. Our EOS for
partially ionized hydrogen plasmas in strong magnetic fields (Potekhin et al. 1999; Potekhin
& Chabrier 2004) exhibits a phase transition with
K and critical
density
at 1 ≲ B12 ≲ 103. According to another study (Lai & Salpeter 1997; Lai 2001), Tcrit for hydrogen is several
times smaller.
Medin & Lai (2006) performed
density-functional calculations of the cohesive energy Qs of
the condensed phases of H, He, C, and Fe in strong magnetic fields. A comparison with
previous density-functional calculations of other authors prompts that
Qs may vary within a factor of two at
B12 ≳ 1, depending on the approximations (see Medin & Lai 2006 for references and
discussion). In a subsequent study, Medin & Lai
(2007) calculated the equilibrium densities of saturated vapors of He, C, and Fe
atoms and polymers above the condensed surfaces, and obtained
Tcrit at several values of B by equating
the vapor density to ρs. Unlike the previous authors, Medin & Lai (2006, 2007) have taken the electronic band structure of the condensed matter
into account self-consistently, but they did not allow for the atomic motion across the
magnetic field and mostly neglected the contributions of the excited atomic and molecular
states in the gaseous phase. Medin & Lai
obtained ρs assuming that the linear molecular chains form a
rectangular array with sides 2R in the plane perpendicular to
B, and that the distance a between the
nuclei along B in the condensed matter remains the same as
in the gaseous phase, so that
ρs = mi/4aR2
(Medin 2012, priv. comm.). Using Tables 3–5 of Medin
& Lai (2006) for a and R, we can
describe their results for the surface density of 12C
at
and 56Fe at
by Eq. (86) with
and
ξ = 0.55 ± 0.11, respectively.
Medin & Lai (2007) found that the critical
temperature is
Tcrit ≈ 0.08Qs/kB.
Their numerical results for He, C, and Fe can be roughly (within a factor of 1.5)
described as K
at 1 ≲ B12 ≲ 1000. For comparison, the results of Lai & Salpeter (1997) for H at
10 ≲ B12 ≲ 500 suggest
K. The
discrepancies between different estimates of ρs and
Tcrit reflect the current theoretical uncertainties.
The filled dots on the curves for magnetized carbon in both panels of Fig. 7 correspond to the condensed surface position in the
fully-ionized plasma model. In this model K and
ρcrit ≈ ρs,0.47.
With decreasing temperature T below Tcrit,
the surface density increases and tends to the limit
ρs,1 given using Eq. (86) as
ρs ≈ ρs,1/ [1 + 1.1 (T/Tcrit)5]
(cf. Fig. 1). At smaller densities, there is a
thermodynamically unstable region in this model, therefore the curves in Fig. 7 are not continued to the left beyond this point. For
the magnetized iron model, the melting curve does not cross the surface because
Tm > Tcrit.
In this case, the open circles mark the density that the condensed surface would have at
much smaller temperatures
T ~ 106 K ≪ Tcrit. The parts
of the curves to the left of the open circles cannot be reached in a stationary stellar
envelope.
7.4. Thermal structure of a magnetar envelope
The results presented above have a direct application to the calculations of the thermal and mechanical structure of neutron-star envelopes with strong magnetic fields. Figure 8 illustrates the structure of a typical magnetar envelope with the ground-state nuclear composition. For illustration we have assumed that the magnetar has mass 1.4 M⊙ and radius 12 km, and the considered patch of the stellar surface has effective temperature 106.5 K and magnetic field B = 1015 G inclined at 45°. The top panel shows the thermal structure of the envelope, which has been calculated by numerical solution of the system of heat balance equations, taking the general relativity effects and neutrino emission into account (Potekhin et al. 2007). The middle panel presents the ion charge Z as function of ρ (Rüster et al. 2006). In the bottom panel (analogous to Fig. 6) we plot several reduced thermodynamic functions of ρ and T along the thermal profile (i.e., taking T from the top panel), starting at the condensed solid surface.
![]() |
Fig. 8 Structure of a magnetar envelope having the ground-state nuclear composition, the effective temperature 106.5 K, and magnetic field B = 1015 G, inclined at 45° to the surface. Top panel: thermal profiles calculated using the present EOS (solid red line) and the EOSs where the Coulomb nonideality is either neglected (dotted green line) or treated without account of magnetic quantization (dashed blue line, which is superposed on the solid red line). The melting temperature is drawn by the oblique dot-dashed line. Thin vertical dot-dashed lines mark the points of phase transitions from solid to liquid and back to solid state. Asterisks mark the ends of the convective segment, which is indicated by the arrow. Middle panel: ion charge (Rüster et al. 2006). Bottom panel: reduced pressure, heat capacity, and logarithmic derivatives of pressure. |
The temperature quickly grows at the solid surface and reaches the melting point at the depth z ≈ 7 cm. Thus, at the given conditions, the liquid ocean of a magnetar turns out to be covered by a thin layer of “ice” (solid substance). We treat the solid crust as immobile, but the liquid layer below the “ice” is convective up to the depth z ~ 1 m. We treat the convective heat transport through this layer in the adiabatic approximation (Schwarzschild 1958). The change of the heat-transport mechanism from conduction to convection causes the break of the temperature profile at the melting point. We underline that this treatment is only an approximation. In reality, the superadiabatic growth of temperature can lead to a hydrostatic instability of the shell of “ice” and eventually to its cracking and fragmentation into turning-up “ice floes”. This can result in transient enhancements of the thermal luminosity of magnetars.
The temperature profile flattens with density increase, and the Coulomb plasma freezes again at the interface between the layers of 66Ni and 86Kr at ρ = 1.5 × 109 g cm-3 (z = 73.8 m). These phase transitions do not cause any substantial breaks in χρ, χT, or CV/NikB, because the Coulomb plasmas have similar structure factors in the liquid and crystalline phases in the melting region (cf. Baiko et al. 1998).
At the boundaries between layers composed of different chemical elements, the reduced thermodynamic functions do not exhibit substantial discontinuities, except for the abrupt increases in P/nikBT at the interfaces 66Ni/86Kr (ρ = 1.5 × 109 g cm-3) and 78Ni/124Mo (ρ = 1.32 × 1011 g cm-3), which are caused by the decreases in ni with the large jumps in A (of course, the non-normalized pressure P is continuous). The specific heat per ion CV/NikB is almost continuous at these interfaces, which means that heat capacity of unit volume abruptly decreases. The drop in χT at the 78Ni/128Mo interface is due to the same decrease in ni, which leads to the decrease in the ionic contribution that mostly determines ∂P/∂T at the strong degeneracy.
The oscillations of the reduced thermodynamic functions (most noticeable for χρ and χT) correspond to consecutive population of excited Landau levels by degenerate electrons with density increase, analogous to the oscillations in Fig. 6.
The magnetic effects on the nonideal part of the plasma thermodynamic functions have almost no influence on the temperature profile in the magnetar envelope, as illustrated in the upper panel of Fig. 8 where the corresponding solid and dashed lines virtually coincide. For comparison, the dotted line in the upper panel shows the result of a calculation totally neglecting the Coulomb nonideality. In this case, the profile is quite different at low densities, where there is no longer a solid surface. However, even in this case the thermal profile is almost the same at large ρ. This means that the Coulomb nonideality has a minor impact on the relation between the internal and effective temperatures and therefore on the cooling curves (Yakovlev & Pethick 2004), but it can be important for the shape of the thermal spectrum (cf., e.g., Potekhin et al. 2012).
8. Conclusions
We have systematically reviewed analytical approximations for the EOS of fully-ionized electron-ion plasmas in magnetic fields and described several improvements to the previously published approximations, taking nonideality attributable to ion-ion, electron-electron, and electron-ion interactions into account. The presented formulae are applicable in a wide range of plasma parameters, including the domains of nondegenerate and degenerate, nonrelativistic and relativistic electrons, weakly and strongly coupled Coulomb liquids, classical and quantum Coulomb crystals. As an application, we have calculated and discussed the behavior of thermodynamic functions, melting, and latent heat at crystallization of strongly coupled Coulomb plasmas with the parameters appropriate for cooling white dwarfs and envelopes of nonmagnetized and strongly magnetized neutron stars. We have also shown that a typical outer envelope of a magnetar can have a liquid layer beneath the solid surface.
Acknowledgments
We are grateful to José Pons for pointing out a bug in a previous version of one of our subroutines, to D. A. Baiko and D. G. Yakovlev for making their results available to us prior to publication and for useful discussions, and to D. G. Yakovlev for valuable remarks on a preliminary version of this paper. The work of A.Y.P. was partially supported by the Ministry of Education and Science of the Russian Federation (Agreement No. 8409, 2012), the Russian Foundation for Basic Research (RFBR grant 11-02-00253-a), and the Russian Leading Scientific Schools program (grant NSh-3769.2010.2).
References
- Alastuey, A., & Jancovici, B. 1980, Phys. A, 102, 327 [Google Scholar]
- Baiko, D. A. 2000, Ph.D. Thesis (St. Petersburg: Ioffe Phys.-Tech. Inst.) [in Russian] [Google Scholar]
- Baiko, D. A. 2002, Phys. Rev. E, 66, 056405 [NASA ADS] [CrossRef] [Google Scholar]
- Baiko, D. A. 2009, Phys. Rev. E, 80, 046405 [NASA ADS] [CrossRef] [Google Scholar]
- Baiko, D. A., Kaminker, A. D., Potekhin, A. Y., & Yakovlev, D. G. 1998, Phys. Rev. Lett., 81, 5556 [NASA ADS] [CrossRef] [Google Scholar]
- Baiko, D. A., Potekhin, A. Y., & Yakovlev, D. G. 2001, Phys. Rev. E, 64, 057402 [NASA ADS] [CrossRef] [Google Scholar]
- Banerjee, B., Constantinescu, D. H., & Rehak, P. 1974, Phys. Rev. D, 10, 2384 [NASA ADS] [CrossRef] [Google Scholar]
- Berestetskiĭ, V. B., Lifshitz, E. M., & Pitaevskiĭ, L. P., 1982, Quantum Electrodynamics (Oxford: Butterworth-Heinemann) [Google Scholar]
- Blandford, R. D., & Hernquist, L. 1982, J. Phys. C: Solid State Phys., 15, 6233 [Google Scholar]
- Blinnikov, S. I., Dunina-Barkovskaya, N. V., & Nadyozhin, D. K. 1996, ApJS, 106 171; erratum: 1998, ApJS, 118, 603 [NASA ADS] [CrossRef] [Google Scholar]
- Brami, B., Hansen, J. P., & Joly, F. 1979, Physica A, 95, 505 [NASA ADS] [CrossRef] [Google Scholar]
- Broderick, A., Prakash, M., & Lattimer, J. M. 2000, ApJ, 537, 351 [NASA ADS] [CrossRef] [Google Scholar]
- Caillol, J. M. 1999, J. Chem. Phys., 111, 6538 [NASA ADS] [CrossRef] [Google Scholar]
- Canuto, V., & Chiu, H.-Y. 1968, Phys. Rev., 173, 1220 [NASA ADS] [CrossRef] [Google Scholar]
- Chabrier, G., & Ashcroft, N. W. 1990, Phys. Rev. A, 42, 2284 [NASA ADS] [CrossRef] [Google Scholar]
- Chabrier, G., & Baraffe, I. 2000, ARA&A, 38, 337 [NASA ADS] [CrossRef] [Google Scholar]
- Chabrier, G., & Potekhin, A. Y. 1998, Phys. Rev. E, 58, 4941 [NASA ADS] [CrossRef] [Google Scholar]
- Chabrier, G., Douchin, F., & Potekhin, A. Y. 2002, J. Phys., Condensed Matter, 14, 9133 [NASA ADS] [CrossRef] [Google Scholar]
- Chandrasekhar, S. 1957, An Introduction to the Study of Stellar Structure (New York: Dover, 1957), Chap. X [Google Scholar]
- Cornu, F. 1998, Phys. Rev. E, 58, 5293 [NASA ADS] [CrossRef] [Google Scholar]
- Danz, R. W., & Glasser, M. L. 1971, Phys. Rev. B, 4, 94 [NASA ADS] [CrossRef] [Google Scholar]
- DeWitt, H. E., & Slattery, W. 2003, Contrib. Plasma Phys., 43, 279 [NASA ADS] [CrossRef] [Google Scholar]
- DeWitt, H. E., Slattery, W., & Chabrier, G. 1996, Phys. A, 228, 21 [Google Scholar]
- Farouki, R. T., & Hamaguchi, S. 1993, Phys. Rev. E, 47, 4330 [NASA ADS] [CrossRef] [Google Scholar]
- Fortov, V. E. 2009, Phys. Usp., 52, 615 [NASA ADS] [CrossRef] [Google Scholar]
- Fushiki, I., Gudmundsson, E. H., & Pethick, C. J. 1989, ApJ, 342, 958 [NASA ADS] [CrossRef] [Google Scholar]
- Galam, S., & Hansen, J. P. 1976, Phys. Rev. A, 14, 816 [NASA ADS] [CrossRef] [Google Scholar]
- Girifalco, L. A. 1973, Statistical Physics of Materials (New York: Wiley-Interscience, 1973), App. V [Google Scholar]
- Griffith, D. J. 1999, Introduction to Electrodynamics, 3rd ed. (London: Prentice-Hall), Chap. 6. [Google Scholar]
- Haensel, P., Potekhin, A. Y., & Yakovlev, D. G. 2007, Neutron Stars 1: Equation of State and Structure (New York: Springer) [Google Scholar]
- Hamaguchi, S., Farouki, R. T., & Dubin, D. H. E. 1997, Phys. Rev. E, 56, 4671 [NASA ADS] [CrossRef] [Google Scholar]
- Hansen, B. M. S. 2004, Phys. Rep., 399, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Hansen, J. P., & Vieillefosse, P. 1975, Phys. Lett. A, 53, 187 [NASA ADS] [CrossRef] [Google Scholar]
- Hansen, J. P., Torrie, G. M., & Vieillefosse, P. 1977, Phys. Rev. A, 16, 2153 [NASA ADS] [CrossRef] [Google Scholar]
- Hughto, J., Horowitz, C. J., Schneider, A. S., et al. 2012, Phys. Rev. E, accepted [arXiv:1211.0891] [Google Scholar]
- Ichimaru, S., Iyetomi, H., & Tanaka, S. 1987, Phys. Rep., 149, 91 [NASA ADS] [CrossRef] [Google Scholar]
- Jancovici, B. 1962, Nuovo Cimento, 25, 428 [Google Scholar]
- Johnson, M. H., & Lippmann, B. A. 1949, Phys. Rev., 76, 828 [NASA ADS] [CrossRef] [Google Scholar]
- Kaplan, J. L., & Glasser, M. L. 1972, Phys. Rev. Lett., 28, 1077 [NASA ADS] [CrossRef] [Google Scholar]
- Kittel, C. 1963, Quantum Theory of Solids (New York: Wiley) [Google Scholar]
- Lai, D. 2001, Rev. Mod. Phys., 73, 629 [NASA ADS] [CrossRef] [Google Scholar]
- Lai, D., & Salpeter, E. E. 1997, ApJ, 491, 270 [NASA ADS] [CrossRef] [Google Scholar]
- Landau, L. D. 1930, Z. f. Physik, 64, 629 [Google Scholar]
- Landau, L. D., & Lifshitz, E. M. 1977, Quantum Mechanics. Non-Relativistic Theory, 3rd edn. (Oxford: Pergamon) [Google Scholar]
- Landau, L. D., & Lifshitz, E. M. 1980, Statistical Physics, 3rd edn. (Oxford: Butterworth-Heinemann) [Google Scholar]
- Medin, Z., & Cumming, A. 2010, Phys. Rev. E, 81, 036107 [NASA ADS] [CrossRef] [Google Scholar]
- Medin, Z., & Lai, D. 2006, Phys. Rev. A, 74, 062508 [NASA ADS] [CrossRef] [Google Scholar]
- Medin, Z., & Lai, D. 2007, MNRAS, 382, 1833 [NASA ADS] [Google Scholar]
- Militzer, B., & Graham, R. L. 2006, J. Phys. Chem. Sol., 67, 2136 [NASA ADS] [CrossRef] [Google Scholar]
- Morbec, J. M., & Capelle, K. 2008, Phys. Rev. B, 78, 085107 [NASA ADS] [CrossRef] [Google Scholar]
- Nagai, T., & Fukuyama, H. 1982, J. Phys. Soc. Japan, 51, 3431 [NASA ADS] [CrossRef] [Google Scholar]
- Nagai, T., & Fukuyama, H. 1983, J. Phys. Soc. Japan, 52, 44 [NASA ADS] [CrossRef] [Google Scholar]
- Ogata, S., Iyetomi, H., Ichimaru, S., & Van Horn, H. M. 1993, Phys. Rev. E, 48, 1344 [NASA ADS] [CrossRef] [Google Scholar]
- Paxton, B., Bildsten, L., Dotter, A., et al. 2011, ApJS, 192, 3 [Google Scholar]
- Pollock, E. L., & Hansen, J. P. 1973, Phys. Rev. A, 8, 3110 [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., & Chabrier, G. 2000, Phys. Rev. E, 62, 8554 (Paper I) [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., & Chabrier, G. 2004, ApJ, 600, 317 [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., & Chabrier, G. 2010, Contrib. Plasma Phys., 50, 82 (Paper II) [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., Chabrier, G., & Shibanov, Yu. A. 1999, Phys. Rev. E, 60, 2193 [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., Chabrier, G., & Yakovlev, D. G. 2007, Ap&SS, 308, 353 (electronic version with corrected misprints [arXiv:astro-ph/0611014v3] [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., Chabrier, G., & Rogers, F. J. 2009a, Phys. Rev. E, 79, 016411 [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., Chabrier, G., Chugunov, A. I., DeWitt, H. E., & Rogers, F. J. 2009b, Phys. Rev. E, 80, 047401 [NASA ADS] [CrossRef] [Google Scholar]
- Potekhin, A. Y., Suleimanov, V. F., van Adelsberg, M., & Werner, K. 2012, A&A, 546, A121 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Rögnvaldsson, Ö. E., Fushiki, I., Gudmundsson, E. H., Pethick, C. J., & Yngvason, J. 1993, ApJ, 416, 276 [NASA ADS] [CrossRef] [Google Scholar]
- Rosenfeld, Y. 1996, Phys. Rev. E, 54, 2827 [NASA ADS] [CrossRef] [Google Scholar]
- Ruderman, M. A. 1971, Phys. Rev. Lett., 27, 1306 [NASA ADS] [CrossRef] [Google Scholar]
- Rüster, S. B., Hempel, M., & Schaffner-Bielich, J. 2006, Phys. Rev. C, 73, 035804 [NASA ADS] [CrossRef] [Google Scholar]
- Salpeter, E. E. 1961 ApJ, 134, 669 [Google Scholar]
- Schwarzschild, M. 1958, Structure and Evolution of the Stars (Princeton: Princeton Univ. Press) [Google Scholar]
- Schwinger, J. 1988, Particles, Sources, and Fields (Redwood: Addison-Wesley) [Google Scholar]
- Sharma, R., & Reddy, S. 2011, Phys. Rev. C, 83, 025803 [NASA ADS] [CrossRef] [Google Scholar]
- Steinberg, M., Ortner, J., & Ebeling, W. 1998, Phys. Rev. E, 58, 3806 [NASA ADS] [CrossRef] [Google Scholar]
- Steinberg, M., Ebeling, W., & Ortner, J. 2000, Phys. Rev. E, 61, 2290 [NASA ADS] [CrossRef] [Google Scholar]
- Suh, I.-S., & Mathews, G. J. 2001, ApJ, 546, 1126 [NASA ADS] [CrossRef] [Google Scholar]
- Tanaka, S., Mitake, S., & Ichimaru, S. 1896, Phys. Rev. A, 32 [Google Scholar]
- Thorolfsson, A., Rögnvaldsson, Ö. E., Yngvason, J., & Gudmundsson, E. H. 1998, ApJ, 502, 847 [NASA ADS] [CrossRef] [Google Scholar]
- Timmes, F. X., & Arnett, D. 1999, ApJS, 125, 277 [NASA ADS] [CrossRef] [Google Scholar]
- Timmes, F. X., & Swesty, F. D. 2000, ApJS, 126, 501 [NASA ADS] [CrossRef] [Google Scholar]
- Usov, N. A., Grebenshchikov, Yu. B., & Ulinich, F. R. 1980, Sov. Phys.–JETP, 51, 148 [NASA ADS] [Google Scholar]
- van Leeuwen, J. H. 1921, J. Phys. Radium, Ser. VI, 2, 361 [Google Scholar]
- Wickramasinghe, D. T., & Ferrario, L. 2000, PASP, 112, 873 [NASA ADS] [CrossRef] [Google Scholar]
- Wigner, E. P. 1932, Phys. Rev. 40, 749 [Google Scholar]
- Yakovlev, D. G., & Kaminker, A. D. 1994, in The Equation of State in Astrophysics, eds. G. Chabrier, & E. Schatzman (Cambridge: Cambridge University Press), Proc. IAU Colloq., 147, 214 [Google Scholar]
- Yakovlev, D. G., & Pethick, C. J. 2004, ARA&A, 42, 169 [NASA ADS] [CrossRef] [EDP Sciences] [MathSciNet] [Google Scholar]
- Yakovlev, D. G., & Shalybkov, D. A. 1989, Astrophys. Space Phys. Rev. 7, 311 [Google Scholar]
Appendix A: Nonideal part of the free energy of electrons
Tanaka et al. (1985) calculated the interaction
energy of the electron fluid at finite T and presented a fitting formula
that reproduced their numerical results as well as the results of other authors in various
limits. Subsequently the behavior of the fit at
T ≪ TF was improved by Ichimaru et al. (1987A.1). The result reads
(A.1)where
A = b − gd,
B = a − g,
C = 2 − d2/f,
,
and a, b, d, f, and
g are the following functions of
θ = T/TF
(at B = 0):
The
accuracy of Eq. (A.1) is 1%.
In a quantizing magnetic field, we replace the argument θ in these expressions by the quantity θ ∗ defined by Eq. (74), as explained in Sect. 5.1.
Appendix B: Nonideal part of the free energy of ions in the rigid background
B.1. Coulomb liquid
For the reduced free energy
fii ≡ Fii/NikBT
of the classic OCP, we have the following analytical formula (Paper I): (B.1)where
A1 = −0.907347,
A2 = 0.62849, B1 = 0.0045,
B2 = 170,
B3 = −8.4 × 10-5,
B4 = 0.0037, and
.
The derivative
(B.2)reproduces Monte Carlo
calculations of the reduced internal energy
Uii/NikBT
at 1 ≤ Γ ≤ 190 (Caillol 1999) within the accuracy
of these calculations, ≲10-3. For any values of the coupling parameter in a
liquid OCP,
,
the fractional error of the approximation (B.1) does not exceed 2 × 10-4.
The classical treatment of ion motion is justified at
T ≫ Tp. One can extend the applicability
range of the analytical EOS to T ~ Tp by
the Wigner-Kirkwood quantum corrections (Wigner
1932; Landau & Lifshitz 1980).
The lowest-order correction to the reduced free energy is (B.3)The next-order
correction ∝ ħ4 was obtained by Hansen
& Vieillefosse (1975). It can be written as
(B.4)This
expression, unlike Eq. (B.3), is not
exact. Both corrections have limited applicability, because as soon as
η becomes large, the Wigner expansion diverges and the plasma forms a
quantum liquid, whose free energy is not known in an analytical form. Therefore we use
only the lowest-order correction (B.3),
i.e.,
. In
a magnetic field, Eq. (B.3) is replaced
by Eq. (75) (Sect. 5.2).
B.2. Coulomb crystal
The reduced free energy of an OCP in the crystalline phase is given by Eq. (32), where the first three terms describe
the harmonic lattice model (Baiko et al. 2001).
For the bcc crystal, we have C0 = −0.895 929 255 68
and u1 = 0.511 3875, and for
fth the following fitting formula can be used:
(B.5)where
α1 = 0.932446, α2 = 0.334547,
α3 = 0.265764,
The
Taylor expansion of Eq. (B.5) at small
η is consistent with the Wigner correction (B.3). However, the next Taylor term
~η3 is absent in the Wigner expansion, and therefore
Eq. (B.5) does not reproduce
higher-order Wigner corrections. Nevertheless, approximation (B.5) is very accurate: it reproduces the
numerical results in Baiko et al. (2001) with
fractional deviations within 5 × 10-6, and its first and second derivatives
reproduce the calculated contributions to the internal energy and heat capacity with
deviations up to several parts in 105. Other types of simple lattices are
described by the same expressions with slightly different parameters (see Baiko et al. 2001).
Anharmonic corrections for Coulomb lattices were studied in a number of works (see
Papers I and II for references). In the classical regime η → 0, we have
chosen one of the 11 parametrizations proposed by Farouki & Hamaguchi (1993): (B.6)where
a1 = 10.9, a2 = 247, and
a3 = 1.765 × 105. A continuation to arbitrary
η, which is consistent with available analytical and numerical
results for quantum crystals, reads (Paper II)
(B.7)Superstrong magnetic
fields can significantly change these expressions under the conditions
ζi ≳ 1 and η ≳ 1. Analytical
approximations for the free energy of a harmonic Coulomb crystal in quantizing magnetic
fields are derived in Sect. 6. Analogous results
for the anharmonic corrections are currently unavailable.
Appendix C: Electron polarization corrections
C.1. Coulomb liquid
The screening contribution to the reduced free energy of the Coulomb liquid at
0 < Γ ≲ 300 has been calculated by the HNC technique and
fitted by the expression (Paper I) (C.1)where
ensures
exact transition to the Debye-Hückel limit at Γ → 0,
cTF = (18/175) (12/π)2/3Z7/3(1 − Z − 1/3 + 0.2 Z−1/2(
fits the numerical data at large Γ and reproduces the Thomas-Fermi limit (Salpeter 1961) at Z → ∞, the
parameters a = 1.11 Z0.475,
b = 0.2 + 0.078 (lnZ)2,
and ν = 1.16 + 0.08lnZ provide a low-order
approximation to Fie for intermediate
rs and Γ. The functions
improve
the fit at relatively large rs. Finally, the function
is
the relativistic correction, as is
in the denominator.
C.2. Coulomb crystal
The screening contribution to the reduced free energy of the Coulomb crystals was
evaluated using the semiclassical perturbation approach with an effective structure
factor (Paper I) and fitted by the expression (Paper II) (C.2)where
the
parameter aTF = 0.00352 is related to
cTF in Eq. (C.1), and parameters s and
b1 – b4 depend on
Z:
Here,
the numerical parameters are given for the bcc crystal; their values for the fcc lattice
are slightly different (Paper I).
All Figures
![]() |
Fig. 1 Characteristic density-temperature domains at B = 1012 G (blue online) and 1015 G (red online) for fully-ionized iron. Solid lines indicate the Fermi temperature as function of density, the dotted line shows the plasma temperature, the dot-dashed line shows the melting temperature as function of density, short and long dashes delimit the domains of strongly quantizing magnetic field and of magnetic condensation, respectively, and the heavy dots mark the critical point for the condensation (Sect. 7.3). |
In the text |
![]() |
Fig. 2 Thermal phonon contribution to the reduced internal energy uth = Uth/NikBT as a function of log (T/Tp) = −log η at β = ħωci/kBTp = 0, 1, 10, 100, and 103 (numbers near the lines). The analytical approximation in Eq. (76) (dotted lines) and in Eq. (78) (short-dashed lines) are compared with the numerical results of Baiko (2009) (solid lines for β = 1, 10, and 100). |
In the text |
![]() |
Fig. 3 Thermal phonon contribution to the reduced entropy
sth = Sth/NikBT
as a function of
log (T/Tp) at
β = ħωci/kBTp = 0,
0.1, 1, 10, 100, and 104 (numbers near the lines). The analytical
approximations in Eq. (77) (dotted
lines) and in Eq. (79) (dashed
lines) are compared with the numerical results of Baiko (2009) (solid lines for |
In the text |
![]() |
Fig. 4 Thermal phonon contribution to the reduced heat capacity CV,lat/NikBT as a function of log (T/Tp) at β = ħωci/kBTp = 0, 1, 10, 100, and 103 (numbers near the lines). The analytical approximation in Eq. (80) (short-dashed lines) is compared with the numerical results of Baiko (2009) (solid lines for β = 0, 1, 10, and 100). The dotted lines correspond to the first term on the r.h.s. of Eq. (80). |
In the text |
![]() |
Fig. 5 Reduced first moment of phonon frequencies |
In the text |
![]() |
Fig. 6 Reduced thermodynamic functions
P/nikBT,
S/NikB,
CV/NikB,
χρ, and
χT for a fully-ionized nonmagnetic
(dashed lines) and magnetized (B = 1012 K, solid lines)
iron plasma at T = 107 K. The vertical dotted lines mark
the densities at which (1) |
In the text |
![]() |
Fig. 7 Characteristics of the melting transition of nonideal carbon and iron plasmas at different field strengths B (marked near the curves). Lower panel: the value Γm of the Coulomb coupling parameter Γ at the melting point as function of mass density ρ. Upper panel: normalized latent heat per ion at the melting transition. The dot-dashed and dashed segments of the curves correspond to the domains of nonperturbative quantum effects (T < 0.5 Tp) and electron response (Z2 Ry > 0.1 ϵF), respectively. The dotted horizontal lines mark the OCP values. The filled and open circles mark the positions of the real and virtual condensed surfaces (see text in Sect. 7.3). |
In the text |
![]() |
Fig. 8 Structure of a magnetar envelope having the ground-state nuclear composition, the effective temperature 106.5 K, and magnetic field B = 1015 G, inclined at 45° to the surface. Top panel: thermal profiles calculated using the present EOS (solid red line) and the EOSs where the Coulomb nonideality is either neglected (dotted green line) or treated without account of magnetic quantization (dashed blue line, which is superposed on the solid red line). The melting temperature is drawn by the oblique dot-dashed line. Thin vertical dot-dashed lines mark the points of phase transitions from solid to liquid and back to solid state. Asterisks mark the ends of the convective segment, which is indicated by the arrow. Middle panel: ion charge (Rüster et al. 2006). Bottom panel: reduced pressure, heat capacity, and logarithmic derivatives of pressure. |
In the text |
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.