Entrainment effects in neutron-proton mixtures within the nuclear energy-density functional theory. II. Finite temperatures and arbitrary currents.
Abstract
Mutual entrainment effects in hot neutron-proton superfluid mixtures are studied in the framework of the self-consistent nuclear energy-density functional theory. The local mass currents in homogeneous or inhomogeneous nuclear systems, which we derive from the time-dependent Hartree-Fock-Bogoliubov equations at finite temperatures, are shown to have the same formal expression as the ones we found earlier in the absence of pairing at zero temperature. Analytical expressions for the entrainment matrix are obtained for application to superfluid neutron-star cores. Results are compared to those obtained earlier using Landau’s theory. Our formulas, valid for arbitrary temperatures and currents, are applicable to various types of functionals including the Brussels-Montreal series for which unified equations of state have been already calculated, thus laying the ground for a fully consistent microscopic description of superfluid neutron stars.
I Introduction
Formed from the gravitational core-collapse of massive stars during supernova explosions, neutron stars are initially very hot but rapidly cool down by emitting neutrinos. Their very dense matter is thus expected to undergo various phase transitions Blaschke and Chamel (2018). In particular, the core of a mature neutron star is thought to contain a neutron-proton superfluid mixture (see, e.g., Ref. Chamel (2017a) for a recent review). Because a superfluid can flow without resistance and carries no heat, the dynamics of a neutron star must be described by three distinct components at least: the neutron superfluid, the proton superconductor, and the normal fluid. Due to strong nuclear interactions, neutrons and protons cannot flow completely independently and are mutually entrained, similarly to superfluid 4He-3He mixtures Andreev and Bashkin (1975).
Although a fully relativistic treatment is required for an accurate description of the global dynamics of neutron stars, the flows of neutrons and protons remain essentially nonrelativistic at the nuclear length scales of interest here (at such scales, spacetime curvature can also be safely ignored, as shown e.g. in Ref. Glendenning (2000)). Therefore, we shall consider here nonrelativistic superfluid dynamics. The mass current of one nucleon species ( for neutron, proton) is expressible as a combination of the “superfluid velocities” (momenta per unit mass) of both species and of the normal velocity of thermal excitations, as Gusakov and Haensel (2005)
(1) | |||||
(2) |
It follows from Eqs. (1) and (2) that the normal fluid carries a momentum density given by
(3) |
As shown in Ref. Gusakov and Andersson (2006), the relativistic entrainment matrix, denoted by and relating the nucleon four-currents to the superfluid four-velocities, can be inferred from its nonrelativistic counterpart .
The (symmetric) entrainment matrix is a key microscopic ingredient for modeling the dynamics of neutron stars, see, e.g. Refs. Andersson and Comer (2007); Chamel (2017a); Graber et al. (2017); Haskell and Sedrakian (2018) and references therein. It should be stressed that the entrainment matrix itself may depend on the superfluid flows, and this could play an important role on the dynamics of neutron stars Gusakov and Kantor (2013). Although observed neutron stars are usually cold, meaning that their internal temperature is much lower than the Fermi temperatures , thermal effects may still be significant for the superfluid dynamics since the associated critical temperatures are typically much lower than . In particular, it has been recently shown that the temperature-dependence of the entrainment matrix may have implications for neutron-star oscillations Dommes et al. (2019); Kantor et al. (2020).
The entrainment matrix of a neutron-proton superfluid mixture at finite temperatures was first calculated in Ref. Gusakov and Haensel (2005) within Landau’s theory of Fermi liquids suitably extended to deal with superfluid systems Larkin and Migdal (1963); Leggett (1965). Calculations were performed assuming are small compared to the corresponding Fermi velocities , and thus considering first-order current perturbations of the static state. The same approach was later extended to relativistic mixtures allowing for the presence of hyperons Gusakov et al. (2009). Landau parameters were calculated using a relativistic mean-field model including scalar self-interactions but ignoring pairing. Numerical results for the relativistic entrainment matrix were obtained using the Lagrangian parametrization of Ref. Glendenning (1985), and employing the same empirical fits for the dependence on the critical temperatures as in Ref. Gusakov and Haensel (2005). More recently, some nonlinear effects of the superfluid flows have been taken into account in the calculations of the neutron-proton entrainment matrix Leinson (2018). However, Landau’s theory, on which all these studies rely, is not self-consistent, as emphasized in Ref. Gusakov and Haensel (2005). Moreover this approach cannot be easily transposed to inhomogeneous systems such as the inner crust of neutron stars, where superfluid neutrons coexist with nuclear clusters.
We have recently calculated the entrainment matrix of neutron-proton mixtures at low temperatures Chamel and Allard (2019) within the self-consistent nuclear energy-density functional theory Duguet (2014). This theory has been already applied by different groups to determine the equation of state of cold dense matter using the same functional in all regions of neutron stars (crust, mantle, and core), thus ensuring a unified and thermodynamically consistent treatment, see, e.g. Refs. Douchin and Haensel (2001); Potekhin et al. (2013); Sharma et al. (2015); Pearson et al. (2018); Mutafchieva et al. (2019); Pearson et al. (2020); Mondal et al. (2020). Moreover, this theory has been also employed to compute the properties of superfluid neutrons in neutron-star crusts (see, e.g., Refs. Monrozeau et al. (2007); Chamel et al. (2010); Pastore (2015) for pairing gaps, critical temperatures, and specific heat; Refs. Chamel (2017b); Watanabe and Pethick (2017); Kashiwaba and Nakatsukasa (2019) for superfluid fractions) and the dynamics of quantized vortices Avogadro et al. (2007); Wlazłowski et al. (2016); Bulgac (2019). In this second paper, we extend our previous analysis to finite temperatures and arbitrary currents. The dependence of the entrainment matrix on temperature and superflows are taken into account fully self-consistently within the time-dependent Hartree-Fock-Bogoliubov (TDHFB) method.
After introducing the TDHFB method in Sect. II and deriving the general expressions for the local currents and superfluid velocities valid for any (homogeneous or inhomogeneous) nuclear system, calculations of the entrainment matrix in the outer core of neutron stars are presented in Sect. III. Landau’s approximations are also discussed. Throughout the paper, we shall ignore the small difference between the neutron and proton masses, and the nucleon mass will be denoted by .
II Time-dependent Hartree-Fock-Bogoliubov equations
II.1 Matrix formulation
The TDHFB method is discussed, e.g. in the classical textbook of Ref. Blaizot and Ribka (1986). The energy of a nucleon-matter element of volume is expressed as a function of the one-body density matrix and pairing tensor defined by the following thermal averages of the creation and destruction operators, and (using the symbol for Hermitian conjugation), for nucleons of charge type in a quantum state (using the symbol for complex conjugation):
(4) |
(5) |
Introducing the generalized Bogoliubov transformation111In Ref. Blaizot and Ribka (1986), the matrices were denoted by and .
(6) |
such that ( is the Kronecker’s symbol) and , where and are creation and destruction operators of a quasiparticle in a quantum state , the TDHFB equations, which formally take the same form at any temperature, can be written as Blaizot and Ribka (1986)
(7) |
(8) |
where denotes the chemical potentials. The matrices and of the single-particle Hamiltonian and the pair potential, respectively, are defined as
(9) |
(10) |
The fermionic algebra of the particle operators ( and ) and the quasiparticle operators ( and ) yields the following identities
(11) |
(12) |
The one-body density matrix and the pairing tensor can be expressed in terms of the quasiparticle components as
(13) |
(14) |
II.2 Coordinate-space formulation
The energy is generally further assumed to depend on and only through the following local densities and currents222The energy may be a functional of other densities and currents. We consider here only those relevant for calculating the entrainment couplings in homogeneous nuclear matter using the most popular functionals.:
(i) the nucleon number density at position and time
(15) |
(ii) the pair density (in general a complex number) at position and time
(16) |
(ii) the kinetic energy density (in units of ) at position and time
(17) |
(iii) and the momentum density (in units of ) at position and time
(18) |
where the particle and pair density matrices in coordinate space are respectively defined by Dobaczewski et al. (1984)
(19) |
(20) |
where and are the creation and destruction operators for nucleons of charge type at position with spin at time t. Introducing single-particle basis wavefunctions , these matrices can be alternatively written in terms of and as
(21) |
(22) |
Examples of nuclear energy-density functionals depending on the above local densities and currents are those constructed from zero-range effective nucleon-nucleon interactions of the Skyrme type Bender et al. (2003); Stone and Reinhard (2007). The present formalism is also applicable to more general classes of functionals, such as those proposed in Ref. Chamel et al. (2009). The pair density matrix is related to the order parameter of the superfluid phase at position and time as follows (see, e.g., Eq.(2.4.24) of Ref. Leggett (2006))
(23) |
The matrices (9) and (10) of the single-particle Hamiltonian and pair potential are given by
(24) | |||||
(25) | |||||
where
(26) |
(27) |
(28) |
The factor of in Eq. (28) arises from the antisymmetry of the pairing tensor (taking the derivative with respect to is equivalent to taking the derivative with respect to ). Because is a vector, it must obviously depend itself on . This may also be the case for all the other fields. For instance, the potential derived from the Brussels-Montreal functionals BSk19-26 Goriely et al. (2010, 2013) depends on , and .
Since the energy is real, it can only depend on the pair density through its square modulus . The pairing potential (28) can thus be written as
(29) |
The last equality shows that has the same phase as the order parameter . Using Eq. (22), the matrix elements (25) of the pairing field (29) will thus generally take the following form:
(30) |
with
(31) |
Let us note that . In the traditional formulation of the TDHFB equations (see, e.g., Ref. Blaizot and Ribka (1986)), represents the matrix elements of the effective (in-medium) two-body pairing interaction.
II.3 Mass currents and superfluid velocities
The TDHFB Eqs. (7) and (8) can be conveniently rewritten as Blaizot and Ribka (1986)
(32) |
(33) |
As shown in Appendix A, Eq. (32) can be translated in coordinate space by the same continuity equations as in the absence of pairing (a similar conclusion was previously drawn in Ref. Gusakov (2010) within Landau’s theory)
(34) |
where the nucleon mass current is given by Chamel and Allard (2019)
(35) |
The effective mass and the vector field are defined by Eq. (27). As shown in our previous work Chamel and Allard (2019), the nucleon mass current can be expressed solely in terms of the momentum densities as
(36) | |||||
where represents the nuclear-energy terms contributing to the mass currents, and we have introduced the following fields
(37) |
(38) |
Let us recall that the subscripts and denote isoscalar and isovector quantities, respectively, namely sums over neutrons and protons for the former (e.g. ) and differences between neutrons and protons for the latter (e.g. ).
III Entrainment effects in neutron-star cores
III.1 Exact solution of the TDHFB equations in homogeneous matter
In this section, we consider an homogeneous neutron-proton mixture with stationary nucleon currents in the normal rest frame (). The latter assumption ensures that the entropy densities are independent of time. Recalling that Blaizot and Ribka (1986)
(41) |
where denotes Boltzmann’s constant, the quasiparticle distribution functions are therefore also independent of time.
The single-particle wave functions are given by plane waves:
(42) |
where denotes the Pauli spinor. As can be seen from Eqs. (21), and (24), the density matrix (4) and the single-particle Hamiltonian matrix (9) are both diagonal in this basis.
The superfluid velocities, which are necessarily spatially uniform and independent of time, are conveniently written as
(43) |
This corresponds to a superfluid phase (modulo some arbitrary constant term without any physical consequence). Inserting Eq. (42) in Eq. (23) using Eqs. (16) and (22), the order parameter (23) thus reduces to
(44) |
One simple choice is to consider that is non-zero only if the states and have opposite spins and wave vectors and are such that . We can always arrange states such that and . For convenience, we introduce the following shorthand notation
(45) |
These quantum numbers define the conjugate states that are paired. Indeed, it follows from the definition (10) that the only nonzero matrix elements of the pair potential are of the form . In the absence of current (), the conjugate state is the time-reversed state of . The presence of a non-vanishing current () breaks the time-reversal symmetry. The nonzero elements of the and matrices satisfying Eqs. (11) and (12) are of the form and . Substituting by the quasiparticle energies , the TDHFB Eqs. (7) and (8) finally reduce to
(46) |
where ( denoting the eigenvalues of the single-particle Hamiltonian matrix) are explicitly given by
(47) |
The matrix elements are given by the following equation
(48) |
Using Eqs. (13), (15), (18), and (21), the nucleon number and momentum densities read
(49) |
(50) |
respectively, where the quasiparticle distribution is given by Blaizot and Ribka (1986)
(51) |
where .
The solutions of Eq. (46), readily obtained by diagonalizing the HFB matrix, are given by333Equation (46) actually admits two kinds of solutions corresponding to the eigenvalues . Those associated with are such that the expressions (53) and (54) of and are swapped. However, these solutions lead to the same values for the nucleon number density (49) and the momentum density (50) using the fact that . Therefore, they give the same mass current (36) and, consequently, the same entrainment matrix.
(52) |
(53) |
(54) |
where
(55) |
Using Eq. (47), we have
(56) |
(57) |
where we have introduced the effective superfluid velocity
(58) |
For standard Skyrme functionals, the effective mass depends only on the nucleon densities, whereas the potential depends on the pairing gaps (which in turn depend also on and ). For the extended Skyrme functionals proposed in Ref. Chamel et al. (2009), the potential depends also explicitly on , and . In general, the dependence of the energies (56) on the superfluid velocities may be therefore quite complicated.
Using Eqs. (31) and (42), it can be shown that the effective pairing interaction is independent of the wave vectors, and depends only on the spins. The only nonzero elements are given by (denoting the spins with arrows for clarity)
(59) |
and any other element obtained by permutation of the spin indices. It thus follows from Eq. (48), that is also independent of the wave vectors (but still depends on the given wave vectors ). Dropping the wave vector as a subscript, introducing the pairing gap
(60) |
and using Eqs. (53) and (54), Eq. (48) reduces to the gap equation
(61) |
The summation is only over the wave vectors , the summation over the spins has been already carried out. Note that the right-hand side of this equation explicitly depends on the wave vectors through Eqs. (51), (52), (56) and (57).
III.2 Entrainment matrix from the TDHFB solution
Using the solution of the TDHFB equations, the momentum density (50) can be alternatively written as
(62) |
The first term in the right-hand side coincides with the expression (29) adopted in our previous work Chamel and Allard (2019), thus demonstrating the validity of this identification. The second term accounts for quasiparticle excitations (note that the summation over spin states has been already carried out). Remarking that the component of orthogonal to does not contribute to the sum (since in this case), the momentum density can be expressed as
(63) |
with the function defined as
(64) |
Substituting Eq. (51) yields
(65) |
In terms of the effective superfluid velocity (58), the mass current (35) reduces to
(66) |
while the momentum density becomes
(67) |
Using this expression of combined with the general expression of the vector field Chamel and Allard (2019)
(68) |
leads to a self-consistent system of equations for and , whose solutions are
(69) |
(70) |
(71) |
(72) |
(73) |
(74) |
Substituting Eq. (69) into (58), the entrainment matrix can be readily obtained from Eq. (66):
(75) |
(76) |
(77) |
Using Eqs. (72) and (73), one can notice that the entrainment matrix is manifestly symmetric (i.e. ). Let us emphasize that Eqs. (75)-(77) give the exact expression for the entrainment matrix within the TDHFB theory in homogeneous nuclear matter. No approximation has been made at this point. In particular, the full dependence of the entrainment matrix elements on the currents is taken into account.
If the currents are small enough such that , the quasiparticle distributions (51) vanish at , hence also the functions , as can be seen from Eq. (64). In this regime, Eqs. (75)-(77) reduce to the expressions we derived earlier ignoring nuclear pairing within the time-dependent Hartree-Fock (TDHF) equations Chamel and Allard (2019), namely
(78) |
(79) |
(80) |
This allows us to rewrite Eq. (66) into the following alternative form
(81) |
III.3 Landau’s approximations
In previous studies of entrainment effects Gusakov and Haensel (2005); Leinson (2018), the mass current was defined as
(85) |
(, and were, respectively denoted by , and , in Ref. Gusakov and Haensel (2005), and by , and , in Ref. Leinson (2018)). Note that represents the group velocity of the single-particle state . Using Eqs. (47) and (49), Eq. (85) can be expressed as
(86) |
It can be seen from Eqs. (49) and (50) that Eq (III.3) coincides with our general expression (35) of the mass current. To compare our results to those obtained earlier within Landau’s theory of Fermi liquids, we assume that the superfluid velocities are small compared to the corresponding Fermi velocities , where denotes the Fermi wave number. We thus expand the quasiparticle energy (52) as follows:
(87) |
(88) |
where the single-particle energy is now evaluated in the static ground state ignoring any dependence on the pairing gaps, and expanding linearly around the Fermi surface
(89) |
which coincides with Eq. (6) of Ref. Leinson (2018). In this expression, the reduced chemical potential defined by has been replaced by the Fermi energy at , namely (in general, depends on the temperature, the gaps and the currents). The quasiparticle distribution function thus becomes
(90) |
Similarly, . In the continuum limit, replacing the discrete summation over by an integral, the function introduced in Eq. (64) thus reads
(91) |
where the first integration is over the solid angle in -space, is the angle between and , and is the level density444Note that denotes the level density per one spin state since summation over spins is already taken into account in Eq. (64). In the Landau’s theory, the level density is further assumed to be constant and approximated by its value on the Fermi surface, namely . Moreover, the term is also evaluated on the Fermi surface, and therefore is dropped. With all these approximations, the function finally reduces to Eq. (70) of Ref. Leinson (2018) (where was denoted by ):
(92) |
Note that a factor of 2 was absorbed by integrating over half the energy domain since the function to integrate is invariant under the change . Similar approximations for the gap equations (61) yield
(93) |
Let us remark that the gap equations are thus decoupled from the particle number conservation (49) since the reduced chemical potentials are approximated by the corresponding Fermi energies at . Further assuming , Eq. (93) can be expressed as Leinson (2018)
(94) |
where denotes the solution of Eq. (93) at in the absence of currents.
Substituting Eq. (92) in Eqs. (70)-(III.2), and evaluating all other quantities for the superfluids at rest, the entrainment matrix (75)-(77) can be recast in a form similar to Eq. (30) of Ref. Leinson (2018):
(95) |
(96) |
(97) |
(98) |
(99) |
where denotes the dimensionless Landau parameter (derivatives are calculated in the absence of currents)
(100) |
and the function is given by
(101) |
Simplifying the exact solution (75)-(77) using Landau’s approximations, we have thus recovered the entrainment matrix derived earlier in Refs. Leinson (2018); Gusakov and Haensel (2005). It should be stressed that the above formulas do not account for the full dependence of the entrainment matrix on the currents unlike Eqs. (75)-(77). In particular, the nonlinear effects contained in the single-particles energies have been ignored, compare Eqs. (56) and (89) recalling that the mean fields themselves have a highly nonlinear dependence on the currents.
IV Conclusions
We have studied the dynamics of nuclear superfluid systems at finite temperatures and finite currents in the framework of the self-consistent time-dependent nuclear-energy density functional theory. Considering the TDHFB equations in coordinate space, we have derived general expressions for the local nucleon mass currents (35) and local superfluid velocities (39), which are valid for both homogeneous systems (such as the outer core of neutron stars) and inhomogeneous systems (such as the crust of neutron stars, atomic nuclei and vortices). Remarkably, the mass currents are found to have the same formal expressions at any temperature, and coincide with the ones we derived earlier in the absence of pairing from the TDHF equations at zero temperature Chamel and Allard (2019).
Focusing on homogeneous neutron-proton superfluid mixtures, we have shown that the TDHFB equations can be solved exactly for arbitrary temperatures and currents. Using this solution, we have been able to express the Andreev-Bashkin mutual entrainment matrix in analytic form. The formal simplicity of our expression (75)-(77) are however deceptive: the entrainment coupling coefficients depend themselves on the currents in a very complicated way through the self-consistency of the TDHFB equations. We have explicitly demonstrated that our expression reduces to that obtained earlier in Ref. Leinson (2018) within Landau’s theory.
Our formulas are applicable to a large class of nuclear energy-density functionals. These include the Brussels-Montreal functionals based on generalized Skyrme effective interactions, for which unified equations of state for all regions of neutron stars have been calculated Potekhin et al. (2013); Mutafchieva et al. (2019); Pearson et al. (2018, 2020), thus paving the way for a fully consistent microscopic treatment of the dynamics of superfluid neutron stars. The formalism we have developed here in the nuclear context may also be easily transposed to the less exotic kinds of superfluids studied in terrestrial laboratories and described by similar TDHFB equations.
The relativistic entrainment matrix introduced in Ref. Gusakov and Andersson (2006) is given by Eqs. (82), (83), and (84). However, it would be worth carrying out the same analysis within a fully relativistic self-consistent microscopic framework, using the relativistic finite temperature HFB method Li et al. (2015).
Acknowledgements.
The authors thank T. Duguet and J. Sauls for valuable discussions. This work was financially supported by the Fonds de la Recherche Scientifique (Belgium) under Grant No. PDR T.004320. This work was also partially supported by the European Cooperation in Science and Technology action (EU) CA16214.Appendix A Continuity equations
Making use of the completeness relations
(102) |
and using Eqs. (24) and (25), we can demonstrate the following identities:
(103) |
(104) |
Multiplying Eq. (32) by and summing over indices and yields
(105) | ||||
(106) | ||||
(107) |
The summation in (105) yields the density matrix , as can be seen from Eq. (21). The next two summations in Eq. (106) can be simplified using Eq. (103) and the orthonormality property of the single-particle wavefunctions
(108) |
We can thus write
(109) |
Similarly, we have
(110) |
Let us recall that involves the operator so that it cannot be factored out of the integral. However, the hermiticity of the Hamiltonian matrix, , implies the following identity Chamel and Allard (2019):
(111) |
Finally, Eq. (110) becomes
(112) |
Likewise, the two summations in Eq. (107) can be expressed as follows using Eqs. (22) and (104):
(113) |
(114) |
Collecting terms, multiplying by , integrating over and summing over spins lead to
(115) |
As shown in Ref. Chamel and Allard (2019), the integral can be equivalently expressed as the divergence of a particle flux. The last two terms cancel each other if the pairing potential is calculated self-consistently from Eq. (29). Finally, multiplying by leads to the same Eq. (34) as previously derived in Ref. Chamel and Allard (2019) ignoring pairing.
References
- Blaschke and Chamel (2018) D. Blaschke and N. Chamel, in Astrophysics and Space Science Library, edited by L. Rezzolla, P. Pizzochero, D. I. Jones, N. Rea, and I. Vidaña (2018), vol. 457 of Astrophysics and Space Science Library, p. 337.
- Chamel (2017a) N. Chamel, Journal of Astrophysics and Astronomy 38, 43 (2017a).
- Andreev and Bashkin (1975) A. F. Andreev and E. P. Bashkin, Sov. Phys. JETP 42, 164 (1975).
- Glendenning (2000) N. K. Glendenning, Compact stars : nuclear physics, particle physics, and general relativity (Springer, 2000).
- Gusakov and Haensel (2005) M. E. Gusakov and P. Haensel, Nuclear Physics A 761, 333 (2005).
- Gusakov and Andersson (2006) M. E. Gusakov and N. Andersson, MNRAS 372, 1776 (2006).
- Andersson and Comer (2007) N. Andersson and G. L. Comer, Living Reviews in Relativity 10, 1 (2007).
- Graber et al. (2017) V. Graber, N. Andersson, and M. Hogg, International Journal of Modern Physics D 26, 1730015-347 (2017).
- Haskell and Sedrakian (2018) B. Haskell and A. Sedrakian, in Astrophysics and Space Science Library, edited by L. Rezzolla, P. Pizzochero, D. I. Jones, N. Rea, and I. Vidaña (2018), vol. 457 of Astrophysics and Space Science Library, p. 401.
- Gusakov and Kantor (2013) M. E. Gusakov and E. M. Kantor, MNRAS 428, L26 (2013).
- Dommes et al. (2019) V. A. Dommes, E. M. Kantor, and M. E. Gusakov, MNRAS 482, 2573 (2019).
- Kantor et al. (2020) E. M. Kantor, M. E. Gusakov, and V. A. Dommes, Phys. Rev. Lett. 125, 151101 (2020).
- Larkin and Migdal (1963) A. I. Larkin and A. B. Migdal, Sov. Phys. JETP-USSR 17, 1146 (1963).
- Leggett (1965) A. J. Leggett, Physical Review 140, 1869 (1965).
- Gusakov et al. (2009) M. E. Gusakov, E. M. Kantor, and P. Haensel, Phys. Rev. C 80, 015803 (2009).
- Glendenning (1985) N. K. Glendenning, ApJ 293, 470 (1985).
- Leinson (2018) L. B. Leinson, MNRAS 479, 3778 (2018).
- Chamel and Allard (2019) N. Chamel and V. Allard, Phys. Rev. C 100, 065801 (2019).
- Duguet (2014) T. Duguet, The Nuclear Energy Density Functional Formalism (2014), vol. 879, p. 293.
- Douchin and Haensel (2001) F. Douchin and P. Haensel, Astronomy& Astrophysics 380, 151 (2001).
- Potekhin et al. (2013) A. Y. Potekhin, A. F. Fantina, N. Chamel, J. M. Pearson, and S. Goriely, A&A 560, A48 (2013).
- Sharma et al. (2015) B. K. Sharma, M. Centelles, X. Viñas, M. Baldo, and G. F. Burgio, A&A 584, A103 (2015).
- Pearson et al. (2018) J. M. Pearson, N. Chamel, A. Y. Potekhin, A. F. Fantina, C. Ducoin, A. K. Dutta, and S. Goriely, MNRAS 481, 2994 (2018).
- Mutafchieva et al. (2019) Y. D. Mutafchieva, N. Chamel, Z. K. Stoyanov, J. M. Pearson, and L. M. Mihailov, Phys. Rev. C 99, 055805 (2019).
- Pearson et al. (2020) J. M. Pearson, N. Chamel, and A. Y. Potekhin, Phys. Rev. C 101, 015802 (2020).
- Mondal et al. (2020) C. Mondal, X. Viñas, M. Centelles, and J. N. De, Phys. Rev. C 102, 015802 (2020).
- Monrozeau et al. (2007) C. Monrozeau, J. Margueron, and N. Sandulescu, Phys. Rev. C 75, 065807 (2007).
- Chamel et al. (2010) N. Chamel, S. Goriely, J. M. Pearson, and M. Onsi, Phys. Rev. C 81, 045804 (2010).
- Pastore (2015) A. Pastore, Phys. Rev. C 91, 015809 (2015), eprint 1412.8410.
- Chamel (2017b) N. Chamel, Journal of Low Temperature Physics 189, 328 (2017b).
- Watanabe and Pethick (2017) G. Watanabe and C. J. Pethick, Phys. Rev. Lett. 119, 062701 (2017).
- Kashiwaba and Nakatsukasa (2019) Y. Kashiwaba and T. Nakatsukasa, Phys. Rev. C 100, 035804 (2019).
- Avogadro et al. (2007) P. Avogadro, F. Barranco, R. A. Broglia, and E. Vigezzi, Phys. Rev. C 75, 012805 (2007).
- Wlazłowski et al. (2016) G. Wlazłowski, K. Sekizawa, P. Magierski, A. Bulgac, and M. M. Forbes, Phys. Rev. Lett. 117, 232701 (2016), eprint 1606.04847.
- Bulgac (2019) A. Bulgac, Physica Status Solidi B Basic Research 256, 1800592 (2019).
- Blaizot and Ribka (1986) J. Blaizot and G. Ribka, Quantum Theory of Finite Systems (The MIT Press, 1986).
- Dobaczewski et al. (1984) J. Dobaczewski, H. Flocard, and J. Treiner, Nucl. Phys. A 422, 103 (1984).
- Bender et al. (2003) M. Bender, P.-H. Heenen, and P.-G. Reinhard, Reviews of Modern Physics 75, 121 (2003).
- Stone and Reinhard (2007) J. R. Stone and P. G. Reinhard, Progress in Particle and Nuclear Physics 58, 587 (2007).
- Chamel et al. (2009) N. Chamel, S. Goriely, and J. M. Pearson, Phys. Rev. C 80, 065804 (2009).
- Leggett (2006) A. Leggett, Quantum Liquids (Oxford University Press, 2006).
- Goriely et al. (2010) S. Goriely, N. Chamel, and J. M. Pearson, Phys. Rev. C 82, 035804 (2010).
- Goriely et al. (2013) S. Goriely, N. Chamel, and J. M. Pearson, Phys. Rev. C 88, 024308 (2013).
- Gusakov (2010) M. E. Gusakov, Phys. Rev. C 81, 025804 (2010).
- Li et al. (2015) J. J. Li, J. Margueron, W. H. Long, and N. Van Giai, Phys. Rev. C 92, 014302 (2015).