-
PDF
- Split View
-
Views
-
Cite
Cite
M. E. Gusakov, E. M. Kantor, A. I. Chugunov, L. Gualtieri, Dissipation in relativistic superfluid neutron stars, Monthly Notices of the Royal Astronomical Society, Volume 428, Issue 2, 11 January 2013, Pages 1518–1536, https://doi.org/10.1093/mnras/sts129
- Share Icon Share
Abstract
We analyse damping of oscillations of general relativistic superfluid neutron stars. To this aim we extend the method of decoupling of superfluid and normal oscillation modes first suggested in Gusakov & Kantor. All calculations are made self-consistently within the finite temperature superfluid hydrodynamics. The general analytic formulas are derived for damping times due to the shear and bulk viscosities. These formulas describe both normal and superfluid neutron stars and are valid for oscillation modes of arbitrary multipolarity. We show that (i) use of the ordinary one-fluid hydrodynamics is a good approximation, for most of the stellar temperatures, if one is interested in calculation of the damping times of normal f modes, (ii) for radial and p modes such an approximation is poor and (iii) the temperature dependence of damping times undergoes a set of rapid changes associated with resonance coupling of neighbouring oscillation modes. The latter effect can substantially accelerate viscous damping of normal modes in certain stages of neutron-star thermal evolution.
1 INTRODUCTION
Neutron stars (NSs) are compact objects with the mass M ∼ M⊙, circumferential radius R ∼ 10 km and the central density ρc several times higher than the nuclear density ρ0 ≈ 2.8 × 1014 g cm−3. They are interesting because of extreme conditions in their interiors and a wide variety of associated astrophysical phenomena. In particular, internal instabilities or external perturbations can excite NS oscillations, which are potentially detectable by the next-generation gravitational wave interferometers (see e.g. Andersson & Kokkotas 2001; Andersson 2003; Owen 2010). It is very probable that quasi-periodic oscillations of electromagnetic radiation observed in the tails of the giant gamma-ray flares are connected with oscillations in NS crust (e.g. Israel et al. 2005; Strohmayer & Watts 2005, 2006; Watts & Strohmayer 2007), and that seismology would become a significant source of information about NSs in the nearest future (Abbot et al. 2007; Andersson et al. 2011; Watts 2011).
For the correct interpretation of already existing and future observations one requires a well-developed theory of oscillating NSs. It should, in particular, (i) be based on the general relativity theory, since NSs are relativistic objects; (ii) employ an adequate model of superdense matter, including realistic equation of state and parameters of baryon superfluidity; and (iii) correctly account for the effects of baryon superfluidity on the hydrodynamics of NS matter.
Let us discuss briefly a (key) role of superfluidity. According to numerous microscopic calculations (see e.g. Lombardo & Schulze 2001), baryon matter in the internal layers of NSs becomes super-fluid at T ≲ 108–1010 K. It is very difficult to interpret the observa-tional data on pulsar glitches (see e.g. Chamel & Haensel 2008) and cooling of NSs (Yakovlev, Levenfish & Shibanov 1999; Yakovlev & Pethick 2004) without invoking baryon superfluidity. Recent real-time observations of cooling NS in Cassiopea A supernova remnant (Heinke & Ho 2010) also present a strong argument in favour of the existence of baryon superfluidity in the NS core. The observations were explained by Shternin et al. (2011) and Page et al. (2011) within a scenario, suggested for the first time in Gusakov et al. (2004) and Page et al. (2004), and assuming mild neutron superfluidity (with maximum neutron critical temperatures Tcn max ∼ 7-9 × 108 K) and strong proton superconductivity (with maximum proton critical temperatures Tcp max ≳ 2-3 × 109 K) in the NS core.
Combined analysis of all the three factors (i)–(iii) is a formidable task for the oscillation theory so in the literature they were considered successively. The foundations of the relativistic theory of stellar oscillations were laid 50 yr ago by Chandrasekhar (1964) and Thorne & Campolattaro (1967) and were further developed in many subsequent papers (see e.g. Detweiler & Ipser 1973; Ipser & Thorne 1973; Lindblom & Detweiler 1983; Detweiler & Lindblom 1985; Cutler & Lindblom 1987; Cutler, Lindblom & Splinter 1990; Chandrasekhar & Ferrari 1991; Kokkotas & Schutz 1992; Yoshida & Lee 2003a; Lin, Andersson & Comer 2008 and a review of Kokkotas & Schmidt 1999). When studying oscillations of NSs, most of these works considered ordinary one-fluid relativistic hydrodynamics.
Meanwhile it is well known that superfluidity leads to appearance of additional velocity fields, describing the superfluid degrees of freedom (e.g. Khalatnikov & Lebedev 1982; Khalatnikov 1989; Carter & Khalatnikov 1992). This substantially complicates the hydrodynamics of NS matter, making it multifluid (Mendell 1991a,b; Gusakov & Andersson 2006). In addition, superfluidity affects the kinetic coefficients (such as bulk and shear viscosities) and also requires additional viscous coefficients to be introduced (see Gusakov 2007; Gusakov & Kantor 2008 for details).
Oscillations of superfluid NSs have been studied actively only in the last two decades (see e.g. Lee 1995; Lindblom & Mendell 2000; Prix & Rieutord 2002; Yoshida & Lee 2003b; Prix, Comer & Andersson 2004; Samuelsson & Andersson 2009; Wong, Lin & Leung 2009; Passamonti & Andersson 2011, 2012), starting from the pioneering papers by Epstein (1988) and Lindblom & Mendell (1994). However, most of these works neglect general relativity effects and employ zero temperature (T = 0) limit of superfluid hydrodynamics (i.e. hydrodynamics, applicable only at T = 0). Within the general relativity theory oscillations were discussed only by Comer, Langlois & Lin (1999), Andersson, Comer & Langlois (2002), Yoshida & Lee (2003a), Gusakov & Andersson (2006), Lin et al. (2008), Kantor & Gusakov (2011) and Chugunov & Gusakov (2011), but most of these works used zero temperature hydrodynamics. Moreover, in some of these papers (e.g. Andersson et al. 2002; Lin et al. 2008) the presence of superfluid component was modelled by an artificial (polytropic) equation of state which does not represent any specific microphysical model.
Only in the recent papers (Gusakov & Andersson 2006; Chugunov & Gusakov 2011; Kantor & Gusakov 2011) an attempt was made to self-consistently calculate the oscillation spectra using a realistic model of superdense matter and allowing for the effects of finite stellar temperatures. It was shown that in many cases an approximation T = 0 is not justified and, moreover, it can lead to qualitatively incorrect results (Chugunov & Gusakov 2011; Kantor & Gusakov 2011).
Of particular interest is the question of how superfluidity influences dissipation of NS oscillations. It is of extreme importance, for instance, for understanding physical conditions under which a rotating NS becomes unstable with respect to excitation of various oscillations (e.g. r modes), and for estimating gravitational radiation from such stars (e.g. Andersson & Kokkotas 2001).
There were several serious and successful attempts to allow for the effects of superfluidity when studying the dissipation of oscillations in NSs (see e.g. Lindblom & Mendell 1995, 2000; Lee & Yoshida 2003; Andersson, Glampedakis &Haskell 2009; Haskell, Andersson & Passamonti 2009; Haskell & Andersson 2010; Passamonti & Glampedakis 2012), but all of them considered Newtonian stars and used the T = 0 superfluid hydrodynamics. The self-consistent analysis of dissipation in superfluid NSs was only recently performed for a simple case of a radially oscillating NS (Kantor & Gusakov 2011).
The aim of this paper is to fill this gap and to consider, for the first time, dissipation of non-radial oscillations in general relativistic superfluid NSs employing realistic microphysics input with accurate treatment of the effects of finite stellar temperatures.
The paper is organized as follows. Relativistic superfluid hydrodynamics is briefly reviewed in Section 2. Section 3 discusses an unperturbed star and introduces variables describing small deviations of NS from equilibrium. In Section 4 we derive expressions for the oscillation energy and its dissipation rates due to bulk and shear viscosities. In Section 5 the equations that govern oscillations of superfluid NSs are explicitly written out. Section 6 describes the approach to study dissipation of superfluid NS oscillations. This approach is applied for a detailed numerical analysis of realistic models of oscillating NSs in Section 7. Section 8 presents a summary of our results.
In what follows, we use the system of units in which c = kB = 1, where c is the speed of light and kB is the Boltzmann constant.
2 DISSIPATIVE SUPERFLUID HYDRODYNAMICS
In this paper we consider, for simplicity, npe matter in NS cores, which is matter composed of neutrons (n), protons (p) and electrons (e). Because both protons and neutrons can be in the superfluid state, one has to use the relativistic hydrodynamics of superfluid mixtures to study oscillations of NSs. Here we briefly discuss the corresponding equations to establish notations and to make the presentation more self-contained. Our consideration closely follows the papers by Gusakov & Andersson (2006), Gusakov (2007) and, especially, Kantor & Gusakov (2011). The reader is referred to these works for more details.
Together with the quasi-neutrality condition (ne = np) and equation (3), the equations of superfluid hydrodynamics include (Gusakov 2007) the following.
- Continuity equations for baryons (b) and electrons (e):(8)\begin{eqnarray} j^{\mu }_{ ({\rm b}) ; \, \mu } &=& 0, \end{eqnarray}(9)\begin{eqnarray} j^{\mu }_{({\rm e}) ; \, \mu } &=& 0. \end{eqnarray}
- Energy–momentum conservation:(10)\begin{eqnarray} T^{\mu \nu }_{; \,\mu } &=& 0, \end{eqnarray}(11)\begin{eqnarray} T^{\mu \nu } &=& (P+\varepsilon ) \, u^{\mu } u^{\nu } + P g^{\mu \nu } + Y_{ik} \left( w^{\mu }_{(i)} w^{\nu }_{(k)} + \mu _i \, w^{\mu }_{(k)} u^{\nu }\right.\nonumber\\[5pt] &&\left. + \mu _k \, w^{\nu }_{(i)} u^{\mu } \right)+\tau ^{\mu \nu }, \end{eqnarray}(12)\begin{eqnarray} \tau ^{\mu \nu } &=& - \eta \, H^{\mu \gamma } \, H^{\nu \delta } \,\, \left( u_{\gamma ; \delta } + u_{\delta ; \gamma } - {2 \over 3} \,\, g_{\gamma \delta } \,\, u^{\varepsilon }_{;\varepsilon } \right)\nonumber\\ && - \xi _{1 {\rm n}} \, H^{\mu \nu } \, \left[ Y_{{\rm n} k} w^{\gamma }_{(k)} \right]_{; \gamma } - \xi _2 \, H^{\mu \nu } \, u^{\gamma }_{;\gamma }. \end{eqnarray}
- Potentiality condition for superfluid motion of neutrons:(13)\begin{eqnarray} \mathrm{\partial} _{\nu } \left[ w_{({\rm n}) \mu } + (\mu _{\rm n} + \varkappa _{\rm n}) u_{\mu } \right] &=& \mathrm{\partial} _{\mu } \left[ w_{({\rm n}) \nu } +(\mu _{\rm n} + \varkappa _{\rm n}) u_{\nu } \right], \end{eqnarray}(iv) The second law of thermodynamics:(14)\begin{eqnarray} \varkappa _{\rm n} &=& - \xi _{3{{\rm n}}} \, \left[ Y_{{\rm n}k} w^{\mu }_{(k)} \right]_{;\mu } - \xi _{4{\rm n}} \, u^{\mu }_{;\mu }. \end{eqnarray}In formulas (8)–(15) gμν is the metric tensor; Hμν ≡ gμν + uμuν; ∂μ ≡ ∂/(∂xμ); P, ε, S and μe are the pressure, energy density, entropy density and relativistic electron chemical potential, respectively. These quantities are related by the formula:(15)\begin{equation} {\rm d} \varepsilon = T \, {\rm d} S + \mu _{\rm e} \, {\rm d} n_{\rm e} + \mu _i \, {\rm d} n_i + \frac{Y_{ik}}{2} \, {\rm d} \left[w^{\alpha }_{(i)} w_{(k) \alpha } \right]. \end{equation}Finally, η is the shear viscosity coefficient and ξ1n, ξ2, ξ3n and ξ4n are the bulk viscosity coefficients. Because of the Onsager symmetry principle, one has(16)\begin{equation} P=-\varepsilon + \mu _{\rm e} n_{\rm e} + \mu _i n_i + {\it TS}. \end{equation}Moreover, if the bulk viscosities are generated solely by the direct or modified URCA processes, one has an additional constraint (Gusakov 2007)(17)\begin{equation} \xi _{1{\rm n}}=\xi _{4{\rm n}}. \end{equation}In the absence of superfluidity the only non-zero coefficient is ξ2 – the ordinary bulk viscosity.(18)\begin{equation} \xi _{1{\rm n}}^2=\xi _2 \xi _{3{\rm n}}. \end{equation}
3 BASIC EQUATIONS
3.1 An unperturbed star
An equilibrium configuration of a non-rotating superfluid NS was analysed in detail in section 3 of Gusakov & Andersson (2006). Here we present only the main results of this analysis, which will be used in what follows.
3.2 Small departures from equilibrium
4 DAMPING OF OSCILLATIONS DUE TO THE BULK AND SHEAR VISCOSITIES: GENERAL FORMULAS
4.1 Mechanical energy
In the absence of superfluidity Wsfl = Vsfl = 0, Wb = W and Vb = V. In that case equation (72) gives a mechanical energy of a non-superfluid star that coincides, up to notations, with the corresponding expression (29) of Thorne & Campolattaro (1967).
4.2 Dissipation rates
The damping time τgrav due to radiation of gravitational waves can be obtained from the equations, describing linear oscillations of NSs (see Section 5). The goal of the present section is to determine the dissipation rate of oscillation energy due to the bulk |$\mathfrak {W}_{\rm bulk}$| and shear |$\mathfrak {W}_{\rm shear}$| viscosities and, as a consequence, the damping times τbulk and τshear.
5 OSCILLATION EQUATIONS
In order to calculate τbulk and τshear one has to determine the oscillation eigenfrequencies σ and eigenfunctions H0, H1, H2, K, Wb, Vb, Wsfl and Vsfl. To do that one needs to formulate oscillation equations. Since the dissipation is weak, when deriving the oscillation equations one can neglect the dissipative terms in the superfluid hydrodynamics of Section 2 and put τμν = 0 and ϰn = 0.
As it was shown in Gusakov & Kantor (2011), equations describing small linear oscillations of an NS include the following.
- Continuity equations for baryons (8) and electrons (9) that can be written in terms of the baryon and electron number density perturbations, δnb and δne, as(83)\begin{eqnarray} \delta n_{\rm b} &=& \frac{{\rm i}}{\omega \, {\rm e}^{-\nu /2}} \left[ \mathrm{\partial} _j (n_{\rm b}) \, U_{\rm (b)}^j + n_{\rm b} \, U^\mu _{\rm (b)\, ; \mu } \right], \end{eqnarray}where j is the spatial index and we defined(84)\begin{eqnarray} \delta n_{\rm e} &=& \delta n_{\rm e \, ({\rm norm})} + \delta n_{\rm e \, ({\rm sfl})}, \end{eqnarray}(85)\begin{eqnarray} \delta n_{\rm e \, ({\rm norm})} &\equiv & \frac{{\rm i}}{\omega \, {\rm e}^{-\nu /2}} \left[ \mathrm{\partial} _j (n_{\rm e}) \, U_{\rm (b)}^j + n_{\rm e} \, U^\mu _{\rm (b)\, ; \mu } \right], \end{eqnarray}(86)\begin{eqnarray} \delta n_{\rm e \, ({\rm sfl})} &\equiv & -\frac{{\rm i}}{\omega \, {\rm e}^{-\nu /2}} \left[ \mathrm{\partial} _j (n_{\rm e}) \, X^j + n_{\rm e} \, X^\mu _{; \mu } \right]. \end{eqnarray}
- Einstein equations, which can schematically be presented aswhere the perturbation δTμν of the energy–momentum tensor (11) can be expressed in terms of the perturbations of baryon four-velocity δUμ(b), metric δgμν, pressure δP and energy density δε as(87)\begin{equation} \delta (R^{\mu \nu } - 1/2 \,\, g^{\mu \nu } \, R) =8 \pi G\, \, \delta T^{\mu \nu }, \end{equation}In equation (87) Rμν and R are the Ricci tensor and scalar curvature, respectively, and G is the gravitation constant.(88)\begin{eqnarray} \delta T^{\mu \nu } &=&\displaystyle (\delta P + \delta \varepsilon ) \, U^{\mu }_{({\rm b})} U^{\nu }_{({\rm b})} + (P + \varepsilon ) \left[U^{\mu }_{({\rm b})} \, \delta U^{\nu }_{({\rm b})}+ U^{\nu }_{({\rm b})} \, \delta U^{\mu }_{({\rm b})} \right]\nonumber\\ && +\; \delta P \, g^{\mu \nu } + P \, \delta g^{\mu \nu }. \end{eqnarray}
- ‘Superfluid’ equation that can be derived from equations (10) and (13) of Section 2 (here we present only the spatial components j of this equation)6Expressing the vectors wj(i) through Xj in this equation (see equations 3 and 4), and introducing the redshifted imbalance of chemical potentials δμ∞ ≡ eν/2 δμ, one can rewrite equation (89) as(89)\begin{equation} {\rm i} \, \omega \, (\mu _{\rm n} \, Y_{{\rm n}k} \, w_{(k)j} - n_{\rm b} \, w_{({\rm n})j}) = n_{\rm e} \,\, \mathrm{\partial} _j ({\rm e}^{\nu /2} \,\delta \mu ). \end{equation}where y is defined by equation (68). Note that this equation dictates the most general form of the superfluid Lagrangian displacement ξj(sfl) that was already obtained in equation (52) from the symmetry arguments.(90)\begin{equation} X_j = \frac{{\rm i} \, n_{\rm e}}{\mu _{\rm n} n_{\rm b} \, \omega \, y} \,\mathrm{\partial} _j (\delta \mu ^\infty ), \end{equation}
When deriving equation (93) we took into account that [∂ε(nb, ne)/∂ne] δne = −δμ δne is a quadratically small quantity, because δμ = 0 in equilibrium.7
Finally, let us mention one important property that follows from the oscillation equations and quasi-neutrality condition (3). If neutrons in a non-rotating non-magnetized star are normal (i.e. Ynn = Ynp = 0), while protons are superfluid (Ypp ≠ 0), then oscillation eigenfrequencies and eigenfunctions for such star will be indistinguishable from that for a normal star of the same mass (where both protons and neutrons are non-superfluid).
6 OUR APPROACH
6.1 Decoupling of superfluid and normal modes
In principle, equations (83)–(101) allow one to study the non-radial oscillations of superfluid NSs and thus to determine the spectrum of eigenfrequencies ω, eigenfunctions H0, H1,…, Wsfl and Vsfl, and hence the damping times τgrav, τbulk and τshear. However, this task can be significantly simplified, if one notes that the dimensionless coupling parameter s (94) is small for realistic equations of state of superdense matter (Gusakov & Kantor 2011). For example, for the equation of state APR (Akmal, Pandharipande & Ravenhall 1998) employed below s ∼ 0.01-0.05. This means that one can look for the solution to the system of equations (83)–(101) in the form of a series in s. Since s is small, the approximation s = 0 is already quite accurate. Indeed, as it was shown in Gusakov & Kantor (2011) with the example of radial oscillations, the eigenfrequencies calculated in this approximation differ from the exact ones, on average, by ∼ 1.5-2 per cent. Thus, in what follows all calculations are performed assuming s = 0.
How this approach simplifies the problem? As it was first demonstrated in Gusakov & Kantor (2011), in the s = 0 approximation superfluid degrees of freedom (vectors Xj) completely decouple from the ‘normal’ degrees of freedom [metric perturbations δgμν and baryon four-velocities δUμ(b)]. That is, one has two distinct classes of oscillations: ‘superfluid’ and ‘normal’ modes, which are described by independent equations. For superfluid-type oscillations the metric and baryon velocity are not perturbed [δgμν = 0 and δUμ(b) = 0]; hence these modes do not emit gravitational waves; moreover, they are entirely localized in the SFL-region. At the same time, the frequencies of normal modes are indistinguishable from those of a normal (non-superfluid) star of the same mass.8 Below we discuss in more detail decoupling of superfluid and normal oscillation modes and how this property can be used to calculate the characteristic damping times.
6.2 A strategy to calculate the damping times
- A star oscillates at a frequency which is not an eigenfrequency of the Einstein equations (87). In that case, to satisfy equation (87), one has to demandFrom equation (101) it follows then that δμnorm l = 0 and the superfluid equation (98) decouples from the Einstein equations. As a result we arrive at the ‘source-free’ equation (with the right-hand side vanished), first derived in Chugunov & Gusakov (2011),10(103)\begin{equation} H_0=H_1=H_2=K=W_{\rm b}=V_{\rm b}=0. \end{equation}This equation describes superfluid modes and should be solved in the stellar region where neutrons are superfluid (SFL-region). It should be supplemented with a number of boundary conditions, discussed in Chugunov & Gusakov (2011) (similar, but more general boundary conditions for equation 98 are presented in Appendix A). Having solved equation (104) for δμl, it is easy to determine the functions Wsfl and Vsfl using equations (48), (52) and (90). Using Wsfl and Vsfl, one can find the functions W and V from equations (53), (54) and (103),(104)\begin{eqnarray} \delta \mu _l^{\prime \prime }&+&\left(\frac{h^\prime }{h} -\frac{\lambda ^\prime }{2} +\frac{2}{r}\right)\delta \mu _l^\prime\nonumber\\ & -&{\rm e}^\lambda \left[\frac{l(l+1)}{r^2}+\mathrm e^{-\nu /2}\frac{\omega ^2}{h \, \mathfrak {B}} \right] \delta \mu _l =0. \end{eqnarray}This information is sufficient to calculate τbulk and τshear from equations (77) and (78) (as follows from equation 103, τgrav = ∞ for superfluid modes in the s = 0 approximation).(105)\begin{equation} W=-W_{\rm sfl}, \quad V=-V_{\rm sfl}. \end{equation}
A star oscillates at a frequency which is an eigenfrequency of Einstein equations (87). In that case, the eigenfrequency and eigenfunctions H0, H1, H2, K, Wb and Vb are indistinguishable from the corresponding eigenfrequency and eigenfunctions for an oscillating non-superfluid NS (we recall that for the non-superfluid star Wb = W, Vb = V, because Wsfl = Vsfl = 0, see equations 53 and 54). There is, however, one very important difference: for a superfluid star the functions Wsfl and Vsfldo not vanish in the SFL-region and are comparable there to Wb and Vb. As follows from equations (77) and (78), the damping times τbulk and τshear depend on these functions (as well as on W = Wb − Wsfl and V = Vb − Vsfl) that is why the determination of Wsfl and Vsfl is a necessary task.
To determine these functions we make use of equation (98). Since the oscillation frequency ω = σ + i/τgrav and the eigenfunctions H0, H1, H2, K, Wb and Vb are already known, we can, using equation (101), calculate δμnorm l and determine a ‘source’ in the right-hand side of equation (98). This source plays a role of an external driving force that makes the superfluid equation (98) ‘oscillate’ at the frequency ω, which is not an eigenfrequency for this equation.11 As a result, the function δμl(r) will be non-zero. To determine it one has to specify the boundary conditions for equation (98); they are formulated in Appendix A. Having solved equation (98) numerically and having defined δμl(r), one can calculate the functions Wsfl and Vsfl, using equations (48), (52), (90) and (97).
Summarizing, in the approximation s = 0 the eigenfrequencies and eigenfunctions H0, H1, H2, K, Wb and Vb (and hence τgrav) for the normal modes appear to be the same as for a non-superfluid star. At the same time the eigenfunctions Wsfl and Vsfl are non-zero in the SFL-region and should be determined from equation (98). As a result, the damping times τbulk and τshear, defined by equations (77) and (78), will differ from the corresponding times, calculated using the ordinary (non-superfluid) hydrodynamics (even if one takes into account the effects of superfluidity on the kinetic coefficients).
7 RESULTS
Let us apply the approach, suggested in the previous section, to determine the frequency spectrum and damping times for an oscillating superfluid NS. But first let us discuss its equilibrium model.
7.1 Microphysics input and equilibrium model
As mentioned in Section 2, we consider the simplest npe-composition of NS core. We adopt APR equation of state (Akmal et al. 1998) parametrized by Heiselberg & Hjorth-Jensen (1999) in the core and the equation of state by Negele & Vautherin (1973) in the crust.
All numerical results presented here are obtained for an NS with the mass M = 1.4 M⊙. The circumferential radius for such star is R = 12.2 km, and the central density is ρc = 9.26 × 1014 g cm−3. The crust–core interface lies at the distance of Rcc = 10.9 km from the centre.
When modelling the effects of superfluidity we assume the triplet pairing of neutrons and singlet pairing of protons in the NS core. The neutron superfluidity in the stellar crust is neglected; it should not affect strongly the global oscillations of NSs.
We consider two models of nucleon superfluidity: model ‘1’ (simplified) and model ‘2’ (more realistic). In the model 1 the redshifted proton critical temperature is constant over the core, T∞cp ≡ Tcp eν/2 = 5 × 109 K; the redshifted neutron critical temperature T∞cn ≡ Tcn eν/2 increases with the density ρ and reaches the maximum value T∞cn max = 6 × 108 K at the stellar centre (r =0). This model corresponds to the model 3 of Kantor & Gusakov (2011).
In the model 2 both critical temperatures Tcn and Tcp are density dependent. This model does not contradict the results of microscopic calculations (see e.g. Yakovlev et al. 1999; Lombardo & Schulze 2001) and is similar to the nucleon pairing models used to explain observations of the cooling NS in Cassiopea A supernova remnant (Shternin et al. 2011).
The models 1 and 2 are shown in Figs 1 and 2, respectively.12 The function Tci(ρ) in both figures is shown in the left-hand panels, while the right-hand panels demonstrate the dependence T∞ci(r) (i = n and p). With the decrease of the redshifted temperature T∞ the size of the SFL-region [given by the condition T < Tcn(r) or, equivalently, T∞ < T∞cn(r)] increases or remains unchanged. For instance, the SFL-region corresponding to T∞ = 4 × 108 K is shaded in Figs 1 and 2. One can see that for the model 2 there can be three-layer configurations of a star with no neutron superfluidity in the centre and in the outer region but with superfluid intermediate region. On the contrary, in the model 1 only two-layer configurations are possible.

Left-hand panel: nucleon critical temperatures Tck (k = n, p) versus density ρ for model 1. Right-hand panel: redshifted critical temperatures T∞ck versus radial coordinate r (in units of R) for model 1.

The entrainment matrix Yik is calculated for the superfluidity models 1 and 2 in a way similar to how it was done in Kantor & Gusakov (2011).
When analysing viscous dissipation in oscillating NSs we allow for the damping due to shear and bulk viscosities. For the shear viscosity coefficient η we take the electron shear viscosity ηe, calculated in Shternin & Yakovlev (2008). We neglect the nucleon shear viscosity because (i) it is poorly known even for non-superfluid matter and (ii) it appears to be less than the electron shear viscosity in the core at T ≪ Tcp (Shternin & Yakovlev 2008).
The bulk viscosity coefficients are calculated as described by Gusakov (2007), Gusakov & Kantor (2008) and Kantor & Gusakov (2011). Since the direct URCA process is closed for our stellar model with M = 1.4 M⊙, the main contributor to the bulk viscosity is the modified URCA process.
7.2 Oscillations of a non-superfluid star
As follows from Section 6.2, before considering oscillations of a superfluid NS one should study those of a normal (non-superfluid) star of the same mass. To this aim, we have determined the eigenfrequencies and eigenfunctions of the radial and non-radial oscillation modes for a non-superfluid NS of mass M = 1.4 M⊙ and equation of state APR (see Section 7.1). We have solved the equations describing radial and non-radial perturbations of a non-rotating star in general relativity. These equations are derived by expanding the perturbed Einstein's equations in tensorial spherical harmonics in an appropriate gauge, and are integrated in the frequency domain.
Stellar modes are defined as solutions of the perturbed equations which are regular at the centre and with vanishing Lagrangian pressure perturbation at the surface, and (if l > 1) which behave as a pure outgoing wave at infinity; as discussed above, such solutions have complex frequencies ω = σ + i/τ. If l ≤ 1, instead, the frequency is real and the mode is not associated with gravitational emission.
The oscillation modes are classified according to the source of the restoring force which prevails in bringing the perturbed element of fluid back to the equilibrium position; for instance, we have a g mode if the restoring force is mainly provided by buoyancy, a p mode if it is due to a gradient of pressure and so on.
The radial modes are calculated as described in Gusakov et al. (2005). To calculate the non-radial modes we follow the formulation of Lindblom & Detweiler (1983) and Detweiler & Lindblom (1985). In their formulation, the equations for non-radial perturbations can be expressed, inside the star, as a system of first-order differential equations in the variables H0, H1, H2, K, Wb and Vb defined in Section 5. Outside the star, they reduce to a simple, second-order differential equation (the Zerilli equation). By numerical integration of these equations (the procedure we have followed is described in detail e.g. in Burgio et al. 2011) we find, for each value of the multipolarity l, the (complex) eigenfrequencies ω and the corresponding eigenfunctions H0(r), H1(r), H2(r), K(r), Wb(r) and Vb(r). The results of our computations are summarized in Table 1 and illustrated in Figs 3 and 4.

The function δμnorm l (in units of 107 K) versus r for fundamental radial F mode as well as for p1 and f modes with multipolarities l = 1, 2 and 3 (see the footnote 13). The energy of each oscillation mode is 1043 erg. Shaded region corresponds to crust, where δμnorm l is not defined and was not plotted.

Damping times τb + s ≡ (τ− 1bulk + τ− 1shear)− 1 versus T∞ for various oscillation modes. The effects of superfluidity are partially taken into account, as described in the text. Thick and thin solid lines correspond to radial (l = 0) F and 1H modes, respectively; dot–dashed line corresponds to dipole (l = 1) p1 mode; thick and thin dashes correspond to quadrupole (l = 2) f and p1 modes, respectively; and thick and thin dots correspond to octupole (l = 3) f and p1 modes, respectively.
Frequency σ (in units of 104 s−1 and in units of |$\tilde{\sigma }=c/R \approx 2.46 \times 10^4$| s−1) and the damping time τgrav (in seconds) for various oscillation modes of a non-superfluid NS. The first column shows the multipolarity l of modes and their names.
l, mode . | σ/(104 s−1) . | |$\sigma /\tilde{\sigma }$| . | τgrav (s) . |
---|---|---|---|
0, F | 1.703 | 0.691 | ∞ |
0, 1H | 4.080 | 1.656 | ∞ |
0, 2H | 5.732 | 2.327 | ∞ |
1, p1 | 2.893 | 1.175 | ∞ |
2, f | 1.155 | 0.469 | 0.212 |
2, p1 | 3.720 | 1.510 | 3.799 |
3, f | 1.554 | 0.631 | 18.24 |
3, p1 | 4.360 | 1.770 | 33.26 |
l, mode . | σ/(104 s−1) . | |$\sigma /\tilde{\sigma }$| . | τgrav (s) . |
---|---|---|---|
0, F | 1.703 | 0.691 | ∞ |
0, 1H | 4.080 | 1.656 | ∞ |
0, 2H | 5.732 | 2.327 | ∞ |
1, p1 | 2.893 | 1.175 | ∞ |
2, f | 1.155 | 0.469 | 0.212 |
2, p1 | 3.720 | 1.510 | 3.799 |
3, f | 1.554 | 0.631 | 18.24 |
3, p1 | 4.360 | 1.770 | 33.26 |
Frequency σ (in units of 104 s−1 and in units of |$\tilde{\sigma }=c/R \approx 2.46 \times 10^4$| s−1) and the damping time τgrav (in seconds) for various oscillation modes of a non-superfluid NS. The first column shows the multipolarity l of modes and their names.
l, mode . | σ/(104 s−1) . | |$\sigma /\tilde{\sigma }$| . | τgrav (s) . |
---|---|---|---|
0, F | 1.703 | 0.691 | ∞ |
0, 1H | 4.080 | 1.656 | ∞ |
0, 2H | 5.732 | 2.327 | ∞ |
1, p1 | 2.893 | 1.175 | ∞ |
2, f | 1.155 | 0.469 | 0.212 |
2, p1 | 3.720 | 1.510 | 3.799 |
3, f | 1.554 | 0.631 | 18.24 |
3, p1 | 4.360 | 1.770 | 33.26 |
l, mode . | σ/(104 s−1) . | |$\sigma /\tilde{\sigma }$| . | τgrav (s) . |
---|---|---|---|
0, F | 1.703 | 0.691 | ∞ |
0, 1H | 4.080 | 1.656 | ∞ |
0, 2H | 5.732 | 2.327 | ∞ |
1, p1 | 2.893 | 1.175 | ∞ |
2, f | 1.155 | 0.469 | 0.212 |
2, p1 | 3.720 | 1.510 | 3.799 |
3, f | 1.554 | 0.631 | 18.24 |
3, p1 | 4.360 | 1.770 | 33.26 |
Table 1 presents the real parts of the eigenfrequencies Re(ω) = σ (measured in units of 104 s−1 and in units of |$\tilde{\sigma } \equiv c/R \approx 2.46\times 10^4$| s−1) and the characteristic gravitational damping times τgrav (in seconds) for the modes with l = 0 (fundamental F mode and first two overtones 1H and 2H), l = 1 (dipole p1 mode), l = 2 (quadrupole f and p1 modes) and l = 3 (octupole f and p1 modes).13 One can see that σ ≫ 1/τgrav in all these cases. That is, damping due to emission of gravitational waves occurs on a time-scale much longer than the oscillation period.
Using definition (99) and equation (101) we have determined, in terms of the eigenfunctions H0(r), H1(r),…,Vb(r), the function δμnorm l(r) and, consequently, the quantity δμ∞norm(r, θ) = δμnorm l(r) Y0l(θ) for each mode. As follows from equations (92) and (100), for a non-superfluid star δμ∞norm(r, θ) is simply a redshifted imbalance of chemical potentials, δμ∞ = δμ∞norm. The function δμnorm l(r), entering equation (98), is shown in Fig. 3 for the oscillation modes from Table 1. It is normalized such that the mechanical energy of oscillations is 1043 erg. The shaded region corresponds to the crust of the star, where δμnorm l(r) is not defined (protons are bound in nuclei there). As seen in the figure, |δμnorm l(r)| for f modes is about one order of magnitude smaller than for p modes. This is not surprising, since matter is only weakly compressed during f mode oscillations, so that a deviation from beta-equilibrium (when δμ∞ = 0) is small. The functions δμnorm l(r) are employed to calculate the damping times of a superfluid NS in Section 7.4.
Fig. 4 shows the viscous damping time τb + s ≡ (τ− 1bulk + τ− 1shear)− 1 as a function of T∞ for a set of oscillation modes. The solid lines correspond to radial (l = 0) modes F and 1H, dot–dashed line corresponds to dipole (l = 1) mode p1, dashed lines correspond to quadrupole (l = 2) modes f and p1, and dotted lines correspond to octupole (l = 3) modes f and p1. To calculate τb+s we used the formulas for τbulk and τshear, applicable for the ordinary hydrodynamics of a non-superfluid liquid.14 However, we allow for the effects of superfluidity when calculating the kinetic coefficients η and ξ2 (the other bulk viscous coefficients do not appear in the normal-fluid hydrodynamics). To calculate η and ξ2 we adopt the nucleon superfluidity model 2 (see Section 7.1). Such an approximate approach to accounting for the effects of superfluidity is commonly used in the literature, but it is not fully consistent. The results of a more consistent approach (see Section 6) are discussed in Section 7.4.
As follows from Fig. 4, the dependence of τb+s on T∞ is a power law at T∞ ≲ 6 × 108 K. At such T∞ the proton superfluidity is ‘strong’ (T∞ ≪ T∞cp). In that case the bulk viscosity is exponentially suppressed (Haensel, Levenfish & Yakovlev 2001), while the shear viscosity η ∝ 1/(T∞)2 (Shternin & Yakovlev 2008) and dominates. As a result, τb+s ∝ (T∞)2. At high enough temperatures T∞ ≳ 6 × 108 K the damping due to the bulk viscosity starts to prevail; this results in decreasing of τb+s with growing T∞ (the curves in Fig. 4 bend down). At such T∞ the neutrons are normal and the proton superfluidity is weak or absent. Neglecting the proton superfluidity, one obtains ξ2 ∝ (T∞)6 (Haensel et al. 2001), hence τb+s ∝ 1/(T∞)6.
Let us note that the curves for f modes in Fig. 4 (thick dashed line and thick dots) bend down later than others; for them the shear viscosity is the dominant mechanism of damping up to T∞ ≈ 2.0 × 109 K. This is not surprising, since, as it was noted above, for f modes the deviation from beta-equilibrium is small (δμ∞ is reduced by an order of magnitude in comparison to p modes, see Fig. 3); hence damping due to the bulk viscosity is suppressed (the relation between δμ∞ and τbulk was discussed in detail e.g. in Gusakov et al. 2005). As a result, τb+s approaches its ‘bulk viscosity’ asymptote τb+s ∝ 1/T6 at higher temperatures T∞ > 2.0 × 109 K.
7.3 Frequency spectrum for superfluid NSs
First of all let us consider the frequency spectrum for radial oscillations of a superfluid NS employing the simplified model 1 of nucleon superfluidity. For such model this problem was discussed in detail by Kantor & Gusakov (2011), where it was solved exactly. Here we compare this exact solution with the approximate calculations obtained in the s = 0 approximation (see Section 6). Such a comparison is very useful, since it allows one to make a conclusion about applicability of the approximate approach in the case of non-radial oscillations, where the exact solution is not attempted.
The eigenfrequencies σ of radial pulsations (in units of |$\tilde{\sigma }$|) versus T∞8 = T∞/(108 K) are shown in Figs 5(a)–(c). In Fig. 5(a) this dependence was obtained assuming that superfluid and normal modes are completely decoupled (s = 0 approximation). The thick solid lines demonstrate the first three normal (non-superfluid) radial modes F, 1H and 2H. As one expects, their frequencies do not depend on T∞. The dashes are for the first six superfluid modes 1, …, 6, which are the solutions to equation (104). These modes, on the contrary, strongly depend on T∞ and approach their temperature-independent asymptotes only at T∞ ≲ 5 × 107 K (when the entire NS core is superfluid and Yik does not depend on T∞). At T∞ > T∞cn max = 6 × 108 K all neutrons are normal so that superfluid modes do not exist.

Eigenfrequencies σ (in units of |$\tilde{\sigma }$|) of radial oscillations versus T∞8 = T∞/(108 K) for model 1 of nucleon superfluidity. (a) Approximate spectrum. First three normal modes (F, 1H and 2H) are shown by the solid lines; first six superfluid modes 1, …, 6 are shown by dashes. (b) Exact spectrum. Alternate solid and dashed lines show the first six exact modes (I, …, VI) of a radially oscillating star. No spectrum was plotted in the shaded region. (c) Approximate (dashed lines) and exact (solid lines) spectra. At T∞ > T∞cn max = 6 × 108 K all neutrons are normal and the spectrum is that of a non-superfluid star.
Fig. 5(b) demonstrates the results of the exact solution to equations (83)–(101) obtained by Kantor & Gusakov (2011) for radial oscillations of a superfluid NS. The frequencies σ of the first six oscillation modes (I,…,VI) as functions of T∞8 are shown by alternate solid and dashed lines. No spectrum is plotted in the grey-shaded area. One can observe that the approximate spectrum (Fig. 5 a) is very similar to the exact spectrum (Fig. 5 b). However, there is one important difference: instead of crossings of superfluid and normal modes in Fig. 5(a) we have avoided crossings of the modes in Fig. 5(b). At these points the superfluid mode turns into the normal one and vice versa. As it was discussed in details in Gusakov & Kantor (2011) and Kantor & Gusakov (2011), this is not surprising, since in a vicinity of avoided crossings the Einstein equations (87) and superfluid equation (98) interact resonantly, so that approximation of completely decoupled superfluid and normal modes (s = 0) is inapplicable.15
For comparison, in Fig. 5(c) we plot both the approximate (dashed lines) and exact (solid lines) spectra. The agreement between both spectra is very good: the difference is less than a few per cent.
Such a close agreement of the exact and approximate results for radial oscillations allows us to analyse the spectrum of non-radial oscillations using the same approximation s = 0. The results of this analysis are shown in Fig. 6 for more realistic model 2 of nucleon superfluidity (see Section 7.1 and Fig. 2). Superfluid modes shown in this figure have been already studied in detail in our recent paper (Chugunov & Gusakov 2011). Thus, here we discuss them only briefly.

Eigenfrequencies σ versus T∞ for model 2 of nucleon superfluidity and for multipolarities l = 0, 1, 2 and 3. For each l we plot first few normal modes (solid lines) and first 25 superfluid modes (dashed lines), whose eigenfunctions δμl differ by the number of radial nodes n. At T∞ ≤ T∞cn(0) ≈ 2 × 108 K (see the left vertical dotted line), neutron superfluidity occupies the stellar centre. The bottom panel demonstrates the variation of the SFL-region (shown by hatching) with T∞. Shaded area in all panels shows the region where all neutrons are normal.
Fig. 6 contains five panels. Four upper panels present eigenfrequencies σ as functions of T∞8 for normal modes from Table 1 (thick horizontal lines) and for superfluid modes (dashes) with multipolarities l = 0, 1, 2 and 3. For each l there is an infinite set of superfluid modes whose eigenfunctions δμl differ by the number of radial nodes n; in the figure we plot the first 25 of them. The lower panel demonstrates broadening of the SFL-region with decreasing T∞8 (SFL-region is shown by hatches). For model 2 (which we employ here) the redshifted neutron critical temperature T∞cn(r) has a maximum at T∞cn max ≈ 5.1 × 108 K (right vertical dotted line). The neutron superfluidity reaches the stellar centre at T∞ = T∞cn(0) ≈ 2 × 108 K (left vertical dotted line). At T∞ > T∞cn max all neutrons are normal, hence only normal modes exist in the star. At T∞ < T∞cn(0) the core is completely occupied by the neutron superfluidity. One can see that the behaviour of superfluid modes differs strongly at T∞ > T∞cn(0) and at T∞ < T∞cn(0). This feature was discussed in Chugunov & Gusakov (2011) and Kantor & Gusakov (2011), where it was demonstrated that (roughly speaking) the frequencies σ of superfluid modes scale with Ynn and Rsfl as |$\sigma \sim \sqrt{Y_{\rm nn}}/R_{\rm sfl}$|, where Rsfl is the size of the SFL-region. With the increasing of temperature Ynn decreases, while the size of the SFL-region can either decrease [at T∞ > T∞cn(0)] or remain constant [at T∞ < T∞cn(0)]. As a result, there is a partial compensation of these two tendencies at T∞ > T∞cn(0) (hence, the frequency changes only weakly), while at T∞ < T∞cn(0) the effect of decreasing of Ynn is not compensated (hence, σ decreases with growing T∞). At T∞ ≲ 5 × 107 K Yik does not depend on T∞, and, as in the case of radial pulsations, the frequencies approach their low-temperature asymptotes.
7.4 Damping times for superfluid NSs
As in the case of eigenfrequencies, we first consider the e-folding times τ− 1b + s ≡ τbulk− 1 + τ− 1shear for radial (l = 0) pulsations for the simplified model 1 of nucleon superfluidity (see Fig. 1).
In Figs 7(a) and (d) we present the functions σ(T∞) and τb+s(T∞), obtained using the approximate method of Section 6.2. The frequencies and damping times are plotted for normal F mode (thick solid line) as well as for the first four superfluid modes 1, …, 4 (dashed lines).16 In the region shaded in grey the function τb+s(T∞) for the normal mode was not plotted (there are too many merging resonances in this region). The dotted curve in Figs 7(d)–(f) labelled Fnfh (‘nfh’ is the abbreviation for ‘normal-fluid hydrodynamics’) shows the damping time calculated using the ordinary hydrodynamics of non-superfluid liquid but taking into account the effects of superfluidity on the bulk and shear viscosities. This curve is analogous to the thick solid curve in Fig. 4, obtained under the same conditions but for the model 2 of nucleon superfluidity. The vertical dotted line in Figs 7(a) and (d) indicates a temperature at which frequencies of normal F mode and the first superfluid mode coincide.

Eigenfrequencies σ (panels a–c) and damping times τb+s (panels d–f) of a radially oscillating NS versus T∞ for model 1 of nucleon superfluidity. Panels (a) and (d): approximate solution (normal F mode and first four superfluid modes 1, …, 4 are shown by solid and dashed lines, respectively); panels (b) and (e): exact solution (first four exact modes I, …, IV are shown by solid, dashed, dot–dashed and dotted lines, respectively); panels (c) and (f): both approximate (dashed lines) and exact (solid lines) solutions. Panels (a)–(c) are the same spectra as those plotted, respectively, in Figs 5(a)–(c). Normal F mode is not shown in the shaded region because of technical reasons (too many resonances). Dotted lines in panels (d)–(f) show damping times for F mode calculated using ordinary normal-fluid hydrodynamics (see the text for more details). Part of the mode IV (at T∞ < 6 × 107 K) is shown by dots in the panel (f), as described in the text.
We present a detailed analysis of Fig. 7(d) in what follows, together with description of the approximate solutions for non-radial oscillation modes (Figs 8 and 9).

Damping times τb+s versus T∞ for various oscillation modes for model 2 of nucleon superfluidity. On each panel we plot one normal mode (shown by solid line; its multipolarity and name is indicated) and first 15 superfluid modes (dashed lines). Dotted lines show τb+s(T∞) for normal modes calculated using normal-fluid hydrodynamics (see the text for more details). In the shaded area all neutrons are normal and superfluid modes do not exist.

Eigenfrequencies σ (upper panels) and damping times τb+s (lower panels) versus T∞ for quadrupole (l = 2) oscillation modes in an increasingly larger scale. The normal p1 mode is shown by solid lines. Lower left-hand panel coincides with Fig. 8(e). Other lower panels are zoomed in versions of Fig. 8(e). Notations are the same as in Figs 6 and 8.
For comparison, Figs 7(b) and (e) demonstrate the results of the exact calculation of frequencies σ(T∞) and damping times τb+s(T∞) for the first four (I,…,IV) oscillation modes of the superfluid NS [the modes are shown by solid (I), dashed (II), dot–dashed (III) and dotted (IV) lines].
To see how well the approximate solution (Figs 7 a and d) agrees with the exact one (Figs 7 b and e), both solutions are presented in Figs 7(c) and (f). Dashes correspond to approximate solution, and solid lines correspond to exact solution. A portion of the mode IV in Fig. 7(f) is shown by thick dots because the corresponding approximate solution (the mode 1H) is not plotted. One sees that the agreement between the approximate and exact solutions is reasonable everywhere (average error does not exceed 10–25 per cent) except for the resonances (see below) and an interval of temperatures T∞ ≲ 3 × 107 K where the mode III of exact solution deviates from the second superfluid mode of approximate solution. To explain this deviation let us note that, as follows from Fig. 5(a), at such T∞ the frequency of the normal mode 1H practically coincides with that of the second superfluid mode. In that case equations (87) and (98) interact resonantly, so that the approximation of independent superfluid and normal modes is poor even though parameter s is small.17
Let us now consider the non-radial oscillations. Fig. 8 presents an approximate solution for the function τb+s(T∞), which is obtained for a realistic nucleon superfluidity model 2. By dashes we show superfluid modes, and solid lines correspond to normal modes. Each panel in the figure is plotted for one normal mode (its name and multipolarity l are indicated) and for the first 15 superfluid modes with the same l. By dots, as in Figs 7(d)–(f), we plot τb+s for a corresponding normal modes calculated using the ordinary normal-fluid hydrodynamics. In the shaded region superfluid modes were not plotted because all neutrons are normal there and the star oscillates as a non-superfluid.
In more detail damping times are demonstrated for quadrupole (l = 2) oscillation modes in Fig. 9. In particular, the normal p1 mode is shown there by solid lines. In the three lower panels we plot the dependence τb+s(T∞) in an increasingly larger scale. In the three upper panels we plot, in the same scale, the oscillation frequencies σ(T∞) (the corresponding spectrum was already presented in Fig. 6 in linear scale). Lower left-hand panel of Fig. 9 coincides with Fig. 8(e).
Let us discuss the main conclusions that can be drawn from the analysis of Figs 7(d), 8 and 9.
For any normal mode the dependence τb+s(T∞) (solid lines in these figures) has a set of resonance features (spikes) concentrated (for radial and p modes) to the critical temperature T∞cn(0) at which neutron superfluidity in the core centre dies out. For model 1 T∞cn(0) = Tcn max∞ = 6 × 108 K (see Fig. 1), and for model 2 T∞cn(0) ≈ 2 × 108 K (see Fig. 2). The resonances appear when frequency of the normal mode approaches the frequency of one of the superfluid modes. For instance, solid line in Fig. 7(a) crosses superfluid modes four times (in Figs 7 a and d, the temperature T∞ of the first crossing is shown by the vertical dotted line and equals T∞ ≈ 108 K). Correspondingly, four resonances appear in Fig. 7(d). A similar situation can be observed in Figs 8 and 9. Near resonances τb+s for normal mode rapidly decreases by one to two orders of magnitude (see item 2) and, in the resonance point, it becomes strictly equal to τb+s for the corresponding superfluid mode.
Such behaviour of the approximate solution τb+s(T∞) for normal modes in the vicinity of resonances can be easily understood. In resonance points, in which the frequencies of superfluid and normal modes coincide, equation (98) has a non-trivial solution even in the absence of the source δμnorm l. For it to be satisfied with the source, the oscillation amplitude δμ must be infinitely large. In other words, in resonance points all the energy must be contained in superfluid degrees of freedom (in particular, near resonances Wsfl ≫ Wb and Vsfl ≫ Vb). Formally, this means that in the resonance point the damping time τb+s should be exactly the same as for the superfluid mode.
Another important point that is worth noting is that, as follows from Fig. 7(f), the approximate solution for the normal radial F mode describes qualitatively well the exact solution near resonances (the latter is shown by solid lines). We expect that the same is also true for non-radial modes for which the exact solution was not attempted. At first glance such an agreement between the approximate and exact solutions seems surprising because the approximation s = 0 should not work in the vicinity of resonances, where the frequencies of superfluid and normal modes are close to each other. Nevertheless, one verifies that this approximation is still suitable for a qualitatively correct description of the function τb+s(T∞) if one bears in mind that (i) close to any resonance the exact solution is a linear superposition of independent solutions describing (intersecting) superfluid and normal modes and (ii) τb+s for the superfluid mode is much less than for the normal mode.
Items (i) and (ii) mean that, in the exact solution, the main contribution to τb+s comes from the superfluid mode (while the contribution from the normal mode is small). This leads us to conclusion that the superfluid modes are the main sources of viscous dissipation in the vicinity of resonance points. The same conclusion was already drawn above using the approximate method of Section 6.2. This explains why the approximate method gives qualitatively correct results for τb+s(T∞) near resonances.
In order to avoid confusion let us emphasize that the function τb+s(T∞) contains resonance features (spikes) for normal modes only in the approximate solution (see Figs 7 d, 8 and 9). In the exact solution any normal oscillation mode turns into a superfluid one near resonance (and vice versa). This leads to an abrupt decreasing (increasing) of τb+s and formation of a ‘step-like’ structure rather than spike (see Fig. 7 e).
It was already mentioned above that, as follows from Figs 7(d), 8 and 9, normal modes (far from resonances) damp out by one to two orders of magnitude slower than those superfluid modes with which they can have equal frequencies (i.e. intersect in the σ-T∞ plane).
What is the reason for such a fast damping of superfluid modes? To be more concrete, below we consider a low-temperature case, T∞ ≲ 3 × 107 K. There are three main factors: (i) for superfluid modes eigenfunctions Wsfl and Vsfl have a maximum in the central regions of a star where the shear viscosity is maximal. On the contrary, for normal modes the maximum of eigenfunctions Wb and Vb lies closer to the NS surface, where the shear viscosity coefficient can be substantially (five and more times) smaller. As a consequence, |$\mathfrak {W}_{\rm shear}$| for superfluid modes turns out to be greater (and hence τshear smaller) than for normal modes. (ii) The energy of superfluid modes is given by equation (73) and depends on the quantity y (see equation 68 for the definition of y). At low T∞ the parameter y is small, y ∼ np/nn ∼ 0.04-0.09, which also results in decreasing of the characteristic damping times for superfluid modes.18 (iii) This factor is particularly important for radial oscillations (l = 0) and is related to a coefficient α1 in expression (78) for the damping time τshear due to shear viscosity. This coefficient is given by equation (81) which is a sum of four terms. It turns out that for the normal radial modes the first term is well compensated by the third term H2, while the other terms vanish. For the superfluid modes such compensation does not occur because for them H2 = 0.
At low enough T∞ the damping times for normal radial and p modes can be several times larger or smaller than τb+s, calculated employing ordinary hydrodynamics of non-superfluid liquid but accounting for the effects of superfluidity on the bulk and shear viscosity coefficients (dotted lines in Figs 7 d, 8 a, d, e and f, and 9). Let us inspect, for example, Fig. 8(a). One sees that at T∞ ≲ 108 K, τb+s, calculated in the frame of non-superfluid hydrodynamics, is approximately four times larger than τb+s determined self-consistently. This difference arises because to plot the dotted curve we used the formulas of Section 4 in which Wsfl = Vsfl = 0. As T∞ grows, however, the difference in two ways of calculating τb+s rapidly decreases because the SFL-region becomes smaller and hence its contribution to τb+s becomes less and less pronounced.
Unlike the radial and p modes, the agreement between dotted and solid lines for normal f modes is very good (see Figs 8 b and c), which means that for these modes use of the non-superfluid hydrodynamics (far from resonances) is well justified. The reason for such a good agreement of damping times is related to a relatively weak compression–decompression of matter in the course of the f-type oscillations. As a consequence, for the normal f modes the source δμnorm l in equation (98) is small, so that far from the resonances δμ ≈ 0 and the superfluid degrees of freedom are almost not excited (Wsfl ≈ Vsfl ≈ 0, see equations 48, 52 and 90). This result confirms, extends and, we think, provides a deeper understanding of the results previously obtained in a Newtonian framework by e.g. Lindblom & Mendell (1994) and Andersson et al. (2009).
At T∞ → T∞cn max = 6 × 108 K, one can observe the rapid increasing of τb+s for superfluid modes in Fig. 7. It is bounded from above by τbulk and is related with the tendency of τshear to grow to infinity in this limit. Such a behaviour of τshear was discussed in detail in Kantor & Gusakov (2011) and is specific for model 1 of nucleon superfluidity.
8 SUMMARY
In this paper we, for the first time, self-consistently analyse the effects of nucleon superfluidity on damping of oscillations of non-rotating general relativistic NSs. Our main results are summarized below.
The analytic formulas are derived for the oscillation energy Emech (71) and for the characteristic damping times τbulk (77) and τshear (78) due to the bulk and shear viscosities. These expressions are valid for oscillations of arbitrary multipolarity l. Expression (71) for Emech is the generalization of formula (26) of Thorne & Campolattaro (1967), written for a non-superfluid NS. Expressions (77) and (78) are the generalizations, to the case of superfluidity, of formulas (5) and (6) in Cutler et al. (1990). Note that the damping times, calculated using the formulas of Cutler et al. (1990), appear to be two times smaller than our τbulk and τshear, calculated from equations (77) and (78) under assumption that superfluid degrees of freedom are suppressed (i.e. Wsfl = Vsfl = 0).
An approximate method is developed in detail and applied, which allows one to easily determine the eigenfrequencies and eigenfunctions of an oscillating superfluid NS, provided that they are known for a normal (non-superfluid) star of the same mass (see Section 6.2). The method is based on the approximate decoupling of equations describing superfluid and normal oscillation modes and exploits the ideas first formulated in Gusakov & Kantor (2011) and Chugunov & Gusakov (2011).
Using radial oscillations as an example, and adopting the simplified model 1 of nucleon superfluidity (Fig. 1), we demonstrate that this method leads to oscillation frequencies and characteristic damping times that agree well with the results of exact calculation.
The approximate method of Section 6.2 is applied to study non-radial oscillations of a superfluid NS assuming the realistic model 2 of nucleon superfluidity (Fig. 2). A number of normal and superfluid oscillation modes with multipolarities l = 0, …, 3 are considered. In particular, the following normal modes are analysed: F mode for l = 0, p1 mode for l = 1 and f and p1 modes for l = 2 and 3.
We demonstrate the following.
As a rule, for any given normal mode (whose frequency σ coincides with the corresponding frequency of a non-superfluid NS and does not depend on the internal redshifted stellar temperature T∞), the viscous damping time τb + s ≡ (τ− 1bulk + τ− 1shear)− 1 is one order of magnitude greater than τb+s for those superfluid modes that can intersect the normal mode in the σ-T∞ plane. This effect is non-local (occurs only after integration over the NS volume) and is determined by a number of factors (see item 2 of Section 7.4).
The function τb+s(T∞) for any normal mode contains resonance features. In resonance points the frequency σ of a normal mode coincides with that of some of the superfluid modes (their σ depends on T∞). When passing a resonance (e.g. with growing T∞), τb+s initially rapidly decreases (by one to two orders of magnitude) until it reaches the value of τb+s for this superfluid mode and, after that, it increases again (see Figs 7 d, 8 and 9).
Resonance features (spikes) appear only in the approximate treatment of Section 6, in which the normal and superfluid modes intersect at resonance points (see e.g. Fig. 5 a). In the exact solution instead of crossings one has avoided crossings of modes (Fig. 5 b). Near avoided crossings any real mode changes its behaviour from normal-like to superfluid-like (and vice versa). As a result, instead of spikes one has a very rapid step-like decreasing (increasing) of τb+s (cf. Figs 7d and e).
Sufficiently far from the resonances τb+s for normal radial and p modes, determined self-consistently employing the hydrodynamics of a superfluid liquid, can differ several fold from τb+s, calculated using the ordinary normal-fluid hydrodynamics (but accounting for the effects of superfluidity on the shear and bulk viscosities). The latter approximation is often adopted in the literature devoted to oscillations of NSs.
In contrast to radial and p modes, for f modes far from the resonances, use of the ordinary hydrodynamics of non-superfluid liquid for calculation of τb+s is well justified. The reason is that for f-type oscillations the imbalance δμ of chemical potentials is relatively small (matter does not compress significantly during oscillations). Thus, superfluid degrees of freedom are almost not excited (see Sections 6.2 and 7.4).
Since for f modes far from the resonances δμ is small (i.e. deviation from the beta-equilibrium is weak), bulk viscous damping of f modes is suppressed in comparison to p modes.
Though here we only considered oscillations of superfluid non-rotating NSs, we expect that the main conclusions of this work will also remain (mostly) unchanged for rotating NSs. Our results indicate that dissipative evolution of oscillating NSs may follow quite different scenarios than those usually considered in the literature. This is especially true if one is interested in the combined analysis of damping of oscillations and thermal evolution of an NS or in the analysis of instability windows, which are the values of T∞ and rotation frequency at which a star becomes unstable with respect to the emission of gravitational waves (e.g. the r mode instability; see Andersson 1998; Friedman & Morsink 1998). These issues are extremely interesting and important, but we left them beyond the scope of this paper and will address the related topics in our subsequent publication.
This study was supported by the Dynasty Foundation, Ministry of Education and Science of Russian Federation (contract No. 11.G34.31.0001 with SPbSPU and leading scientist G. G. Pavlov, and agreement No. 8409, 2012), RFBR (11-02-00253-a, 12-02-31270-mol-a), FASI (grant NSh-4035.2012.2), RF president programme (grant MK-857.2012.2), by the RAS presidium programme ‘Support for young scientists’, and by CompStar, a Research Networking Programme of the European Science Foundation.
Note that, in Gusakov (2007), there is an additional term in the expression for Sμ, so that
This equality holds true only if one neglects the thermal conductivity, as we assume in equation (12).
However, beyond the linear approximation, equations (4), (5), (19) and (20) yield U(b) μUμ(b) = −1 + YniYnk w(i) μwμ(k)/n2b. The normalization condition (45) is generally not fulfilled because the reference frame in which Uμ(b) = (1, 0, 0, 0) is not comoving [i.e. jμ(b)U(b) μ ≠ −nb in that reference frame]. As it was already indicated in Section 2, all thermodynamic variables are measured in the reference frame, in which uμ = (1, 0, 0, 0).
It is worth to make a number of comments on equation (89): (i) in Gusakov & Kantor (2011) this equation was derived under the assumption that the only superfluid species are neutrons (i.e. Ypi = 0). A generalization of this equation to the case of possible proton superfluidity is presented in (Chugunov & Gusakov 2011, see their equation 3). (ii) In both papers (Chugunov & Gusakov 2011; Gusakov & Kantor 2011), this equation is written with the same mistake. In particular, in Chugunov & Gusakov (2011) one should write ne ∂j(eν/2 δμ) instead of ne eν/2 ∂j(δμ) in the right-hand side of equation (3).
The equality ∂ε(nb, ne)/∂ne = −δμ follows from the second law of thermodynamics (15), which can be rewritten in our case as dε = μn dnb − δμ dne.
Here and below by ‘normal modes’ we mean oscillation modes (of approximate solution) that also exist in the normal (non-superfluid) star.
This is the main advantage of treating |$\tilde{s}$| in a non-perturbative way. Note, however, that this trick leads to somewhat ‘excessive’ accuracy of the approximate solution to oscillation equations: the retained terms depending on |$\tilde{s}$| may lead to smaller correction to the solution than the s-dependent terms which were ignored. Bearing this in mind and with the aim to simplify consideration, in Gusakov & Kantor (2011) it was assumed that both parameters s and |${\tilde{s}}$| vanish in the s = 0 approximation. Such an approach is also possible. In that case, strictly speaking, the resulting Einstein equations would differ slightly from the equations describing oscillations of a non-superfluid NSs. In particular, instead of the standard adiabatic index of the ‘frozen’ npe-matter γfr = (nb/P)[∂P(nb, xe)/∂nb], the new index would appear, γ = (nb/P)[∂P(nb, ne)/∂nb]. However, this difference is not essential, because s and |$\tilde{s}$| are small.
The f mode is absent in the case of l = 1.
Thus, it would not be correct to say that any real oscillation mode of a superfluid star is either purely superfluid or purely normal: for some T∞ it can show itself as a superfluid, but for other T∞ it can behave as a normal mode (see Fig. 5b).
For a quantitatively correct description of the function τb+s(T∞) in that case it is, in principle, straightforward to develop a perturbation theory similar to the degenerate perturbation theory of quantum mechanics (see below the discussion of resonances in Figs 7–9).
To get an estimate for y we made use of the sum rule μnYnn + μpYnp = nn valid at T∞ = 0 (Gusakov, Kantor & Haensel 2009a), and neglected the small matrix element Ynp in comparison with Ynn.
REFERENCES
APPENDIX A: BOUNDARY CONDITIONS TO EQUATION (98)
- The boundary (one of the boundaries) between the SFL-region and non-superfluid matter lies inside the core and is defined by the condition T = Tcn(Rb) (Rb is the radial coordinate of the boundary). Then at the boundary Ynn(Rb) = Ynp(Rb) = 0 and from equations (90) and (98), one has(A3)\begin{equation} \delta \mu _l^\prime =\frac{{\rm e^{\lambda -\nu /2} \, \omega ^2}}{h^\prime \, \mathfrak {B}}\, (\delta \mu _l - \delta \mu _{{\rm norm} \, l}). \end{equation}
- Outer boundary of the SFL-region coincides with the crust–core interface (Rb = Rcc). In that case T < Tcn(Rcc) [i.e. Ynn(Rcc) and Ynp(Rcc) are non-zero] and from equation (90) it follows thatThe conditions (A1)–(A4) are necessary and sufficient for solving equation (98).(A4)\begin{equation} \delta \mu _{l}^\prime (R_{\rm cc})=0. \end{equation}