[1,2]\fnmPiotr \surMagierski

\equalcont

These authors contributed equally to this work.

\equalcont

These authors contributed equally to this work.

\equalcont

These authors contributed equally to this work.

\equalcont

These authors contributed equally to this work.

[1]\orgdivFaculty of Physics, \orgnameWarsaw University of Technology, \orgaddress\streetulica Koszykowa 75, \cityWarsaw, \postcode00-662, \countryPoland

2]\orgdivPhysics Department, \orgnameUniversity of Washington, \orgaddress\street3910 15th Ave. NE, \citySeattle, \postcodeWA 98195-1560, \countryUSA

3]\orgnameInstitute of Physics, Polish Academy of Sciences, \orgaddress\streetAleja Lotnikow 32/46, \cityWarsaw, \postcode02-668, \countryPoland

Quantum vortices in fermionic superfluids: from ultracold atoms to neutron stars.

piotr.magierski@pw.edu.pl    \fnmAndrea \surBarresi andrea.barresi.dokt@pw.edu.pl    \fnmAndrzej \surMakowski andrzej.makowski2.dokt@pw.edu.pl    \fnmDaniel \surPęcak daniel.pecak@pw.edu.pl    \fnmGabriel \surWlazłowski gabriel.wlazlowski@pw.edu.pl * [ [
Abstract

Superfluid dilute neutron matter and ultracold gas, close to the unitary regime, exhibit several similarities. Therefore, to a certain extent, fermionic ultracold gases may serve as emulators of dilute neutron matter, which forms the inner crust of neutron stars and is not directly accessed experimentally. Quantum vortices are one of the most significant properties of neutron superfluid, essential for comprehending neutron stars’ dynamics. The structure and dynamics of quantum vortices as a function of pairing correlations’ strength are being investigated experimentally and theoretically in ultracold gases. Certain aspects of these studies are relevant to neutron stars. We provide an overview of the characteristics of quantum vortices in s-wave-type fermionic and electrically neutral superfluids. The main focus is on the dynamics of fermionic vortices and their intrinsic structure.

keywords:
fermionic superfluid, quantum vortices, vortex structure, vortex dynamics

1 Introduction

Pairing correlations in nuclear systems are essential for describing both static and dynamic properties of nuclei. Such phenomena as odd-even mass difference or back bending of moments of inertia of atomic nuclei cannot be explained without invoking the concept of pair correlations (see  [1, 2, 3] and references threrein). However, the precise extraction of the pairing magnitude is difficult due to the finite size of the system. The information about pairing gap is obscured by polarization effects associated with adding or substracting a single nucleon (see eg. [4] for discussion of the interplay between pairing and mean-field effects in the odd-even mass staggering). For most applications, it is sufficient to consider one particular feature of pairing correlations: the presence of the energy gap at the Fermi surface. Various manifestations of superfluidity can be traced back to this single quantity, usually treated as a parameter of mean-field models (see e.g. [5]). Therefore, it is still an open question whether pairing in nuclear systems can be treated as a condensate of pairs, giving rise to the superfluid phase, or it is merely an incoherent pair correlation. Clearly, the condition of nonvanishing off-diagonal long-range order cannot be applied to atomic nuclei due to their finite size. However, if a superfluid phase exists, it has to be seen in phenomena requiring a well-defined phase factor of the pairing field. These involve analogs of the Josephson effect [6, 7, 8, 9], both in DC and AC realizations, and solitonic excitations [10, 11]. Manifestations of these effects were reported to be seen in nuclear collisions through enhanced neutron pair transfer [12, 13, 14], photon radiation of particular energy [15, 16, 17], or an increase of the barrier for capture of colliding nuclei [18].

On the other hand, even stronger pairing correlations are predicted to be present in nuclear matter at subnuclear densities [19]. Unfortunately, dilute nuclear, particularly neutron matter, cannot be directly accessed experimentally. They form outer layers of neutron stars — the so-called inner crust [20]. However, since the system is of macroscopic size and is characterized by strong pairing correlations, the superfluid phase (of s-wave type) is likely to be formed [21, 22, 23, 24]. Moreover, the dynamics of a rotating star raises the natural question concerning the role of quantum vortices in the dynamics of the crust, especially in the context of observed glitch phenomenon [25, 26]. The presence of quantum vortices, which require a particular configuration of pairing field, is possible only under the assumption that the neutron matter is superfluid.

The properties of vortices in neutron stars can be modeled using Density Functional Theory (DFT) based on the properties of known nuclear systems [27, 28]. However, since neutron matter at subnuclear densities is relatively close to the so-called unitary regime, one can expect that the properties of vortices are close to those observed in ultracold atomic gases. Therefore, theoretical studies confronted with experiments on ultracold atomic gases can help provide constraints for the theory of quantum vortices in neutron matter [29]. Fermionic ultracold gases provide a particularly useful playground to investigate properties of vortices [30], since they provide an example of a system that is electrically neutral, contrary to metal superconductors, which is close to the conditions in the inner crust (proton fraction is small and localized in impurities). Therefore, all effects related to interactions with magnetic and electric fields can be ignored.

In this review, we summarize the main characteristics of vortices and their dynamics predicted to exist in neutron matter and those studied in ultracold gases, supported by experiments that are relevant for neutron stars.

2 Vortex structure and thermodynamic properties

The discovery of He-II triggered studies of quantum vortices, which could be rather easily generated in this system. However, in superfluid He-II, the healing length is relatively small, and the vortex core sizes do not exceed a few Å  [31]. Consequently, they cannot be directly observed in experiments. The situation is different in ultracold atomic gases, which are dilute. Therefore, the vortex core size can be about 103104superscript103superscript10410^{3}-10^{4}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT times larger, depending on the value of the scattering length. Attempts to cool down fermionic gases eventually led to the observation of quantum vortices, which provided unambiguous evidence of superfluidity in the system [32]. The regular lattice of vortices has been detected in fermionic gases in a wide range of scattering lengths ranging from Bardeen-Cooper-Schrieffer (BCS) regime through unitary limit to molecular Bose-Einstein Condensate (BEC) regime. The versatility of ultracold atomic gases allows for considering properties of vortices within the whole range of BEC-BCS crossover. It is particularly important as one can smoothly transform fermionic vortices into bosonic vortices, varying the value of the scattering length.

Both in the BEC and BCS regimes, the presence of a quantum vortex is associated with a particular configuration of the order parameter, which vanishes in the vortex core. The same is valid for superfluid systems, e.g. He-II, where the "condensate wave function" plays the role of the order parameter. In both cases, the order parameter is responsible for the stability and dynamics of quantum vortices. The main difference, however, between bosonic and fermionic vortices lies in their structure, which, in the latter case, is more complex. In the bosonic systems at T=0𝑇0T=0italic_T = 0 the vortex core is essentially empty. With increasing temperature, thermal excitations can be generated, forming an admixture of a normal component inside the core. Regarding fermionic vortices, the particle density in the core does not vanish even at T=0𝑇0T=0italic_T = 0. Indeed, the order parameter in the vicinity of the center of the core creates a particular shape of the pairing potential and does not prevent fermions from occupying the core, leading merely to the emergence of quantized in-gap states. The first estimates for fermion-bound states in the core have been given in Ref. [33], and therefore, they are sometimes referred to as Caroli-De Gennes-Matricon states. It was also conjectured that due to these states, thermodynamic properties of superconductors would be affected, namely the specific heat at temperature kBT<|Δ|subscript𝑘𝐵𝑇Δk_{B}T<|\Delta|italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T < | roman_Δ | (|Δ|Δ|\Delta|| roman_Δ | denotes the magnitude of pairing field far from the vortex) will contain contribution behaving linearly as a function of T𝑇Titalic_T. When one goes from a deep BCS regime to the unitary regime, the spectrum of states in the core gets gradually more sparse. The in-gap states are expected to disappear eventually beyond the unitary regime on the BEC side [34].

Before discussing properties of the core states let us first summarize briefly Bogoliubov-de Gennes (BdG) formalism, which allow to capture the main properties of the quantum vortex. The explicit form of BdG equations for spin-imbalanced system reads (no spin-orbit coupling is considered):

(un,(𝒓)un,(𝒓)vn,(𝒓)vn,(𝒓))=En(un,(𝒓)un,(𝒓)vn,(𝒓)vn,(𝒓)),=(h(𝒓)μ00Δ(𝒓)0h(𝒓)μΔ(𝒓)00Δ(𝒓)h(𝒓)+μ0Δ(𝒓)00h(𝒓)+μ),formulae-sequencematrixsubscript𝑢𝑛𝒓subscript𝑢𝑛𝒓subscript𝑣𝑛𝒓subscript𝑣𝑛𝒓subscript𝐸𝑛matrixsubscript𝑢𝑛𝒓subscript𝑢𝑛𝒓subscript𝑣𝑛𝒓subscript𝑣𝑛𝒓matrixsubscript𝒓subscript𝜇00Δ𝒓0subscript𝒓subscript𝜇Δ𝒓00superscriptΔ𝒓subscriptsuperscript𝒓subscript𝜇0superscriptΔ𝒓00subscriptsuperscript𝒓subscript𝜇\displaystyle\begin{gathered}{\cal H}\begin{pmatrix}u_{n,\uparrow}(\bm{r})\\ u_{n,\downarrow}(\bm{r})\\ v_{n,\uparrow}(\bm{r})\\ v_{n,\downarrow}(\bm{r})\end{pmatrix}=E_{n}\begin{pmatrix}u_{n,\uparrow}(\bm{r% })\\ u_{n,\downarrow}(\bm{r})\\ v_{n,\uparrow}(\bm{r})\\ v_{n,\downarrow}(\bm{r})\end{pmatrix},\\ {\cal H}=\begin{pmatrix}h_{\uparrow}(\bm{r})-\mu_{\uparrow}&0&0&\Delta(\bm{r})% \\ 0&h_{\downarrow}(\bm{r})-\mu_{\downarrow}&-\Delta(\bm{r})&0\\ 0&-\Delta^{*}(\bm{r})&-h^{*}_{\uparrow}(\bm{r})+\mu_{\uparrow}&0\\ \Delta^{*}(\bm{r})&0&0&-h^{*}_{\downarrow}(\bm{r})+\mu_{\downarrow}\end{% pmatrix},\end{gathered}start_ROW start_CELL caligraphic_H ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) = italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) , end_CELL end_ROW start_ROW start_CELL caligraphic_H = ( start_ARG start_ROW start_CELL italic_h start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ( bold_italic_r ) - italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL roman_Δ ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_h start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ( bold_italic_r ) - italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_CELL start_CELL - roman_Δ ( bold_italic_r ) end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL - roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_r ) end_CELL start_CELL - italic_h start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ( bold_italic_r ) + italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_r ) end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - italic_h start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ( bold_italic_r ) + italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) , end_CELL end_ROW (3)

where μ,subscript𝜇\mu_{\uparrow,\downarrow}italic_μ start_POSTSUBSCRIPT ↑ , ↓ end_POSTSUBSCRIPT are chemical potentials for spin-up and spin-down particles, respectively. Single particle hamiltonian h=h=22m2subscriptsubscriptsuperscriptPlanck-constant-over-2-pi22𝑚superscript2h_{\uparrow}=h_{\downarrow}=-\frac{\hbar^{2}}{2m}\nabla^{2}italic_h start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT = italic_h start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT = - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT consists of the kinetic term and the mean-field potential, but we will omit the latter, as of the secondary importance in this case. In principle, the single particle hamiltonian hσsubscript𝜎h_{\sigma}italic_h start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT may explicitly depend on the spin state (see eg. Ref. [35]). The pairing gap is related to quasi-particle wave-functions:

Δ(𝒓)=geff20<En<Ec(un,(𝒓)vn,(𝒓)un,(𝒓)vn,(𝒓)),Δ𝒓subscript𝑔eff2subscript0subscript𝐸𝑛subscript𝐸𝑐subscript𝑢𝑛𝒓superscriptsubscript𝑣𝑛𝒓subscript𝑢𝑛𝒓superscriptsubscript𝑣𝑛𝒓\Delta(\bm{r})=-\dfrac{g_{\textrm{eff}}}{2}\sum_{0<E_{n}<E_{c}}(u_{n,\uparrow}% (\bm{r})v_{n,\downarrow}^{*}(\bm{r})-u_{n,\downarrow}(\bm{r})v_{n,\uparrow}^{*% }(\bm{r})),\\ roman_Δ ( bold_italic_r ) = - divide start_ARG italic_g start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT 0 < italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT < italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_u start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) italic_v start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_r ) - italic_u start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) italic_v start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_r ) ) , (4)

where geffsubscript𝑔effg_{\textrm{eff}}italic_g start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT is a regularized coupling constant and Ecsubscript𝐸𝑐E_{c}italic_E start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is cut-off energy scale, see [35] for details of the regularization scheme.

The BdG equations (3) decouple into two independent sets:

(h(𝒓)μΔ(𝒓)Δ(𝒓)h(𝒓)+μ)(un,(𝒓)vn,(𝒓))=En+(un,(𝒓)vn,(𝒓)),matrixsubscript𝒓𝜇Δ𝒓superscriptΔ𝒓subscriptsuperscript𝒓𝜇matrixsubscript𝑢𝑛𝒓subscript𝑣𝑛𝒓subscript𝐸limit-from𝑛matrixsubscript𝑢𝑛𝒓subscript𝑣𝑛𝒓\displaystyle\begin{gathered}\begin{pmatrix}h_{\uparrow}(\bm{r})-\mu&\Delta(% \bm{r})\\ \Delta^{*}(\bm{r})&-h^{*}_{\downarrow}(\bm{r})+\mu\end{pmatrix}\begin{pmatrix}% u_{n,\uparrow}(\bm{r})\\ v_{n,\downarrow}(\bm{r})\end{pmatrix}=E_{n+}\begin{pmatrix}u_{n,\uparrow}(\bm{% r})\\ v_{n,\downarrow}(\bm{r})\end{pmatrix},\end{gathered}start_ROW start_CELL ( start_ARG start_ROW start_CELL italic_h start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ( bold_italic_r ) - italic_μ end_CELL start_CELL roman_Δ ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_r ) end_CELL start_CELL - italic_h start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ( bold_italic_r ) + italic_μ end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) = italic_E start_POSTSUBSCRIPT italic_n + end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) , end_CELL end_ROW (6)
(h(𝒓)μΔ(𝒓)Δ(𝒓)h(𝒓)+μ)(un,(𝒓)vn,(𝒓))=En(un,(𝒓)vn,(𝒓)),matrixsubscript𝒓𝜇Δ𝒓superscriptΔ𝒓subscriptsuperscript𝒓𝜇matrixsubscript𝑢𝑛𝒓subscript𝑣𝑛𝒓subscript𝐸limit-from𝑛matrixsubscript𝑢𝑛𝒓subscript𝑣𝑛𝒓\displaystyle\begin{gathered}\begin{pmatrix}h_{\downarrow}(\bm{r})-\mu&-\Delta% (\bm{r})\\ -\Delta^{*}(\bm{r})&-h^{*}_{\uparrow}(\bm{r})+\mu\end{pmatrix}\begin{pmatrix}u% _{n,\downarrow}(\bm{r})\\ v_{n,\uparrow}(\bm{r})\end{pmatrix}=E_{n-}\begin{pmatrix}u_{n,\downarrow}(\bm{% r})\\ v_{n,\uparrow}(\bm{r})\end{pmatrix},\end{gathered}start_ROW start_CELL ( start_ARG start_ROW start_CELL italic_h start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ( bold_italic_r ) - italic_μ end_CELL start_CELL - roman_Δ ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL - roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_italic_r ) end_CELL start_CELL - italic_h start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT ( bold_italic_r ) + italic_μ end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) = italic_E start_POSTSUBSCRIPT italic_n - end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) , end_CELL end_ROW (8)

where μ=12(μ+μ)𝜇12subscript𝜇subscript𝜇\mu=\frac{1}{2}(\mu_{\uparrow}+\mu_{\downarrow})italic_μ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT + italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ) denotes mean chemical potential and En±=En±Δμ2subscript𝐸limit-from𝑛plus-or-minusplus-or-minussubscript𝐸𝑛Δ𝜇2E_{n\pm}=E_{n}\pm\frac{\Delta\mu}{2}italic_E start_POSTSUBSCRIPT italic_n ± end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ± divide start_ARG roman_Δ italic_μ end_ARG start_ARG 2 end_ARG with Δμ=μμΔ𝜇subscript𝜇subscript𝜇\Delta\mu=\mu_{\uparrow}-\mu_{\downarrow}roman_Δ italic_μ = italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT. Solutions of equations (6) and (8) are connected through the symmetry relation, namely if vector φ+=(un,vn)Tsubscript𝜑superscriptsubscript𝑢𝑛absentsubscript𝑣𝑛absent𝑇\varphi_{+}=\left(u_{n\uparrow},v_{n\downarrow}\right)^{T}italic_φ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT = ( italic_u start_POSTSUBSCRIPT italic_n ↑ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_n ↓ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT represents a solution of Eq. (6) with eigenvalue Ensubscript𝐸𝑛E_{n}italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, then vector φ=(vn,un)Tsubscript𝜑superscriptsuperscriptsubscript𝑣𝑛absentsuperscriptsubscript𝑢𝑛absent𝑇\varphi_{-}=(v_{n\uparrow}^{*},u_{n\downarrow}^{*})^{T}italic_φ start_POSTSUBSCRIPT - end_POSTSUBSCRIPT = ( italic_v start_POSTSUBSCRIPT italic_n ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_u start_POSTSUBSCRIPT italic_n ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT is a solution of Eq. (8) with eigenvalue Ensubscript𝐸𝑛-E_{n}- italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. Therefore it is sufficient to solve equations (6) only (for all quasiparticle energy states), and then solutions with positive quasiparticle energies contribute to the spin-down densities, whereas solutions with negative energies to the spin-up densities.

For the solution describing a vortex the particular form of pairing field is required Δ(𝒓)=|Δ(r)|exp(iϕ)Δ𝒓Δ𝑟𝑖italic-ϕ\Delta(\boldsymbol{r})=|\Delta(r)|\exp(i\phi)roman_Δ ( bold_italic_r ) = | roman_Δ ( italic_r ) | roman_exp ( italic_i italic_ϕ ) (for the vortex rotating clockwise), where (r,ϕ,z)𝑟italic-ϕ𝑧(r,\phi,z)( italic_r , italic_ϕ , italic_z ) are cylindical coordinates with r𝑟ritalic_r being the distance from the center of the vortex core and z𝑧zitalic_z - the coordinate along the vortex line (which is assumed to be a straight line). Therefore the pairing field does not depend on z𝑧zitalic_z and the general solution of eq. (3) describing the vortex reads:

(un,(𝒓)vn,(𝒓))=(unmkz,(r)eimϕeikzzvnmkz,(r)ei(m1)ϕeikzz),matrixsubscript𝑢𝑛𝒓subscript𝑣𝑛𝒓matrixsubscript𝑢𝑛𝑚subscript𝑘𝑧𝑟superscript𝑒𝑖𝑚italic-ϕsuperscript𝑒𝑖subscript𝑘𝑧𝑧subscript𝑣𝑛𝑚subscript𝑘𝑧𝑟superscript𝑒𝑖𝑚1italic-ϕsuperscript𝑒𝑖subscript𝑘𝑧𝑧\displaystyle\begin{gathered}\begin{pmatrix}u_{n,\uparrow}(\bm{r})\\ v_{n,\downarrow}(\bm{r})\end{pmatrix}=\begin{pmatrix}u_{nmk_{z},\uparrow}(r)e^% {im\phi}e^{ik_{z}z}\\ v_{nmk_{z},\downarrow}(r)e^{i(m-1)\phi}e^{ik_{z}z}\end{pmatrix},\end{gathered}start_ROW start_CELL ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) = ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_n italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , ↑ end_POSTSUBSCRIPT ( italic_r ) italic_e start_POSTSUPERSCRIPT italic_i italic_m italic_ϕ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_z end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_n italic_m italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , ↓ end_POSTSUBSCRIPT ( italic_r ) italic_e start_POSTSUPERSCRIPT italic_i ( italic_m - 1 ) italic_ϕ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_z end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ) , end_CELL end_ROW (10)

The other pair of solutions can be obtained using the symmetry properties of BdG equations. Note that (10) implies that the angular momentum of spin-up particles (holes) differs from spin-down holes (particles), of the same energy, by Planck-constant-over-2-pi\hbarroman_ℏ.

The spectrum of quantized states inside the vortex core has a peculiar structure, reflecting the behavior of the order parameter. In the deep BCS limit, the main features of the spectrum can be reproduced within the semiclassical Andreev approximation [36]. In this approximation one assumes the separation of scales related to the density and the pairing field spatial modulation. The former is set by the Fermi wavelength, whereas the latter - by the coherence length ξ𝜉\xiitalic_ξ. Therefore one may decompose the variation of u𝑢uitalic_u and v𝑣vitalic_v components of wave-functions (see Eq. (3)) at the Fermi surface into rapidly oscillating parts associated with the Fermi momentum kFsubscript𝑘Fk_{\textrm{F}}italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT and smooth variations governed by the coherence length, i.e. u(𝒓)=ei𝒌F𝒓u~(𝒓)𝑢𝒓superscript𝑒𝑖subscript𝒌F𝒓~𝑢𝒓u(\bm{r})=e^{i\bm{k}_{\textrm{F}}\cdot\bm{r}}\tilde{u}(\bm{r})italic_u ( bold_italic_r ) = italic_e start_POSTSUPERSCRIPT italic_i bold_italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT ⋅ bold_italic_r end_POSTSUPERSCRIPT over~ start_ARG italic_u end_ARG ( bold_italic_r ) with |𝒌F|=kFsubscript𝒌Fsubscript𝑘F|\bm{k}_{\textrm{F}}|=k_{\textrm{F}}| bold_italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT | = italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT, and similarly for the v𝑣vitalic_v component. The Andreev approximation can be used also in the case of spin imbalance systems, providing the local polarization is relatively weak Δμ=μμ12(μ+μ)εF=kF2/2Δ𝜇subscript𝜇subscript𝜇much-less-than12subscript𝜇subscript𝜇subscript𝜀𝐹superscriptsubscript𝑘F22\Delta\mu=\mu_{\uparrow}-\mu_{\downarrow}\ll\frac{1}{2}(\mu_{\uparrow}+\mu_{% \downarrow})\approx\varepsilon_{F}=k_{\textrm{F}}^{2}/2roman_Δ italic_μ = italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ≪ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT + italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT ) ≈ italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2. We will focus only on one set of BdG equations, which in Andreev approximation describing states close to the Fermi surface acquire the form (we set =m=1Planck-constant-over-2-pi𝑚1\hbar=m=1roman_ℏ = italic_m = 1):

(i𝒌FΔ(r)Δ(r)i𝒌F)(u~n,(r)v~n,(r))=E~n+(u~n,(r)v~n,(r)),𝑖subscript𝒌Fbold-∇ΔrsuperscriptΔr𝑖subscript𝒌Fbold-∇subscript~𝑢𝑛rmissing-subexpressionsubscript~𝑣𝑛rmissing-subexpressionsubscript~𝐸limit-from𝑛subscript~𝑢𝑛rmissing-subexpressionsubscript~𝑣𝑛rmissing-subexpression\left(\begin{array}[]{cc}-i\bm{k}_{\textrm{F}}\cdot\bm{\nabla}&\Delta(\textbf{% r})\\ \Delta^{*}(\textbf{r})&i\bm{k}_{\textrm{F}}\cdot\bm{\nabla}\end{array}\right)% \left(\begin{array}[]{cc}\tilde{u}_{n,\uparrow}(\textbf{r})\\ \tilde{v}_{n,\downarrow}(\textbf{r})\end{array}\right)=\tilde{E}_{n+}\left(% \begin{array}[]{cc}\tilde{u}_{n,\uparrow}(\textbf{r})\\ \tilde{v}_{n,\downarrow}(\textbf{r})\end{array}\right),( start_ARRAY start_ROW start_CELL - italic_i bold_italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT ⋅ bold_∇ end_CELL start_CELL roman_Δ ( r ) end_CELL end_ROW start_ROW start_CELL roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( r ) end_CELL start_CELL italic_i bold_italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT ⋅ bold_∇ end_CELL end_ROW end_ARRAY ) ( start_ARRAY start_ROW start_CELL over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( r ) end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( r ) end_CELL start_CELL end_CELL end_ROW end_ARRAY ) = over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n + end_POSTSUBSCRIPT ( start_ARRAY start_ROW start_CELL over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( r ) end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL over~ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( r ) end_CELL start_CELL end_CELL end_ROW end_ARRAY ) , (11)

where E~n+=En++Δμ2subscript~𝐸limit-from𝑛subscript𝐸limit-from𝑛Δ𝜇2\tilde{E}_{n+}=E_{n+}+\frac{\Delta\mu}{2}over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n + end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_n + end_POSTSUBSCRIPT + divide start_ARG roman_Δ italic_μ end_ARG start_ARG 2 end_ARG. The second pair of equations for u~n,(r)subscript~𝑢𝑛r\tilde{u}_{n,\downarrow}(\textbf{r})over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_n , ↓ end_POSTSUBSCRIPT ( r ) and v~n,(r)subscript~𝑣𝑛r\tilde{v}_{n,\uparrow}(\textbf{r})over~ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n , ↑ end_POSTSUBSCRIPT ( r ) has similar form and correspond to E~n=EnΔμ2subscript~𝐸limit-from𝑛subscript𝐸limit-from𝑛Δ𝜇2\tilde{E}_{n-}=E_{n-}-\frac{\Delta\mu}{2}over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n - end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_n - end_POSTSUBSCRIPT - divide start_ARG roman_Δ italic_μ end_ARG start_ARG 2 end_ARG.

Although, Andreev approximation requires that |Δ|/εF1much-less-thanΔsubscript𝜀𝐹1|\Delta|/\varepsilon_{F}\ll 1| roman_Δ | / italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ≪ 1 (εFsubscript𝜀𝐹\varepsilon_{F}italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT denotes the Fermi energy), the spectrum of low-lying states can be qualitatively reproduced using a schematic model with pairing field defined as Δ(r,ϕ)=|Δ|eiϕθ(rrv)Δ𝑟italic-ϕΔsuperscript𝑒𝑖italic-ϕ𝜃𝑟subscript𝑟𝑣\Delta(r,\phi)=|\Delta|e^{i\phi}\theta(r-r_{v})roman_Δ ( italic_r , italic_ϕ ) = | roman_Δ | italic_e start_POSTSUPERSCRIPT italic_i italic_ϕ end_POSTSUPERSCRIPT italic_θ ( italic_r - italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ) and shown in Fig. 1(a). The reason is that, the most important ingredient of the model is related to the particular phase pattern of the pairing potential, whereas the asymptotic behavior of the wave functions play only a minor role (for a comprehensive discussion of the quality of Andreev approximation see Ref. [37]).

Refer to caption
Figure 1: A schematic picture of sections through the vortex core. Section perpendicular to the vortex line is shown in the left subfigure a). The classical trajectory representing particle of momentum kFsubscript𝑘𝐹k_{F}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT is denoted by a red solid line, and the retroreflected hole is shown as a brown dashed line. Note that the angular momentum component Lzsubscript𝐿𝑧L_{z}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT corresponding to the trajectory is negative, and the vortex rotates counterclockwise. Section along to the vortex line is shown in the right subfigure b). The classical trajectory representing particle of momentum kpsubscript𝑘𝑝k_{p}italic_k start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is denoted by a red solid line, and the reflected hole of momentum khsubscript𝑘k_{h}italic_k start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT is shown as a brown dashed line.

Then, the quantization condition, originating from the perfect particle-hole retroreflection, reads [38]:

EnεFkFrv1(LzkFrv)2+arccos(LzkFrv)arccosEn|Δ|=πn,subscript𝐸𝑛subscript𝜀𝐹subscript𝑘Fsubscript𝑟𝑣1superscriptsubscript𝐿𝑧subscript𝑘Fsubscript𝑟𝑣2subscript𝐿𝑧subscript𝑘Fsubscript𝑟𝑣subscript𝐸𝑛Δ𝜋𝑛\displaystyle\frac{E_{n}}{\varepsilon_{F}}k_{\textrm{F}}r_{v}\sqrt{1-\left(% \frac{L_{z}}{k_{\textrm{F}}r_{v}}\right)^{2}}+\arccos{\left(\frac{-L_{z}}{k_{% \textrm{F}}r_{v}}\right)}-\arccos{\frac{E_{n}}{|\Delta|}}=\pi n,divide start_ARG italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT square-root start_ARG 1 - ( divide start_ARG italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + roman_arccos ( divide start_ARG - italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG ) - roman_arccos divide start_ARG italic_E start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG | roman_Δ | end_ARG = italic_π italic_n , (12)

where n{0,±1,±2,n\in\{0,\pm 1,\pm 2,\dotsitalic_n ∈ { 0 , ± 1 , ± 2 , …}, rvsubscript𝑟𝑣r_{v}italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT denotes the radius of the vortex core, kFsubscript𝑘Fk_{\textrm{F}}italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT - Fermi momentum, and |Lz|=rkFsubscript𝐿𝑧𝑟subscript𝑘F|L_{z}|=rk_{\textrm{F}}| italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT | = italic_r italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT. Note that only the states with n=0𝑛0n=0italic_n = 0 correspond to core states, i.e. En=0|Δ|less-than-or-similar-tosubscript𝐸𝑛0ΔE_{n=0}\lesssim|\Delta|italic_E start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT ≲ | roman_Δ |111While solving the full BdG eqs. the states corresponding to n0𝑛0n\neq 0italic_n ≠ 0 can still be found at energies smaller than |Δ|Δ|\Delta|| roman_Δ |, but they are very close to continuum.. The limit |E||Δ|much-less-than𝐸Δ|E|\ll|\Delta|| italic_E | ≪ | roman_Δ | can be quite accurately approximated by the expression:

E±,n=0,m|Δ|2εFrvξ(rvξ+1)m,subscript𝐸formulae-sequenceplus-or-minus𝑛0𝑚superscriptΔ2subscript𝜀𝐹subscript𝑟𝑣𝜉subscript𝑟𝑣𝜉1𝑚E_{\pm,n=0,m}\approx-\frac{|\Delta|^{2}}{\varepsilon_{F}\frac{r_{v}}{\xi}\left% (\frac{r_{v}}{\xi}+1\right)}m,italic_E start_POSTSUBSCRIPT ± , italic_n = 0 , italic_m end_POSTSUBSCRIPT ≈ - divide start_ARG | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT divide start_ARG italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_ξ end_ARG ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_ξ end_ARG + 1 ) end_ARG italic_m , (13)

where m𝑚mitalic_m is the magnetic quantum number associated with Lz=msubscript𝐿𝑧Planck-constant-over-2-pi𝑚L_{z}=\hbar mitalic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = roman_ℏ italic_m, pointing along the vortex axis and ξ=εF/(kF|Δ|)𝜉subscript𝜀𝐹subscript𝑘FΔ\xi=\varepsilon_{F}/(k_{\textrm{F}}|\Delta|)italic_ξ = italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT / ( italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT | roman_Δ | ) is a coherence length. Minus sign in front of the rhs expression (13) is related to the counterclockwise rotation of the vortex. Linear dependence ELzproportional-to𝐸subscript𝐿𝑧E\propto L_{z}italic_E ∝ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT holds until the energy approaches ΔΔ\Deltaroman_Δ, and subsequently, it is bent towards large angular momenta. This part of the spectrum cannot be reproduced within Andreev approximation. In the regime close to unitarity and the inner crust of neutron stars, the linear part of the spectrum consists of a few states only, as shown in the inset of Fig. 2. The results presented in this figure have been obtained by solving numerically Hartree-Fock-Bogoliubov equations (25) (having the same form as eqs. (3)) on the spatial lattice (in 3D) for the pure neutron matter. The BSk energy density functional with the local pairing field Δ(𝒓)Δ𝒓\Delta(\boldsymbol{r})roman_Δ ( bold_italic_r ) has been used (see Section 3). The resulting profile of the pairing field and the neutron density distribution ρ(𝒓)𝜌𝒓\rho(\boldsymbol{r})italic_ρ ( bold_italic_r ), have been obtained selfconsistently through the total energy minimization, with the requirement that the pairing field possess the structure characteristic for the vortex located in the center of the cylindrical box. Thus the results represent more realistic structure of the vortex core. The quasiparticle energies as a function of the angular momentum along the vortex symmetry axis have been plotted in the inset. The plotted energies correspond to kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 in expression (10).

Refer to caption
Figure 2: The cross-section through the center of the vortex in neutron matter of bulk density ρ=0.0059subscript𝜌0.0059\rho_{\infty}=0.0059italic_ρ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 0.0059 fm-3 and corresponding bulk pairing gap Δ=1.33subscriptΔ1.33\Delta_{\infty}=1.33roman_Δ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = 1.33 MeV. The parameters correspond to the inner crust region of a neutron star. The cross-section of (a) density and pairing field (b) (both quantities normalized to the corresponding bulk value). Two characteristic length scales are shown: inverse Fermi momentum kF1superscriptsubscript𝑘𝐹1k_{F}^{-1}italic_k start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and the superfluid coherence length ξ𝜉\xiitalic_ξ. Inset in (b): the energy spectrum of quantum states (of particle and hole types) and their quantized angular momenta (the relative shift of angular momenta between particles and holes is the consequence of relation (10)). The energy scales governing the physics of the system — bulk energy gap ΔsubscriptΔ\Delta_{\infty}roman_Δ start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT, as well as so-called minigap Emgsubscript𝐸𝑚𝑔E_{mg}italic_E start_POSTSUBSCRIPT italic_m italic_g end_POSTSUBSCRIPT — are shown. The results are obtained with DFT framework for nuclear systems [28].

Note that in the inner crust of the neutron star, in the region where the s-wave pairing is predicted to be the strongest, i.e. with the value of the pairing gap within the interval (1.2,1.6)1.21.6(1.2,1.6)( 1.2 , 1.6 ) MeV, the size of the vortex core is much smaller (at T0.1less-than-or-similar-to𝑇0.1T\lesssim 0.1italic_T ≲ 0.1 MeV) than the typical modulation of the nuclear density, induced by the presence of nuclear lattice. Namely, the radii of the vortex cores do not exceed 12121212 fm [28], whereas typical lattice constants in this region range between 40404040 and 70707070 fm [39]. Therefore, for the nuclear densities (0.0060.034)0.0060.034(0.006-0.034)( 0.006 - 0.034 ) fm-3 it is possible to place the vortex core in the region, where the matter is uniform. The situation is different, however, if one approaches the boundaries of the inner crust. In the limit of small densities, at the border of the outer crust region, the lattice constant reaches values of about 100100100100 fm, whereas the size of the vortex cores can be arbitrarily large due to the vanishing density of the neutron gas outside nuclei. The similar situation occurs in the region, where pasta phases are predicted to exist, i.e. at the densities 0.050.050.050.05 fm-3 and higher. The typical length scale of nuclear density modulation is about 20202020 fm and the vortex core can be of similar size. In these cases one has to take into account that the density inside the vortex core is in addition modulated due to the particular nuclear structure, independent of the shape of the pairing field associated with the vortex. This effect will in turn introduce additional modification of the structure of the in-gap states, due to the scattering of quasiparticles on nuclear inhomogeneities. Moreover, in these region, the scattering of quasiparticles (inside the vortex core) on nuclear inhomogeneities may provide another efficient source of energy dissipation for a moving vortex (see discussion in Section 3).

There are few immediate consequences to the existence of the so-called chiral branch corresponding to n=0𝑛0n=0italic_n = 0. Since the quasiparticle excitation spectrum consists of states with m<0𝑚0m<0italic_m < 0 (for a vortex rotating counterclockwise), the energy of the first excited state Emg|Δ|22εFsubscript𝐸𝑚𝑔superscriptΔ22subscript𝜀𝐹E_{mg}\approx\frac{|\Delta|^{2}}{2\varepsilon_{F}}italic_E start_POSTSUBSCRIPT italic_m italic_g end_POSTSUBSCRIPT ≈ divide start_ARG | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG (assuming rv=ξsubscript𝑟𝑣𝜉r_{v}=\xiitalic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = italic_ξ) sets the energy scale (minigap energy) associated with the excitation of the vortex core. Moreover, the chiral branch’s structure determines the density behavior in the vortex core. Specifically, since the m=0𝑚0m=0italic_m = 0 state is unoccupied, the density reveals depletion in the core. It can be deduced from BdG eqs. that density ρ(r)r2+constproportional-to𝜌𝑟superscript𝑟2const\rho(r)\propto r^{2}+\textrm{const}italic_ρ ( italic_r ) ∝ italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + const in the vicinity of the center of the vortex. The vortex core structure, obtained within the framework of Density Functional Theory (see section 3), is shown in Fig. 2. It has to be emphasized that the depletion of the density inside a vortex core is crucial for the experimental detection of vortices in ultracold Fermi gases.

The presence of a chiral branch has an impact on thermodynamic properties as well. For temperatures within the range EmgkBT|Δ|much-less-thansubscript𝐸𝑚𝑔subscript𝑘𝐵𝑇much-less-thanΔE_{mg}\ll k_{B}T\ll|\Delta|italic_E start_POSTSUBSCRIPT italic_m italic_g end_POSTSUBSCRIPT ≪ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ≪ | roman_Δ |, it was predicted [33] that the specific heat should be a linear function of temperature. It is, however, true in deep BCS limit only (ΔεFmuch-less-thanΔsubscript𝜀𝐹\Delta\ll\varepsilon_{F}roman_Δ ≪ italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT). Due to the limited number of in-gap states in strongly interacting systems, the temperature range fulfilling the above criterion may not exist. Moreover, the dependence of the size of the core on temperature makes the functional dependence of the specific heat more complicated [28]. Another modification of the specific heat comes from the superflow surrounding a vortex due to the rearrangement of the quasiparticle spectrum induced by the flow. Namely, the quasiparticle dispersion relation in the presence of superflow with velocity vssubscript𝑣𝑠v_{s}italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT reads:

E±(𝒌,𝒗𝒔)=𝒌𝒗𝒔±(2k22Mμ~)2+|Δ|2,subscript𝐸plus-or-minus𝒌subscript𝒗𝒔plus-or-minusPlanck-constant-over-2-pi𝒌subscript𝒗𝒔superscriptsuperscriptPlanck-constant-over-2-pi2superscript𝑘22𝑀~𝜇2superscriptΔ2E_{\pm}(\boldsymbol{k},\boldsymbol{v_{s}})=\hbar\boldsymbol{k}\cdot\boldsymbol% {v_{s}}\pm\sqrt{\left(\frac{\hbar^{2}k^{2}}{2M}-\tilde{\mu}\right)^{2}+|\Delta% |^{2}},italic_E start_POSTSUBSCRIPT ± end_POSTSUBSCRIPT ( bold_italic_k , bold_italic_v start_POSTSUBSCRIPT bold_italic_s end_POSTSUBSCRIPT ) = roman_ℏ bold_italic_k ⋅ bold_italic_v start_POSTSUBSCRIPT bold_italic_s end_POSTSUBSCRIPT ± square-root start_ARG ( divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_M end_ARG - over~ start_ARG italic_μ end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (14)

where μ~=μ12Mvs2~𝜇𝜇12𝑀superscriptsubscript𝑣𝑠2\tilde{\mu}=\mu-\frac{1}{2}Mv_{s}^{2}over~ start_ARG italic_μ end_ARG = italic_μ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_M italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT represents the correction to the chemical potential due to the superflow. We consider the case kFvs/|Δ|1much-less-thanPlanck-constant-over-2-pisubscript𝑘Fsubscript𝑣𝑠Δ1\hbar k_{\textrm{F}}v_{s}/|\Delta|\ll 1roman_ℏ italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / | roman_Δ | ≪ 1 corresponding to velocities smaller than the critical velocity. In that case E+(𝒌,𝒗𝒔)>0subscript𝐸𝒌subscript𝒗𝒔0E_{+}(\boldsymbol{k},\boldsymbol{v_{s}})>0italic_E start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( bold_italic_k , bold_italic_v start_POSTSUBSCRIPT bold_italic_s end_POSTSUBSCRIPT ) > 0 and one can determine the correction to the specific heat for the moving superfluid. Keeping terms up to vs2superscriptsubscript𝑣𝑠2v_{s}^{2}italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT one gets:

CV(vs)2T2VN(0)2πT|Δ|exp(|Δ|T)subscript𝐶𝑉subscript𝑣𝑠2superscript𝑇2𝑉𝑁02𝜋𝑇ΔΔ𝑇\displaystyle C_{V}(v_{s})\approx\frac{2}{T^{2}}VN(0)\sqrt{2\pi T|\Delta|}\exp% \left(-\frac{|\Delta|}{T}\right)italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≈ divide start_ARG 2 end_ARG start_ARG italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_V italic_N ( 0 ) square-root start_ARG 2 italic_π italic_T | roman_Δ | end_ARG roman_exp ( - divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG )
×[|Δ|2+μMvs23((|Δ|T)24|Δ|T+234(|Δ|μ)2)2],absentdelimited-[]superscriptΔ2𝜇𝑀superscriptsubscript𝑣𝑠23superscriptsuperscriptΔ𝑇24Δ𝑇234superscriptΔ𝜇22\displaystyle\times\left[|\Delta|^{2}+\frac{\mu Mv_{s}^{2}}{3}\left(\left(% \frac{|\Delta|}{T}\right)^{2}-4\frac{|\Delta|}{T}+2-\frac{3}{4}\left(\frac{|% \Delta|}{\mu}\right)^{2}\right)^{2}\right],× [ | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG italic_μ italic_M italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG ( ( divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 4 divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG + 2 - divide start_ARG 3 end_ARG start_ARG 4 end_ARG ( divide start_ARG | roman_Δ | end_ARG start_ARG italic_μ end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] , (15)

where

N(0)=1(2π)2(2M2)3/2μ𝑁01superscript2𝜋2superscript2𝑀superscriptPlanck-constant-over-2-pi232𝜇N(0)=\frac{1}{(2\pi)^{2}}\left(\frac{2M}{\hbar^{2}}\right)^{3/2}\sqrt{\mu}italic_N ( 0 ) = divide start_ARG 1 end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( divide start_ARG 2 italic_M end_ARG start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT square-root start_ARG italic_μ end_ARG (16)

is the density of states at the Fermi surface. The last term, proportional to (|Δ|/μ)2superscriptΔ𝜇2\left(|\Delta|/\mu\right)^{2}( | roman_Δ | / italic_μ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, is a correction due to the modification of the chemical potential μ~~𝜇\tilde{\mu}over~ start_ARG italic_μ end_ARG in the presence of superflow. It is, however, at least an order of magnitude smaller than the other terms and can be neglected. One may apply this formula to the flow induced by a vortex within the local density approximation. Namely, one can divide the volume around the vortex line into infinitesimal concentric cylindrical shells (of length L𝐿Litalic_L), in which the superfluid velocity is constant, and subsequently integrate the contributions to the specific heat coming from each shell. For the cylinder of radius Routsubscript𝑅𝑜𝑢𝑡R_{out}italic_R start_POSTSUBSCRIPT italic_o italic_u italic_t end_POSTSUBSCRIPT one gets the correction to specific heat per unit length of the vortex:

ΔCVflowLΔsuperscriptsubscript𝐶𝑉𝑓𝑙𝑜𝑤𝐿absent\displaystyle\frac{\Delta C_{V}^{flow}}{L}\approxdivide start_ARG roman_Δ italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f italic_l italic_o italic_w end_POSTSUPERSCRIPT end_ARG start_ARG italic_L end_ARG ≈ 13(2π)3/2N(0)|Δ|TμT22Mln(Routrv)[(|Δ|T)24|Δ|T+2]exp(|Δ|T).13superscript2𝜋32𝑁0Δ𝑇𝜇𝑇superscriptPlanck-constant-over-2-pi22𝑀subscript𝑅𝑜𝑢𝑡subscript𝑟𝑣delimited-[]superscriptΔ𝑇24Δ𝑇2Δ𝑇\displaystyle\frac{1}{3}(2\pi)^{3/2}N(0)\sqrt{\frac{|\Delta|}{T}}\frac{\mu}{T}% \frac{\hbar^{2}}{2M}\ln\left(\frac{R_{out}}{r_{v}}\right)\left[\left(\frac{|% \Delta|}{T}\right)^{2}-4\frac{|\Delta|}{T}+2\right]\exp\left(-\frac{|\Delta|}{% T}\right).divide start_ARG 1 end_ARG start_ARG 3 end_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT italic_N ( 0 ) square-root start_ARG divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG end_ARG divide start_ARG italic_μ end_ARG start_ARG italic_T end_ARG divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_M end_ARG roman_ln ( divide start_ARG italic_R start_POSTSUBSCRIPT italic_o italic_u italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG ) [ ( divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 4 divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG + 2 ] roman_exp ( - divide start_ARG | roman_Δ | end_ARG start_ARG italic_T end_ARG ) . (17)

As discussed in Ref. [28], Routsubscript𝑅𝑜𝑢𝑡R_{out}italic_R start_POSTSUBSCRIPT italic_o italic_u italic_t end_POSTSUBSCRIPT is of the order of inter-vortex distance in the case of vortex lattice.

Consequently, the modification of the specific heat per unit volume, due to the presence of vortices, reads:

1VΔCV=σvor(ΔCVcoreL+ΔCVflowL),1𝑉Δsubscript𝐶𝑉subscript𝜎𝑣𝑜𝑟Δsuperscriptsubscript𝐶𝑉𝑐𝑜𝑟𝑒𝐿Δsuperscriptsubscript𝐶𝑉𝑓𝑙𝑜𝑤𝐿\frac{1}{V}\Delta C_{V}=\sigma_{vor}\left(\frac{\Delta C_{V}^{core}}{L}+\frac{% \Delta C_{V}^{flow}}{L}\right),divide start_ARG 1 end_ARG start_ARG italic_V end_ARG roman_Δ italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT italic_v italic_o italic_r end_POSTSUBSCRIPT ( divide start_ARG roman_Δ italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c italic_o italic_r italic_e end_POSTSUPERSCRIPT end_ARG start_ARG italic_L end_ARG + divide start_ARG roman_Δ italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f italic_l italic_o italic_w end_POSTSUPERSCRIPT end_ARG start_ARG italic_L end_ARG ) , (18)

where σvor=Nvor/Ssubscript𝜎𝑣𝑜𝑟subscript𝑁𝑣𝑜𝑟𝑆\sigma_{vor}=N_{vor}/Sitalic_σ start_POSTSUBSCRIPT italic_v italic_o italic_r end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_v italic_o italic_r end_POSTSUBSCRIPT / italic_S (and V=S×L𝑉𝑆𝐿V=S\times Litalic_V = italic_S × italic_L) is the surface vortex density and the two terms: ΔCVcoreΔsuperscriptsubscript𝐶𝑉𝑐𝑜𝑟𝑒\Delta C_{V}^{core}roman_Δ italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c italic_o italic_r italic_e end_POSTSUPERSCRIPT, ΔCVflowΔsuperscriptsubscript𝐶𝑉𝑓𝑙𝑜𝑤\Delta C_{V}^{flow}roman_Δ italic_C start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_f italic_l italic_o italic_w end_POSTSUPERSCRIPT represent modification of the specific heat due to the vortex core and the flow induced by the vortex, respectively. Due to small average vortex density σvor1021subscript𝜎𝑣𝑜𝑟superscript1021\sigma_{vor}\approx 10^{-21}italic_σ start_POSTSUBSCRIPT italic_v italic_o italic_r end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 21 end_POSTSUPERSCRIPT fm-2, the modification of the specific heat of the neutron star crust is negligible for the temperatures T0.1greater-than-or-approximately-equals𝑇0.1T\gtrapprox 0.1italic_T ⪆ 0.1 MeV. However, if the vortex lattice is perturbed, giving rise to the turbulent-like behavior, one can expect that the vortex density will exhibit significant spatial fluctuations [40]. Using the results of Ref. [28] (see Fig. 10 therein) one can estimate that the local increase of vortex density to σvor102subscript𝜎𝑣𝑜𝑟superscript102\sigma_{vor}\approx 10^{-2}italic_σ start_POSTSUBSCRIPT italic_v italic_o italic_r end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT fm-2 leads to approximately twice larger specific heat at T0.1𝑇0.1T\approx 0.1italic_T ≈ 0.1 MeV for a wide range of neutron densities in the crust (0.0060.034)0.0060.034(0.006-0.034)( 0.006 - 0.034 ) fm-3. For even lower T𝑇Titalic_T, specific heat will increase exponentially.

The core states, discussed above, correspond to 2D vortex, i.e., they are characterized by kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0, where kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is a momentum along the vortex line. In 3D, corresponding to the infinite vortex line, each state forming the chiral branch represents a band. The slope of the band is directly related to the effective mass of quasiparticle excitations, and it can be traced back to the quasiparticle scattering properties along the vortex line. In Fig. 1(b), the schematic picture of Andreev scattering, leading to the motion of quasiparticles along the vortex line, has been shown. Contrary to the quantization condition, which resulted from the assumption that the hole(particle) is reflected exactly backward (which is true if the incoming particle(hole) is at the Fermi surface), here we need to release this constraint to determine the effective mass. Consequently one needs to take into account corrections related to εF±Eplus-or-minussubscript𝜀𝐹𝐸\varepsilon_{F}\pm Eitalic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ± italic_E. As a result of the momentum conservation along the vortex line, the reflection law reads: εF+Esinα=εFEsinβsubscript𝜀𝐹𝐸𝛼subscript𝜀𝐹𝐸𝛽\sqrt{\varepsilon_{F}+E}\sin\alpha=\sqrt{\varepsilon_{F}-E}\sin\betasquare-root start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT + italic_E end_ARG roman_sin italic_α = square-root start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT - italic_E end_ARG roman_sin italic_β, where kp=2(εF+E)subscript𝑘𝑝2subscript𝜀𝐹𝐸k_{p}=\sqrt{2(\varepsilon_{F}+E)}italic_k start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = square-root start_ARG 2 ( italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT + italic_E ) end_ARG and kh=2(εFE)subscript𝑘2subscript𝜀𝐹𝐸k_{h}=\sqrt{2(\varepsilon_{F}-E)}italic_k start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT = square-root start_ARG 2 ( italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT - italic_E ) end_ARG are particle and hole momenta, respectively. The effective velocity along the vortex line can be evaluated as vz=SΔT=2(εF+E)sinαsin(βα)/sin(β+α)subscript𝑣𝑧𝑆Δ𝑇2subscript𝜀𝐹𝐸𝛼𝛽𝛼𝛽𝛼v_{z}=\frac{S}{\Delta T}=\sqrt{2(\varepsilon_{F}+E)}\sin\alpha\sin(\beta-% \alpha)/\sin(\beta+\alpha)italic_v start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = divide start_ARG italic_S end_ARG start_ARG roman_Δ italic_T end_ARG = square-root start_ARG 2 ( italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT + italic_E ) end_ARG roman_sin italic_α roman_sin ( italic_β - italic_α ) / roman_sin ( italic_β + italic_α ), where ΔTΔ𝑇\Delta Troman_Δ italic_T denotes the time interval between two consecutive particle-to-hole reflections. Finally one gets [38]:

vz=kzkp2kz2kh2kz2kp2kz2+kh2kz2,subscript𝑣𝑧subscript𝑘𝑧superscriptsubscript𝑘𝑝2superscriptsubscript𝑘𝑧2superscriptsubscript𝑘2superscriptsubscript𝑘𝑧2superscriptsubscript𝑘𝑝2superscriptsubscript𝑘𝑧2superscriptsubscript𝑘2superscriptsubscript𝑘𝑧2v_{z}=k_{z}\frac{\sqrt{k_{p}^{2}-k_{z}^{2}}-\sqrt{k_{h}^{2}-k_{z}^{2}}}{\sqrt{% k_{p}^{2}-k_{z}^{2}}+\sqrt{k_{h}^{2}-k_{z}^{2}}},italic_v start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT divide start_ARG square-root start_ARG italic_k start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - square-root start_ARG italic_k start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG square-root start_ARG italic_k start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + square-root start_ARG italic_k start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG , (19)

where kz=kpsinα=khsinβsubscript𝑘𝑧subscript𝑘𝑝𝛼subscript𝑘𝛽k_{z}=k_{p}\sin\alpha=k_{h}\sin\betaitalic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_sin italic_α = italic_k start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT roman_sin italic_β is the momentum component along the vortex line. Considering the linear term in kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and E𝐸Eitalic_E on the rhs, one obtains the effective mass Meff1E/2εFsuperscriptsubscript𝑀eff1𝐸2subscript𝜀𝐹M_{\textrm{eff}}^{-1}\approx E/2\varepsilon_{F}italic_M start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ≈ italic_E / 2 italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT, which agrees with the effective mass derived from the formula for the dispersion relation in the BCS limit E(kz)=E(0)/1kz2/(2εF)𝐸subscript𝑘𝑧𝐸01superscriptsubscript𝑘𝑧22subscript𝜀𝐹E(k_{z})=E(0)/\sqrt{1-k_{z}^{2}/(2\varepsilon_{F})}italic_E ( italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) = italic_E ( 0 ) / square-root start_ARG 1 - italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) end_ARG [33]. One may also estimate the magnitude of the effective mass component along the vortex line, corresponding to angular momentum Lz=msubscript𝐿𝑧Planck-constant-over-2-pi𝑚L_{z}=\hbar mitalic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = roman_ℏ italic_m: Meff1(m)2|m|3(ΔεF)2superscriptsubscript𝑀eff1𝑚2𝑚3superscriptΔsubscript𝜀𝐹2M_{\textrm{eff}}^{-1}(m)\approx\frac{2|m|}{3}\left(\frac{\Delta}{\varepsilon_{% F}}\right)^{2}italic_M start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_m ) ≈ divide start_ARG 2 | italic_m | end_ARG start_ARG 3 end_ARG ( divide start_ARG roman_Δ end_ARG start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Note that, in deep BCS limit, the inverse of the effective mass will be exponentially small since Δ/εFeπ/2|a|kFproportional-toΔsubscript𝜀𝐹superscript𝑒𝜋2𝑎subscript𝑘F\Delta/\varepsilon_{F}\propto e^{-\pi/2|a|k_{\textrm{F}}}roman_Δ / italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ∝ italic_e start_POSTSUPERSCRIPT - italic_π / 2 | italic_a | italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, and clearly, the departure from the flat band behavior will be significant at the unitarity, where Δ/εF0.5Δsubscript𝜀𝐹0.5\Delta/\varepsilon_{F}\approx 0.5roman_Δ / italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ≈ 0.5. The band flatness, increasing the effective mass, will affect the propagation of the confined quasiparticle excitations, e.g., in the form of local spin-polarization, along the vortex line. It was shown in Ref. [41] that the locally induced spin-polarization in the vortex core (e.g., due to the reconnection process with spin-polarized vortex) can hardly propagate along the vortex line. Indeed, the motion along the vortex line is characterized by the velocity vz=k0/Meffk0(ΔεF)2subscript𝑣𝑧subscript𝑘0subscript𝑀effproportional-tosubscript𝑘0superscriptΔsubscript𝜀𝐹2v_{z}=k_{0}/M_{\textrm{eff}}\propto k_{0}\left(\frac{\Delta}{\varepsilon_{F}}% \right)^{2}italic_v start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ∝ italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( divide start_ARG roman_Δ end_ARG start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where k0subscript𝑘0k_{0}italic_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial momentum of the wave packet. Similarly, the wave packet width, in the limit of long times, behaves as (zvzt)2t(ΔεF)2proportional-todelimited-⟨⟩superscript𝑧subscript𝑣𝑧𝑡2𝑡superscriptΔsubscript𝜀𝐹2\sqrt{\langle(z-v_{z}t)^{2}\rangle}\propto t\left(\frac{\Delta}{\varepsilon_{F% }}\right)^{2}square-root start_ARG ⟨ ( italic_z - italic_v start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_ARG ∝ italic_t ( divide start_ARG roman_Δ end_ARG start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and leads to an effective suppression of the polarization propagation, see Ref. [38].

The structure of the vortex core is sensitive to spin imbalance. When the chemical potentials for spin-up (μsubscript𝜇\mu_{\uparrow}italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT) and spin-down (μsubscript𝜇\mu_{\downarrow}italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT) fermions begin to differ, the core of the vortex will be affected first, before the superfluid will be modified in bulk. This is due to the fact that Emg<|Δ|subscript𝐸𝑚𝑔ΔE_{mg}<|\Delta|italic_E start_POSTSUBSCRIPT italic_m italic_g end_POSTSUBSCRIPT < | roman_Δ |. Indeed, unpaired fermions tend to accumulate at the core, as shown in [42, 43]. Consequently, the chiral branch splits into two components corresponding to spin-up and spin-down particles. The main features of this splitting can be described introducing a slight modification of the expression (12), which reads:

E~n±εFkFrv1(LzkFrv)2+arccos(LzkFrv)arccosE~n±|Δ|=πn,subscript~𝐸limit-from𝑛plus-or-minussubscript𝜀𝐹subscript𝑘Fsubscript𝑟𝑣1superscriptsubscript𝐿𝑧subscript𝑘Fsubscript𝑟𝑣2subscript𝐿𝑧subscript𝑘Fsubscript𝑟𝑣subscript~𝐸limit-from𝑛plus-or-minusΔ𝜋𝑛\displaystyle\frac{\tilde{E}_{n\pm}}{\varepsilon_{F}}k_{\textrm{F}}r_{v}\sqrt{% 1-\left(\frac{L_{z}}{k_{\textrm{F}}r_{v}}\right)^{2}}+\arccos{\left(\frac{-L_{% z}}{k_{\textrm{F}}r_{v}}\right)}-\arccos{\frac{\tilde{E}_{n\pm}}{|\Delta|}}=% \pi n,divide start_ARG over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n ± end_POSTSUBSCRIPT end_ARG start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT square-root start_ARG 1 - ( divide start_ARG italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + roman_arccos ( divide start_ARG - italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG ) - roman_arccos divide start_ARG over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n ± end_POSTSUBSCRIPT end_ARG start_ARG | roman_Δ | end_ARG = italic_π italic_n , (20)

where E~n±=En±±Δμ2subscript~𝐸limit-from𝑛plus-or-minusplus-or-minussubscript𝐸limit-from𝑛plus-or-minusΔ𝜇2\tilde{E}_{n\pm}=E_{n\pm}\pm\frac{\Delta\mu}{2}over~ start_ARG italic_E end_ARG start_POSTSUBSCRIPT italic_n ± end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_n ± end_POSTSUBSCRIPT ± divide start_ARG roman_Δ italic_μ end_ARG start_ARG 2 end_ARG and Δμ=μμΔ𝜇subscript𝜇subscript𝜇\Delta\mu=\mu_{\uparrow}-\mu_{\downarrow}roman_Δ italic_μ = italic_μ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT is the difference between chemical potentials of spin-up and spin-down particles, respectively. A schematic representation of the solutions corresponding to the chiral branch (n=0𝑛0n=0italic_n = 0) is shown in Fig. 3.

Refer to caption
Figure 3: Chiral branches (red solid line) of low-lying in-gap states in the vortex core for clockwise rotating vortex. Parts of branches that became empty and occupied due to induced spin imbalance have been shown in green and brown colors, respectively.

The immediate consequence of this splitting results in occupying states, by majority spin particles, which rotate in opposite directions than the vortex. The highest value of angular momentum corresponding to the opposite rotation Lzmax=moppositesuperscriptsubscript𝐿𝑧𝑚𝑎𝑥Planck-constant-over-2-pisubscript𝑚𝑜𝑝𝑝𝑜𝑠𝑖𝑡𝑒L_{z}^{max}=\hbar m_{opposite}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m italic_a italic_x end_POSTSUPERSCRIPT = roman_ℏ italic_m start_POSTSUBSCRIPT italic_o italic_p italic_p italic_o italic_s italic_i italic_t italic_e end_POSTSUBSCRIPT is related to spin imbalance and reads:

max|mopposite|12εF|Δ|2rvξ(rvξ+1)Δμ.subscript𝑚opposite12subscript𝜀𝐹superscriptΔ2subscript𝑟𝑣𝜉subscript𝑟𝑣𝜉1Δ𝜇\max|m_{\textrm{opposite}}|\approx\frac{1}{2}\frac{\varepsilon_{F}}{|\Delta|^{% 2}}\frac{r_{v}}{\xi}\left(\frac{r_{v}}{\xi}+1\right)\Delta\mu.roman_max | italic_m start_POSTSUBSCRIPT opposite end_POSTSUBSCRIPT | ≈ divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_ε start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT end_ARG start_ARG | roman_Δ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_ξ end_ARG ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_ξ end_ARG + 1 ) roman_Δ italic_μ . (21)

It is, therefore, clear that the reversed flow of these particles partly cancels the flow of the unpolarized component. However, as was shown in Ref.[38], the cancellation is not exact and leads to a reversal of the total flow in the core. The splitting of the chiral branch, induced by the spin imbalance, which rearranges the occupation of low-lying states, has another effect. Namely, the m=0𝑚0m=0italic_m = 0 state becomes now occupied, and the density is no longer depleted inside a polarized vortex. It makes the experimental detection of vortices in spin imbalanced system far more complex.

In the case of quantum vortices in the inner crust of a neutron star, the mechanism that may lead to the polarization of the vortex core is due to the magnetic field. It couples to the neutron magnetic moment and may induce a spin flip. In Ref. [28] the minimal value of the magnetic field required to produce such an excitation in the vortex core has been estimated to range between about 10151016Gsuperscript1015superscript1016𝐺10^{15}-10^{16}G10 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 16 end_POSTSUPERSCRIPT italic_G, depending on neutron density. Note, that this value is more than an order of magnitude smaller than magnetic field needed to destroy superfluidity in the neutron star crust, which was estimated as B1017G𝐵superscript1017𝐺B\geq 10^{17}Gitalic_B ≥ 10 start_POSTSUPERSCRIPT 17 end_POSTSUPERSCRIPT italic_G [44]. There are known magnetars where such conditions may exist [45].

The important effect of polarization in the vortex core is that it eventually causes the minigap to vanish. As was shown in Ref. [38], this is due to the fact that the spectrum of states with kz=0subscript𝑘𝑧0k_{z}=0italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0 is asymmetric with respect to Fermi surface (see Fig. 3). However, the spectrum of states corresponding to kzkFmuch-greater-thansubscript𝑘𝑧subscript𝑘Fk_{z}\gg k_{\textrm{F}}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≫ italic_k start_POSTSUBSCRIPT F end_POSTSUBSCRIPT becomes symmetric, which is apparent by inspecting the structure of the BdG Hamiltonian, describing the infinite vortex line. Therefore, one may infer that at certain values of kz=±kz1,±kz2,subscript𝑘𝑧plus-or-minussubscript𝑘𝑧1plus-or-minussubscript𝑘𝑧2k_{z}=\pm k_{z1},\pm k_{z2},\dotsitalic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = ± italic_k start_POSTSUBSCRIPT italic_z 1 end_POSTSUBSCRIPT , ± italic_k start_POSTSUBSCRIPT italic_z 2 end_POSTSUBSCRIPT , … the quasiparticle energies vanishes E(±kzi)=0𝐸plus-or-minussubscript𝑘𝑧𝑖0E(\pm k_{zi})=0italic_E ( ± italic_k start_POSTSUBSCRIPT italic_z italic_i end_POSTSUBSCRIPT ) = 0. When the quasiparticle energy changes from negative to positive value, the particle state vsubscript𝑣v_{\uparrow}italic_v start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT with momentum m𝑚mitalic_m is converted into the hole state usubscript𝑢u_{\uparrow}italic_u start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT with momentum m+1𝑚1-m+1- italic_m + 1, inducing the change of m𝑚mitalic_m: Δm=|2m1|Δ𝑚2𝑚1\Delta m=|2m-1|roman_Δ italic_m = | 2 italic_m - 1 |. Therefore, the spin-imbalanced vortex is characterized by a series of quasiparticle level crossings at the Fermi surface, i.e., points at which the minigap vanishes. Eventually, if the spin imbalance increases, it is predicted that modulation of the pairing field, similar to the Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) phase, will be seen [46].

All these characteristics of the fermionic vortex core structure should have an impact on dynamics and, in particular, on dissipative processes associated with the vortex motion. These will be discussed in the next section.

3 Vortex dynamics

The problem of vortex dynamics is essential for understanding the behavior of superfluids, particularly the role of dissipative processes in the decay of quantum turbulence [47, 48]. The apparent differences between the structure of bosonic and fermionic vortices, discussed in the previous section, raise a question about the impact on vortex dynamic. The existence of another energy scale associated with minigap and the presence of the chiral branch is expected to generate additional contributions to the dissipative force. Last but not least, there is still an open question associated with vortex inertia, which in the case of bosonic vortices is usually assumed to be negligible [49].

Answers to these questions should result in formulating the effective equation of motion of a fermionic vortex, where all terms will acquire microscopic underpinning. If we consider a 2D vortex, i.e., neglecting degrees of freedom associated with the susceptibility of vortex lines to bending and generating Kelvin waves, the problem is still complex as the most general equation of motion comprises the following terms:

MVd2𝑹Vdt2=γd𝑹Vdt+ωd𝑹Vdt×𝒆z+𝑭superfluid+𝑭pinning,subscript𝑀𝑉superscript𝑑2subscript𝑹𝑉𝑑superscript𝑡2𝛾𝑑subscript𝑹𝑉𝑑𝑡𝜔𝑑subscript𝑹𝑉𝑑𝑡subscript𝒆𝑧subscript𝑭superfluidsubscript𝑭pinningM_{V}\frac{d^{2}\boldsymbol{R}_{V}}{dt^{2}}=\gamma\frac{d\boldsymbol{R}_{V}}{% dt}+\omega\frac{d\boldsymbol{R}_{V}}{dt}\times\boldsymbol{e}_{z}+\boldsymbol{F% }_{\textrm{superfluid}}+\boldsymbol{F}_{\textrm{pinning}},italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = italic_γ divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG + italic_ω divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG × bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + bold_italic_F start_POSTSUBSCRIPT superfluid end_POSTSUBSCRIPT + bold_italic_F start_POSTSUBSCRIPT pinning end_POSTSUBSCRIPT , (22)

where 𝑹Vsubscript𝑹𝑉\boldsymbol{R}_{V}bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT is the vortex position, 𝒆zsubscript𝒆𝑧\boldsymbol{e}_{z}bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is the unit vector perpendicular to the plane of vortex motion, 𝑭pinningsubscript𝑭pinning\boldsymbol{F}_{\textrm{pinning}}bold_italic_F start_POSTSUBSCRIPT pinning end_POSTSUBSCRIPT describes the interaction with inhomogeneities and 𝑭superfluidsubscript𝑭superfluid\boldsymbol{F}_{\textrm{superfluid}}bold_italic_F start_POSTSUBSCRIPT superfluid end_POSTSUBSCRIPT is the component of the force acting on a vortex but independent of its velocity. The equation arises from the general expression of forces that can be present in the system and being in agreement with Galilean invariance. The latter requirement implies that the force either depends on (d𝑹Vdt𝒗s)𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑠(\frac{d\boldsymbol{R}_{V}}{dt}-\boldsymbol{v}_{s})( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) or on (d𝑹Vdt𝒗n)𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑛(\frac{d\boldsymbol{R}_{V}}{dt}-\boldsymbol{v}_{n})( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ), where 𝒗ssubscript𝒗𝑠\boldsymbol{v}_{s}bold_italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 𝒗nsubscript𝒗𝑛\boldsymbol{v}_{n}bold_italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT describe the velocity of superfluid and normal components, respectively. Consequently, the total force on rhs of (22) reads:

𝑭𝑭\displaystyle\boldsymbol{F}bold_italic_F =\displaystyle== A𝒆z×(d𝑹Vdt𝒗s)+B𝒆z×(d𝑹Vdt𝒗n)+C(d𝑹Vdt𝒗n)+𝐴subscript𝒆𝑧𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑠𝐵subscript𝒆𝑧𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑛limit-from𝐶𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑛\displaystyle A\boldsymbol{e}_{z}\times\left(\frac{d\boldsymbol{R}_{V}}{dt}-% \boldsymbol{v}_{s}\right)+B\boldsymbol{e}_{z}\times\left(\frac{d\boldsymbol{R}% _{V}}{dt}-\boldsymbol{v}_{n}\right)+C\left(\frac{d\boldsymbol{R}_{V}}{dt}-% \boldsymbol{v}_{n}\right)+italic_A bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT × ( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + italic_B bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT × ( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) + italic_C ( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) + (23)
+\displaystyle++ 𝑭pinning(𝑹V,d𝑹Vdt,𝒆z×d𝑹Vdt),subscript𝑭pinningsubscript𝑹𝑉𝑑subscript𝑹𝑉𝑑𝑡subscript𝒆𝑧𝑑subscript𝑹𝑉𝑑𝑡\displaystyle\boldsymbol{F}_{\textrm{pinning}}\left(\boldsymbol{R}_{V},\frac{d% \boldsymbol{R}_{V}}{dt},\boldsymbol{e}_{z}\times\frac{d\boldsymbol{R}_{V}}{dt}% \right),bold_italic_F start_POSTSUBSCRIPT pinning end_POSTSUBSCRIPT ( bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT , divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG , bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT × divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG ) ,

where the last term, describing the pinning force, breaks Galilean invariance as it corresponds to the interaction with impurities222More precisely: the pinning force describes all effects related to the momentum transfer between the vortex and impurities. It may depend on the location of the vortex, as well as its velocity [50]. The dependence on the vortex velocity becomes important in the limit of large core sizes (as compared to the impurity size), and originates from the effect of scattering of quasiparticles, bound in the vortex core. It gives rise to the so-called Kopnin-Kravtsov force [51, 52]. In the case of the neutron star crust this effect may become important in the region of pasta phases, where the core size becomes larger (due to weaker pairing correlations) than the typical scale of spatial density modulation. The pinning force is particularly important in the context of neutron star crust [53, 54], where the pinning-unpinning mechanism is essential for understanding the glitch phenomenon [26, 48]. The coefficients A,B,C𝐴𝐵𝐶A,B,Citalic_A , italic_B , italic_C depends on densities ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and ρnsubscript𝜌𝑛\rho_{n}italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, describing superfluid and normal components of the system. The equation (22) is obtained by separating terms proportional to d𝑹Vdt𝑑subscript𝑹𝑉𝑑𝑡\frac{d\boldsymbol{R}_{V}}{dt}divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG, and introducing 𝑭superfluidsubscript𝑭superfluid\boldsymbol{F}_{\textrm{superfluid}}bold_italic_F start_POSTSUBSCRIPT superfluid end_POSTSUBSCRIPT, which depends on 𝒗ssubscript𝒗𝑠\boldsymbol{v}_{s}bold_italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 𝒗nsubscript𝒗𝑛\boldsymbol{v}_{n}bold_italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT only (see eg. Ref. [55]). For the sake of generality we include the vortex mass MVsubscript𝑀𝑉M_{V}italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT, which, however, is routinely assumed to be negligible. Coefficients γ𝛾\gammaitalic_γ and ω𝜔\omegaitalic_ω need to be derived from microscopic theory and are functions of densities ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and ρnsubscript𝜌𝑛\rho_{n}italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT.

There are still many uncertainties concerning each term of this equation, particularly for fermionic vortices. Apart from the Magnus force 𝑭Mρs𝒆z×(d𝑹Vdt𝒗s)proportional-tosubscript𝑭𝑀subscript𝜌𝑠subscript𝒆𝑧𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑠\boldsymbol{F}_{M}\propto\rho_{s}\boldsymbol{e}_{z}\times(\frac{d\boldsymbol{R% }_{V}}{dt}-\boldsymbol{v}_{s})bold_italic_F start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT ∝ italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT × ( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ), which has similar form in both bosonic and fermionic case, the form and magnitude of dissipative forces, both transversal and longitudinal, are still debatable [56, 57, 58]. In the microscopic description, these forces have to be related to quasiparticles and phonon scattering on the vortex core. It includes the so-called Iordanskii force, which is the transverse force related to the relative motion with respect to the normal component of the superfluid: 𝑭I𝒆z×(d𝑹Vdt𝒗n)proportional-tosubscript𝑭𝐼subscript𝒆𝑧𝑑subscript𝑹𝑉𝑑𝑡subscript𝒗𝑛\boldsymbol{F}_{I}\propto\boldsymbol{e}_{z}\times\left(\frac{d\boldsymbol{R}_{% V}}{dt}-\boldsymbol{v}_{n}\right)bold_italic_F start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ∝ bold_italic_e start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT × ( divide start_ARG italic_d bold_italic_R start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG - bold_italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ). It is predicted to emerge from the scattering of elementary excitations of superfluid on the vortex core [59]. In the BEC case, it is related to the scattering of phonons, whereas in Fermi systems it would include also quasiparticle excitations. However the role and the magnitude of this force are still not clear (see eg. the argument in Ref. [60] against the existence of the transverse force, associated with the normal component).

Usually, to date, the determination of dissipative forces relied on a semiclassical approach [61, 62]. It consisted of modeling the presence of the classical Hamilton function: H(Lz)=αLz𝐻subscript𝐿𝑧𝛼subscript𝐿𝑧H(L_{z})=\alpha L_{z}italic_H ( italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) = italic_α italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT grasping the main feature of the chiral branch. The evolution of quasiparticles has been modeled through the semiclassical distribution function f(Lz,ϕ)𝑓subscript𝐿𝑧italic-ϕf(L_{z},\phi)italic_f ( italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_ϕ ) , where ϕitalic-ϕ\phiitalic_ϕ is an angle being canonically conjugate variable to Lzsubscript𝐿𝑧L_{z}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. Subsequently, one may search for a solution of Boltzmann eq. describing small deviations from equilibrium distribution. The quasiparticle collisions have been treated in relaxation time approximation, which introduces another parameter to the description. This approach is, however, expected to work in the deep BCS regime, where one can ignore the presence of discrete states in the band, and the effects related to mini-gap are of secondary importance.

Recently, a microscopic approach rooted in density functional theory has become possible. It provides a framework for comprehensive studies of static and time-dependent problems including thermal effects. Its popularity in nuclear physics grows with time, and today it is used for describing the properties of nuclei across the whole nuclear chart, nuclear reactions, and properties of neutron star crust [63, 64, 65, 66]. The modern DFT variants takes into account superfluid correlations, with the pairing field treated as a dynamical variable. One of the variants that proved its usability, especially in the context of ultracold Fermi gases, is known as Superfluid Local Density Approximation (SLDA) [35]. The name refers to the class of the energy density functionals \mathcal{E}caligraphic_E that depend only on local densities, such that the energy of system can be written as

E=[ρq(𝒓),νq(𝒓),τq(𝒓),𝒋q(𝒓),]𝑑𝒓.𝐸subscript𝜌𝑞𝒓subscript𝜈𝑞𝒓subscript𝜏𝑞𝒓subscript𝒋𝑞𝒓differential-d𝒓E=\int\mathcal{E}[\rho_{q}(\bm{r}),\nu_{q}(\bm{r}),\tau_{q}(\bm{r}),\bm{j}_{q}% (\bm{r}),\ldots]\,d\bm{r}.italic_E = ∫ caligraphic_E [ italic_ρ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) , italic_ν start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) , italic_τ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) , bold_italic_j start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) , … ] italic_d bold_italic_r . (24)

In this expression, q𝑞qitalic_q is a generalized index that in the context of nuclear systems refers to protons and neutrons (q=n,p𝑞𝑛𝑝q=n,pitalic_q = italic_n , italic_p), while in the context of cold fermionic gases - to atomic species. The functional is expressed through various densities, where the most common are: normal density (ρqsubscript𝜌𝑞\rho_{q}italic_ρ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT), kinetic density (τqsubscript𝜏𝑞\tau_{q}italic_τ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT), current density (𝒋qsubscript𝒋𝑞\bm{j}_{q}bold_italic_j start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT) and anomalous density (νqsubscript𝜈𝑞\nu_{q}italic_ν start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT). Other densities, not listed here, are indicated by dots. The anomalous density is crucial for the description of superfluidity as it quantifies the presence of Cooper pairs and defines the order parameter Δq(𝒓)=gνq(𝒓)subscriptΔ𝑞𝒓𝑔subscript𝜈𝑞𝒓\Delta_{q}(\bm{r})=g\nu_{q}(\bm{r})roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) = italic_g italic_ν start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) where g𝑔gitalic_g is coupling constant. The order parameter, within this framework, is treated as a complex field, allowing for the description of quantum vortices.

The precise form of the functional depends on the considered system; however, for all local functionals that belong to the SLDA class, the energy minimization condition leads to equations that formally have the same structure as Hartree-Fock-Bogoliubov or Bogoliubov-de Gennes equations. For static problems, they have a generic form (spin indices are omitted):

(hq(𝒓)Δq(𝒓)Δq(𝒓)hq(𝒓))(uq,n(𝒓)vq,n(𝒓))=Eq,n(uq,n(𝒓)vq,n(𝒓)),matrixsubscript𝑞𝒓subscriptΔ𝑞𝒓subscriptsuperscriptΔ𝑞𝒓subscriptsuperscript𝑞𝒓matrixsubscript𝑢𝑞𝑛𝒓subscript𝑣𝑞𝑛𝒓subscript𝐸𝑞𝑛matrixsubscript𝑢𝑞𝑛𝒓subscript𝑣𝑞𝑛𝒓\displaystyle\begin{pmatrix}h_{q}(\bm{r})&\Delta_{q}(\bm{r})\\ \Delta^{*}_{q}(\bm{r})&-h^{*}_{q}(\bm{r})\end{pmatrix}\begin{pmatrix}u_{q,n}(% \bm{r})\\ v_{q,n}(\bm{r})\end{pmatrix}=E_{q,n}\begin{pmatrix}u_{q,n}(\bm{r})\\ v_{q,n}(\bm{r})\end{pmatrix},( start_ARG start_ROW start_CELL italic_h start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL start_CELL roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL roman_Δ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL start_CELL - italic_h start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) = italic_E start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ( start_ARG start_ROW start_CELL italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW start_ROW start_CELL italic_v start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ( bold_italic_r ) end_CELL end_ROW end_ARG ) , (25)

where [uq,n,uq,n]Tsuperscriptsubscript𝑢𝑞𝑛subscript𝑢𝑞𝑛𝑇[u_{q,n},u_{q,n}]^{T}[ italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT are quasiparticle orbitals expressed as mixtures of particles (vq,nsubscript𝑣𝑞𝑛v_{q,n}italic_v start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT) and holes (uq,nsubscript𝑢𝑞𝑛u_{q,n}italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT). One has to remember that both hq(𝒓)subscript𝑞𝒓h_{q}(\bm{r})italic_h start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) and Δq(𝒓)subscriptΔ𝑞𝒓\Delta_{q}(\bm{r})roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) are in fact 2×2222\times 22 × 2 matrix in spin space and therefore [uq,n,uq,n]Tsuperscriptsubscript𝑢𝑞𝑛subscript𝑢𝑞𝑛𝑇[u_{q,n},u_{q,n}]^{T}[ italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT is in fact four component vector (see eqs. (3)). They are used to construct densities, e.g., ρq(𝒓)=Eq,n>0|vq,n(𝒓)|2subscript𝜌𝑞𝒓subscriptsubscript𝐸𝑞𝑛0superscriptsubscript𝑣𝑞𝑛𝒓2\rho_{q}(\bm{r})=\sum_{E_{q,n}>0}|v_{q,n}(\bm{r})|^{2}italic_ρ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( bold_italic_r ) = ∑ start_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT > 0 end_POSTSUBSCRIPT | italic_v start_POSTSUBSCRIPT italic_q , italic_n end_POSTSUBSCRIPT ( bold_italic_r ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The single-particle mean-fields hqsubscript𝑞h_{q}italic_h start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT and pairing potentials ΔqsubscriptΔ𝑞\Delta_{q}roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT, are defined via functional derivatives over densities:

hqsubscript𝑞\displaystyle h_{q}italic_h start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT =δδτq+δδρqi2{δδ𝒋q,},absentbold-∇𝛿𝛿subscript𝜏𝑞bold-∇𝛿𝛿subscript𝜌𝑞𝑖2𝛿𝛿subscript𝒋𝑞bold-∇\displaystyle=-\bm{\nabla}\frac{\delta\mathcal{E}}{\delta\tau_{q}}\bm{\nabla}+% \frac{\delta\mathcal{E}}{\delta\rho_{q}}-\frac{i}{2}\left\{\frac{\delta% \mathcal{E}}{\delta\bm{j}_{q}},\bm{\nabla}\right\},= - bold_∇ divide start_ARG italic_δ caligraphic_E end_ARG start_ARG italic_δ italic_τ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG bold_∇ + divide start_ARG italic_δ caligraphic_E end_ARG start_ARG italic_δ italic_ρ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG - divide start_ARG italic_i end_ARG start_ARG 2 end_ARG { divide start_ARG italic_δ caligraphic_E end_ARG start_ARG italic_δ bold_italic_j start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG , bold_∇ } , (26)
ΔqsubscriptΔ𝑞\displaystyle\Delta_{q}roman_Δ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT =δδνq,absent𝛿𝛿subscriptsuperscript𝜈𝑞\displaystyle=-\frac{\delta\mathcal{E}}{\delta\nu^{*}_{q}},= - divide start_ARG italic_δ caligraphic_E end_ARG start_ARG italic_δ italic_ν start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG , (27)

where {}\left\{\ldots\right\}{ … } denotes the anticommutator and δδ𝒋q𝛿𝛿subscript𝒋𝑞\frac{\delta\mathcal{E}}{\delta\bm{j}_{q}}divide start_ARG italic_δ caligraphic_E end_ARG start_ARG italic_δ bold_italic_j start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT end_ARG represents a vector constructed by variations over three components of the current 𝒋qsubscript𝒋𝑞\bm{j}_{q}bold_italic_j start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT. One may notice that expressions, like kinetic term 12m212𝑚superscript2\frac{1}{2m}\nabla^{2}divide start_ARG 1 end_ARG start_ARG 2 italic_m end_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, or mean-field potential U𝑈Uitalic_U are replaced by generalized forms, and by adequately choosing the form of the functional one can go beyond the mean-field approach, while keeping the complexity of the framework at the same level as the mean-field formulation. This flexibility can be used to create the functional, which reproduces selected observables, derived by other methods or taken from experiments. In the case of ultracold atoms the SLDA functional generates the correct equation of state (and related thermodynamic properties), strength of the pairing correlations, and effective mass of quasiparticles [67, 35, 68]. The functionals for nuclear systems are more complex, and they are constructed usually in such a way to reproduce nuclear masses and radii and/or equation of state of neutron matter or symmetric nuclear matter. For example, the results presented in Fig. 2 were obtained with Brussels-Montreal Skyrme functional (BSk), which was specifically constructed for astrophysical applications [69, 70, 71, 72, 73]. It assures correct reproduction of the equation of state and have been fitted to reproduce the magnitude of the pairing gap as a function of neutron density, obtained within Brueckner theory, including screening effects [74]. Consequently, DFT provides a reliable source of microscopic information about the vortices [34, 75, 76, 77, 28]. The approach can be extended to time-dependent phenomena by replacing Eq,ki/tsubscript𝐸𝑞𝑘𝑖Planck-constant-over-2-pi𝑡E_{q,k}\rightarrow i\hbar\partial/\partial titalic_E start_POSTSUBSCRIPT italic_q , italic_k end_POSTSUBSCRIPT → italic_i roman_ℏ ∂ / ∂ italic_t, which allows studying the vortex dynamics [41, 27, 78, 79, 80, 81]. Consequently, the time-dependent studies can provide microscopic insight into origin of forces acting on the vortex.

It has been conjectured that an accelerating vortex may be subject to dissipative forces originating from quasiparticles trapped in the core [62]. The conjecture, based on a semiclassical approach, presents a mechanism in which the acceleration of a vortex will lead to heating up a gas of quasiparticles in the core. As a result, some of them will be emitted from the core, taking away energy carried by the vortex line. Although the semiclassical approach is applicable in deep BCS regime, the experiment has been recently performed for a system close to the unitarity, where significant dissipation effects have been observed [82, 83]. However, the presence of finite temperature did not allow to distinguish the source of dissipation clearly. Through time-dependent SLDA (TDSLDA) simulations of collisions of two pairs of vortex-antivortex dipoles, it has been possible to identify dissipative mechanism due to quasiparticle ejection [78]. The application of the TDSLDA framework revealed that the dissipative mechanism via excitations of the vortex core, while present, emerges to be of secondary importance, and thermal effects dominate the dynamics in experimental realization.

It has to be noted that, in the context of neutron stars, there is an additional dissipation channel associated with neutron quasiparticles bound in the vortex core. It is due to the interaction both with electrons and protons and has been discussed in Ref. [84]. In the case of electrons the coupling is provided by interaction with the neutron magnetic moment, whereas protons couple strongly through the nuclear interaction. Detailed dynamics is complicated due to the coupling between electron dynamics and protons [85]. However, the results indicate that coupling of proton-electron plasma to neutron vortices lead to short relaxation times implying that magnetars superfluid cores will couple to the remaining stellar plasma on short dynamical time-scales once proton superconductivity is suppressed.

Refer to caption
Figure 4: Vortex dipole (vortex-antivortex pair) propagating in a spin imbalanced unitary Fermi gas (UFG). Spin imbalance induces the existence of nodal lines of the pairing field, where the majority spin particles are accumulated. Therefore the local polarization is increased along the nodal lines, which is seen as white stripes (local polarization is defined as: p(r)=ρρρ+ρ𝑝𝑟subscript𝜌subscript𝜌subscript𝜌subscript𝜌p(r)=\frac{\rho_{\uparrow}-\rho_{\downarrow}}{\rho_{\uparrow}+\rho_{\downarrow}}italic_p ( italic_r ) = divide start_ARG italic_ρ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT - italic_ρ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT ↑ end_POSTSUBSCRIPT + italic_ρ start_POSTSUBSCRIPT ↓ end_POSTSUBSCRIPT end_ARG). Arrows indicate the local current and shows the position of the vortex cores. Lower subfigure b) show the configuration after time interval t=250ϵF1𝑡250superscriptsubscriptitalic-ϵ𝐹1t=250\epsilon_{F}^{-1}italic_t = 250 italic_ϵ start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT with respect to initial configuration a).

Another successful application of TDSLDA approach has been the determination of the pinning force acting between a vortex and a nucleus immersed in superfluid neutron matter. To date, studies of this effect were based on comparing the energy of the vortex pinned and unpinned from the impurity [86, 87, 88, 89, 75, 90, 91, 92]. However, these results were not very accurate due to small energy difference between configurations. One must also remember that the pinning process requires energy dissipation. Otherwise, the vortex will simply orbit the nucleus as a result of the Magnus force. Therefore, the proper description of dissipative processes is essential. Ref. [27] described how the pinning force can be accurately determined by dragging the impurity against the vortex with constant velocity. Then, by combining the information about the force with the vortex-nucleus separation, one could extract the force as a function of vortex-nucleus separation. It was found that the extracted force has a negligible tangential component and is of predominantly central character. The effective range of the force is about 10fm10fm10\textrm{fm}10 fm for the lower density, increasing to about 15fm15fm15\textrm{fm}15 fm for the higher density, consistent with an increasing coherence length with density and decreasing neutron pairing gap. The behavior of the total force for small separations demonstrates that it is not merely a function of distance. At small separations, the deformation of the vortex line and the nuclear deformation become important degrees of freedom.

4 Summary and open questions

The vortex structure and dynamics discussed in the previous sections are crucial for understanding the role of vortices in neutron stars, particularly in the glitch phenomenon, and more generally, to understand the decay pattern of fermionic quantum turbulence.

Vortex structure is relatively well known, although its modifications due to spin imbalance are still not fully understood. The situation is different when one considers dynamical properties. To date, most investigations were performed based on very simplified assumptions, which at most can be valid in deep BCS regimes. However, the possibility of using the unconstrained time-dependent BdG approach, and more generally, the TDSLDA framework, paved the way to extract components of forces exerted on a vortex moving through a superfluid. It whets our appetite concerning the determination of the microscopic underpinning of vortex dynamics in a fermionic environment. It is particularly important that these studies are performed in close collaboration with experiments in ultracold atoms, which allows the testing of the applicability of the theoretical framework. To date, the two aforementioned examples of applications of TDSLDA have provided very promising outcomes.

The degree of freedom related to spin imbalance will open a new avenue for studying vortex dynamics. The motion of the vortex in the presence of a large number of majority spin particles, may change the vortex dynamics qualitatively. Indeed, their scattering off the core is expected to enhance dissipative processes considerably [93]. As an example, in the Fig. 4, the results of TDSLDA simulations of moving vortex dipole have been shown. In the lower subfigure, one can notice the significant distortion of the initial vortex dipole (upper subfigure), which occurs due to interaction with majority spin particles accumulated in the pairing nodal lines. As a result some quasiparticles have been trapped in the vortex core. Therefore, it is expected that the motion of the vortex in the spin-imbalanced system may shed light on the magnitude of the Iordanskii force, which should be significantly enhanced in this case. As a result, we may encounter new, unexpected features of vortex dynamics that have not yet been observed.

\bmhead

Acknowledgments

This work was financially supported by the (Polish) National Science Center Grants under Contracts No. UMO-2021/43/B/ST2/01191 (PM, AM), UMO-2022/45/B/ST2/00358 (GW, ABa) and UMO-2021/40/C/ST2/00072 (DP). This work used computational resources of Tsubame3.0 supercomputer provided by Tokyo Institute of Technology through the HPCI System Research Project (Project ID: hp230081).

References

  • \bibcommenthead
  • Ring and Schuck [2004] Ring, P., Schuck, P.: The Nuclear Many-body Problem. Springer, Berlin, Germany (2004)
  • Dean and Hjorth-Jensen [2003] Dean, D.J., Hjorth-Jensen, M.: Pairing in nuclear systems: from neutron stars to finite nuclei. Rev. Mod. Phys. 75, 607–656 (2003) https://doi.org/10.1103/RevModPhys.75.607
  • Brink and Broglia [2023] Brink, D.M., Broglia, R.A.: Nuclear Superfluidity: Pairing in Finite Systems. Cambridge University Press, Cambridge, England (2023)
  • Dobaczewski et al. [2001] Dobaczewski, J., Magierski, P., Nazarewicz, W., Satuła, W., Szymański, Z.: Odd-even staggering of binding energies as a consequence of pairing and mean-field effects. Phys. Rev. C 63, 024308 (2001) https://doi.org/10.1103/PhysRevC.63.024308
  • Möller and Nix [1992] Möller, P., Nix, J.: Nuclear pairing models. Nuclear Physics A 536(1), 20–60 (1992) https://doi.org/10.1016/0375-9474(92)90244-E
  • Goldanskii and Larkin [1968] Goldanskii, V.I., Larkin, A.I.: An analog of the josephson effect in nuclear transformations. Sov. Phys. JETP 26, 617 (1968)
  • Dietrich [1970] Dietrich, K.: On a nuclear josephson effect in heavy ion scattering. Physics Letters B 32(6), 428–430 (1970) https://doi.org/10.1016/0370-2693(70)90372-2
  • Dietrich [1971] Dietrich, K.: Semiclassical theory of a nuclear josephson effect in reactions between heavy ions. Annals of Physics 66(2), 480–508 (1971) https://doi.org/10.1016/0003-4916(71)90067-4
  • Magierski [2021] Magierski, P.: The tiniest superfluid circuit in nature. Physics 14, 27 (2021)
  • Magierski et al. [2017] Magierski, P., Sekizawa, K., Wlazłowski, G.: Novel role of superfluidity in low-energy nuclear reactions. Phys. Rev. Lett. 119, 042501 (2017) https://doi.org/%****␣main.tex␣Line␣1000␣****10.1103/PhysRevLett.119.042501
  • Magierski et al. [2022] Magierski, P., Makowski, A., Barton, M.C., Sekizawa, K., Wlazłowski, G.: Pairing dynamics and solitonic excitations in collisions of medium-mass, identical nuclei. Phys. Rev. C 105, 064602 (2022) https://doi.org/10.1103/PhysRevC.105.064602
  • Mermaz [1987] Mermaz, M.C.: Weak evidence for a nuclear josephson effect in the S34superscriptS34{}^{34}\mathrm{S}start_FLOATSUPERSCRIPT 34 end_FLOATSUPERSCRIPT roman_S (S32superscriptS32{}^{32}\mathrm{S}start_FLOATSUPERSCRIPT 32 end_FLOATSUPERSCRIPT roman_S, S32superscriptS32{}^{32}\mathrm{S}start_FLOATSUPERSCRIPT 32 end_FLOATSUPERSCRIPT roman_S) elastic scattering reaction. Phys. Rev. C 36, 1192–1193 (1987) https://doi.org/10.1103/PhysRevC.36.1192
  • Mermaz and Girod [1996] Mermaz, M.C., Girod, M.: Neutron pair and proton pair transfer reactions between identical cores in the sulfur region. Phys. Rev. C 53, 1819–1823 (1996) https://doi.org/10.1103/PhysRevC.53.1819
  • Sugiyama et al. [1997] Sugiyama, Y., Tomita, Y., Yamanouti, Y., Hamada, S., Ikuta, T., Fujita, H., Napoli, D.R.: Elastic two-neutron transfer reactions of Ni58+60Nisuperscript60superscriptNi58Ni{}^{58}\mathrm{Ni}{+}^{60}\mathrm{Ni}start_FLOATSUPERSCRIPT 58 end_FLOATSUPERSCRIPT roman_Ni + start_POSTSUPERSCRIPT 60 end_POSTSUPERSCRIPT roman_Ni and Ni62+64Nisuperscript64superscriptNi62Ni{}^{62}\mathrm{Ni}{+}^{64}\mathrm{Ni}start_FLOATSUPERSCRIPT 62 end_FLOATSUPERSCRIPT roman_Ni + start_POSTSUPERSCRIPT 64 end_POSTSUPERSCRIPT roman_Ni around the coulomb barrier. Phys. Rev. C 55, 5–7 (1997) https://doi.org/10.1103/PhysRevC.55.R5
  • Montanari et al. [2016] Montanari, D., et al.: Pair neutron transfer in Ni60+116Snsuperscript116superscriptNi60Sn{}^{60}\text{Ni}+\phantom{\rule{1.60004pt}{0.0pt}}^{116}\text{Sn}start_FLOATSUPERSCRIPT 60 end_FLOATSUPERSCRIPT Ni + start_POSTSUPERSCRIPT 116 end_POSTSUPERSCRIPT Sn probed via γ𝛾\gammaitalic_γ-particle coincidences. Phys. Rev. C 93, 054623 (2016) https://doi.org/10.1103/PhysRevC.93.054623
  • Potel et al. [2021] Potel, G., Barranco, F., Vigezzi, E., Broglia, R.A.: Quantum entanglement in nuclear cooper-pair tunneling with gamma rays. Phys. Rev. C 103, 021601 (2021) https://doi.org/%****␣main.tex␣Line␣1100␣****10.1103/PhysRevC.103.L021601
  • [17] Broglia, R., Barranco, F., Corradi, L., Potel, G., Szilner, S., Vigezzi, E.: Nuclear josephson-like γ𝛾\gammaitalic_γ - emission. arXiv:2206.05351 https://doi.org/10.48550/arXiv.2206.05351
  • Scamps [2018] Scamps, G.: Examining empirical evidence of the effect of superfluidity on the fusion barrier. Phys. Rev. C 97, 044611 (2018) https://doi.org/10.1103/PhysRevC.97.044611
  • Cao et al. [2006] Cao, L., Lombardo, U., Schuck, P.: Screening effects in superfluid nuclear and neutron matter within brueckner theory. Phys. Rev. C 74(6), 064301 (2006) https://doi.org/10.1103/PhysRevC.74.064301
  • Chamel and Haensel [2008] Chamel, N., Haensel, P.: Physics of Neutron Star Crusts. Living Reviews in Relativity 11(1), 10 (2008) https://doi.org/10.12942/lrr-2008-10
  • Migdal [1959] Migdal, A.B.: Superfluidity and the moments of inertia of nuclei. Nuclear Physics 13(5), 655–674 (1959) https://doi.org/10.1016/0029-5582(59)90264-0
  • Baym et al. [1969] Baym, G., Pethick, C., Pines, D.: Superfluidity in Neutron Stars. Nature 224, 673–674 (1969) https://doi.org/10.1038/224673a0
  • Pines and Alpar [1985] Pines, D., Alpar, M.A.: Superfluidity in neutron stars. Nature 316(6023), 27–32 (1985) https://doi.org/10.1038/316027a0
  • Chamel [2017] Chamel, N.: Superfluidity and Superconductivity in Neutron Stars. J. Astrophys. Astron. 38, 43 (2017) https://doi.org/10.1007/s12036-017-9470-9
  • Anderson and Itoh [1975] Anderson, P.W., Itoh, N.: Pulsar glitches and restlessness as a hard superfluidity phenomenon. Nature 256(5512), 25–27 (1975) https://doi.org/10.1038/256025a0
  • Haskell and Melatos [2015] Haskell, B., Melatos, A.: Models of pulsar glitches. International Journal of Modern Physics D 24(03), 1530008 (2015) https://doi.org/10.1142/S0218271815300086
  • Wlazłowski et al. [2016] Wlazłowski, G., Sekizawa, K., Magierski, P., Bulgac, A., Forbes, M.M.: Vortex pinning and dynamics in the neutron star crust. Phys. Rev. Lett. 117, 232701 (2016) https://doi.org/10.1103/PhysRevLett.117.232701
  • Pęcak et al. [2021] Pęcak, D., Chamel, N., Magierski, P., Wlazłowski, G.: Properties of a quantum vortex in neutron matter at finite temperatures. Phys. Rev. C 104(5), 055801 (2021) https://doi.org/10.1103/PhysRevC.104.055801
  • Poli et al. [2023] Poli, E., Bland, T., White, S.J.M., Mark, M.J., Ferlaino, F., Trabucco, S., Mannarelli, M.: Glitches in rotating supersolids. Phys. Rev. Lett. 131, 223401 (2023) https://doi.org/%****␣main.tex␣Line␣1300␣****10.1103/PhysRevLett.131.223401
  • Giorgini et al. [2008] Giorgini, S., Pitaevskii, L.P., Stringari, S.: Theory of ultracold atomic fermi gases. Rev. Mod. Phys. 80, 1215–1274 (2008) https://doi.org/10.1103/RevModPhys.80.1215
  • Ortiz and Ceperley [1995] Ortiz, G., Ceperley, D.M.: Core structure of a vortex in superfluid He4superscriptHe4{}^{4}\mathrm{He}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT roman_He. Phys. Rev. Lett. 75, 4642–4645 (1995) https://doi.org/10.1103/PhysRevLett.75.4642
  • Zwierlein et al. [2005] Zwierlein, M.W., Abo-Shaeer, J.R., Schirotzek, A., Schunck, C.H., Ketterle, W.: Vortices and superfluidity in a strongly interacting fermi gas. Nature 435, 1047–1051 (2005) https://doi.org/10.1038/nature03858
  • Caroli et al. [1964] Caroli, C., De Gennes, P.G., Matricon, J.: Bound fermion states on a vortex line in a type ii superconductor. Physics Letters 9(4), 307–309 (1964) https://doi.org/10.1016/0031-9163(64)90375-0
  • Sensarma et al. [2006] Sensarma, R., Randeria, M., Ho, T.-L.: Vortices in superfluid fermi gases through the bec to bcs crossover. Phys. Rev. Lett. 96, 090403 (2006) https://doi.org/10.1103/PhysRevLett.96.090403
  • Bulgac et al. [2012] Bulgac, A., Forbes, M. McNeil, Magierski, P.: The unitary Fermi gas: From monte carlo to density functionals. In: Lecture Notes on Physics: The BCS-BEC Crossover and the Unitary Fermi Gas, p. 305. Springer, Berlin Heidelberg (2012). https://doi.org/10.1007/978-3-642-21978-8
  • Andreev [1964] Andreev, A.F.: The thermal conductivity of the intermediate state in superconductors. Zh. Eksp. Teor. Fiz. 46(5), 1823–1828 (1964)
  • Adagideli and Goldbart [2002] Adagideli, I., Goldbart, P.: Quantal andreev billiards: Semiclassical approach to mesoscale oscillations in the density of states. International Journal of Modern Physics B 16 (2002) https://doi.org/10.1142/S0217979202010300
  • Magierski et al. [2022] Magierski, P., Wlazłowski, G., Makowski, A., Kobuszewski, K.: Spin-polarized vortices with reversed circulation. Phys. Rev. A 106, 033322 (2022) https://doi.org/10.1103/PhysRevA.106.033322
  • Douchin, F. and Haensel, P. [2001] Douchin, F., Haensel, P.: A unified equation of state of dense matter and neutron star structure. A&A 380(1), 151–167 (2001) https://doi.org/10.1051/0004-6361:20011402
  • Wlazłowski et al. [2024] Wlazłowski, G., Forbes, M.M., Sarkar, S.R., Marek, A., Szpindler, M.: Fermionic quantum turbulence: Pushing the limits of high-performance computing. PNAS Nexus 3(5), 160 (2024) https://doi.org/10.1093/pnasnexus/pgae160 https://academic.oup.com/pnasnexus/article-pdf/3/5/pgae160/58004759/pgae160.pdf
  • Tylutki and Wlazłowski [2021] Tylutki, M., Wlazłowski, G.: Universal aspects of vortex reconnections across the bcs-bec crossover. Phys. Rev. A 103, 051302 (2021) https://doi.org/10.1103/PhysRevA.103.L051302
  • Takahashi et al. [2006] Takahashi, M., Mizushima, T., Ichioka, M., Machida, K.: Vortex-core structure in neutral fermion superfluids with population imbalance. Phys. Rev. Lett. 97(18), 180407 (2006) https://doi.org/10.1103/PhysRevLett.97.180407
  • Hu et al. [2007] Hu, H., Liu, X.-J., Drummond, P.D.: Visualization of vortex bound states in polarized fermi gases at unitarity. Phys. Rev. Lett. 98(6), 060406 (2007) https://doi.org/10.1103/PhysRevLett.98.060406
  • Stein et al. [2016] Stein, M., Sedrakian, A., Huang, X.-G., Clark, J.W.: Spin-polarized neutron matter: Critical unpairing and bcs-bec precursor. Phys. Rev. C 93, 015802 (2016) https://doi.org/10.1103/PhysRevC.93.015802
  • Kaspi and Beloborodov [2017] Kaspi, V.M., Beloborodov, A.M.: Magnetars. Annual Review of Astronomy and Astrophysics 55, 261–301 (2017) https://doi.org/10.1146/annurev-astro-081915-023329
  • Inotani et al. [2021] Inotani, D., Yasui, S., Mizushima, T., Nitta, M.: Radial fulde-ferrell-larkin-ovchinnikov-like state in a population-imbalanced fermi gas. Phys. Rev. A 103, 053308 (2021) https://doi.org/10.1103/PhysRevA.103.053308
  • Tsubota [2008] Tsubota, M.: Quantum turbulence. Journal of the Physical Society of Japan 77(11), 111006 (2008) https://doi.org/10.1143/JPSJ.77.111006
  • Haskell et al. [2020] Haskell, B., Antonopoulou, D., Barenghi, C.: Turbulent, pinned superfluids in neutron stars and pulsar glitch recoveries. Monthly Notices of the Royal Astronomical Society 499(1), 161–170 (2020) https://doi.org/10.1093/mnras/staa2678
  • Nakamura et al. [2012] Nakamura, K., Babajanov, D., Matrasulov, D., Kobayashi, M.: Dynamics of inertial vortices in multicomponent bose-einstein condensates. Phys. Rev. A 86, 053613 (2012) https://doi.org/10.1103/PhysRevA.86.053613
  • Sonin [1997] Sonin, E.B.: Magnus force in superfluids and superconductors. Phys. Rev. B 55, 485–501 (1997) https://doi.org/10.1103/PhysRevB.55.485
  • Kopnin and Kravtsov [1976 [JETP Lett. 23, 578 (1976)]] Kopnin, N.B., Kravtsov, V.E.: Conductivity and hall effect of pure type-ii superconductors at low temperature. Pis’ma Zh. Eksp. Teor. Fiz. 23, 631–634 (1976 [JETP Lett. 23, 578 (1976)])
  • Kopnin and Lopatin [1997] Kopnin, N.B., Lopatin, A.V.: Two relaxation times in mutual friction of superfluid 3he. Phys. Rev. B 56, 766–779 (1997) https://doi.org/10.1103/PhysRevB.56.766
  • Link [2009] Link, B.: Dynamics of quantum vorticity in a random potential. Physical review letters 102(13), 131101 (2009)
  • Haskell and Melatos [2016] Haskell, B., Melatos, A.: Pinned vortex hopping in a neutron star crust. Monthly Notices of the Royal Astronomical Society 461(2), 2200–2211 (2016)
  • Geller et al. [1998] Geller, M.R., Wexler, C., Thouless, D.J.: Transverse force on a quantized vortex in a superconductor. Phys. Rev. B 57, 8119–8122 (1998) https://doi.org/10.1103/PhysRevB.57.R8119
  • Stone [1996] Stone, M.: Spectral flow, magnus force, and mutual friction via the geometric optics limit of andreev reflection. Phys. Rev. B 54, 13222–13229 (1996) https://doi.org/10.1103/PhysRevB.54.13222
  • Stone [2000] Stone, M.: Iordanskii force and the gravitational aharonov-bohm effect for a moving vortex. Phys. Rev. B 61 (2000) https://doi.org/10.1103/PhysRevB.61.11780
  • Sonin [2013] Sonin, E.B.: Transverse force on a vortex and vortex mass: Effects of free bulk and vortex-core bound quasiparticles. Phys. Rev. B 87, 134515 (2013) https://doi.org/10.1103/PhysRevB.87.134515
  • Iordanskii [1965 [Sov. Phys. JETP, 22, 160 (1966)]] Iordanskii, S.V.: Mutual friction force in a rotating bose gas. Zh. Eksp. Teor. Fiz. 49, 225–236 (1965 [Sov. Phys. JETP, 22, 160 (1966)])
  • Wexler [1997] Wexler, C.: Magnus and iordanskii forces in superfluids. Phys. Rev. Lett. 79, 1321–1324 (1997) https://doi.org/10.1103/PhysRevLett.79.1321
  • Kopnin [2002] Kopnin, N.B.: Vortex dynamics and mutual friction in superconductors and fermi superfluids. Reports on Progress in Physics 65(11), 1633 (2002) https://doi.org/10.1088/0034-4885/65/11/202
  • Silaev [2012] Silaev, M.A.: Universal mechanism of dissipation in fermi superfluids at ultralow temperatures. Phys. Rev. Lett. 108, 045303 (2012) https://doi.org/10.1103/PhysRevLett.108.045303
  • Nakatsukasa et al. [2016] Nakatsukasa, T., Matsuyanagi, K., Matsuo, M., Yabana, K.: Time-dependent density-functional description of nuclear dynamics. Rev. Mod. Phys. 88, 045004 (2016) https://doi.org/10.1103/RevModPhys.88.045004
  • Colò [2020] Colò, G.: Nuclear density functional theory. Advances in Physics: X 5(1), 1740061 (2020) https://doi.org/10.1080/23746149.2020.1740061
  • Bulgac [2019] Bulgac, A.: Time-dependent density functional theory for fermionic superfluids: From cold atomic gases - to nuclei and neutron stars crust. physica status solidi (b) 256(7), 1800592 (2019) https://doi.org/10.1002/pssb.201800592
  • Magierski [2019] Magierski, P.: Nuclear reactions and superfluid time depen-dent density functional theory. Progress of Time-Dependent Nuclear Reaction Theory 2, 57–71 (2019) https://doi.org/%****␣main.tex␣Line␣1850␣****10.2174/9781681087641119020008
  • Bulgac [2007] Bulgac, A.: Local-density-functional theory for superfluid fermionic systems: The unitary gas. Phys. Rev. A 76, 040502 (2007) https://doi.org/10.1103/PhysRevA.76.040502
  • Boulet et al. [2022] Boulet, A., Wlazłowski, G., Magierski, P.: Local energy density functional for superfluid fermi gases from effective field theory. Phys. Rev. A 106, 013306 (2022) https://doi.org/10.1103/PhysRevA.106.013306
  • Goriely et al. [2016] Goriely, S., Chamel, N., Pearson, J.M.: Further explorations of skyrme-hartree-fock-bogoliubov mass formulas. xvi. inclusion of self-energy effects in pairing. Phys. Rev. C 93(3), 034337 (2016) https://doi.org/10.1103/PhysRevC.93.034337
  • Chamel et al. [2009] Chamel, N., Goriely, S., Pearson, J.M.: Further explorations of skyrme-hartree-fock-bogoliubov mass formulas. xi. stabilizing neutron stars against a ferromagnetic collapse. Phys. Rev. C 80(6), 065804 (2009) https://doi.org/10.1103/PhysRevC.80.065804
  • Chamel et al. [2008] Chamel, N., Goriely, S., Pearson, J.M.: Further explorations of Skyrme Hartree Fock Bogoliubov mass formulas. IX: Constraint of pairing force to 1S0 neutron-matter gap. Nucl. Phys. A 812(1-4), 72–98 (2008) https://doi.org/10.1016/j.nuclphysa.2008.08.015
  • Goriely et al. [2009] Goriely, S., Chamel, N., Pearson, J.M.: Skyrme-hartree-fock-bogoliubov nuclear mass formulas: Crossing the 0.6 mev accuracy threshold with microscopically deduced pairing. Phys. Rev. Lett. 102, 152503 (2009) https://doi.org/10.1103/PhysRevLett.102.152503
  • Chamel [2010] Chamel, N.: Effective contact pairing forces from realistic calculations in infinite homogeneous nuclear matter. Phys. Rev. C 82(1), 014313 (2010) https://doi.org/10.1103/PhysRevC.82.014313
  • Cao et al. [2006] Cao, L.G., Lombardo, U., Schuck, P.: Screening effects in superfluid nuclear and neutron matter within brueckner theory. Phys. Rev. C 74, 064301 (2006) https://doi.org/10.1103/PhysRevC.74.064301
  • Avogadro et al. [2007] Avogadro, P., Barranco, F., Broglia, R.A., Vigezzi, E.: Quantum calculation of vortices in the inner crust of neutron stars. Phys. Rev. C 75, 012805 (2007) https://doi.org/10.1103/PhysRevC.75.012805
  • Bulgac and Yu [2003] Bulgac, A., Yu, Y.: Vortex state in a strongly coupled dilute atomic fermionic superfluid. Phys. Rev. Lett. 91, 190404 (2003) https://doi.org/10.1103/PhysRevLett.91.190404
  • Yu and Bulgac [2003] Yu, Y., Bulgac, A.: Spatial structure of a vortex in low density neutron matter. Phys. Rev. Lett. 90, 161101 (2003) https://doi.org/10.1103/PhysRevLett.90.161101
  • Barresi et al. [2023] Barresi, A., Boulet, A., Magierski, P., Wlazłowski, G.: Dissipative dynamics of quantum vortices in fermionic superfluid. Phys. Rev. Lett. 130, 043001 (2023) https://doi.org/10.1103/PhysRevLett.130.043001
  • Hossain et al. [2022] Hossain, K., Kobuszewski, K., Forbes, M.M., Magierski, P., Sekizawa, K., Wlazłowski, G.: Rotating quantum turbulence in the unitary fermi gas. Phys. Rev. A 105, 013304 (2022) https://doi.org/10.1103/PhysRevA.105.013304
  • Wlazłowski et al. [2015] Wlazłowski, G., Bulgac, A., Forbes, M.M., Roche, K.J.: Life cycle of superfluid vortices and quantum turbulence in the unitary fermi gas. Phys. Rev. A 91, 031602 (2015) https://doi.org/10.1103/PhysRevA.91.031602
  • Bulgac et al. [2014] Bulgac, A., Forbes, M.M., Kelley, M.M., Roche, K.J., Wlazłowski, G.: Quantized superfluid vortex rings in the unitary fermi gas. Phys. Rev. Lett. 112, 025301 (2014) https://doi.org/10.1103/PhysRevLett.112.025301
  • Kwon et al. [2021] Kwon, W.J., Del Pace, G., Xhani, K., Galantucci, L., Muzi Falconi, A., Inguscio, M., Scazza, F., Roati, G.: Sound emission and annihilations in a programmable quantum vortex collider. Nature (London) 600(7887), 64 (2021)
  • Autti et al. [2020] Autti, S., Ahlstrom, S.L., Haley, R.P., Jennings, A., Pickett, G.R., Poole, M., Schanen, R., Soldatov, A.A., Tsepelin, V., Vonka, J., Wilcox, T., Woods, A.J., Zmeev, D.E.: Fundamental dissipation due to bound fermions in the zero-temperature limit. Nat. Commun. 11(1), 4742 (2020)
  • Sedrakian and Clark [2019] Sedrakian, A., Clark, J.W.: Superfluidity in nuclear systems and neutron stars. The European Physical Journal A 55, 167 (2019) https://doi.org/10.1140/epja/i2019-12863-6
  • Haskell and Sedrakian [2018] Haskell, B., Sedrakian, A.: In: Rezzolla, L., Pizzochero, P., Jones, D.I., Rea, N., Vidaña, I. (eds.) Superfluidity and Superconductivity in Neutron Stars, pp. 401–454. Springer, Cham (2018). https://doi.org/10.1007/978-3-319-97616-7_8 . https://doi.org/10.1007/978-3-319-97616-7_8
  • Pizzochero et al. [1997] Pizzochero, P.M., Viverit, L., Broglia, R.A.: Vortex-nucleus interaction and pinning forces in neutron stars. Phys. Rev. Lett. 79, 3347–3350 (1997) https://doi.org/10.1103/PhysRevLett.79.3347
  • Donati and Pizzochero [2003] Donati, P., Pizzochero, P.M.: Is there nuclear pinning of vortices in superfluid pulsars? Phys. Rev. Lett. 90, 211101 (2003) https://doi.org/10.1103/PhysRevLett.90.211101
  • Donati and Pizzochero [2004] Donati, P., Pizzochero, P.M.: Fully consistent semi-classical treatment of vortex-nucleus interaction in rotating neutron stars. Nuclear Physics A 742(3), 363–379 (2004) https://doi.org/10.1016/j.nuclphysa.2004.07.002
  • Donati and Pizzochero [2006] Donati, P., Pizzochero, P.M.: Realistic energies for vortex pinning in intermediate-density neutron star matter. Physics Letters B 640(3), 74–81 (2006) https://doi.org/10.1016/j.physletb.2006.07.047
  • Avogadro et al. [2008] Avogadro, P., Barranco, F., Broglia, R.A., Vigezzi, E.: Vortex-nucleus interaction in the inner crust of neutron stars. Nuclear Physics A 811(3), 378–412 (2008) https://doi.org/10.1016/j.nuclphysa.2008.07.010
  • Seveso et al. [2015] Seveso, S., Pizzochero, P.M., Grill, F., Haskell, B.: Mesoscopic pinning forces in neutron star crusts. Monthly Notices of the Royal Astronomical Society 455(4), 3952–3967 (2015) https://doi.org/10.1093/mnras/stv2579 https://academic.oup.com/mnras/article-pdf/455/4/3952/4094721/stv2579.pdf
  • Klausner et al. [2023] Klausner, P., Barranco, F., Pizzochero, P.M., Roca-Maza, X., Vigezzi, E.: Microscopic calculation of the pinning energy of a vortex in the inner crust of a neutron star. Phys. Rev. C 108, 035808 (2023) https://doi.org/10.1103/PhysRevC.108.035808
  • [93] Barresi, A., et al. (in preparation)