Relaxation time for the alignment between quark spin and
angular velocity in a rotating QCD medium
Abstract
We compute the relaxation times for massive quarks and antiquarks to align their spins with the angular velocity in a rigidly rotating medium at finite temperature and baryon density. The rotation effects are implemented using a fermion propagator immersed in a cylindrical rotating environment. The relaxation time is computed as the inverse of the interaction rate to produce an asymmetry between the quark (antiquark) spin components along and opposite to the angular velocity. For conditions resembling heavy-ion collisions, the relaxation times for quarks are smaller than for antiquarks. For semicentral collisions the relaxation time is within the possible life-time of the QGP for all collision energies. However, for antiquarks this happens only for collision energies GeV. The results are quantified in terms of the intrinsic quark and antiquark polarizations, namely, the probability to build the spin asymmetry as a function of time. Our results show that these intrinsic polarizations tend to 1 with time at different rates given by the relaxation times with quarks reaching a sizable asymmetry at a faster pace. These are key results to further elucidate the mechanisms of hyperon polarization in relativistic heavy-ion collisions.
I Introduction
Relativistic heavy-ion collisions are the best tool to explore, in a controlled manner, the properties of strongly interacting matter under extreme conditions. The study of the different observables emerging from these reactions has produced a wealth of results revealing an ever more complete picture of these properties for temperatures and densities close to or above the deconfinement transition. However, some other phenomena still miss a clearer understanding and pose a challenge for the evolving standard model of heavy-ion reactions. One of these observables is the relatively large degree of polarization of and hyperons measured in semicentral collision for energies 2.5 GeV GeV, which shows an increasing trend as the energy and centrality of the collision decreases. The raising trend is different for s than s STAR:2017ckg; STAR:2021beb; STAR:2023nvo; HADES:2022enx. For semicentral collisions, the matter density profile in the transverse plane induces the development of a global angular momentum, quantified in terms of the thermal vorticity Becattini2008; Becattini:2016gvu. Such angular momentum could be transferred to spin degrees of freedom and be responsible for the observed global polarization Becattini:2022zvf. This expectation is supported by the relation between rotation and spin, nowadays referred to as the Barnett effect, whereby a spinning ferromagnet experiences a change of its magnetization 1915PhRv....6..239B and the closely related Einstein–de Haas effect, based on the observation that a change in the magnetic moment of a free body causes this body to rotate 1915KNAB...18..696E. As a consequence, significant efforts have been devoted to quantify how this vorticity may be responsible for the magnitude of the observed polarization, assuming that the medium rotation is transferred to the spin polarization regardless of the microscopic mechanisms responsible for the effect Karpenko:2021wdm; DelZanna:2013eua; Karpenko:2013wva; Karpenko:2016jyx; Ivanov:2019ern; Wei:2018zfb; Vitiuk:2019rfv; Xie:2019jun; Ivanov:2020udj. However, the transferring of rotational motion to spin can only happen provided the medium induced reactions occur fast enough so that the alignment of the spin and angular velocity takes place on average within the lifetime of the medium. In the recent literature, this question has been addressed using different approaches Montenegro:2018bcf; Kapusta:2019ktm; Kapusta:2019sad; Montenegro:2020paq; Kapusta:2020dco; Kapusta:2020npk; Torrieri:2022ogj.
In a couple of recent works, we have explored whether this relaxation time for the alignment is short enough so that the observed polarization of hyperons can be attributed to the transferring of rotation to spin degrees of freedom Ayala:2020ndx; Ayala:2019iin. This is achieved by computing the interaction rate for the spin of a strange quark to align with the thermal vorticity, assuming an effective spin-vorticity coupling in a thermal QCD medium. The findings have been used to compute the and polarization in the context of a core-corona model Ayala:2023xyn; Ayala:2022yyx; Ayala:2021xrn; Ayala:2020soy. The calculation resorts to computing the imaginary part of the self-energy of a vacuum quark whose propagator does not experience the effects of the rotational motion. To improve the description, also in a recent work, we have computed the propagation of a spin one-half fermion immersed in a rigid, cylindrical rotating environment Ayala:2021osy. For these purposes, we have followed the method introduced in Ref. Iablokov:2020upc which requires knowledge of the explicit solutions of the Dirac equation. These have been previously studied in different contexts by imposing different boundary conditions Chodos:1974je; Chernodub:2016kxh; Ambrus:2015lfr; Chen:2015hfc; Ebihara:2016fwa; Ambrus:2014uqa; Yamamoto:2013zwa; Gaspar:2023nqk.
In this work we use the propagator found in Ref. Ayala:2021osy to compute the imaginary part of the self-energy of a quark immersed in a rotating QCD medium at finite temperature () and baryo-chemical potential (). We show that for values of and where the chiral symmetry restoration/deconfinement transition is thought to take place, the relaxation time for quarks turns out to be small enough, compared to the medium life-time, for the inferred, commonly accepted values of the medium angular velocity, after a semicentral heavy-ion collision. However, this is not the case for the antiquarks except for collision energies GeV. The work is organized as follows: In Sec. II we briefly revisit the derivation of the fermion propagator in a rotating environment. In Sec. III we use this propagator to compute the interaction rate for a quark spin to align with the vorticity in a QCD rotating medium at finite temperature and baryo-chemical potential. In Sec. LABEL:IV we compute the relaxation time for values of and close to the chiral symmetry restoration/deconfinement transition and show that for quarks this relaxation time is within the putative life-time of the system produced in the reaction, although this is not the case for antiquarks except for large collision energies. We finally summarize and conclude in Sec. LABEL:concl.
II Propagator for a spin one-half fermion in a rotating environment
The physics within a relativistic rotating frame is most easily described in terms of a metric tensor resembling that of a curved space-time. We consider that the interaction region can be thought of as a rigid cylinder rotating around the -axis with constant angular velocity which is produced in semicentral collisions. We can thus write the metric tensor as
(1) |
A fermion with mass within the cylinder is described by the Dirac equation
(2) |
where is the affine connection. In this context, the -matrices in Eq. (2) correspond to the Dirac matrices in the rotating frame, which satisfy the usual anti-commutation relations
(3) |
The relation between the gamma matrices in the rotating frame and the usual gamma matrices are
(4) | ||||
In this notation, refers to the rotating frame while refers to the local rest frame. Therefore, Eq. (2) can be written as
(5) | ||||
In the Dirac representation,
(6) |
where is the Pauli matrix associated with the third component of the spin. Therefore, we can rewrite Eq. (5) as
(7) |
where
(8) |
This expression defines the total angular momentum in the direction. The term represents the orbital angular momentum, whereas is the spin. On the other hand, the term is the usual momentum operator. We can find solutions to Eq. (7) in the form
(9) |
and then, the function satisfies a Klein-Gordon like equation
(10) |
Notice that the spin operator when applied to produces eigenvalues 1/2. Consequently, conservation of the total angular momentum expressed in terms of the eigenvalues imposes solutions with for 1/2 and for 1/2. With these considerations, the solution of Eq. (10) can be written in cylindrical coordinates as
(11) |
where are Bessel functions of the first kind,
(12) |
is the transverse momentum squared and we have defined , representing the fermion energy observed from the inertial frame. Hence, the solution of Eq. (9) is
(13) | ||||
Before writing the expression for the fermion propagator in the rotating environment, it is important to highlight some features of the solution. First, causality requires that , where is the radius of the cylinder. Therefore, the solution is valid as long as . Second, we can simplify the solution assuming that the fermion is totally dragged by the vortical motion such that the angular position is determined by the product of the angular velocity and the time, specifically . This is a reasonable approximation when considering that during the early stages of a peripheral heavy-ion collision, particle interactions have not yet produced the development of a radial expansion. With this approximation, the propagator is translational invariant and can be simply Fourier transformed. With these features in mind, we write the fermion propagator as
(14) |
where
(15) |
In this last expression, and represent the eigenvalues and eigenvectors of Eq. (10). Taking as independent quantum numbers, the closure relation is written as
(16) |
Hence, Eq. (14) becomes
(17) |
where
(18) |
and we have defined
(19) | ||||
with
(20) |
Carrying out the integration and the summation, and taking the Fourier transform, we obtain
(21) |
We can write Eq. (21) in terms of the Dirac-gamma matrices as
(22) |
where
(23) |
is the spin projection operator. Notice that the derivation of the fermion propagator is performed in vacuum. To use this propagator including a finite temperature and chemical potential, recall that in equilibrium it is sufficiently general to make the replacement where are Matsubara frequencies for fermions. Equation (22) represents our approximation for the fermion propagator in a cylindrical rigidly rotating environment. We now use this propagator to compute the relaxation time for the fermion spin to align with the angular velocity in the rotating medium.
III Interaction rate for a quark spin to align with the angular velocity in a QCD rotating medium
In a QCD plasma in thermal equilibrium at temperature and baryon chemical potential , the interaction rates for a quark with spin components in the direction of and four-momentum to align its spin in the direction of the angular momentum vector can be expressed in terms of the total interaction rate, which in turn is given by the probability (per unit time) for a transition between the same quantum quark state , represented by properly normalized spinors with a definite spin projection () along the direction of the angular velocity. This transition is mediated by the imaginary part of the self-energy, Im. In symbols,
(24) |
Shuffling the indexes around, we can also write
(25) | |||||
where we used that the spin projection operators are given by
(26) |
To extract the creation rate for e for spin-aligned quark states from the total interaction rate, as discussed in Ref. LeBellac, we multiply the total interaction rate by the fermion distribution function , for a grand-canonical ensemble in the presence of a conserved charge to which a quark chemical potential is associated, namely,
(27) |
In previous analyses Ayala:2020ndx; Ayala:2019iin the interaction has been modeled using an effective vertex coupling the thermal vorticity and the quark spin. To improve the description, hereby we consider the case where the fermion is subject to the effect of a rotation within a rigid cylinder. The one-loop contribution to , depicted in Fig. 1, is given by
(28) |
where is the quark propagator in a rotating environment obtained in Eq. (22), is the effective gluon propagator in the thermal medium and are the group generators. The four-momenta are for the fermion and for the gluon, with being the gluon Matsubara frequencies. Also, and in the hard thermal loop (HTL) approximation is given by
(29) |
where are the polarization tensors for three-dimensional longitudinal and transverse gluons LeBellac. The gluon propagator functions for longitudinal and transverse modes, , are given by
(30) |
(31) | |||||
where
(32) |
and is the gluon thermal mass squared given by
(33) |
where and are the Casimir factors for the adjoint and fundamental representations of .
It is convenient to first look at the sum over Matsubara frequencies for the products of the propagator functions for longitudinal and transverse gluons, with , and the Matsubara propagator for the quark in a rotating environment , wich is described in Ref. LeBellac, can be obtained as the inverse of the denominator of each of the components of Eq. 22 with the replacement .
(34) |
The sum can be performed introducing the spectral densities and for the gluon and fermion, respectively. The imaginary part of can be written as
(35) | |||||
where is the Bose-Einstein distribution. The spectral densities are obtained from the imaginary part of after the analytic continuation and contain the discontinuities of the gluon propagator across the real axis. Their support depends on the ratio . For , have support on the (timelike) quasi-particle poles. For , their support coincides with the branch cut of and corresponds to Landau damping. On the other hand, the fermion spectral density is
(36) |
We now concentrate on the trace factors required for the computation of Eq. (27). The term proportional to the fermion momentum and angular velocity
(37) |
vanishes identically, whereas the terms proportional to the fermion mass are given by
(38) |
The delta functions in Eqs. (35) and (36) restrict the integration over gluon energies to the spacelike region, . Therefore, the parts of the gluon spectral densities that contribute to the interaction rate are given by
(39) |
(40) |
With all these ingredients we write the interaction rate as
{IEEEeqnarray}rCl
Γ^±(p_0) & = g2m CFπ2 ∫d3k(2π)3 ∫_-∞ ^∞dk02π ∫_-∞ ^∞dp_0 ^′
×
f(k_0)(4ρ_L(k_0) + 8ρ_T(k_0) )
×~f(p_0 ^′- μ∓Ω/ 2) δ(p_0 - k_0 - p_0 ^′)
× δ((p_0 ^′±Ω/2)^2 - E^2),
with
(41) |
Notice that
(42) |
Therefore, we can integrate Eq. (1) over to obtain
{IEEEeqnarray}rCl
Γ^±(p_0) & = g2m CFπ2 ∫d3k(2π)3 ∫_-∞ ^∞12Edk02π f(k_0)
× (4ρ_L(k_0) + 8ρ_T(k_0) )
× [~f(E - μ∓Ω)δ(p_0 - k_0 - E ±Ω/2)
+ ~f(-E - μ∓Ω)δ(p_0 - k_0 + E ±Ω/2)].
Notice that for the considered angular velocities appropriate to the early stages of the collision (10 MeV MeV), and for a strange quark mass MeV, the combination can always be safely regarded as being positive. The kinematical constraint imposed by the first of the delta functions in Eq. (III) corresponds to the rate to produce rotating and thermalized quarks originated by the dispersion of vacuum nonrotating quarks as a result of dispersion with medium quarks. This is depicted in Fig. 2.
We then single out this contribution from the total rate which can then be written as
{IEEEeqnarray}rCl
Γ^±(p_0) & = g2m CFπ2 ∫k2dk d(cosθ)dϕ(2π)3 ∫_-∞ ^∞12Edk02π
× (4ρ_L(k_0) + 8ρ_T(k_0) )δ(p_0 - k_0 - E ±Ω/2)
× f(k_0) ~f(E -μ∓Ω).
The kinematical restrictions for the integration
translate into integration regions . After integrating over the angle between and , and over the azimuthal angle , and finally using that , we obtain
{IEEEeqnarray}rCl
Γ^±(p_0) & = g2m CFπ2∫_0 ^∞dk k2(2π)3∫_R^±dk_0 f(k0)2pk
×