This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Relaxation time for the alignment between quark spin and
angular velocity in a rotating QCD medium

Alejandro Ayala1{}^{1}    Santiago Bernal-Langarica1{}^{1}    Isabel Domínguez Jiménez2{}^{2}    Ivonne Maldonado3{}^{3}    José Jorge Medina-Serna1{}^{1}    Javier Rendón1{}^{1}    María Elena Tejeda-Yeomans4{}^{4} 1{}^{1}Instituto de Ciencias Nucleares, Universidad Nacional Autónoma de México, Apartado Postal 70-543, CdMx 04510, Mexico.
2{}^{2}Facultad de Ciencias Físico-Matemáticas, Universidad Autánoma de Sinaloa, Avenida de las Américas y Boulevard Universitarios, Ciudad Universitaria, C.P. 80000, Culiacán, Sinaloa, México.
3{}^{3}Joint Institute for Nuclear Research, Dubna 141980, Russia.
4{}^{4}Facultad de Ciencias - CUICBAS, Universidad de Colima, Bernal Díaz del Castillo No. 340, Col. Villas San Sebastián, 28045 Colima, Mexico.
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 sNN50\sqrt{s_{NN}}\gtrsim 50 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 Λ\Lambda and Λ¯\overline{\Lambda} hyperons measured in semicentral collision for energies 2.5 GeV sNN27\lesssim\sqrt{s_{NN}}\lesssim 27 GeV, which shows an increasing trend as the energy and centrality of the collision decreases. The raising trend is different for Λ\Lambdas than Λ¯\overline{\Lambda}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 Λ\Lambda and Λ¯\overline{\Lambda} 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 (TT) and baryo-chemical potential (μB\mu_{B}). We show that for values of TT and μB\mu_{B} 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 sNN50\sqrt{s_{NN}}\gtrsim 50 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 TT and μB\mu_{B} 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 z^\hat{z}-axis with constant angular velocity Ω\Omega which is produced in semicentral collisions. We can thus write the metric tensor as

gμν=(1(x2+y2)Ω2yΩxΩ0yΩ100xΩ0100001).g_{\mu\nu}=\begin{pmatrix}1-(x^{2}+y^{2})\Omega^{2}&y\Omega&-x\Omega&0\\ y\Omega&-1&0&0\\ -x\Omega&0&-1&0\\ 0&0&0&-1\\ \end{pmatrix}. (1)

A fermion with mass mm within the cylinder is described by the Dirac equation

[iγμ(μ+Γμ)m]Ψ=0,\left[i\gamma^{\mu}\left(\partial_{\mu}+\Gamma_{\mu}\right)-m\right]\Psi=0, (2)

where Γμ\Gamma_{\mu} is the affine connection. In this context, the γμ\gamma^{\mu}-matrices in Eq. (2) correspond to the Dirac matrices in the rotating frame, which satisfy the usual anti-commutation relations

{γμ,γν}=2gμν.\{\gamma^{\mu},\gamma^{\nu}\}=2g^{\mu\nu}. (3)

The relation between the gamma matrices in the rotating frame and the usual gamma matrices are

γt=γ0,γx=γ1+yΩγ0,\displaystyle\gamma^{t}=\gamma^{0},\;\;\;\;\;\;\;\gamma^{x}=\gamma^{1}+y\Omega\gamma^{0}, (4)
γz=γ3,γy=γ2xΩγ0.\displaystyle\gamma^{z}=\gamma^{3},\;\;\;\;\;\;\;\gamma^{y}=\gamma^{2}-x\Omega\gamma^{0}.

In this notation, μ={t,x,y,z}\mu=\{t,x,y,z\} refers to the rotating frame while μ={0,1,2,3}\mu=\{0,1,2,3\} refers to the local rest frame. Therefore, Eq. (2) can be written as

[iγ0\displaystyle\Big{[}i\gamma^{0} (txΩy+yΩxi2Ωσ12)\displaystyle\left(\partial_{t}-x\Omega\partial_{y}+y\Omega\partial_{x}-\frac{i}{2}\Omega\sigma^{12}\right) (5)
+iγ1x+iγ2y+iγ3zm]Ψ=0.\displaystyle+i\gamma^{1}\partial_{x}+i\gamma^{2}\partial_{y}+i\gamma^{3}\partial_{z}-m\Big{]}\Psi=0.

In the Dirac representation,

σ12=(σ300σ3),\sigma^{12}=\begin{pmatrix}\sigma^{3}&0\\ 0&\sigma^{3}\end{pmatrix}, (6)

where σ3=diag(1,1)\sigma^{3}=\mbox{diag}(1,-1) is the Pauli matrix associated with the third component of the spin. Therefore, we can rewrite Eq. (5) as

[iγ0(t+ΩJ^z)+iγm]Ψ=0,\displaystyle\left[i\gamma^{0}\left(\partial_{t}+\Omega\hat{J}_{z}\right)+i\vec{\gamma}\cdot\vec{\nabla}-m\right]\Psi=0, (7)

where

J^zL^z+S^z=i(xyyx)+12σ12.\hat{J}_{z}\equiv\hat{L}_{z}+\hat{S}_{z}=-i(x\partial_{y}-y\partial_{x})+\frac{1}{2}\sigma^{12}. (8)

This expression defines the total angular momentum in the z^\hat{z} direction. The term L^z\hat{L}_{z} represents the orbital angular momentum, whereas S^z\hat{S}_{z} is the spin. On the other hand, the term i-i\vec{\nabla} is the usual momentum operator. We can find solutions to Eq. (7) in the form

Ψ(x)=[iγ0(t+ΩJ^z)+iγ+m]ϕ(x),\Psi(x)=\left[i\gamma^{0}\left(\partial_{t}+\Omega\hat{J}_{z}\right)+i\vec{\gamma}\cdot\vec{\nabla}+m\right]\phi(x), (9)

and then, the function ϕ(x)\phi(x) satisfies a Klein-Gordon like equation

[(it+ΩJ^z)2+2x+2y+2zm2]ϕ(x)=0.\left[\left(i\partial_{t}+\Omega\hat{J}_{z}\right)^{2}+\partial^{2}_{x}+\partial^{2}_{y}+\partial^{2}_{z}-m^{2}\right]\phi(x)=0. (10)

Notice that the spin operator S^z\hat{S}_{z} when applied to ϕ(x)\phi(x) produces eigenvalues s=±s=\pm1/2. Consequently, conservation of the total angular momentum expressed in terms of the eigenvalues j=s+lj=s+l imposes solutions with ll for s=s=1/2 and l+1l+1 for s=s=-1/2. With these considerations, the solution of Eq. (10) can be written in cylindrical coordinates (t,x,y,z)(t,ρsinφ,ρcosφ,z)(t,x,y,z)\to(t,\rho\sin\varphi,\rho\cos\varphi,z) as

ϕ(x)=(Jl(kρ)Jl+1(kρ)eiφJl(kρ)Jl+1(kρ)eiφ)eEt+ikzz+ilφ,\phi(x)=\begin{pmatrix}J_{l}(k_{\perp}\rho)\\ J_{l+1}(k_{\perp}\rho)e^{i\varphi}\\ J_{l}(k_{\perp}\rho)\\ J_{l+1}(k_{\perp}\rho)e^{i\varphi}\end{pmatrix}e^{-Et+ik_{z}z+il\varphi}, (11)

where JlJ_{l} are Bessel functions of the first kind,

k2=E~2kz2m2,\displaystyle k_{\perp}^{2}=\tilde{E}^{2}-k_{z}^{2}-m^{2}, (12)

is the transverse momentum squared and we have defined E~E+jΩ\tilde{E}\equiv E+j\>\Omega, representing the fermion energy observed from the inertial frame. Hence, the solution of Eq. (9) is

Ψ(x)=\displaystyle\Psi(x)= ([E+jΩ+mkz+ik]Jl(kρ)[E+jΩ+m+kzik]Jl+1(kρ)eiφ[EjΩ+mkz+ik]Jl(kρ)[EjΩ+m+kzik]Jl+1(kρ)eiφ)\displaystyle\begin{pmatrix}\left[E+j\Omega+m-k_{z}+ik_{\perp}\right]J_{l}(k_{\perp}\rho)\\ \left[E+j\Omega+m+k_{z}-ik_{\perp}\right]J_{l+1}(k_{\perp}\rho)e^{i\varphi}\\ \left[-E-j\Omega+m-k_{z}+ik_{\perp}\right]J_{l}(k_{\perp}\rho)\\ \left[-E-j\Omega+m+k_{z}-ik_{\perp}\right]J_{l+1}(k_{\perp}\rho)e^{i\varphi}\end{pmatrix} (13)
×e(E+jΩ)t+ikzz+ilφ.\displaystyle\times e^{-(E+j\Omega)t+ik_{z}z+il\varphi}.

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 ΩR<1\Omega R<1, where RR is the radius of the cylinder. Therefore, the solution is valid as long as Ω<1/R\Omega<1/R. 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 φ+Ωt=0\varphi+\Omega t=0. 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 S(x,x)S(x,x^{\prime}) as

S(x,x)=[iγ0(t+ΩJ^z)+iγ+m]G(x,x),S(x,x^{\prime})=\left[i\gamma^{0}\left(\partial_{t}+\Omega\hat{J}_{z}\right)+i\vec{\gamma}\cdot\vec{\nabla}+m\right]G(x,x^{\prime}), (14)

where

G(x,x)=(i)0dτλexp([iτλ])ϕλ(x)ϕλ(x).G(x,x^{\prime})=(-i)\int_{-\infty}^{0}d\tau\sum_{\lambda}\exp{\left[-i\tau\lambda\right]}\phi_{\lambda}(x)\phi_{\lambda}^{\dagger}(x). (15)

In this last expression, λ\lambda and ϕλ(x)\phi_{\lambda}(x) represent the eigenvalues and eigenvectors of Eq. (10). Taking E,k,kz,lE,k_{\perp},k_{z},l as independent quantum numbers, the closure relation is written as

λϕλ(x)ϕλ(x)=l=dEdkzdkk(2π)3ϕ(x)ϕ(x)=(1010010110100101)δ4(xx).\sum_{\lambda}\phi_{\lambda}(x)\phi_{\lambda}^{\dagger}(x)=\sum_{l=\infty}^{\infty}\int\frac{dEdk_{z}dk_{\perp}k_{\perp}}{(2\pi)^{3}}\phi(x)\phi^{\dagger}(x)\\ =\begin{pmatrix}1&0&1&0\\ 0&1&0&1\\ 1&0&1&0\\ 0&1&0&1\end{pmatrix}\delta^{4}(x-x^{\prime}). (16)

Hence, Eq. (14) becomes

S(x,x)=l=dEdkzdkk(2π)3Φ(ρ,ρ)ei(E(l+1/2)Ω)(tt)eikz(zz)eil(φφ)E2kz2m2k2+iϵ,S(x,x^{\prime})=\sum_{l=\infty}^{\infty}\int\frac{dEdk_{z}dk_{\perp}k_{\perp}}{(2\pi)^{3}}\Phi(\rho,\rho^{\prime})\frac{e^{-i(E-(l+1/2)\Omega)(t-t^{\prime})}e^{ik_{z}(z-z^{\prime})}e^{il(\varphi-\varphi^{\prime})}}{E^{2}-k_{z}^{2}-m^{2}-k_{\perp}^{2}+i\epsilon}, (17)

where

Φ(ρ,ρ)=(𝒜𝒥l,l𝒜𝒥l,l+1eiφ𝒜𝒥l,l𝒜𝒥l,l𝒥l,l+1eiφ𝒥l+1,l+1ei(φφ)𝒥l+1,leiφ𝒥l+1,l+1ei(φφ)𝒞𝒥l,l𝒞𝒥l,l+1eiφ𝒞𝒥l,l𝒞𝒥l,l𝒟𝒥l,l+1eiφ𝒟𝒥l+1,l+1ei(φφ)𝒟𝒥l+1,leiφ𝒟𝒥l+1,l+1ei(φφ)),\Phi(\rho,\rho^{\prime})=\begin{pmatrix}\mathcal{A}\mathcal{J}_{l,l}&\mathcal{A}\mathcal{J}_{l,l+1}e^{-i\varphi^{\prime}}&\mathcal{A}\mathcal{J}_{l,l}&\mathcal{A}\mathcal{J}_{l,l}\\ \mathcal{B}\mathcal{J}_{l,l+1}e^{i\varphi}&\mathcal{B}\mathcal{J}_{l+1,l+1}e^{i(\varphi-\varphi^{\prime})}&\mathcal{B}\mathcal{J}_{l+1,l}e^{i\varphi}&\mathcal{B}\mathcal{J}_{l+1,l+1}e^{i(\varphi-\varphi^{\prime})}\\ \mathcal{C}\mathcal{J}_{l,l}&\mathcal{C}\mathcal{J}_{l,l+1}e^{-i\varphi^{\prime}}&\mathcal{C}\mathcal{J}_{l,l}&\mathcal{C}\mathcal{J}_{l,l}\\ \mathcal{D}\mathcal{J}_{l,l+1}e^{i\varphi}&\mathcal{D}\mathcal{J}_{l+1,l+1}e^{i(\varphi-\varphi^{\prime})}&\mathcal{D}\mathcal{J}_{l+1,l}e^{i\varphi}&\mathcal{D}\mathcal{J}_{l+1,l+1}e^{i(\varphi-\varphi^{\prime})}\end{pmatrix}, (18)

and we have defined

𝒜\displaystyle\mathcal{A} [E+mkz+ik],\displaystyle\equiv\left[E+m-k_{z}+ik_{\perp}\right], (19)
\displaystyle\mathcal{B} [E+m+kzik],\displaystyle\equiv\left[E+m+k_{z}-ik_{\perp}\right],
𝒞\displaystyle\mathcal{C} [E+mkzk+ik],\displaystyle\equiv\left[-E+m-k_{z}k+ik_{\perp}\right],
𝒟\displaystyle\mathcal{D} [E+m+kzkik],\displaystyle\equiv\left[-E+m+k_{z}k-ik_{\perp}\right],

with

𝒥l,lJl(kρ)Jl(kρ).\mathcal{J}_{l,l^{\prime}}\equiv J_{l}(k_{\perp}\rho)J_{l^{\prime}}(k_{\perp}\rho^{\prime}). (20)

Carrying out the integration and the summation, and taking the Fourier transform, we obtain

S(p)=(p0+Ω/2pz+m+ip(p0+Ω/2)2p2m2+iϵ0p0+Ω/2pz+m+ip(p0+Ω/2)2p2m2+iϵ00p0+Ω/2+pz+mip(p0Ω/2)2p2m2+iϵ0p0+Ω/2+pz+mip(p0Ω/2)2p2m2+iϵ(p0+Ω/2)pz+m+ip(p0+Ω/2)2p2m2+iϵ0(p0+Ω/2)pz+m+ip(p0+Ω/2)2p2m2+iϵ00(p0+Ω/2)+pz+mip(p0Ω/2)2p2m2+iϵ0(p0+Ω/2)+pz+mip(p0Ω/2)2p2m2+iϵ).S(p)=\begin{pmatrix}\frac{p_{0}+\Omega/2-p_{z}+m+ip_{\perp}}{\left(p_{0}+\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}&0&\frac{p_{0}+\Omega/2-p_{z}+m+ip_{\perp}}{\left(p_{0}+\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}&0\\ 0&\frac{p_{0}+\Omega/2+p_{z}+m-ip_{\perp}}{\left(p_{0}-\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}&0&\frac{p_{0}+\Omega/2+p_{z}+m-ip_{\perp}}{\left(p_{0}-\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}\\ \frac{-(p_{0}+\Omega/2)-p_{z}+m+ip_{\perp}}{\left(p_{0}+\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}&0&\frac{-(p_{0}+\Omega/2)-p_{z}+m+ip_{\perp}}{\left(p_{0}+\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}&0\\ 0&\frac{-(p_{0}+\Omega/2)+p_{z}+m-ip_{\perp}}{\left(p_{0}-\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}&0&\frac{-(p_{0}+\Omega/2)+p_{z}+m-ip_{\perp}}{\left(p_{0}-\Omega/2\right)^{2}-p^{2}-m^{2}+i\epsilon}\end{pmatrix}. (21)

We can write Eq. (21) in terms of the Dirac-gamma matrices as

S(P)=(p0+Ω/2pz+ip)(γ0+γ3)+m(1+γ5)(p0+Ω/2)2p2m2+iϵ𝒪++(p0Ω/2+pzip)(γ0γ3)+m(1+γ5)(p0Ω/2)2p2m2+iϵ𝒪,S(P)=\frac{\left(p_{0}+\Omega/2-p_{z}+ip_{\perp}\right)\left(\gamma_{0}+\gamma_{3}\right)+m\left(1+\gamma_{5}\right)}{(p_{0}+\Omega/2)^{2}-p^{2}-m^{2}+i\epsilon}\mathcal{O}^{+}+\frac{(p_{0}-\Omega/2+p_{z}-ip_{\perp})(\gamma_{0}-\gamma_{3})+m(1+\gamma_{5})}{(p_{0}-\Omega/2)^{2}-p^{2}-m^{2}+i\epsilon}\mathcal{O}^{-}, (22)

where

𝒪±=12[1±iγ1γ2]\mathcal{O}^{\pm}=\frac{1}{2}\left[1\pm i\gamma^{1}\gamma^{2}\right] (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 p0iω~n+μp_{0}\to i\tilde{\omega}_{n}+\mu where ωn=(2n+1)πT\omega_{n}=(2n+1)\pi T 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 TT and baryon chemical potential μB\mu_{B}, the interaction rates Γ±\Gamma^{\pm} for a quark with spin components s=±1/2s=\pm 1/2 in the direction of Ω\vec{\Omega} and four-momentum P=(p0,p)P=(p_{0},\vec{p}) 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 u±u^{\pm}, represented by properly normalized spinors with a definite spin projection (±\pm) along the direction of the angular velocity. This transition is mediated by the imaginary part of the self-energy, ImΣ\Sigma. In symbols,

Γ±(p0)u¯a±ImΣabub±.\Gamma^{\pm}(p_{0})\sim\bar{u}_{a}^{\pm}\text{Im}\Sigma_{ab}u_{b}^{\pm}. (24)

Shuffling the indexes around, we can also write

Γ±(p0)\displaystyle\Gamma^{\pm}(p_{0}) \displaystyle\sim ub±u¯a±ImΣab\displaystyle u_{b}^{\pm}\bar{u}_{a}^{\pm}\text{Im}\Sigma_{ab} (25)
=\displaystyle= Tr[𝒪±ImΣab],\displaystyle\text{Tr}[\mathcal{O}^{\pm}\text{Im}\Sigma_{ab}],

where we used that the spin projection operators 𝒪±\mathcal{O}^{\pm} are given by

𝒪±u±u¯±.\displaystyle\mathcal{O}^{\pm}\equiv u^{\pm}\bar{u}^{\pm}. (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 f~\tilde{f}, for a grand-canonical ensemble in the presence of a conserved charge to which a quark chemical potential μ=1/3μB\mu=1/3\mu_{B} is associated, namely,

Γ±(p0)=f~(p0μΩ/2)Tr[O±ImΣ].\displaystyle\Gamma^{\pm}(p_{0})=\tilde{f}(p_{0}-\mu\mp\Omega/2)\text{Tr}\left[O^{\pm}\;\text{Im}\Sigma\right]. (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 Σ\Sigma, depicted in Fig. 1, is given by

{feynhand}\vertex\vertex\vertex\vertex\vertex\vertex\propagPP\propagPKP-K\propagPKP-K\propagPP\propagKK\propagKK
Figure 1: One-loop quark self-energy diagram that defines the kinematics. The gluon line with a blob represents the effective gluon propagator at finite density and temperature. The open circle on the fermion propagator represents the effect of the rotating environment.
Σ(P)=Tnd3k(2π)3γμTaS(PK)γνTbGμνab,\Sigma(P)=T\sum_{n}\int\frac{d^{3}k}{(2\pi)^{3}}\gamma^{\mu}T_{a}S(P-K)\gamma^{\nu}T_{b}\;^{*}G_{\mu\nu}^{ab}, (28)

where SS is the quark propagator in a rotating environment obtained in Eq. (22), Gμνab{}^{*}G_{\mu\nu}^{ab} is the effective gluon propagator in the thermal medium and TaT_{a} are the SU(3)SU(3) group generators. The four-momenta are P=(iω~,p)P=(i\tilde{\omega},\vec{p}) for the fermion and K=(iωn,k)K=(i\omega_{n},\vec{k}) for the gluon, with ωn\omega_{n} being the gluon Matsubara frequencies. Also, Gμνab=Gμνδab{}^{*}G_{\mu\nu}^{ab}=^{*}\!\!G_{\mu\nu}\delta^{ab} and in the hard thermal loop (HTL) approximation Gμν{}^{*}G_{\mu\nu} is given by

Gμν(K)=ΔL(K)PLμν+ΔT(K)PTμν{}^{*}G_{\mu\nu}(K)=\Delta_{L}(K)P_{L\;\mu\nu}+\Delta_{T}(K)P_{T\;\mu\nu} (29)

where PL,TμνP_{L,T\;\mu\nu} are the polarization tensors for three-dimensional longitudinal and transverse gluons LeBellac. The gluon propagator functions for longitudinal and transverse modes, ΔL,T(K)\Delta_{L,T}(K), are given by

ΔL1(K)=K2+2mT2K2k2[1iωnkQ0(iωnk)],\Delta_{L}^{-1}(K)=K^{2}+2m_{T}^{2}\frac{K^{2}}{k^{2}}\left[1-\frac{i\omega_{n}}{k}Q_{0}\left(\frac{i\omega_{n}}{k}\right)\right], (30)
ΔT1(K)\displaystyle\Delta_{T}^{-1}(K) =\displaystyle= K2mT2(iωnk)\displaystyle-K^{2}-m_{T}^{2}\left(\frac{i\omega_{n}}{k}\right) (31)
×\displaystyle\times {[1(iωnk)2]Q0(iωnk)+(iωnk)},\displaystyle\left\{\left[1-\left(\frac{i\omega_{n}}{k}\right)^{2}\right]Q_{0}\left(\frac{i\omega_{n}}{k}\right)+\left(\frac{i\omega_{n}}{k}\right)\right\},

where

Q0(x)=12ln(x+1x1),Q_{0}(x)=\frac{1}{2}\ln\left(\frac{x+1}{x-1}\right), (32)

and mT2m_{T}^{2} is the gluon thermal mass squared given by

mT2=16g2𝒞AT2+112g2𝒞F(T2+3π2μ2),m_{T}^{2}=\frac{1}{6}g^{2}\mathcal{C}_{A}T^{2}+\frac{1}{12}g^{2}\mathcal{C}_{F}(T^{2}+\frac{3}{\pi^{2}}\mu^{2}), (33)

where 𝒞A\mathcal{C}_{A} and 𝒞F\mathcal{C}_{F} are the Casimir factors for the adjoint and fundamental representations of SU(3)SU(3).

It is convenient to first look at the sum over Matsubara frequencies for the products of the propagator functions for longitudinal and transverse gluons, Δi\Delta_{i} with i=L,Ti=L,T, and the Matsubara propagator for the quark in a rotating environment Δ~\tilde{\Delta}, 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 p0iω~n+μp_{0}\to i\tilde{\omega}_{n}+\mu.

Si(iω)=TnΔi(iωn)Δ~(i(ωωn)).S_{i}(i\omega)=T\sum_{n}\Delta_{i}(i\omega_{n})\tilde{\Delta}(i(\omega-\omega_{n})). (34)

The sum can be performed introducing the spectral densities ρi\rho_{i} and ρF\rho_{F} for the gluon and fermion, respectively. The imaginary part of SiS_{i} can be written as

Im(Si)\displaystyle\text{Im}(S_{i})\!\!\! =\displaystyle= π(eβ(p0μΩ/2)+1)dk02πdp02πf(k0)\displaystyle\!\!\!\pi(e^{\beta(p_{0}-\mu-\Omega/2)}+1)\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}\frac{dk_{0}}{2\pi}\frac{dp^{\prime}_{0}}{2\pi}f(k_{0}) (35)
×\displaystyle\times f~(p0μΩ/2)δ(p0k0p0)\displaystyle\!\!\tilde{f}(p^{\prime}_{0}-\mu\mp\Omega/2)\delta(p_{0}-k_{0}-p^{\prime}_{0})
×\displaystyle\times ρi(k0,k)ρF(p0,pk),\displaystyle\rho_{i}(k_{0},k)\rho_{F}(p^{\prime}_{0},p-k),

where f(k0)f(k_{0}) is the Bose-Einstein distribution. The spectral densities ρi\rho_{i} are obtained from the imaginary part of Δi(iωn)\Delta_{i}(i\omega_{n}) after the analytic continuation iωnk0+iϵi\omega_{n}\rightarrow k_{0}+i\epsilon and contain the discontinuities of the gluon propagator across the real k0k_{0} axis. Their support depends on the ratio x=k0/kx=k_{0}/k. For |x|>1|x|>1, ρi\rho_{i} have support on the (timelike) quasi-particle poles. For |x|<1|x|<1, their support coincides with the branch cut of Q0(x)Q_{0}(x) and corresponds to Landau damping. On the other hand, the fermion spectral density is

ρF(p0,p)=2πδ((p0±Ω/2)2p2m2),\rho_{F}(p_{0}^{\prime},p)=-2\pi\delta\left((p_{0}^{\prime}\pm\Omega/2)^{2}-p^{2}-m^{2}\right),\\ (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

PL,TμνTr[γμ(γ0±γ3)(1±iγ1γ2)γν]=0,\displaystyle P_{L,T\;\mu\nu}\text{Tr}\left[\gamma^{\mu}\left(\gamma^{0}\pm\gamma^{3}\right)\left(1\pm i\gamma^{1}\gamma^{2}\right)\gamma^{\nu}\right]=0, (37)

vanishes identically, whereas the terms proportional to the fermion mass are given by

PLμνTr[γμ(1+γ5)(1±iγ1γ2)γν]\displaystyle P_{L\;\mu\nu}\text{Tr}\left[\gamma^{\mu}\left(1+\gamma^{5}\right)\left(1\pm i\gamma^{1}\gamma^{2}\right)\gamma^{\nu}\right] =\displaystyle= 4\displaystyle-4
PTμνTr[γμ(1+γ5)(1±iγ1γ2)γν]\displaystyle P_{T\;\mu\nu}\text{Tr}\left[\gamma^{\mu}\left(1+\gamma^{5}\right)\left(1\pm i\gamma^{1}\gamma^{2}\right)\gamma^{\nu}\right] =\displaystyle= 8.\displaystyle-8. (38)

The delta functions in Eqs. (35) and (36) restrict the integration over gluon energies to the spacelike region, |x|<1|x|<1. Therefore, the parts of the gluon spectral densities that contribute to the interaction rate are given by

ρL(k0,k)=2πmT2xθ(1x2)[k2+2mT2(1x2ln|x+1x1|)]2+π2mT4x2,\rho_{L}(k_{0},k)=\frac{2\pi m_{T}^{2}x\theta(1-x^{2})}{\left[k^{2}+2m_{T}^{2}\left(1-\frac{x}{2}\ln\left|\frac{x+1}{x-1}\right|\right)\right]^{2}+\pi^{2}m_{T}^{4}x^{2}}, (39)
ρT(k0,k)=2πmT2x(1x2)θ(1x2)2[(k2(x21)mT2[x2+x(1x2)2ln|x+1x1|])2+π24mT4x2(1x2)2].\rho_{T}(k_{0},k)=\frac{2\pi m_{T}^{2}x(1-x^{2})\theta(1-x^{2})}{2\left[\left(k^{2}(x^{2}-1)-m_{T}^{2}\left[x^{2}+\frac{x(1-x^{2})}{2}\ln\left|\frac{x+1}{x-1}\right|\right]\right)^{2}+\frac{\pi^{2}}{4}m_{T}^{4}x^{2}(1-x^{2})^{2}\right]}. (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

E2=|pk|2m2.E^{2}=|\vec{p}-\vec{k}|^{2}-m^{2}. (41)

Notice that

δ((p0±Ω/2)2E2)=\displaystyle\delta\left(\left(p_{0}^{\prime}\pm\Omega/2\right)^{2}-E^{2}\right)=
12E[δ(p0±Ω/2E)+δ(p0±Ω/2+E)].\displaystyle\frac{1}{2E}\Big{[}\delta(p_{0}^{\prime}\pm\Omega/2-E)+\delta(p_{0}^{\prime}\pm\Omega/2+E)\Big{]}. (42)

Therefore, we can integrate Eq. (1) over p0p_{0}^{\prime} 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 Ω14\lesssim\Omega\lesssim 14 MeV), and for a strange quark mass m100m\sim 100 MeV, the combination E±Ω/2E\pm\Omega/2 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.

{feynhand}\vertex\vertex\vertex\vertex\vertex\vertex\propagPP\propag\propagPKP-K\propag\propagKK
Figure 2: Feynman diagram representing a process whereby an initially nonrotating quark is dragged by the medium and aligns its spin either parallel or antiparallel to the angular velocity by means of its interactions with medium particles mediated by soft thermal gluons.

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 k0k_{0} integration translate into integration regions ±\mathcal{R^{\pm}}. After integrating over the angle θ\theta between p\vec{p} and k\vec{k}, and over the azimuthal angle ϕ\phi, and finally using that E2=p2+m2=|pk|2m2=p2+k22pkcos(θ)+m2E^{2}=p^{2}+\ m^{2}=|\vec{p}-\vec{k}|^{2}-m^{2}=p^{2}+k^{2}-2pk\cos{\theta}+m^{2}, we obtain {IEEEeqnarray}rCl Γ^±(p_0) & = g2m CFπ2∫_0 ^∞dk k2(2π)3∫_R^±dk_0 f(k0)2pk
×