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

Polarized Photons from the Early Stages of Relativistic Heavy-Ion Collisions

Sigtryggur Hauksson [email protected] Institut de Physique Théorique, CEA/Saclay, Université Paris-Saclay, 91191, Gif sur Yvette, France    Charles Gale [email protected] Department of Physics, McGill University, 3600 University Street, Montréal, QC, Canada H3A 2T8
Abstract

The polarization of real photons emitted from early-time heavy-ion collisions is calculated, concentrating on the contribution from bremsstrahlung and quark-antiquark annihilation processes at leading order in the strong coupling. The effect of an initial momentum space anisotropy of the parton distribution is evaluated using a model for the non-equilibrium scattering kernel for momentum broadening. The effect on the photon polarization is reported for different degrees of anisotropy. The real photons emitted early during in-medium interactions will be dominantly polarized along the beam axis.

I Introduction

The theory of the nuclear strong interaction, QCD, features a transition from a phase where the relevant degrees of freedom are quarks and gluons – at high temperature – to one where the appropriate basis consists of composite hadrons at a lower temperature. This transition has been predicted by several theoretical approaches, including the non-perturbative field-theoretical framework of lattice QCD [[See, forexample, ][, andreferencestherein.]Aarts:2023vsf]. Decades of intense theoretical effort have revealed that the transition from confined hadrons to partons is not a thermodynamic phase transition in the proper sense but rather an analytic crossover [[See, forexample, ][, andreferencestherein.]Aoki:2006we]. On the experimental side, this exotic state of strongly interacting matter – the quark-gluon plasma (QGP) – has been observed in the relativistic collisions of nuclei (“heavy-ions”) performed at the Relativistic Heavy-Ion Collider (RHIC), and its existence has been later confirmed by experiments performed at the Large Hadron Collider (LHC) [3]. There also is strong evidence supporting the presence of QGP in smaller systems [4].

Even though new aspects of the QGP are continuously being discovered, it is fair to assert that the field is well poised to enter a phase of quantitative characterization, owing in large part to the large variety of experimental observables being measured by the experimental collaborations. Measurements designed to probe the QGP have reported results on soft hadron collective behavior [5], on QCD jet modification and energy loss [6], on photon [7] and dilepton [8] production, and on many other aspects as well. The theoretical tools developed to study the dynamics of nuclear collisions and the formation of the QGP typically consist of multistage models, rendered necessary by the complexity of the nuclear reaction. Recent examples of such a composite theoretical approach are studies [9, 10, 11, 12] where the initial state model (e.g. TRENTo [13], IP-Glasma [14]) preceded a fluid-dynamical phase ([15, 16, 17, 18]. A hadronic cascade afterburner (e.g. UrQMD [19], SMASH [20]), evolve the final state hadrons until their measurement. Such composite models can make statistically significant statements about transport parameters such as shear and bulk viscosity, as well as quantify the energy loss of energetic QCD jets [21].

As the multistage modeling of relativistic heavy-ion collisions covers a variety of dynamical conditions ranging from far from equilibrium initial states to almost ideal fluid dynamics, it is important to critically examine its different eras. In searching for observables capable of revealing the different modeling epochs, penetrating probes such as real and virtual photons impose themselves. Electromagnetic variables are emitted at all stages of the collision and as such can report on local conditions at their creation point [22]. The largest uncertainty in the chain of models currently lie at the beginning, in the time span preceding “hydrodynamization”. Early in the history of the collision, photon emission is liable to occur in media far from equilibrium which necessitate a dedicated theoretical treatment. The emission of photons from non-equilibrium environments has received some recent attention [23, 24].

This study will consider photon emission in the early stages of heavy-ion collisions, and will focus more specifically on the polarization states of those photons as a probe of the medium at early times. As our calculation only relies on the medium having a pressure anisotropy, or equivalently a momentum anisotropy in parton distribution, it holds both before hydrodynamization as well as in the beginning of the hydrodynamic stage while pressure anisotropy still persists. Some previous estimates for photon polarization at early times considered leading order direct photon production channels like those of the Compton process and qq¯q\bar{q} annihilation [25, 26, 27]. It is known that an equally important contribution as those two – at the same order in αs\alpha_{s}, the strong coupling constant – is that associated with the Landau-Pomeranchuk-Migdal effect (LPM) [28, 29, 30]. That contribution, evaluated for a medium out of equilibrium forms the basis of this work. It is fair to remind readers that the measurement of real photon polarization states is challenging, owing to the complications related to the external conversions into lepton pairs. The angular distribution of this pairs will reflect the polarization state. A more realistic proposition is a measurement of virtual photon polarization states, as measured through an internal conversion process leading to a dilepton final state. Consequently, the goal of our work is to first set the foundations for subsequent such evaluations and to perform a first estimate of the polarization signature of an early, non-equilibrium, strongly interacting medium.

Our paper is organized as follows: Section II lays out the building blocks of our non-equilibrium formalism. The section following that one discusses the collision kernel used to model the medium interactions. The numerical methods used to obtain the production rate of polarized photons are discussed in Section IV. Results and conclusion constitute Sections V and VI, respectively. We present analytical and numerical details in the Appendices.

II Polarized photon emission

Refer to caption
Figure 1: Our choice of coordinate system. The zz-axis is chosen to be along the beam axis. For a photon emitted at midrapidity we choose the xx-axis to be along its momentum. The photon (in red) can either be transversely polarized along the beam axis (zz-axis) or transversely polarized orthogonal to the beam axis, i.e. along the yy-axis.
Refer to caption
Figure 2: Photon radiation through medium-induced bremsstrahlung off a quark
Refer to caption
Figure 3: Photon radiation through medium-induced quark-antiquark pair annihilation

The quark-gluon plasma radiates photons through two different processes at leading order in perturbation theory. (See [31] for higher order corrections.) These processes are two-to-two scattering with a photon in the final state [32, 33], and bremsstrahlung and pair annihilation with a resulting photon [28, 30]. Two-to-two scattering on one hand and bremsstrahlung and pair annihilation on the other hand give roughly equal contribution to photon yield in the plasma [29]. While photons from two-to-two scattering have been studied extensively in an anisotropic medium and have been shown to be polarized [25], bremsstrahlung and pair annihilation photons have been studied much less in an anisotropic plasma due to the more complicated physics involved. We consider this now. Polarized photon emission has also been studied in other contexts, including in holography where a background magnetic field is included [34, 35, 36], due to vortical flow in the plasma [37], and due to the chiral magnetic effect [38, 39]. Dilepton polarization has furthermore been considered in [40, 41, 42]

In bremsstrahlung an on-shell photon is emitted collinearly off a quark or an antiquark, see Fig. 2. In vacuum this process would be kinematically forbidden as an on-shell quark cannot emit an on-shell photon. However, in a medium the process is made possible to due soft gluon kicks from the medium that bring the quark slightly off-shell. These kicks have momentum gΛ\sim g\Lambda where Λ\Lambda is a hard scale akin to temperature and gg is the coupling constant. The off-shellness of the quark is therefore of order P2g2Λ2P^{2}\sim g^{2}\Lambda^{2} meaning that the emission takes time tp/P21/g2Λt\sim p/P^{2}\sim 1/g^{2}\Lambda where pp is the quark momentum. During this time the quark can receive arbitrarily many soft gluon kicks since the mean-free time between two such kicks is also of order 1/g2Λ1/g^{2}\Lambda. All of these kicks need to be included at leading order in perturbation theory [28, 30, 43]. In an analogous fashion a quark-antiquark pair can annihilate and radiate a photon due to medium kicks, see Fig. 3.

We will now show that photons emitted through bremsstrahlung and pair annihilation are polarized in an anisotropic medium. This polarization can be described by extending the framework developed in [28, 30, 23], see App. A. To fix ideas we choose the coordinate system in Fig. 1. The zz-axis lies along the beam axis in a heavy-ion collision. We consider a photon at midrapidity and orient the coordinate system so that the xx-axis lies along its momentum 𝐤\mathbf{k}. The momentum of the photon can of course have any direction in the plane transverse to the beam axis; aligning it with the short axis of the plasma as in Fig. 1 is simply for illustration. 111In this work, the net polarization of photons emitted from a fluid cell is independent of the angular orientation in the transverse plane as we focus on the effect of longitudinal expansion. This could be generalized to include transverse expansion which breaks this symmetry. Our formalism could also easily be extended to photons at finite rapidity. As an on-shell photon is transversely polarized, the polarization basis can be choosen as ϵz=(0,0,0,1)\epsilon_{z}=(0,0,0,1) and ϵy=(0,0,1,0)\epsilon_{y}=(0,0,1,0). The photon is thus polarized along the beam axis or transverse to the beam axis.

In Appendix A we show that the rate of producing zz-polarized photons with momentum kk through bremsstrahlung is

kdΓzd3𝐤=6αEMsqs2(2π)3 20𝑑pxk8px2(k+px)\displaystyle k\frac{d\Gamma_{z}}{d^{3}\mathbf{k}}=\frac{6\alpha_{EM}\sum_{s}q_{s}^{2}}{(2\pi)^{3}}\,2\int_{0}^{\infty}dp_{x}\;\frac{k}{8p_{x}^{2}(k+p_{x})}
×nf(k+px)[1nf(px)](Fin(ζ)Az+Fout(ζ)Ay)\displaystyle\times n_{f}(k+p_{x})\left[1-n_{f}(p_{x})\right]\left(F_{\mathrm{in}}(\zeta)A_{z}+F_{\mathrm{out}}(\zeta)A_{y}\right) (1)

where momenta are defined in Fig. 2 and ζ=k/(px+k)\zeta=k/(p_{x}+k) is the momentum fraction of the photon. Similarly, the rate of producing yy-polarized photons is

kdΓyd3𝐤=6αEMsqs2(2π)320𝑑pxk8px2(k+px)\displaystyle k\frac{d\Gamma_{y}}{d^{3}\mathbf{k}}=\frac{6\alpha_{EM}\sum_{s}q_{s}^{2}}{(2\pi)^{3}}2\int_{0}^{\infty}dp_{x}\;\frac{k}{8p_{x}^{2}(k+p_{x})}
×nf(k+px)[1nf(px)](Fin(ζ)Ay+Fout(ζ)Az)\displaystyle\times n_{f}(k+p_{x})\left[1-n_{f}(p_{x})\right]\left(F_{\mathrm{in}}(\zeta)A_{y}+F_{\mathrm{out}}(\zeta)A_{z}\right) (2)

where AzA_{z} and AyA_{y} have been interchanged relative to Eq. (II). Here AzA_{z} and AyA_{y} quantify the amount of momentum broadening in the zz- and yy-direction and are defined as

Az\displaystyle A_{z} =Red2𝐩(2π)2 2pzfz(𝐩)\displaystyle=\mathrm{Re}\,\int\frac{d^{2}\mathbf{p}_{\perp}}{(2\pi)^{2}}\;2p_{z}f_{z}(\mathbf{p}_{\perp})
Ay\displaystyle A_{y} =Red2𝐩(2π)2 2pyfy(𝐩)\displaystyle=\mathrm{Re}\,\int\frac{d^{2}\mathbf{p}_{\perp}}{(2\pi)^{2}}\;2p_{y}f_{y}(\mathbf{p}_{\perp}) (3)

where 𝐟\mathbf{f} solves an integro-differential equation given below. Furthermore,

Fin(ζ)=(2ζ)2ζF_{\mathrm{in}}(\zeta)=\frac{(2-\zeta)^{2}}{\zeta} (4)

and

Fout(ζ)=ζ,F_{\mathrm{out}}(\zeta)=\zeta, (5)

are polarized splitting functions [44]. Finally, there are momentum factors nfn_{f} for the incoming and outgoing quarks. 222We have assumed that there is no chiral imbalance in the medium and that the baryon chemical potential vanishes, so that quarks and antiquarks of both helicities have the same momentum distribution nfn_{f}. The analogous expressions for quark-antiquark pair annihilation are given in Appendix A.

Refer to caption
Figure 4: Definition of the quantities 𝐧\mathbf{n} (the vector between the outgoing quark and photon) and ϕ𝐪\phi_{\mathbf{q}} (the angle defining a soft gluon kick of magnitude qq_{\perp}). The photon can be polarized in the yy- or the zz-directions which are transverse to its direction of motion, see also Fig. 1.

Eqs. (II) and (II) for polarized photon emission can be understood intuitively. We consider a zz-polarized photon for concreteness. As the photon travels in the xx-direction, the splitting plane of the photon and the outgoing quark is spanned by 𝐞^x\widehat{\mathbf{e}}_{x} and a vector orthogonal to 𝐞^x\widehat{\mathbf{e}}_{x} which we call 𝐧^\widehat{\mathbf{n}}, see Fig. 4. Eqs. (II) and (II) show that we can project the vector 𝐧^\widehat{\mathbf{n}} on to the yy- and the zz-axes and sum over the contributions. For 𝐧^\widehat{\mathbf{n}} in the zz-direction, the zz-polarized photon is polarized in the splitting plane and the hard splitting function is FinF_{\mathrm{in}}. Momentum broadening is quantified by the zz-component AzA_{z}. On the other hand, for 𝐧^\widehat{\mathbf{n}} in the yy-direction, the zz-polarized photon is polarized out of the splitting plane and the hard splitting function is FoutF_{\mathrm{out}}. Momentum broadening is then quantified by AyA_{y}.

The rate equations for photon emission depend on the function 𝐟=(fz,fy)\mathbf{f}=(f_{z},f_{y}) which quantifies momentum broadening. It solves the integro-differential equation

2𝐩=iδE𝐟(𝐩)+d2q(2π)2𝒞(𝐪)[𝐟(𝐩)𝐟(𝐩+𝐪)]2\mathbf{p}_{\perp}=i\delta E\mathbf{f}(\mathbf{p}_{\perp})+\int\frac{d^{2}q_{\perp}}{(2\pi)^{2}}\;\mathcal{C}(\mathbf{q}_{\perp})\left[\mathbf{f}(\mathbf{p}_{\perp})-\mathbf{f}(\mathbf{p}_{\perp}+\mathbf{q}_{\perp})\right] (6)

where

δE=k2p(p+k)[p2+m2].\delta E=\frac{k}{2p(p+k)}\left[p_{\perp}^{2}+m_{\infty}^{2}\right]. (7)

The central ingredient in this equation is the collision kernel 𝒞(𝐪)\mathcal{C}(\mathbf{q}_{\perp}) which gives the rate for a quark to receive soft gluon kicks of transverse momentum 𝐪\mathbf{q}_{\perp}. The collision kernel gives rise to a gain term and a loss term in Eq. (6).

In an isotropic medium, the collision kernel is by definition isotropic, 𝒞(𝐪)=𝒞(q)\mathcal{C}(\mathbf{q}_{\perp})=\mathcal{C}(q_{\perp}) and one can show that 𝐟=𝐩f^(p)\mathbf{f}=\mathbf{p}_{\perp}\widehat{f}(p_{\perp}). This means that Az=AyA_{z}=A_{y} and thus there is no net polarization of photons emitted. In an anisotropic medium, 𝒞(𝐪)\mathcal{C}(\mathbf{q}_{\perp}) depends not only on the magnitude of the kick qq_{\perp} but also on its orientation. In other words, when writing

𝐪=(qz,qy)=q(cosϕ𝐪,sinϕ𝐪)\mathbf{q}_{\perp}=\left(q_{z},q_{y}\right)=q_{\perp}(\cos\phi_{\mathbf{q}},\sin\phi_{\mathbf{q}}) (8)

the collision kernel depends on both ϕ𝐪\phi_{\mathbf{q}} and qq_{\perp}, see Fig. 4. This leads to 𝐟\mathbf{f} having more complicated angular dependence so that AzAyA_{z}\neq A_{y}. Therefore, photon emission from an anisotropic medium is polarized.

III Model of the collision kernel in an anisotropic plasma

As previously argued, the collision kernel for soft gluon kicks 𝒞(𝐪)\mathcal{C}(\mathbf{q}_{\perp}) is anisotropic in heavy-ion collisions which leads to polarization of photons emitted through bremsstrahlung. The ultimate source of the anisotropy in the kernel is longitudinal expansion of the medium along the beam axis. Such a longitudinal expansion gives pressure anisotropy at early and intermediate times with longitudinal pressure PLP_{L} less than transverse pressure PTP_{T} . On a microscopic level, this means that quark and gluon quasiparticles have an anisotropic momentum distribution with pz2<px2,py2\langle p_{z}^{2}\rangle<\langle p^{2}_{x}\rangle,\langle p^{2}_{y}\rangle. This is captured by the distribution introduced in Ref. [45]:

f(𝐩)=1+ξfiso(𝐩2+ξpz2)f(\mathbf{p})=\sqrt{1+\xi}f_{\mathrm{iso}}(\sqrt{\mathbf{p}^{2}+\xi p_{z}^{2}}) (9)

where fisof_{\mathrm{iso}} is an isotropic distribution, and ξ>0\xi>0 quantifies the degree of anisotropy. The prefactor 1+ξ\sqrt{1+\xi} ensures that the number density of quarks and gluons is the same as in equilibrium. A simple calculation gives the pressure anisotropy PT/PLP_{T}/P_{L} in terms of the momentum space anisotropy ξ\xi, linking the macroscopic and microscopic descriptions.333 Specifically, PT/PL=12ξ+(ξ1)arctanξarctanξξ/(1+ξ)P_{T}/P_{L}=\frac{1}{2}\frac{\sqrt{\xi}+(\xi-1)\arctan\sqrt{\xi}}{\arctan\sqrt{\xi}-\sqrt{\xi}/(1+\xi)}, see e.g. [46].

Ideally, one would want to calculate the collision kernel for momentum broadening, 𝒞(𝐪)\mathcal{C}(\mathbf{q}_{\perp}), directly from Eq. (9). Such a calculation would use that the hard quasiparticles in Eq. (9) radiate soft gluons which are then responsible for momentum broadening. This would allow to quantify the degree of photon polarization in a non-equilibrium medium from first principles. Unfortunately, going from Eq. (9) to the collision kernel is difficult in practice, partially due to instabilities that can be present in an anisotropic plasma [47].

In this study, we will use a simple model of the collision kernel in a longitudinally expanding medium. We take inspiration from results for the collision kernel in thermal equilibrium at leading order in perturbation theory [48],

𝒞eq(𝐪)=g2CFT(1q21q2+mD02)\mathcal{C}_{\mathrm{eq}}(\mathbf{q}_{\perp})=g^{2}C_{F}T\left(\frac{1}{q_{\perp}^{2}}-\frac{1}{q_{\perp}^{2}+m_{D0}^{2}}\right) (10)

Here mD02m_{D0}^{2} is the equilibrium Debye mass which describes screening of electric fields. (The equilibrium kernel has furthermore been evaluated at next-to-leading order [49], as well as on the lattice, see e.g. [50, 51].) In our anisotropic model, we replace the equilibrium Debye mass by its anisotropic extension444In [45] this quantity is referred to as m+2m_{+}^{2}. We have expanded in small ξ\xi in Eq. (III) but this is simply for convenience and not fundamental to the setup we use. found in [45]

mD2(ϕ𝐪)=(12ξ3)mD02+ξmD02cos2ϕ𝐪.m_{D}^{2}(\phi_{\mathbf{q}})=\left(1-\frac{2\xi}{3}\right)m_{D0}^{2}+\xi m_{D0}^{2}\cos^{2}\phi_{\mathbf{q}}. (11)

so that

𝒞(𝐪)=g2CFΛ(1q21q2+mD2(ϕ𝐪))\mathcal{C}(\mathbf{q}_{\perp})=g^{2}C_{F}\Lambda\Big{(}\frac{1}{q_{\perp}^{2}}-\frac{1}{q_{\perp}^{2}+m_{D}^{2}(\phi_{\mathbf{q}})}\Big{)} (12)

The anisotropic correction has an angular dependence with more broadening in the zz-direction than in the yy-direction. To simplify calculations, we expand the collision kernel in ξ\xi, writing

𝒞(𝐪)g2CFΛ\displaystyle\mathcal{C}(\mathbf{q}_{\perp})\approx g^{2}C_{F}\Lambda (1q21q2+mD02\displaystyle\left(\frac{1}{q_{\perp}^{2}}-\frac{1}{q_{\perp}^{2}+m_{D0}^{2}}\right.
+2ξmD02/3+ξmD02cos2ϕ𝐪(q2+mD02)2)\displaystyle+\left.\frac{-2\xi m_{D0}^{2}/3+\xi m_{D0}^{2}\cos^{2}\phi_{\mathbf{q}}}{\left(q_{\perp}^{2}+m_{D0}^{2}\right)^{2}}\right) (13)

where Λ\Lambda is a hard scale, akin to temperature.

Eq. (III) is a toy model for the collision kernel, intended to illustrate how an anisotropic kernel leads to polarized photon emission and to estimate the magnitude of this effect. This toy model only includes changes to the screening of electric fields in an anisotropic medium and not the myriad other non-equilibrium effects that can arise. Nevertheless, this collision kernel can be motivated by theoretical arguments making us belief that it captures some of the salient features of the full non-equilibrium kernel.

In general the collision kernel is defined by

𝒞(𝐪)=g2CFdq0dqx(2π)2Drrμν(Q)vμvν 2πδ(vQ)\mathcal{C}(\mathbf{q}_{\perp})=g^{2}C_{F}\int\frac{dq^{0}dq_{x}}{(2\pi)^{2}}\;D_{rr}^{\mu\nu}(Q)v_{\mu}v_{\nu}\,2\pi\delta(v\cdot Q) (14)

where vμ=(1,1,0,0)v^{\mu}=(1,1,0,0) is the four-velocity of the quark emitting a photon. The kernel depends on the statistical correlator for gluons in the medium,

Drr(Q):=12{A,A}(Q)=Dret(Q)Π(Q)Dadv(Q)D_{rr}(Q):=\frac{1}{2}\left\langle\left\{A,A\right\}\right\rangle(Q)=D_{\mathrm{ret}}(Q)\Pi(Q)D_{\mathrm{adv}}(Q) (15)

which characterizes the occupation density of a pair of soft gluons. We have omitted Lorentz indices for simplicity. This statistical correlator contains information on how the soft gluons are emitted by hard quasiparticles with rate Π(Q)\Pi(Q) and then propagate in the medium according to the retarded propagator Dret(Q)=i/(Q2Πret)D_{\mathrm{ret}}(Q)=i/(Q^{2}-\Pi_{\mathrm{ret}}) and the advanced propagator Dadv=DretD_{\mathrm{adv}}=-D_{\mathrm{ret}}^{*}.

Making some heuristic assumptions allows one to employ a sum rule in [48] to motivate the toy model for the collision kernel in Eq. (III), starting from the definitions in Eqs. (14) and (15). The goal is to include anisotropic corrections to the screening of chromoelectric fields, while ignoring anisotropic corrections to the density of gluons and to change in polarization that occurs during propagation. We work strictly at small anisotropy ξ1\xi\ll 1, including only effects of order 𝒪(ξ)\mathcal{O}(\xi).

The first heuristic assumption is to employ the identity Π(Q)=Λq02ImΠret\Pi(Q)=\frac{\Lambda}{q^{0}}2\mathrm{Im}\,\Pi_{\mathrm{\mathrm{ret}}} which is known as the KMS identity and which expresses detailed balance between production and decay of soft gluons. This is not strictly valid in a non-equilibrium medium and amounts to ignoring anisotropic corrections to the density of gluons. Then one can write

Drr(Q):=Λq0(DretDadv).D_{rr}(Q):=\frac{\Lambda}{q^{0}}\left(D_{\mathrm{ret}}-D_{\mathrm{adv}}\right). (16)

At small anisotropy the retarded propagator in Eq. (16) is

Dretμν(Q)PTμνQ2ΠT+PLμνQ2ΠL.D_{\mathrm{ret}}^{\mu\nu}(Q)\approx\frac{P_{T}^{\mu\nu}}{Q^{2}-\Pi_{T}}+\frac{P_{L}^{\mu\nu}}{Q^{2}-\Pi_{L}}. (17)

Here we have only included anisotropic corrections to the screening as given by ΠT\Pi_{T} and ΠL\Pi_{L}, see App. B.

Our second heuristic approximation is to focus on the anisotropic correction to ΠL\Pi_{L} and ignore those in ΠT\Pi_{T}. Comparing with the equilibrium calculation [48], this amounts to calculating anisotropic corrections to the term 1/(q2+mD02)1/(q_{\perp}^{2}+m_{D0}^{2}) in Eq. (10) while leaving the term 1/q21/q_{\perp}^{2} as is. This means that we include anisotropic corrections to the screening of chromoelectric fields as given by a Debye mass but do not include anistropic corrections to the screening of chromomagnetic fields.

The reason we use this approximation is that the sum rule we employ does not work for the transverse screening in ΠT\Pi_{T}. This is because the term 1/(Q2ΠT)1/(Q^{2}-\Pi_{T}) has a pole in the upper half complex plane of q0q_{0} corresponding to Weibel instabilities [52]. A formal use of the sum rule would lead to a contribution of the form 1/(q2m2)1/(q_{\perp}^{2}-m^{2}) which is ill-defined at q=mq_{\perp}=m. The solution to this issue is to use a retarded propagator for soft gluons that includes the mechanism by which non-Abelian interaction caps the growth of the unstable soft gluon modes. This is beyond the scope of this project.

Given these two approximations, one can use the sum rule in [48] nearly directly. The longitudinal retarded self-energy at small anisotropy ξ\xi is

ΠL(Q)=ΠL0(x)+ξ[16(1+3cos2θn)Q2𝐪2mD02\displaystyle\Pi_{L}(Q)=\Pi_{L}^{0}(x)+\xi\left[\frac{1}{6}\left(1+3\cos 2\theta_{n}\right)\frac{Q^{2}}{\mathbf{q}^{2}}m_{D0}^{2}\right.
+ΠL0(x)(cos2θnx22(1+3cos2θn))]\displaystyle\left.+\Pi_{L}^{0}(x)\left(\cos 2\theta_{n}-\frac{x^{2}}{2}\left(1+3\cos 2\theta_{n}\right)\right)\right] (18)

where θn\theta_{n} is the angle between 𝐪\mathbf{q} and the anisotropy vector 𝐧=𝐞z\mathbf{n}=\mathbf{e}_{z} which defines a preferred direction in Eq. (9) [45]. Here

ΠL0(x)=(x21)mD02[x2logx+1+iϵx1+iϵ1]\Pi_{L}^{0}(x)=\left(x^{2}-1\right)m_{D0}^{2}\left[\frac{x}{2}\log\frac{x+1+i\epsilon}{x-1+i\epsilon}-1\right] (19)

is the equilibrium value. We next do a change of variables in Eq. (14) to x=q0/q=qx/qx2+q2x=q_{0}/q=q_{x}/\sqrt{q_{x}^{2}+q_{\perp}^{2}} so that qx2=x2q2/(1x2)q_{x}^{2}=x^{2}q_{\perp}^{2}/(1-x^{2}) and q2=q2/(1x2)q^{2}=q_{\perp}^{2}/(1-x^{2}). Then one should substitute cos2θn=2(qz/q)21=2(1x2)cos2ϕ𝐪1\cos 2\theta_{n}=2(q_{z}/q)^{2}-1=2(1-x^{2})\cos^{2}\phi_{\mathbf{q}}-1 to get dependence only on xx, qq_{\perp} and ϕ𝐪\phi_{\mathbf{q}}. This gives that the longitudinal contribution to the collision kernel is

Λg2CFπ01dxxImΠ~L(x,ϕ𝐪)(𝐪2+ReΠ~L(x,ϕ𝐪))2+(ImΠ~L(x,ϕ𝐪))2.\frac{\Lambda g^{2}C_{F}}{\pi}\int_{0}^{1}\frac{dx}{x}\frac{\mathrm{Im}\,\widetilde{\Pi}_{L}(x,\phi_{\mathbf{q}})}{\left(\mathbf{q}_{\perp}^{2}+\mathrm{Re}\,\widetilde{\Pi}_{L}(x,\phi_{\mathbf{q}})\right)^{2}+\left(\mathrm{Im}\,\widetilde{\Pi}_{L}(x,\phi_{\mathbf{q}})\right)^{2}}. (20)

where

Π~L(x,ϕ𝐪)=ΠL0(x)+ξ[16(1+3cos2θn)(x21)mD02\displaystyle\widetilde{\Pi}_{L}(x,\phi_{\mathbf{q}})=\Pi_{L}^{0}(x)+\xi\left[\frac{1}{6}\left(1+3\cos 2\theta_{n}\right)(x^{2}-1)m_{D0}^{2}\right.
+ΠL0(x)(cos2θnx22(1+3cos2θn))]|cos2θn=2(1x2)cos2ϕ𝐪1\displaystyle\left.+\Pi_{L}^{0}(x)\left(\cos 2\theta_{n}-\frac{x^{2}}{2}\left(1+3\cos 2\theta_{n}\right)\right)\right]\bigg{\rvert}_{\cos 2\theta_{n}=2(1-x^{2})\cos^{2}\phi_{\mathbf{q}}-1} (21)

has no explicit dependence on qq_{\perp}.

@articleAurenche:2002pd, @articleAurenche:2002pd, An argument nearly identical555 The retarded propagator 1/[x2q2q2(1x2)ΠL(x,ϕ)]1/\left[x^{2}q_{\perp}^{2}-q_{\perp}^{2}-(1-x^{2})\Pi_{L}(x,\phi)\right] in Eq. 7 in [48] has an extra pole in the upper half complex plane in the anisotropic case. This pole can be seen by taking the limit x=k0/kx=k_{0}/k\rightarrow\infty in which case the propagator becomes 1/(1x2)/(q2+(1313ξcos2ϕ)mD02+ξ3x2cos2ϕmD02)\sim 1/(1-x^{2})/(q_{\perp}^{2}+\left(\frac{1}{3}-\frac{1}{3}\xi\cos^{2}\phi\right)m_{D0}^{2}+\frac{\xi}{3}x^{2}\cos^{2}\phi\,m_{D0}^{2}) which has a pole which is parametrically of the form x±iq/mDξx\sim\pm iq_{\perp}/m_{D}\sqrt{\xi} and thus far from the real axis when ξ1\xi\ll 1. This is not in contradiction with the usual properties of the retarded propagator as we have imposed q0=qxq_{0}=q_{x} and then search for poles in q0q_{0}. One can then show that the correction due to this pole to the sum rule in Eq. 9 in [48] is 𝒪(ξ3/2)\mathcal{O}(\xi^{3/2}) which is subleading to the 𝒪(ξ)\mathcal{O}(\xi) contributions we consider. to the one in [48] then shows that Eq. (20) is

g2CFΛ[1q2+limxΠ~L(x,ϕ)1q2+Π~L(0,ϕ)]\displaystyle g^{2}C_{F}\Lambda\left[\frac{1}{q_{\perp}^{2}+\lim_{x\rightarrow\infty}\widetilde{\Pi}_{L}(x,\phi)}-\frac{1}{q_{\perp}^{2}+\widetilde{\Pi}_{L}(0,\phi)}\right]
=g2CFΛ1q2+mD2(ϕ𝐪)\displaystyle=-g^{2}C_{F}\Lambda\frac{1}{q_{\perp}^{2}+m_{D}^{2}(\phi_{\mathbf{q}})} (22)

since Π~L(0,ϕ)=mD2(ϕ𝐪)\widetilde{\Pi}_{L}(0,\phi)=m_{D}^{2}(\phi_{\mathbf{q}}). Our conclusion is therefore that given our heuristic approximations, the collision kernel is given by Eq. (III). We emphasize that this collision kernel is not intended to capture all of the non-equilibrium physics but to focus on anisotropic corrections to the screening of chromoelectric fields.

IV Numerical method

We wish to evaluate the rate of polarized photon emission in an anisotropic medium such as that found in a longitudinally expanding quark-gluon plasma. The starting point is Eqs. (II) and (II) which require solving the integro-differential equation in Eq. (6) numerically, assuming the model for the collision kernel given in Eq. (III). To solve Eq. (6) we go to impact parameter space, i.e. the space Fourier conjugate to 𝐩\mathbf{p}_{\perp}. Defining

𝐟(𝐛)=d2p(2π)2ei𝐩𝐛𝐟(𝐩).\mathbf{f}(\mathbf{b})=\int\frac{d^{2}p_{\perp}}{(2\pi)^{2}}\;e^{i\mathbf{p}_{\perp}\cdot\mathbf{b}}\,\mathbf{f}(\mathbf{p}_{\perp}). (23)

the equation we wish to solve becomes

2ibδ(2)(𝐛)=ik2p(p+k)[b2+m2]𝐟(𝐛)+𝒞(𝐛)𝐟(𝐛).-2i\nabla_{b}\delta^{(2)}(\mathbf{b})=\frac{ik}{2p(p+k)}\left[-\nabla_{b}^{2}+m_{\infty}^{2}\right]\mathbf{f}(\mathbf{b})+\mathcal{C}(\mathbf{b})\mathbf{f}(\mathbf{b}). (24)

where

𝒞(𝐛)=d2p(2π)2[1ei𝐩𝐛]𝒞(𝐩).\mathcal{C}(\mathbf{b})=\int\frac{d^{2}p_{\perp}}{(2\pi)^{2}}\;\left[1-e^{i\mathbf{p}_{\perp}\cdot\mathbf{b}}\,\right]\mathcal{C}(\mathbf{p}_{\perp}). (25)

A straightforward calculation shows that the collision kernel from Eq. (III) is

𝒞(𝐛)=𝒞0(b)+ξ𝒞1(a)(b)+ξcos2β𝒞1(b)(b)\mathcal{C}(\mathbf{b})=\mathcal{C}_{0}(b)+\xi\,\mathcal{C}_{1}^{(a)}(b)+\xi\cos 2\beta\,\mathcal{C}_{1}^{(b)}(b) (26)

in impact parameter space where 𝐛=(bz,by)=(cosβ,sinβ)b\mathbf{b}=(b_{z},b_{y})=\left(\cos\beta,\sin\beta\right)b. The terms of the collision kernel are given by

𝒞0(𝐛)=g2CFT2π[K0(mD0b)+γE+logmD0b2],\mathcal{C}_{0}(\mathbf{b})=\frac{g^{2}C_{F}T}{2\pi}\left[K_{0}(m_{D0}b)+\gamma_{E}+\log\frac{m_{D0}b}{2}\right], (27)
𝒞1(a)(b)=g2CFT18πM2mD02(mD0bK1(mD0b)1)\mathcal{C}_{1}^{(a)}(b)=g^{2}C_{F}T\frac{1}{8\pi}\frac{M^{2}}{m_{D0}^{2}}\left(m_{D0}bK_{1}(m_{D0}b)-1\right) (28)

and

𝒞1(b)(b)=g2CFTM2b24π[2(mD0b)4\displaystyle\mathcal{C}_{1}^{(b)}(b)=g^{2}C_{F}T\frac{M^{2}b^{2}}{4\pi}\Bigg{[}\frac{2}{(m_{D0}b)^{4}}
12mD0bK1(mD0b)1(mD0b)2K2(mD0b)]\displaystyle-\frac{1}{2m_{D0}b}K_{1}(m_{D0}b)-\frac{1}{\left(m_{D0}b\right)^{2}}K_{2}(m_{D0}b)\Bigg{]} (29)

In order to solve Eq. (24) we do an expansion in small ξ\xi, giving

𝐟(𝐛)=𝐟0(𝐛)+ξ𝐟1(𝐛)+\mathbf{f}(\mathbf{b})=\mathbf{f}_{0}(\mathbf{b})+\xi\mathbf{f}_{1}(\mathbf{b})+\dots (30)

The zeroth order solution in ξ\xi satisfies the usual isotropic equation

ik2p(p+k)[b2+m2]𝐟0(𝐛)+𝒞0(b)𝐟0(𝐛)=2ibδ(2)(𝐛)\frac{ik}{2p(p+k)}\left[-\nabla_{b}^{2}+m_{\infty}^{2}\right]\mathbf{f}_{0}(\mathbf{b})+\mathcal{C}_{0}(b)\mathbf{f}_{0}(\mathbf{b})=-2i\nabla_{b}\delta^{(2)}(\mathbf{b}) (31)

and can be shown to have angular dependence 𝐟0(𝐛)(cosβ,sinβ)\mathbf{f}_{0}(\mathbf{b})\sim(\cos\beta,\sin\beta). The first order satisfies

ik2p(p+k)\displaystyle\frac{ik}{2p(p+k)} [b2+m2]𝐟1(𝐛)+𝒞0(b)𝐟1(𝐛)\displaystyle\left[-\nabla_{b}^{2}+m_{\infty}^{2}\right]\mathbf{f}_{1}(\mathbf{b})+\mathcal{C}_{0}(b)\mathbf{f}_{1}(\mathbf{b}) (32)
=[𝒞1(a)(b)+cos2β𝒞1(b)(b)]𝐟0(𝐛).\displaystyle=-\left[\mathcal{C}_{1}^{(a)}(b)+\cos 2\beta\,\mathcal{C}_{1}^{(b)}(b)\right]\mathbf{f}_{0}(\mathbf{b}).

Due to the angular dependence of the right hand side we can write in full generality

f1z(𝐛)=cosβf1(1z)(b)+cos3βf1(3)(b)f_{1z}(\mathbf{b})=\cos\beta f_{1}^{(1z)}(b)+\cos 3\beta\,f_{1}^{(3)}(b) (33)

and

f1y(𝐛)=sinβf1(1y)(b)+sin3βf1(3)(b)f_{1y}(\mathbf{b})=\sin\beta f_{1}^{(1y)}(b)+\sin 3\beta\,f_{1}^{(3)}(b) (34)

where these functions solve

𝒦[f1(1z)(𝐛)]+𝒞0(b)f1(1z)(𝐛)=[𝒞1(a)(b)+12𝒞1(b)(b)]f0,\mathcal{K}\left[f_{1}^{(1z)}(\mathbf{b})\right]+\mathcal{C}_{0}(b)f_{1}^{(1z)}(\mathbf{b})=-\left[\mathcal{C}_{1}^{(a)}(b)+\frac{1}{2}\mathcal{C}_{1}^{(b)}(b)\right]f_{0}, (35)
𝒦[f1(1y)(𝐛)]+𝒞0(b)f1(1y)(𝐛)=[𝒞1(a)(b)12𝒞1(b)(b)]f0\mathcal{K}\left[f_{1}^{(1y)}(\mathbf{b})\right]+\mathcal{C}_{0}(b)f_{1}^{(1y)}(\mathbf{b})=-\left[\mathcal{C}_{1}^{(a)}(b)-\frac{1}{2}\mathcal{C}_{1}^{(b)}(b)\right]f_{0} (36)

with

𝒦[𝐟(𝐛)]=ik2p(p+k)[d2db2+1bddb1b2m2]𝐟(𝐛).\mathcal{K}\left[\mathbf{f}(\mathbf{b})\right]=-\frac{ik}{2p(p+k)}\left[\frac{d^{2}}{db^{2}}+\frac{1}{b}\frac{d}{db}-\frac{1}{b^{2}}-m_{\infty}^{2}\right]\mathbf{f}(\mathbf{b}). (37)

(The differential equation for f1(3)f_{1}^{(3)} is given in App. C.)

Our goal is to evaluate

Az=2Imbzfz(𝐛)|b=0=2Im𝐛^𝐟0+ξf1(1z)b|b=0A_{z}=2\mathrm{Im}\,\partial_{b_{z}}f_{z}(\mathbf{b})\bigg{\rvert}_{b=0}=2\mathrm{Im}\,\frac{\mathbf{\widehat{b}}\cdot\mathbf{f}_{0}+\xi\,f_{1}^{(1z)}}{b}\bigg{\rvert}_{b=0} (38)

and

Ay=2Imbyfy(𝐛)|b=0=2Im𝐛^𝐟0+ξf1(1y)b|b=0A_{y}=2\mathrm{Im}\,\partial_{b_{y}}f_{y}(\mathbf{b})\bigg{\rvert}_{b=0}=2\mathrm{Im}\,\frac{\mathbf{\widehat{b}}\cdot\mathbf{f}_{0}+\xi\,f_{1}^{(1y)}}{b}\bigg{\rvert}_{b=0} (39)

which were defined in Eqs. (II) and (II). Thus, we only need to know f1(1z)(b)/bf_{1}^{(1z)}(b)/b and f1(1y)(b)/bf_{1}^{(1y)}(b)/b in the limit b0b\rightarrow 0 where it must be finite. This gives the boundary condition that This is done by demanding that f1(1z)f_{1}^{(1z)} and f1(1y)f_{1}^{(1y)} vanish at b=0b=0. The other boundary condition is that the functions vanish as bb\rightarrow\infty as can be seen from Eq. (23).

To evaluate f1(1z)(b)/bf_{1}^{(1z)}(b)/b and f1(1y)(b)/bf_{1}^{(1y)}(b)/b at b=0b=0, we demand that the functions f1(1z)f_{1}^{(1z)} and f1(1y)f_{1}^{(1y)} vanish at very large bb and evolve the functions numerically to small bb using Eqs. (35) and (36). In practice, this means that we start the evolution at a large but finite value of bb where f1(1z)f_{1}^{(1z)} and f1(1y)f_{1}^{(1y)} are initialized to a small value. A typical numerical solution for f11zf_{1}^{1z} and f11yf_{1}^{1y} then blows up as evolved towards b0b\rightarrow 0. We must then extract the finite part of our numerical solution. This is done by matching with known, analytic solutions of the differential equations in the small bb limit.

For instance, focusing on Eq. (35), we call the particular solution w(b)w(b) and the two independent solutions of the homogeneous equation w1(b)w_{1}(b) and w2(b)w_{2}(b). These are known analytically at small bb, see Appendix C. We can write our numerical solution in full generality at small bb as

f1(1z)(b)=w(b)+α1w1(b)+α2w2(b)f_{1}^{(1z)}(b)=w(b)+\alpha_{1}w_{1}(b)+\alpha_{2}w_{2}(b) (40)

where α1\alpha_{1} and α2\alpha_{2} are found numerically. To extract from this a solution with the right behaviour as b0b\rightarrow 0, one must in essence subtract a linear combination of w1w_{1} and w2w_{2} which satisfies the boundary condition at bb\rightarrow\infty. Then one is left with a solution which satisfies boundary conditions both at b=0b=0 and bb\rightarrow\infty and which gives f1(1z)(b)/bf_{1}^{(1z)}(b)/b at b=0b=0. This procedure is explained in further detail in Appendix C, see also [53, 48, 54] for earlier work in the isotropic case. A major difference with the isotropic case is that Eqs. (35) and (36) have a non-trivial right hand side which complicates the matching procedure. For instance, one must find an analytic solution w(b)w(b) of the full differential equations at small bb, including the right hand side. Furthermore, cancellation errors between solutions of the full differential equation and the homogeneous equation must carefully be avoided to get reliable results, see Appendix C.

V Results

Refer to caption
Figure 5: Spectrum of photons coming from bremsstrahlung and pair annihilation in a plasma at effective temperature Λ\Lambda. The anisotropic plasma has ξ=1.0\xi=1.0.
Refer to caption
Figure 6: Degree of polarization of photons emitted from bremsstrahlung and pair annihilation in an anisotropic plasma with ξ=1.0\xi=1.0. The quantity RR is defined in Eq. (43).

Figs. 5 and 6 are the main results of this work. They show the rate of photon production through bremsstrahlung and pair-annihilation in an anisotropic quark-gluon plasma with fixed anisotropy ξ\xi. The collision kernel is given by Eq. (III) while the momentum distribution of medium quarks is

f(𝐩)=1+ξep2+ξpz2/Λ+1f(\mathbf{p})=\frac{\sqrt{1+\xi}}{e^{\sqrt{p^{2}+\xi p_{z}^{2}}/\Lambda}+1} (41)

where Λ\Lambda can be seen as an effective temperature. Fig. 5 shows the total rate for producing photons at mid-rapidity and with momentum kk, i.e.

kdΓd3𝐤=kdΓzd3𝐤+kdΓyd3𝐤,k\frac{d\Gamma}{d^{3}\mathbf{k}}=k\frac{d\Gamma_{z}}{d^{3}\mathbf{k}}+k\frac{d\Gamma_{y}}{d^{3}\mathbf{k}}, (42)

where kdΓzd3𝐤k\frac{d\Gamma_{z}}{d^{3}\mathbf{k}} is the rate of producing photons polarized along the beam axis and kdΓyd3𝐤k\frac{d\Gamma_{y}}{d^{3}\mathbf{k}} is the rate for photons polarized orthogonal to the beam axis and to the photon momentum. Fig. 6 shows the degree of polarization at different momenta defined as

R=kdΓzd3𝐤kdΓyd3𝐤kdΓzd3𝐤+kdΓyd3𝐤.R=\frac{k\frac{d\Gamma_{z}}{d^{3}\mathbf{k}}-k\frac{d\Gamma_{y}}{d^{3}\mathbf{k}}}{k\frac{d\Gamma_{z}}{d^{3}\mathbf{k}}+k\frac{d\Gamma_{y}}{d^{3}\mathbf{k}}}. (43)

These quantities are shown for three values of anisotropy parameter, ξ=0.3\xi=0.3, ξ=0.6\xi=0.6 and ξ=0.9\xi=0.9, which correspond to pressure anisotropy of PL/PT0.81P_{L}/P_{T}\approx 0.81, PL/PT0.68P_{L}/P_{T}\approx 0.68 and PL/PT0.57P_{L}/P_{T}\approx 0.57 respectively. These are rather moderate values of pressure anisotropy which can be found in the early and intermediate stages of heavy-ion collisions.

The spectrum in an anisotropic medium is higher than that in an equilibrium medium at the same effective temperature Λ\Lambda, as can be seen in Fig. 5. This is due to the factor 1+ξ\sqrt{1+\xi} in the momentum distribution in Eq. (9) which increases the number of quarks with pz=0p_{z}=0 which can emit photons at midrapidity. This effect is partially compensated by anisotropic corrections which reduce the collision kernel 𝒞(𝐪)\mathcal{C}(\mathbf{q}_{\perp}) meaning that a given quark receives less momentum broadening, reducing the rate at which it emits photons collinearly.

More interesting is the polarization RR as a function of momentum kk. As seen in Fig. 6, the polarization has different signs for lower and higher values of the photon energy kk: it is along the beam axis for lower values of kk while it is orthogonal to the beam axis at higher values of kk. This owes to the interplay between bremsstrahlung and pair annihilation. As shown in Appendix A, because of the different polarized splitting functions, bremsstrahlung tends to give photons polarized along the zz-axis, while pair annihilation tends to give polarization along the xx-axis, see Fig. 1. As bremsstrahlung is suppressed at high kk (there are few medium quarks with energy higher than kk) xx-polarization dominates in that regime. On the contrary, pair annihilation is suppressed at low kk since the number of quarks with energy less than k/2k/2 is phase space suppressed. This gives zz-polarized photons in that regime.

Despite the complicated dependence of polarization on photon momentum, polarization along the beam axis dominates. This is simply because there are many more photons at lower kk and thus their polarization is dominant. Furthermore, in [25] it was shown that photons from two-to-two scattering, which are equally important as bremsstrahlung and pair annihilation photons, are also predominantly polarized along the beam axis, with an even greater magnitude of polarization. Thus a definite prediction of our work is that medium photons are polarized along the beam axis.

To make contact with potential experiments on photon polarization more work is needed. Firstly, all photon sources that cannot be subtracted in experiments need to be included, such as prompt photons and photons from the hadronic stage. Secondly, calculation of the rate of photon production in an anisotropic medium need to be folded with hydrodynamic or kinetic theory simulations of the medium to get a realistic evolution of the medium anisotropy.

VI Conclusion

In this work, we have calculated for the first time the degree of polarization for photons emitted in bremsstrahlung and off-shell pair annihilation processes in a hot medium consisting of quarks and gluons. Our evaluation includes the LPM regime and is at complete leading order in the strong coupling. The polarization of the real photons originating from bremsstrahlung and annihilation processes depends on the anisotropy of the original parton distribution, and therefore the polarization can instruct us on the dynamics in an environment that is not accessible to the vast majority of probes and observables measured in relativistic heavy-ion collisions. Specifically, it gives a measure of the pressure anisotropy at early times.

We trust that the methods and techniques developed and used here will be useful in the evaluations of polarization signatures of real and virtual photons, evaluated with scattering kernels for momentum broadening derived from microscopic theories and using time-evolution models based in QCD.

Acknowledgements.
C. G. acknowledges the support of the Natural Sciences and Engineering Research Council of Canada (NSERC), SAPIN-2020-00048.

Appendix A Derivation of rate equations

We wish to show that the total rate for polarized photon production through bremsstrahlung and pair annihilation can be written as

kdΓzd3𝐤\displaystyle k\frac{d\Gamma_{z}}{d^{3}\mathbf{k}} =6αEMsqs2(2π)3𝑑pzF(𝐤+𝐩)[1F(𝐩)]\displaystyle=\frac{6\alpha_{EM}\sum_{s}q_{s}^{2}}{(2\pi)^{3}}\int_{-\infty}^{\infty}dp^{z}\;F(\mathbf{k}+\mathbf{p})\left[1-F(\mathbf{p})\right] (44)
×1214(px)2(px+k)2[(2px+k)2Az+k2Ay]\displaystyle\times\frac{1}{2}\frac{1}{4(p^{x})^{2}(p^{x}+k)^{2}}\left[(2p^{x}+k)^{2}A_{z}+k^{2}A_{y}\right]

and

kdΓyd3𝐤\displaystyle k\frac{d\Gamma_{y}}{d^{3}\mathbf{k}} =6αEMsqs2(2π)3𝑑pzF(𝐤+𝐩)[1F(𝐩)]\displaystyle=\frac{6\alpha_{EM}\sum_{s}q_{s}^{2}}{(2\pi)^{3}}\int_{-\infty}^{\infty}dp^{z}\;F(\mathbf{k}+\mathbf{p})\left[1-F(\mathbf{p})\right] (45)
×1214(px)2(px+k)2[k2Az+(2px+k)2Ay]\displaystyle\times\frac{1}{2}\frac{1}{4(p^{x})^{2}(p^{x}+k)^{2}}\left[k^{2}A_{z}+(2p^{x}+k)^{2}A_{y}\right]

where AzA_{z} and AyA_{y} are defined in Eq. (II). Here the momentum distributions are contained in

F(px)=θ(px)f(𝐩)+θ(px)(1f(𝐩)).F(p^{x})=\theta(p^{x})f(\mathbf{p})+\theta(-p^{x})\left(1-f(-\mathbf{p})\right). (46)

Looking at the momentum factors in Eqs. (44) and (45), we see that bremsstrahlung off a quark or an antiquark corresponds to px>0p^{x}>0 and px<kp^{x}<-k. These cases can be rewritten to give Eqs. (II) and (II). Furthermore, k<px<0-k<p^{x}<0 corresponds to quark-antiquark pair annihilation. As the quark moves in the same direction as the photon we can set e.g. f(𝐩)=f(px)f(\mathbf{p})=f(p^{x}).

The derivation of polarized photon emission in Eqs. (44) and (45) is similar to that for unpolarized emission found in [28, 30, 23]. The polarized rate comes from evaluating the diagram in Fig. 8. The gluon ladders which represent soft kicks from the medium are evaluated in the same way as for unpolarized photon emission. Physically, this is because the soft kicks do not have enough energy to flip the helicity of quarks. The gluon ladders are summed to all orders to give the integral equation in Eq. (6). On the contrary, one must keep track of polarization at the hard emission vertices to evaluate the polarized emission rate.

For a bare quark loop without soft gluon kicks, Fig. 7, the hard emission vertices for a zz-polarized photon take the form

Hzz:=ϵμzϵνzTr[γμ(+)γν].H^{zz}:=\epsilon^{z}_{\mu}\epsilon^{z\,*}_{\nu}\mathrm{Tr}\,\left[\gamma^{\mu}(\not{K}+\not{P})\gamma^{\nu}\not{P}\right]. (47)

Here ϵμz\epsilon^{z}_{\mu} is the photon polarization tensor for zz-polarization. This trace is most easily evaluated by using that e.g.

=sus(𝐩)u¯s(𝐩)\not{P}=\sum_{s}u^{s}(\mathbf{p})\overline{u}^{s}(\mathbf{p}) (48)

where the sum is over spin states and

us(𝐩)=2p[1σ𝐩^2ξs1+σ𝐩^2ξs]u^{s}(\mathbf{p})=\sqrt{2p}\begin{bmatrix}\frac{1-\mathbf{\sigma}\cdot\widehat{\mathbf{p}}}{2}\xi^{s}\\ \frac{1+\mathbf{\sigma}\cdot\widehat{\mathbf{p}}}{2}\xi^{s}\end{bmatrix} (49)

are helicity states of quarks [55]. Here ξs\xi^{s} form a basis for two-component spin states. One can then show by an explicit calculation that

Hzz\displaystyle H^{zz} =s,t[γzus(𝐩+𝐤)u¯s(𝐩+𝐤)γzut(𝐩)u¯t(𝐩)]\displaystyle=\sum_{s,t}\left[\gamma^{z}u^{s}(\mathbf{p}+\mathbf{k})\overline{u}^{s}(\mathbf{p}+\mathbf{k})\gamma^{z}u^{t}(\mathbf{p})\overline{u}^{t}(\mathbf{p})\right] (50)
=8p(p+k)p2(p+k)2[(2p+k)2pz2+k2py2].\displaystyle=\frac{8p(p+k)}{p^{2}(p+k)^{2}}\left[(2p+k)^{2}p_{\perp\,z}^{2}+k^{2}p_{\perp\,y}^{2}\right].

Similarly, the hard emission vertices for emission of yy-polarized photons are

Hyy=8p(p+k)p2(p+k)2[(2p+k)2py2+k2pz2].H^{yy}=\frac{8p(p+k)}{p^{2}(p+k)^{2}}\left[(2p+k)^{2}p_{\perp\,y}^{2}+k^{2}p_{\perp\,z}^{2}\right]. (51)

This is the same as Eq. (50), except that pz2p_{\perp\,z}^{2} and py2p_{\perp\,y}^{2} have been interchanged.

To include soft gluon as in Fig. 8 one simply replaces one of the hard vertices by the dressed vertex 𝐟(𝐩)\mathbf{f}(\mathbf{p}_{\perp}) which includes gluons rungs and obeys the integral equation in Eq. (6). This means that in Eqs. (50) and (51) one replaces pz2pzfzp_{\perp\,z}^{2}\longrightarrow p_{\perp\,z}f_{z} and py2pyfyp_{\perp\,y}^{2}\longrightarrow p_{\perp\,y}f_{y}. This reproduces Eqs. (44) and (45).

Refer to caption
Figure 7: Definition of momenta in photon emission.
Refer to caption
Figure 8: Diagram for medium-induced photon emission. The gluon rungs are responsible for momentum broadening.

Appendix B The retarded gluon propagator at small anisotropy

In a system whose hard quasiparticles have the momentum distribution in Eq. (9), the retarded propagator for soft gluons is [45, 47]

Dretμν(Q)\displaystyle D_{\mathrm{ret}}^{\mu\nu}(Q) =(PTμνCμν)DretB\displaystyle=\left(P_{T}^{\mu\nu}-C^{\mu\nu}\right)D_{\mathrm{ret}}^{B}
+[(Q2Πc)PLμν+(Q2ΠL)Cμν+ΠdDμν]DretA\displaystyle+\left[\left(Q^{2}-\Pi_{c}\right)P_{L}^{\mu\nu}+\left(Q^{2}-\Pi_{L}\right)C^{\mu\nu}+\Pi_{d}D^{\mu\nu}\right]D_{\mathrm{ret}}^{A} (52)

where

DretA=1(Q2ΠL)(Q2Πc)Q2𝐪2q02Πd2D_{\mathrm{ret}}^{A}=\frac{1}{\left(Q^{2}-\Pi_{L}\right)\left(Q^{2}-\Pi_{c}\right)-\frac{Q^{2}\mathbf{q}^{2}}{q_{0}^{2}}\Pi_{d}^{2}} (53)

and

DretB=1Q2Πe.D_{\mathrm{ret}}^{B}=\frac{1}{Q^{2}-\Pi_{e}}. (54)

The tensors PTμνP_{T}^{\mu\nu} and PLμνP_{L}^{\mu\nu} are the same as in thermal equilibrium while CμνC^{\mu\nu} and DμνD^{\mu\nu} are new tensors that depend on the anisotropy vector 𝐧\mathbf{n}. The self-energy components Πe\Pi_{e}, ΠL\Pi_{L}, ΠT\Pi_{T} and Πd\Pi_{d} are given in [47] which contains further details.

As Πd=𝒪(ξ)\Pi_{d}=\mathcal{O}(\xi) we can approximate

DretA1(Q2ΠL)(Q2Πc)D_{\mathrm{ret}}^{A}\approx\frac{1}{\left(Q^{2}-\Pi_{L}\right)\left(Q^{2}-\Pi_{c}\right)} (55)

up to order 𝒪(ξ)\mathcal{O}(\xi). We furthermore ignore 𝒪(ξ)\mathcal{O}(\xi) corrections in the numerator as they do not describe anisotropic corrections to screening which are the non-equilibrium corrections we focus on. Then

Dretμν(Q)\displaystyle D_{\mathrm{ret}}^{\mu\nu}(Q) (PTμνCμν)DretB\displaystyle\approx\left(P_{T}^{\mu\nu}-C^{\mu\nu}\right)D_{\mathrm{ret}}^{B}
+[(Q2Πc)PLμν+(Q2ΠL)Cμν]DretA\displaystyle\qquad+\left[\left(Q^{2}-\Pi_{c}\right)P_{L}^{\mu\nu}+\left(Q^{2}-\Pi_{L}\right)C^{\mu\nu}\right]D_{\mathrm{ret}}^{A}
=PTμνCμνQ2Πe+PLμνQ2ΠL+CμνQ2Πc\displaystyle=\frac{P_{T}^{\mu\nu}-C^{\mu\nu}}{Q^{2}-\Pi_{e}}+\frac{P_{L}^{\mu\nu}}{Q^{2}-\Pi_{L}}+\frac{C^{\mu\nu}}{Q^{2}-\Pi_{c}} (56)

The terms with the tensor CμνC^{\mu\nu} go like

Cμν[1Q2Πe+1Q2Πc]=CμνΠcΠe(Q2Πe)(Q2Πc)C^{\mu\nu}\left[\frac{-1}{Q^{2}-\Pi_{e}}+\frac{1}{Q^{2}-\Pi_{c}}\right]=C^{\mu\nu}\frac{\Pi_{c}-\Pi_{e}}{(Q^{2}-\Pi_{e})(Q^{2}-\Pi_{c})} (57)

which can be ignored as ΠcΠe=𝒪(ξ)\Pi_{c}-\Pi_{e}=\mathcal{O}(\xi) in the numerator. We are thus left with

Dretμν(Q)PTμνQ2Πe+PLμνQ2ΠL.D_{\mathrm{ret}}^{\mu\nu}(Q)\approx\frac{P_{T}^{\mu\nu}}{Q^{2}-\Pi_{e}}+\frac{P_{L}^{\mu\nu}}{Q^{2}-\Pi_{L}}. (58)

at small anisotropy where we only include anisotropic corrections in the denominators. The tensors PLP_{L} and PTP_{T} are the same as in equilibrium while Πe\Pi_{e} and ΠL\Pi_{L} have anisotropic corrections. (We call Πe=ΠT\Pi_{e}=\Pi_{T} in the main text of the paper.)

Appendix C Details of numerical method

In this Appendix we discuss how to solve Eqs. (35) and (36) numerically. Unlike the isotropic equation, Eq. (31), the anisotropic equation has a non-vanishing source term on the right hand side for all values of bb. One thus needs a different numerical solution method than that developed in [53, 48, 54] for the equilibrium case. We note that the differential equation for the function f1(3)f_{1}^{(3)} is

𝒦[f1(3)(𝐛)]\displaystyle\mathcal{K}\left[f_{1}^{(3)}(\mathbf{b})\right] +ik2p(p+k)8b2f1(3)(𝐛)+𝒞0(b)f1(3)(𝐛)\displaystyle+\frac{ik}{2p(p+k)}\frac{8}{b^{2}}f_{1}^{(3)}(\mathbf{b})+\mathcal{C}_{0}(b)f_{1}^{(3)}(\mathbf{b})
=12𝒞1(b)(b)f0\displaystyle=-\frac{1}{2}\mathcal{C}_{1}^{(b)}(b)f_{0} (59)

but we will not discuss this function further as it is not needed to evaluate AzA_{z} and AyA_{y} in Eqs. (38) and (39)

For concreteness, we focus on solving Eq. (35). Defining a scaled function

G=π2kp(p+k)m2f1(1z)(b)b,G=\frac{\pi}{2}\frac{k}{p(p+k)m_{\infty}^{2}}\frac{f_{1}^{(1z)}(b)}{b}, (60)

as well as scaled collision kernels 𝒞¯(t)=2p(p+k)k1m2𝒞(b)\overline{\mathcal{C}}(t)=\frac{2p(p+k)}{k}\frac{1}{m_{\infty}^{2}}\mathcal{C}(b) and variable t=mbt=m_{\infty}b, this equation becomes

i\displaystyle-i [d2Gdt2+3tdGdtG]+𝒞¯0(t)G\displaystyle\left[\frac{d^{2}G}{dt^{2}}+\frac{3}{t}\frac{dG}{dt}-G\right]+\overline{\mathcal{C}}_{0}(t)G
=[𝒞¯1(a)(t)+12𝒞¯1(b)(t)]f¯0(t)\displaystyle=-\left[\overline{\mathcal{C}}_{1}^{(a)}(t)+\frac{1}{2}\overline{\mathcal{C}}_{1}^{(b)}(t)\right]\overline{f}_{0}(t) (61)

where f¯0(t)=π2kp(p+k)m2𝐛𝐟0/b2\overline{f}_{0}(t)=\frac{\pi}{2}\frac{k}{p(p+k)m_{\infty}^{2}}\mathbf{b}\cdot\mathbf{f}_{0}/b^{2}. As shown in [48], we can write f¯0(t)=K1(t)/t+f¯0rest(t)\overline{f}_{0}(t)=K_{1}(t)/t+\overline{f}_{0}^{\mathrm{rest}}(t) where f¯0rest(t)\overline{f}_{0}^{\mathrm{rest}}(t) is function that is finite in the limit t0t\rightarrow 0 and which we know numerically using the methods of [53, 48].

We need to solve Eq. (C), imposing the boundary conditions that G(t)0G(t)\rightarrow 0 as tt\rightarrow\infty, as well as that G(t)G(t) is finite as t0t\rightarrow 0. These boundary conditions are difficult to satisfy simultaneously for a numerical solution. Instead we find a numerical solutions g(t)g(t) of Eq. (C) with g(t)=0g(t\rightarrow\infty)=0 and a solution of the homogeneous equation without the source term that also satisfies g0(t)=0g_{0}(t\rightarrow\infty)=0. In general, both g(t)g(t) and g0(t)g_{0}(t) blow up as t0t\rightarrow 0. However, we know that the solution we are seeking can be written as

G(t)=g(t)+Ag0(t)G(t)=g(t)+Ag_{0}(t) (62)

where AA is chosen so that G(0)G(0) is finite.

We can find an explicit expression of AA by using analytic solutions of Eq. (C) for t1t\ll 1. In that limit the right hand side is

[𝒞¯1(a)(t)+12𝒞¯1(b)(t)]f¯0(t)\displaystyle\left[\overline{\mathcal{C}}_{1}^{(a)}(t)+\frac{1}{2}\overline{\mathcal{C}}_{1}^{(b)}(t)\right]\overline{f}_{0}(t)
[𝒞¯1(a)(t)+12𝒞¯1(b)(t)]K1(t)/ta+blogt\displaystyle\approx\left[\overline{\mathcal{C}}_{1}^{(a)}(t)+\frac{1}{2}\overline{\mathcal{C}}_{1}^{(b)}(t)\right]K_{1}(t)/t\approx a+b\log t (63)

where aa and bb are constants that depend on the momenta kk and pp as well as the masses mD2m_{D}^{2} and m2m_{\infty}^{2}. The differential equation becomes

i[d2gdt2+3tdgdtg]+𝒞¯0(t)g=a+blogt-i\left[\frac{d^{2}g}{dt^{2}}+\frac{3}{t}\frac{dg}{dt}-g\right]+\overline{\mathcal{C}}_{0}(t)g=a+b\log t (64)

which has general solution

g(t)=w(t)+α1w1(t)+α2w2(t)g(t)=w(t)+\alpha_{1}w_{1}(t)+\alpha_{2}w_{2}(t) (65)

where

w(t)=iaib(logt+2t2)w(t)=-ia-ib(\log t+\frac{2}{t^{2}}) (66)

is an exact particular solution and w2(t)=2J1(it)/(it)w_{2}(t)=2J_{1}(it)/(it) and w1(t)=π2Y1(it)/(it)14(2γE2log2+iπ)w2(t)w_{1}(t)=\frac{\pi}{2}Y_{1}(-it)/(-it)-\frac{1}{4}\left(2\gamma_{E}-2\log 2+i\pi\right)w_{2}(t) are solutions of the homogenous equation such that w1(t)=1/t2+12logt+𝒪(1)w_{1}(t)=1/t^{2}+\frac{1}{2}\log t+\mathcal{O}(1) and w2(t)=𝒪(1)w_{2}(t)=\mathcal{O}(1) for small tt. Similarly, the homogeneous equation can be written as

g0(t)=β1w1(t)+β2w2(t).g_{0}(t)=\beta_{1}w_{1}(t)+\beta_{2}w_{2}(t). (67)

The coefficients α1\alpha_{1}, α2\alpha_{2}, β1\beta_{1} and β2\beta_{2} can be found from Eqs. (65) and (67) by equating the numerical solutions g(t)g(t) and g0(t)g_{0}(t) and their derivatives with the analytic functions at some small value tmin1t_{\mathrm{min}}\ll 1.

The small tt behaviour of Eqs. (65) and (67) shows that

A=2ibα1β1A=\frac{2ib-\alpha_{1}}{\beta_{1}} (68)

which makes

G(0)=ia12ib+α2+(2ibα1)β2β1G(0)=-ia-\frac{1}{2}ib+\alpha_{2}+\frac{(2ib-\alpha_{1})\beta_{2}}{\beta_{1}} (69)

finite. This is the expression that we need in order to evaluate Eq. (38). All quantities are known for numerical solutions g(t)g(t) and g0(t)g_{0}(t).

Eq. (69) suffers from numerical cancellation errors in the terms α2α1β2/β1\alpha_{2}-\alpha_{1}\beta_{2}/\beta_{1}. It is thus better to rewrite these terms using Eqs. (65) and (67) which shows that

α2α1β2β1=gg0gg0+g0wg0wg0w2g0w2.\alpha_{2}-\frac{\alpha_{1}\beta_{2}}{\beta_{1}}=\frac{g^{\prime}g_{0}-gg_{0}^{\prime}+g_{0}^{\prime}w-g_{0}w^{\prime}}{g_{0}w_{2}^{\prime}-g_{0}^{\prime}w_{2}}. (70)

where all quantities are evaluated at tmin1t_{\mathrm{min}}\ll 1. The culprit behind cancellation errors is the term gg0gg0g^{\prime}g_{0}-gg_{0}^{\prime}. It can be evaluated more precisely by noting that

𝒢W(t)=g(t)g0(t)g(t)g0(t).\mathcal{G}_{W}(t)=g^{\prime}(t)g_{0}(t)-g(t)g_{0}^{\prime}(t). (71)

solves the equation

d(t3𝒢W)dt=it3[𝒞¯1(a)(t)+12𝒞¯1(b)(t)]f¯0(t)g0\frac{d(t^{3}\mathcal{G}_{W})}{dt}=-it^{3}\left[\overline{\mathcal{C}}_{1}^{(a)}(t)+\frac{1}{2}\overline{\mathcal{C}}_{1}^{(b)}(t)\right]\overline{f}_{0}(t)g_{0} (72)

which can be integrated to give 𝒢W(tmin)\mathcal{G}_{W}(t_{\mathrm{min}}) and thus a reliable value of G(0)G(0).

References