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

Vector meson’s spin alignments in high energy reactions

Jin-Hui Chen [email protected] Key Laboratory of Nuclear Physics and Ion-beam Application (MoE), Institute of Modern Physics, Fudan University, Shanghai 200433, China    Zuo-Tang Liang [email protected] Institute of Frontier and Interdisciplinary Science, Key Laboratory of Particle Physics and Particle Irradiation (MOE), Shandong University, Qingdao, Shandong 266237, China    Yu-Gang Ma [email protected] Key Laboratory of Nuclear Physics and Ion-beam Application (MoE), Institute of Modern Physics, Fudan University, Shanghai 200433, China    Xin-Li Sheng [email protected] INFN-Firenze, 50019 Sesto Fiorentino FI, Italy    Qun Wang [email protected] Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China School of Mechanics and Physics, Anhui University of Science and Technology, Huainan, Anhui 232001, China
Abstract

The global spin alignment of vector mesons has been observed by the STAR collaboration at the Relativistic Heavy Ion Collider (RHIC) at Brookhaven National Laboratory (BNL). It provides a unique opportunity to probe the correlation between the polarized quark and antiquark in the strongly coupled quark-gluon plasma (sQGP) produced in relativistic heavy ion collisions, opening a new window to explore the properties of sQGP. In addition, spin alignments of vector mesons have also been observed in other high-energy particle collisions such as e+ee^{+}e^{-} annihilations at high energies where hadron production is dominated by quark fragmentation mechanism. The results obtained are quite different from those obtained in heavy ion collisions where quark coalescence/combination mechanism dominates suggesting strong dependence on hadronization mechanisms. So comprehensive studies in different hadronization processes are needed. In this article, we present a brief review of theoretical and experimental advances in the study of vector meson’s spin alignments in a variety of high-energy particle collisions, with emphasis on hadronization mechanisms.

Keywords: Vector meson, spin alignment, global polarization, relativistic heavy ion collisions, fragmentation
PACS: 12.38.Mh,12.38.–t, 13.88.+e, 13.66.Bc, 25.75.–q, 25.75.Nq

I Introduction

Quantum Chromodynamics (QCD) is the fundamental theory for strong interaction that is responsible for binding quarks and gluons together to form protons, neutrons, and other hadrons. Relativistic Heavy Ion Collisions (RHIC) are a powerful tool for studying QCD and properties of strongly interacting matter under extreme conditions Braun-Munzinger:2015hba ; Florkowski:2018fap ; Ko ; Rapp ; Zhang1 ; ZhaoJ . In these collisions, heavy nuclei are accelerated to nearly light speed and then collide to achieve immense energy density that recreates conditions similar to those just microseconds after the Big Bang, allowing scientists to study quark-gluon plasma (QGP) - a state of matter in which quarks and gluons are decoupled from hadrons Pandav:2022xxx ; HeWB ; Chen:2023mel ; MaYG ; PangLG ; Sun1 ; Chen1 .

Spin is an intrinsic form of angular momentum carried by elementary particles, and it is a (pseudo-)vector quantity that can point to different directions. As elementary particles of strong interaction, quarks and gluons carry spins. As composite particles of strong interaction, many hadrons also carry spins. The spin polarization refers to the alignment of the spin along a specific direction (called spin quantization direction). Spin degrees of freedom have been playing an important role in the development of modern physics since its discovery in 1925. The global polarization effect (GPE) of the quark gluon plasma (QGP) produced in high energy heavy-ion collisions is a new spin effect in particle and nuclear physics. The theoretical prediction Liang:2004ph ; Liang:2004xn ; Gao:2007bc was made almost two decades ago and attracted immediate attention Voloshin:2004ha ; Betz:2007kg ; Ipp:2007ng ; Becattini:2007sr ; Liang:2007ma ; Chen:2007zzq ; Selyuzhenkov:2007ab ; STAR:2008lcm ; STAR:2007ccu . However the enthusiasm was soon dampened by STAR’s earlier attempts Chen:2007zzq ; Selyuzhenkov:2007ab ; STAR:2008lcm ; STAR:2007ccu that gave null results for global Λ\Lambda-hyperon polarization as well as global spin alignments of vector mesons due to limited statistics in data at that time. The earlier results show that the GPE, even if it exists, could be very small, so it is unclear if it is within the scope and resolution of current experiments. The enthusiasm was aroused by non-vanishing results for the GPE for Λ\Lambda hyperons in STAR’s beam-energy-scan experiments STAR:2017ckg , which showed that the GPE decreases monotonically with collision energies. Now the GPE has grown to be a sub-field in heavy-ion collisions, see Refs. Wang:2017jpl ; Liang:2019clf ; Gao:2020lxh ; Huang:2020dtn ; Gao:2020vbh ; Becattini:2020ngo ; Becattini:2022zvf ; Becattini:2024uha for recent reviews.

Another surprise came five years later from STAR’s measurements on the global spin alignment of ϕ\phi mesons with high statistics STAR:2022fan . The STAR’s results STAR:2022fan seem to conflict with hyperon’s polarization as predicted in Ref. Liang:2004xn that the vector meson’s spin alignment is proportional to quark polarization squared and thus should be much smaller than that observed in STAR’s experiment STAR:2022fan . So STAR’s results are definitely non-trivial and may have deep implication for hadronization mechanism and properties of QCD Yang:2017sdk ; Sheng:2019kmk ; Sheng:2022ffb ; Sheng:2022wsy . When a system of particles exhibits spin polarization, it means that the spins of these particles are aligned more often in a particular direction than randomly oriented. Now the global spin alignment of vector mesons has attracted a broad interest in experimental and theoretical communities Wang:2023fvy ; Chen:2023hnb ; Xin-Li:2023gwh ; Jian-Hua:2023cna ; Li-Juan:2023bws ; Chen:2024bik . Spin polarization in the context of QCD is a critical aspect of understanding the internal structure and dynamics of hadrons. Through both theoretical and experimental approaches, physicists aim to unravel the complexities of how quarks and gluons contribute to the spin and other properties of hadrons. Several theoretical interpretations have been proposed and new measurements are underway.

In addition to relativistic heavy-ion collisions, the vector meson’s spin alignment has also been studied in other high energy processes such as e+ee^{+}e^{-}, epe^{-}p and pppp collisions DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs ; ZEUS:1999wqh ; NOMAD:2006kuc ; STAR:2008lcm . The results show quite different features and theoretical efforts have also been made to describe them Anselmino:1997ui ; Anselmino:1998jv ; Anselmino:1999cg ; Xu:2001hz ; Xu:2003fq ; Chen:2016moq ; Chen:2016iey ; Chen:2020pty . It is therefore desirable to summarize these experimental and theoretical results and make comparison among different high energy reactions as guidance for future studies.

Firstly, we briefly review spin density matrices for vector mesons as well as measurement methods for the spin alignment, and then review experimental results in different high energy collisions. Secondly, we summarize theoretical approaches in two different hadronization mechanisms as well as the linear response theory for the spin alignment of vector mesons in thermalized QGP. Finally, we present a short summary and outlook.

II Spin density matrices and measurement methods for spin alignment

II.1 Spin density matrices

The spin polarization of particles produced in high energy reaction can be described by the spin density matrix ρ^\hat{\rho}. For particles with spin-1/21/2 such as quarks and anti-quarks, ρ^q\hat{\rho}_{q} is a 2×22\times 2 Hermitian matrix which can be expanded as

ρ^q=12(1+Pqσ),\displaystyle\hat{\rho}_{q}=\frac{1}{2}(1+\vec{P}_{q}\cdot\vec{\sigma}), (1)

where σ=(σx,σy,σz)\vec{\sigma}=(\sigma_{x},\sigma_{y},\sigma_{z}) are Pauli matrices, and Pq=Tr(σρ^q)\vec{P}_{q}={\rm Tr}(\vec{\sigma}\hat{\rho}_{q}) is the mean spin polarization vector in the quark’s rest frame.

For spin-1 particles, the spin density matrix is a 3×33\times 3 Hermitian matrix. With the spin quantization axis specified, the spin states are denoted as |jm|jm\rangle and the spin density matrix can be put into the form Bacchetta:2000jk

ρ^V=\displaystyle\hat{\rho}_{V}= (ρ11ρ10ρ11ρ01ρ00ρ01ρ11ρ10ρ11)\displaystyle\left(\begin{array}[]{ccc}\rho_{11}&\rho_{10}&\rho_{1-1}\\ \rho_{01}&\rho_{00}&\rho_{0-1}\\ \rho_{-11}&\rho_{-10}&\rho_{-1-1}\end{array}\right) (5)
=\displaystyle= 13+12PiΣi+TijΣij,\displaystyle\frac{1}{3}+\frac{1}{2}P_{i}\Sigma_{i}+T_{ij}\Sigma_{ij}, (6)

where i,ji,j=1,2,3, and Σi\Sigma_{i} and Σij\Sigma_{ij} are 3×\times3 traceless matrices defined as Xin-Li:2023gwh ; Becattini:2024uha

Σ1=\displaystyle\Sigma_{1}= 12(010101010),\displaystyle\frac{1}{\sqrt{2}}\left(\begin{array}[]{ccc}0&1&0\\ 1&0&1\\ 0&1&0\end{array}\right), (10)
Σ2=\displaystyle\Sigma_{2}= 12(0i0i0i0i0),\displaystyle\frac{1}{\sqrt{2}}\left(\begin{array}[]{ccc}0&-i&0\\ i&0&-i\\ 0&i&0\end{array}\right), (14)
Σ3=\displaystyle\Sigma_{3}= (100000001),\displaystyle\left(\begin{array}[]{ccc}1&0&0\\ 0&0&0\\ 0&0&-1\end{array}\right), (18)
Σij=\displaystyle\Sigma_{ij}= 12(ΣiΣj+ΣjΣi)23δij.\displaystyle\frac{1}{2}(\Sigma_{i}\Sigma_{j}+\Sigma_{j}\Sigma_{i})-\frac{2}{3}\delta_{ij}. (19)

The polarization vector εm\vec{\varepsilon}_{m} in the rest frame of the spin-1 particle in the spin state |1m|1m\rangle can be defined as

εm=0=z^,\displaystyle\vec{\varepsilon}_{m=0}=\hat{z}, (20)
εm=±1=12(x^±iy^),\displaystyle\vec{\varepsilon}_{m=\pm 1}=\mp\frac{1}{\sqrt{2}}(\hat{x}\pm i\hat{y}), (21)

where the spin quantization direction is assumed to be z^\hat{z} direction. For the measurement of global spin alignment, the spin quantization direction is taken as the direction of global orbital angular momentum, i.e., εm=0=y^\vec{\varepsilon}_{m=0}=\hat{y}, while other polarization vectors are determined by conditions εmεm=δmm\vec{\varepsilon}_{m}^{\,*}\cdot\vec{\varepsilon}_{m^{\prime}}=\delta_{mm^{\prime}} and εm=±1=εm=1\vec{\varepsilon}_{m=\pm 1}^{\,*}=-\vec{\varepsilon}_{m=\mp 1}. Another widely used choice is the direction of meson’s three-momentum in the laboratory frame, εm=0=p/|p|\vec{\varepsilon}_{m=0}=\vec{p}/|\vec{p}|, corresponding to the helicity polarization. The spin density matrix for different choices of spin quantization directions are related by the following transformation in the spin space,

ρ^V=Rρ^VR,\hat{\rho}_{V}=R\,\hat{\rho}^{\prime}_{V}\,R^{\dagger}, (22)

where ρ^V\hat{\rho}_{V} and ρ^V\hat{\rho}^{\prime}_{V} are two spin density matrices corresponding to two set of polarization vectors εm=0,±1\vec{\varepsilon}_{m=0,\pm 1} and εm=0,±1\vec{\varepsilon}^{\,\prime}_{m=0,\pm 1}, respectively. The transformation matrix RR is a 3×33\times 3 matrix in the spin space with elements given by the inner product of polarization vectors, RmmεmεmR_{mm^{\prime}}\equiv\vec{\varepsilon}^{\,*}_{m}\cdot\vec{\varepsilon}^{\,\prime}_{m^{\prime}}.

The spin vector S\vec{S} is related to the polarization vector ε\vec{\varepsilon} by S=Im(εm×εm)\vec{S}={\rm Im}({\vec{\varepsilon}_{m}}^{\,*}\times\vec{\varepsilon}_{m}). If the vector meson is in the pure spin state with m=0m=0, the spin vector is vanishing, S=ε0×ε0=0\vec{S}=\vec{\varepsilon}_{0}\times\vec{\varepsilon}_{0}=0. If the vector meson is in the pure spin state with m=±1m=\pm 1 we have S=Im(ε±1×ε±1)=±z^\vec{S}={\rm Im}({\vec{\varepsilon}_{\pm 1}}^{\,*}\times\vec{\varepsilon}_{\pm 1})=\pm\hat{z}, which is parallel or anti-parallel to the spin quantization direction.

II.2 Measurement of spin density matrix through angular distribution of decay product

In high energy reactions, the spin polarization of a particle AA is mainly measured through the angular distribution of its decay product in two-body decay A1+2A\to 1+2 in the rest frame of AA. The corresponding formulae are derived using symmetry properties and conservation laws in the decay process.

In the rest frame of AA, the momenta of the particle 11 and 22 are denoted as p=p1=p2\vec{p}=\vec{p}_{1}=-\vec{p}_{2}, and their spin states are labeled by their helicities λ1\lambda_{1} and λ2\lambda_{2} respectively. The decay amplitude is given by,

Am(p;λ1λ2)=\displaystyle A_{m}(\vec{p};\lambda_{1}\lambda_{2})= p;λ1λ2|U^|jAmA,\displaystyle\langle\vec{p};\lambda_{1}\lambda_{2}|\hat{U}|j_{A}m_{A}\rangle, (23)

where U^\hat{U} stands for the transition operator, |jAmA|j_{A}m_{A}\rangle denotes the spin state of AA, and |p;λ1λ2|\vec{p};\lambda_{1}\lambda_{2}\rangle is the helicity state of decay daughters.

We can insert the completeness identity for the helicity states of the two-particle system |E;jm;λ1λ2|E;jm;\lambda_{1}\lambda_{2}\rangle with fixed EE (the energy of the system), jj (total angular momentum quantum number) and mm (the eigenvalue of jzj_{z}). From energy and angular momentum conservation laws in the decay process in AA’s rest frame, we have E=MAE=M_{A} (the mass of AA), j=jAj=j_{A} and m=mAm=m_{A}. So we obtain,

Am(p;λ1λ2)=\displaystyle A_{m}(\vec{p};\lambda_{1}\lambda_{2})= p;λ1λ2|MA;jAmA;λ1λ2\displaystyle\langle\vec{p};\lambda_{1}\lambda_{2}|M_{A};j_{A}m_{A};\lambda_{1}\lambda_{2}\rangle
×MA;jAmA;λ1λ2|U^|jAmA.\displaystyle\times\langle M_{A};j_{A}m_{A};\lambda_{1}\lambda_{2}|\hat{U}|j_{A}m_{A}\rangle. (24)

The space rotation invariance demands that the helicity amplitude

MA;jAmA;λ1λ2|U^|jAmAHA(λ1,λ2),\displaystyle\langle M_{A};j_{A}m_{A};\lambda_{1}\lambda_{2}|\hat{U}|j_{A}m_{A}\rangle\equiv H_{A}(\lambda_{1},\lambda_{2}), (25)

is independent of mAm_{A}. So we obtain,

Am(p;λ1λ2)\displaystyle A_{m}(\vec{p};\lambda_{1}\lambda_{2})
=p;λ1λ2|MA;jAmA;λ1λ2HA(λ1,λ2).\displaystyle=\langle\vec{p};\lambda_{1}\lambda_{2}|M_{A};j_{A}m_{A};\lambda_{1}\lambda_{2}\rangle H_{A}(\lambda_{1},\lambda_{2}). (26)

We see that the angular dependence is solely given by the inner product p;λ1λ2|MA;jAmA;λ1,λ2\langle\vec{p};\lambda_{1}\lambda_{2}|M_{A};j_{A}m_{A};\lambda_{1},\lambda_{2}\rangle, which we calculate as follows. First, we calculate it for p\vec{p} in zz-direction, i.e., p,0,0;λ1λ2|MA;jAmA;λ1λ2\langle p,0,0;\lambda_{1}\lambda_{2}|M_{A};j_{A}m_{A};\lambda_{1}\lambda_{2}\rangle, which gives a constant (2jA+1)/4π\sqrt{(2j_{A}+1)/4\pi} independent of mAm_{A}. We then transform |p,0,0;λ1λ2|p,0,0;\lambda_{1}\lambda_{2}\rangle to |p,θ,φ;λ1λ2|p,\theta,\varphi;\lambda_{1}\lambda_{2}\rangle in the direction of p\vec{p} by a spatial rotation. The rotation can be achieved by three successive rotations in Euler angles (α,β,γ)=(φ,θ,φ)(\alpha,\beta,\gamma)=(\varphi,\theta,-\varphi) that rotate p=(p,0,0)\vec{p}=(p,0,0) to p=(p,θ,φ)\vec{p}=(p,\theta,\varphi) with the rotation operator Jacob:1959at

R^(φ,θ,φ)=eiφJ^zeiθJ^yeiφJ^z.\hat{R}(\varphi,\theta,-\varphi)=e^{-i\varphi\hat{J}_{z}}e^{-i\theta\hat{J}_{y}}e^{i\varphi\hat{J}_{z}}. (27)

Hence, we obtain the inner product as

p;λ1λ2|MA;jAmA;λ1λ2\displaystyle\langle\vec{p};\lambda_{1}\lambda_{2}|M_{A};j_{A}m_{A};\lambda_{1}\lambda_{2}\rangle
=2jA+14πdmAλjA(θ)ei(mAλ)φ,\displaystyle=\sqrt{\frac{2j_{A}+1}{4\pi}}d_{m_{A}\lambda}^{j_{A}*}(\theta)e^{i(m_{A}-\lambda)\varphi}, (28)

where dmmj(θ)=jm|eiθJ^y|jmd_{mm^{\prime}}^{j}(\theta)=\langle jm|e^{-i\theta\hat{J}_{y}}|jm^{\prime}\rangle is the element of the Wigner rotation matrix and λ=λ1λ2\lambda=\lambda_{1}-\lambda_{2}. We note that the definition of the rotation operator R^\hat{R} in (27) introduces an additional rotation γ=φ\gamma=-\varphi relative to that in Ref. Chung:1971ri .

The spin density matrix of the system of particles 1 and 2 can be defined as

ρ^12=U^ρ^AU^,\displaystyle\hat{\rho}_{12}=\hat{U}\hat{\rho}^{A}\hat{U}^{\dagger}, (29)

where ρA\rho^{A} is the spin density matrix of A. Then the angular distribution is given by

W(θ,φ)=\displaystyle W(\theta,\varphi)= Nλ1,λ2p;λ1,λ2|ρ^12|p;λ1,λ2\displaystyle N\sum_{\lambda_{1},\lambda_{2}}\langle\vec{p};\lambda_{1},\lambda_{2}|\hat{\rho}_{12}|\vec{p};\lambda_{1},\lambda_{2}\rangle
=\displaystyle= Nλ1,λ2;mA,mA|HA(λ1,λ2)|2ρmAmAA\displaystyle N\sum_{\lambda_{1},\lambda_{2};m_{A},m^{\prime}_{A}}|H_{A}(\lambda_{1},\lambda_{2})|^{2}\rho^{A}_{m_{A}m^{\prime}_{A}}
×ei(mAmA)φdmAλjA(θ)dmAλjA(θ),\displaystyle\times e^{i(m_{A}-m^{\prime}_{A})\varphi}d_{m_{A}\lambda}^{j_{A}*}(\theta)d_{m^{\prime}_{A}\lambda}^{j_{A}}(\theta), (30)

where ρmAmAA=mA|ρ^A|mA\rho^{A}_{m_{A}m^{\prime}_{A}}=\langle m_{A}|\hat{\rho}^{A}|m^{\prime}_{A}\rangle denotes elements of ρ^A\hat{\rho}^{A} and NN is the normalization constant.

Since the helicity amplitude HA(λ1,λ2)H_{A}(\lambda_{1},\lambda_{2}) is generally unknown, only in some special cases one can use W(θ,φ)W(\theta,\varphi) to determine elements of ρ^A\hat{\rho}^{A}. Here are three such cases.

(1) jA=1/2j_{A}=1/2, λ1=±1/2\lambda_{1}=\pm 1/2, λ2=0\lambda_{2}=0, as in a spin-1/2 hyperon’s decay into a spin-1/2 baryon and a pion, HBπH\to B\pi, where HH denotes a hyperon. In this case, we have two independent helicity amplitudes HA(±1/2,0)H_{A}(\pm 1/2,0) and W(θ,φ)W(\theta,\varphi) is given by

W(θ,φ)=\displaystyle W(\theta,\varphi)= 14π(1+αAPAn),\displaystyle\frac{1}{4\pi}\left(1+\alpha_{A}\vec{P}_{A}\cdot\vec{n}\right), (31)

where PA=Tr(σρ^A)\vec{P}_{A}={\rm Tr}(\vec{\sigma}\hat{\rho}_{A}) is the polarization vector of AA, n=p/|p|\vec{n}=\vec{p}/|\vec{p}~{}| is the momentum direction and

αA=HA(1/2,0)HA(1/2,0)HA(1/2,0)+HA(1/2,0),\alpha_{A}=\frac{H_{A}(1/2,0)-H_{A}(-1/2,0)}{H_{A}(1/2,0)+H_{A}(-1/2,0)}, (32)

is the decay parameter for A1+2A\to 1+2. We also see that if parity is conserved in the decay process so that HA(1/2,0)=HA(1/2,0)H_{A}(1/2,0)=H_{A}(-1/2,0), we then have αA=0\alpha_{A}=0 and the isotropic distribution W(θ,φ)=1/4πW(\theta,\varphi)={1}/{4\pi}. So Eq. (31) can be used to determine PA\vec{P}_{A} only in weak decays.

(2) jA=1j_{A}=1, λ1=λ2=0\lambda_{1}=\lambda_{2}=0, such as in the vector meson’s decay into two pions, VππV\to\pi\pi. In this case, the helicity amplitude HA(0,0)H_{A}(0,0) is a trivial constant that can be absorbed into the normalization constant, so that

W(θ,φ)=\displaystyle W(\theta,\varphi)= 38π{(1ρ00A)+(3ρ00A1)cos2θ\displaystyle\frac{3}{8\pi}\Bigl{\{}(1-\rho^{A}_{00})+(3\rho^{A}_{00}-1)\cos^{2}\theta
2sin2θ[cos2φReρ11Asin2φImρ11A]\displaystyle-{2}\sin^{2}\theta\bigl{[}\cos 2\varphi{\rm Re}\rho^{A}_{1-1}-\sin 2\varphi{\rm Im}\rho^{A}_{1-1}\bigr{]}
2sin2θ[cosφ(Reρ10AReρ10A)\displaystyle-\sqrt{2}\sin 2\theta\bigl{[}\cos\varphi({\rm Re}\rho^{A}_{10}-{\rm Re}\rho^{A}_{-10})
sinφ(Imρ10A+Imρ10A)]},\displaystyle-\sin\varphi({\rm Im}\rho^{A}_{10}+{\rm Im}\rho^{A}_{-10})\bigr{]}\Bigr{\}}, (33)

If we integrate over φ\varphi, we obtain,

W(θ)=34\displaystyle W(\theta)=\frac{3}{4} {(1ρ00A)+(3ρ00A1)cos2θ}.\displaystyle\Bigl{\{}(1-\rho^{A}_{00})+(3\rho^{A}_{00}-1)\cos^{2}\theta\Bigr{\}}. (34)

We see that by measuring the distribution in θ\theta one can extract the value of ρ00A\rho^{A}_{00}.

(3) jA=1j_{A}=1, λ1=±1/2,λ2=±1/2\lambda_{1}=\pm 1/2,\lambda_{2}=\pm 1/2, such as in the vector meson’s decay into a dilepton pair, Ve+eV\to e^{+}e^{-}. Here, we have four combinations of λ1\lambda_{1} and λ2\lambda_{2}. In this case, only if parity and helicity conservation are valid so that HA(λ1,λ2)=HA(λ1,λ2)H_{A}(-\lambda_{1},-\lambda_{2})=H_{A}(\lambda_{1},\lambda_{2}) and HA(λ1,λ2)0H_{A}(\lambda_{1},\lambda_{2})\not=0 only for λ1=λ2\lambda_{1}=-\lambda_{2}, only one non-vanishing helicity amplitude is left and can be absorbed into the normalization constant. We then obtain the angular distribution as

W(θ,φ)=\displaystyle W(\theta,\varphi)= 38π(1+ρ00A){1+λθcos2θ\displaystyle\frac{3}{8\pi(1+\rho^{A}_{00})}\Bigl{\{}1+\lambda_{\theta}\cos^{2}\theta
+λφsin2θcos2φ+λθφsin2θsin2φ\displaystyle+\lambda_{\varphi}\sin^{2}\theta\cos 2\varphi+\lambda_{\theta\varphi}\sin 2\theta\sin 2\varphi
+λφsin2θsin2φ+λθφsin2θsinφ},\displaystyle+\lambda^{\perp}_{\varphi}\sin^{2}\theta\sin 2\varphi+\lambda^{\perp}_{\theta\varphi}\sin 2\theta\sin\varphi\Bigr{\}}, (35)

where the λ\lambda coefficients are related to elements of ρA\rho^{A} by

λθ=13ρ00A1+ρ00A,\displaystyle\lambda_{\theta}=\frac{1-3\rho^{A}_{00}}{1+\rho^{A}_{00}}, (36)
λφ=2Reρ11A1+ρ00A,\displaystyle\lambda_{\varphi}=\frac{2{\rm Re}\rho^{A}_{1-1}}{1+\rho^{A}_{00}}, (37)
λθφ=2Re(ρ10Aρ10A)1+ρ00A,\displaystyle\lambda_{\theta\varphi}=\frac{\sqrt{2}{\rm Re}(\rho^{A}_{10}-\rho^{A}_{-10})}{1+\rho^{A}_{00}}, (38)
λφ=2Imρ11A1+ρ00A,\displaystyle\lambda_{\varphi}^{\perp}=-\frac{2{\rm Im}\rho^{A}_{1-1}}{1+\rho^{A}_{00}}, (39)
λθφ=2Im(ρ10A+ρ10A)1+ρ00A.\displaystyle\lambda_{\theta\varphi}^{\perp}=-\frac{\sqrt{2}{\rm Im}(\rho^{A}_{10}+\rho^{A}_{-10})}{1+\rho^{A}_{00}}. (40)

In principle, one can extract elements of ρA\rho^{A} from W(θ,φ)W(\theta,\varphi) through these coefficients.

Measurements on hyperon polarization and vector meson’s spin alignment have been carried out in different high energy reactions using Eqs. (31-35) (see e.g. Refs. Lesnik:1975my ; Bunce:1976yb ; Bensinger:1983vc ; Gourlay:1986mf ; TASSO:1984nda ; ALEPH:1996oew ; OPAL:1997oem , DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs ; NOMAD:2006kuc and Chen:2007zzq ; Selyuzhenkov:2007ab ; STAR:2008lcm ; STAR:2007ccu ). In this brief review, we concentrate on spin alignment of vector mesons with light flavors in comparison with hyperon polarization, thus only Eqs. (31) and (33) are involved.

III Experimental results

III.1 Vector meson’s spin alignments in e+ee^{+}e^{-}-annihilations

The earliest measurements of vector meson’s spin alignments in high energy reactions might be made at the Large Electron-Position collider (LEP) at European Organization for Nuclear Research (CERN) in the 1990s DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs , where one of the popular polarization axis was the helicity axis defined by the individual particle’s momentum direction. Measurements have been carried out by DELPHI and OPAL Collaborations for K0K^{*0} and ϕ\phi mesons and even for heavy flavor meson such as BB and DD^{*} DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs . They also measured the off-diagonal element such as ρ1,1\rho_{1,-1}. As an example, we show the results from Ref. OPAL:1997vmw in Fig. 1.

Refer to caption
Figure 1: Experimental data obtained by OPAL Collaboration at LEP for the spin alignment ρ00\rho_{00} and off-diagonal element ρ11\rho_{1-1} of K0K^{*0}. This figure is taken from Ref. OPAL:1997vmw .

From these measurements DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs in e+ee^{+}e^{-} collisions, we see clearly that ρ00\rho_{00} is significantly larger than 1/3 in the fragmentation region xp>0.3x_{p}>0.3 (xpp/pbeamx_{p}\equiv p/p_{\rm beam}) which show that the spin of vector mesons is significantly aligned in that region. However, at small fractional momenta xp0.3x_{p}\leq 0.3, null results for K0K^{*0} and ϕ\phi mesons were reported in e+ee^{+}e^{-} collisions DELPHI:1997ruo ; OPAL:1997vmw .

The data have attaracted much theoretical attention Anselmino:1997ui ; Anselmino:1998jv ; Anselmino:1999cg ; Xu:2001hz ; Xu:2003fq ; Chen:2016moq ; Chen:2016iey ; Chen:2020pty and we will come back to this later in Sec. VI.

III.2 Global spin alignments of vector mesons in heavy-ion collisions

New measurements of the vector meson’s spin alignment in heavy-ion collisions are different from conventional studies, where a new polarization axis along the nucleus-nucleus system’s orbital angular momentum is defined Liang:2004ph ; Liang:2004xn , and this is the so-called global spin alignment. In experiment, the quantization axis is determined by the normal of the reaction plane, which can be reconstructed by using the charge particle momentum distribution collected in the detector STAR:2022fan . Particles of interest, for example, the vector-meson ϕ\phi and K0K^{*0} are observed by paring of their decay daughters (K±K^{\pm}, π±\pi^{\pm}) with subtraction of the combinatorial background. Then the polar angle distribution of Eq. (34) is analyzed, and the ρ00\rho_{00} is extracted after correction for detection efficiency and acceptance STAR:2022fan .

Refer to caption
Figure 2: Measurements of the vector meson’s spin density matrix element ρ00\rho_{00} as functions of transverse momentum (pTp_{\rm T}) for K0K^{*0} and ϕ\phi mesons in Au+Au collisions at different energies. Data points are STAR’s measurements STAR:2022fan , bands are theoretical calculations for ϕ\phi mesons Sheng:2022wsy .

In 2008 the STAR Collaboration performed a measurement of the vector meson’s global spin alignment in Au+Au collisions at sNN\sqrt{s_{\rm NN}} = 200 GeV. Due to limited statistics and only covered top RHIC collision energies, no significant results were reported STAR:2008lcm . Since 2010, the collaboration has been collecting and analyzing data of higher statistics and at lower collision energies, including data in the Beam Energy Scan Phase I (BES-I) runs and high statistics Au+Au runs at sNN\sqrt{s_{\rm NN}} = 200 GeV. The analysis was focused on mid-central collisions (20-60%) where larger system angular momenta are expected in comparison with the values in central or peripheral collisions. Fig. 2 presents the transverse momentum dependence of ρ00\rho_{00} for K0K^{*0} and ϕ\phi in 20-60% central Au+Au collisions at sNN\sqrt{s_{\rm NN}} = 11.5, 19.6, 27, 39, 62.4 (54.4), 200 GeV. The ρ00\rho_{00} results show a non-trivial pTp_{\rm T} dependence. For K0K^{*0} mesons at sNN\sqrt{s_{\rm NN}} = 54.4 and 200 GeV, they are larger than 1/3 with about 2σ2\sigma significance at intermediate pTp_{\rm T}. At low beam energies the statistics is not sufficient to observe any significant deviation from 1/3. From the calculation in Ref. Liang:2004xn , one naively expect ρ00\rho_{00} to be smaller than 1/3 due to hadronization of the polarized quark and antiquark via quark combination, and larger than 1/3 due to fragmentation of the quark and antiquark. On the other hand, for ϕ\phi mesons for all energies considered, we see that the departure of ρ00\rho_{00} from 1/3 mainly occurs at pTp_{\rm T}\sim 1.0 - 2.4 GeV/c, while at higher pTp_{\rm T} the result can be regarded as being consistent with 1/3 within 2σ\sim 2\sigma or less significance.

The STAR collaboration also studied the collision energy dependence by integrating ρ00(pT)\rho_{00}(p_{\rm T}) with the weight 1/(stat. error)2. Fig. 3 presents ρ00\rho_{00} for K0K^{*0} and ϕ\phi mesons in 20-60% central Au+Au collisions at collision energies ranging from sNN\sqrt{s_{\rm NN}} = 11.5 to 200 GeV STAR:2022fan . We see that ρ00\rho_{00} for ϕ\phi mesons increases with decreasing the collision energy, while ρ00\rho_{00} for K0K^{*0} mesons fluctuates around 1/3 with the collision energy. To quantify the effect, the average of ρ00\rho_{00} is taken over lower collision energies for K0K^{*0} and ϕ\phi. The ρ00\rho_{00} for ϕ\phi mesons, averaged over beam energies between 11.5 and 62.4 GeV, is 0.3512 ±\pm 0.0017 (stat.) ±\pm 0.0017 (syst.). Taking the total uncertainties as the sum in quadrature of statistical and systematic uncertainties, the result indicates that ρ00\rho_{00} for ϕ\phi mesons is above 1/3 with a significance of 7.4σ7.4\sigma, representing the first observation of the global spin alignment. The ρ00\rho_{00} for K0K^{*0} mesons, averaged over beam energies between 11.5 and 54.4 GeV, is 0.3356 ±\pm 0.0034 (stat.) ±\pm 0.0043 (syst.) and is consistent with 1/3. The measurements of the ALICE collaboration in Pb+Pb collisions at 2.76 TeV ALICE:2019aid , taken from the closest data points to the mean pTp_{\rm T} for the pTp_{\rm T} range used in STAR’s measurements, are also shown for comparison in Fig. 3. The ρ00\rho_{00} data point for K0K^{*0} and ϕ\phi mesons from ALICE collaboration is more or less consistent with 1/3 with large uncertainties.

Refer to caption
Figure 3: Measurements of ρ00\rho_{00} with respect to the system orbital angular momentum in high energy heavy-ion collisions. Stars represent the data for ϕ\phi mesons STAR:2022fan ; ALICE:2019aid , circles represent the data for K0K^{*0} mesons STAR:2022fan ; ALICE:2019aid . The solid red line is the prediction from a non-relativistic coalescence model with strong force fields Sheng:2019kmk . More sophisticated and complete calculations using the relativistic coalescence model and spin kinetic equation Sheng:2022ffb were presented in Ref. Sheng:2022wsy ; Sheng:2023urn .

According to various studies, there are many sources that contribute to the global spin alignment of ϕ\phi mesons including vortical flows Yang:2017sdk , electromagnetic fields Yang:2017sdk generated by the electric currents carried by the colliding nuclei, local spin alignment Xia:2020tyd , quark polarization along the direction of its momentum (helicity polarization) Gao:2021rom , the spin alignment by fragmentation Liang:2004xn of polarized quarks, and the shear stress tensor Li:2022vmb ; Wagner:2022gza ; Dong:2023cng . However, these conventional mechanisms are not sufficient to account for the observed ρ00\rho_{00} for ϕ\phi mesons. It was also proposed that local correlations or fluctuations in turbulent color fields Muller:2021hpe and glasma fields Kumar:2023ghs can also generate a significant contribution to ρ00\rho_{00}. A recent theoretical development based on local correlations or fluctuations of ϕ\phi vector fields can describe the experimental data. The ϕ\phi vector field is the ’33’ component of the SU(3) vector multiplet induced by currents of pseudo-Goldstone bosons Manohar:1983md , and it can polarize ss and s¯\bar{s} quarks in the same way as the electromagnetic field does. The solid curve of Fig. 3 is such a fit to the data. We see that the calculation describes the data reasonably well. The correlation of the ϕ\phi vector field can be quantified by Eϕ2\langle{\vec{E}}_{\phi}^{2}\rangle and Bϕ2\langle{\vec{B}}_{\phi}^{2}\rangle and can be extracted by fitting data, where Eϕ{\vec{E}}_{\phi} and Bϕ{\vec{B}}_{\phi} are electric and magnetic components of the ϕ\phi vector field respectively. We will discuss more details about the theoretical model in Sec. IV.2.

IV Vector meson’s spin alignments in quark combination models

It was well-known that huge angular momenta are generated in non-central high energy heavy-ion collisions Liang:2004ph ; Gao:2007bc . Due to spin-orbit couplings in QCD interaction, such huge angular momenta lead to the global polarization in quark scatterings in the form of the hyperon polarization Liang:2004ph ; Gao:2007bc and vector meson’s spin alignment Liang:2004xn , the so-called GPE.

The GPE depends not only on the global quark polarization but also on the hadronization mechanism. In relativistic heavy-ion collisions, it is envisaged that a system of decomfined quarks and anti-quarks is created in the central rapidity and moderate transverse momentum region. Various aspects of experimental data suggest that hadronization of the system proceeds via combination of quarks and anti-quarks, a mechanism phrased as “quark recombination”, “quark coalescence” or simply as “quark combination”. We simply refer to it as “quark combination model” or “quark coalescence model” in this paper. We will summarize the results of vector meson’s spin alignments by the quark combination model in this section.

IV.1 Vector meson’s spin alignments in non-relativistic quark combination models

In a non-relativistic quark combination model, the physics in the process that qq (quark) and q¯\bar{q} (antiquark) combine to MM (meson) can be demonstrated in a clear way. Here, it is envisaged that in the combination process the vector meson’s spin is just the sum over the quark’s and anti-quark’s spins. Hence the spin density matrix and the spin alignment of the vector meson can be calculated from that of the quark and anti-quark. Such a calculation is straightforward and was carried out in Ref. Liang:2004xn , which we will summarize in this subsection.

The global quark polarization was taken as a constant so that the spin density matrix for the quark and antiquark takes the diagonal form Liang:2004xn

ρ^q=12(1+Pq001Pq).\displaystyle\hat{\rho}^{q}=\frac{1}{2}\left(\begin{array}[]{cc}1+P_{q}&0\\ 0&1-P_{q}\end{array}\right). (43)

The spin density matrix of the q1q¯2q_{1}\bar{q}_{2} system was taken as a direct product of the quark’s and anti-quark’s,

ρ^q1q¯2=ρ^q1ρ^q¯2.\displaystyle\hat{\rho}^{q_{1}\bar{q}_{2}}=\hat{\rho}^{q_{1}}\otimes\hat{\rho}^{\bar{q}_{2}}. (44)

The elements of the spin density matrix ρ^V\hat{\rho}^{V} for the vector meson is obtained from ρ^q1q¯2\hat{\rho}^{q_{1}\bar{q}_{2}} as

ρmmV=jm|ρ^q1q¯2|jm,\displaystyle\rho^{V}_{m^{\prime}m}=\langle jm^{\prime}|\hat{\rho}^{q_{1}\bar{q}_{2}}|jm\rangle, (45)

which leads to

ρmmV=\displaystyle\rho^{V}_{m^{\prime}m}= m1,m2,m1,m2ρm1m2,m1m2q1q¯2\displaystyle\sum_{m_{1},m_{2},m_{1}^{\prime},m_{2}^{\prime}}\rho^{q_{1}\bar{q}_{2}}_{m_{1}^{\prime}m_{2}^{\prime},m_{1}m_{2}}
×jVm|m1m2m1m2|jVm,\displaystyle\times\langle j_{V}m^{\prime}|m^{\prime}_{1}m^{\prime}_{2}\rangle\langle m_{1}m_{2}|j_{V}m\rangle, (46)

where |jVm|j_{V}m\rangle is the vector meson’s spin state in the constituent quark model with jV=1j_{V}=1 and m=0,±1m=0,\pm 1, and jVm|m1m2\langle j_{V}m|m_{1}m_{2}\rangle are Clebsch-Gordan coefficients. After a straightforward calculation, we obtain the normalized spin alignment ρ00V\rho^{V}_{00} as Liang:2004xn ,

ρ00V=1Pq1Pq¯23+Pq1Pq¯2.\rho^{V}_{00}=\frac{1-P_{q_{1}}P_{\bar{q}_{2}}}{3+P_{q_{1}}P_{\bar{q}_{2}}}. (47)

If we take Pq1=Pq¯2=PqP_{q_{1}}=P_{\bar{q}_{2}}=P_{q} (flavor blind for quarks and antiquarks), we simply obtain

ρ00V=1Pq23+Pq2.\rho^{V}_{00}=\frac{1-P_{q}^{2}}{3+P_{q}^{2}}. (48)

In exact the same way, we obtain the global hyperon polarization PH=PqP_{H}=P_{q} for Λ\Lambda, Σ\Sigma, and Ξ\Xi Liang:2004ph .

From Eqs. (47) and (48), we see that, in contrast to PHP_{H}, ρ00V\rho^{V}_{00} is quadratic in PqP_{q} and should be less than 1/31/3. We emphasize that the case considered in Ref. Liang:2004xn is over simplified in the fact that only the spin degree of freedom for the quark and antiquark is considered, neglecting other degree(s) of freedom and the correlation among the quark’s and antiquark’s polarization. So it might not be a surprise that the STAR data STAR:2017ckg ; STAR:2022fan for the global spin alignment of ϕ\phi mesons show a large deviation from 1/3, far beyond the value estimated from PΛP_{\Lambda}’s data by its square. In fact, if we make a step forward by considering the dependence of PqP_{q} on other degree of freedom Sheng:2019kmk ; Sheng:2020ghv ; Sheng:2022ffb ; Sheng:2022wsy , Pq1Pq¯2P_{q_{1}}P_{\bar{q}_{2}} in Eq. (47) should be replaced by Pq1Pq¯2\langle P_{q_{1}}P_{\bar{q}_{2}}\rangle, so we have

ρ00V=1Pq1Pq¯23+Pq1Pq¯2.\rho^{V}_{00}=\frac{1-\langle P_{q_{1}}P_{\bar{q}_{2}}\rangle}{3+\langle P_{q_{1}}P_{\bar{q}_{2}}\rangle}. (49)

The STAR data STAR:2017ckg ; STAR:2022fan indicate

Pq1Pq¯2Pq1Pq¯2.\langle P_{q_{1}}P_{\bar{q}_{2}}\rangle\not=\langle P_{q_{1}}\rangle\langle P_{\bar{q}_{2}}\rangle. (50)

This means that there should be a strong correlation between the polarization of the quark and antiquark. Hence, the study of the global spin alignment for the vector meson in heavy-ion collisions provides a unique opportunity for exploring the local correlation in the quark’s and antiquark’s polarization, a new window for the study of properties of QGP.

It is also emphasized in Ref. Liang:2023talk that the average in Eq. (49) is two-folded: the first average is taken inside the vector meson’s wave function (local correlation) and the second average is taken in the range outside the vector meson’s wave function (long-range correlation). More measurements are needed to tell the difference between two types of correlation Liang:2023talk . A systematic formulation has been presented in Lv:2024uev . It has bee extended to spin 3/2 baryons in Zhang:2024hyq .

IV.2 Vector meson’s spin alignments in quark combination models with phase space dependence

IV.2.1 Non-relativistic model

In this subsection, we consider the non-relativistic quark combination model with phase space dependence Sheng:2020ghv ; Xin-Li:2023gwh . We extend the density operator (1) to the phase space as

ρ^q=\displaystyle\hat{\rho}^{q}= rs[d3p][d3q]d3xeiqx\displaystyle\sum_{rs}\int[d^{3}{\vec{p}}][d^{3}{\vec{q}}]\int d^{3}{\vec{x}}e^{-i{\vec{q}}\cdot{\vec{x}}}
×wrsq(x,p)|r,p+q2s,pq2|,\displaystyle\times w_{rs}^{q}({\vec{x}},{\vec{p}})\left|r,{\vec{p}}+\frac{{\vec{q}}}{2}\right\rangle\left\langle s,{\vec{p}}-\frac{{\vec{q}}}{2}\right|\,, (51)

where r,s=±r,s=\pm denote spin states, |r,p\left|r,{\vec{p}}\right\rangle denotes the spin-momentum state, and the momentum measure is defined as [d3p]d3p/(2π)3[d^{3}{\vec{p}}]\equiv d^{3}{\vec{p}}/(2\pi)^{3}. The weight function wrsq(x,p)w_{rs}^{q}({\vec{x}},{\vec{p}}) is the Wigner function in the spin-phase space and can be obtained as

wrsq(x,p)=[d3q]eiqxr,p+q2|ρ^q|s,pq2.w_{rs}^{q}({\vec{x}},{\vec{p}})=\int[d^{3}{\vec{q}}]e^{i{\vec{q}}\cdot{\vec{x}}}\left\langle r,{\vec{p}}+\frac{{\vec{q}}}{2}\right|\hat{\rho}^{q}\left|s,{\vec{p}}-\frac{{\vec{q}}}{2}\right\rangle\,. (52)

It is a 2×22\times 2 Hermitian matrix in the spin space and can be parameterized as

wrsq(x,p)=12fq(x,p)[δrs+σrsPq(x,p)].w_{rs}^{q}({\vec{x}},{\vec{p}})=\frac{1}{2}f_{q}({\vec{x}},{\vec{p}})\left[\delta_{rs}+{\vec{\sigma}}_{rs}\cdot{\vec{P}}_{q}({\vec{x}},{\vec{p}})\right]\,. (53)

Here fq(x,p)f_{q}({\vec{x}},{\vec{p}}) is the un-polarized distribution function, and Pq(x,p){\vec{P}}_{q}({\vec{x}},{\vec{p}}) is the spin polarization in phase space. The density operator for the anti-quark can be defined in the same way. The density operator for a quark-antiquark pair q1q¯2q_{1}\overline{q}_{2} is defined from the direct product of quark’s and anti-quark’s one as ρ^q1q¯2ρ^q1ρ^q¯2\hat{\rho}^{q_{1}\bar{q}_{2}}\equiv\hat{\rho}^{q_{1}}\otimes\hat{\rho}^{\bar{q}_{2}}. The Wigner function can be normalized as

sd3x[d3p]wss(x,p)=1.\sum_{s}\int d^{3}{\vec{x}}\int[d^{3}{\vec{p}}]w_{ss}({\vec{x}},{\vec{p}})=1. (54)

The quark’s spin quantization direction can be any direction, e.g., the zz-direction, without loss of generality.

The elements of the vector meson’s density matrix are obtained as

ρmmV(x,p)\displaystyle\rho_{m^{\prime}m}^{V}({\vec{x}},{\vec{p}}) (55)
=\displaystyle= 1N(x,p)[d3q]eiqx\displaystyle\frac{1}{N({\vec{x}},{\vec{p}})}\int[d^{3}{\vec{q}}]e^{i{\vec{q}}\cdot{\vec{x}}}
×jVm;p+q2|ρ^q1ρ^q¯2|jVm;pq2\displaystyle\times\left\langle j_{V}m^{\prime};{\vec{p}}+\frac{{\vec{q}}}{2}\right|\hat{\rho}^{q_{1}}\otimes\hat{\rho}^{\bar{q}_{2}}\left|j_{V}m;{\vec{p}}-\frac{{\vec{q}}}{2}\right\rangle
=\displaystyle= 1N(x,p)d3xb[d3pb][d3qb]exp(iqbxb)\displaystyle\frac{1}{N({\vec{x}},{\vec{p}})}\int d^{3}{\vec{x}}_{b}[d^{3}{\vec{p}}_{b}][d^{3}{\vec{q}}_{b}]\exp(-i{\vec{q}}_{b}\cdot{\vec{x}}_{b})
×φV(pb+qb2)φV(pbqb2)\displaystyle\times\varphi_{V}^{\ast}\left({\vec{p}}_{b}+\frac{{\vec{q}}_{b}}{2}\right)\varphi_{V}\left({\vec{p}}_{b}-\frac{{\vec{q}}_{b}}{2}\right)
×r1,r2,s1,s2wm1m1q1(x1,p1)wm2m2q¯2(x2,p2)\displaystyle\times\sum_{r_{1},r_{2},s_{1},s_{2}}w_{m_{1}^{\prime}m_{1}}^{q_{1}}({\vec{x}}_{1},{\vec{p}}_{1})w_{m_{2}^{\prime}m_{2}}^{\bar{q}_{2}}({\vec{x}}_{2},{\vec{p}}_{2})
×jVmm1m2m1m2jVm,\displaystyle\times\left\langle j_{V}m^{\prime}\mid m_{1}^{\prime}m_{2}^{\prime}\right\rangle\left\langle m_{1}m_{2}\mid j_{V}m\right\rangle\,,

where x1,2x±xb/2{\vec{x}}_{1,2}\equiv{\vec{x}}\pm{\vec{x}}_{b}/2 and p1,2p/2±pb{\vec{p}}_{1,2}\equiv{\vec{p}}/2\pm{\vec{p}}_{b} are positions and momenta for the quark q1q_{1} and the antiquark q¯2\overline{q}_{2} respectively, and N(x,p)N({\vec{x}},{\vec{p}}) is the normalization factor that ensures m=0,±1ρmmV(x,p)=1\sum_{m=0,\pm 1}\rho_{mm}^{V}({\vec{x}},{\vec{p}})=1. Since we work in the non-relativistic limit, the spin and momentum can be decoupled. Therefore Clebsch-Gordan coefficients m1m2jVm\left\langle m_{1}m_{2}\mid j_{V}m\right\rangle and jVmm1m2\left\langle j_{V}m^{\prime}\mid m_{1}^{\prime}m_{2}^{\prime}\right\rangle as well as the meson’s wave function φV\varphi_{V} appear in Eq. (55). By further neglecting the phase-space dependence of unpolarized distribution functions and choosing φV\varphi_{V} as a Gaussian distribution with the width aVa_{V} in the momentum space, we derive the spin alignment and other parameters in the angular distribution (33),

ρ00V(x,p)1323Pq1zPq¯2zV+29Pq1Pq¯2V,\displaystyle\rho_{00}^{V}({\vec{x}},{\vec{p}})\approx\frac{1}{3}-\frac{2}{3}\left\langle P_{q_{1}}^{z}P_{\bar{q}_{2}}^{z}\right\rangle_{V}+\frac{2}{9}\left\langle{\vec{P}}_{q_{1}}\cdot{\vec{P}}_{\bar{q}_{2}}\right\rangle_{V}\,,
Reρ1,1V(x,p)13Pq1xPq¯2xPq1yPq¯2yV,\displaystyle\text{Re}\rho_{1,-1}^{V}({\vec{x}},{\vec{p}})\approx\frac{1}{3}\left\langle P_{q_{1}}^{x}P_{\bar{q}_{2}}^{x}-P_{q_{1}}^{y}P_{\bar{q}_{2}}^{y}\right\rangle_{V},
Imρ1,1V(x,p)13Pq1xPq¯2y+Pq1yPq¯2xV,\displaystyle-\text{Im}\rho_{1,-1}^{V}({\vec{x}},{\vec{p}})\approx\frac{1}{3}\left\langle P_{q_{1}}^{x}P_{\bar{q}_{2}}^{y}+P_{q_{1}}^{y}P_{\bar{q}_{2}}^{x}\right\rangle_{V},
Re[ρ1,0ρ1,0](x,p)23Pq1zPq¯2x+Pq1xPq¯2zV,\displaystyle\text{Re}\left[\rho_{1,0}-\rho_{-1,0}\right]({\vec{x}},{\vec{p}})\approx\frac{\sqrt{2}}{3}\left\langle P_{q_{1}}^{z}P_{\bar{q}_{2}}^{x}+P_{q_{1}}^{x}P_{\bar{q}_{2}}^{z}\right\rangle_{V},
Im[ρ1,0+ρ1,0](x,p)23Pq1yPq¯2z+Pq1zPq¯2yV,\displaystyle-\text{Im}\left[\rho_{1,0}+\rho_{-1,0}\right]({\vec{x}},{\vec{p}})\approx\frac{\sqrt{2}}{3}\left\langle P_{q_{1}}^{y}P_{\bar{q}_{2}}^{z}+P_{q_{1}}^{z}P_{\bar{q}_{2}}^{y}\right\rangle_{V},

where the correlation between quark and antiquark’s polarizations is defined as

Pq1iPq¯2jV\displaystyle\left\langle P_{q_{1}}^{i}P_{\bar{q}_{2}}^{j}\right\rangle_{V}\equiv 1π3d3xbd3pbexp(pb2aV2aV2xb2)\displaystyle\frac{1}{\pi^{3}}\int d^{3}{\vec{x}}_{b}d^{3}{\vec{p}}_{b}\exp\left(-\frac{{\vec{p}}_{b}^{2}}{a_{V}^{2}}-a_{V}^{2}{\vec{x}}_{b}^{2}\right)
×Pq1i(x1,p1)Pq¯2j(x2,p2).\displaystyle\times P_{q_{1}}^{i}({\vec{x}}_{1},{\vec{p}}_{1})P_{\bar{q}_{2}}^{j}({\vec{x}}_{2},{\vec{p}}_{2})\,. (57)

We note that in Eq. (LABEL:eq:tensor-polar) the spin quantization direction is set to the zz-direction.

IV.2.2 Relativistic model in quantum kinetic approach

For a relativistic quark or antiquark, its polarization four-vector is always perpendicular to its four-momentum, implying that the spin with momentum cannot be decoupled as in the non-relativistic case. Consequently, the vector meson’s spin cannot be obtained by the constituent quark’s and anti-quark’s spins through the angular momentum coupling in non-relativistic quantum mechanics. Based on the Kadanoff-Baym equation in the closed-time-path formalism, a relativistic spin kinetic theory for vector mesons has been constructed Sheng:2022ffb to explain experimental data on the spin alignment STAR:2017ckg .

In order to describe spin transport phenomena, we use the matrix-valued spin-dependent distribution (MVSD) fm1m2V(x,p)f_{m_{1}m_{2}}^{V}(x,{\vec{p}}) for the vector meson. The spin Boltzmann equation with coalescence and dissociation collision terms reads,

pxfm1m2V(x,p)\displaystyle p\cdot\partial_{x}f_{m_{1}m_{2}}^{V}(x,{\vec{p}}) (58)
=\displaystyle= 116m1,m2[ϵμ(m1,p)ϵν(m1,p)δm2m2\displaystyle\frac{1}{16}\sum_{m_{1}^{\prime},m_{2}^{\prime}}\left[\epsilon_{\mu}^{\ast}(m_{1},{\vec{p}})\epsilon_{\nu}(m_{1}^{\prime},{\vec{p}})\delta_{m_{2}m_{2}^{\prime}}\right.
+δm1m1ϵμ(m2,p)ϵν(m2,p)]\displaystyle\left.+\delta_{m_{1}m_{1}^{\prime}}\epsilon_{\mu}^{\ast}(m_{2}^{\prime},{\vec{p}})\epsilon_{\nu}(m_{2},{\vec{p}})\right]
×𝒞m1m2μν(x,p),\displaystyle\times\mathcal{C}_{m_{1}^{\prime}m_{2}^{\prime}}^{\mu\nu}(x,{\vec{p}}),

where ϵμ(m,p)\epsilon^{\mu}(m,{\vec{p}}) denotes the meson’s normalized polarization vector perpendicular to pμp^{\mu}, and m1m_{1}, m2m_{2}, m1m_{1}^{\prime}, and m2m_{2}^{\prime} label the vector meson’s spin states along the spin quantization direction with three values 0 and ±1\pm 1. Here we have assumed that the vector meson is on its mass-shell p0=mV2+p2p^{0}=\sqrt{m_{V}^{2}+{\vec{p}}^{2}}. The gain and loss terms in 𝒞m1m2μν\mathcal{C}_{m_{1}^{\prime}m_{2}^{\prime}}^{\mu\nu} correspond to the coalescence and dissociation processes for the quark and antiquark, respectively. In the dilute limit when distribution functions of the meson, the constituent quark and antiquark are much less than unity, Eq. (58) is simplified as

pxfm1m2V(x,p)=\displaystyle p\cdot\partial_{x}f_{m_{1}m_{2}}^{V}(x,{\vec{p}})= 18[ϵμ(m1,p)ϵν(m2,p)𝒞coalμν(x,p)\displaystyle\frac{1}{8}\left[\epsilon_{\mu}^{\ast}(m_{1},{\vec{p}})\epsilon_{\nu}(m_{2},{\vec{p}})\mathcal{C}_{\text{coal}}^{\mu\nu}(x,{\vec{p}})\right.
𝒞diss(p)fm1m2V(x,p)],\displaystyle\left.-\mathcal{C}_{\text{diss}}({\vec{p}})f_{m_{1}m_{2}}^{V}(x,{\vec{p}})\right], (59)

where the contribution from the dissociation is proportional to the MVSD with the coefficient 𝒞diss\mathcal{C}_{\text{diss}} being independent of the MVSD. The coalescence kernel 𝒞coalμν\mathcal{C}_{\text{coal}}^{\mu\nu} can be obtained from the Kadanoff-Baym equation,

𝒞coalμν(x,p)\displaystyle\mathcal{C}_{\text{coal}}^{\mu\nu}(x,{\vec{p}}) =\displaystyle= d3p(2π)2δ(EpVEpq¯Eppq)Epq¯Eppq\displaystyle\int\frac{d^{3}{\vec{p}}^{\prime}}{(2\pi)^{2}}\frac{\delta(E_{{\vec{p}}}^{V}-E_{{\vec{p}}^{\prime}}^{\bar{q}}-E_{{\vec{p}}-{\vec{p}}^{\prime}}^{q})}{E_{{\vec{p}}^{\prime}}^{\bar{q}}E_{{\vec{p}}-{\vec{p}}^{\prime}}^{q}} (60)
×Tr{Γν(pγMq¯)\displaystyle\times\text{Tr}\left\{\Gamma^{\nu}(p^{\prime}\cdot\gamma-M_{\bar{q}})\right.
×[1+γ5γPq¯(x,p)]\displaystyle\times[1+\gamma_{5}\gamma\cdot P_{\bar{q}}(x,{\vec{p}}^{\prime})]
×Γμ[(pp)γ+Mq]\displaystyle\times\Gamma^{\mu}[(p-p^{\prime})\cdot\gamma+M_{q}]
×[1+γ5γPq(x,pp)]}\displaystyle\times\left.[1+\gamma_{5}\gamma\cdot P_{q}(x,{\vec{p}}-{\vec{p}}^{\prime})]\right\}
×fq¯(x,p)fq(x,pp),\displaystyle\times f_{\bar{q}}(x,{\vec{p}}^{\prime})f_{q}(x,{\vec{p}}-{\vec{p}}^{\prime})\,,

where MqM_{q} and Mq¯M_{\bar{q}} are quark and antiquark masses respectively, EpVE_{{\vec{p}}}^{V}, EppqE_{{\vec{p}}-{\vec{p}}^{\prime}}^{q}, and Epq¯E_{{\vec{p}}^{\prime}}^{\bar{q}} are energies for the vector meson, the constituent quark and antiquark respectively, and Γμ\Gamma^{\mu} is the effective vertex of quark-antiquark-meson. Here fq/q¯f_{q/\bar{q}} and Pq/q¯μP_{q/\bar{q}}^{\mu} are the unpolarized and spin polarization distributions for the quark/antiquark respectively. By neglecting the space dependence (the distributions are homogeneous in space) and setting fm1m2V=0f_{m_{1}m_{2}}^{V}=0 before the hadronization stage in heavy-ion collisions, one can obtain a formal solution to Eq. (59). Spin density matrix elements can be derived as Sheng:2022wsy ,

ρm1m2V(x,p)=ϵμ(m1,p)ϵν(m2,p)𝒞coalμν(x,p)m=0,±1ϵα(m,p)ϵβ(m,p)𝒞coalαβ(x,p).\rho_{m_{1}m_{2}}^{V}(x,{\vec{p}})=\frac{\epsilon_{\mu}^{\ast}(m_{1},{\vec{p}})\epsilon_{\nu}(m_{2},{\vec{p}})\mathcal{C}_{\text{coal}}^{\mu\nu}(x,{\vec{p}})}{\sum_{m=0,\pm 1}\epsilon_{\alpha}^{\ast}(m,{\vec{p}})\epsilon_{\beta}(m,{\vec{p}})\mathcal{C}_{\text{coal}}^{\alpha\beta}(x,{\vec{p}})}. (61)

This is a relativistic formula for the vector meson’s spin density matrix built from the coalescence kernel (60) encoding the polarization distributions of the constituent quark and antiquark. Clearly the vector meson’s spin density matrix has non-trivial dependence on the spin polarization and momenta of the quark and antiquark.

We consider that ss and s¯\bar{s} are polarized in a thermal medium by an effective vector field called the ϕ\phi field Sheng:2019kmk ; Sheng:2022wsy ; Sheng:2022ffb induced by currents of pseudo-Goldstone bosons Manohar:1983md . The spin polarization of ss and s¯\bar{s} is then given by

Pq/q¯μ(x,p)±gϕ4Ms(up)ThϵμναβpνFαβϕ(x),P_{q/\bar{q}}^{\mu}(x,{\vec{p}})\approx\pm\frac{g_{\phi}}{4M_{s}(u\cdot p)T_{\text{h}}}\epsilon^{\mu\nu\alpha\beta}p_{\nu}F_{\alpha\beta}^{\phi}(x), (62)

where uμu^{\mu} denotes the velocity of the thermal background at the effective hadronization temperature ThT_{\text{h}}, gϕg_{\phi} is the effective coupling constant of the ss¯ϕs\bar{s}\phi vertex, and Fαβϕ(x)F_{\alpha\beta}^{\phi}(x) is the ϕ\phi field’s strength tensor. Substituting (62) into Eq. (60) and assuming uμ=(1,0,0,0)u^{\mu}=(1,0,0,0) in the meson’s rest frame, we obtain the spin alignment of the ϕ\phi meson

ρ00ϕ(x,p)\displaystyle\rho_{00}^{\phi}(x,{\vec{p}})
\displaystyle\approx 134gϕ2Mϕ2ThC1[13BϕBϕ(ϵ0Bϕ)2]\displaystyle\frac{1}{3}-\frac{4g_{\phi}^{2}}{M_{\phi}^{2}T_{\text{h}}}C_{1}\left[\frac{1}{3}{\vec{B}}_{\phi}^{\prime}\cdot{\vec{B}}_{\phi}^{\prime}-({\vec{\epsilon}}_{0}\cdot{\vec{B}}_{\phi}^{\prime})^{2}\right]
4gϕ2Mϕ2ThC2[13EϕEϕ(ϵ0Eϕ)2],\displaystyle-\frac{4g_{\phi}^{2}}{M_{\phi}^{2}T_{\text{h}}}C_{2}\left[\frac{1}{3}{\vec{E}}_{\phi}^{\prime}\cdot{\vec{E}}_{\phi}^{\prime}-({\vec{\epsilon}}_{0}\cdot{\vec{E}}_{\phi}^{\prime})^{2}\right], (63)

where MϕM_{\phi} is the mass of the ϕ\phi meson, and Bϕ{\vec{B}}_{\phi}^{\prime} and Eϕ{\vec{E}}_{\phi}^{\prime} are electric and magnetic components of the ϕ\phi field in the meson’s rest frame respectively. Here ϵ0{\vec{\epsilon}}_{0} denotes the spin quantization direction. The coefficients C1C_{1} and C2C_{2} depend only on quark’s and meson’s masses,

C1\displaystyle C_{1} =\displaystyle= 8Ms4+16Ms2Mϕ2+3Mϕ4120Ms2(Mϕ2+2Ms2),\displaystyle\frac{8M_{s}^{4}+16M_{s}^{2}M_{\phi}^{2}+3M_{\phi}^{4}}{120M_{s}^{2}(M_{\phi}^{2}+2M_{s}^{2})},
C2\displaystyle C_{2} =\displaystyle= 8Ms414Ms2Mϕ2+3Mϕ4120Ms2(Mϕ2+2Ms2).\displaystyle\frac{8M_{s}^{4}-14M_{s}^{2}M_{\phi}^{2}+3M_{\phi}^{4}}{120M_{s}^{2}(M_{\phi}^{2}+2M_{s}^{2})}. (64)

We note that the momentum dependence can be obtained by expressing Bϕ{\vec{B}}_{\phi}^{\prime} and Eϕ{\vec{E}}_{\phi}^{\prime} in Eq. (63) in terms of laboratory frame fields through Lorentz transformation.

Bϕ\displaystyle{\vec{B}}_{\phi}^{\prime} =\displaystyle= γBϕγv×Eϕ+(1γ)vBϕv2v,\displaystyle\gamma{\vec{B}}_{\phi}-\gamma{\vec{v}}\times{\vec{E}}_{\phi}+(1-\gamma)\frac{{\vec{v}}\cdot{\vec{B}}_{\phi}}{v^{2}}{\vec{v}},
Eϕ\displaystyle{\vec{E}}_{\phi}^{\prime} =\displaystyle= γEϕ+γv×Bϕ+(1γ)vEϕv2v,\displaystyle\gamma{\vec{E}}_{\phi}+\gamma{\vec{v}}\times{\vec{B}}_{\phi}+(1-\gamma)\frac{{\vec{v}}\cdot{\vec{E}}_{\phi}}{v^{2}}{\vec{v}}, (65)

where γ=Epϕ/Mϕ\gamma=E_{{\vec{p}}}^{\phi}/M_{\phi} is the Lorentz factor and v=p/Epϕ{\vec{v}}={\vec{p}}/E_{{\vec{p}}}^{\phi} is the velocity of the ϕ\phi meson. Then ρ00ϕ\rho_{00}^{\phi} can be rewritten as

ρ00ϕ(x,p)\displaystyle\rho_{00}^{\phi}(x,{\vec{p}}) \displaystyle\approx 134gϕ23Mϕ4Th2i=1,2,3IB,i(p)(Biϕ)2\displaystyle\frac{1}{3}-\frac{4g_{\phi}^{2}}{3M_{\phi}^{4}T_{\mathrm{h}}^{2}}\sum_{i=1,2,3}I_{B,i}({\vec{p}})({\vec{B}}_{i}^{\phi})^{2} (66)
4gϕ23Mϕ4Th2i=1,2,3IE,i(p)(Eiϕ)2,\displaystyle-\frac{4g_{\phi}^{2}}{3M_{\phi}^{4}T_{\mathrm{h}}^{2}}\sum_{i=1,2,3}I_{E,i}({\vec{p}})({\vec{E}}_{i}^{\phi})^{2},

where IB,i(p)I_{B,i}({\vec{p}}) and IE,i(p)I_{E,i}({\vec{p}}) are momentum functions. These functions are derived by substituting transformations (IV.2.2) into Eq. (63) and then separating different field components. For the global spin alignment, the explicit expression of IB,iI_{B,i} reads Sheng:2022ffb

IB,i(p)\displaystyle I_{B,i}(p) =\displaystyle= C1[(Epϕ)2+(6Epϕ(p2)2mϕ+Epϕ3(Epϕ)2)δ2,i\displaystyle C_{1}\left[(E_{p}^{\phi})^{2}+\left(\frac{6E_{p}^{\phi}(p_{2})^{2}}{m_{\phi}+E_{p}^{\phi}}-3(E_{p}^{\phi})^{2}\right)\delta_{2,i}\right. (67)
(1+3(p2)2(mϕ+Epϕ)2)(pi)2]\displaystyle\left.-\left(1+\frac{3(p_{2})^{2}}{(m_{\phi}+E_{p}^{\phi})^{2}}\right)(p_{i})^{2}\right]
+C2[pp(pi)23(j=1,2,3ϵ2ijpj)2],\displaystyle+C_{2}\left[\vec{p}\cdot\vec{p}-(p_{i})^{2}-3\left(\sum_{j=1,2,3}\epsilon_{2ij}p_{j}\right)^{2}\right],

where ϵ\epsilon in the last line is the rank-3 antisymmetric tensor. The explicit expression of IE,i(p)I_{E,i}(p) can be obtained from IB,iI_{B,i} by exchanging C1C_{1} with C2C_{2}. One can further take the space and momentum average of ρ00ϕ(x,p)\rho_{00}^{\phi}(x,{\vec{p}})

ρ00ϕ(x,p)x,p\displaystyle\left\langle\rho_{00}^{\phi}(x,{\vec{p}})\right\rangle_{x,{\vec{p}}}
\displaystyle\approx 134gϕ23Mϕ4Th2i=1,2,3IB,i(p)p(Biϕ)2x\displaystyle\frac{1}{3}-\frac{4g_{\phi}^{2}}{3M_{\phi}^{4}T_{\mathrm{h}}^{2}}\sum_{i=1,2,3}\left\langle I_{B,i}({\vec{p}})\right\rangle_{{\vec{p}}}\left\langle({\vec{B}}_{i}^{\phi})^{2}\right\rangle_{x}
4gϕ23Mϕ4Th2i=1,2,3IE,i(p)p(Eiϕ)2x,\displaystyle-\frac{4g_{\phi}^{2}}{3M_{\phi}^{4}T_{\mathrm{h}}^{2}}\sum_{i=1,2,3}\left\langle I_{E,i}({\vec{p}})\right\rangle_{{\vec{p}}}\left\langle({\vec{E}}_{i}^{\phi})^{2}\right\rangle_{x}, (68)

where the momentum average is defined as

O(p)p=d3p(Epϕ)1O(p)fϕ(p)d3p(Epϕ)1fϕ(p),\left\langle O({\vec{p}})\right\rangle_{\vec{p}}=\frac{\int d^{3}{\vec{p}}(E_{{\vec{p}}}^{\phi})^{-1}O({\vec{p}})f_{\phi}({\vec{p}})}{\int d^{3}{\vec{p}}(E_{\vec{p}}^{\phi})^{-1}f_{\phi}({\vec{p}})}, (69)

if we want to obtain momentum-integrated data for ρ00ϕ\rho_{00}^{\phi}. Here fϕ(p)f_{\phi}({\vec{p}}) is its momentum distribution which may contain information about collective flows such as v1v_{1} and v2v_{2}. If we want to obtain the transverse momentum spectra of ρ00ϕ\rho_{00}^{\phi}, we have to integrate over the azimuthal angle and rapidity and keep pTp_{T}, i.e. to replace d3p/Epϕd^{3}{\vec{p}}/E_{p}^{\phi} in (69) by dydφdyd\varphi. We see in Eq. (69) that the average ρ00ϕ\rho_{00}^{\phi} depends on the space average field squared which quantifies the field fluctuation. The theoretical results for ρ00ϕ\left\langle\rho_{00}^{\phi}\right\rangle as functions of transverse momenta, collision energies and centralities are presented in Ref. Sheng:2022wsy , which are in a good agreement with recent STAR data STAR:2022fan .

V Linear response theory for spin alignment of vector mesons in thermal media

In this section, we will show how to calculate the spin alignment of the vector meson from the Kubo formula in linear response theory zubarev1996statistical ; zubarev1997statistical ; Zubarev_1979 ; Kapusta:2006pm in thermalized QGP. The detailed discussion of this topic is in Ref. Dong:2023cng .

The Closed-Time-Path (CTP) formalism is a field-theory based method for many-particle systems in off-equilibrium as well in equilibrium Chou:1984es ; Blaizot:2001nr ; Wang:2001dm ; Berges:2004yj ; Cassing:2008nn ; Crossley:2015evo . When it is used for systems in equilibrium, it is actually the real time formalism of the thermal field theory Kapusta:2006pm ; Kapusta:2023eix . Wigner functions can be obtained from two-point Green’s functions on the CTP Sheng:2021kfc ; Sheng:2022ffb ; Hidaka:2022dmn ; Dong:2023cng ; Becattini:2024uha . The on-shell Wigner function for the vector meson is related to its MVSD (proportional to the spin density matrix) and defined as

Wμν(x,pon)=\displaystyle W^{\mu\nu}(x,p_{\mathrm{on}})= Epπ0𝑑p0G<μν(x,p)\displaystyle\frac{E_{p}}{\pi}\int_{0}^{\infty}dp_{0}G_{<}^{\mu\nu}(x,p)
=\displaystyle= m1,m2ϵμ(m1,p)ϵν(m2,p)fm1m2V(x,p),\displaystyle\sum_{m_{1},m_{2}}\epsilon^{\mu}\left(m_{1},{\vec{p}}\right)\epsilon^{\nu\ast}\left(m_{2},{\vec{p}}\right)f_{m_{1}m_{2}}^{V}(x,{\vec{p}}), (70)

where G<μν(x,p)G_{<}^{\mu\nu}(x,p) is the “++-” component of two-point Green’s functions on the CTP, and fV=Tr(fV)ρ^Vf_{V}=\mathrm{Tr}(f_{V})\hat{\rho}^{V} with fVfm1m2Vf_{V}\equiv f_{m_{1}m_{2}}^{V} being the short-hand notation for the MVSD, Tr(fV)=mfmmV\mathrm{Tr}(f_{V})=\sum_{m}f_{mm}^{V} being the trace of the MVSD, and ρ^V\hat{\rho}_{V} being the spin density matrix in Eq. (6). One can check that Wμν(x,pon)W^{\mu\nu}(x,p_{\mathrm{on}}) is always transverse to the on-shell momentum, pμonWμν(x,pon)=0p_{\mu}^{\mathrm{on}}W^{\mu\nu}(x,p_{\mathrm{on}})=0. The on-shell Wigner function can be decomposed into the scalar (𝒮\mathcal{S}), polarization (W[μν]W^{[\mu\nu]}) and tensor polarization (𝒯μν\mathcal{T}^{\mu\nu}) parts as Sheng:2023chinphyb ; Becattini:2024uha

Wμν(x,pon)=\displaystyle W^{\mu\nu}(x,p_{\mathrm{on}})= W[μν]+W(μν)\displaystyle W^{[\mu\nu]}+W^{(\mu\nu)}
=\displaystyle= 13Δμν(pon)𝒮+W[μν]+𝒯μν,\displaystyle-\frac{1}{3}\Delta^{\mu\nu}(p_{\mathrm{on}})\mathcal{S}+W^{[\mu\nu]}+\mathcal{T}^{\mu\nu}, (71)

where each part is defined as

W[μν]\displaystyle W^{[\mu\nu]}\equiv 12(WμνWνμ),\displaystyle\frac{1}{2}(W^{\mu\nu}-W^{\nu\mu}),
W(μν)\displaystyle W^{(\mu\nu)}\equiv 12(Wμν+Wνμ),\displaystyle\frac{1}{2}(W^{\mu\nu}+W^{\nu\mu}),
𝒯μν\displaystyle\mathcal{T}^{\mu\nu}\equiv W(μν)+13Δμν(pon)𝒮.\displaystyle W^{(\mu\nu)}+\frac{1}{3}\Delta^{\mu\nu}(p_{\mathrm{on}})\mathcal{S}. (72)

With Eq. (71) one can show that both W[μν]W^{[\mu\nu]} and 𝒯μν\mathcal{T}^{\mu\nu} are traceless, gμνW[μν]=gμν𝒯μν=0g_{\mu\nu}W^{[\mu\nu]}=g_{\mu\nu}\mathcal{T}^{\mu\nu}=0. Using Eqs. (6) and (70), we obtain

𝒮=\displaystyle\mathcal{S}= Tr(fV)=Δμν(pon)Wμν,\displaystyle\mathrm{Tr}(f_{V})=-\Delta^{\mu\nu}(p_{\mathrm{on}})W_{\mu\nu},
W[μν]=\displaystyle W^{[\mu\nu]}= 12Tr(fV)λ1,λ2ϵμ(λ1,p)ϵν(λ2,p)PiΣλ1λ2i,\displaystyle\frac{1}{2}\mathrm{Tr}(f_{V})\sum_{\lambda_{1},\lambda_{2}}\epsilon^{\mu}\left(\lambda_{1},{\vec{p}}\right)\epsilon^{\nu\ast}\left(\lambda_{2},{\vec{p}}\right)P_{i}\Sigma_{\lambda_{1}\lambda_{2}}^{i},
𝒯μν=\displaystyle\mathcal{T}^{\mu\nu}= Tr(fV)λ1,λ2ϵμ(λ1,p)ϵν(λ2,p)TijΣλ1λ2ij.\displaystyle\mathrm{Tr}(f_{V})\sum_{\lambda_{1},\lambda_{2}}\epsilon^{\mu}\left(\lambda_{1},{\vec{p}}\right)\epsilon^{\nu\ast}\left(\lambda_{2},{\vec{p}}\right)T_{ij}\Sigma_{\lambda_{1}\lambda_{2}}^{ij}. (73)

We see that W[μν]W^{[\mu\nu]} is related to PiP_{i} while 𝒯μν\mathcal{T}^{\mu\nu} is related to TijT_{ij}.

We can extract f00ρ00f_{00}\propto\rho_{00} by projecting

Lμν(pon)=ϵμ,(0,p)ϵν(0,p)+13Δμν(pon),L^{\mu\nu}(p_{\mathrm{on}})=\epsilon^{\mu,*}\left(0,{\vec{p}}\right)\epsilon^{\nu}\left(0,{\vec{p}}\right)+\frac{1}{3}\Delta^{\mu\nu}(p_{\mathrm{on}}), (74)

onto WμνW^{\mu\nu} in Eq. (70) as

Lμν(pon)Wμν\displaystyle L_{\mu\nu}(p_{\mathrm{on}})W^{\mu\nu}
=\displaystyle= λ1,λ2Lμν(pon)ϵμ(λ1,p)ϵν(λ2,p)fλ1λ2V(x,p)\displaystyle\sum_{\lambda_{1},\lambda_{2}}L_{\mu\nu}(p_{\mathrm{on}})\epsilon^{\mu}\left(\lambda_{1},{\vec{p}}\right)\epsilon^{\nu\ast}\left(\lambda_{2},{\vec{p}}\right)f_{\lambda_{1}\lambda_{2}}^{V}(x,{\vec{p}})
=\displaystyle= f00V(x,p)+13λ1,λ2ϵμ(λ1,p)ϵμ(λ2,p)fλ1λ2V(x,p)\displaystyle f_{00}^{V}(x,{\vec{p}})+\frac{1}{3}\sum_{\lambda_{1},\lambda_{2}}\epsilon^{\mu}\left(\lambda_{1},{\vec{p}}\right)\epsilon_{\mu}^{\ast}\left(\lambda_{2},{\vec{p}}\right)f_{\lambda_{1}\lambda_{2}}^{V}(x,{\vec{p}})
=\displaystyle= f00V(x,p)13Tr(fV).\displaystyle f_{00}^{V}(x,{\vec{p}})-\frac{1}{3}\mathrm{Tr}(f_{V}). (75)

In (74), ϵμ(0,p)\epsilon^{\mu}(0,{\vec{p}}) is the polarization vector along the spin quantization direction. With the first line of Eq. (73) and Eq. (75), we obtain

Lμν(pon)WμνΔμν(pon)Wμν=f00V(x,p)Tr[fV(x,p)]13=ρ0013.\frac{L_{\mu\nu}(p_{\mathrm{on}})W^{\mu\nu}}{-\Delta^{\mu\nu}(p_{\mathrm{on}})W_{\mu\nu}}=\frac{f_{00}^{V}(x,{\vec{p}})}{\mathrm{Tr}[f_{V}(x,{\vec{p}})]}-\frac{1}{3}=\rho_{00}-\frac{1}{3}. (76)

The above formula relates the Wigner function to ρ00\rho_{00}.

The medium effects can be described by retarded and advanced two-point Green’s functions through spectral functions. For vector mesons interacting with thermal quarks, the spectral function can be defined through the imaginary part of the retarded two-point Green’s function (propagator in momentum space) as

ImG~Rμν(p)=ΔTμνρT(p)+ΔLμνρL(p).\mathrm{Im}\widetilde{G}_{R}^{\mu\nu}(p)=\Delta_{T}^{\mu\nu}\rho_{T}(p)+\Delta_{L}^{\mu\nu}\rho_{L}(p). (77)

The definition of two-point Green’s functions GG and Σ\Sigma differs by a factor i=1i=\sqrt{-1} from the usual one in quantum field theory, which are related by G=iG~G=i\widetilde{G} and Σ=iΣ~\Sigma=i\widetilde{\Sigma}. In Eq. (77), we defined

ΔTμν=\displaystyle\Delta_{T}^{\mu\nu}= gμ0gν0+gμν+𝐩μ𝐩ν|p|2,\displaystyle-g^{\mu 0}g^{\nu 0}+g^{\mu\nu}+\frac{\mathbf{p}^{\mu}\mathbf{p}^{\nu}}{|{\vec{p}}|^{2}},
ΔLμν=\displaystyle\Delta_{L}^{\mu\nu}= ΔμνΔTμνgμνpμpνp2ΔTμν,\displaystyle\Delta^{\mu\nu}-\Delta_{T}^{\mu\nu}\equiv g^{\mu\nu}-\frac{p^{\mu}p^{\nu}}{p^{2}}-\Delta_{T}^{\mu\nu}, (78)

as the transverse and longitudinal projector respectively with 𝐩μ=(0,p)\mathbf{p}^{\mu}=(0,{\vec{p}}), and ρT,L\rho_{T,L} are spectral functions in the transverse and longitudinal directions given by Dong:2023cng

ρT(p)=\displaystyle\rho_{T}(p)= Im1p2m2+Σ~(p)+isgn(p0)ε,\displaystyle-\mathrm{Im}\frac{1}{p^{2}-m^{2}+\widetilde{\Sigma}_{\perp}(p)+i\,\mathrm{sgn}(p_{0})\varepsilon},
ρL(p)=\displaystyle\rho_{L}(p)= Im1p2m2+p2|𝐩|2Σ~00(p)+isgn(p0)ε,\displaystyle-\mathrm{Im}\frac{1}{p^{2}-m^{2}+\frac{p^{2}}{|\mathbf{p}|^{2}}\widetilde{\Sigma}_{00}(p)+i\,\mathrm{sgn}(p_{0})\varepsilon}, (79)

where Σ~\widetilde{\Sigma}_{\perp} and Σ~00\widetilde{\Sigma}_{00} are from Σ~Rμν\widetilde{\Sigma}_{R}^{\mu\nu}: Σ~(1/2)ΔμνTΣ~Rμν\widetilde{\Sigma}_{\perp}\equiv-(1/2)\Delta_{\mu\nu}^{T}\widetilde{\Sigma}_{R}^{\mu\nu} and Σ~00=Σ~R00\widetilde{\Sigma}_{00}=\widetilde{\Sigma}_{R}^{00}, sgn(p0)\mathrm{sgn}(p_{0}) is the sign of p0p_{0}, and ε\varepsilon is an infinitesimal positive number. One can check in Eq. (78) that pμΔTμν=pμΔLμν=0p_{\mu}\Delta_{T}^{\mu\nu}=p_{\mu}\Delta_{L}^{\mu\nu}=0. In Eq. (79), one can verify that the real parts of Σ~\widetilde{\Sigma}_{\perp} and Σ~00\widetilde{\Sigma}_{00} contribute to the mass correction while the imaginary parts of Σ~\widetilde{\Sigma}_{\perp} and Σ~00\widetilde{\Sigma}_{00} determines the width or life-time of the quasi-particle mode.

The two-point function G<μνG_{<}^{\mu\nu} is related to GAμνG_{A}^{\mu\nu} and GRμνG_{R}^{\mu\nu} through fetter2003quantum ; zubarev1997statistical

G<μν(p)\displaystyle G_{<}^{\mu\nu}(p) =\displaystyle= inB(p0)[G~Rμν(p)G~Aμν(p)]\displaystyle in_{B}(p_{0})\left[\widetilde{G}_{R}^{\mu\nu}(p)-\widetilde{G}_{A}^{\mu\nu}(p)\right] (80)
=\displaystyle= 2nB(p0)ImG~Rμν(p),\displaystyle-2n_{B}(p_{0})\mathrm{Im}\widetilde{G}_{R}^{\mu\nu}(p),

where nB(p0)=1/(eβp0βμV1)n_{B}(p_{0})=1/(e^{\beta p_{0}-\beta\mu_{V}}-1) is the Bose-Einstein distribution with the inverse temperature β1/T\beta\equiv 1/T and the vector meson’s chemical potential μV\mu_{V}. Inserting Eq. (80) into (70) we obtain the on-shell Wigner function Wμν(pon)W^{\mu\nu}(p_{\mathrm{on}}) from which the spin alignment ρ00\rho_{00} can be extracted through Eq. (76).

For free vector mesons, the spectral functions are ρT(0)=ρL(0)=πsgn(p0)δ(p2m2)\rho_{T}^{(0)}=\rho_{L}^{(0)}=\pi\mathrm{sgn}(p_{0})\delta(p^{2}-m^{2}), which give G<μν(p)G_{<}^{\mu\nu}(p) and ImG~Rμν(p)\mathrm{Im}\widetilde{G}_{R}^{\mu\nu}(p) for the free vector meson following Eqs. (77) and (80).

The non-equilibrium correction to G<μν(p)G_{<}^{\mu\nu}(p) can be calculated through the Kubo formula in linear response theory. The Kubo formula has been derived in Zubarev’s approach to non-equilibrium density operator Zubarev_1979 ; Hosoya:1983id ; Becattini:2019dxo . Detailed derivation is given in Ref. Becattini:2019dxo . Considering the thermal shear tensor ξμν=1/2(μβν+νβμ)\xi_{\mu\nu}=1/2(\partial_{\mu}\beta_{\nu}+\partial_{\nu}\beta_{\mu}) as a perturbation from local equilibrium, the next-to-leading order correction of G<μνG_{<}^{\mu\nu} can be written as Dong:2023cng

δG<μν(x,p)\displaystyle\delta G_{<}^{\mu\nu}(x,p)
=\displaystyle= 4TlimKμ0K0Imdp10dp202π\displaystyle 4T\lim_{K^{\mu}\rightarrow 0}\frac{\partial}{\partial K_{0}}\mathrm{Im}\int\frac{dp_{1}^{0}dp_{2}^{0}}{2\pi}
×nB(p10)nB(p20)p10p20+K0+iϵδ(p0p10+p202)ξγλ\displaystyle\times\frac{n_{B}(p_{1}^{0})-n_{B}(p_{2}^{0})}{p_{1}^{0}-p_{2}^{0}+K^{0}+i\epsilon}\delta\left(p^{0}-\frac{p_{1}^{0}+p_{2}^{0}}{2}\right)\xi_{\gamma\lambda}
×a,b=L,Tρa(p1)ρb(p2)Iabμνγλ(p1,p2),\displaystyle\times\sum_{a,b=L,T}\rho_{a}(p_{1})\rho_{b}(p_{2})I_{ab}^{\mu\nu\gamma\lambda}(p_{1},p_{2}), (81)

where p1=(p10,pK/2),p2=(p20,p+K/2)p_{1}=(p_{1}^{0},{\vec{p}}-{\vec{K}}/2),p_{2}=(p_{2}^{0},{\vec{p}}+{\vec{K}}/2). The tensor Iabμνγλ(p1,p2)I_{ab}^{\mu\nu\gamma\lambda}(p_{1},p_{2}) is expressed as Dong:2023cng

Iabμνγλ(p1,p2)\displaystyle I_{ab}^{\mu\nu\gamma\lambda}(p_{1},p_{2})
=\displaystyle= (p1λp2γ+p1γp2λ)Δa,αν(p1)Δbμα(p2)\displaystyle(p_{1}^{\lambda}p_{2}^{\gamma}+p_{1}^{\gamma}p_{2}^{\lambda})\Delta_{a,\alpha}^{\nu}(p_{1})\Delta_{b}^{\mu\alpha}(p_{2})
+(p1,αp2αmV2)\displaystyle+(p_{1,\alpha}p_{2}^{\alpha}-m_{V}^{2})
×[Δaγν(p1)Δbμλ(p2)+Δaλν(p1)Δbμγ(p2)]\displaystyle\times\left[\Delta_{a}^{\gamma\nu}(p_{1})\Delta_{b}^{\mu\lambda}(p_{2})+\Delta_{a}^{\lambda\nu}(p_{1})\Delta_{b}^{\mu\gamma}(p_{2})\right]
[p1γp2αΔa,αν(p1)Δbμλ(p2)\displaystyle-\left[p_{1}^{\gamma}p_{2}^{\alpha}\Delta_{a,\alpha}^{\nu}(p_{1})\Delta_{b}^{\mu\lambda}(p_{2})\right.
+p2γp1αΔaλν(p1)Δb,αμ(p2)]\displaystyle\left.+p_{2}^{\gamma}p_{1}^{\alpha}\Delta_{a}^{\lambda\nu}(p_{1})\Delta_{b,\alpha}^{\mu}(p_{2})\right]
[p1,αp2λΔaγν(p1)Δbμα(p2)\displaystyle-\left[p_{1,\alpha}p_{2}^{\lambda}\Delta_{a}^{\gamma\nu}(p_{1})\Delta_{b}^{\mu\alpha}(p_{2})\right.
+p1λp2,αΔaαν(p1)Δbμγ(p2)]\displaystyle\left.+p_{1}^{\lambda}p_{2,\alpha}\Delta_{a}^{\alpha\nu}(p_{1})\Delta_{b}^{\mu\gamma}(p_{2})\right]
gγλ[gβα(p2,ρp1ρmV2)p1,βp2,α]\displaystyle-g^{\gamma\lambda}\left[g_{\beta\alpha}(p_{2,\rho}p_{1}^{\rho}-m_{V}^{2})-p_{1,\beta}p_{2,\alpha}\right]
×Δaαν(p1)Δbμβ(p2),\displaystyle\times\Delta_{a}^{\alpha\nu}(p_{1})\Delta_{b}^{\mu\beta}(p_{2}), (82)

where ΔTμν(p)\Delta_{T}^{\mu\nu}(p) and ΔLμν(p)\Delta_{L}^{\mu\nu}(p) are given by Eq. (78). Then the spin alignment of vector mesons can be obtained from Eq. (76) through Eqs. (70) and (81).

Under the quasi-particle approximation (QPA), the self-energies and spectral functions can be calculated analytically by expanding p0p_{0} in powers of δp0=p0EpV\delta p_{0}=p_{0}-E_{p}^{V}. This expansion requires ΔE/EpVΓ/EpVϵ1\Delta E/E_{p}^{V}\sim\Gamma/E_{p}^{V}\sim\epsilon\ll 1, where ΔE\Delta E and Γ\Gamma are the mass shift and width of the vector meson respectively. Then the spin alignment can be expressed as

δρ00(p)=δρ00(ξ=0)(p)+ξμνCμν(p),\delta\rho_{00}({\vec{p}})=\delta\rho_{00}^{(\xi=0)}({\vec{p}})+\xi_{\mu\nu}C^{\mu\nu}({\vec{p}}), (83)

where δρ00(ξ=0)\delta\rho_{00}^{(\xi=0)} is the spin alignment independent of the shear stress tensor and CμνC^{\mu\nu} is the dimensionless tensor coefficient in the shear stress term. The results under the QPA are compared with full numerical results are shown in Fig. (4). For the values of parameters we choose, the magnitudes of δρ00(ξ=0)\delta\rho_{00}^{(\xi=0)} and Cμν(p)C^{\mu\nu}({\vec{p}}) are 𝒪(103)\mathcal{O}(10^{-3}) and 𝒪(102103)\mathcal{O}(10^{-2}\sim 10^{-3}), respectively. As a consequence, if the thermal shear tensor is 𝒪(102)\mathcal{O}(10^{-2}), the contribution from the thermal shear tensor to the ϕ\phi meson’s spin alignment is 𝒪(104105)\mathcal{O}(10^{-4}\sim 10^{-5}), which is much smaller than the contribution from the strong force fields Sheng:2022wsy .

Refer to caption
Figure 4: The full numerical and QPA results for CμνC^{\mu\nu} as functions of vector mesons’ momenta. In the upper panel the momentum is perpendicular to the spin quantization direction zz, while in the lower panel the momentum is parallel to the spin quantization direction.

VI Vector meson’s spin alignments in quark fragmentation

In high energy reactions, hadron production is normally described by fragmentation functions (FFs). The hadron‘s Polarization can be realized in terms of polarized FFs Chen:2016moq ; Chen:2016iey ; Chen:2020pty . These FFs are defined via quark-quark correlators. Among different high energy reactions, the clean and ideal place to study FFs is in high energy e+ee^{+}e^{-} collisions. In Ref. Chen:2016moq , the complete decomposition of quark-quark correlators for spin-1 hadrons with systematic results for hadron polarization in terms of FFs up to twist-3 in e+eVπXe^{+}e^{-}\to V\pi X have been presented. Experimental measurements have been carried out many years ago for the longitudinal polarization of Λ\Lambda hyperons ALEPH:1996oew ; OPAL:1997oem and for the spin alignments of vector mesons DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs in the inclusive production process e+ehXe^{+}e^{-}\to hX at LEP. Significant effects have been observed. These experimental data have attracted many phenomenological studies Anselmino:1997ui ; Anselmino:1998jv ; Anselmino:1999cg ; Xu:2001hz ; Xu:2003fq ; Chen:2016moq ; Chen:2016iey ; Chen:2020pty and parameterization of corresponding FFs has been obtained Chen:2016moq ; Chen:2016iey ; Chen:2020pty . In this section we will summarize these results.

VI.1 FFs from quark-quark correlator

For the fragmentation of the quark (or anti-quark), the quark-quark correlator as a 4×44\times 4 matrix in Dirac indices is defined by

Ξ^(k;p,S)=\displaystyle\hat{\Xi}(k;p,S)= Xd4ξ2πeikξhX|ψ¯(ξ)(ξ;)|0\displaystyle\sum_{X}\int\frac{d^{4}\xi}{2\pi}e^{-ik\xi}\langle hX|\bar{\psi}(\xi)\mathcal{L}(\xi;\infty)|0\rangle
×0|(0;)ψ(0)|hX,\displaystyle\times\langle 0|\mathcal{L}^{\dagger}(0;\infty)\psi(0)|hX\rangle, (84)

where kk and pp are four-momenta of the quark and hadron respectively, SS denotes the hadron’s spin, and (ξ;)\mathcal{L}(\xi;\infty) is the gauge link defined as

(ξ;)=Pexp[igξ𝑑ηA+(η;ξ+,ξ)],\mathcal{L}(\xi;\infty)=P\mathrm{exp}\left[ig\int_{\xi^{-}}^{\infty}d\eta^{-}A^{+}(\eta^{-};\xi^{+},\vec{\xi}_{\perp})\right], (85)

which guarantees the gauge invariance of FFs. In the rest part of the paper, for notational simplicity, we suppress the gauge link in forthcoming equations. In this case, the field operator ψ(ξ)\psi(\xi) just stands for (ξ;)ψ(ξ)\mathcal{L}^{\dagger}(\xi;\infty)\psi(\xi).

The FFs are obtained from Ξ^(k;p,S)\hat{\Xi}(k;p,S) in the following way. First, we expand Ξ^(k;p,S)\hat{\Xi}(k;p,S) in terms of Γ\Gamma matrices as

Ξ^(k;\displaystyle\hat{\Xi}(k; p,S)=12[Ξ(k;p,S)+iγ5Ξ~(k;p,S)\displaystyle p,S)=\frac{1}{2}\Bigl{[}\Xi(k;p,S)+i\gamma_{5}\tilde{\Xi}(k;p,S)
+γαΞα(k;p,S)+γ5γαΞ~α(k;p,S)\displaystyle+\gamma^{\alpha}\Xi_{\alpha}(k;p,S)+\gamma_{5}\gamma^{\alpha}\tilde{\Xi}_{\alpha}(k;p,S)
+iσαβγ5Ξαβ(k;p,S)],\displaystyle+i\sigma^{\alpha\beta}\gamma_{5}\Xi_{\alpha\beta}(k;p,S)\Bigr{]}, (86)

where coefficients are all real functions and are Lorentz scalar, pseudo-scalar, vector, axial-vector and tensor respectively. Then we expand these coefficient functions according to their respective Lorentz transformation properties in terms of basic Lorentz covariants multiplied by scalar functions which are constructed from available variables. These scalar functions are the FFs.

Clearly, basic Lorentz covariants to be constructed depend strongly on what variables that we have at hand. Besides pp and kk, we have variables for spin states which are different for spin-0, 1/2 or 1 hadrons. For this purpose, we need to decompose the spin density matrix in terms of known operators multiplied by Lorentz covariants. For spin-1/2 hadrons, the spin density matrix ρ\rho is decomposed as in Eq. (1), but S\vec{S} in the rest frame of the hadron should be generalized to a covariant form S=(0,S)S=(0,\vec{S}). For spin-1 hadrons, the 3×33\times 3 density matrix ρ^\hat{\rho} is usually decomposed as in Eq. (6). The tensor polarization Tij=Tr(ρΣij)T^{ij}={\rm Tr}(\rho\Sigma^{ij}) can be parameterized as

[T]=12(23SLL+STTxxSTTxySLTxSTTxy23SLLSTTxxSLTySLTxSLTy43SLL).\displaystyle[T]=\frac{1}{2}\left(\begin{array}[]{ccc}-\frac{2}{3}S_{LL}+S_{TT}^{xx}&S_{TT}^{xy}&S_{LT}^{x}\\ S_{TT}^{xy}&-\frac{2}{3}S_{LL}-S_{TT}^{xx}&S_{LT}^{y}\\ S_{LT}^{x}&S_{LT}^{y}&\frac{4}{3}S_{LL}\end{array}\right). (90)

We see that, for spin-1 hadrons, besides the polarization vector SS, there is also a tensor polarization part. The polarization vector is similar to that for spin-1/2 hadrons. The tensor polarization part has five components, namely a Lorentz scalar SLLS_{LL}, a Lorentz vector SLTμ=(0,SLTx,SLTy,0)S_{LT}^{\mu}=(0,S_{LT}^{x},S_{LT}^{y},0) and a Lorentz tensor STTμνS_{TT}^{\mu\nu} that has two components STTxx=STTyyS_{TT}^{xx}=-S_{TT}^{yy} and STTxy=STTyxS_{TT}^{xy}=S_{TT}^{yx}. The vector meson’s spin alignment ρ00\rho_{00} is only related to SLLS_{LL} through ρ00=(12SLL)/3\rho_{00}=(1-2S_{LL})/3.

The quark-quark correlator given by Eq. (84) is un-integrated, i.e. it depends on the four-momentum kk. If we consider three- or one-dimensional FFs, we need to integrate it over kk^{-} and k\vec{k}_{\perp}. We take one-dimensional FFs as an example. In this case, after integrating over kk^{-} and k\vec{k}_{\perp}, we obtain,

Ξ^(z;p,S)=\displaystyle\hat{\Xi}(z;p,S)= Xdξ2πeik+ξ\displaystyle\sum_{X}\int\frac{d\xi^{-}}{2\pi}e^{-ik^{+}\xi^{-}}
×hX|ψ¯(ξ)|00|ψ(0)|hX,\displaystyle\times\langle hX|\bar{\psi}(\xi^{-})|0\rangle\langle 0|\psi(0)|hX\rangle, (91)

where zp+/k+z\equiv p^{+}/k^{+}. After Lorentz decomposition, we obtain terms related to SLLS_{LL} as,

zΞα(z;p,S)=\displaystyle z\Xi_{\alpha}(z;p,S)= p+n¯α[D1(z)+SLLD1LL(z)]+\displaystyle p^{+}\bar{n}_{\alpha}[D_{1}(z)+S_{LL}D_{1LL}(z)]+
+powersuppressedterms.\displaystyle+{\rm power~{}suppressed~{}terms}. (92)

We can obtain the expression for D1(z)+SLLD1LL(z)D_{1}(z)+S_{LL}D_{1LL}(z) by reversely solving Eqs. (91) and (92),

D1(z)+\displaystyle D_{1}(z)+ SLLD1LL(z)=Xzdξ2πp+eik+ξ\displaystyle S_{LL}D_{1LL}(z)=\sum_{X}\int\frac{zd\xi^{-}}{2\pi p^{+}}e^{-ik^{+}\xi^{-}}
×hX|ψ¯(ξ)γ+|00|ψ(0)|hX,\displaystyle\times\langle hX|\bar{\psi}(\xi^{-})\gamma^{+}|0\rangle\langle 0|\psi(0)|hX\rangle, (93)

For comparison, we present the corresponding equations for G1LG_{1L} that describes the longitudinal spin transfer in the fragmentation process

zΞ~α(z;p,S)=\displaystyle z\tilde{\Xi}_{\alpha}(z;p,S)= λp+n¯αG1L(z)+powersuppressedterms\displaystyle\lambda p^{+}\bar{n}_{\alpha}G_{1L}(z)+{\rm power~{}suppressed~{}terms}
λG1L(z)=\displaystyle\lambda G_{1L}(z)= Xzdξ2πp+eik+ξ\displaystyle\sum_{X}\int\frac{zd\xi^{-}}{2\pi p^{+}}e^{-ik^{+}\xi^{-}}
×hX|ψ¯(ξ)γ+γ5|00|ψ(0)|hX.\displaystyle\times\langle hX|\bar{\psi}(\xi^{-})\gamma^{+}\gamma_{5}|0\rangle\langle 0|\psi(0)|hX\rangle. (94)

The difference between D1(z)+SLLD1LL(z)D_{1}(z)+S_{LL}D_{1LL}(z) in Eq. (93) and λG1L\lambda G_{1L} in Eq. (94) is that γ+\gamma^{+} is involved in the former while γ+γ5\gamma^{+}\gamma_{5} is involved in the latter.

VI.2 Vector meson’s spin alignments from FFs

From Eqs. (93), we see clearly that, similar to the well-known unpolarized FF D1(z)D_{1}(z), D1LLD_{1LL} is independent of the quark’s polarization because it is a sum over the quark’s spin states

D1(z)+SLLD1LL(z)=Xzdξ2πp+eik+ξ\displaystyle D_{1}(z)+S_{LL}D_{1LL}(z)=\sum_{X}\int\frac{zd\xi^{-}}{2\pi p^{+}}e^{-ik^{+}\xi^{-}}
×λq=L,RhX|ψ¯λq(ξ)γ+|00|ψλq(0)|hX,\displaystyle\times\sum_{\lambda_{q}=L,R}\langle hX|\bar{\psi}_{\lambda_{q}}(\xi^{-})\gamma^{+}|0\rangle\langle 0|\psi_{\lambda_{q}}(0)|hX\rangle, (95)

where λq=L,R\lambda_{q}=L,R denotes the quark’s helicity (or chirality), and ψL/R=(1±γ5)ψ/2\psi_{L/R}=(1\pm\gamma_{5})\psi/2. In contrast, the result for G1LG_{1L} gives

λG1L(z)=\displaystyle\lambda G_{1L}(z)= Xzdξ2πp+eik+ξ\displaystyle\sum_{X}\int\frac{zd\xi^{-}}{2\pi p^{+}}e^{-ik^{+}\xi^{-}}
×\displaystyle\times [hX|ψ¯L(ξ)γ+|00|ψL(0)|hX\displaystyle[\langle hX|\bar{\psi}_{L}(\xi^{-})\gamma^{+}|0\rangle\langle 0|\psi_{L}(0)|hX\rangle
hX|ψ¯R(ξ)γ+|00|ψR(0)|hX],\displaystyle-\langle hX|\bar{\psi}_{R}(\xi^{-})\gamma^{+}|0\rangle\langle 0|\psi_{R}(0)|hX\rangle], (96)

which depends on the quark’s spin explicitly.

We can draw an important conclusion from Eq. (95) for the spin alignment ρ00=(12SLL)/3\rho_{00}=(1-2S_{LL})/3 for the vector meson produced in the fragmentation process qVXq\rightarrow VX: it is determined by D1LLD_{1LL} and independent of the initial polarization of the quark. The conclusion is rather unexpected because the vector meson’s spin alignment in high energy reactions was first observed in e+eVXe^{+}e^{-}\rightarrow VX at LEP DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs where initial quarks and anti-quarks are longitudinally polarized. However, this is consistent with space reflection in fragmentation processes where ρ00\rho_{00} is space reflection invariant while the helicity of the initial quark changes the sign. This conclusion is rather solid since it follows from the general principle of QCD. It can also be tested easily in experiments. In the following we present numerical results for e+eVXe^{+}e^{-}\rightarrow VX and ppVXpp\rightarrow VX as examples.

VI.3 Vector meson’s spin alignments in e+eVXe^{+}e^{-}\rightarrow VX

Suppose that the quark fragmentation mechanism dominates hadron production in e+ee^{+}e^{-} annihilation at high energies, we obtain the vector meson’s alignment in e+eVXe^{+}e^{-}\rightarrow VX as

SLL(z,y)=qWq(y)D1LL(z)2qWq(y)D1(z),\displaystyle\langle S_{LL}\rangle(z,y)=\frac{\sum_{q}W_{q}(y)D_{1LL}(z)}{2\sum_{q}W_{q}(y)D_{1}(z)}, (97)
λ(z,y)=qPq(y)Wq(y)G1L(z)qWq(y)D1(z),\displaystyle\langle\lambda\rangle(z,y)=\frac{\sum_{q}P_{q}(y)W_{q}(y)G_{1L}(z)}{\sum_{q}W_{q}(y)D_{1}(z)}, (98)

where we also show the result of hyperon polarization for comparison. In Eqs. (97) and (98), Pq(y)P_{q}(y) and Wq(y)W_{q}(y) are the quark polarization and production weight at the vertex of e+ee^{+}e^{-} annihilation respectively which are determined by the quark’s electric charge, cVec_{V}^{e} and cAec_{A}^{e} in the weak interaction current, and so on (see e.g.  Chen:2016iey for the detailed expressions), y=l2pq/[(l1+l2)pq]y=l_{2}\cdot p_{q}/[(l_{1}+l_{2})\cdot p_{q}], where l1l_{1} and l2l_{2} are the four-momenta of the incident ee^{-} and e+e^{+} respectively, and pqp_{q} is the four-momentum of the produced quark.

The FFs are extracted from data available at a given scale and their scale dependence is determined by the QCD evolution equation, the DGLAP equation Dokshitzer:1977sg ; Gribov:1972ri ; Altarelli:1977zs ; Owens:1978qz ; Georgi:1977mg ; Uematsu:1978yw ; Ravindran:1996ri ; Ravindran:1996jd . In Ref. Chen:2016iey ; Chen:2020pty , such parameterizations of D1LLD_{1LL} for vector mesons have been obtained for the first time by fitting available data DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj . As comparison, parameterizations of G1LG_{1L} for Λ\Lambda are also given therein Chen:2016iey ; Chen:2020pty by fitting the LEP data ALEPH:1996oew ; OPAL:1997oem . Here, we show the fitted results obtained there Chen:2016iey ; Chen:2020pty in Figs. 5 and 6.

Refer to caption
Figure 5: Longitudinal polarization of Λ\Lambda in e+eΛXe^{+}e^{-}\to\Lambda X at high energies. The LEP data are taken from Refs. ALEPH:1996oew ; OPAL:1997oem . The solid line is the fit obtained in Ref. Chen:2016iey at the LEP energy while those at other energies are calculated results using the DGLAP equation for FFs and energy dependence of PqP_{q}. The figure is taken from Ref. Chen:2016iey .
Refer to caption
Figure 6: The spin alignments of K0K^{*0} and ρ0\rho^{0} in e+eVXe^{+}e^{-}\to VX at the ZZ-pole fitted in Ref. Chen:2020pty compared with experimental data DELPHI:1997ruo ; OPAL:1997vmw . The solid line is the fit in Ref. Chen:2020pty at the LEP energy while those at other energies are the results using the DGLAP equation for FFs. The figure is taken from Ref. Chen:2016iey .

From Figs. 5 and 6, we see a distinct feature that there is a strong energy dependence for PLΛP_{L\Lambda}, while the energy dependence for ρ00K\rho_{00}^{K^{*}} is weak. The former comes mainly from the energy dependence of PqP_{q} while the latter comes mainly from the QCD evolution of FFs. The relatively rapid change in the energy region around the mass of ZZ boson comes from the influence of WqW_{q}. This feature can be tested in future experiments.

VI.4 Vector meson’s spin alignments in ppVXpp\rightarrow VX

Assuming the universality of FFs, one can calculate vector meson’s spin alignments in other high energy reactions where the fragmentation mechanism dominates. This applies to hadron production at high transverse momentum in pppp collisions and deeply inelastic scatterings. In Ref. Chen:2020pty , such calculations have been carried out in pppp collisions. As examples, we show in Fig. 7 the results obtained there for K0K^{*0} and ρ0\rho^{0} in two rapidity regions as functions of pTp_{T}.

Refer to caption
Figure 7: Spin alignments of vector mesons in pppp collisions at s=200\sqrt{s}=200 GeV and s=5.02\sqrt{s}=5.02 TeV for K0K^{*0} and ρ0\rho^{0} in two rapidity regions as functions of pTp_{T}. This figure is taken from Ref. Chen:2020pty .

We see that ρ00V\rho^{V}_{00} for vector mesons deviate from 1/31/3 significantly showing large spin alignments. Such results can be tested in experiments at RHIC and LHC.

VII Summary and outlook

Spin alignments for vector mesons have been observed in high energy reactions STAR:2022fan ; DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs ; NOMAD:2006kuc . There are two types of spin alignments: global spin alignments in heavy-ion collisions STAR:2022fan and helicity-basis spin alignments in e+ee^{+}e^{-} collisions DELPHI:1997ruo ; OPAL:1997vmw ; OPAL:1997nwj ; OPAL:1999hxs . These measurements show different features of spin alignments arising from different hadronization mechanisms in collisions of heavy-ion and e+ee^{+}e^{-}.

In high energy heavy-ion collisions in the region of central rapidity and small transverse momentum, hadron production is mainly through quark combination or coalescence. The global spin alignment in this case not only depends on the polarization of quarks and anti-quarks but is also sensitive to the local correlation between the polarization of quarks and that of anti-quarks. Measurements of the global spin alignment in this case provide a good opportunity to study the polarization correlation in the sQGP produced in heavy-ion collisions. The STAR’s measurements of ϕ\phi meson’s spin alignments STAR:2022fan is consistent with the theoretical description based on ϕ\phi vector fields that lead to a strong local correlation between PsP_{s} and Ps¯P_{\bar{s}} Yang:2017sdk ; Sheng:2020ghv ; Sheng:2022ffb ; Sheng:2022wsy . The data were used to extract the strength of the local fluctuation in ϕ\phi vector fields Sheng:2022ffb ; Sheng:2022wsy .

In e+ee^{+}e^{-} and pppp collisions at high energies, hadron production is dominated by fragmentation with fragmentation functions. It has been shown that the spin alignment of vector mesons in the helicity basis is independent of the polarization of the initial quark in the quark fragmentation process. With the parameterization of corresponding fragmentation functions the prediction on spin alignments of vector mesons in high energy e+ee^{+}e^{-} and pppp collisions has been made Chen:2016moq ; Chen:2016iey ; Chen:2020pty .

It is impressive and interesting to see the distinction between the global spin alignments in high energy heavy-ion collisions and in e+ee^{+}e^{-} or pppp collisions which are dominated by different hadronization mechanisms. It is important to extend the measurements to different high energy reactions at different energies to test the universality of these properties. Further measurements on global spin alignments of different vector mesons in relativistic heavy ion collisions are expected to be carried out in the near future by the STAR Collaboration at RHIC and ALICE Collaboration at LHC. More studies are also planned at GSI, HIAF and NICA in the lower energy regions. Also studies in pppp collisions are expected by STAR at RHIC, and in e+ee^{+}e^{-} by BESIII at BEPC and Belle II at KEK. They can also be measured in lepton-nucleon scatterings at future EIC. Obviously such studies provide new insights into properties of sQGP and hadronization mechanisms.

Acknowledgements

This work was supported in part by the National Key Research and Development Program of China under Contract No. 2022YFA1604900, by the National Natural Science Foundation of China (NSFC) under Contract Nos. 12025501, 11890710, 11890713, 11890714, 12147101 and 12135011, by the Strategic Priority Research Program of the Chinese Academy of Sciences (CAS) under Grant No. XDB34030102 and by the Natural Science Foundation of Shandong Province.

References

  • (1) P. Braun-Munzinger, V. Koch, T. Schäfer, and J. Stachel, Phys. Rept. 621, 76 (2016), 1510.00442.
  • (2) W. Florkowski, A. Kumar, and R. Ryblewski, Prog. Part. Nucl. Phys. 108, 103709 (2019), 1811.04409.
  • (3) C. M. Ko, Nucl. Sci. Tech. 34, 80 (2023).
  • (4) R. Rapp, Nucl. Sci. Tech. 34, 63 (2023).
  • (5) Y. Zhang, D. W. Zhang, and X. F. Luo, Nucl. Tech. 46, 040001 (2023).
  • (6) J. Zhao, J.-H. Chen, X.-G. Huang, and Y.-G. Ma, Nucl. Sci. Tech. 35, 20 (2024).
  • (7) A. Pandav, D. Mallick, and B. Mohanty, Prog. Part. Nucl. Phys. 125, 103960 (2022), 2203.07817.
  • (8) W.-B. He, Y.-G. Ma, L.-G. Pang, H.-C. Song, and K. Zhou, Nucl. Sci. Tech. 34, 88 (2023).
  • (9) J. Chen, X. Dong, Y.-G. Ma, and Z. Xu, Sci. Bull. 68, 3252 (2023), 2311.09877.
  • (10) Y. G. Ma, Nucl. Sci. Tech. 34, 97 (2023).
  • (11) L. G. Pang and X. N. Wang, Nucl. Sci. Tech. 34, 194 (2023).
  • (12) K. J. Sun et al., Nucl. Tech. 46, 040012 (2023).
  • (13) Q. Chen, G. L. Ma, and J. H. Chen, Nucl. Tech. 46, 040013 (2023).
  • (14) Z.-T. Liang and X.-N. Wang, Phys. Rev. Lett. 94, 102301 (2005), nucl-th/0410079, [Erratum: Phys.Rev.Lett. 96, 039901 (2006)].
  • (15) Z.-T. Liang and X.-N. Wang, Phys. Lett. B 629, 20 (2005), nucl-th/0411101.
  • (16) J.-H. Gao et al., Phys. Rev. C 77, 044902 (2008), 0710.2943.
  • (17) S. A. Voloshin, (2004), nucl-th/0410089.
  • (18) B. Betz, M. Gyulassy, and G. Torrieri, Phys. Rev. C 76, 044901 (2007), 0708.0035.
  • (19) A. Ipp, A. Di Piazza, J. Evers, and C. H. Keitel, Phys. Lett. B 666, 315 (2008), 0710.5700.
  • (20) F. Becattini, F. Piccinini, and J. Rizzo, Phys. Rev. C 77, 024906 (2008), 0711.1253.
  • (21) Z.-t. Liang, J. Phys. G 34, S323 (2007), 0705.2852.
  • (22) STAR, J. H. Chen, J. Phys. G 34, S331 (2007).
  • (23) STAR, I. Selyuzhenkov, J. Phys. G 34, S1099 (2007), nucl-ex/0701034.
  • (24) STAR, B. I. Abelev et al., Phys. Rev. C 77, 061902 (2008), 0801.1729.
  • (25) STAR, B. I. Abelev et al., Phys. Rev. C 76, 024915 (2007), 0705.1691, [Erratum: Phys.Rev.C 95, 039906 (2017)].
  • (26) STAR, L. Adamczyk et al., Nature 548, 62 (2017), 1701.06657.
  • (27) Q. Wang, Nucl. Phys. A 967, 225 (2017), 1704.04022.
  • (28) Z.-T. Liang, M. A. Lisa, and X.-N. Wang, Nucl. Phys. News 30, 10 (2020), 1912.07822.
  • (29) J.-H. Gao, Z.-T. Liang, Q. Wang, and X.-N. Wang, Lect. Notes Phys. 987, 195 (2021), 2009.04803.
  • (30) X.-G. Huang, J. Liao, Q. Wang, and X.-L. Xia, Lect. Notes Phys. 987, 281 (2021), 2010.08937.
  • (31) J.-H. Gao, G.-L. Ma, S. Pu, and Q. Wang, Nucl. Sci. Tech. 31, 90 (2020), 2005.10432.
  • (32) F. Becattini and M. A. Lisa, Ann. Rev. Nucl. Part. Sci. 70, 395 (2020), 2003.03640.
  • (33) F. Becattini, Rept. Prog. Phys. 85, 122301 (2022), 2204.01144.
  • (34) F. Becattini et al., (2024), 2402.04540.
  • (35) STAR, M. S. Abdallah et al., Nature 614, 244 (2023), 2204.02302.
  • (36) Y.-G. Yang, R.-H. Fang, Q. Wang, and X.-N. Wang, Phys. Rev. C 97, 034917 (2018), 1711.06008.
  • (37) X.-L. Sheng, L. Oliva, and Q. Wang, Phys. Rev. D 101, 096005 (2020), 1910.13684, [Erratum: Phys.Rev.D 105, 099903 (2022)].
  • (38) X.-L. Sheng, L. Oliva, Z.-T. Liang, Q. Wang, and X.-N. Wang, Phys. Rev. D 109, 036004 (2024), 2206.05868.
  • (39) X.-L. Sheng, L. Oliva, Z.-T. Liang, Q. Wang, and X.-N. Wang, Phys. Rev. Lett. 131, 042304 (2023), 2205.15689.
  • (40) X.-N. Wang, Nucl. Sci. Tech. 34, 15 (2023), 2302.00701.
  • (41) J. Chen, Z.-T. Liang, Y.-G. Ma, and Q. Wang, Sci. Bull. 68, 874 (2023), 2305.09114.
  • (42) X.-L. Sheng, Z.-T. Liang, and Q. Wang, Acta Phys. Sin. 72, 072502 (2023).
  • (43) J.-H. Gao, X.-G. Huang, Z.-T. Liang, Q. Wang, and X.-N. Wang, Acta Phys. Sin. 72, 072501 (2023).
  • (44) L.-J. Ruan, Z.-B. Xu, and C. Yang, Acta Phys. Sin. 72, 112401 (2023).
  • (45) J. Chen et al., Nucl. Phys. News 34, 17 (2024).
  • (46) DELPHI, P. Abreu et al., Phys. Lett. B 406, 271 (1997).
  • (47) OPAL, K. Ackerstaff et al., Phys. Lett. B 412, 210 (1997), hep-ex/9708022.
  • (48) OPAL, K. Ackerstaff et al., Z. Phys. C 74, 437 (1997).
  • (49) OPAL, G. Abbiendi et al., Eur. Phys. J. C 16, 61 (2000), hep-ex/9906043.
  • (50) ZEUS, J. Breitweg et al., Eur. Phys. J. C 12, 393 (2000), hep-ex/9908026.
  • (51) NOMAD, A. Chukanov et al., Eur. Phys. J. C 46, 69 (2006), hep-ex/0604050.
  • (52) M. Anselmino, M. Bertini, F. Murgia, and P. Quintairos, Eur. Phys. J. C 2, 539 (1998), hep-ph/9704420.
  • (53) M. Anselmino, M. Bertini, F. Murgia, and B. Pire, Phys. Lett. B 438, 347 (1998), hep-ph/9805234.
  • (54) M. Anselmino, M. Bertini, F. Caruso, F. Murgia, and P. Quintairos, Eur. Phys. J. C 11, 529 (1999), hep-ph/9904205.
  • (55) Q.-h. Xu, C.-x. Liu, and Z.-t. Liang, Phys. Rev. D 63, 111301 (2001), hep-ph/0103267.
  • (56) Q.-h. Xu and Z.-t. Liang, Phys. Rev. D 67, 114013 (2003), hep-ph/0304125.
  • (57) K.-b. Chen, W.-h. Yang, S.-y. Wei, and Z.-t. Liang, Phys. Rev. D 94, 034003 (2016), 1605.07790.
  • (58) K.-b. Chen, W.-h. Yang, Y.-j. Zhou, and Z.-t. Liang, Phys. Rev. D 95, 034009 (2017), 1609.07001.
  • (59) K.-b. Chen, Z.-t. Liang, Y.-k. Song, and S.-y. Wei, Phys. Rev. D 102, 034001 (2020), 2002.09890.
  • (60) A. Bacchetta and P. J. Mulders, Phys. Rev. D 62, 114004 (2000), hep-ph/0007120.
  • (61) M. Jacob and G. C. Wick, Annals Phys. 7, 404 (1959).
  • (62) S. U. Chung, (1971).
  • (63) A. Lesnik et al., Phys. Rev. Lett. 35, 770 (1975).
  • (64) G. Bunce et al., Phys. Rev. Lett. 36, 1113 (1976).
  • (65) J. Bensinger et al., Phys. Rev. Lett. 50, 313 (1983).
  • (66) S. A. Gourlay et al., Phys. Rev. Lett. 56, 2244 (1986).
  • (67) TASSO, M. Althoff et al., Z. Phys. C 27, 27 (1985).
  • (68) ALEPH, D. Buskulic et al., Phys. Lett. B 374, 319 (1996).
  • (69) OPAL, K. Ackerstaff et al., Eur. Phys. J. C 2, 49 (1998), hep-ex/9708027.
  • (70) ALICE, S. Acharya et al., Phys. Rev. Lett. 125, 012301 (2020), 1910.14408.
  • (71) X.-L. Sheng, S. Pu, and Q. Wang, Phys. Rev. C 108, 054902 (2023), 2308.14038.
  • (72) X.-L. Xia, H. Li, X.-G. Huang, and H. Zhong Huang, Phys. Lett. B 817, 136325 (2021), 2010.01474.
  • (73) J.-H. Gao, Phys. Rev. D 104, 076016 (2021), 2105.08293.
  • (74) F. Li and S. Y. F. Liu, (2022), 2206.11890.
  • (75) D. Wagner, N. Weickgenannt, and E. Speranza, Phys. Rev. Res. 5, 013187 (2023), 2207.01111.
  • (76) W.-B. Dong, Y.-L. Yin, X.-L. Sheng, S.-Z. Yang, and Q. Wang, Phys. Rev. D 109, 056025 (2024), 2311.18400.
  • (77) B. Müller and D.-L. Yang, Phys. Rev. D 105, L011901 (2022), 2110.15630, [Erratum: Phys.Rev.D 106, 039904 (2022)].
  • (78) A. Kumar, B. Müller, and D.-L. Yang, Phys. Rev. D 108, 016020 (2023), 2304.04181.
  • (79) A. Manohar and H. Georgi, Nucl. Phys. B 234, 189 (1984).
  • (80) X.-L. Sheng, Q. Wang, and X.-N. Wang, Phys. Rev. D 102, 056013 (2020), 2007.05106.
  • (81) Z.-t. Liang, Off diagonal elements of the spin density matrix of vector mesons in heavy ion collisions, in The 7th International Conference on Chirality, Vorticity and magnetic field in heavy ion collisions, July 2023, Beijing, China, 2023, https://indico.ihep.ac.cn/event/16043/.
  • (82) J.-p. Lv, Z.-h. Yu, Z.-t. Liang, Q. Wang, and X.-N. Wang, Phys. Rev. D 109, 114003 (2024), 2402.13721.
  • (83) Z. Zhang, J.-p. Lv, Z.-h. Yu, and Z.-t. Liang, (2024), 2406.03840.
  • (84) D. Zubarev, V. Morozov, and G. Röpke, Statistical Mechanics of Nonequilibrium Processes, Basic Concepts, Kinetic TheoryStatistical Mechanics of Nonequilibrium Processes (Wiley-VCH Berlin, 1996).
  • (85) D. Zubarev, V. Morozov, and G. Röpke, Statistical Mechanics of Nonequilibrium Processes, Statistical Mechanics of Nonequilibrium Processes. Volume 2: Relaxation and Hydrodynamic ProcessesStatistical Mechanics of Nonequilibrium Processes (Wiley-VCH Berlin, 1997).
  • (86) D. N. Zubarev, A. V. Prozorkevich, and S. A. Smolyanskii, Theoretical and Mathematical Physics 40, 821 (1979).
  • (87) J. I. Kapusta and C. Gale, Finite-temperature field theory: Principles and applicationsCambridge Monographs on Mathematical Physics (Cambridge University Press, Cambridge, 2011).
  • (88) K.-c. Chou, Z.-b. Su, B.-l. Hao, and L. Yu, Phys. Rept. 118, 1 (1985).
  • (89) J.-P. Blaizot and E. Iancu, Phys. Rept. 359, 355 (2002), hep-ph/0101103.
  • (90) Q. Wang, K. Redlich, H. Stoecker, and W. Greiner, Phys. Rev. Lett. 88, 132303 (2002), nucl-th/0111040.
  • (91) J. Berges, AIP Conf. Proc. 739, 3 (2004), hep-ph/0409233.
  • (92) W. Cassing, Eur. Phys. J. ST 168, 3 (2009), 0808.0715.
  • (93) M. Crossley, P. Glorioso, and H. Liu, JHEP 09, 095 (2017), 1511.03646.
  • (94) J. I. Kapusta and C. Gale, Finite-Temperature Field TheoryCambridge Monographs on Mathematical Physics (Cambridge University Press, 2023).
  • (95) X.-L. Sheng, N. Weickgenannt, E. Speranza, D. H. Rischke, and Q. Wang, Phys. Rev. D 104, 016029 (2021), 2103.10636.
  • (96) Y. Hidaka, S. Pu, Q. Wang, and D.-L. Yang, Prog. Part. Nucl. Phys. 127, 103989 (2022), 2201.07644.
  • (97) X.-L. Sheng, Z.-T. Liang, and Q. Wang, Acta Phys. Sin. (in Chinese) 72, 072502 (2023).
  • (98) A. Fetter and J. Walecka, Quantum Theory of Many-particle SystemsDover Books on Physics (Dover Publications, New York, 2003).
  • (99) A. Hosoya, M.-a. Sakagami, and M. Takao, Annals Phys. 154, 229 (1984).
  • (100) F. Becattini, M. Buzzegoli, and E. Grossi, Particles 2, 197 (2019), 1902.01089.
  • (101) Y. L. Dokshitzer, Sov. Phys. JETP 46, 641 (1977).
  • (102) V. N. Gribov and L. N. Lipatov, Sov. J. Nucl. Phys. 15, 438 (1972).
  • (103) G. Altarelli and G. Parisi, Nucl. Phys. B 126, 298 (1977).
  • (104) J. F. Owens, Phys. Lett. B 76, 85 (1978).
  • (105) H. Georgi and H. D. Politzer, Nucl. Phys. B 136, 445 (1978).
  • (106) T. Uematsu, Phys. Lett. B 79, 97 (1978).
  • (107) V. Ravindran, Nucl. Phys. B 490, 272 (1997), hep-ph/9607384.
  • (108) V. Ravindran, Phys. Lett. B 398, 169 (1997), hep-ph/9606273.