logo

An introduction to relativistic spin hydrodynamics

INVITED REVIEW

An introduction to relativistic spin hydrodynamics

Xu-Guang Huang
Nuclear Science and TechniquesVol.36, No.11Article number 208Published in print Nov 2025Available online 17 Aug 2025
12400

Spin polarization and spin transport are common phenomena in many quantum systems. Relativistic spin hydrodynamics provides an effective low-energy framework to describe these processes in quantum many-body systems. The fundamental symmetry underlying relativistic spin hydrodynamics is angular momentum conservation, which naturally leads to inter-conversion between spin and orbital angular momenta. This inter-conversion is a key feature of relativistic spin hydrodynamics, which is closely related to entropy production and introduces ambiguity in the construction of constitutive relations. In this article, we present a pedagogical introduction of relativistic spin hydrodynamics. We demonstrate how to derive constitutive relations by applying local thermodynamic laws and explore several distinctive aspects of spin hydrodynamics. These include pseudo-gauge ambiguity, the behavior of the system in the presence of strong vorticity, and the challenges of modeling the freeze-out of spin in heavy-ion collisions. We also outline some future prospects for spin hydrodynamics.

Heavy-ion collisionSpin hydrodynamicsSpin polarization
1

Introduction

Spin is a fundamental property of particles arising from quantum mechanics, and it plays a central role in numerous phenomena within the quantum regime. As a form of angular momentum, spin naturally couples to rotation, allowing it to become polarized by rotational motion. Similarly, for a charged particle with nonzero spin, or a neutral particle with a non-trivial charge form factor, spin can couple to an external magnetic field as well. Additionally, for a particle in motion (i.e., with finite momentum), its spin may couple to acceleration, electric fields, or gradients of external potentials, such as chemical potential and temperature. In the case of massless particles, the spin state is specified by its helicity state, meaning that it is intrinsically slaved by the motion of the particle. As a result, spin can be manipulated by rotating fields, magnetic fields, electric fields, and several other external influences. Conversely, detecting the spin of a particle provides invaluable insights into the environment or underlying dynamics of the system.

In heavy-ion collision physics, the primary interest lies in the creation of deconfined quark-gluon matter, commonly referred to as the quark-gluon plasma (QGP) [1-5]. To uncover the properties of the QGP in heavy-ion collision experiments, it is essential to design specific hadronic observables that are sensitive to particular features of the QGP. Because charged particles are typically the easiest to detect, many observables rely on the charge of the hadrons. For instance, the total multiplicity of the detected charged hadrons reflects the initial energy of QGP. Meanwhile, the anisotropy in the momentum-space distribution of charged hadrons corresponds to the initial anisotropy in the spatial distribution of partons, leading to well-known harmonic flow parameters [6]. By measuring these hadronic observables, researchers have revealed several novel properties of hot and dense matter created during heavy-ion collisions. One significant finding is that the QGP must be extremely hot, with a typical temperature reaching 300-500 MeV at RHIC and LHC, indicating an extremely high energy density. Additionally, the QGP medium was found to interact strongly, with a very small shear viscosity to entropy density ratio η/s. This low ratio is required to explain the observed harmonic flow parameters [4, 5]. In fact, the η/s of the QGP is the lowest among all known fluids.

Since 2017, it has been established that the spin degree of freedom can be used to probe the properties of QGP [7]. This is achieved by measuring the spin polarization of spinful hadrons, such as hyperons and vector mesons [8-10]. Notably, it has been observed that the Λ and Λ¯ hyperons can exhibit significant spin polarization at collision energies of tens of GeV [7, 11-14]. Similarly, the ϕ and J/ψ mesons exhibited considerable spin alignment [15, 16]1. These discoveries open new avenues for studying the QGP through the spin degree of freedom. For instance, we now understand that the global spin polarization (i.e., the total amount of spin polarization with respect to the reaction plane) of hyperons arises from angular momentum conservation through the formation of fluid vortices within the QGP: In non-central heavy-ion collisions, the system possesses substantial orbital angular momentum, which subsequently induces strong fluid vorticity in the QGP [17-19], thereby polarizing the spins of quarks via spin-rotation coupling [20-35]. However, to fully understand the spin polarization phenomenon, a dynamical theory of spin polarization and spin transport in a hot medium is essential, analogous to the necessity of a dynamical theory of the bulk medium for understanding harmonic flows. Naturally, such a dynamical theory of spin transport can be derived from either kinetic or hydrodynamic theory. In recent years, both spin kinetic theory and spin hydrodynamics have made significant advancements. In this article, we focus on spin hydrodynamics, and refer readers to Refs. [36] for a review of spin kinetic theory, and Refs. [37-44, 44-49] for a review of spin polarization phenomena in heavy-ion collisions. In addition, we will focus only on spin polarization in hot and dense medium rather than in systems created in, for example, electron-ion collisions [50].

Throughout this article, we use the natural units c==kB=1 and the metric convention ημν=ημν=diag(1,1,1,1).

2

Relativistic hydrodynamics as an effective theory

Before discussing spin hydrodynamics, let us first briefly review the general structure of relativistic hydrodynamics from the perspective of effective field theory. The hydrodynamic theory describes the low-energy behavior of interacting many-body systems, where only conserved charge densities exhibit their dynamics. Because the conserved charge densities do not vanish, they redistribute themselves in space according to their equations of motion (EOMs). When expressed in a manner of spatial gradient expansion, these EOMs constitute hydrodynamic equations.

Let us consider the hydrodynamic theory of a system with space-time translation symmetry and global U(1) symmetry. The corresponding conserved charge densities are the energy density ε(x), momentum density πi(x),i=13, and the U(1) charge density n(x). We want to derive dynamical equations for these conserved charge densities. Sometimes, it is more convenient to work with potential variables conjugated to charge densities. These are the temperature T(x) (or its inverse β(x)=1/T(x)), fluid velocity uμ(x) normalized as uμuνημν=1, and chemical potential of n(x), μ(x). These conserved charge densities (or equivalently their conjugates) are hydrodynamic variables in hydrodynamics. Our starting point is conservation laws: μΘμν=0, (1) μJμ=0, (2) where Θμv is the energy-momentum tensor and is the U(1) current. As an effective field theory, we express Θμv and in terms of the conserved charged density (or equivalently, their conjugates) and their various gradient orders. We assume spatial isotropy of the system, that is, there are no external forces breaking the SO(3) symmetry. The building blocks are the fluid velocity and various quantities that can be classified into different representations of SO(3) in the fluid rest frame. Up to the first-order gradients, these quantities are Scalar:ε,n,Dε,Dn,θu=u,Vector:Duμ,με,μn,ωμν(1/2)(μuννuμ),Tensor:σμν(1/2)[μuν+νuμ(2/3)Δμνθ], (3) where Du is the co-moving time derivative, θ is the expansion rate of the fluid, μ=Δμνν is the spatial gradient operator, Δμνημνuμuν is the spatial projector, σμν is the shear tensor that is traceless, and ωμν is the vorticity tensor. Note that the co-moving time derivatives will eventually be replaced by spatial gradients using the EOMs in the leading order. Note that the vorticity tensor transforms in the same way as a three-vector under proper three-rotations (i.e., a three-rotation R with det R=1) because it can be substituted by a three pseudo-vector ωμ(1/2)ϵμνρσuνωρσ. Consequently, we can write the most general structure decomposition up to O() for Θμv and as follows2: Θμν=(a0+b0εDε+b0nDn+b0uθ)uμuν+c0(uμνε+uνμε)+d0(uμνn+uνμn)+e0(uμDuν+uνDuμ)+f0(uμων+uνωμ)+(g0+h0εDε+h0nDn+h0uθ)Δμν+i0σμν+j0(uμνεuνμε)+k0(uμνnuνμn)+l0(uμDuνuνDuμ)+m0(uμωνuνωμ)+n0ϵμνρσuρσε+o0ϵμνρσuρσn+p0ϵμνρσuρDuσ+q0ωμν+O(2), (4) Jμ=(A0+B0εDε+B0nDn+B0uθ)uμ+C0με+D0μn+E0Duμ+F0ωμ+O(2), (5) with all the coefficients (playing the roles of the Wilson coefficients in the effective field theory, as short-distance physics are encoded in these coefficients) functions of ε and n. They are constructed by first decomposing with respect to and then with respect to different representations of SO(3). In these decompositions, the terms with f0, m0, n0, o0, p0 in Θμv and F0 in as coefficients transform differently from Θμv and under parity (P), respectively, meaning that they can appear only when the system contains parity violating content. Under time reversal transformation (T), all the terms of first-order gradients on the right-hand sides of Θμv and except for terms with coefficients f0, m0, n0, o0, p0, F0 transform differently from Θμv and , respectively. This means that these terms must be dissipative (i.e., these terms are responsible for entropy generation in the fluid), while terms with coefficients f0, m0, n0, o0, p0, F0 can appear without generating entropy, that is, they could arise in ideal hydrodynamics despite being at first order in gradients. Thus, the terms with coefficients f0, m0, n0, o0, p0, F0 are especially interesting. In fact, some of them have been intensively studied, and it was found that they contain very rich quantum phenomena (usually dubbed chiral anomalous transports) that are closely related to the chiral anomaly of the system if the underlying physics is governed by the quantum gauge theory. Recently, such chiral anomalous transport has become an active subject in condensed matter physics, astrophysics, and heavy-ion collision physics (see Refs. [41, 51-56] for recent reviews focusing on heavy-ion collision physics). Similarly, we can also examine the balance between the right-hand and left-hand sides of Eqs. (4)-(5) under charge conjugation (C) transformation. The terms with coefficients b0n, d0, h0n, k0, o0, A0, B0ε, B0u, C0, E0, F0 must vanish if there is no environmental charge-conjugation violation (naturally, the presence of a nonzero charge density n violates the C symmetry and allows these terms to be present). The antisymmetric terms in Θμv are particularly interesting. To reveal their meaning, we consider the angular momentum conservation law: μMμνρ=0, (6) where Mμνρ is the angular momentum tensor Mμνρ=xνΘμρxρΘμν+Σμνρ, (7) and Σμνρ is the spin tensor. We can re-write Eq. (6) in the following form: μΣμνρ=ΘρνΘνρ. (8) Thus, the antisymmetric part of Θμv provides a source for spin generation (one may more clearly see this by integrating Eq. (8) over space). This will be the focus of this study, and we return to it in the next section. In the remainder of this section, for the purpose of demonstrating the construction of the hydrodynamic theory, we simply assume that the system does not possess a spin tenor, so that Θμv is symmetric, Θμv, and assume that there is no environmental parity violation, so that terms with coefficients f0, m0, n0, o0, p0, F0 must vanish. Thus, the most general decomposition of Θμv and up to the first order in gradients into different components with respect to , and subsequently, for the components orthogonal to , with respect to different irreducible tensor structures under SO(3) are as follows: Θμν=(a0+b0εDε+b0nDn+b0uθ)uμuν+c0(uμνε+uνμε)+d0(uμνn+uνμn)+e0(uμDuν+uνDuμ)+(g0+h0εDε+h0nDn+h0uθ)Δμν+i0σμν+O(2), (9) Jμ=(A0+B0εDε+B0nDn+B0uθ)uμ+C0με+D0μn+E0Duμ+O(2). (10) Up to this point, expressions (9)–(10) are merely parameterizations of Θμv and , and such a parameterization is ambiguous at O() order (and higher orders in gradients). To see this, we consider to re-express Θμv and in terms of a redefinition of the hydrodynamic variables ε, n, uμ which differ from ε, n, uμ by O()-order shifts: ε=ε+δε,n=n+δn,uμ=uμ+δuμ, (11) where δε, δn, δuμ are order-O() quantities and uμδuμ=O(2) so that u2=1 is maintained at O(). This can be seen by noting that u2=u2+2uμδuμ+δu2 which leads to 2uμδuμ+δu2=O(2) and therefore uμδuμ must be O(2). In terms of the primed variables, we obtain Θμν=[a0(a0εδε+a0nδn)+b0εDε+b0nDn+b0uθ]uμuν+c0(uμνε+uνμε)+d0(uμνn+uνμn)+e0(uμDuν+uνDuμ)+(g0a0)(δuμuν+uμδuν)+[g0(g0εδε+g0nδn)+h0εDε+h0nDn+h0uθ]Δμν+i0σμν+O(2), (12) Jμ=[A0(A0εδε+A0nδn)+B0εDε+B0nDn+B0uθ]uμA0δuμ+C0με+D0μn+E0Duμ+O(2), (13) where a0=a0(ε, n) and similarly for g0, A0 and all second order terms are omitted. By observing the expressions in the three square brackets, one can see that by suitably choosing δε and δn, one can eliminate the first-order terms in two of the three square brackets. For example, one can solve out δε and δn by requiring the first-order terms in square brackets in Θμv to vanish. However, it is more convenient to eliminate the first-order terms in the coefficients of uμuν in Θμv and uμ in . Similarly, by suitably choosing δuμ, one can eliminate either the second line in Θμv (this choice is called the Landau-Lifshitz frame for ) or the second line in (this choice is called the Eckart frame for ). Therefore, we can always choose the following simpler forms for Θμv and (Landau-Lifshitz frame), Θμν=a0uμuν+(g0+h0εDε+h0nDn+h0uθ)Δμν+i0σμν+O(2), (14) Jμ=A0uμ+C0με+D0μn+E0Duμ+O(2). (15) Contracting with , we can identify that a0=uμuνΘμν which is the local energy density ε and A0=uJ which is the local U(1) charge density n. [Sometimes, this is also considered as the matching condition because this means that uμuνΘμν=uμuνΘ(0)μν and uμJμ=uμJ(0)μ with Θ(0)μν and J(0)μ is the zeroth order energy-momentum tensor and charge current.]

Let us first consider the zeroth-order terms, which, as we have already discussed, correspond to ideal hydrodynamics: Θ(0)μν=εuμuν+g0Δμν, (16) J(0)μ=nuμ. (17) In the rest frame of the fluid, =(1, 0), it becomes Θ(0)μν=diag(ε, g0, g0, g0) which identifies g0 as the thermodynamic pressure P. In the zeroth order, the conservation laws are (ε+P)DuμμP=0, (18) Dε+(ε+P)θ=0, (19) Dn+nθ=0. (20) To close these equations, we need to know the thermodynamic relation among P,ε,n, that is, the equation of state, P=P(ε,n).

Let us then consider the first-order terms that correspond to dissipative hydrodynamics. From Eqs. (18)-(20), we notice that we could replace Dε and Dn in the first order terms by (ε+P)θ and -n and Duμ by μP/(ε+P). This allows us to re-write the energy-momentum tensor and charge current as Θμν=εuμuν(P+h0θ)Δμν+i0σμν+O(2), (21) Jμ=nuμ+C0με+D0μn+O(2), (22) with h0=(ε+P)h0ε+nh0nh0u, C0=C0+E0(P/ε)n/(ε+P), and D0=D0+E0(P/n)ε/(ε+P). Further constraints can be imposed, based on the laws of local thermodynamics. For a fluid at rest, we have the first law of thermodynamics as Tds+μdn=dε, (23) Ts+μn=ε+P, (24) where s denotes the entropy density. To proceed, we propose the covariant generalization of the second one (Gibbs-Duhem relation): sμ=Pβμ+ΘμνβναJμ, (25) where βμ=βuμ(β=1/T), α=μ/T, and is the entropy current, such that us=s. The divergence of (multiplied by T) can be calculated directly as: Tμsμ=Θ(1)μνμuνTJ(1)μμα. (26) The second law of local thermodynamics requires that Tμsμ0 for any configuration of the velocity field , temperature T, and chemical potential μ, which imposes the following constraints: h0=ζ0,i0=2η0,J(1)μ=σμα, (27) where ζ and η are the bulk and shear viscosities, respectively, and σ is charge conductivity. This also shows that the coefficients C0 and D0 are fixed in such a way that C0με+D0μn=σμα. The EOMs of the first-order dissipative hydrodynamics are then read (ε+Pζθ)Duμμ(Pζθ)+2ηΔνμρσνρ=0, (28) Dε+(ε+Pζθ)θ2ησμνσμν=0, (29) Dn+nθ+σ2α=0. (30) The first equation is the relativistic Navier-Stokes equation. The above procedure can continue to a higher order in gradients and provide higher-order hydrodynamics. However, we did not discuss these more complicated situations. Readers can find discussions in Refs. [57-61].

3

Construction of relativistic spin hydrodynamics

With the above preparation, we now discuss the construction of relativistic spin hydrodynamics, in which the conservation of angular momentum is explicitly encoded within a (quasi)-hydrodynamic framework. The fundamental conservation laws are the energy-momentum conservation (1) and angular momentum conservation (8). Before delving into the detailed construction, we note that if we assign spin density Sμν=uρΣρμν as a dynamic variable in our framework, Eq. (8) indicates that it is generally not conserved. This reflects the fact that the spin angular momentum can be transformed into orbital angular momentum, thus disqualifying it as a true hydrodynamic mode. Consequently, spin hydrodynamics is not a strict hydrodynamic theory for the gapless modes. Instead, it should be categorized as quasi-hydrodynamics, where the low-energy dynamic variables comprise true hydrodynamic modes and some gapped modes (quasi-hydrodynamic modes) whose gap in the low-momentum region is parametrically small compared with other microscopic modes (the hard modes of the system) [62]. This results in spectrum separation; for physics at energy scales comparable to these modes, we can only consider the quasi-hydrodynamic modes alongside the true hydrodynamic modes. Generalized hydrodynamics [63] and Hydro+ [64] near the QCD critical point fall into this category. The spin hydrodynamics that we will discuss also belongs to this type of theory. This framework requires that spin excitations, despite being gapped, remain low-energy excitations compared with other microscopic modes [62]. For instance, if the system contains massive fermions, the spins of these fermions are difficult to relax because the spin-orbit coupling is inversely suppressed by the mass of the fermions compared to the typical energy transfer [65-67]. Thus, these spins are quasi-conserved, and we can formulate a quasi-hydrodynamic theory for it, which is called spin hydrodynamics.

We consider a charge-neutral system such as the quark gluon plasma or the usual electric plasma, in which some of the constituent particles are spinful particles. The symmetry considered is pacetime translation symmetry and Lorentz symmetry. This leads to the energy-momentum conservation and angular momentum conservation, as given by Eq. (1) and Eq. (8). Now, the spin tensor Σμρσ plays the role of the charge current and we can write it as Σμρσ=Sρσuμ+higherorderterms, with the spin density Sρσ playing a similar role to the charge density n in Eq. (22). To proceed, we need to choose a suitable power-counting scheme for all (quasi-)hydrodynamic variables. If we consider the QGP in heavy ion collisions, from the measurements of global spin polarization of hyperons, we know that the spin density in the QGP should be small because the hyperon spin polarization is only a few percent. Thus, it is reasonable to assume that the spin density Sρσ is parametrically smaller than the true hydrodynamic modes described by variables ε and . Thus, we take the following power-counting scheme: ε, P, T, uμO(1), (31) SρσO(). (32) Analogous to the fact that the chemical potential μ is conjugate to the charge density n, we can introduce the spin potential μρσ to be conjugate thermodynamically to the spin density Sρσ and propose the first law for local thermodynamics as (analogous to Eq. (23)): Tds+12μμνdSμν=dε, (33) Ts+12μμνSμν=ε+P. (34) Following the discussions on the fluid local frame, we realize that the same discussions are still valid for the symmetric part of the energy-momentum tensor; thus, we still choose the definition of such that it is the eigenvector of the symmetric part of the energy-momentum tensor, Θsμν (we still call it the Landau-Lifshitz frame)3: Θsμνuν=εuμ. (35) Because the EOM for spin density involves only the antisymmetric part of the energy-momentum tensor Θaμν, the symmetric part Θsμν, still takes the same tensor decomposition upto the first order in gradients, as in Eq. (21): Θsμν=εuμuν(Pζθ)Δμν+2ησμν+O(2). (36) To determine the form of Θaμν, we used the second law of local thermodynamics. The covariant entropy current is (an analog of Eq. (25)) sμ=Pβμ+Θμνβν12αρσΣμρσ, (37) with αρσ=μρσ/T. The production rate of entropy then reads Tμsμ=Θs(1)μν(μuν)+Θaμν(μμν+T[μβν])+O(3). (38) The semi-positiveness of the first term on the right-hand side is guaranteed when both the bulk and shear viscosities are semi-positive. The requirement of the semi-positiveness of the second term gives the constitutive relation for Θaμν at O() order [68] Θaμν=qμuνqνuμ+ϕμν, (39) qμ=λ[βμT+Duμ2μμνuν], (40) ϕμν=ηsΔμρΔνσ(μρσTϖρσ). (41) The quantity ϖμν=(1/2)(νβμμβν) (42) is the thermal vorticity tensor. The quantities λ and ηs must be semi-positive to guarantee the semi-positivity of the entropy production. These are called boost heat conductivity and rotational viscosity, respectively [68]. Using these constitutive relations, we obtain the spin hydrodynamic equations up to O(2) order: (ε+Pζθ)Duμμ(Pζθ)+2ηΔνμρσνρ+quμΔνμDqνqμθ+Δρμνϕνρ=0, (43) Dε+(ε+Pζθ)θ2ησμνσμν+q+qμDuμ+ϕμνωμν=0, (44) DSρσ+Sρσθ+2Θaρσ=0. (45) In this section, we present a detailed derivation of the constitutive relations for relativistic spin hydrodynamics up to first order. For related discussions that follow a similar approach, see Refs.[62, 69-78]. Other methodologies for deriving and analyzing the constitutive relations of spin hydrodynamics have also been discussed in the literature, including utilizing the hydrostatic partition function with constraints from the entropy current and Onsager relations [79, 80], using local equilibrium and non-equilibrium statistical operators [73, 77, 81-83], and employing kinetic theories [76, 84-95]. Relativistic spin hydrodynamics have become a vibrant area of research, attracting intense discussion in recent years. In the following section, we will explore some of these developments; further insights can be found in Refs. [96-110].

4

Discussions

We developed spin hydrodynamics based on local thermodynamic laws. Spin hydrodynamics exhibit several novel features that differ significantly from those found in conventional relativistic hydrodynamics for other types of conservation laws (e.g., the energy-momentum conservation and baryon number conservation). In this section, we explore and discuss certain intriguing characteristics.

4.1
Pseudo-gauge ambiguity

The definition of conserved current is not unique. One example is the magnetization current and dipole charge density. Let = (ρ, J) represent the conserved conductive electric current. For a polarizable and magnetizable material, the total charge density and electric current are given by ρ˜=ρ+P and J˜=J+×M, respectively, where P is the electric dipole density, and M is the magnetization density. In the covariant form, we have: J˜μ=Jμ+νMμνwithMμν=Mνμ. (46) Obviously, the total current J˜μ is conserved if the conduction current is conserved, and the total electric charge remains unchanged provided the surface dipole density vanishes. The transformation of a conserved current that preserves both the original conservation law and total conserved charge is called a pseudo-gauge transformation. The example above demonstrates that the total current and conduction current differ by a pseudo-gauge transformation (with the magnetization Mμν serving as the pseudo-gauge field). This example also highlights that a pseudo-gauge transformation is not a true gauge transformation because it alters the physical content of the transformed current. Further insight into the pseudo-gauge transformation can be obtained by examining Maxwell’s equations: μFμν=J˜ν. (47) One could subtract ρMνρ from both sides and find μHμν=Jν, (48) where the new field-strength tensor is defined as HμνFμν+Mμν. This demonstrates that without imposing additional constraints, the two sets of fields, (Fμν, J˜μ) and (Hμν, Jμ), describe the same physical laws, and one can freely choose which set to use. (If further constraints are imposed, such as the Bianchi equation [ρFμν]=0, which is not preserved under a general pseudo-gauge transformation, then only certain pseudo-gauges that respect the Bianchi equation are permitted.)

Similarly, let us consider angular momentum conservation (note the analogy with Eq. (48), where Σμνρ and Θρv - Θ are analogous to Hμν and in Eq. (48)): μΣμνρ=ΘρνΘνρ, (49) which is preserved under the transformation ΣμρσΣ˜μρσΣμρσΦμρσ, (50) ΘμνΘ˜μνΘμν+12λΦλμν, (51) with Φλμν=Φλνμ denotes an arbitrary local field. However, this transformation violates the conservation law of the energy-momentum tensor. It can be modified as following: ΣμρσΣ˜μρσΣμρσΦμρσ, (52) ΘμνΘ˜μνΘμν+12λ(ΦλμνΦμλνΦνλμ), (53) which preserves both Eq. (49) and Eq. (1). Given a spacelike hypersurface Ξ, the total energy-momentum and total angular momentum across Ξ are Pν=dΞμΘμν, (54) Mρσ=dΞμMμρσ           =dΞμ(xρΘμσxσΘμρ+Σμρσ). (55) One can check that and Mρσ are invariant under pseudo-gauge transformation (52) and (53) if the pseudo-gauge field Φμρσ vanishes at the boundary of Ξ4.

One consequence of the pseudo-gauge transformation is the freedom to choose the symmetry properties of the spin tensor. To illustrate this, we consider an example in which we aim to transform the general spin tensor Σμρσ=Σμσρ into a completely antisymmetric form. We can choose Φμρσ=Σ(μρ)σ12Σσμρ. After applying the pseudo-gauge transformation, we obtain ΣμρσΣ˜μρσ=12(ΣμρσΣρμσ+Σσμρ), (56) ΘμνΘ˜μν=Θμν+14λ(3Σνμλ+ΣμνλΣλνμ). (57) Note that the obtained Σ˜μρσ is totally antisymmetric; therefore, it can be parameterized as Σ˜μρσ=ϵμρσνS˜ν, (58) where S˜μ denotes the corresponding spin (pseudo)vector. Thus, the spin density tensor is thus S˜μν=ϵμνρσuρS˜σ. The main difference between this spin density tensor and that used in Sect. 3 is that S˜μν contains three degrees of freedom corresponding to the three spatial spin vectors, whereas Sμν has six degrees of freedom, with three for spatial spin and three for boost. Thus, in some cases, it is more convenient to use Σ˜μρσ to construct the spin hydrodynamics. By following a procedure similar to that adopted in Sect. 3, we can derive constitutive relations in this context. In doing so, we decompose S˜μ into S˜μ=σμ+n5uμ, where σμ represents the spatial spin with the condition σu=0, and n5 is a pseudoscalar field (hence subscript 5). We also decompose Θ˜μν into: Θ˜μν=εuμuνPΔμν+Θ˜s(1)μν+q˜μuνq˜νuμ+ϕ˜μν, (59) where, as in Sect. 3, we assumed the Landau-Lifshitz frame Θ˜sμνuν=εuμ, (60) such that Θ˜s(1)μν is purely transverse to . It is important to note that although we use the same symbols ε, P, and as in Sect. 3, their actual values may differ because the energy-momentum tensors and spin tensors in these two cases are different (but connected by the pseudo-gauge transformations (56) and (57)). We adopt a power counting scheme similar to the one we chose in Sect. 3, ε, P, T, uμO(1), (61) S˜μ, q˜μ, ϕ˜μνO(). (62) Using the same form for the entropy current and the first law for local thermodynamics presented in Sect. 3, one can then find the divergence of the entropy current to be Tμsμ=Θ˜s(1)μν(μuν)+Θ˜aμν(μ˜μν+T[μβν])+O(3). (63) We note that in deriving this result, we have utilized the fact that contracting the equation of motion (49) with uρ reveals that q˜μ is not an independent current, but is determined by S˜μ through the following relation: q˜μ=12ϵμνρσuνρS˜σ. (64) This is because when the spin tensor is completely antisymmetric, the components responsible for the boost are gauged away, meaning that the corresponding torque for the boost in the antisymmetric part of the energy-momentum tensor cannot be an independent current either. Owing to this relationship, we can show that n5=S˜u is actually an O(3) quantity (and thus does not appear on the right-hand side of Eq. (63)). In fact, through direct calculation, one can find that the higher order terms that are neglected in Eq. (63) contains only one term: n5, 12n5ϵμνρσuσ[μ(βμ˜νρ)+μuρ(νβ+Dβν)], (65) which infers that n5ϵμνρσuσ[μ(βμ˜νρ)+μuρ(νβ+Dβν)]O(3) and thus can be neglected [62]. Therefore, from Eq. (63), we derive the constitutive relations for spin hydrodynamics with a completely antisymmetric spin tensor as follows [62]: Θ˜s(1)μν=ζθΔμν+2ησμν, (66) Θ˜a(1)μν=ϕ˜(1)μν=ηsΔμρΔνσ(μρσTϖρσ). (67) Although these relations take the same form as those obtained in Sect. 3, it is important to note that they apply specifically to the pseudo-gauge of a completely antisymmetric spin tensor. These relations are particularly convenient for describing the evolution of spatial spin degrees of freedom.

Thus, choosing different forms for the spin tensor (loosely referred to as different pseudo-gauges) leads to different forms for the constitutive relations. In an extreme case, one might even select Φμρσ=Σμρσ, which completely eliminates the spin tensor and renders the energy-momentum tensor totally symmetric (this choice is commonly referred to as the Belinfante gauge [111-113]). While this may seem to eliminate all information about spin in hydrodynamics, the energy density, viscous tensors, and heat current remain influenced by spin, meaning that the dynamics of spin are still embedded within these quantities. For discussions on the transformation from canonical to Belinfante gauges, see Refs. [69, 71, 75, 99, 114, 115]. In addition, other pseudo-gauges have been employed and discussed in the context of spin hydrodynamics [85, 86, 88, 116-119].

4.2
Spin hydrodynamics for strong vorticity

The power counting scheme employed in the previous discussions is motivated by the observation that at global equilibrium, the spin potential μμν is determined by the thermal vorticity ϖμν=(νβμμβν)/2, which is naturally assumed to be an O() quantity. However, this assumption may not hold true because the global equilibrium allows for arbitrarily large rotations (vorticity). When the vorticity is large, the assignment ϖμνO() becomes inadequate; instead, it is more appropriate to consider that ϖμνO(1). We explore this situation in this subsection, following closely the discussions in Ref. [72]. Before going into the details, it is useful to compare spin hydrodynamics with magnetohydrodynamics (MHD), in which the magnetic field is treated as an O(1) quantity (See Ref. [120] for a review of relativistic MHD).

The MHD describes the coupled evolution of fluid energy-momentum (or temperature and velocity) and the electromagnetic field in the low-energy and long-wavelength regimes. The fundamental equations consist of the conservation laws for the energy-momentum tensor and Maxwell’s equations. Owing to the screening effect, the electric fields within the fluid are gapped and parametrically small compared to the magnetic field. This renders the electric field inactive in the low-energy, long-wavelength regime. In contrast, there is no screening of the magnetic field, allowing it to exhibit its own dynamics even in this regime. Consequently, the magnetic field can be large and is treated as an O(1) quantity, despite the fact that B=×A involves one spatial gradient. The presence of an O(1) magnetic field breaks the SO(3) symmetry in the constitutive relations for Θμv, introducing anisotropy even in ideal hydrodynamics. Specifically, we can define a normalized vector bμ=Bμ/B, where B=BμBμ, satisfying b2=-1 and bu=0, as an additional building block for hydrodynamic constitutive relations. For example, for a partiy-even and charge neutral fluid, the energy-momentum tensor can be decomposed into Θμν=εuμuνPΞμν+Pbμbν+Θ(1)μν, (68) where Ξμν=Δμν+bνbν is a projector transverse to both and . The terms P and P represent the pressures in directions transverse and parallel to the magnetic field, respectively. Note that when we allow an environmental parity violation (e.g., when there is a density imbalance between the right- and left-hand particles in the fluid) and a finite charge density, the term u(μbν) can appear in the zeroth order. The term Θ(1)μν (which is assumed to be symmetric because the spin degree of freedom is typically disregarded in MHD) denotes a collection of terms that are at least of order O() in the gradient expansion and consistent with the Onsager relations. For a parity-even fluid, all such terms are expressed as Θ(1)μν=i=17λiηiμνρσρuσ, where λi is the corresponding transport coefficient [121-23]: η1μνρσ=bμbνbρbσ, (69a) η2μνρσ=ΞμνΞρσ, (69b) η3μνρσ=ΞμνbρbσΞρσbμbν, (69c) η4μνρσ=2[b(μΞν)ρbσ+b(μΞν)σbρ], (69d) η5μνρσ=2Ξρ(μΞν)σΞμνΞρσ, (69e) η6μνρσ=b(μbν)ρbσb(μbν)σbρ, (69f) η7μνρσ=Ξρ(μbν)σ+Ξσ(μbν)ρ, (69g) where bμν=ϵμνρσuρbσ is a cross projector that appears only when charge-conjugation symmetry is violated (e.g., when a net charge density is presented).

Similar to the discussions above regarding MHD, we can consider a scenario for spin hydrodynamics where the vorticity is treated as zeroth-order in gradients, while the gradients of other thermodynamic quantities are treated as first-order. In line with the MHD, this framework has been referred to as gyrohydrodynamics in Ref. [72]. To simplify the notation, we reuse to denote the unit vector along the vorticity, bμ=ϖμ/ϖμϖμ=ωμ/ωμωμ, (70) with ϖμ=ϵμνρσuνρβσ/2=βωμ the thermal vorticity vector. We chose the pesudo-gauge such that the spin tensor is totally antisymmetric. Using uμ,bμ as well as gμν,ϵμνρσ as building blocks, we can decompose Θμv and Σμνρ into the following irreducible forms: Θμν=εuμuνPΞμν+Pbμbν+P×bμν +qμuνuμqν+Θs(1)μν+ϕμν, (71) Σμνρ=ϵμνρλSλ=ϵμνρλ(n5uλSbλ+Sλ), (72) where P,× represent pressures (which will be counted as O(1) quantities in gradient expansion) in different directions, whose physical meaning will become clear shortly. The quantity S=bS denotes the spin component in the direction of the vorticity, while Sμ=ΞμνSν denotes the spin component transverse to the vorticity. As before, we chose the Landau-Lifshitz frame, with , Θs(1)μν, ϕμν, and Sμ transverse to . Note again that, with this fully antisymmetric choice of spin tensor, the vector is no longer independent, but is determined by through Eq. (64).

The power counting scheme is such that S is counted as order one, whereas ϕμν=ϕνμ, Sμ, n5, and (see Eq. (64)) are counted as at least O(). Additionally, we will count as O() (since spin is totally quantum in nature) in comparison to other thermodynamic quantities, which can appear even at the classical level and are therefore assigned O(0). This allows for a double expansion in both and . For the entropy current, we can write sμ=suμ+s(1)μ and use Eq. (33). It is straightforward to derive the divergence of the entropy current, and after some calculations, it was found that up to O(2,3) [72]: μsμ=[sβ(ε+P)]θ(PPμS)bμbνμβν+P×bμν(μβν+βμμν)+Θs(1)μν(μβν)+ϕμν([μβν]+βμμν)+μ(s(1)μβμμn5)+O(2,3). (73) The first line provides the zero-order contribution to the entropy production, which is expected to vanish so that they represent non-dissipative contributions. This gives ε+P=Ts,P=P+μS,P×=0. (74) The first relation is the Gibbs-Duhem relation, indicating that P can be interpreted as the thermodynamic pressure. The second relation shows that the pressure along the vorticity direction differs from the thermodynamic pressure by an amount due to spin polarization μS. This term is similar to the MB term in the magnetohydrodynamic constitutive relation. The third relation shows that there is no spin torque at the leading order.

At O() order, the requirement of a semi-positive entropy production gives that Θs(1)μν=Tημνρσ(ρβσ)+Tξμνρσ([ρβσ]+βμρσ), (75) ϕμν=Tγμνρσ([ρβσ]+βμρσ)+Tξ'μνρσ(ρβσ), (76) s(1)μ=βμμn5, (77) where ημνρσ and γμνρσ are the usual and rotational viscous tensors representing the response of the symmetric and antisymmetric parts of the energy-momentum tenor to fluid shear and expansion, and the difference between vorticity and spin potential, respectively, and ξμνρσ and ξμνρσ are two cross viscous tensors. Note that the cross viscous tensors are not independent of each other but inter-related according to Onsager’s reciprocal principle, ξ'μνρσ(b)=ξρ(b). By decomposing these tensors into irreducible structures, one obtains a number of new transport coefficients (viscosities) that characterize the response of the fluid to gradients of fluid velocity and spin potential [72]: ημνρσ=ζΞμνΞρσ+ζbμbνbρbσ+ζ×(bμbνΞρσ+Ξμνbρbσ)+η(ΞμρΞνσ+ΞμσΞνρΞμνΞρσ)+2η(bμΞν(ρbσ)+bνΞμ(ρbσ))+2ηH(Ξμ(ρbσ)ν+Ξν(ρbσ)μ)+2ηH(bμbν(ρbσ)+bνbμ(ρbσ)), (78) γμνρσ=γ(ΞμρΞνσΞμσΞνρ)+2γ(bμΞν[ρbσ]bνΞμ[ρbσ])+2γH(bμbν[ρbσ]bνbμ[ρbσ]), (79) ξμνρσ=2ξ(bμΞν[ρbσ]+bνΞμ[ρbσ])+ζHΞμνbρσ+ζHbμbνbρσ+2ξH(bμbν[ρbσ]+bνbμ[ρbσ]), (80) where the η’s, ζ’s, γ’s, and Ξ’s are transport coefficients. Especially, those with subscript “H” are Hall-type transport coefficients which do not contribute to the entropy production and thus their sign are not constrained by the second law of local thermodynamics. One may wonder why the term bμνbρσ (such term would contribute to an O() analog of P× term in Eq. (71)) does not appear in γμνρσ. This is because it is not independent of the other terms in γμνρσ [121]. Note that the expression for ξμνρσ is different from that in Ref. [72] but equivalently yields the same constitutive relations once substituted into Eq. (75).

4.3
A spin Cooper-Frye formula

To apply spin hydrodynamics to specific physical systems, we need to know the appropriate observables for the detection of spin degrees of freedom in the fluid. In principle, the presence of the spin degree of freedom in the fluid should modify the usual hydrodynamic quantities such as the energy density and fluid velocity, but when the spin density is not large (nevertheless it is always suppressed by comparing to the traditional hydrodynamic quantities), such modification is small. In heavy ion collisions, the natural observable is the spin polarization of hadrons, including spin-1/2 hyperons and spin-1 vector mesons. Hyperons are of special interest because they primarily decay via weak interactions such that the momentum of one of the daughter particles tends to align with the spin direction of the hyperon. To obtain the spin polarization observables of a hadron from spin hydrodynamics, machinery is required to convert the outcomes of spin hydrodynamics, such as fluid velocity, temperature, and spin potential, to measurable hadronic observables.

In the application of traditional hydrodynamics to heavy-ion collisions, the hadron momentum spectra are typically obtained using the Cooper-Frye formula: EpdNidp3=ΞdΞμ(x)pμfi(x,p), (81) where the integral is over the freeze-out hypersurface (where particlization occurs) Ξ, and fi(x,p) is the distribution function of species i of the hadrons in the fluid. Any possible degeneracy of the hadrons should be accounted for in fi. For example, when dissipative effects are neglected, the distribution function fi is typically taken as the Fermi-Dirac or Bose-Einstein functions fF,B(pβμi) with μi is the chemical potential. The above Cooper-Frye formula has been widely used in hydrodynamic simulations in heavy-ion collisions and has proven to be very successful. Therefore, to extend traditional hydrodynamics to spin hydrodynamics, we also need to generalize the above Cooper-Frye formula to a spin Cooper-Frye formula.

Let us consider a system in which thermal equilibrium is reached locally but not necessarily globally. The density operator ρ^ for the description of such an ensemble is obtained by maximizing the entropy functional under the constraints of the given energy-momentum and angular momentum (or spin) densities: S[ρ^]=Tr(ρ^lnρ^)+λ(Trρ^1)ΞdΞμ[Tr(ρ^Θ^μν)Θμν]βν  +12ΞdΞμ[Tr(ρ^Σ^μνρ)Σμνρ]μνρ, (82) where Θμν(x) and Σμνρ(x) are the actual local energy-momentum tensor and spin tensor, respectively, and βν(x) and μνρ(x) are the corresponding Lagrange multipliers. The Lagrange multiplier λ is introduced to normalize ρ^ and is related to the partition function Z as exp(1λ)=Z. The resultant density operator is the local-equilibrium density operator [124-127]: ρ^LE=1ZLEexp{ΞdΞμ(x)[Θ^μν(x)βν(x)12Σ^μρσ(x)μρσ(x)]}, (83) where ZLE denote the local-equilibrium partition function. Now, we see that ρ^LE is determined by the local thermodynamic quantities βμ and μρσ. If we calculate the spin density μρσ(x) using this density operator, we obtain a relationship between μρσ(x) and the local thermodynamic quantities (and possibly their derivatives). However, this is not particularly useful in the context of heavy-ion collisions because what is measured is the spin density in momentum space, rather than the coordinate space. To express such a relation for the spin density in momentum space or phase space, the most natural approach is to use the Wigner function.

To illustrate how this can be achieved, we consider a Dirac fermion system as an example. The Wigner operator is defined as follows: W^(x,p)=d4seipsψ^¯(x+s2)ψ^(xs2), (84) where [ψ^¯ψ^]abψ^¯bψ^a with a, b spinor indices. We choose the canonical pseudo-gauge, in which the energy-momentum tensor operator and spin tensor operator are given by Θ^μν=ψ^¯iγμνψ^ημνL^, (85) Σ^μνρ=14ψ^¯{γμ,σρσ}ψ^=12ϵμνρσψ^¯γσγ5ψ^. (86) where L^ is the Lagrangian (in the following, we consider free fermions, so that L=ψ¯(iγμμm)ψ is in quadratic form, and the second term in Θ^μν vanishes when using the equation of motion of the field operator) and σρσ=i[γρ,γσ]/2. Note that the second equation indicates that the spin vector S^σ=(1/2)ψ^¯γσγ5ψ^ is half the axial current. Both Θ^μν(x) and Σ^μνρ(x) are local Heisenberg operators. We can extend them into operators in the phase space by using the Wigner transformation, for example, Σ^μνρ(x,p)=12ϵμνρσd4seipsψ^¯(x+s2)γσγ5ψ^(xs2)=12ϵμνρσTrD[γσγ5W^(x,p)], (87) where TrD is the trace over the Dirac space. It is easy to see that d4p/(2π)4Σ^μνρ(x,p)=Σ^μνρ(x). The integration of Σ^μνρ(x,p) over a certain spacelike hypersurface gives us the spin tensor in momentum space (whose exact meaning will be clarified later), whose ensemble average under ρ^LE is exactly the quantity that we are looking for. Therefore, we must calculate the Wigner function under local equilibrium: W(x,p)=W^(x,p)=Tr[ρ^LEW^(x,p)], (88) where Tr denotes the trace over a complete set of microstates in the system. To proceed, the local-equilibrium density operator can be rewritten as ρ^LE=exp(A^+B^)/ZLE with the abbreviations A^=P^μβμ(x), (89) B^=dΞμ(y)[Θ^μν(y)Δβν(y)12Σ^μρσ(y)μρσ(y)], (90) where P^μ=dΞν(y)Θ^νμ(y), Δβμ(y)=βμ(y)βμ(x). The purpose of rewriting ρ^LE in this form is that, the correlation length between the spin tensor and the energy-momentum tensor is typically small. Within this correlation length, we can assume that local thermodynamic quantities, such as βμ, vary only slightly. Given that μρσ is also small at the hypersurface Ξ (which is a reasonable assumption for heavy-ion collisions, although it may not hold for a strongly polarized medium), we assign ΔβνμρσO(), therefore, A^O(1), B^O(). Using this power-counting scheme, we can expand the right-hand side of Eq. (88) order by order in by applying the identity eA^+B^=eA^+eA^01dλeλA^B^eλA^+, and obtain W(x,p)=W0(x,p)+W1(x,p)+, (91) where W0(x,p)=W^(x,p)01Z0Tr(eA^W^(x,p)), (92) W1(x,p)W^(x,p)(Θ)+W^(x,p)(Σ), (93) with W^(x,p)(Θ)01dλdΞρ(y)Δβν(y)Θ^ρν(yiλβ(x))W^(x,p)0,c,W^(x,p)(Σ)1201dλdΞν(y)μρσ(y)Σ^νρσ(yiλβ(x))W^(x,p)0,c, (94) and Z0=TreA^. Here, 0,c means the connected part of the correlation. The calculation then will depend on the shape of the hypersurface Ξ. For illustration, we consider Ξ to be the 3-space at some time t so that its normal direction is t^μ=(1,0). The calculation is then straightforward using the free field operator ψ^(x)=σ=121(2π)3/2d3k2Ek[uσ(k)eikxa^σ(k)+vσ(k)eikxb^σ(k)], (95) where Ek=k2+m2 and a^σ(k), b^σ(k) are annihilation operators for particles and antiparticles satisfying the anti-commutation relation {a^σ(k),a^σ(q)}={b^σ(k),b^σ(q)}=2Ekδσσδ3(kq) and the relation a^σ(k)a^σ(q)0=b^σ(k)b^σ(q)0=2Ekδσσδ3(kq)nF(kβ). In the following, we consider only the particle branch; the antiparticle branch is completely similar. The zeroth-order Wigner function can be easily obtained: W0(x,p)=2π(p+m)θ(p0)δ(p2m2)nF(pβ), which is spin independent: TrD[γμγ5W0(x,p)]=0.

The first-order Wigner function reads W^(x,p)(Θ/Σ)=2π01dλd3k2Ekd3q2Eqδ4(pq+k2)(γk+m)t^μI(Θ/Σ)μ(γq+m)×eλ(kq)β(x)nF(k)[1nF(q)], (96) where nF(p)=nF[β(x)p], I(Θ)μ=γμpν[λβν(x)]Δβλ[iqβδ3(qk)], and I(Σ)μ=14ϵμνρσγ5γνμρσδ3(qk) with Δμν=ημνt^μt^ν. To obtain this result, we have used dΞμ(y)(yx)αei(pq)(yx)=(2π)3t^μΔβαipβδ3(pq), (97) which is valid when Ξ is a 3-space. In heavy-ion collisions, the true freeze-out hypersurface Ξ is of course not a 3-space and thus correction due to the non-flatness of Ξ would appear; see discussions in Refs. [128, 129].

Using the first-order Wigner function in Eq. (96), the local-equilibrium spin vector in the phase space is directly obtained by finishing the trace over the Dirac space [130, 131]: Sμ(x,p)=4πδ(p2m2)θ(p0)nF(p)[1nF(p)]{14ϵμνρσpνμρσ+Σμνt^[(ξνλ+Δμνλ)pλ]}, (98) where Σμνt^=ϵμνρσpρt^σ/(2pt^), ξμν=(μβν) is the thermal shear tensor, and Δμμν=μμνϖμν is the difference between the spin potential and the thermal vorticity tensor.

With this spin vector in phase space, the spin vector per particle in momentum space is obtained by average over hypersurface Ξ [130, 131]: Sμ(p)=12dΞ(x)pTrD[γμγ5W(x,p)]dΞpTrD[W(x,p)]=dΞp{ϵμναβpνμαβ+4Σμνt^[pλ(ξνλ+Δμνλ)]}nF(1nF)8mdΞpnF, (99) where on the right-hand side is on-shell. This is a Cooper-Frye-type formula for the spin vector, which connects the momentum-space distribution of the mean spin vector of particles emitted from Ξ with the fluid properties characterized by μμν(x) and βμ(x) on Ξ. Thus, once these fluid variables are obtained from spin hydrodynamics, this spin Cooper-Frye formula allows us to convert them into the mean spin vector in momentum space, which is a directly measurable quantity.

We provide several comments before concluding this subsection. First, at local equilibrium, the thermal shear tensor can induce spin polarization, which has important implications for the spin polarization phenomena in heavy-ion collisions [128, 132-134]. Second, when the system is in global equilibrium, the spin potential is determined by the thermal vorticity and the thermal shear tensor ξμν vanishes. In this case, the spin Cooper-Frye formula is reduced to that obtained in Refs. [20-22]. Third, we did not include the effects of finite baryon chemical potential. Its inclusion is straightforward, with the modification that the distribution function nF(pβ)nF(pβα), where α=μ/T. Additionally, a new term 4dΞpμνt^να should be added to the numerator of Eq. (99), which is referred to as the spin Hall effect [135]. Fourth, formula (99) depends on the choice of pseudo-gauge [130, 131]. In particular, it is possible to completely eliminate the contributions of thermal shear by adopting appropriate pseudo-gauges. Therefore, when applying this formula to spin hydrodynamics, it is important to carefully choose a pseudo-gauge to maintain consistency.

5

Summary and outlooks

This article provides a pedagogical introduction to relativistic spin hydrodynamics. First, we demonstrate how one can derive a set of hydrodynamic equations from conservation equations based on the requirements of local thermodynamic laws, primarily the second law of thermodynamics. We then extended this framework to include the conservation of angular momentum, which leads to spin hydrodynamics. In the framework of spin hydrodynamics, the new (quasi-)hydrodynamic variable is spin density. Owing to spin-orbit coupling, the spin density is not a strict hydrodynamic variable but rather a quasi-hydrodynamic variable. It relaxes to a local equilibrium value determined by the local thermal vorticity through dissipative conversion of the spin and orbital angular momenta. We demonstrate how such dissipative processes are characterized by two new transport coefficients: one for boosting heat conductivity and the other for rotational viscosity.

We discuss several interesting aspects of spin hydrodynamics. First, we address the pseudo-gauge ambiguity in defining the spin tensor, which reflects the freedom to separate the total angular momentum into spin and orbital components. One consequence of this pseudo-gauge ambiguity is that we have the flexibility to choose spin tensors with different symmetries in their indices as the starting point for the derivation of spin hydrodynamics, leading to different constitutive relations. Second, we emphasize the importance of derivative power counting in the formulation of spin hydrodynamics. In particular, for a strongly vortical (or strongly spin-polarized) fluid, it is natural to assign the vorticity and spin potential as being of similar strength to other local thermodynamic quantities, such as temperature, in terms of derivative powers. This is analogous to the magnetohydrodynamics. As a result, anisotropy emerges in the constitutive relations both at the zeroth order and the first order in derivatives. This framework is well-suited for describing strongly vortical or spin-polarized fluids. Third, for potential applications of spin hydrodynamics, such as in heavy-ion collisions, we require a method to convert the results of spin hydrodynamics—specifically, the spin density (or spin potential), temperature, and fluid velocity—into momentum-space observables. To this end, we give a spin Cooper-Frye formula for Dirac fermions, and a similar formula can also be derived for spin-one vector bosons.

Spin hydrodynamics is an area of intensive study with many interesting aspects already explored and many more awaiting investigations. We provide a brief discussion of some of these topics.

(1) Spin magnetohydrodynamics. When the constituents of the fluid are charged, the fluid can interact with the electromagnetic fields and behave like a magnetized fluid. In this case, it is convenient to extend spin hydrodynamics to spin magnetohydrodynamics [136-140]. As electric fields are easily screened, they are not typically described as hydrodynamic variables. Therefore, the new hydrodynamic variable is the magnetic field (more precisely, the magnetic flux), Bμ=F˜μνuν, which is counted as an O(1) quantity in derivative power counting. The conservation law is simply a Bianchi identity. μF˜μν=0. (100) Here, F˜μν=(1/2)ϵμνρσFρσ is the dual Maxwell tensor. This equation should be combined with the conservation laws of energy-momentum and angular momentum to form complete equations of motion for the fluid. Expanding F˜μν in terms of hydrodynamic variables yields [120]: F˜μν=BμuνBνuμ+F˜(1)μν, (101) where F˜(1)μν and are transverse to . Local thermodynamic laws can be imposed, for example, the first law and a generalized Gibbs-Duhem relation, as follows: Tds+12μμνdSμν+HμdBμ=dε, (102) Ts+12μμνSμν+HμBμ=ε+P, (103) with Hμ the “magnetic potential” conjugate to the magnetic flux (physically, it can be interpreted as the in-medium magnetic field strength). The convariant form for the Gibbs-Duhem relation is sμ=Pβμ+Θμνβν12Σμρσαρσ+F˜μνγν, (104) with γμ=βHμ. The second law of thermodynamics requires μsμ0, which imposes constraints on the possible forms of the constitutive relations order by order in the gradient expansion. Recently, such a framework for spin magnetohydrodynamics was discussed (see Refs. [139, 140] for further detail).

It would be interesting to extend these studies to include possible parity-violating effects, thereby obtaining spin magnetohydrodynamics in a chiral conducting medium. This provides a bridge between spin magnetohydrodynamics and chiral magnetohydrodynamics. Another issue that may affect the formulation of spin magnetohydrodynamics is pseudo-gauge ambiguity. As we have seen, such an ambiguity is crucial for the formulation of spin hydrodynamics, and it would be interesting to explore how it influences the formulation of spin magnetohydrodynamics. Finally, exploring possible collective modes and instabilities in such a fluid is also important. This would be valuable for potential applications (e.g., possible dynamo mechanisms owing to spin degrees of freedom) in what we might call spin plasma, whether in heavy-ion collisions or astrophysical systems.

(2) Calculation of the new transport coefficients. As we have seen, new transport coefficients appear in spin hydrodynamics, most notably rotational viscosity ηs. Strictly speaking, ηs, unlike the typical shear viscosity η, is not a transport coefficient in the traditional sense. It does not characterize the ability to transport spin within the fluid; rather, it represents how quickly the spin density relaxes to its equilibrium value, which is determined by thermal vorticity. This can be easily understood by rewriting Eq. (8) in the canonical pseudo-gauge and in component form (keeping linear terms in spin density and velocity): tSiηs(μiϖi) where μi=ϵijkμik, which leads to tμi=Γs(μiϖi) with Γs=ηs/χs the spin relaxation rate and χs the spin susceptibility. Nevertheless, the calculation of Γs and ηs is important for understanding the evolution of spin polarization. Recently, Γs has been computed perturbatively for heavy quarks in hot QCD plasma [65, 67] and baryons in hot hadronic plasma [66]. Kinetic theory-based calculations have also been reported [67]. The results show that, for heavy quarks, this parameter can be parametrically small, making the spin degree of freedom a quasi-hydrodynamic mode. In the future, the calculation of other new transport coefficients, such as those arising in gyrohydrodynamics [72], could also be crucial for understanding spin dynamics in different fluids. In addition, it is important to examine and understand the pseudo-gauge dependence of these new transport coefficients.

(3) Simulation of spin hydrodynamics. It is important to develop a suitable numerical framework for performing simulations to apply spin hydrodynamics to heavy-ion collisions. First-order relativistic hydrodynamic equations are known to suffer from numerical instabilities and emergence of acausal modes. The origin of this problem lies in the fact that first-order constitutive relations are non-dynamical, meaning that the response of the fluid to thermodynamic forces is instantaneous. One solution to this problem is to make the constitutive relations more dynamic. For example, the constitutive relation for the shear channel can be modified as τπ(Dπ)μν+πμν=2ησμν, (105) where πμν is the traceless symmetric part of Θ(1)μν, (Dπ)μν(1/2)[ΔμρΔνσ+ΔμσΔνρ(2/3)ΔμνΔρσ]Dπρσ is the traceless part of the co-moving time derivative of πμν, and τπ represents how quickly πμν relaxes into the hydrodynamic constitutive relation. (Note that this procedure introduces a new dynamic mode that is not a hydrodynamic mode and relaxes on a timescale given by τπ.) The use of such a modification has been successful in the numerical simulation of relativistic hydrodynamics. For relativistic spin hydrodynamics, modifications similar to the constitutive relations may be adopted to implement numerical simulations. This has recently been discussed in Refs. [41, 74, 92, 93, 95, 141, 142]. Essentially, the constitutive relation Eq. (41) is replaced by a dynamic relation τϕ(Dϕ)μν+ϕμν=ηsΔμρΔνσ(μρσTϖρσ), (106) where τϕ is the relaxation time for the antisymmetric part of the energy-momentum tensor and (Dϕ)μνΔμρΔνσDϕρσ is the transverse part of the co-moving time derivative of ϕμν. With these modifications, a numerical simulation of relativistic spin hydrodynamics can be performed, which will provide valuable insights into spin polarization phenomena (see the recent progress in Refs. [143, 144]) such as those observed in heavy-ion collisions.

References
1.X.N. Wang (Ed.), Quark-Gluon Plasma 5, (World Scientific, New Jersey, 2016). https://doi.org/10.1142/9533
2.Y. Du, C. Li, C. Shi, et al.,

Review of QCD phase diagram analysis using effective field theories

. Nucl. Tech. (in Chinese) 46,. https://doi.org/10.11889/j.0253-3219.2023.hjs.46.040009
Baidu ScholarGoogle Scholar
3.Y. Zhang, D. Zhang, X. Luo,

Experimental study of the QCD phase diagram in relativistic heavy-ion collisions

. Nucl. Tech. (in Chinese) 46, 040001 (2023). https://doi.org/10.11889/j.0253-3219.2023.hjs.46.040001
Baidu ScholarGoogle Scholar
4.Q.Y. Shou, et al.,

Properties of QCD matter: a review of selected results from ALICE experiment

. Nucl. Sci. Tech. 35, 219 (2024). https://doi.org/10.1007/s41365-024-01583-2
Baidu ScholarGoogle Scholar
5.J. Chen, X. Dong, X.H. He, et al.,

Properties of the QCD matter: review of selected results from the relativistic heavy ion collider beam energy scan (RHIC BES) program

. Nucl. Sci. Tech. 35, 214 (2024). https://doi.org/10.1007/s41365-024-01591-2
Baidu ScholarGoogle Scholar
6.J.Y. Ollitrault,

Anisotropy as a signature of transverse collective flow

. Phys. Rev. D 46, 229-245 (1992). https://doi.org/10.1103/PhysRevD.46.229
Baidu ScholarGoogle Scholar
7.L. Adamczyk, et al.,

Global Λhyperon polarization in nuclear collisions: evidence for the most vortical fluid

. Nature 548, 62-65 (2017). https://doi.org/10.1038/nature23004
Baidu ScholarGoogle Scholar
8.Z.T. Liang, X.N. Wang,

Globally polarized quark-gluon plasma in non-central A+A collisions

. Phys. Rev. Lett. 94, 102301 (2005). [Erratum: Phys.Rev.Lett. 96, 039901 (2006)]. https://doi.org/10.1103/PhysRevLett.94.102301
Baidu ScholarGoogle Scholar
9.S.A. Voloshin,

Polarized secondary particles in unpolarized high energy hadron-hadron collisions?

. arXiv:nucl-th/0410089
Baidu ScholarGoogle Scholar
10.Z.T. Liang, X.N. Wang,

Spin alignment of vector mesons in non-central A+A collisions

. Phys. Lett. B 629, 20-26 (2005). https://doi.org/10.1016/j.physletb.2005.09.060
Baidu ScholarGoogle Scholar
11.J. Adam, et al.,

Global polarization of Λhyperons in Au+Au collisions at sNN−200 GeV

. Phys. Rev. C 98, 014910 (2018). https://doi.org/10.1103/PhysRevC.98.014910
Baidu ScholarGoogle Scholar
12.M.S. Abdallah, et al.,

Global Λ-hyperon polarization in Au+Au collisions at sNN=3 GeV

. Phys. Rev. C 104, L061901 (2021). https://doi.org/10.1103/PhysRevC.104.L061901
Baidu ScholarGoogle Scholar
13.R. Abou Yassine, et al.,

Measurement of global polarization ofΛhyperons in few-GeV heavy-ion collisions

. Phys. Lett. B 835, 137506 (2022). https://doi.org/10.1016/j.physletb.2022.137506
Baidu ScholarGoogle Scholar
14.M.I. Abdulhamid, B.E. Aboona, J. Adam, et al.,

Global polarization ofΛand Λ¯ hyperons in Au + Au collisions at sNN=19.6 and 27 GeV

. Phys. Rev. C 108, 014910 (2023). https://doi.org/10.1103/PhysRevC.108.014910
Baidu ScholarGoogle Scholar
15.M.S. Abdallah, et al.,

Pattern of global spin alignment of ϕ and K*0 mesons in heavy-ion collisions

. Nature 614, 244-248 (2023). https://doi.org/10.1038/s41586-022-05557-5
Baidu ScholarGoogle Scholar
16.S. Acharya, et al.,

Measurement of the J/ΨPolarization with Respect to the Event Plane in Pb-Pb Collisions at the LHC

. Phys. Rev. Lett. 131, 042303 (2023). https://doi.org/10.1103/PhysRevLett.131.042303
Baidu ScholarGoogle Scholar
17.W.T. Deng, X.G. Huang,

Vorticity in Heavy-Ion Collisions

. Phys. Rev. C 93, 064907 (2016). https://doi.org/10.1103/PhysRevC.93.064907
Baidu ScholarGoogle Scholar
18.Y. Jiang, Z.W. Lin, J. Liao,

Rotating quark-gluon plasma in relativistic heavy ion collisions

. Phys. Rev. C 94, 044910 (2016). [Erratum: Phys.Rev.C 95, 049904 (2017)]. https://doi.org/10.1103/PhysRevC.94.044910
Baidu ScholarGoogle Scholar
19.X.G. Deng, X.G. Huang, Y.G. Ma, et al.,

Vorticity in low-energy heavy-ion collisions

. Phys. Rev. C 101, 064908 (2020). https://doi.org/10.1103/PhysRevC.101.064908
Baidu ScholarGoogle Scholar
20.F. Becattini, V. Chandra, L. Del Zanna, et al.,

Relativistic distribution function for particles with spin at local thermodynamical equilibrium

. Annals Phys. 338, 32-49 (2013). https://doi.org/10.1016/j.aop.2013.07.004
Baidu ScholarGoogle Scholar
21.R.H. Fang, L.G. Pang, Q. Wang, et al.,

Polarization of massive fermions in a vortical fluid

. Phys. Rev. C 94, 024904 (2016). https://doi.org/10.1103/PhysRevC.94.024904
Baidu ScholarGoogle Scholar
22.Y.C. Liu, K. Mameda, X.G. Huang,

Covariant Spin Kinetic Theory I: Collisionless Limit

. Chin. Phys. C 44, 094101 (2020). [Erratum: Chin.Phys.C 45, 089001 (2021)]. https://doi.org/10.1088/1674-1137/ac009b
Baidu ScholarGoogle Scholar
23.I. Karpenko, F. Becattini,

Study ofΛpolarization in relativistic nuclear collisions at sNN=7.7−200 GeV

. Eur. Phys. J. C 77, 213 (2017). https://doi.org/10.1140/epjc/s10052-017-4765-1
Baidu ScholarGoogle Scholar
24.Y. Xie, D. Wang, L.P. Csernai,

GlobalΛpolarization in high energy collisions

. Phys. Rev. C 95, 031901 (2017). https://doi.org/10.1103/PhysRevC.95.031901
Baidu ScholarGoogle Scholar
25.H. Li, L.G. Pang, Q. Wang, et al.,

GlobalΛpolarization in heavy-ion collisions from a transport model

. Phys. Rev. C 96, 054908 (2017). https://doi.org/10.1103/PhysRevC.96.054908
Baidu ScholarGoogle Scholar
26.S. Shi, K. Li, J. Liao,

Searching for the Subatomic Swirls in the CuCu and CuAu Collisions

. Phys. Lett. B 788, 409-413 (2019). https://doi.org/10.1016/j.physletb.2018.09.066
Baidu ScholarGoogle Scholar
27.X.L. Xia, H. Li, Z.B. Tang, et al.,

Probing vorticity structure in heavy-ion collisions by localΛpolarization

. Phys. Rev. C 98, 024905 (2018). https://doi.org/10.1103/PhysRevC.98.024905
Baidu ScholarGoogle Scholar
28.D.X. Wei, W.T. Deng, X.G. Huang,

Thermal vorticity and spin polarization in heavy-ion collisions

. Phys. Rev. C 99, 014905 (2019). https://doi.org/10.1103/PhysRevC.99.014905
Baidu ScholarGoogle Scholar
29.O. Vitiuk, L.V. Bravina, E.E. Zabrodin,

Is differentΛand Λ¯ polarization caused by different spatio-temporal freeze-out picture?

Phys. Lett. B 803, 135298 (2020). https://doi.org/10.1016/j.physletb.2020.135298
Baidu ScholarGoogle Scholar
30.Y.B. Ivanov, V.D. Toneev, A.A. Soldatov,

Estimates of hyperon polarization in heavy-ion collisions at collision energies sNN=4−40 GeV

. Phys. Rev. C 100, 014908 (2019). https://doi.org/10.1103/PhysRevC.100.014908
Baidu ScholarGoogle Scholar
31.B. Fu, K. Xu, X.G. Huang, et al.,

Hydrodynamic study of hyperon spin polarization in relativistic heavy ion collisions

. Phys. Rev. C 103, 024903 (2021). https://doi.org/10.1103/PhysRevC.103.024903
Baidu ScholarGoogle Scholar
32.Y. Guo, J. Liao, E. Wang, et al.,

Hyperon polarization from the vortical fluid in low-energy nuclear collisions

. Phys. Rev. C 104, L041902 (2021). https://doi.org/10.1103/PhysRevC.104.L041902
Baidu ScholarGoogle Scholar
33.H. Li, X.L. Xia, X.G. Huang, et al.,

Global spin polarization of multistrange hyperons and feed-down effect in heavy-ion collisions

. Phys. Lett. B 827, 136971 (2022). https://doi.org/10.1016/j.physletb.2022.136971
Baidu ScholarGoogle Scholar
34.X.G. Deng, X.G. Huang, Y.G. Ma,

Lambda polarization in 108Ag+108Ag and 197Au+197Au collisions around a few GeV

. Phys. Lett. B 835, 137560 (2022). https://doi.org/10.1016/j.physletb.2022.137560
Baidu ScholarGoogle Scholar
35.X.Y. Wu, C. Yi, G.Y. Qin, et al.,

Local and global polarization ofΛhyperons across RHIC-BES energies: The roles of spin hall effect, initial condition, and baryon diffusion

. Phys. Rev. C 105, 064909 (2022). https://doi.org/10.1103/PhysRevC.105.064909
Baidu ScholarGoogle Scholar
36.Y. Hidaka, S. Pu, Q. Wang, et al.,

Foundations and applications of quantum kinetic theory

. Prog. Part. Nucl. Phys. 127, 103989 (2022). https://doi.org/10.1016/j.ppnp.2022.103989
Baidu ScholarGoogle Scholar
37.J.H. Gao, G.L. Ma, S. Pu, et al.,

Recent developments in chiral and spin polarization effects in heavy-ion collisions

. Nucl. Sci. Tech. 31, 90 (2020). https://doi.org/10.1007/s41365-020-00801-x
Baidu ScholarGoogle Scholar
38.F. Becattini, M.A. Lisa,

Polarization and Vorticity in the Quark–Gluon Plasma

. Ann. Rev. Nucl. Part. Sci. 70, 395-423 (2020). https://doi.org/10.1146/annurev-nucl-021920-095245
Baidu ScholarGoogle Scholar
39.X.G. Huang, J. Liao, Q. Wang, et al.,

Vorticity and Spin Polarization in Heavy Ion Collisions: Transport Models

. Lect. Notes Phys. 987, 281-308 (2021). https://doi.org/10.1007/978-3-030-71427-7_9
Baidu ScholarGoogle Scholar
40.X.G. Huang,

Vorticity and Spin Polarization — A Theoretical Perspective

. Nucl. Phys. A 1005, 121752 (2021). https://doi.org/10.1016/j.nuclphysa.2020.121752
Baidu ScholarGoogle Scholar
41.Y.C. Liu, X.G. Huang,

Anomalous chiral transports and spin polarization in heavy-ion collisions

. Nucl. Sci. Tech. 31, 56 (2020). https://doi.org/10.1007/s41365-020-00764-z
Baidu ScholarGoogle Scholar
42.F. Becattini,

Spin and polarization: a new direction in relativistic heavy ion physics

. Rept. Prog. Phys. 85, 122301 (2022). https://doi.org/10.1088/1361-6633/ac97a9
Baidu ScholarGoogle Scholar
43.X.L. Sheng, Z.T. Liang, Q. Wang,

Global spin alignment of vector mesons in heavy ion collisions

. Acta Phys. Sin. 72, 072502 (2023). https://doi.org/10.7498/aps.72.20230071
Baidu ScholarGoogle Scholar
44.J.H. Gao, X.G. Huang, Z.T. Liang, et al.,

Spin-orbital coupling in strong interaction and global spin polarization

. Acta Phys. Sin. 72, 072501 (2023). https://doi.org/10.7498/aps.72.20230102
Baidu ScholarGoogle Scholar
45.L.J. Ruan, Z.B. Xu, C. Yang,

Global polarization of hyperons and spin alignment of vector mesons in quark matters

. Acta Phys. Sin. 72, 112401 (2023). https://doi.org/10.7498/aps.72.20230496
Baidu ScholarGoogle Scholar
46.J.H. Gao, X.L. Sheng, Q. Wang, et al.,

Relativistic spin transport theory for spin-1/2 fermions

. Acta Phys. Sin. 72, 112501 (2023). https://doi.org/10.7498/aps.72.20222470
Baidu ScholarGoogle Scholar
47.F. Becattini, M. Buzzegoli, T. Niida, et al.,

Spin polarization in relativistic heavy-ion collisions

. Int. J. Mod. Phys. E 33, 2430006 (2024). https://doi.org/10.1142/S0218301324300066
Baidu ScholarGoogle Scholar
48.J.H. Chen, Z.T. Liang, Y.G. Ma, et al.,

Vector meson’s spin alignments in high energy reactions

. arXiv:2407.06480
Baidu ScholarGoogle Scholar
49.X. Sun, C.S. Zhou, J.H. Chen, et al.,

Measurements of global polarization of QCD matter in heavy-ion collisions

. Acta Phys. Sin. 72, 072401 (2023). https://doi.org/10.7498/aps.72.20222452
Baidu ScholarGoogle Scholar
50.Z. Ji, X.Y. Zhao, A.Q. Guo, et al.,

Lambda polarization at the Electron-ion collider in China

. Nucl. Sci. Tech. 34, 155 (2023). https://doi.org/10.1007/s41365-023-01317-w
Baidu ScholarGoogle Scholar
51.D.E. Kharzeev, J. Liao, S.A. Voloshin, et al.,

Chiral magnetic and vortical effects in high-energy nuclear collisions—A status report

. Prog. Part. Nucl. Phys. 88, 1-28 (2016). https://doi.org/10.1016/j.ppnp.2016.01.001
Baidu ScholarGoogle Scholar
52.X.G. Huang,

Electromagnetic fields and anomalous transports in heavy-ion collisions — A pedagogical review

. Rept. Prog. Phys. 79, 076302 (2016). https://doi.org/10.1088/0034-4885/79/7/076302
Baidu ScholarGoogle Scholar
53.K. Hattori, X.G. Huang,

Novel quantum phenomena induced by strong magnetic fields in heavy-ion collisions

. Nucl. Sci. Tech. 28, 26 (2017). https://doi.org/10.1007/s41365-016-0178-3
Baidu ScholarGoogle Scholar
54.D.E. Kharzeev, J. Liao,

Chiral magnetic effect reveals the topology of gauge fields in heavy-ion collisions

. Nature Rev. Phys. 3, 55-63 (2021). https://doi.org/10.1038/s42254-020-00254-6
Baidu ScholarGoogle Scholar
55.X.L. Zhao, G.L. Ma, Y.G. Ma,

Electromagnetic field effects and anomalous chiral phenomena in heavy-ion collisions at intermediate and high energy

. Acta Phys. Sin. 72, 112502 (2023). https://doi.org/10.7498/aps.72.20230245
Baidu ScholarGoogle Scholar
56.D.E. Kharzeev, J. Liao, P. Tribedy,

Chiral Magnetic Effect in Heavy Ion Collisions: The Present and Future

.. arXiv:2405.05427
Baidu ScholarGoogle Scholar
57.P. Romatschke,

New Developments in Relativistic Viscous Hydrodynamics

. Int. J. Mod. Phys. E 19, 1-53 (2010). https://doi.org/10.1142/S0218301310014613
Baidu ScholarGoogle Scholar
58.S. Jeon, U. Heinz,

Introduction to Hydrodynamics

. Int. J. Mod. Phys. E 24, 1530010 (2015). https://doi.org/10.1142/S0218301315300106
Baidu ScholarGoogle Scholar
59.L. Yan,

A flow paradigm in heavy-ion collisions

. Chin. Phys. C 42, 042001 (2018). https://doi.org/10.1088/1674-1137/42/4/042001
Baidu ScholarGoogle Scholar
60.W. Florkowski, M.P. Heller, M. Spalinski,

New theories of relativistic hydrodynamics in the LHC era

. Rept. Prog. Phys. 81, 046001 (2018). https://doi.org/10.1088/1361-6633/aaa091
Baidu ScholarGoogle Scholar
61.G.S. Rocha, D. Wagner, G.S. Denicol, et al.,

Theories of Relativistic Dissipative Fluid Dynamics

. Entropy 26, 189 (2024). https://doi.org/10.3390/e26030189
Baidu ScholarGoogle Scholar
62.M. Hongo, X.G. Huang, M. Kaminski, et al.,

Relativistic spin hydrodynamics with torsion and linear response theory for spin relaxation

. JHEP 11, 150 (2021). https://doi.org/10.1007/JHEP11(2021)150
Baidu ScholarGoogle Scholar
63.S. Grozdanov, A. Lucas, N. Poovuttikul,

Holography and hydrodynamics with weakly broken symmetries

. Phys. Rev. D 99, 086012 (2019). https://doi.org/10.1103/PhysRevD.99.086012
Baidu ScholarGoogle Scholar
64.M. Stephanov, Y. Yin,

Hydrodynamics with parametric slowing down and fluctuations near the critical point

. Phys. Rev. D 98, 036006 (2018). https://doi.org/10.1103/PhysRevD.98.036006
Baidu ScholarGoogle Scholar
65.M. Hongo, X.G. Huang, M. Kaminski, et al.,

Spin relaxation rate for heavy quarks in weakly coupled QCD plasma

. JHEP 08, 263 (2022). https://doi.org/10.1007/JHEP08(2022)263
Baidu ScholarGoogle Scholar
66.Y. Hidaka, M. Hongo, M.A. Stephanov, et al.,

Spin relaxation rate for baryons in a thermal pion gas

. Phys. Rev. C 109, 054909 (2024). https://doi.org/10.1103/PhysRevC.109.054909
Baidu ScholarGoogle Scholar
67.S. Li, H.U. Yee,

Quantum Kinetic Theory of Spin Polarization of Massive Quarks in Perturbative QCD: Leading Log

. Phys. Rev. D 100, 056022 (2019). https://doi.org/10.1103/PhysRevD.100.056022
Baidu ScholarGoogle Scholar
68.K. Hattori, M. Hongo, X.G. Huang, et al.,

Fate of spin polarization in a relativistic fluid: An entropy-current analysis

. Phys. Lett. B 795, 100-106 (2019). https://doi.org/10.1016/j.physletb.2019.05.040
Baidu ScholarGoogle Scholar
69.K. Fukushima, S. Pu,

Spin hydrodynamics and symmetric energy-momentum tensors – A current induced by the spin vorticity –

. Phys. Lett. B 817, 136346 (2021). https://doi.org/10.1016/j.physletb.2021.136346
Baidu ScholarGoogle Scholar
70.D. She, A. Huang, D. Hou, et al.,

Relativistic viscous hydrodynamics with angular momentum

. Sci. Bull. 67, 2265-2268 (2022). https://doi.org/10.1016/j.scib.2022.10.020
Baidu ScholarGoogle Scholar
71.A. Daher, A. Das, W. Florkowski, et al.,

Canonical and phenomenological formulations of spin hydrodynamics

. Phys. Rev. C 108, 024902 (2023). https://doi.org/10.1103/PhysRevC.108.024902
Baidu ScholarGoogle Scholar
72.Z. Cao, K. Hattori, M. Hongo, et al.,

Gyrohydrodynamics: Relativistic spinful fluid with strong vorticity

. PTEP 2022, 071D01 (2022). https://doi.org/10.1093/ptep/ptac091
Baidu ScholarGoogle Scholar
73.J. Hu,

Cross effects in spin hydrodynamics: Entropy analysis and statistical operator

. Phys. Rev. C 107, 024915 (2023). https://doi.org/10.1103/PhysRevC.107.024915
Baidu ScholarGoogle Scholar
74.R. Biswas, A. Daher, A. Das, et al.,

Relativistic second-order spin hydrodynamics: An entropy-current analysis

. Phys. Rev. D 108, 014024 (2023). https://doi.org/10.1103/PhysRevD.108.014024
Baidu ScholarGoogle Scholar
75.F. Becattini, A. Daher, X.L. Sheng,

Entropy current and entropy production in relativistic spin hydrodynamics

. Phys. Lett. B 850, 138533 (2024). https://doi.org/10.1016/j.physletb.2024.138533
Baidu ScholarGoogle Scholar
76.Z. Drogosz, W. Florkowski, M. Hontarenko,

Hybrid approach to perfect and dissipative spin hydrodynamics

.. arXiv:2408.03106
Baidu ScholarGoogle Scholar
77.S. Dey, A. Das,

Kubo formula for spin hydrodynamics: spin chemical potential as leading order in gradient expansion

.. arXiv:2410.04141
Baidu ScholarGoogle Scholar
78.L. Yang, L. Yan,

Relativistic spin hydrodynamics revisited with general rotation by entropy-current analysis

.. arXiv:2410.07583
Baidu ScholarGoogle Scholar
79.A.D. Gallegos, U. Gürsoy, A. Yarom,

Hydrodynamics of spin currents

. SciPost Phys. 11, 041 (2021). https://doi.org/10.21468/SciPostPhys.11.2.041
Baidu ScholarGoogle Scholar
80.A.D. Gallegos, U. Gursoy, A. Yarom,

Hydrodynamics, spin currents and torsion

. JHEP 05, 139 (2023). https://doi.org/10.1007/JHEP05(2023)139
Baidu ScholarGoogle Scholar
81.J. Hu,

Kubo formulae for first-order spin hydrodynamics

. Phys. Rev. D 103, 116015 (2021). https://doi.org/10.1103/PhysRevD.103.116015
Baidu ScholarGoogle Scholar
82.A. Tiwari, B.K. Patra,

Second-order spin hydrodynamics from Zubarev’s nonequilibrium statistical operator formalism

.. arXiv:2408.11514
Baidu ScholarGoogle Scholar
83.D. She, Y.W. Qiu, D. Hou,

Relativistic Second-order Spin Hydrodynamics: A Kubo-Type Formulation for the Quark-Gluon Plasma

.. arXiv:2410.15142
Baidu ScholarGoogle Scholar
84.W. Florkowski, B. Friman, A. Jaiswal, et al.,

Relativistic fluid dynamics with spin

. Phys. Rev. C 97, 041901 (2018). https://doi.org/10.1103/PhysRevC.97.041901
Baidu ScholarGoogle Scholar
85.W. Florkowski, A. Kumar, R. Ryblewski,

Relativistic hydrodynamics for spin-polarized fluids

. Prog. Part. Nucl. Phys. 108, 103709 (2019). https://doi.org/10.1016/j.ppnp.2019.07.001
Baidu ScholarGoogle Scholar
86.S. Bhadury, W. Florkowski, A. Jaiswal, et al.,

Relativistic dissipative spin dynamics in the relaxation time approximation

. Phys. Lett. B 814, 136096 (2021). https://doi.org/10.1016/j.physletb.2021.136096
Baidu ScholarGoogle Scholar
87.S. Shi, C. Gale, S. Jeon,

From chiral kinetic theory to relativistic viscous spin hydrodynamics

. Phys. Rev. C 103, 044906 (2021). https://doi.org/10.1103/PhysRevC.103.044906
Baidu ScholarGoogle Scholar
88.S. Bhadury, W. Florkowski, A. Jaiswal, et al.,

Dissipative Spin Dynamics in Relativistic Matter

. Phys. Rev. D 103, 014030 (2021). https://doi.org/10.1103/PhysRevD.103.014030
Baidu ScholarGoogle Scholar
89.H.H. Peng, J.J. Zhang, X.L. Sheng, et al.,

Ideal Spin Hydrodynamics from the Wigner Function Approach

. Chin. Phys. Lett. 38, 116701 (2021). https://doi.org/10.1088/0256-307X/38/11/116701
Baidu ScholarGoogle Scholar
90.J. Hu,

Relativistic first-order spin hydrodynamics via the Chapman-Enskog expansion

. Phys. Rev. D 105, 076009 (2022). https://doi.org/10.1103/PhysRevD.105.076009
Baidu ScholarGoogle Scholar
91.N. Weickgenannt, E. Speranza, X.l. Sheng, et al.,

Generating Spin Polarization from Vorticity through Nonlocal Collisions

. Phys. Rev. Lett. 127, 052301 (2021). https://doi.org/10.1103/PhysRevLett.127.052301
Baidu ScholarGoogle Scholar
92.N. Weickgenannt, D. Wagner, E. Speranza, et al.,

Relativistic second-order dissipative spin hydrodynamics from the method of moments

. Phys. Rev. D 106, 096014 (2022). https://doi.org/10.1103/PhysRevD.106.096014
Baidu ScholarGoogle Scholar
93.N. Weickgenannt, D. Wagner, E. Speranza, et al.,

Relativistic dissipative spin hydrodynamics from kinetic theory with a nonlocal collision term

. Phys. Rev. D 106, L091901 (2022). https://doi.org/10.1103/PhysRevD.106.L091901
Baidu ScholarGoogle Scholar
94.N. Weickgenannt,

Linearly stable and causal relativistic first-order spin hydrodynamics

. Phys. Rev. D 108, 076011 (2023). https://doi.org/10.1103/PhysRevD.108.076011
Baidu ScholarGoogle Scholar
95.D. Wagner,

Resummed spin hydrodynamics from quantum kinetic theory

. arXiv:2409.07143
Baidu ScholarGoogle Scholar
96.D. Montenegro, L. Tinti, G. Torrieri,

Ideal relativistic fluid limit for a medium with polarization

. Phys. Rev. D 96, 056012 (2017). [Addendum: Phys.Rev.D 96, 079901 (2017)]. https://doi.org/10.1103/PhysRevD.96.056012
Baidu ScholarGoogle Scholar
97.D. Montenegro, G. Torrieri,

Causality and dissipation in relativistic polarizable fluids

. Phys. Rev. D 100, 056011 (2019). https://doi.org/10.1103/PhysRevD.100.056011
Baidu ScholarGoogle Scholar
98.D. Montenegro, G. Torrieri,

Linear response theory and effective action of relativistic hydrodynamics with spin

. Phys. Rev. D 102, 036007 (2020). https://doi.org/10.1103/PhysRevD.102.036007
Baidu ScholarGoogle Scholar
99.S. Li, M.A. Stephanov, H.U. Yee,

Nondissipative Second-Order Transport, Spin, and Pseudogauge Transformations in Hydrodynamics

. Phys. Rev. Lett. 127, 082302 (2021). https://doi.org/10.1103/PhysRevLett.127.082302
Baidu ScholarGoogle Scholar
100.D.L. Wang, S. Fang, S. Pu,

Analytic solutions of relativistic dissipative spin hydrodynamics with Bjorken expansion

. Phys. Rev. D 104, 114043 (2021). https://doi.org/10.1103/PhysRevD.104.114043
Baidu ScholarGoogle Scholar
101.D.L. Wang, X.Q. Xie, S. Fang, et al.,

Analytic solutions of relativistic dissipative spin hydrodynamics with radial expansion in Gubser flow

. Phys. Rev. D 105, 114050 (2022). https://doi.org/10.1103/PhysRevD.105.114050
Baidu ScholarGoogle Scholar
102.G. Torrieri, D. Montenegro,

Linear response hydrodynamics of a relativistic dissipative fluid with spin

. Phys. Rev. D 107, 076010 (2023). https://doi.org/10.1103/PhysRevD.107.076010
Baidu ScholarGoogle Scholar
103.J. Hu,

Linear mode analysis from spin transport equation

. Phys. Rev. D 106, 036004 (2022). https://doi.org/10.1103/PhysRevD.106.036004
Baidu ScholarGoogle Scholar
104.J. Hu,

Linear mode analysis and spin relaxation

. Phys. Rev. D 105, 096021 (2022). https://doi.org/10.1103/PhysRevD.105.096021
Baidu ScholarGoogle Scholar
105.G. Sarwar, M. Hasanujjaman, J.R. Bhatt, et al.,

Causality and stability of relativistic spin hydrodynamics

. Phys. Rev. D 107, 054031 (2023). https://doi.org/10.1103/PhysRevD.107.054031
Baidu ScholarGoogle Scholar
106.A. Daher, A. Das, R. Ryblewski,

Stability studies of first-order spin-hydrodynamic frameworks

. Phys. Rev. D 107, 054043 (2023). https://doi.org/10.1103/PhysRevD.107.054043
Baidu ScholarGoogle Scholar
107.S. Pu, X.G. Huang,

Relativistic spin hydrodynamics

. Acta Phys. Sin. 72, 071202 (2023). https://doi.org/10.7498/aps.72.20230036
Baidu ScholarGoogle Scholar
108.D. Wagner, M. Shokri, D.H. Rischke,

Damping of spin waves

. Phys. Rev. Res. 6, 043103 (2024). https://doi.org/10.1103/PhysRevResearch.6.043103
Baidu ScholarGoogle Scholar
109.X. Ren, C. Yang, D.L. Wang, et al.,

Thermodynamic stability in relativistic viscous and spin hydrodynamics

. Phys. Rev. D 110, 034010 (2024). https://doi.org/10.1103/PhysRevD.110.034010
Baidu ScholarGoogle Scholar
110.D.L. Wang, L. Yan, S. Pu,

Late-time asymptotic solutions, attractor, and focusing behavior of spin hydrodynamics

. arXiv:2408.03781
Baidu ScholarGoogle Scholar
111.F.J. Belinfante,

On the current and the density of the electric charge, the energy, the linear momentum and the angular momentum of arbitrary fields

. Physica 7, 449-474 (1940). https://doi.org/10.1016/S0031-8914(40)90091-X
Baidu ScholarGoogle Scholar
112.L. Rosenfeld,

On energy-momentum tensror

. Memoires Acad. Roy. De Belgique 18, 1-30 (1940).
Baidu ScholarGoogle Scholar
113.F.W. Hehl, P. Von Der Heyde, G.D. Kerlick, et al.,

General Relativity with Spin and Torsion: Foundations and Prospects

. Rev. Mod. Phys. 48, 393-416 (1976). https://doi.org/10.1103/RevModPhys.48.393
Baidu ScholarGoogle Scholar
114.F. Becattini, W. Florkowski, E. Speranza,

Spin tensor and its role in non-equilibrium thermodynamics

. Phys. Lett. B 789, 419-425 (2019). https://doi.org/10.1016/j.physletb.2018.12.016
Baidu ScholarGoogle Scholar
115.E. Speranza, N. Weickgenannt,

Spin tensor and pseudo-gauges: from nuclear collisions to gravitational physics

. Eur. Phys. J. A 57, 155 (2021). https://doi.org/10.1140/epja/s10050-021-00455-2
Baidu ScholarGoogle Scholar
116.S. Bhadury, J. Bhatt, A. Jaiswal, et al.,

New developments in relativistic fluid dynamics with spin

. Eur. Phys. J. ST 230, 655-672 (2021). https://doi.org/10.1140/epjs/s11734-021-00020-4
Baidu ScholarGoogle Scholar
117.N. Weickgenannt, D. Wagner, E. Speranza,

Pseudogauges and relativistic spin hydrodynamics for interacting Dirac and Proca fields

. Phys. Rev. D 105, 116026 (2022). https://doi.org/10.1103/PhysRevD.105.116026
Baidu ScholarGoogle Scholar
118.R. Singh,

On the non-uniqueness of the energy-momentum and spin currents

. arXiv:2406.02127
Baidu ScholarGoogle Scholar
119.M. Buzzegoli, A. Palermo,

Emergent canonical spin tensor in the chiral symmetric hot QCD

.. arXiv:2407.14345
Baidu ScholarGoogle Scholar
120.K. Hattori, M. Hongo, X.G. Huang,

New Developments in Relativistic Magnetohydrodynamics

. Symmetry 14, 1851 (2022). https://doi.org/10.3390/sym14091851
Baidu ScholarGoogle Scholar
121.X.G. Huang, A. Sedrakian, D.H. Rischke,

Kubo formulae for relativistic fluids in strong magnetic fields

. Annals Phys. 326, 3075-3094 (2011). https://doi.org/10.1016/j.aop.2011.08.001
Baidu ScholarGoogle Scholar
122.S. Grozdanov, D.M. Hofman, N. Iqbal,

Generalized global symmetries and dissipative magnetohydrodynamics

. Phys. Rev. D 95, 096003 (2017). arXiv:1610.07392, https://doi.org/10.1103/PhysRevD.95.096003
Baidu ScholarGoogle Scholar
123.J. Hernandez, P. Kovtun,

Relativistic magnetohydrodynamics

. JHEP 05, 001 (2017). https://doi.org/10.1007/JHEP05(2017)001
Baidu ScholarGoogle Scholar
124.D.N. Zubarev, A.V. Prozorkevich, S.A. Smolyanskii,

Derivation of nonlinear generalized equations of quantum relativistic hydrodynamics

. Theor. Math. Phys. 40, 821-831 (1979). https://doi.org/10.1007/BF01032069
Baidu ScholarGoogle Scholar
125.C.G. van Weert,

Maximum entropy principle and relativistic hydrodynamics

. Annals of Physics 140, 133-162 (1982). https://doi.org/10.1016/0003-4916(82)90338-4
Baidu ScholarGoogle Scholar
126.F. Becattini, L. Bucciantini, E. Grossi, et al.,

Local thermodynamical equilibrium and the beta frame for a quantum relativistic fluid

. Eur. Phys. J. C 75, 191 (2015). https://doi.org/10.1140/epjc/s10052-015-3384-y
Baidu ScholarGoogle Scholar
127.F. Becattini, M. Buzzegoli, E. Grossi,

Reworking the Zubarev’s approach to non-equilibrium quantum statistical mechanics

. Particles 2, 197-207 (2019). https://doi.org/10.3390/particles2020014
Baidu ScholarGoogle Scholar
128.F. Becattini, M. Buzzegoli, A. Palermo,

Spin-thermal shear coupling in a relativistic fluid

. Phys. Lett. B 820, 136519 (2021). https://doi.org/10.1016/j.physletb.2021.136519
Baidu ScholarGoogle Scholar
129.X.L. Sheng, F. Becattini, X.G. Huang, et al.,

Spin polarization of fermions at local equilibrium: Second-order gradient expansion

. arXiv:2407.12130
Baidu ScholarGoogle Scholar
130.Y.C. Liu, X.G. Huang,

Spin polarization formula for Dirac fermions at local equilibrium

. Sci. China Phys. Mech. Astron. 65, 272011 (2022). https://doi.org/10.1007/s11433-022-1903-8
Baidu ScholarGoogle Scholar
131.M. Buzzegoli,

Pseudogauge dependence of the spin polarization and of the axial vortical effect

. Phys. Rev. C 105, 044907 (2022). https://doi.org/10.1103/PhysRevC.105.044907
Baidu ScholarGoogle Scholar
132.S.Y.F. Liu, Y. Yin,

Spin polarization induced by the hydrodynamic gradients

. JHEP 07, 188 (2021). https://doi.org/10.1007/JHEP07(2021)188
Baidu ScholarGoogle Scholar
133.F. Becattini, M. Buzzegoli, G. Inghirami, et al.,

Local Polarization and Isothermal Local Equilibrium in Relativistic Heavy Ion Collisions

. Phys. Rev. Lett. 127, 272302 (2021). https://doi.org/10.1103/PhysRevLett.127.272302
Baidu ScholarGoogle Scholar
134.B. Fu, S.Y.F. Liu, L. Pang, et al.,

Shear-Induced Spin Polarization in Heavy-Ion Collisions

. Phys. Rev. Lett. 127, 142301 (2021). arXiv:2103.10403, https://doi.org/10.1103/PhysRevLett.127.142301
Baidu ScholarGoogle Scholar
135.S.Y.F. Liu, Y. Yin,

Spin Hall effect in heavy-ion collisions

. Phys. Rev. D 104, 054043 (2021). https://doi.org/10.1103/PhysRevD.104.054043
Baidu ScholarGoogle Scholar
136.R. Singh, M. Shokri, S.M.A.T. Mehr,

Relativistic hydrodynamics with spin in the presence of electromagnetic fields

. Nucl. Phys. A 1035, 122656 (2023). https://doi.org/10.1016/j.nuclphysa.2023.122656
Baidu ScholarGoogle Scholar
137.S. Bhadury, W. Florkowski, A. Jaiswal, et al.,

Relativistic Spin Magnetohydrodynamics

. Phys. Rev. Lett. 129, 192301 (2022). https://doi.org/10.1103/PhysRevLett.129.192301
Baidu ScholarGoogle Scholar
138.M. Kiamari, N. Sadooghi, M.S. Jafari,

Relativistic magnetohydrodynamics of a spinful and vortical fluid: Entropy current analysis

. Phys. Rev. D 109, 036024 (2024). https://doi.org/10.1103/PhysRevD.109.036024
Baidu ScholarGoogle Scholar
139.Z. Fang, K. Hattori, J. Hu,

Anisotropic linear waves and breakdown of the momentum expansion in spin magnetohydrodynamics

. arXiv:2409.07096
Baidu ScholarGoogle Scholar
140.Z. Fang, K. Hattori, J. Hu,

First-order spin magnetohydrodynamics

. 2024. arXiv:2410.00721, https://doi.org/10.1016/j.jspc.2024.100003
Baidu ScholarGoogle Scholar
141.X.Q. Xie, D.L. Wang, C. Yang, et al.,

Causality and stability analysis for the minimal causal spin hydrodynamics

. Phys. Rev. D 108, 094031 (2023). https://doi.org/10.1103/PhysRevD.108.094031
Baidu ScholarGoogle Scholar
142.A. Daher, W. Florkowski, R. Ryblewski, et al.,

Stability and causality of rest frame modes in second-order spin hydrodynamics

. Phys. Rev. D 109, 114001 (2024). https://doi.org/10.1103/PhysRevD.109.114001
Baidu ScholarGoogle Scholar
143.S.K. Singh, R. Ryblewski, W. Florkowski,

Spin dynamics with realistic hydrodynamic background for relativistic heavy-ion collisions

. Phys. Rev. C 111, 024907 (2025). https://doi.org/10.1103/PhysRevC.111.024907
Baidu ScholarGoogle Scholar
144.Sapna , S.K. Singh, D. Wagner,

Spin Polarization ofΛhyperons from Dissipative Spin Hydrodynamics

. arXiv:2503.22552
Baidu ScholarGoogle Scholar
Footnote
1

The spin alignment of a vector meson is quantified by the deviation of ρ00 from 1/3, where ρ00 is the 00-component of the vector meson’s spin density matrix.

2

One can start without including the co-moving time-derivative terms as those terms are eventually replaced by the spatial gradients up on using leading-order hydrodynamic EOMs. But we keep them to make the discussions more transparent.

3

Since Sρσ is counted as O() quantities, the term Sρσuμ is unchanged at O() under a re-definition of uμuμ+δuμ with δuμO(). Therefore, Eq. (35) is automatically satisfied at O() up on using the zeroth-order EOM for [68]. But when there appear other conserved charges, such as a global U(1) charge, Eq. (35) is a proposal to fix the local rest frame of the fluid.

4

This can be checked by noting that for Aλμν=Aμλν we have dΞμλAλμν=dΞμλAλμν+dΞnμnλnAλμν=dΞμλAλμν with nμ the norm of Ξ and λ=λnλn. Then one can use the Gauss theorem to transform it to an integral over the boundary of Ξ.