• Aucun résultat trouvé

Helical vortical effects, helical waves, and anomalies of Dirac fermions

N/A
N/A
Protected

Academic year: 2021

Partager "Helical vortical effects, helical waves, and anomalies of Dirac fermions"

Copied!
14
0
0

Texte intégral

(1)Vortical effects in Dirac fluids with vector, chiral and helical charges Victor E. Ambrus, M.N. Chernodub. To cite this version: Victor E. Ambrus, M.N. Chernodub. Vortical effects in Dirac fluids with vector, chiral and helical charges. 2020. �hal-02447887�. HAL Id: hal-02447887 https://hal.archives-ouvertes.fr/hal-02447887 Preprint submitted on 9 Nov 2020. HAL is a multi-disciplinary open access archive for the deposit and dissemination of scientific research documents, whether they are published or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés..

(2) Helical vortical effects, helical waves, and anomalies of Dirac fermions Victor E. Ambrus, 1 and M. N. Chernodub2, 3, 4 1. arXiv:1912.11034v1 [hep-th] 23 Dec 2019. Department of Physics, West University of Timis, oara, Bd. Vasile Pârvan 4, Timis, oara 300223, Romania 2 Institut Denis Poisson, Université de Tours, Tours 37200, France 3 Laboratory of Physics of Living Matter, Far Eastern Federal University, Sukhanova 8, Vladivostok, 690950, Russia 4 Corresponding author (maxim.chernodub@lmpt.univ-tours.fr). (Dated: December 24, 2019) Helicity is a classically conserved quantity which can be used, in addition to and independently of the (vector) charge and chirality, to characterize thermodynamic ensembles of massless Dirac fermions. We identify a symmetry of the Dirac Lagrangian which is responsible, via the Noether theorem, to the classical conservation of the helicity current. We demonstrate the existence of new nondissipative transport phenomena, helical vortical effects, that emerge in a helically-imbalanced rotating fermionic system. These phenomena lead to appearance of a new gapless hydrodynamic excitation, the helical vortical wave. Our results also imply that the helical symmetry suffers from a quantum anomaly. We conjecture the existence of a new type of triangle anomalies in QED which involve the helicity currents in addition to the standard vector and axial currents.. Introduction. Massless or nearly-massless fermions appear in many areas of physics, including theories of fundamental interactions, cosmological models of early Universe, ultra-hot relativistic plasmas, and superfluids, to mention a few [1]. Many relativistic phenomena are now available for experimental verification in recently discovered crystals of Dirac and Weyl semimetals, where the massless fermions appear as quasiparticle excitations [2]. The most important properties of these excitations are usually associated with their vector (gauge) and axial (chiral) symmetries that affect, in the case of semimetals, electromagnetic [3], thermal [4], and elastic [5] responses of these materials. Many unusual features of these semimetals are associated with the quantum anomaly that break the continuous axial symmetry of an underlying classical theory [6]. Similar anomalies lead to exotic transport phenomena of quarks, mediated by the topology of evolving gluon fields in expanding quark-gluon plasma of heavy-ion collisions [7]. In our paper, we remind that, in addition to the vector and axial charges, there is also a third, well-known, and, simultaneously, often-forgotten quantity that characterizes massless fermions: the helicity. The fermionic helicity is sometimes confused with the chirality, even though these quantities reflect different physical properties of fermions [8]. To highlight the importance of helicity, we demonstrate the existence of a set of new transport phenomena emerging in a gas of rotating massless fermions, the Helical Vortical Effects (HVE), that differ substantially from their chiral counterparts, the Chiral Vortical Effects [9–14]. The HVE may see its applications in noncentral ultrarelativistic heavy-ion collisions that create a nearly perfect fluid of quark-gluon plasma, the most vortical fluid ever known [15]. Hydrodynamics of relativistic plasmas with nonzero vorticity has attracted significant attention recently [16–20]. We uncover a new gapless hydrodynamic excitation, the Helical Vortical Wave (HVW) which is analogous to the Chiral Vortical Wave. (CVW) [16]. In the existing low-density quark-gluon plasmas, the HVW propagates faster than CVW, being able to proliferate even in a globally-neutral plasma. Vector-axial-helical triad for massless Dirac fermions. We consider one species of free massless Dirac fermions in a flat (3+1) dimensional Minkowski spacetime, described by the Lagrangian L=. i (ψγ µ ∂µ ψ − ∂µ ψγ µ ψ), 2. (1). where ψ = ψ † γ 0 is the Dirac adjoint of the 4-component spinor ψ and the units ~ = c = 1 are used. The 4 × 4 gamma matrices γ µ are taken, for definiteness, in the Dirac representation, with µ = 0, . . . , 3. The classical Dirac Lagrangian (1) is invariant under the action of the (global) vector symmetry group: U (1)V :. ψ → eiαV ψ,. ψ̄ → e−iαV ψ̄.. (2). According to the Noether theorem, the symmetry (2) implies the existence of the classically conserved vector current, JVµ = ψγ µ ψ, with ∂µ JVµ = 0. If the fermions would carry an electric charge e and were coupled to an electromagnetic field Aµ , then the global symmetry (2) would become a local (gauge) symmetry, αV = αV (x). The coupling to electromagnetism is achieved by adding a source term with the electric current eJVµ to the Lagrangian (1): L → L + eJVµ Aµ . The symmetry (2) is unbroken at the quantum level, thus reflecting the fundamental property of the electric charge conservation in quantum theory. The Lagrangian (1) possesses also the axial symmetry: U (1)A :. 5. ψ → eiαA γ ψ,. 5. ψ̄ → ψ̄eiαA γ ,. (3). where γ 5 = iγ 0 γ 1 γ 2 γ 3 is the fifth gamma matrix. The Noether theorem gives us the classically conserved axial µ (sometimes called “chiral”) current JA = ψγ µ γ 5 ψ. The axial symmetry (3) is broken at the quantum level, leading to nonconservation of the axial current in the presence.

(3) 2 of the electric E and magnetic B background fields [6]: µ ∂µ JA =. e2 e2 Fµν Feµν ≡ − 2 E · B, 2 16π 4π. (4). where Feµν = 12 εµναβ Fαβ , Fµν = ∂µ Aν −∂ν Aµ and we use √ the convention ε0123 = −g for the Levi-Civita tensor. The axial charge (chirality) χ is identified with respect to the eigenstates of the γ 5 matrix, γ 5 ψ = χψ. Due to the property (γ 5 )2 = 1, one distinguishes the right-handed (R) and left-handed (L) chiral eigenstates, respectively: γ 5 ψR = +ψR ,. γ 5 ψL = −ψL .. (5). µ µ The axial current JA = JR − JLµ represents the difference µ in the currents of the right-chiral, JR = ψ̄R γ µ ψR , and µ left-chiral, JL = ψ̄L γ µ ψL , Dirac fermions. The chirality χ of a fermion state is intimately related to the helicity λ of the same state. Classically, the helicity is given by the projection of the spin s on the direction of motion of the fermion given by its momentum p. At the quantum level, the helicity λ is an eigenvalue, hψ = λψ, of the helicity operator:. h=. s·p 1 γ·p ≡ γ5γ0 , p 2 p. (6). where p = |p| is the absolute value of the momentum operator p = −i∂, and si = 12 ε0ijk Σjk is the spin operator which is a part of the covariant tensor Σµν = 4i [γ µ , γ ν ]. Since the fermion is a spin 1/2 particle, the helicity takes two values, λ = ±1/2. One distinguishes the right-handed (↑) and the left-handed (↓) helicity eigenstates [21]: 2hψ↑ = +ψ↑ ,. 2hψ↓ = −ψ↓ .. (7). The chirality and helicity are different quantities. The chirality of a particle is equal to its helicity (for example, a right-chiral particle has a right-handed helicity) while the chirality of an antiparticle is opposite to its helicity (for instance, a right-chiral antiparticle has a lefthanded helicity). For a single fermion the vector (particle/antiparticle) and the axial (right-/left-chiral) charges determine rigidly the helicity of this single fermion. We will see later that for an ensemble of particles a similar relation would not work: the total helicity of the ensemble is not determined by its total vector and axial charges. The chirality operator γ 5 commutes with the helicity operator h given in Eq. (6), while both of them commute with the Dirac Hamiltonian corresponding to the Lagrangian of massless fermions (1): b = α · p, H. (8). 0. where α ≡ γ γ in the original Dirac notations. These commutation relations, [γ 5 , h] = 0,. b γ 5 ] = 0, [H,. b h] = 0, [H,. (9). indicate that all three operators possess the same eigensystem. The last relation implies that the helicity, similarly to the chirality, is a classically conserved number.. What is the symmetry of the Dirac Lagrangian (1) that would lead – via the Noether theorem – to a classical conservation of the helical charge that, in turn, is suggested by the last commutation relation of Eq. (9)? The Lagrangian (1) is invariant under a “helical” symmetry: U (1)H :. ψ → e2iαH h ψ,. ψ̄ → ψ̄e−2iαH h , (10). where the helicity operator h is given in Eq. (6). One may readily check that the helical symmetry (10) leads to µ the classically conserved helicity current: JH = 2ψγ µ hψ. We introduced here the normalization factor 2 for the helicity current in order to enforce the integer eigenvalue 2λ = ±1 for the helicity (7), similarly to the chirality (5). Thus, the vector-axial-helical densities     Z 1 QV  QA  = d3 x ψγ 0 γ 5  ψ. (11) QH 2h form a “triad” of the classically conserved – via the Noether theorem – U (1) quantities of the massless Dirac fermions described by the Lagrangian (1).[22] The behavior of the charges Q` and the associated currents J` of all three quantities ` = V, A, H under the C, P , and T inversions are shown in Table I. QV QA C − + P + − T + +. QH JV − − − − + −. JA + + −. JH − + −. Ω + + −. TABLE I. Parities of the vector (V ), axial (A), and helical (H) charges (Q) and currents (J ) of a massless Dirac fermion, as well as the vorticity Ω, under the C-, P -, and T -inversions. The signs +/− denote the even/odd parities.. A complete set of mode solutions of the Dirac equation, iγ µ ∂µ ψ = 0, includes the particle modes Uj and the corresponding antiparticle modes Vj = iγ 2 Uj∗ , where j is a cumulative label that indexes the eigenmodes. The commutation relations (9) imply that these modes are eigenfunctions of the Hamiltonian, chirality and the helicity operators, simultaneously: b j =Ej Uj , HU 5. γ Uj =χj Uj , hUj =λj Uj ,. b j = − Ej V j , HV 5. γ Vj = − χj Vj , hVj =λj Vj ,. (12a) (12b) (12c). where λj = ±1/2 and χj = ±1, while Ej is the mode energy, satisfying |Ej | = pj (for future convenience, we allow Ej to take negative values). Using Eq. (6) and the Dirac equation for massless particles, the relationship between λj and χj can be readily established: 1 2hUj = γ 5 HUj p. ⇒. λj =. χj Ej . 2|Ej |. (13).

(4) 3 The field operator is right-handed chirality (R). left-handed chirality (L). E. R. left-handed helicity (↓). R. L. j. 𝜇R. spins directed outwards. where canonical anticommutation rules for particle b̂j and antiparticle dˆj operators are assumed. The operators of all three conserved charges (11) are: b V := P (b̂† b̂j − dˆ† dˆj ), :Q (15a) j j j P † ˆ† ˆ b A := :Q (15b) j χj (b̂j b̂j + dj dj ), P † † ˆ ˆ b H := :Q (15c) j 2λj (b̂j b̂j − dj dj ),. R. 𝜇L. L. L. right-handed helicity (↑). p. (a) right-handed helicity (↑). where the colons denote Wick (normal) ordering. The particle and antiparticle states contribute differently to the axial (15b) and helicity (15c) charges. Equations (12) and (15) also imply that the helicities and chiralities are indeed equal (opposite) to each other for particle (antiparticle) modes. Helical chemical potential. Similarly to the vector (15a) and axial (15b) charges, the existence of the third classically conserved number, helicity (15c), requires the introduction of the appropriate thermodynamically conjugated, “helical”, chemical potential µH . Let us recall that a right-chirality particle (antiparticle) has a right-handed (left-handed) helicity with spins parallel (anti-parallel) to its direction of motion. An ensemble of an equal number of right-helicity particles (that all have a right-handed chirality) and right-helicity antiparticles (characterized by a left-handed chirality) has a zero global vector (electric) charge density, hQV i = 0, and a zero total axial (chiral) charge, hQA i = 0, while the global helicity charge of this ensemble is nonzero, hQH i = 6 0. To describe such helicity-imbalanced, but otherwise neutral systems of Dirac fermions, we introduce the helical chemical potential µH , as the use of vector, µV , and axial, µA , chemical potentials is not enough. Formally, the helicity operator (6) is an ambiguous, Lorentz-frame-dependent quantity. However, we consider the Dirac systems at finite temperature T 6= 0 and/or in the presence of a finite helicity with µH 6= 0 that explicitly break the Lorentz invariance. Therefore, our physical environment sets a natural Lorentz frame to define the helicity operator and removes the mentioned ambiguity. It is customary to label the massless Dirac eigenmodes by their chiralities. One usually distinguishes the righthanded and left-handed “chiral Weyl cones” with the appropriate chemical potentials µR and µL , respectively, shown in Fig. 1(a). The difference in the occupation numbers between the modes is controlled by the axial (chiral) chemical potential, µA = (µR − µL )/2. Since the helicity operator shares the same basis with the chirality, we can also use the helicity to characterize the energy branches in an independent, and nonequivalent, manner, Fig. 1(b). Now, the occupation numbers of the “helical Weyl cones” are labelled by the righthelical (µ↑ ) and the left-helical (µ↓ ) chemical potentials.. spins directed inwards. R. left-handed helicity (↓). L. anti-particles. (14). particles. right-handed helicity (↑). left-handed helicity (↓). right-handed chirality (R). E. R. left-handed chirality (L). R. 𝜇↑. L. L. R. R. 𝜇↓. L. left-handed chirality (L). L. particles. X [Uj (x)b̂j + Vj (x)dˆ†j ],. anti-particles. b ψ(x) =. right-handed chirality (R). p. (b) FIG. 1. The dispersion relations Ep = ±|p| for Dirac fermions in (a) chiral and (b) helical basis. The chemical potentials µR/L determine the occupation numbers for the right-/lefthanded chiralities, while the chemical potentials µ↑/↓ dictate the densities of the right-/left-handed helicities. The spin orientations s with respect to the fermion’s momentum p are shown by the horizontal arrows for each energy branch.. The difference between these potentials gives the helical chemical potential, µH = (µ↑ − µ↓ )/2, which determines the helical imbalance of the system. Rigidly-rotating thermal states. To reveal a role of the helical potential, we consider a gas of Dirac fermions at a finite temperature, uniformly rotating about the axis z. It is convenient to introduce a particular basis of kinematic vectors for the rigid motion with the four-velocity u = Γ(∂t + Ω∂ϕ ),. Γ= p. 1 1 − ρ 2 Ω2. ,. (16). where Ω is the angular velocity and Γ is the Lorentz factor. We use cylindrical coordinates (ρ, ϕ, z). The acceleration and vorticity four-vectors are: a = ∇u u = −ρΩ2 Γ2 ∂ρ ,. ω = ω · u = ΩΓ2 ∂z ,. (17). where (∇u u)µ = uν ∇ν uµ , ω µν = 12 εµνλσ ∇λ uσ is the vorticity tensor. The fourth vector, which is orthogonal to u, a and ω, is τ µ = εµνλσ ων aλ uσ , or τ = a · ω · ∆ = −ρΩ3 Γ5 (ρΩ∂t + ρ−1 ∂ϕ ), where ∆µν = g µν − uµ uν .. (18).

(5) 4 The expectation value of an operator  is given by the thermal average over the Fock space [23, 24]: hÂi = Z −1 Tr(%̂Â),. (19). where Z = Tr(%̂) is the partition function,     X b − ΩM cz − b`  , %̂ = exp −β0 H µ`;0 Q. (20). `=V,A,H. J` = σ`z Ω + σ`ϕ Ω2 Ω × ρ + . . .. (ρ → 0), (23). where the ellipsis denotes higher-order terms in the radial distance ρ with respect to the leading terms in Eq. (23). The vortical conductivities along the rotation axis are: µV µA µ3 + 3µH (µ2V + µ2A ) 2µH T ln 2 + + H 2 2 π π 12π 2 T 2 2 µH (ω + 3a ) + + ..., 48π 2 T 2 2 2 T µV µA µH µ + µA + µ2H z σA = + + V + . . . , (24) 2 6 2π 2π 2 T µH µA µ3 + 3µV (µ2H + µ2A ) 2µV T z + V σH = 2 ln 2 + 2 π π 12π 2 T 2 2 µV (ω + 3a ) + + ..., 48π 2 T σVz =. b is the Hamiltonian (8), is the density operator [25], H z c M is the z component of the total angular momentum operator, β0 ≡ 1/T0 and µl,0 are, respectively, the values of the inverse temperature T0 and the chemical potentials at the rotation axis ρ = 0. Using the fact that b M cz ] = 0, the modes (12) are, simultaneously, the [H, cz . These mode solutions have been eigenvectors of M previously derived in Ref. [26] and are reproduced, for convenience, in Eq. (A5) of the Supplementary Material (SM). The SM describes also the analytical approaches used to compute the thermal expectation values below. The rigidly rotating gas of Dirac fermions generates the vector, axial, and helical 4-currents (` = V, A, H), J`µ ≡ h: Jb`µ :i = Q` uµ + σ`ω ω µ + σ`τ τ µ , µ. The rotation generates the vector, axial, and helical (` = V, A, H) spatial currents (21), both along the axis of rotation ez kΩ and along the circumference eϕ kΩ × ρ:. (21). µ. along the 4-velocity u , the 4-vorticity ω , and the 4circumference τ µ . The radial components along the 4acceleration vector aµ are absent for all currents (21). Henceforth, we work in the β (thermometer) frame, by fixing the four-velocity uµ to be equal to the one given in Eq. (16) [27–29]. At the rotation axis (ρ = 0), the 4-velocity uµ = (1, 0) points along the time coordinate, the 4-vorticity ω µ = (0, Ω ez ) is directed along the axis of rotation (17), while the 4-acceleration obviously vanishes, aµ ≡ 0. In a close vicinity of the axis, the vector τ µ = (0, τ ) is aligned along the circular angular coordinate (18), τ ' ρΩ3 eϕ . We calculate the 4-currents (21) in a high-temperature expansion. The charge densities Q` are as follows: µV T 2 4µA µH T µ3V + 3µV (µ2A + µ2H ) + ln 2 + 3 π2 3π 2  2 2  ω +a µA µH + µV + + ..., (22a) 4π 2 2T µA T 2 4µH µV T µ3A + 3µA (µ2H + µ2V ) QA = + ln 2 + 3 π2 3π 2 ω 2 + a2  µH µV  + µA + + ..., (22b) 4π 2 2T µH T 2 4µV µA T µ3 + 3µH (µ2V + µ2A ) QH = + ln 2 + H 2 3 π 3π 2  2 2  ω +a µV µA + µH + + ..., (22c) 2 4π 2T QV =. where we ignore terms of order O(T −1 , Ω2 /T 2 , Ω4 ). In the limit of vanishing µH , our results for QV and QA coincide with those found in Ref. [30].. where we ignored terms of order O(T −3 , Ω4 ). The circular conductivities in Eq. (23) are given by: µA µH µV + + ..., 6π 2 12π 2 T µH µV µA ϕ + ..., σA = 2+ 6π 12π 2 T µH µV µA ϕ σH = 2+ + ..., 6π 12π 2 T σVϕ =. (25a) (25b) (25c). where terms of order O(T −3 , Ω2 ) were ignored. On the rotation axis (ρ = 0), the circumferential currents vanish, and the current (23) points exactly along the vorticity Ω. The vortical transport effects (23) are consistent with the C-, P -, and T -symmetries of the vector, axial, and helical currents and charges, as shown in Table I. All vortical effects are dissipationless phenomena, because the laws (23) and (24) are even under the T -inversion. In addition to the usual chiral vortical effects (CVE’s), the currents (23) exhibit a plethora of new helical vortical effects (HVE’s). For example, the rotating dense (charged) Dirac matter generates the helical current JH that is linearly proportional to the vector chemical potential µV and temperature T (24). On the other hand, the neutral Dirac matter with nonzero helicity (µH 6= 0) generates the vector (charged) current JV . Remarkably, the mentioned helical terms, linearly proportional to a chemical potential and temperature, are allowed for HVE’s and, at the same time, are forbidden for the CVE’s by virtue of the C-, P -, and T -symmetries. Helical anomalies in QED. It is known that the vecz tor σVz and axial σA vortical conductivities (24), at a vanishing helical chemical potential, µH = 0, are determined by the axial quantum anomalies (for a review, see [31]). For example, the µV µA term in the vector vortical conductivity σVz is generated by the axial-vector-vector (AV V ) vertex of the axial anomaly (4), which is also responsible for the µ2V term in the axial vortical conducz tivity σA . Both these terms share similar prefactors with.

(6) 5 z the axial anomaly (4). The axial conductivity σA con2 tains also the µA term due to the axial-axial-axial (AAA) triangular anomaly, as well as the T 2 term which originates from the axial-graviton-graviton (AT T ) vertex of the mixed axial-gravitational anomaly [32]. The presence of the helical component in the vortical conductivities (24) strongly suggests the existence of new types of triangle anomalies which involve the helicity vertex (6). For example, the leading term µH T (µV T ) in the z vector (helical) conductivity σVz (σH ) could have its origin in the new triangle V HT anomaly involving vector (γ µ ), helical (γ µ h), and graviton (Tbµν ) vertices. The new helical anomalies must reveal themselves in the background of a “helical vector field” AH µ which couµ ples with the Dirac fermions via the source term AH µ JH added to the Lagrangian (1). For instance, the quadratic z µH dependence (24) of the axial vortical conductivity σA implies the existence of a particular form of the mixed axial-helical anomaly responsible for the nonconservation of the axial current in the background of the AH µ field. µ 1 H e H,µν The new AHH vertex, ∂µ JA = 16π2 Fµν F , shares similarity with the standard AV V vertex of the axial anomaly (4). A detailed structure of triangular anomalies with helical operators will be explored elsewhere [33]. Helical vortical waves. The emergence of the new degree of freedom, the helicity, allows us to uncover new hydrodynamic excitations in the helical sector, that are similar to the chiral magnetic [34] and chiral vortical [16] waves. To illustrate this fact, we take a globally neutral plasma (n̄` = 0, ` = V, A, H) at finite temperature (T 6= 0), and consider the simplest gapless excitation that propagates along the vorticity vector at the rotational axis, where the mean fluid velocity vanishes, v̄ = 0. The bar over a symbol means a local thermodynamic average. We consider linear modes in hydrodynamic fluctuations. We notice that in a neutral plasma, the local densities, n` = n̄` + δn` = δn` , and the local fluid velocity, v = v̄ + δv = δv, are linear in fluctuations. The “nonanomalous” component of the currents, n` v = δn` δv, vanishes at the first order, and all the currents are thus equal to their “anomalous” contributions (21). The form of the densities (22) implies that in a neutral plasma, the temperature fluctuations δT decouple from hydrodynamics of charge density waves δn` at the linear order (we set T = T̄ ). Moreover, the axial current (24) becomes quadratic in the fluctuations, so that the chiral vortical wave – that involves the vector and axial sectors – does not propagate in the neutral plasma [35–38]. On the contrary, the vector and helical densities (22) and their vortical conductivities (24) are linearly cross-coupled and therefore the vector-helical wave does propagate. We consider the wave excitation along the vorticity (0),z −iωt+ikz z vector, J`z = J` e . We impose the conservation of the vector (` = V ) and helical (` = H) charges, ∂t Q` + ∂z J`z = 0, which is valid in the absence of background fields. Using Eqs. (21), (22), and (24), we ar2 rive to the wave equation (∂t2 − vHVW ∂z2 )Q` = 0, which. describes the propagation of the helical vortical wave (HVW). The HVW is a gapless (massless) hydrodynamic excitation with the dispersion relation ω = vHVW |kz |,. vHVW =. 6 ln 2 |Ω| , π2 T. (26). where vHVW is the HVW velocity. The spectrum of the helical waves will be studied in more detail elsewhere [33]. To estimate the velocity of the HVW in ultrarelativistic heavy-ion collisions, we take the temperature T ' 150 MeV [39] matching the pseudocritical QCD value [40], and the angular frequency Ω ' 6.6 MeV ' 1022 s−1 revealed in a RHIC experiment [15, 42]. We find that at these parameters, the Helical Vortical Wave propagates with the velocity vHVW ' 2 × 10−2 c. It is instructive to compare the velocity of the helical VΩ vortical wave with its chiral analog, vCVW = 3µ π 2 T 2 [16, 35]. To this end, we set for the chemical potential µV = µq ' 30 MeV [41], and obtain vCVW ' 3 × 10−3 c, which falls into a range of the original estimation of Ref. [16]. The helical wave propagates much faster than the chiral 1 µV CVW = 2 ln wave since vvHVW 2 T ≈ 0.7µV /T while µV  T in the quark-gluon plasma being created in the ongoing heavy-ion collision experiments. Conclusions. We stress that the fermionic helicity is an independent characteristic of a massless Dirac fermion similar to its vector and axial degrees of freedom. We construct a classically conserved helicity current and find a variety of new nondissipative helical vortical effects (HVE) in rotating media of Dirac fermions. These effects include, for example, the generation of the vector (helical) current in the presence of the helicity (vector charge) density at finite temperature. The helical vortical effects lead to the appearance of a new gapless hydrodynamic excitation, the helical vortical wave (HVW), that involves coherent oscillation of vector and helical densities. The HVW propagates in a neutral plasma contrary to the chiral vortical wave that requires the presence of matter. In low-density quark-gluon plasmas created in ultrarelativistic heavy-ion collisions, the helical wave propagates much faster than its chiral cousin. We point out that the helical vortical effects have an anomalous origin, and conjecture the existence of a set of new “helical” quantum anomalies in QED that appear at the same footing as the well-known axial and axial-gravitational anomalies.. ACKNOWLEDGMENTS. The authors thank A. Cortijo and K. Landsteiner for useful discussions. The work of V.E.A. is supported by a Grant from the Romanian National Authority for Scientific Research and Innovation, CNCS-UEFISCDI, project number PN-III-P1-1.1-PD-2016-1423. The work of M.N.C. was partially supported by a Grant No. 3.6261.2017/8.9 of the Ministry of Science and Higher Education of Russia..

(7) 6. [1] G. E. Volovik, The Universe in a Helium Droplet (OUP, Oxford, 2009). [2] N. P. Armitage, E. J. Mele, and A. Vishwanath, Rev. Mod. Phys. 90, 015001 (2018). [3] M. M. Vazifeh and M. Franz, Phys. Rev. Lett. 111, 027201 (2013). [4] J. Gooth, A. C. Niemann, T. Meng, A. G. Grushin, K. Landsteiner, B. Gotsmann, F. Menges, M. Schmidt, C. Shekhar, V. Süß, et al., Nature 547, 324 (2017). [5] A. Cortijo, Y. Ferreirós, K. Landsteiner, and M. A. H., Phys. Rev. Lett. 115, 177202 (2015). [6] K. Fujikawa, H. Suzuki, et al., Path integrals and quantum anomalies (Oxford University Press on Demand, 2004). [7] D. E. Kharzeev, Prog. Part. Nucl. Phys. 75, 133 (2014). [8] P. B. Pal, Am. J. Phys. 79, 485 (2011). [9] N. Banerjee, J. Bhattacharya, S. Bhattacharyya, S. Dutta, R. Loganayagam, and P. Surowka, JHEP 01, 094 (2011). [10] J. Erdmenger, M. Haack, M. Kaminski, and A. Yarom, JHEP 01, 055 (2009). [11] M. Torabian and H.-U. Yee, JHEP 08, 020 (2009). [12] D. T. Son and P. Surówka, Phys. Rev. Lett. 103, 191601 (2009). [13] A different type of a Helical vortical effect is suggested to be realized in a vortical fluid supported by the magneticfield background [14]. We consider a purely vortical fluid in the absence of external electromagnetic fields. [14] N. Yamamoto, Phys. Rev. D 93, no. 6, 065017 (2016). [15] L. Adamczyk et al. [STAR Collaboration], Nature 548, 62 (2017). [16] Y. Jiang, X.-G. Huang, and J. Liao, Phys. Rev. D 92, 071501 (2015). [17] A. Boyarsky, J. Frohlich and O. Ruchayskiy, Phys. Rev. D 92, 043004 (2015). [18] P. Pavlovi, N. Leite and G. Sigl, Phys. Rev. D 96, no. 2, 023504 (2017) [19] K. Hattori, Y. Hirono, H. U. Yee and Y. Yin, Phys. Rev. D 100, no. 6, 065023 (2019). [20] D. O. Rybalka, E. V. Gorbar and I. A. Shovkovy, Phys. Rev. D 99, no. 1, 016017 (2019). [21] We use ↑/↓ notation for the right-/left-handed helicity to avoid possible confusion with the R/L states of the right-/left-handed chirality. [22] We leave beside the scope of our work the conformal invariance which reflects a very different kind of symmetry of the theory of massless particles (1). [23] J. I. Kapusta and P. V. Landshoff, J. Phys. G: Nucl. Part. Phys. 15, 267 (1989). [24] M. Laine and A. Vuorinen, Basics of thermal field theory (Springer, 2016). [25] A. Vilenkin, Phys. Rev. D 21, 2260 (1980). [26] V. E. Ambrus, and E. Winstanley, Phys. Lett. B 734, 296 (2014). [27] P. Ván and T. S. Biró, Phys. Lett. B 709, 106 (2012). [28] P. Ván and T. S. Biró, in Proceedings of the 12th Joint European Thermodynamics Conference, Cartolibreria SNOOPY, 2013, ed. by M. Pilotelli, G.P. Beretta, pp. 546–551. arXiv:1305.3190 [gr-qc]. [29] F. Becattini, L. Bucciantini, E. Grossi, and L. Tinti, Eur. Phys. J C 75, 191 (2015).. [30] M. Buzzegoli and F. Becattini, JHEP 12, 002 (2018). [31] K. Landsteiner, E. Megias, and F. Pena-Benitez, Lect. Notes Phys. 871, 433 (2013). [32] K. Landsteiner, E. Megias, and F. Pena-Benitez, Phys. Rev. Lett. 107, 021601 (2011). [33] V. E. Ambrus, and M. N. Chernodub, in preparation. [34] D. E. Kharzeev and H. U. Yee, Phys. Rev. D 83, 085007 (2011). [35] T. Kalaydzhyan and E. Murchikova, Nucl. Phys. B 919, 173 (2017). [36] N. Abbasi, D. Allahbakhshi, A. Davody and S. F. Taghavi, Phys. Rev. D 96, 126002 (2017). [37] At a finite density, µV 6= 0, the helical vortical wave can acquire an energy component and produce new helicalheat waves similarly to the chiral case [20, 35, 36, 38]. [38] M. N. Chernodub, JHEP 1601, 100 (2016). [39] B. V. Jacak and B. Muller, Science 337, 310 (2012). [40] Y. Aoki, G. Endrodi, Z. Fodor, S. D. Katz and K. K. Szabo, Nature 443, 675 (2006). [41] X. G. Huang and T. Koide, Nucl. Phys. A 889, 73 (2012) [42] Q. Wang, Nucl. Phys. A 967, 225 (2017). [43] V. E. Ambrus, and E. Winstanley, arXiv:1908.10244 [hepth]. [44] V. E. Ambrus, , arXiv:1912.09977 [nucl-th]. [45] B. R. Iyer, Phys. Rev. D 26, 1900 (1982)..

(8) 7 Appendix A: Supplementary material. The Supplementary Material presented herein supplements the main text with various technical details. In Subsec. A 1, we present an analysis of the properties of the helicity charge operator : QH : under the CPT transformations, which are summarized in Table I. Then, the details regarding the computation of thermal expectation values are given in Subsec. A 2. The details regarding the computation of the t.e.v.s of the vector and helicity charge currents are given in Subsec. A 3. Finally, the t.e.v.s. of the components of the axial charge current (ACC) and stress-energy tensor (SET) are derived in Subsections A 4 and A 5.. 1.. Discrete symmetries of the helicity charge operator. Taking into accout that P ψ(x)P = γ 0 ψ(−x), it can be seen that: Z −iγ · ∇x 0 P )γ ψ(−x) QH − → d3 x[ψ(−x)γ 0 (γ 0 γ 5 γ 0 p Z iγ · ∇−x = d3 x[ψ(−x)(γ 0 γ 5 γ 0 )ψ(−x) p = − QH , (A1) where ∇x = −∇−x was employed. For charge conjugation, CψC = iγ 2 ψ ∗ and CψC = ∗ −iψ γ 2 . However, the expression needs to be presented in normal order, such that b ∗ (−iγ 2 )(γ 5 γ 0 −iγ · ∇ )iγ 2 ψb∗ ] : d3 x : [ψ p Z b 0 2hψ] b∗ : = − d3 x : [ψγ Z b ∗ :, = d3 x : tr[γ 0 2hψbψ] (A2). C. Z. b H :− :Q →. where the minus sign on the last line is due to changing b the operator order of ψb and ψ. Finally, for time reversal T ψ(t)T = −γ 1 γ 3 ψ(−t), T ψ̄(t)T = ψ̄(−t)γ 1 γ 3 and T γ µ T = (γ µ )∗ , we have: T. Z. QH − → Z =. d3 x T [ψγ 5 γ 0. −iγ · ∇ ψ]T p. d3 x[ψ(−t)γ 1 γ 3 (γ 0 2h)∗ (−γ 1 γ 3 )ψ(−t)]. γ ∗ · i∇ (−γ 1 γ 3 )]ψ(−t)] p Z −iγ · ∇ = d3 x[ψ(−t)γ 0 γ 5 γ 0 [ ]ψ(−t)] p =QH . (A3) Z. =. d3 x[ψ(−t)γ 0 γ 5 γ 0 [γ 1 γ 3. 2.. Modes and thermal expectation values. In this subsection, we discuss the construction of thermal expectation values at finite vector, axial and helicity chemical potentials for massless fermions under rigid rotation. The basis of this construction is Eq. (19), which requires the computation of the trace over Fock space. This trace can be computed by first considering a full set of mode solutions of the Dirac equation, comprised of the simultaneous eigenfunctions of the system formed by the b helicity ĥ, and angular momentum M cz Hamiltonain H, operators, which appear in the expression of the density operator %̂, introduced in Eq. (19). Specifically, the particle and anti-particle mode solutions, Uj and Vj = iγ 2 Uj∗ , must satisfy Eq. (12), together with: (−i∂ϕ + S z )Uj = + mj Uj , (−i∂ϕ + S z )Vj = − mj Vj ,. −i∂z Uj = + kj Uj , −i∂z Vj = − kj Vj .. (A4). In the above, the eigenvalue of the total angular momentum operator, mj = ± 21 , ± 32 , . . . , takes only half-oddinteger values. The eigenvalue of the linear momentum operator along the z axis is kj . The particle mode soλ lutions Uj ≡ UE,k,m satisfying the eigenvalue equations (12) and (A4) can be written as [26]: ! −iEt+ikz e 1 λ UE,k,m (x) = 4π 2λE/|E|  q  i(m− 21 )ϕ 1 (qρ) 1 + 2λk e J m− p 2  , (A5) ⊗ q i(m+ 12 )ϕ J 2iλ 1 − 2λk e m+ 21 (qρ) p p where q = p2 − k 2 . For the construction of rigidly-rotating thermal states, Iyer [45] argued that the modes with positive co-rotating ej = Ej −Ωmj > 0 should be interpreted as partienergy E cle modes. Following this prescription, the field operator can be expanded with respect to the particle modes Uj and the corresponding anti-particle modes Vj = iγ 2 Uj∗ as follows: b ψ(x) =. X. Z ∞ X. λ=±1/2 m=−∞. ∞. −∞. Z. p. e dE |E|θ(E). dk −p. λ λ ˆλ;Ω;† × [UE,k,m (x)b̂λ;Ω E;k;m + VE,k,m (x)dE;k;m ], X ˆΩ;† ≡ [Uj b̂Ω (A6) j + Vj dj ], j. where the integral over E runs over the full real axis, e ensures that only positive while the step function θ(E) e co-rotating energy Ej = Ej −Ωmj > 0 particle states are taken into account. After second quantization, the mode decomposition in Eq. (A6) naturally introduces the rotating vacuum, |0Ω i, which is annihilated by the operators Ω e bΩ j and dj for which Ej > 0..

(9) 8 The decompositions of the charge operators in b and M cz Eq. (15), together with equivalent ones for H given below,. Eq. (A8), the following t.e.v.s can be derived [25]: Ω 0 e hb̂Ω;† j b̂j 0 i =nβ0 (Ej − µλj ,χj ,0 )δ(j, j ), 0 ˆΩ e hdˆΩ;† j dj 0 i =nβ0 (Ej + µλj ,−χj ,0 )δ(j, j ),. b := :H. X. Ej (b†j bj + d†j dj ),. where nβ0 is the Fermi-Dirac factor:. j. cz := :M. X. mj (b†j bj + d†j dj ),. (A10). nβ0 (a) =. (A7). 1 . eβ0 a + 1. (A11). j. An extensive discussion on how to compute the t.e.v.s of various operators using the formalism employed in this paper can be found in Refs. [43, 44]. Briefly, we consider operators Fb for which the t.e.v.s can be put in the form:. allow the following relations to be derived:. h: Fb :Ω i =. %̂b̂jΩ;† %̂−1 =eβ0 (Ej t−µλj ;χj ;0 ) b̂Ω;† j , e. −1 %̂dˆΩ;† =eβ0 (Ej t+µλj ,−χj ,0 ) dˆΩ;† j %̂ j , e. Xh ej − µλ ,χ ,0 ) F(Uj , Uj )nβ0 (E j j j. (A8). i ej + µλ ,−χ ,0 ) , (A12) −F(Vj , Vj )nβ0 (E j j where F(ψ, χ) is a bilinear form with respect to ψ and χ. The normal ordering is taken with respect to the rotating vacuum |0Ω i, i.e.. where. µλj ;χj ;0 = µV ;0 + 2λj µH;0 + χj µA;0 ,. (A9). is the total chemical potential. The chirality χj and helicity λj are related through Eq. (13). Starting from. : Fb :Ω = Fb − h0Ω |Fb|0Ω i .. (A13). In computing the t.e.v. in Eq. (A12), the integration over E in Eq. (A6) can be manipulated such that the range of E spans only the positive real semiaxis, as follows:. Z ∞ Z p ( ∞ h i X X 1 e − µλ;χ;0 ) + −nβ (|E| e + µλ;−χ;0 ) dp p h: Fb :Ω i = dk [F(Uj , Uj ) + F(Vj , Vj )] nβ0 (|E| 0 2 0 −p λ=± 12 m=−∞ ) h i e e e + sgn(E)[F(U , (A14) j , Uj ) − F(Vj , Vj )] nβ0 (|E| − µλ;χ;0 ) + nβ0 (|E| + µλ;−χ;0 ). e As discussed in Ref. [44], the t.e.v. where χ = 2λ sgn(E). b of the operator F can be expressed with respect to the Minkowski (non-rotating) vacuum |0M i, in which case it is given by: h: Fb :M i = h: Fb :Ω i − h0M | : Fb :Ω |0M i ,. (A15). where the subscript M indicates that the normal ordering is taken with respect to the Minkowski vacuum, M |0M i. The one-particle operators bM j and dj that annihilate |0M i can be introduced by considering the following mode expansion: Z ∞ Z p ∞ X X b ψ(x) = dE |E|θ(E) dk λ=±1/2 m=−∞. ×. −∞. λ [UE,k,m (x)b̂λ;M E;k;m. In Eq. (A15), h0M | : Fb :Ω |0M i represents the expectation value of the operator Fb expressed in normal order with respect to the co-rotating vacuum, taken in the Minkowski (stationary) vacuum state. It receives contributions only e < 0 [44]: from the modes for which EE. h0M | : Fb :Ω |0M i = −. X λ=± 12. Z ∞ X m=−∞. ∞. 0. +. (A16). p. dk −p. × θ(−e p) [F(Uj , Uj ) − F(Vj , Vj )] . (A17). −p ;† λ VE,k,m (x)dˆλ;M E;k;m ],. Z dp p. Inserting the above into Eq. (A15) gives:.

(10) 9. Z p ( Z ∞ ∞ X X 1 dk [F(Uj , Uj ) + F(Vj , Vj )] [nβ0 (e p − µλ;2λ;0 ) − nβ0 (e p + µλ;−2λ;0 )] dp p h: Fb :M i = 2 −p 1 m=−∞ 0 λ=± 2. ) + [F(Uj , Uj ) − F(Vj , Vj )] [nβ0 (e p − µλ;2λ;0 ) + nβ0 (e p + µλ;−2λ;0 )] . (A18). The form (A18) is more convenient for analytical manipulations. In what follows, the computation relies on the following steps. First, the Fermi-Dirac factors are expanded with respect to the rotation parameter Ω, as follows: nβ0 (e p ∓ µ) =. ∞ X (−Ωm)n dn nβ (p ∓ µ). n! dpn 0 n=0. n X. + m2n Jm (z) =. m=−∞ ∞ X. j=0 n X. − m2n+1 Jm (z) =. m=−∞. j=0. 0. 3.. 2Γ(j + 12 ) + 2j √ sn,j z , j! π 2Γ(j + √ j! π. Z. 3 2 ) + 2j sn,j z ,. + 2 2 Jm (qρ) =Jm− 1 (qρ) ± J m+ 1 (qρ),. =2J. 2. (qρ)J. m+ 12. (qρ).. (A21). The coefficients s+ n,j can be obtained from: s+ n,j =. 1 d2n+1  α 2j+1 lim . 2 sinh (2j + 1)! α→0 dα2n+1 2. It can be seen that vanishes when j > n. For small values of n − j ≥ 0, the first few coefficients are given by [26]:. s+ j+2,j =. sj+1,j =. 1 (2j + 1)(2j + 2)(2j + 3), 24. 1 (2j + 1)(2j + 2)(2j + 3) 5760 × (2j + 4)(2j + 5)(10j + 3).. (A23). For general values of n > j, the following recurrences can be established:  2 1 + + sn+1,j =sn,j−1 + j + s+ n,j , 2 n−j X 2n + 1 1 + sn,j+1 = s+ (A24) n−k,j . (j + 1)(2j + 3) 2k k=1. JHµ (ψ, χ) = 2ψγ µ hχ.. (A26). JVµ/H (Vj , Vj ) = [JVµ/H (Uj , Uj )]∗ = JVµ/H (Uj , Uj ). (A27) The above equality implies that h: JbVµ/H :Ω i = h: JbVµ/H :M i ,. (A28). since h0M | : Jbµ :Ω |0M i = 0 by virtue of Eq. (A17). Furthermore, it can be seen that JHµ (Uj , Uj ) = 2λj JVµ (Uj , Uj ), such that it is convenient to discuss the t.e.v.s of the VCC and HCC collectively using the notation:. (A22). s+ n,j. s+ j,j =1,. Vector and helicity charge currents. The bilinear forms F(ψ, χ) for the vector charge current (VCC) and helicity charge current (HCC) are:. where, for future convenience, the following notation was introduced:. m− 12. (A25). By noting that Vj = iγ 2 Uj∗ , it is not difficult to see that. n X 2Γ(j + 32 ) + 2j+1 × √ s z m2n+1 Jm (z) = , (A20) (j + 1)! π n,j m=−∞ j=0. 2. √ Γ[(n + 2)/2] π n+1 dk q = p . 2Γ[(n + 3)/2] n. JVµ (ψ, χ) = ψγ µ χ,. ∞ X. × Jm (qρ). p. (A19). The sum over m appearing in Eq. (A18) is then performed using the following formulas: ∞ X. The last step involves the integration with respect to k, which can be performed using:. µ µ . Jb± = JbVµ ± JbH. (A29). Using the explicit expressions for the modes, given in Eq. (A5), the following expressions can be obtained [44]: .  t Z ∞ h: Jb± :Ω i ∞ 1 X   ϕ b dp ρ h: J± :Ω i = 2π 2 m=−∞ 0 z h: Jb± :Ω i × [nβ0 (e p − µV ;0 ∓ µH;0 ∓ µA;0 ) − nβ0 (e p + µV ;0 ± µH;0 ∓ µA;0 )]   + Z p pJm (qρ)   × × dk  qJm (qρ)  . (A30) 0 − ±pJm (qρ) We illustrate the computational procedure for the temporal component. Expanding the Fermi-Dirac factors using.

(11) 10 in the form:. Eq. (A19), we obtain:. h:. t Jb± :Ω i. 2ΓT 3 n t Li3 [−e−(µV ±µH ∓µA )/T ] h: Jb± :Ω i = π2 o. Z ∞ ∞ 1 X Ω2n dp p = 2π 2 n=0 (2n)! 0. −Li3 [−e(µV ±µH ±µA )/T ]. 2n. ×. d [nβ0 (p − µV ;0 ∓ µH;0 ∓ µA;0 ) dp2n − nβ0 (p + µV ;0 ± µH;0 ∓ µA;0 )] Z p ∞ X + × dk m2n Jm (qρ). 0. +. (A31). m=−∞. The sum over m and integration over k can be carried out using Eqs. (A20) and (A25): Z. p. dk 0. ∞ X. + m2n Jm (qρ) =. m=−∞. =. Z n X 2Γ(j + 12 ) + 2j p √ sn,j ρ dk q 2j j! π 0 j=0 n X 2ρ2j p2j+1 j=0. 2j + 1. s+ n,j .. (A32). 1 + e(µV ±µH ±µA )/T Γ3 Ω2 T 2 (4Γ − 1) ln + O(Ω4 ). 12π 2 1 + e−(µV ±µH ∓µA )/T (A36). The high temperature limit can be extracted by noting that 3ζ(3) π 2 a a2 a3 − − ln 2 − + O(a4 ), 4 12 2 12 a a2 ln(1 + ea ) = ln 2 + + + O(a4 ), (A37) 2 8 Li3 (−ea ) = −. where ζ(z) is the Riemann zeta function. The following result is obtained: ( h:. Upon substitution of the above result into Eq. (A31), the summation over n and j can be reversed, in which case the range of j is from 0 to ∞, while n runs from j to ∞. Shifting downwards the summation over n via n → n + j, the following result is obtained: ∞ ∞ 1 X (ρΩ)2j X Ω2n (2j + 2)! + t h: Jb± :Ω i = s 2π 2 j=0 2j + 1 n=0 (2n + 2j)! n+j,j Z ∞ d2n × dp p2 2n [nβ0 (p − µV ;0 ∓ µH;0 ∓ µA;0 ) dp 0 −nβ0 (p + µV ;0 ± µH;0 ∓ µA;0 )] , (A33). j=0. j!. 0. ∞. 4T µA T2 ± ln 2 3 π2. (µV ± µH )2 + 3µ2A + O(T −1 ) 3π 2 ) i 3ω 2 + a2 h µA −2 4 + O(T ) + O(Ω ) , (A38) + 1± 12π 2 2T where ω 2 = Ω2 Γ4 and a2 = ρ2 Ω4 Γ4 = Ω2 Γ2 (Γ2 − 1). Performing the exact same steps for the ϕ component yields: ( T2 4T µA ϕ b h: J± :Ω i = ΩΓ(µV ± µH ) ± ln 2 3 π2 (µV ± µH )2 + 3µ2A + O(T −1 ) 3π 2 ) i ω 2 + 3a2 h µA −2 4 + 1± + O(T ) + O(Ω ) . (A39) 12π 2 2T +. z For h: Jb± :Ω i, the following expression is obtained:. (ρΩ)2 j = n!Γ2n+2 ,. (A34) z h: Jb± :Ω i = ±. where Γ = (1 − ρ2 Ω2 )−1/2 is the Lorentz factor corresponding to rigid rotation. Changing the integration variable to x = β0 p and using Z. = Γ(µV ± µH ) +. where integration by parts was performed 2j times in the integral over p. In the above expression, the series over n contributes corrections of order Ω2n to the t.e.v. Retaining only terms up to O(Ω2 ), the series over n can be truncated at n = 1. The sum over j can be performed at each value of n using the following results: ∞ X (j + n)!. t Jb± :Ω i. xn dx = −n!Lin+1 (−ea ), ex−a + 1. Γ2 ΩT 2 n Li2 [−e−(µV ±µH ∓µA )/T ] π2 o. −Li2 [−e(µV ±µH ±µA )/T ] ±. Γ4 Ω3 (4Γ2 − 3) sinh[(µV ± µH )/T ] + O(Ω4 ). 24π 2 cosh[(µV ± µH )/T ] + cosh(µA /T ) (A40). (A35). where Lin (z) is the polylogarithm function, and n = 0, 1, 2, . . . is a natural number, Eq. (A33) can be put. Using the following expansion for the polylogarithm: Li2 (−ea ) = −. π2 a2 a3 − a ln 2 − − + O(a5 ), 12 4 24. (A41).

(12) 11 the following expansion is obtained: h:. z Jb± :Ω i. ( ΩΓ2 = ± 2 (µV ± µH ) π. (µV ± µH )2 + 3µ2A + O(T −2 ) 12T )  ω 2 + 3a2  −2 4 1 + O(T ) + O(Ω ) . (A42) + 48T. 2T ln 2 ± µA +. τ The charge density Q± and azimuthal conductivity σ± can be obtained by inverting the expansion (21). Noting ϕ t that Q± = u · h: J± :Ω i = Γ(h: J± :Ω i − ρ2 Ω h: J± :Ω i), the following expression is obtained:. ( Q± = (µV ± µH ). T2 4T µA ln 2 ± 3 π2. µ2 + µ2H + 3µ2A µV µH + V ± + O(T −1 ) 3π 2 π2 ) i ω 2 + a2 h µA + 1± + O(T −2 ) + O(Ω4 ) . 4π 2 2T. The results above show that JAt (Uj , Uj ) and JAϕ (Uj , Uj ) are odd under the simultaneous flip (kj , λj ) → (−kj , −λj ). Because of this property, Eq. (A17) shows that ϕ t h0M | : JbA :Ω |0M i = h0M | : JbA :Ω |0M i = 0.. (A43). The charge conductivity can be obtained using µ ϕ t τµ h: J± :Ω i Ω h: J± :Ω i − h: J± :Ω i = 2 3 3 τ Ω Γ i µV ± µH h µA −2 = + O(T ) + O(Ω2 ). 1 ± 6π 2 2T. The above property implies that the term on the first line in Eq. (A12) makes a vanishing contribution to the t.e.v.s of the components of the ACC. Using the explicit expression for the modes given in Eq. (A5), the following results can be obtained [44]:   pj kj − + JAt (Uj , Uj ) = 2 2λj Jm (q ρ) + J (q ρ) , j j j 8π |Ej | pj mj λj qj × JAϕ (Uj , Uj ) = 2 J (qj ρ), 4π pj ρ mj   2λj kj + 1 − (q ρ) + J (q ρ) . (A49) JAz (Uj , Uj ) = 2 Jm j mj j j 8π pj. τ σ± =. (A44). ω can be obtained as Finally, the vortical conductivity σ± follows: ω z :Ω i ωµ h: J± :Ω i h: J± = ω2 ΩΓ2 µ± (µ2± + 3µ2A ) 2µ± T µA µ± =± + O(T −2 ) ln 2 + ± π2 π2 12π 2 T µ± (ω 2 + 3a2 ) ± [1 + O(T −2 )] + O(Ω4 ). (A45) 48π 2 T. Thus, the t.e.v.s of the t and ϕ components of the ACC with respect to the rotating vacuum are the same as when expressed with respect to the Minkowski vacuum. This conclusion does not hold true for the z component, for which there will be non-vanishing vacuum terms in the t.e.v. computed with respect to the Minkowski vacuum. Using Eq. (A49), the following expressions can be obtained [44]:   t Z ∞ :Ω i h: JbA ∞ 1 X X   ϕ dp :Ω i = ρ h: JbA 4π 2 1 m=−∞ 0 z b λ=± h: JA :Ω i 2 × [nβ0 (e p − µλ;2λ;0 ) + nβ0 (e p + µλ;−2λ;0 )]   + Z p (qρ) 2λpJm   × × dk 2λqJm (qρ) . (A51) 0 − pJm (qρ). ω σ± =. The vector and helical quantities can be reconstructed from the above results using: h: JVµ/H :Ω i =. 4.. 1 µ µ (h: J+ :Ω i ± h: J− :Ω i). 2. (A46). The bilinear form F(ψ, χ) for the axial charge current (ACC) is:. ∞ ∞ X 1 X (ρΩ)2j X Ω2n (2j + 2)! t b h: JA :Ω i = 2λ 4π 2 2j + 1 n=0 (2n + 2j)! j=0 λ=± 12 Z ∞ d2n × s+ dp p2 2n [nβ0 (p − µλ;2λ;0 ) n+j,j dp 0 +nβ0 (p + µλ;−2λ;0 )] . (A52). 3. ΓT t h: JbA :Ω i = − 2 π. (A48). h. µ 1 ,1 /T. Li3 (−e. 2. ) − Li3 (−e. µ− 1 ,−1 /T 2. ). i −µ 1 /T −µ 1 /T +Li3 (−e 2 ,−1 ) − Li3 (−e − 2 ,1 ). (A47). Taking into account the charge conjugation symmetry, Vj = iγ 2 Uj∗ , it can be shown that JAµ (Vj , Vj ) = −[JAµ (Uj , Uj )]∗ = JAµ (Uj , Uj ).. Employing the same steps as in Subsec. A 3, we obtain:. The above expression can be approximated up to O(Ω2 ) by taking into account the n = 0 and n = 1 terms. The following result is obtained:. Axial charge current. JAµ (ψ, χ) = ψγ µ γ 5 χ.. (A50). +. H eµA /T (cosh µV +µ + cosh µTA ) Γ 3 Ω2 T 2 T (4Γ − 1) ln H 24π 2 e−µA /T (cosh µV −µ + cosh µTA ) T. + O(Ω4 ). (A53).

(13) 12 In the high temperature limit, the following result is obtained: (" µA T 2 4T µV µH t b ln 2 h: JA :Ω i = Γ + 3 π2 # 3µA (µV + µH )2 + µ3A −1 + O(T ) + 3π 2 ) i 3ω 2 + a2 h µV µH −2 4 + µA + + O(T ) + O(Ω ) . 12π 2 2T (A54). with respect to the rotating vacuum can be obtained by subtracting this term. The result is: " 2 µ2 + µ2H + µ2A z 2 T b h: JA :Ω i = ΩΓ + V 6 2π 2 # µV µH µA −2 4 + + O(T , Ω ) . (A58) 2π 2 T The corresponding vortical conductivity can be obtained as shown in Eq. (A45) and the result is summarised in Eq. (24).. Similarly, the ϕ component is given by: (" ϕ h: JbA :Ω i = ΩΓ. 5.. 4T µV µH µA T 2 ln 2 + 3 π2. For completeness of the overall presentation, we also provide the result for the thermal expectation value (t.e.v.) of the stress-energy tensor (SET),. # 3µA (µV + µH )2 + µ3A −1 + + O(T ) 3π 2 2. i b b b b Tbµν = [Ψγ (µ ∇ν) Ψ − ∇(µ Ψγν) Ψ]. 2. ). 2. Stress-energy tensor. i µV µH ω + 3a h µA + + O(T −2 ) + O(Ω4 ) . + 2 12π 2T (A55) τ The charge density QA and azimuthal conductivity σA can be obtained using the relations (A43) and (A44). The results is summarised in Eqs. (22b) and (25b), respectively. z :Ω i, the following expression is obtained: For h: JbA. (A59). The bilinear form introduced in Eq. (A12) which corresponds to the SET is: Tµν (ψ, χ) =. i [ψγ(µ ∇ν) χ − ∇(µ ψγν) χ]. 2. (A60). It is easy to check that Tµν (Vj , Vj ) = −[Tµν (Uj , Uj )]∗ = −Tµν (Uj , Uj ). (A61). Due to the above property, the first line in Eq. (A14) vanishes, as was the case for the t.e.v. of the ACC disΓ2 ΩT 2 h −µ 1 ,1 /T −µ 1 ,1 /T z b 2 2 h: J± :A i = − Li2 (−e ) + Li2 (−e ) cussed in Sec. A 4. The explicit expressions for the bi2π 2 linears Tµν (Uj , Uj ) are presented in Refs. [43, 44] and for i −µ 1 /T −µ 1 /T brevity, they are not reproduced here. The procedure to + Li2 (−e 2 ,1 ) + Li2 (−e 2 ,1 ) compute " # the t.e.v.s is also detailed in these references, be−µH H eµA /T + cosh µV +µ e−µA /T + cosh µVing Γ4 Ω 3 the one employed for the ACC in Sec. A 4. 2 T T identical to + (4Γ −3) + +O(Ω4 ). µA µVThe −µHhydrodynamic H content of the SET can be extracted 48π 2 cosh µTA + cosh µV +µ cosh + cosh T T T by considering the following decomposition: (A56). The large temperature expansion is easily obtained:. h: Tbµν :i = P (4uµ uν −gµν )+Πµν +uµ Wν +uν Wµ , (A62) which contains the isotropic pressure component,. (" z h: JbA :M i = ΩΓ2. 2. T + 6. µ2V. + µ2H + 2π 2. µV µH µA + + O(T −2 ) 2π 2 T. µ2A. #. )  ω 2 + 3a2  + 1 + O(T −3 ) + O(Ω4 ) . (A57) 24π 2 It can be seen that the term on the last line above is independent of the thermodynamic parameters T and µl (l ∈ {V, H, A}). This term represents the vacuum conz tribution h0M | : JbA :Ω |0M i, which was previously found in many studies [25, 26, 44]. The t.e.v. of Jbz computed. 7π 2 T 4 T2 2 + (µ + µ2V + µ2H ) 180 6 A µ4 + µ4H + µ4A + 6(µ2V µ2H + µ2V µ2A + µ2H µ2A ) + V 12π 2   2 2 3ω + a 3(µ2V + µ2H + µ2A ) + T2 + 72 π2 4µA µH µV T + ln 2 + O(T 0 ), (A63) π2 the anisotropic stress contribution,   ω 2 µ ν a2 µ ν µν µ ν Π = ΠV τ τ − a a − ω ω 2 2 +ΠH (τ µ ω ν + τ ν ω µ ), (A64) P =.

(14) 13 and the heat flux:. The conductivities κτ and κω are as follows:   3(µ2V + µ2H + µ2A ) 1 2 τ + O(T −1 ), T + κ =− 18 π2 µ3A + 3µA (µ2V + µ2H ) µA T 2 4µH µV T ln 2 + κω = + 3 π2 3π 2 2 2 2 µV µH (3µA + µV + µH ) + 6π 2 T 2 2 µV µH ω +a [µA + ] + O(T −3 , Ω4 ), (A66) + 12π 2 2T while the coefficient ΠH is given by: ΠH = − W µ = κτ τ µ + κω ω µ .. (A65). µH µV µA − + O(T −2 , Ω2 ). 2 3π 6π 2 T. (A67). For massless particles, we find that ΠV = 0, while the energy density is E = 3P ..

(15)

Références

Documents relatifs

and 3 of histograms of zones obtained after sliding a window of size 16x16 pixel on the MR image of the brain, while the parameters of the feature vector used in this study is

Elastic wave propagation inside helical waveguides has been analyzed through a SAFE method based on a non-orthogonal curvilinear coordinate system that is

TREYSSEDE, Fabien, 2015, Veering of dispersion curves and prestress effects in multi-wire helical waveguides, 9th GDR Conference on Wave Propagation in Complex Media for

Since the uniform texture below threshold (Po 2 pot) is definitely stable for perturbations with _q + 0, we anticipate that small q2 will remain the most im- portant

The next step is to adapt the optics toolbox to Dirac fermions using p–n junctions as high-transmission diopters and electrostatically-shaped Klein tunneling barriers to implement

Le second modèle proposé (chapitre 5 ) tire parti de l’analyse des forces des deux familles d’approches : d’une part notre modèle utilise la rapidité d’inférence des

Le nombre total d'espèces dans cette zone pro- tégée (124 espèces réparties en 39 familles) est assez faible en comparaison des données récoltées pour d'autres

The seasonal behaviour of the standard emission potential in our study, with high emission potentials in spring and early summer, lower in late summer and higher again towards