Inhomogeneous confinement and chiral symmetry breaking induced by imaginary angular velocity

Shi Chen Kenji Fukushima Yusuke Shimada
Abstract

We investigate detailed properties of imaginary rotating matter with gluons and quarks at high temperature. Previously, we showed that imaginary rotation induces perturbative confinement of gluons at the rotation center. We perturbatively calculate the Polyakov loop potential and find inhomogeneous confinement above a certain threshold of imaginary angular velocity. We also evaluate the quark contribution to the Polyakov loop potential and confirm that spontaneous chiral symmetry breaking occurs in the perturbatively confined phase.

keywords:
Confinement, Rotation, Chiral Symmetry Breaking, Inhomogeneous States
journal: Physics Letters B
\affiliation

[1]organization=School of Physics and Astronomy, University of Minnesota, addressline=Minneapolis, city=MN, postcode=55455, country=U.S.A.

\affiliation

[2]organization=Department of Physics, The University of Tokyo, addressline=7-3-1 Hongo, Bunkyo-ku, city=Tokyo, postcode=113-0033, country=Japan

1 Introduction

The physical mechanism and the interplay of confinement and chiral symmetry breaking are long-standing puzzles in Quantum Chromodynamics (QCD) that is a fundamental theory in terms of quarks and gluons. At finite temperature, if the quark mass is infinitely large to quench dynamical quarks, confinement and deconfinement of gluons are classified by center symmetry. Dynamical quarks explicitly break center symmetry [1], and the order parameter for center symmetry, i.e., the Polyakov loop is only an approximate measure of confinement. In the limit of zero quark mass, chiral symmetry is exact and the chiral condensate characterizes the state of matter. Various phases of QCD have been considered as functions of external environmental parameters such as the temperature T𝑇Titalic_T, the quark chemical potential μ𝜇\muitalic_μ, the isospin chemical potential [2, 3], the electric and magnetic fields [4, 5, 6], and the rotation angular velocity ω𝜔\omegaitalic_ω [7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29], and their mixture [30, 31, 32]. In this way, the QCD phase diagram has been intensively studied in theoretical and experimental contexts [33].

Among various parameters, the angular velocity ω𝜔\omegaitalic_ω has attracted much attention in recent years. An extraordinary value of ω1022s1similar-to𝜔superscript1022superscripts1\omega\sim 10^{22}\ \mathrm{s}^{-1}italic_ω ∼ 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is estimated based on the data in the heavy-ion collision [34]. In nonrelativistic theories rotation effects are similar to those induced by magnetic fields, while relativistic rotation of quark matter has features analogous to finite density [30, 8, 9, 35]. In the same way as chiral symmetry restoration at high density, chiral effective models such as the (Polyakov–)Nambu–Jona-Lasino model exhibit the chiral phase transition induced by rotation [7, 8, 9, 10, 11]. Interestingly, rotation directly affects gluons, which makes a contrast to effects of finite density and magnetic fields, and the angular velocity turns out to be a useful probe to confinement and deconfinement. For the purpose to investigate a chance of (de)confinement caused by rotation, the hadron resonance gas model and the holographic QCD model have been adopted [11, 12, 13, 14, 15, 16], which predicted that real rotation favors deconfinement. Thus, it is expected that the deconfinement critical temperature should decrease for larger ω𝜔\omegaitalic_ω.

The numerical results from lattice-QCD simulations [18, 19, 20] have invoked controversies reporting a trend opposite to the model predictions. That is, the Polyakov loop decreases as a result of real rotation, so that the deconfinement critical temperature should increase. A technical subtlety lies in the treatment of analytical continuation; the sign problem can be evaded with imaginary angular velocity, ΩIsubscriptΩI\Omega_{\text{I}}roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT, and then the critical temperature is inferred from Tc(ω2)=Tc(ΩI2)subscript𝑇csuperscript𝜔2subscript𝑇csuperscriptsubscriptΩI2T_{\text{c}}(\omega^{2})=T_{\text{c}}(-\Omega_{\text{I}}^{2})italic_T start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ( italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = italic_T start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ( - roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) if the Wick rotation does not hit singularities.

The subtle connection between real and imaginary rotation has been studied recently [10, 20, 23, 24, 28, 29] but the validity condition for analytical continuation is only partially clarified for the moment. There are various scenarios; some support the lattice results [21, 22], some propose a modified model which can explain the lattice results [23], some suggest a non-monotonic function of Tc(ω)subscript𝑇c𝜔T_{\text{c}}(\omega)italic_T start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ( italic_ω ) [24, 25].

Previously in Ref. [36], we exploited the perturbative Polyakov loop potential to investigate rotation effects based on the first-principles approach. Thanks to the asymptotic freedom, QCD or the pure gluonic theory can be perturbatively analyzed if the temperature is sufficiently high. Without rotation, the perturbative Polyakov loop potential, known as the GPY-Weiss potential [37, 38, 39, 40, 41], spontaneously breaks center symmetry, which is consistent with deconfinement expected at high temperature; for a review, see Ref. [42]. We obtained the Polyakov loop potential in the rotating pure gluonic system and found that the Wick rotation, ΩI=iωsubscriptΩIi𝜔\Omega_{\text{I}}=-\mathrm{i}\omegaroman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = - roman_i italic_ω, is hindered by singularity which corresponds to the causality violation. To cure this obstacle, the system must have a finite boundary, and then a closed analytical expression is no longer available.

Interestingly enough, the theory with imaginary angular velocity contains rich physics. In our previous work [36], we discovered exotic realization of confinement even at arbitrarily high temperature, i.e., the perturbatively confined phase. Thus, ΩIsubscriptΩI\Omega_{\text{I}}roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT is a novel theoretical device to tackle the confinement mechanism. In the present work, we will discuss two nontrivial extensions from our previous study. One is exploring inhomogeneous structures away from the imaginary rotation center. In Ref. [36], we focused on r=0𝑟0r=0italic_r = 0 only (where r𝑟ritalic_r is the radial distance) and argued that confinement is homogeneously realized for SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) at ΩI/T=πsubscriptΩI𝑇𝜋\Omega_{\text{I}}/T=\piroman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT / italic_T = italic_π. There are some related works [27, 28, 29] in favor of inhomogeneous confinement/deconfinement with real and imaginary rotation. We will show that the perturbative Polyakov loop potential exhibits an inhomogeneous distribution of the Polyakov loop at r0𝑟0r\neq 0italic_r ≠ 0. Another extension is the inclusion of dynamical quarks, with which we can consider chiral symmetry breaking in the perturbatively confined phase. The relation between confinement and chiral symmetry breaking is not fully understood, but under reasonable assumptions, one can see that confinement leads to chiral symmetry breaking or spontaneous generation of quark mass. We will verify that quark mass is indeed nonzero in the perturbatively confined phase.

2 Inhomogeneous confinement and deconfinement

Real rotation is a real-time effect, while imaginary rotation is interpreted as a geometrical effect. The latter has a theoretical advantage as explicated below. In the presence of an angular velocity vector, 𝝎𝝎\bm{\omega}bold_italic_ω, a thermal system is described by the following partition function:

𝒵T,𝝎=treβ(H^+𝝎𝑱^),subscript𝒵𝑇𝝎tracesuperscripte𝛽^𝐻𝝎^𝑱\mathcal{Z}_{T,\bm{\omega}}=\tr\mathrm{e}^{-\beta(\hat{H}+\bm{\omega}\cdot\hat% {\bm{J}})}\,,caligraphic_Z start_POSTSUBSCRIPT italic_T , bold_italic_ω end_POSTSUBSCRIPT = roman_tr roman_e start_POSTSUPERSCRIPT - italic_β ( over^ start_ARG italic_H end_ARG + bold_italic_ω ⋅ over^ start_ARG bold_italic_J end_ARG ) end_POSTSUPERSCRIPT , (1)

where β=1/T𝛽1𝑇\beta=1/Titalic_β = 1 / italic_T is the inverse temperature and 𝑱^^𝑱\hat{\bm{J}}over^ start_ARG bold_italic_J end_ARG denotes the total angular momentum operator. We can regard the above expression as a thermal expectation value of the topological operator, eβ𝝎𝑱^superscripte𝛽𝝎^𝑱\mathrm{e}^{-\beta\bm{\omega}\cdot\hat{\bm{J}}}roman_e start_POSTSUPERSCRIPT - italic_β bold_italic_ω ⋅ over^ start_ARG bold_italic_J end_ARG end_POSTSUPERSCRIPT. For pure imaginary 𝝎=i𝛀I𝝎isubscript𝛀I\bm{\omega}=\mathrm{i}{\bm{\Omega}}_{\text{I}}bold_italic_ω = roman_i bold_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT, this topological operator,

eiβ𝛀I𝑱^,superscriptei𝛽subscript𝛀I^𝑱\mathrm{e}^{-\mathrm{i}\beta{\bm{\Omega}}_{\text{I}}\cdot\hat{\bm{J}}}\,,roman_e start_POSTSUPERSCRIPT - roman_i italic_β bold_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_italic_J end_ARG end_POSTSUPERSCRIPT , (2)

is a unitary operator that generates a rotation by β|𝛀I|𝛽subscript𝛀I\beta|{\bm{\Omega}}_{\text{I}}|italic_β | bold_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT | angle along the rotation axis 𝛀I\parallel\!{\bm{\Omega}}_{\text{I}}∥ bold_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT. In the Euclidean path integral, we can take account of 𝛀Isubscript𝛀I{\bm{\Omega}}_{\text{I}}bold_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT in the partition function (1) by imposing an aperiodic thermal boundary condition which in cylindrical coordinates takes the form,

(r,θ,z,τ)(r,θβΩI,z,τ+β).similar-to𝑟𝜃𝑧𝜏𝑟𝜃𝛽subscriptΩI𝑧𝜏𝛽(r,\theta,z,\tau)\sim(r,\theta-\beta\Omega_{\text{I}},z,\tau+\beta)\,.( italic_r , italic_θ , italic_z , italic_τ ) ∼ ( italic_r , italic_θ - italic_β roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT , italic_z , italic_τ + italic_β ) . (3)

Alternatively, we can perform the coordinate transformation to change the condition (3) to the ordinary periodic one, but then the metric gains nontrivial components in a rotating frame, i.e.,

gμν=(10000r20r2ΩI00100r2ΩI01+r2ΩI2).subscript𝑔𝜇𝜈matrix10000superscript𝑟20superscript𝑟2subscriptΩI00100superscript𝑟2subscriptΩI01superscript𝑟2superscriptsubscriptΩI2\displaystyle g_{\mu\nu}=\begin{pmatrix}1&0&0&0\\ 0&r^{2}&0&r^{2}\Omega_{\text{I}}\\ 0&0&1&0\\ 0&r^{2}\Omega_{\text{I}}&0&1+r^{2}\Omega_{\text{I}}^{2}\end{pmatrix}\,.italic_g start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT = ( start_ARG start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL 0 end_CELL start_CELL italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL start_CELL 1 + italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ) . (4)

Both approaches lead us to the same partition function.

We shall discuss inhomogeneous structures in the perturbatively confined phase in SU(Nc)𝑆𝑈subscript𝑁cSU(N_{\text{c}})italic_S italic_U ( italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ) pure gluonic matter with Nc=2subscript𝑁c2N_{\text{c}}=2italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 2 and 3333. Although we have already given a derivation in Ref. [36], we quickly review the formulation to make this paper self-contained. In our calculation, we take the cylindrical coordinates, (r,θ,z,τ)𝑟𝜃𝑧𝜏(r,\theta,z,\tau)( italic_r , italic_θ , italic_z , italic_τ ), and assume rigidly rotating matter along the z𝑧zitalic_z-axis.

We take a constant AB4subscript𝐴B4A_{\text{B}4}italic_A start_POSTSUBSCRIPT B 4 end_POSTSUBSCRIPT background and then τsubscript𝜏\partial_{\tau}∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT is replaced by the covariant derivative Dτsubscript𝐷𝜏D_{\tau}italic_D start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT. The inhomogeneity will be studied within the local density approximation (in which the spatial derivatives on AB4subscript𝐴B4A_{\text{B}4}italic_A start_POSTSUBSCRIPT B 4 end_POSTSUBSCRIPT are neglected). Using the Cartan subalgebra 𝔥𝔥\mathfrak{h}fraktur_h of Lie algebra 𝔤𝔤\mathfrak{g}fraktur_g of gauge group G𝐺Gitalic_G, the covariant derivative is

Dτ=τ+iϕ𝑯β,subscript𝐷𝜏subscript𝜏ibold-italic-ϕ𝑯𝛽D_{\tau}=\partial_{\tau}+\mathrm{i}\frac{\bm{\phi}\cdot\bm{H}}{\beta}\,,italic_D start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT = ∂ start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT + roman_i divide start_ARG bold_italic_ϕ ⋅ bold_italic_H end_ARG start_ARG italic_β end_ARG , (5)

where 𝑯𝑯\bm{H}bold_italic_H is a vector of bases of 𝔥𝔥\mathfrak{h}fraktur_h and ϕ𝑯bold-italic-ϕ𝑯{\bm{\phi}}\cdot\bm{H}bold_italic_ϕ ⋅ bold_italic_H is normalized AB4subscript𝐴B4A_{\text{B}4}italic_A start_POSTSUBSCRIPT B 4 end_POSTSUBSCRIPT. Using this background field, we fix the gauge by setting DμAμ=0subscript𝐷𝜇subscript𝐴𝜇0D_{\mu}A_{\mu}=0italic_D start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = 0.

In the actual calculations, the scalar Laplacian, Ds2=(DτΩIθ)2r1r(rr)r2θ2z2subscriptsuperscript𝐷2ssuperscriptsubscript𝐷𝜏subscriptΩIsubscript𝜃2superscript𝑟1subscript𝑟𝑟subscript𝑟superscript𝑟2superscriptsubscript𝜃2superscriptsubscript𝑧2-D^{2}_{\mathrm{s}}=-\quantity(D_{\tau}-\Omega_{\text{I}}\partial_{\theta})^{2% }-r^{-1}\partial_{r}\quantity(r\partial_{r})-r^{-2}\partial_{\theta}^{2}-% \partial_{z}^{2}- italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = - ( start_ARG italic_D start_POSTSUBSCRIPT italic_τ end_POSTSUBSCRIPT - roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_r start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ( start_ARG italic_r ∂ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ) - italic_r start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ∂ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, is the basic building block. We solve the equation of motion, Ds2Ψ=λΨsubscriptsuperscript𝐷2sΨ𝜆Ψ-D^{2}_{\mathrm{s}}\Psi=\lambda\Psi- italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT roman_Ψ = italic_λ roman_Ψ, to find the spectrum as

λn,l,k,𝜶=(2πn+ϕ𝜶βΩIl)2+k2+kz2,subscript𝜆𝑛𝑙𝑘𝜶superscript2𝜋𝑛bold-italic-ϕ𝜶𝛽subscriptΩI𝑙2superscriptsubscript𝑘perpendicular-to2superscriptsubscript𝑘𝑧2\displaystyle\lambda_{n,l,k,\bm{\alpha}}=\quantity(\frac{2\pi{n}+\bm{\phi}\!% \cdot\!{\bm{\alpha}}}{\beta}-\Omega_{\text{I}}{l})^{2}+{k_{\perp}}^{2}+{k_{z}}% ^{2}\,,italic_λ start_POSTSUBSCRIPT italic_n , italic_l , italic_k , bold_italic_α end_POSTSUBSCRIPT = ( start_ARG divide start_ARG 2 italic_π italic_n + bold_italic_ϕ ⋅ bold_italic_α end_ARG start_ARG italic_β end_ARG - roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT italic_l end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (6)

where quantum numbers are n,l𝑛𝑙n,l\in\mathbb{Z}italic_n , italic_l ∈ blackboard_Z, k+subscript𝑘perpendicular-tosuperscriptk_{\perp}\in\mathbb{R}^{+}italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ∈ blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, kzsubscript𝑘𝑧k_{z}\in\mathbb{R}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ∈ blackboard_R, and 𝜶Φ𝜶Φ\bm{\alpha}\in\Phibold_italic_α ∈ roman_Φ with ΦΦ\Phiroman_Φ denoting the union of the 𝔰𝔲(Nc)𝔰𝔲subscript𝑁c\mathfrak{su}(N_{\text{c}})fraktur_s fraktur_u ( italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ) root system and the zero roots. The ghost contribution needs the determinant of Ds2subscriptsuperscript𝐷2s-D^{2}_{\mathrm{s}}- italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, which is given in terms of λn,l,k,𝜶subscript𝜆𝑛𝑙𝑘𝜶\lambda_{n,l,k,\bm{\alpha}}italic_λ start_POSTSUBSCRIPT italic_n , italic_l , italic_k , bold_italic_α end_POSTSUBSCRIPT.

The contribution from the gauge field requires the vector Laplacian, Dv2subscriptsuperscript𝐷2v-D^{2}_{\mathrm{v}}- italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT. For the explicit form of Dv2subscriptsuperscript𝐷2v-D^{2}_{\mathrm{v}}- italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_v end_POSTSUBSCRIPT, see Ref. [36]. The eigenvalue spectrum is the same as the scalar one. Each eigenmode has four polarization degrees of freedom, and two out of four are canceled by the ghost contribution. After some calculations, we arrive at the Polyakov loop potential resulting from the two physical (transverse) modes as

Vg(ϕ;Ω~I)=T4π2𝜶Φlsubscript𝑉𝑔bold-italic-ϕsubscript~ΩI𝑇4superscript𝜋2subscript𝜶Φsubscript𝑙\displaystyle V_{g}({\bm{\phi}};\tilde{\Omega}_{\text{I}})=\frac{T}{4\pi^{2}}% \sum_{{\bm{\alpha}}\in\Phi}\sum_{l\in\mathbb{Z}}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( bold_italic_ϕ ; over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ) = divide start_ARG italic_T end_ARG start_ARG 4 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT bold_italic_α ∈ roman_Φ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_l ∈ blackboard_Z end_POSTSUBSCRIPT 0kdk+dkzsuperscriptsubscript0subscript𝑘perpendicular-todifferential-dsubscript𝑘perpendicular-tosuperscriptsubscriptdifferential-dsubscript𝑘𝑧\displaystyle\int_{0}^{\infty}\!\!k_{\perp}\mathrm{d}{k_{\perp}}\int_{-\infty}% ^{+\infty}\!\!\!\mathrm{d}{k_{z}}∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT roman_d italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + ∞ end_POSTSUPERSCRIPT roman_d italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT
×[Jl12(kr)+Jl+12(kr)]Reln(1eiϕ𝜶iΩ~Ilβ|𝒌|).absentdelimited-[]subscriptsuperscript𝐽2𝑙1subscript𝑘perpendicular-to𝑟subscriptsuperscript𝐽2𝑙1subscript𝑘perpendicular-to𝑟Re1superscripteibold-italic-ϕ𝜶isubscript~ΩI𝑙𝛽𝒌\displaystyle\times\Bigl{[}J^{2}_{l-1}(k_{\perp}r)+J^{2}_{l+1}(k_{\perp}r)% \Bigr{]}\;\mathrm{Re}\ln\quantity(1-\mathrm{e}^{\mathrm{i}\bm{\phi}\cdot\bm{% \alpha}-\mathrm{i}\tilde{\Omega}_{\text{I}}l-\beta|\bm{k}|})\,.× [ italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l - 1 end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_r ) + italic_J start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l + 1 end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_r ) ] roman_Re roman_ln ( start_ARG 1 - roman_e start_POSTSUPERSCRIPT roman_i bold_italic_ϕ ⋅ bold_italic_α - roman_i over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT italic_l - italic_β | bold_italic_k | end_POSTSUPERSCRIPT end_ARG ) . (7)

For notational brevity, we introduced a dimensionless imaginary angular velocity; Ω~I=ΩI/Tsubscript~ΩIsubscriptΩI𝑇\tilde{\Omega}_{\text{I}}=\Omega_{\text{I}}/Tover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = roman_Ω start_POSTSUBSCRIPT I end_POSTSUBSCRIPT / italic_T. By expanding the logarithm, we can perform the momentum integration to simplify the above form into

Vg(ϕ;Ω~I)=2T4π2𝜶Φn=1cos(nϕ𝜶)cos(nΩ~I){n2+2r~2[1cos(nΩ~I)]}2.subscript𝑉𝑔bold-italic-ϕsubscript~ΩI2superscript𝑇4superscript𝜋2subscript𝜶Φsuperscriptsubscript𝑛1𝑛bold-italic-ϕ𝜶𝑛subscript~ΩIsuperscriptsuperscript𝑛22superscript~𝑟2delimited-[]1𝑛subscript~ΩI2\displaystyle V_{g}({\bm{\phi}};\tilde{\Omega}_{\text{I}})=-\frac{2T^{4}}{\pi^% {2}}\sum_{{\bm{\alpha}}\in\Phi}\sum_{n=1}^{\infty}\frac{\cos(n{\bm{\phi}}\cdot% {\bm{\alpha}})\cos(n\tilde{\Omega}_{\text{I}})}{\Bigl{\{}n^{2}+2\tilde{r}^{2}% \bigl{[}1-\cos(n\tilde{\Omega}_{\text{I}})\bigr{]}\Bigr{\}}^{2}}\,.italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( bold_italic_ϕ ; over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ) = - divide start_ARG 2 italic_T start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT bold_italic_α ∈ roman_Φ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG roman_cos ( start_ARG italic_n bold_italic_ϕ ⋅ bold_italic_α end_ARG ) roman_cos ( start_ARG italic_n over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT end_ARG ) end_ARG start_ARG { italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 over~ start_ARG italic_r end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [ 1 - roman_cos ( start_ARG italic_n over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT end_ARG ) ] } start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (8)

The potential is dependent on the dimensionless radial distance; r~=rT~𝑟𝑟𝑇\tilde{r}=rTover~ start_ARG italic_r end_ARG = italic_r italic_T. The potential is minimized at the optimal value of ϕbold-italic-ϕ{\bm{\phi}}bold_italic_ϕ, and the Polyakov loop expectation value, L(ϕ)𝐿bold-italic-ϕL({\bm{\phi}})italic_L ( bold_italic_ϕ ), is evaluated accordingly. The denominator has singularities at r~0~𝑟0\tilde{r}\neq 0over~ start_ARG italic_r end_ARG ≠ 0 even for ϕ=0bold-italic-ϕ0{\bm{\phi}}=0bold_italic_ϕ = 0 (i.e., free theory), which is consistent with Ref. [43].

Specifically, for the SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) Yang-Mills theory, the background gauge field is AB4=(ϕ1T3+ϕ2T8)/gβsubscript𝐴B4subscriptitalic-ϕ1superscript𝑇3subscriptitalic-ϕ2superscript𝑇8𝑔𝛽A_{\text{B}4}=(\phi_{1}T^{3}+\phi_{2}T^{8})/g\betaitalic_A start_POSTSUBSCRIPT B 4 end_POSTSUBSCRIPT = ( italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT ) / italic_g italic_β, where T3superscript𝑇3T^{3}italic_T start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and T8superscript𝑇8T^{8}italic_T start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT constitute the Cartan subalgebra of 𝔰𝔲(3)𝔰𝔲3\mathfrak{su}(3)fraktur_s fraktur_u ( 3 ). The traced Polyakov loop in the fundamental representation of SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) is

|L|=13|trexp(ig0βAB4𝑑τmissing)|=134cos2(ϕ12)+4cos(ϕ12missing)cos(3ϕ22missing)+1.𝐿13tracei𝑔superscriptsubscript0𝛽subscript𝐴B4differential-d𝜏missing134superscript2subscriptitalic-ϕ124subscriptitalic-ϕ12missing3subscriptitalic-ϕ22missing1|L|=\frac{1}{3}\biggl{|}\tr\exp\biggl(\mathrm{i}g\int_{0}^{\beta}A_{\text{B}4}% \,d\tau\biggr{missing})\biggr{|}=\frac{1}{3}\sqrt{4\cos^{2}\Bigl{(}\frac{\phi_% {1}}{2}\Bigr{)}+4\cos\Bigl(\frac{\phi_{1}}{2}\Bigr{missing})\cos\Bigl(\frac{% \sqrt{3}\phi_{2}}{2}\Bigr{missing})+1}\,.| italic_L | = divide start_ARG 1 end_ARG start_ARG 3 end_ARG | roman_tr roman_exp ( start_ARG roman_i italic_g ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT B 4 end_POSTSUBSCRIPT italic_d italic_τ roman_missing end_ARG ) | = divide start_ARG 1 end_ARG start_ARG 3 end_ARG square-root start_ARG 4 roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) + 4 roman_cos ( start_ARG divide start_ARG italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG roman_missing end_ARG ) roman_cos ( start_ARG divide start_ARG square-root start_ARG 3 end_ARG italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG roman_missing end_ARG ) + 1 end_ARG . (9)

In our previous work [36], we showed that confinement, |L|=0𝐿0|L|=0| italic_L | = 0, is realized for Ω~Iπ/2subscript~ΩI𝜋2\tilde{\Omega}_{\text{I}}\geq\pi/2over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ≥ italic_π / 2 at r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0.

Refer to captionRefer to caption
Figure 1: (Left) Polyakov loop |L|𝐿|L|| italic_L | for SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) as a function of dimensionless radial distance r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG and dimensionless imaginary angular velocity Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT. (Right) The same plot of |L|𝐿|L|| italic_L | for SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ). The confined phase with |L|=0𝐿0|L|=0| italic_L | = 0 is bounded by the first-order phase transitions.

It is a straightforward exercise to find the global minima of the potential (8) for r~0~𝑟0\tilde{r}\neq 0over~ start_ARG italic_r end_ARG ≠ 0. Figure 1 shows the results from such extensive analyses of Eq. (8) for SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) (left) and SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) (right). It is notable that both cases generally develop r𝑟ritalic_r-dependent spatial structures. For the SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) case as shown in the left of Fig. 1, the Polyakov loop changes to zero, indicating confinement, with second-order phase transition as Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT grows up. In the previous paper [36], we focused on two edges of r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0 and Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π only. Our present results imply that, for Ω~I3π/4similar-to-or-equalssubscript~ΩI3𝜋4\tilde{\Omega}_{\text{I}}\simeq 3\pi/4over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ≃ 3 italic_π / 4 for example, there should appear a spatial interface separating the confined phase for r~0.5less-than-or-similar-to~𝑟0.5\tilde{r}\lesssim 0.5over~ start_ARG italic_r end_ARG ≲ 0.5 and the deconfined phase for r~0.5greater-than-or-equivalent-to~𝑟0.5\tilde{r}\gtrsim 0.5over~ start_ARG italic_r end_ARG ≳ 0.5. We can confirm a qualitatively similar trend for the SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) case, but the detailed shape looks different as shown in the right of Fig. 1. In this case of SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ), the homogeneously confined phase is realized around Ω~I3π/4less-than-or-similar-tosubscript~ΩI3𝜋4\tilde{\Omega}_{\text{I}}\lesssim 3\pi/4over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ≲ 3 italic_π / 4, and the interface between confinement and deconfinement emerges only when Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT becomes larger.

We should emphasize that this r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG dependence of inhomogeneity is qualitatively consistent with the lattice-QCD results [29]. As closely discussed in our previous work [36], our Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT dependence of the Polyakov loop is opposite to the lattice-QCD case; |L|𝐿|L|| italic_L | goes smaller with larger imaginary rotation in our perturbative calculation, while |L|𝐿|L|| italic_L | goes smaller with larger real rotation in the lattice-QCD simulation. However, in the both cases of ours and the lattice-QCD calculations, |L|𝐿|L|| italic_L | becomes larger (smaller) with larger r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG for a given imaginary (real) angular velocity. This makes a sharp contrast to preceding works [30, 14, 15, 27, 28] and might be a key observation to resolve the controversy of rotating QCD matter.

Refer to caption
Figure 2: SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) effective potential as a function of ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT for Ω~I=π/2subscript~ΩI𝜋2\tilde{\Omega}_{\text{I}}=\pi/2over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π / 2 (left) and Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π (right) with r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0 fixed. The dark (light) colored region has smaller (larger) potential values. The triangular domain indicated by the red line is sufficient for the minimum search.

Now, let us turn our attention to the Polyakov loop potential and the symmetry properties. Since we sum up all the roots in ΦΦ\Phiroman_Φ, different (ϕ1,ϕ2)subscriptitalic-ϕ1subscriptitalic-ϕ2(\phi_{1},\phi_{2})( italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) pairs can have the same potential and the Polyakov loop value according to Weyl symmetry of the SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) root lattice. We can see the characteristic patterns of the potential minima in Fig. 2. The repetition of the minima reflects Weyl symmetry. The red triangle region with three edges, (ϕ1,ϕ2)=(0,0)subscriptitalic-ϕ1subscriptitalic-ϕ200(\phi_{1},\phi_{2})=(0,0)( italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ( 0 , 0 ), (2π,2π/3)2𝜋2𝜋3(2\pi,2\pi/\sqrt{3})( 2 italic_π , 2 italic_π / square-root start_ARG 3 end_ARG ), (2π,2π/3)2𝜋2𝜋3(2\pi,-2\pi/\sqrt{3})( 2 italic_π , - 2 italic_π / square-root start_ARG 3 end_ARG ), is the fundamental domain and we can identify the state of matter from the minimum inside this triangular domain. The triangle has S6subscript𝑆6S_{6}italic_S start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT geometric symmetry, including three-fold rotational symmetry, which manifests center symmetry, and two-fold reflective symmetry, which manifests charge conjugation. The potential shape visualized by the pattern in Fig. 2 does not change by 2π/32𝜋32\pi/32 italic_π / 3 rotation but the phase of the Polyakov loop, L𝐿Litalic_L, does. The center of the triangle at (4π/3,0)4𝜋30(4\pi/3,0)( 4 italic_π / 3 , 0 ) corresponds to L=0𝐿0L=0italic_L = 0, that is, a center symmetric vacuum. Although the potential minima may break center symmetry, the charge conjugation symmetry is never broken.

We point out a nontrivial observation in Fig. 2; an emergent symmetry is realized at Ω~I=π/2subscript~ΩI𝜋2\tilde{\Omega}_{\text{I}}=\pi/2over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π / 2. We observe a reflective mirror on the line of ϕ1=πsubscriptitalic-ϕ1𝜋\phi_{1}=\piitalic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_π. This emergent symmetry comes from the vanishing of odd-n𝑛nitalic_n terms in the one-loop potential (8) at Ω~I=π/2subscript~ΩI𝜋2\tilde{\Omega}_{\text{I}}=\pi/2over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π / 2 and exists for not only r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0 but any radius. It could be either a one-loop artifact or a genuine symmetry. In the latter case, it has to be a non-invertible symmetry like that in the 2D critical Ising model since it exchanges the unbroken and broken vacua.

Refer to caption
Figure 3: SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) effective potential for r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0 (left) and r~=0.5~𝑟0.5\tilde{r}=0.5over~ start_ARG italic_r end_ARG = 0.5 (right) with Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π fixed.

In this work, we also quantify the spatial inhomogeneity. In Fig. 3, we plot the Polyakov loop potential for r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0 (confined phase) and r~=0.5~𝑟0.5\tilde{r}=0.5over~ start_ARG italic_r end_ARG = 0.5 (deconfined phase) at Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π in the SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) case. It is intriguing that the potential minima are located in a different way from the ordinary perturbative vacuum where three vertices of the triangle minimize the potential. In this sense, this deconfined phase discovered in the upper right region in the right panel of Fig. 1 may have exotic properties different from the ordinary one.

We mention the difference from the previous arguments [27] based on the Tolman-Ehrenfest (TE) effect. In our calculation, we treat T𝑇Titalic_T as a Lagrange multiplier to conserve the total energy. By construction, T𝑇Titalic_T is a global variable without r𝑟ritalic_r dependence. Therefore, in our study, we do not need to introduce the apparent temperature as a result of the TE effect. Nevertheless, the calculations naturally lead to r𝑟ritalic_r dependent structures.

3 Chiral symmetry breaking in the perturbatively confined phase

We can repeat similar calculations including dynamical quark contributions that break center symmetry explicitly. We can also address a relation between confinement and chiral symmetry breaking from two (approximate) order parameters, namely, the Polyakov loop and the dynamical quark mass, m𝑚mitalic_m, which is rooted in the chiral condensate. In the perturbative treatment, the pressure (the free energy) is maximized (minimized) for m=0𝑚0m=0italic_m = 0, and the dynamical mass generation is energetically disfavored. It is quite interesting what would happen in the perturbatively confined phase with imaginary rotation.

The theoretical treatments of fermions in the rotating frame are found in Refs. [30, 44, 7]. We should be careful about the fact that the Polyakov loop coupling with quarks is given by the fundamental representation. We can calculate the fermionic partition function by imposing an aperiodic thermal boundary condition or equivalently considering the ordinary anti-periodic boundary condition in the rotating frame. In this paper, for fermions, we choose the latter and the fermionic partition function is the determinant of the Dirac operator γμGBμ+msuperscript𝛾𝜇subscript𝐺B𝜇𝑚\gamma^{\mu}G_{{\rm B}\,\mu}+mitalic_γ start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT italic_G start_POSTSUBSCRIPT roman_B italic_μ end_POSTSUBSCRIPT + italic_m in the rotating frame, i.e.,

𝒵fT,ω=Det(γμGBμ+m).subscript𝒵f𝑇𝜔Detsuperscript𝛾𝜇subscript𝐺B𝜇𝑚\mathcal{Z}_{{\rm f}T,\omega}=\mathrm{Det}(\gamma^{\mu}G_{{\rm B}\,\mu}+m)\,.caligraphic_Z start_POSTSUBSCRIPT roman_f italic_T , italic_ω end_POSTSUBSCRIPT = roman_Det ( italic_γ start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT italic_G start_POSTSUBSCRIPT roman_B italic_μ end_POSTSUBSCRIPT + italic_m ) . (10)

Here, GBμ=DμΓμsubscript𝐺B𝜇subscript𝐷𝜇subscriptΓ𝜇G_{{\rm B}\,\mu}=D_{\mu}-\Gamma_{\mu}italic_G start_POSTSUBSCRIPT roman_B italic_μ end_POSTSUBSCRIPT = italic_D start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT - roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT is the covariant derivative including the AB4subscript𝐴B4A_{{\rm B}4}italic_A start_POSTSUBSCRIPT B4 end_POSTSUBSCRIPT background field with Γμ=i4σijωμijsubscriptΓ𝜇𝑖4superscript𝜎𝑖𝑗subscript𝜔𝜇𝑖𝑗\Gamma_{\mu}=-\frac{i}{4}\sigma^{ij}\,\omega_{\mu ij}roman_Γ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = - divide start_ARG italic_i end_ARG start_ARG 4 end_ARG italic_σ start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT italic_μ italic_i italic_j end_POSTSUBSCRIPT, where σij=i2[γ^i,γ^j]superscript𝜎𝑖𝑗𝑖2superscript^𝛾𝑖superscript^𝛾𝑗\sigma^{ij}=\frac{i}{2}[\hat{\gamma}^{i},\hat{\gamma}^{j}]italic_σ start_POSTSUPERSCRIPT italic_i italic_j end_POSTSUPERSCRIPT = divide start_ARG italic_i end_ARG start_ARG 2 end_ARG [ over^ start_ARG italic_γ end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT , over^ start_ARG italic_γ end_ARG start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ] and ωμij=gρσeiρ(μejσ+Γμνσejν)subscript𝜔𝜇𝑖𝑗subscript𝑔𝜌𝜎superscriptsubscript𝑒𝑖𝜌subscript𝜇subscriptsuperscript𝑒𝜎𝑗subscriptsuperscriptΓ𝜎𝜇𝜈subscriptsuperscript𝑒𝜈𝑗\omega_{\mu ij}=g_{\rho\sigma}\,e_{i}^{\ \rho}\,\quantity(\partial_{\mu}e^{\ % \sigma}_{j}+\Gamma^{\sigma}_{\mu\nu}\,e^{\ \nu}_{j})italic_ω start_POSTSUBSCRIPT italic_μ italic_i italic_j end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT italic_ρ italic_σ end_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_ρ end_POSTSUPERSCRIPT ( start_ARG ∂ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + roman_Γ start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ). We denote the gamma matrices of the flat space-time by γ^isuperscript^𝛾𝑖\hat{\gamma}^{i}over^ start_ARG italic_γ end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT and γμ=eiμγ^isuperscript𝛾𝜇superscriptsubscript𝑒𝑖𝜇superscript^𝛾𝑖\gamma^{\mu}=e_{i}^{\ \mu}\hat{\gamma}^{i}italic_γ start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT = italic_e start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_μ end_POSTSUPERSCRIPT over^ start_ARG italic_γ end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT. After all, we arrive at the expression for the Polyakov loop potential per one fermion (particle or anti-particle but without flavor degrees of freedom) as

Vf(ϕ;Ω~I)subscript𝑉𝑓bold-italic-ϕsubscript~ΩI\displaystyle V_{f}({\bm{\phi}};\tilde{\Omega}_{\text{I}})italic_V start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( bold_italic_ϕ ; over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ) =T4π2𝝁Φfl0kdkdkzabsent𝑇4superscript𝜋2subscript𝝁subscriptΦ𝑓subscript𝑙superscriptsubscript0subscript𝑘perpendicular-todifferential-dsubscript𝑘perpendicular-tosuperscriptsubscriptdifferential-dsubscript𝑘𝑧\displaystyle=-\frac{T}{4\pi^{2}}\sum_{\bm{\mu}\in\Phi_{f}}\sum_{l\in\mathbb{Z% }}\int_{0}^{\infty}k_{\perp}\,\mathrm{d}k_{\perp}\int_{-\infty}^{\infty}% \mathrm{d}k_{z}= - divide start_ARG italic_T end_ARG start_ARG 4 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∑ start_POSTSUBSCRIPT bold_italic_μ ∈ roman_Φ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_l ∈ blackboard_Z end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT roman_d italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT
×[Jl2(kr)+Jl+12(kr)]Reln(1+eiϕ𝝁iΩ~I(l+1/2)β𝒌2+m2missing),absentdelimited-[]superscriptsubscript𝐽𝑙2subscript𝑘perpendicular-to𝑟superscriptsubscript𝐽𝑙12subscript𝑘perpendicular-to𝑟1superscripteibold-italic-ϕ𝝁isubscript~ΩI𝑙12𝛽superscript𝒌2superscript𝑚2missing\displaystyle\qquad\times\Bigl{[}J_{l}^{2}(k_{\perp}r)+J_{l+1}^{2}(k_{\perp}r)% \Bigr{]}\real\ln\Bigl(1+\mathrm{e}^{\mathrm{i}{\bm{\phi}}\cdot\bm{\mu}-\mathrm% {i}\tilde{\Omega}_{\text{I}}(l+1/2)-\beta\sqrt{\bm{k}^{2}+m^{2}}}\Bigr{missing% })\,,× [ italic_J start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_r ) + italic_J start_POSTSUBSCRIPT italic_l + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_k start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT italic_r ) ] start_OPERATOR roman_Re end_OPERATOR roman_ln ( start_ARG 1 + roman_e start_POSTSUPERSCRIPT roman_i bold_italic_ϕ ⋅ bold_italic_μ - roman_i over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ( italic_l + 1 / 2 ) - italic_β square-root start_ARG bold_italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT roman_missing end_ARG ) , (11)

where ΦfsubscriptΦ𝑓\Phi_{f}roman_Φ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT denotes the set of the fundamental weights of 𝔰𝔲(Nc)𝔰𝔲subscript𝑁c\mathfrak{su}(N_{\text{c}})fraktur_s fraktur_u ( italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ). The limit of ϕ=0bold-italic-ϕ0{\bm{\phi}}=0bold_italic_ϕ = 0 renders the above expression to the free quark and anti-quark gas energy with imaginary rotation derived from statistical mechanics.

First, we shall consider the interplay of |L|𝐿|L|| italic_L | and m𝑚mitalic_m at r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0 by setting Nc=Nf=2subscript𝑁csubscript𝑁f2N_{\text{c}}=N_{\text{f}}=2italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT f end_POSTSUBSCRIPT = 2. Then, the total effective potential is given by Vg+2NfVfsubscript𝑉𝑔2subscript𝑁fsubscript𝑉𝑓V_{g}+2N_{\text{f}}V_{f}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT + 2 italic_N start_POSTSUBSCRIPT f end_POSTSUBSCRIPT italic_V start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT (where 2222 comes from the sum over particle and anti-particle). In our numerical calculation, we consider ϕ[0,2π]italic-ϕ02𝜋\phi\in[0,2\pi]italic_ϕ ∈ [ 0 , 2 italic_π ] and m~=m/T[0,10]~𝑚𝑚𝑇010\tilde{m}=m/T\in[0,10]over~ start_ARG italic_m end_ARG = italic_m / italic_T ∈ [ 0 , 10 ]. We note that the former region for ϕitalic-ϕ\phiitalic_ϕ is sufficient thanks to SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) Weyl symmetry. Because we do not model quark interactions to keep our analysis as model independent as possible, m𝑚mitalic_m blows up in the phase where chiral symmetry is fully broken. We thus set a numerical cutoff on m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG in the practical calculation.

Refer to captionRefer to caption
Figure 4: Potential contour as functions of ϕitalic-ϕ\phiitalic_ϕ (Polyakov loop) and m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG (dimensionless dynamical mass) for Nc=Nf=2subscript𝑁csubscript𝑁f2N_{\text{c}}=N_{\text{f}}=2italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT f end_POSTSUBSCRIPT = 2, Ω~I=π/2subscript~ΩI𝜋2\tilde{\Omega}_{\text{I}}=\pi/2over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π / 2 (left) and Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π (right). The darker (lighter) color indicates the smaller (larger) value and the potential minimum has the darkest color.

Figure 4 shows the potential contour as functions of two order parameters in the SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) case. The minimum of the potential is given by the darkest spot in the figures. In the left of Fig. 4 for Ω~I=π/2subscript~ΩI𝜋2\tilde{\Omega}_{\text{I}}=\pi/2over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π / 2, the minimum is still located at m~=0~𝑚0\tilde{m}=0over~ start_ARG italic_m end_ARG = 0 and ϕ=0italic-ϕ0\phi=0italic_ϕ = 0 (i.e., L=1𝐿1L=1italic_L = 1). This is different from the pure gluonic case in the left of Fig. 1; fermions generally favor deconfinement. Interestingly in the right of Fig. 4 for Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π, however, the minimum moves to ϕπsimilar-to-or-equalsitalic-ϕ𝜋\phi\simeq\piitalic_ϕ ≃ italic_π and m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG blows up to the cutoff value, which signifies confinement and spontaneous chiral symmetry breaking.

Refer to captionRefer to caption
Figure 5: Order parameters, |L|𝐿|L|| italic_L | and m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG, as functions of Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT for Nf=2subscript𝑁f2N_{\text{f}}=2italic_N start_POSTSUBSCRIPT f end_POSTSUBSCRIPT = 2, Nc=2subscript𝑁c2N_{\text{c}}=2italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 2 (left) and Nc=3subscript𝑁c3N_{\text{c}}=3italic_N start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = 3 (right). The blue solid line represents the Polyakov loop |L|𝐿|L|| italic_L |, and the green dashed line represents the dynamical mass m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG.

We make the plots for the behavior of the Polyakov loop |L|𝐿|L|| italic_L | and the dynamical mass m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG in Fig. 5. The blue solid line represents |L|𝐿|L|| italic_L |, while the green dashed line represents m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG. With increasing Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT, the Polyakov loop goes smaller and the dynamical mass grows up. Then, a confined and chiral symmetry broken state is favored near Ω~Iπsimilar-to-or-equalssubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}\simeq\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT ≃ italic_π. For the SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) case as shown in the left of Fig. 5, the behavior is 2π2𝜋2\pi2 italic_π periodic because the SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) gauge group covers the fermionic parity ()Fsuperscript𝐹(-)^{F}( - ) start_POSTSUPERSCRIPT italic_F end_POSTSUPERSCRIPT. Also we see that the behavior is symmetric around Ω~I=πsubscript~ΩI𝜋\tilde{\Omega}_{\text{I}}=\piover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT = italic_π. For the SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) case in the right of Fig. 5, in contrast, the order parameters depend on Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT in a complicated way and the confined phase seems to be less favored. Also the periodicity becomes 4π4𝜋4\pi4 italic_π because the SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) gauge group does not contain ()Fsuperscript𝐹(-)^{F}( - ) start_POSTSUPERSCRIPT italic_F end_POSTSUPERSCRIPT.

Refer to captionRefer to caption
Figure 6: Order parameters in the SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) case; the Polyakov loop (left) and the dynamical mass (right) as functions of dimensionless radial distance r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG and dimensionless imaginary angular velocity Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT.

In the same way as the pure gluonic theory, we have quantified the inhomogeneous structures in Fig. 6. The left and the right panels show the Polyakov loop and the dynamical mass, respectively, as functions of r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG and Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT. We have performed this calculation for the SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) case only. As compared to the left of Fig. 1 in the pure gluonic case, the perturbatively confined region is shrunk to the upper left corner. Interestingly, in view of the right of Fig. 6, confinement and chiral symmetry breaking are locked together and the phase transitions occur simultaneously even at large imaginary rotation. In addition, there appear some substructures, but the numerical convergence seems to be insufficient. We have tried, but the potential shape is quite complicated and it is difficult to obtain smooth results in the whole region at r~0.5greater-than-or-equivalent-to~𝑟0.5\tilde{r}\gtrsim 0.5over~ start_ARG italic_r end_ARG ≳ 0.5. This complication might be a precursory of possible fractals [45]. More investigations about these substructures as well as the full SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) calculations are left as future problems.

4 Conclusions

We considered the inhomogeneous structures and the chiral symmetry breaking in the perturbatively confined phase which is realized by rotation with imaginary angular velocity, Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT. In both color SU(2)𝑆𝑈2SU(2)italic_S italic_U ( 2 ) and SU(3)𝑆𝑈3SU(3)italic_S italic_U ( 3 ) cases, the perturbatively confined phase is induced at the rotation center, r~=0~𝑟0\tilde{r}=0over~ start_ARG italic_r end_ARG = 0, for Ω~Isubscript~ΩI\tilde{\Omega}_{\text{I}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT I end_POSTSUBSCRIPT above a certain threshold. We found that the perturbative confinement may break down in the r~0~𝑟0\tilde{r}\neq 0over~ start_ARG italic_r end_ARG ≠ 0 region away from the rotation center. This indicates the existence of phase interface between confinement and deconfinement and the critical temperature is also r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG dependent accordingly. Our results predict that the critical temperature should decrease with increasing r~~𝑟\tilde{r}over~ start_ARG italic_r end_ARG in the presence of imaginary rotation. This trend agrees with the latest results from the lattice simulation. Although the analytical continuation to real rotation has some subtle points, the inhomogeneity we have confirmed from the perturbative effective potential can presumably persist in real rotating systems.

We then added the quark contribution to the Polyakov loop effective potential, which is also a function of the dynamical quark mass, m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG. Our analysis of the potential minimum search showed that confinement and chiral symmetry breaking are tightly correlated in such unconventional systems with imaginary rotation and even without fermionic interaction. In the presence of quarks that explicitly break center symmetry, the Polyakov loop is only an approximate order parameter. Still, in the region where the Polyakov loop is vanishingly small, m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG departs from zero signifying spontaneous breaking of chiral symmetry. In our treatment, in fact, m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG blows up when chiral symmetry is broken, so that large m~~𝑚\tilde{m}over~ start_ARG italic_m end_ARG suppresses center symmetry breaking. In this way, we established simultaneous realization of confinement and chiral symmetry breaking. We also found highly complicated substructures in the deconfined regions. The imaginary-rotating inhomogeneous states with dynamical quarks should deserve further investigations in the future.

Acknowledgement

The authors thank Maxim Chernodub and Xu-Guang Huang for useful discussions. This work was supported by Japan Society for the Promotion of Science (JSPS) KAKENHI Grant Nos. 22H01216 (K.F.) and 22H05118 (K.F.) and JST SPRING, Grant Number JPMJSP2108 (Y.S.).

References