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

aainstitutetext: School of Nuclear Science and Technology, University of Chinese Academy of Sciences, Beijing 100049, Chinabbinstitutetext: Few-Body Systems in Physics Laboratory, RIKEN Nishina Center, Wako 351-0198, Japanccinstitutetext: Research Center for Nuclear Physics, Osaka University, Ibaraki 567-0048, Japan

Fate of the topological susceptibility in two-color dense QCD

Mamiya Kawaguchi b,c    Daiki Suenaga [email protected] [email protected]
Abstract

We explore the topological susceptibility at finite quark chemical potential and zero temperature in two-color QCD (QC2D) with two flavors. Through the Ward-Takahashi identities of QC2D, we find that the topological susceptibility in the vacuum solely depends on three observables: the pion decay constant, the pion mass, and the η\eta mass in the low-energy regime of QC2D. Based on the identities, we numerically evaluate the topological susceptibility at finite quark chemical potential using the linear sigma model with the approximate Pauli-Gursey SU(4)SU(4) symmetry. Our findings indicate that, in the absence of U(1)AU(1)_{A} anomaly effects represented by the Kobayashi-Maskawa-’t Hooft-type determinant interaction, the topological susceptibility vanishes in both the hadronic and baryon superfluid phases. On the other hand, when the U(1)AU(1)_{A} anomaly effects are present, the constant and nonzero topological susceptibility is induced in the hadronic phase, reflecting the mass difference between the pion and η\eta meson. Meanwhile, in the superfluid phase it begins to decrease smoothly. The asymptotic behavior of the decrement is fitted by the continuous reduction of the chiral condensate in dense QC2D, which is similar to the behavior observed in hot three-color QCD matter. In addition, effects from the finite diquark source on the topological susceptibility are discussed. We expect that the present study provides a clue to shed light on the role of the U(1)AU(1)_{A} anomaly in cold and dense QCD matter.

1 Introduction

In Quantum Chromodynamics (QCD), the U(1)AU(1)_{A} anomaly, i.e., the non-conservation of the U(1)AU(1)_{A} axial current caused by the gluonic quantum corrections, plays crucial roles in the low-energy physics governed by the spontaneous breaking of the chiral symmetry. For instance, the U(1)AU(1)_{A} anomaly affects the hadron mass spectrum to yield the heavy η\eta^{\prime} meson [1] and the order of the chiral phase transition in QCD matter [2]. In addition to these low-energy aspects, the U(1)AU(1)_{A} anomaly is also closely related with topological vacuum structures of QCD [3], which is described by the anomalous gluonic operator tagged with the θ\theta parameter. The characteristics of the θ\theta-dependent QCD vacuum is captured by the topological susceptibility: the curvature of the QCD effective potential with respect to θ\theta.

The symmetry breakings in QCD are reflected in meson susceptibility functions defined by two-point functions of quark composite operators in the low-energy limit. At the hadronic level, the so-called chiral-partner structure would be an indication of this property [4]. That is, at high temperature and/or density where chiral symmetry tends to be restored, masses of the mesons related by the chiral transformation become degenerate, and so do the corresponding meson susceptibility functions. This implies that, indeed, the meson susceptibility functions can be regarded as alternative probes to measure the strength of chiral symmetry breaking and restoration. In a similar way, the effective restoration of U(1)AU(1)_{A} symmetry can be quantified by the degeneracies of the meson susceptibility functions connected by the U(1)AU(1)_{A} transformation.

Making use of the Ward-Takahashi identity (WTI) associated with chiral symmetry, one can show that the topological susceptibility is also correlated with the chiral- and U(1)AU(1)_{A}-partner structures in the meson susceptibility functions [5, 6, 7, 8, 9]. Thereby, the topological susceptibility can also be referred to as the indicator for the breaking strength of U(1)AU(1)_{A} symmetry through the chiral phase transition. In fact, lattice QCD simulations at the physical quark masses support that the magnitude of the topological susceptibility smoothly decreases at high temperatures [10, 11, 12]. Furthermore, a strong correlation with the chiral restoration has also been studied through the meson susceptibility functions within 22 flavor QCD [13, 14, 15, 16, 17] and 2+12+1 flavor QCD [18, 19, 20].

Thus far, the susceptibilities in hot QCD matter have been explored by both lattice simulations [13, 14, 18, 19, 20, 15, 10, 11, 12, 16, 17] and effective model analyses [7, 21, 8] in order to gain deeper insights into the symmetry properties of QCD in the extreme environment. However, at finite quark chemical potential μq\mu_{q} lattice QCD simulations with three colors suffer from the sign problem, and then the first-principle numerical computations cannot apply in baryonic matter straightforwardly [22]. For this reason, our understanding of QCD at low-temperature and high-density regime is still limited compared to that in hot medium.

In light of the difficulty of three-color QCD on lattice simulations with finite μq\mu_{q}, two-color QCD (QC2D{\rm QC_{2}D}) with two flavors provides us with a valuable testing ground. This is because the sign problem is resolved in such QCD-like theory owing to its pseudo-real property [23]. Focusing on this fact, many efforts from lattice simulations are being devoted to understandings of, e.g., phase structures, thermodynamics quantities, electromagnetic responses, the hadron mass spectrum, and gluon propagators in cold and dense QC2D matter [24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49]. In association with such numerical examinations, theoretical investigations of QC2D at finite μq\mu_{q} based on effective models have been done [50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62, 63, 64, 65].

In QC2D, diquarks composed of two quarks are treated as color-singlet baryons. In other words, baryons exhibit bosonic behavior similarly to mesons. Reflecting this fact in QC2D, SU(2)L×SU(2)RSU(2)_{L}\times SU(2)_{R} chiral symmetry is extended to the so-called Pauli-Gursey SU(4)SU(4) symmetry, which allows us to describe diquark baryons and light mesons in the single multiplets. Accordingly, the spontaneously symmetry-breaking pattern caused by the chiral condensate is changed to SU(4)Sp(4)SU(4)\to Sp(4) [50, 51]. Despite such an extension of chiral symmetry, symmetry structures of the U(1)AU(1)_{A} axial anomaly induced by gluonic configurations essentially do not differ from those in ordinary three-color QCD.

To shed light on the role of the U(1)AU(1)_{A} axial anomaly in baryonic QCD matter, the μq\mu_{q} dependence of the topological susceptibility has been numerically measured by lattice numerical simulations of QC2[30, 33, 41, 45]. The recent lattice result in Ref. [41] indicates that the effect of μq\mu_{q} does not exert any influence on the behavior of the topological susceptibility in baryonic matter, resulting in an approximately constant value. In contrast, the other group shows that the topological susceptibility is suppressed in high-density regions [45]. Hence, there exist discrepancies among the lattice simulations at finite quark chemical potentials, and the fate of topological susceptibility at high-density regions is still controversial.

In this paper, motivated by the above puzzle, we investigate the topological susceptibility in zero-temperature QC2D at finite μq\mu_{q} based on an effective-model approach. In particular, we employ the linear sigma model based on the approximate Pauli-Gursey SU(4)SU(4) symmetry invented in Ref. [65]. Notably, this model is capable of treating the η\eta meson which plays a significant role in describing the U(1)AU(1)_{A} anomaly structures consistently with other light mesons and diquark baryons. In QC2D, since the diquarks obey the Bose-Einstein statistics, when the mass of the ground-state diquark becomes zero they begin to exhibit the Bose-Einstein condensates (BECs), leading to the emergence of the diquark condensed phase [50, 51]. This phase is also referred to as the baryon superfluid phase due to the violation of U(1)BU(1)_{B} baryon-number symmetry. Meanwhile, the stable phase with no such BECs connected to the vacuum, i.e., zero temperature and zero chemical potential, is called the hadronic phase. The former nontrivial phase triggers a rich hadron mass spectrum such as a mixing among hadrons sharing the identical quantum numbers except for the baryon number.

Within the linear sigma model, the influences of the U(1)AU(1)_{A} anomaly on hadrons are described by the so-called Kobayashi-Maskawa-’t Hooft (KMT)-type determinant interaction [66, 67, 68, 69], which only breaks U(1)AU(1)_{A} symmetry but preserves the Pauli-Gursey SU(4)SU(4) one. This interaction induces a mass difference between the pion and η\eta meson in the vacuum. Thus, in the present analysis we particularly focus on the strength of the KMT-type interaction, in other words, the mass difference between the pion and η\eta meson, in order to quantify roles of the U(1)AU(1)_{A} anomaly in the topological susceptibility. Besides, in lattice simulations source contributions with respect to the diquark condensate would be left sizable, so in this paper we also investigate the diquark source effects so as to facilitate the comparison with lattice data.

This paper is organized as follows. In Sec. 2 we present general properties associated with the topological susceptibility in QC2D by focusing on the underlying QC2D theory, and discuss symmetry partner structures of the meson susceptibility functions. In Sec. 3, the emergence of the Pauli-Gursey SU(4)SU(4) symmetry in QC2D is briefly explained, and our linear sigma model regarded as a low-energy theory of QC2D is introduced. In Sec. 4 we show how the topological susceptibility within the linear sigma model is evaluated by explicitly demonstrating the matching between underlying QC2D and the linear sigma model. Based on it, in Sec. 5 we show our numerical results on the topological susceptibility at finite μq\mu_{q}. In order to facilitate the comparison with lattice simulations, in Sec. 6 we also exhibit the results in the presence of the diquark source contributions. Finally, in Sec. 7 we conclude our present study.

2 Topological susceptibility based on Ward-Takahashi identities of QC2D

Our main aim of this paper is to reveal properties of the topological susceptibility in zero-temperature QC2D with finite quark chemical potential μq\mu_{q}. In this section, we present an analytic formula of the topological susceptibility based on the underlying QC2D Lagrangian [5, 6, 7, 8, 9], which is useful for the investigation within the effective-model framework of the linear sigma model.

The topological susceptibility is one of indicators to measure the magnitude of the U(1)AU(1)_{A} anomaly, which is related to nontrivial gluonic configurations such as the instantons [3]. Hence, we need to return to QCD Lagrangian where such microscopic degrees of freedom are treated manifestly. In two-flavor QC2D, the Lagrangian including the so-called QCD θ\theta-term in Minkowski spacetime is of the form

QC2D=ψ¯(iγμDμml)ψ14GμνaGμν,a+θg264π2ϵμνρσGμνaGρσa.\displaystyle{\cal L}_{\rm QC_{2}D}=\bar{\psi}(i\gamma^{\mu}D_{\mu}-m_{l})\psi-\frac{1}{4}G_{\mu\nu}^{a}G^{\mu\nu,a}+\theta\frac{g^{2}}{64\pi^{2}}\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a}\ . (2.1)

As for the first term, ψ=(u,d)T\psi=(u,d)^{T} denotes the two-flavor quark doublet and Dμψ=(μiμqδμ0igAμaTca)ψD_{\mu}\psi=(\partial_{\mu}-i\mu_{q}\delta_{\mu 0}-igA_{\mu}^{a}T_{c}^{a})\psi is the covariant derivative incorporating effects from a quark chemical potential μq\mu_{q} and interactions with a gluon field AμaA_{\mu}^{a}. The 2×22\times 2 matrix Tca=τca/2T_{c}^{a}=\tau_{c}^{a}/2 is the generator of SU(2)cSU(2)_{c} color group with τca\tau_{c}^{a} being the Pauli matrix. Besides, gg and mlm_{l} are the QCD coupling constant and a current quark mass where the isospin symmetric limit is taken, mu=mdmlm_{u}=m_{d}\equiv m_{l}. The second term in Eq. (2.1) is a gluon kinetic term where Gμνa=μAνaνAμa+gϵabcAμbAνbG_{\mu\nu}^{a}=\partial_{\mu}A_{\nu}^{a}-\partial_{\nu}A_{\mu}^{a}+g\epsilon^{abc}A_{\mu}^{b}A_{\nu}^{b} is the field strength of gluons. The last ingredient of QC2D{\rm QC_{2}D} Lagrangian in Eq. (2.1) is the θ\theta-term of QC2D, which is described by a flavor-singlet topological operator Q(g2/64π2)ϵμνρσGμνaGρσaQ\equiv(g^{2}/64\pi^{2})\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a} tagged with the θ\theta-parameter. Our purpose in this subsection is to derive useful identities with respect to the topological susceptibility, so that only the θ\theta-dependent term in Eq. (2.1), which is gauge invariant, plays significant roles. For this reason, the gauge-fixing terms and the corresponding Faddeev-Popov determinant, which do not affect the following discussions, have been omitted in Eq. (2.1).

The generating functional of QC2D{\cal L}_{\rm QC_{2}D} in the path-integral formulation is given by

ZQC2D=[dψ¯dψ][dA]exp[id4xQC2D],\displaystyle Z_{\rm QC_{2}D}=\int[d\bar{\psi}d\psi][dA]\exp\Biggl{[}i\int d^{4}x{\cal L}_{\rm QC_{2}D}\Biggl{]}\ , (2.2)

and the θ\theta-dependent effective action of QC2D{\rm QC_{2}D} is evaluated as

ΓQC2D=ilnZQC2D.\displaystyle\Gamma_{{\rm QC_{2}D}}=-i\ln Z_{{\rm QC_{2}D}}\ . (2.3)

The topological susceptibility χtop\chi_{\rm top} is defined by the curvature of ΓQC2D\Gamma_{{\rm QC_{2}D}}, i.e., a second derivative with respect to θ\theta at θ=0\theta=0:

χtop\displaystyle\chi_{\rm top} =\displaystyle= d4xδ2ΓQC2Dδθ(x)δθ(0)|θ=0.\displaystyle-\int d^{4}x\frac{\delta^{2}\Gamma_{{\rm QC_{2}D}}}{\delta\theta(x)\delta\theta(0)}\Biggl{|}_{\theta=0}. (2.4)

Thus, from a straightforward calculation of Eq. (2.4) based on the QC2D Lagrangian in Eq. (2.1), one can find that the topological susceptibility is described by a two-point correlation function of the topological operator Q=(g2/64π2)ϵμνρσGμνaGρσaQ=(g^{2}/64\pi^{2})\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a}:

χtop\displaystyle\chi_{\rm top} =\displaystyle= id4x0|TQ(x)Q(0)|0,\displaystyle-i\int d^{4}x\langle 0|TQ(x)Q(0)|0\rangle\ , (2.5)

with TT denoting the time ordered product. It should be noted that contributions stemming from a product of Q\langle Q\rangle have been omitted from Eq. (2.5) due to the parity conservation. The topological susceptibility in Eq. (2.5) is written in terms of the gluonic operator QQ, which would not be a manageable expression since our task in this paper is to evaluate χtop\chi_{\rm top} from a low-energy effective model involving only hadronic degrees of freedom. Difficulties in matching the susceptibility from the effective models with that from underlying QC2D are, however, remedied by utilizing the U(1)AU(1)_{A} axial rotation properly as demonstrated below.

Under the U(1)AU(1)_{A} rotation with a rotation angle αA\alpha_{A}, the quark doublet transforms as

ψexp(iαA/2γ5)ψ.\displaystyle\psi\to\exp(i\alpha_{A}/2\,\gamma_{5})\psi\ . (2.6)

Meanwhile, within the path-integral formalism the gluonic quantum anomaly is generated by the fermionic measure [dψ¯dψ][d\bar{\psi}d\psi] according to the Fujikawa’s method [70], resulting in that the rotated generating functional reads

ZQC2D\displaystyle Z_{\rm QC_{2}D} \displaystyle\to [dψ¯dψ][dA]exp[id4x(ψ¯iγμDμψmlψ¯exp(iαAγ5)ψ\displaystyle\int[d\bar{\psi}d\psi][dA]\exp\Biggl{[}i\int d^{4}x\Biggl{(}\bar{\psi}i\gamma^{\mu}D_{\mu}\psi-m_{l}\bar{\psi}\exp(i\alpha_{A}\gamma_{5})\psi (2.7)
14GμνaGμν,a+(θ2αA)g264π2ϵμνρσGμνaGρσa)].\displaystyle-\frac{1}{4}G_{\mu\nu}^{a}G^{\mu\nu,a}+(\theta-2\alpha_{A})\frac{g^{2}}{64\pi^{2}}\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a}\Biggl{)}\Biggl{]}\ .

Thus, when choosing the rotation angle to be αA=θ/2\alpha_{A}=\theta/2, the θ\theta-dependence of the QCD θ\theta-term is transferred into the quark mass term as

ZQC2D\displaystyle Z_{\rm QC_{2}D} =\displaystyle= [dψ¯dψ][dA]exp[id4x(ψ¯iγμDμψmlψ¯exp(iθ/2γ5)ψ14GμνaGμν,a)].\displaystyle\int[d\bar{\psi}d\psi][dA]\exp\Biggl{[}i\int d^{4}x\Biggl{(}\bar{\psi}i\gamma^{\mu}D_{\mu}\psi-m_{l}\bar{\psi}\exp\left(i\theta/2\,\gamma_{5}\right)\psi-\frac{1}{4}G_{\mu\nu}^{a}G^{\mu\nu,a}\Biggl{)}\Biggl{]}\ .

Following the procedure in Eqs. (2.3) and (2.4) with the rotated generating functional (LABEL:gene_func_mass_theta), the topological susceptibility χtop\chi_{\rm top} is now expressed by fermionic operators as

χtop=14[mlψ¯ψ+iml2χη],\displaystyle\chi_{\rm top}=-\frac{1}{4}\left[m_{l}\langle\bar{\psi}\psi\rangle+im_{l}^{2}\chi_{\eta}\right]\ , (2.9)

where ψ¯ψ\langle\bar{\psi}\psi\rangle is the chiral condensate serving as an order parameter of the spontaneous chiral-symmetry breaking, and χη\chi_{\eta} denotes an η\eta-meson susceptibility function defined by

χη=d4x0|T(iψ¯γ5ψ)(x)(iψ¯γ5ψ)(0)|0.\displaystyle\chi_{\eta}=\int d^{4}x\langle 0|T(i\bar{\psi}\gamma_{5}\psi)(x)(i\bar{\psi}\gamma_{5}\psi)(0)|0\rangle\ . (2.10)

It should be noted that, from the rotated generating functional in Eq. (2.7), a non-conservation law of the U(1)AU(1)_{A} axial current jAμ=ψ¯γμγ5ψj^{\mu}_{A}=\bar{\psi}\gamma^{\mu}\gamma_{5}\psi is also obtained as

μjAμ=2mlψ¯iγ5ψ+g216π2ϵμνρσGμνaGρσa.\displaystyle\partial_{\mu}j_{A}^{\mu}=2m_{l}\bar{\psi}i\gamma_{5}\psi+\frac{g^{2}}{16\pi^{2}}\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a}\ . (2.11)

The topological susceptibility (2.9) is further reduced to a handleable form. In fact, using the SU(2)L×SU(2)RSU(2)_{L}\times SU(2)_{R} chiral-partner relation shown in Appendix A, a chiral WTI with respect to the chiral condensate ψ¯ψ\langle\bar{\psi}\psi\rangle is derived as in Appendix B, which reads

ψ¯ψ=imlχπ.\displaystyle\langle\bar{\psi}\psi\rangle=-im_{l}\chi_{\pi}\ . (2.12)

In this identity, χπ\chi_{\pi} is a pion susceptibility function defined by

χπδab=d4x0|T(iψ¯γ5τfaψ)(x)(iψ¯γ5τfbψ)(0)|0,\displaystyle\chi_{\pi}\delta^{ab}=\int d^{4}x\langle 0|T(i\bar{\psi}\gamma_{5}\tau^{a}_{f}\psi)(x)(i\bar{\psi}\gamma_{5}\tau^{b}_{f}\psi)(0)|0\rangle\ , (2.13)

with τfa\tau^{a}_{f} being the Pauli matrix in the flavor space. Therefore, inserting Eq. (2.12) into Eq. (2.9), the topological susceptibility is found to be determined in terms of a difference of χπ\chi_{\pi} and χη\chi_{\eta} as

χtop=iml24(χπχη).\displaystyle\chi_{\rm top}=\frac{im_{l}^{2}}{4}(\chi_{\pi}-\chi_{\eta})\ . (2.14)

This expression is identical to the one obtained in ordinary three-color QCD through the WTI [5, 6, 7, 8, 9]. Here, to facilitate an understanding of the role of topological susceptibility, we insert the scalar meson susceptibilities χσ\chi_{\sigma} and χa0\chi_{a_{0}} in Eq. (2.14):

χtop\displaystyle\chi_{\rm top} =\displaystyle= iml24[(χπχσ)(χηχσ)],\displaystyle\frac{im_{l}^{2}}{4}\left[(\chi_{\pi}-\chi_{\sigma})-(\chi_{\eta}-\chi_{\sigma})\right]\ ,
χtop\displaystyle\chi_{\rm top} =\displaystyle= iml24[(χπχa0)(χηχa0)],\displaystyle\frac{im_{l}^{2}}{4}\left[(\chi_{\pi}-\chi_{a_{0}})-(\chi_{\eta}-\chi_{a_{0}})\right]\ , (2.15)

where χσ\chi_{\sigma} and χa0\chi_{a_{0}} are the susceptibility functions made of the composite operators ψ¯ψ\bar{\psi}\psi and ψ¯τfaψ\bar{\psi}\tau_{f}^{a}\psi, respectively. Indeed, under the chiral SU(2)L×SU(2)RSU(2)_{L}\times SU(2)_{R} rotation and the U(1)AU(1)_{A} rotation, the meson susceptibility functions are transformed into each other:

χπ{\chi_{\pi}}χσ{\chi_{\sigma}}χa0{\chi_{a_{0}}}χη{\chi_{\eta}}U(1)AU(1)_{A}SU(2)SU(2)U(1)AU(1)_{A}

as explicitly shown in Appendix A. With this transformation, one can realize that the topological susceptibility in Eq. (2.15) is described by the combinations of the chiral SU(2)SU(2) partner χπχσ\chi_{\pi}\leftrightarrow\chi_{\sigma} (χa0χη\chi_{a_{0}}\leftrightarrow\chi_{\eta}) and the U(1)U(1) axial partner χπχa0\chi_{\pi}\leftrightarrow\chi_{a_{0}} (χσχη\chi_{\sigma}\leftrightarrow\chi_{\eta}) . When chiral symmetry is restored and the order parameter of the spontaneous chiral symmetry breaking vanishes ψ¯ψ0\langle\bar{\psi}\psi\rangle\to 0, the chiral partner becomes (approximately) degenerate:

SU(2)L×SU(2)R restoration limit:\displaystyle\mbox{$SU(2)_{L}\times SU(2)_{R}$ restoration limit}: {χπχσ0χa0χη0.\displaystyle\begin{cases}\chi_{\pi}-\chi_{\sigma}\to 0\\ \chi_{a_{0}}-\chi_{\eta}\to 0\end{cases}. (2.16)

After the chiral restoration, the topological susceptibility is dominated by the U(1)AU(1)_{A} axial partner: χtopχηχσ\chi_{\rm top}\sim\chi_{\eta}-\chi_{\sigma} (χtopχπχa0\chi_{\rm top}\sim\chi_{\pi}-\chi_{a_{0}}), so that χtop\chi_{\rm top} acts as the indicator for the breaking strength of U(1)AU(1)_{A} symmetry. It should be noted that χtop\chi_{\rm top} trivially vanishes in the chiral limit (ml=0m_{l}=0) as seen from Eq. (2.14). In this limit, the topological susceptibility is no longer regarded as the indicator. This can also be understood by the fact that when ml=0m_{l}=0, the θ\theta dependence of the generating functional in Eq. (LABEL:gene_func_mass_theta) disappears, resulting in the vanishing topological susceptibility defined by a second derivative with respect to θ\theta.

When studying with a low-energy effective model, the analytical expression of (2.14) is useful for evaluating the topological susceptibility χtop\chi_{\rm top}. Here, we show another expression of χtop\chi_{\rm top} so as to see contributions from the chiral condensate ψ¯ψ\langle\bar{\psi}\psi\rangle clearly. That is, from the identity (2.12) one can rewrite Eq. (2.14) into

χtop=mlψ¯ψ4δm.\displaystyle\chi_{\rm top}=-\frac{m_{l}\langle\bar{\psi}\psi\rangle}{4}\delta_{m}\ . (2.17)

In this expression, the dimensionless quantity δm\delta_{m} is defined by

δm1χηχπ,\displaystyle\delta_{m}\equiv 1-\frac{\chi_{\eta}}{\chi_{\pi}}\ , (2.18)

which measures the variation of the susceptibility functions χπ\chi_{\pi} and χη\chi_{\eta}. Equation (2.17) indicates, indeed, that the topological susceptibility is proportional to the chiral condensate ψ¯ψ\langle\bar{\psi}\psi\rangle and δm\delta_{m}; the explicit chiral-symmetry breaking is entangled with the U(1)AU(1)_{A} anomaly contribution captured by the quantity δm\delta_{m} in χtop\chi_{\rm top}. This structure plays an important role in determining the asymptotic behavior of χtop\chi_{\rm top} at sufficiently large μq\mu_{q} where chiral symmetry is restored.

Furthermore, the Gell-Mann-Oakes-Renner (GOR) relation: fπ2mπ2=mlψ¯ψ/2f_{\pi}^{2}m_{\pi}^{2}=-m_{l}\langle\bar{\psi}\psi\rangle/2 [71], enables us to rewrite the topological susceptibility in Eq. (2.17) as

χtop=fπ2mπ22δm,\displaystyle\chi_{\rm top}=\frac{f_{\pi}^{2}m_{\pi}^{2}}{2}\delta_{m}\ , (2.19)

where fπf_{\pi} is the pion decay constant and mπm_{\pi} is the pion mass. It is obvious from its derivation that Eq. (2.19) holds model-independently.#1#1#1The GOR relation fπ2mπ2=mlψ¯ψ/2f_{\pi}^{2}m_{\pi}^{2}=-m_{l}\langle\bar{\psi}\psi\rangle/2 is derived model-independently but with an assumption that the two-point function of the pseudoscalar channel Dπδabd4x0|T(iψ¯γ5τfaψ)(x)(iψ¯γ5τfbψ)(0)|0eipxD_{\pi}\delta^{ab}\equiv\int d^{4}x\langle 0|T(i\bar{\psi}\gamma_{5}\tau^{a}_{f}\psi)(x)(i\bar{\psi}\gamma_{5}\tau^{b}_{f}\psi)(0)|0\rangle{\rm e}^{-ip\cdot x} is dominated by the lightest pseudoscalar-meson pole : Dπi/(p2mπ2)D_{\pi}\propto i/(p^{2}-m_{\pi}^{2}), as in the case of three-color QCD. Accordingly, the relation (2.19) also holds upon the pole dominance of the lightest pseudoscalar meson. Notably, the quantity δm\delta_{m} in the vacuum is solely determined by the masses of pion and η\eta meson as

δm1mπ2mη2,\displaystyle\delta_{m}\to 1-\frac{m_{\pi}^{2}}{m_{\eta}^{2}}\ , (2.20)

as long as we stick to the low-energy regime of QC2D where χπi/mπ2\chi_{\pi}\sim-i/m_{\pi}^{2} and χηi/mη2\chi_{\eta}\sim-i/m_{\eta}^{2} can apply. Therefore, Eq. (2.19) implies that the topological susceptibility in the vacuum is expressed by three basic observables in low-energy QC2D: fπf_{\pi}, mπm_{\pi} and mηm_{\eta}. We note that δm1\delta_{m}\to 1 corresponds to the significantly large anomaly effects, while δm0\delta_{m}\to 0 implies no such effects. We also note that Leutwyler and Smilga obtained the following form based on the Chiral Perturbation Theory (ChPT) in three-color QCD [72]:

χtop(LS)=mlψ¯ψ4=fπ2mπ22,\displaystyle\chi^{\rm(LS)}_{\rm top}=-\frac{m_{l}\langle\bar{\psi}\psi\rangle}{4}=\frac{f_{\pi}^{2}m_{\pi}^{2}}{2}\ , (2.21)

where the χη\chi_{\eta} contributions are missing. Indeed, in Eq. (2.21), the large anomaly is accidentally taken into account: δm=1\delta_{m}=1. Even in the case of QC2D, the Leutwyler-Smilga relation was also found in  [73].

3 Low-energy effective-model description of two-flavor QC2D{\rm QC_{2}D}

In QC2D{\rm QC_{2}D}, diquarks (antidiquarks) carrying the quark number +2+2 (2-2) are treated as color-singlet baryons, namely, baryons become bosonic similarly to mesons. Accordingly, the so-called Pauli-Gursey SU(4)SU(4) symmetry, which enables us to treat both the baryons and mesons in a consistent way, emerges [50, 51]. In this section, we briefly explain how the Pauli-Gursey SU(4)SU(4) symmetry manifests itself from QC2D Lagrangian, and based on the symmetry we present the linear sigma model which describes couplings among the baryons and mesons.

Thanks to pseudoreal properties of the SU(2)SU(2) generators for color and Dirac spaces, Tca=τ2(Tca)Tτ2T_{c}^{a}=-\tau^{2}(T_{c}^{a})^{T}\tau^{2} and σi=σ2(σi)Tσ2\sigma^{i}=-\sigma^{2}(\sigma^{i})^{T}\sigma^{2} (σi\sigma^{i} is the Pauli matrix in the Dirac space), one can show that the kinetc term of quarks coupling with gauge fields in QC2D, i.e., the first piece in Eq. (2.1), is rewritten to

Q2CD(kin)=ΨiσμDμΨ,\displaystyle{\cal L}_{\rm Q_{2}CD}^{(\rm kin)}=\Psi^{\dagger}i\sigma^{\mu}D_{\mu}\Psi\ , (3.1)

with σμ=(𝟏,σi)\sigma^{\mu}=({\bm{1}},\sigma^{i}), in the Weyl representation. In Eq. (3.1) the quark field Ψ\Psi is given by a four-component column vector in the flavor space as

Ψ=(ψRψ~L)=(uRdRu~Ld~L),\displaystyle\Psi=\begin{pmatrix}\psi_{R}\\ \tilde{\psi}_{L}\end{pmatrix}=\begin{pmatrix}u_{R}\\ d_{R}\\ \tilde{u}_{L}\\ \tilde{d}_{L}\end{pmatrix}\ , (3.2)

where ψL(R)=1γ52ψ\psi_{L(R)}=\frac{1\mp\gamma_{5}}{2}\psi denotes the left-handed (right-handed) quark field and ψ~L(R)\tilde{\psi}_{L(R)} is the conjugate one:

ψ~L(R)=σ2τc2ψL(R).\displaystyle\tilde{\psi}_{L(R)}=\sigma^{2}\tau^{2}_{c}\psi_{L(R)}^{*}\ . (3.3)

Equation (3.1) implies that the quark kinetic term in QC2D is invariant under an SU(4)SU(4) transformation for the quark field as

ΨgΨ,\displaystyle\Psi\to g\Psi\ , (3.4)

with gSU(4)g\in SU(4). Thus, it is proven that SU(2)L×SU(2)RSU(2)_{L}\times SU(2)_{R} chiral symmetry in QC2D is extended to the SU(4)SU(4) one which is often referred to as the Pauli-Gursey SU(4)SU(4) symmetry [50, 51].

Similarly to the kinetic part, the quark mass term, i.e., the second piece in Eq. (2.1), is also expressed in terms of the four-component field Ψ\Psi, which reads

QC2D(mass)=ml2(ΨTσ2τc2EΨ+Ψσ2τc2ETΨ).\displaystyle{\cal L}_{\rm QC_{2}D}^{(\rm mass)}=\frac{m_{l}}{2}\left(\Psi^{T}\sigma^{2}\tau_{c}^{2}E\Psi+\Psi^{\dagger}\sigma^{2}\tau_{c}^{2}E^{T}\Psi^{*}\right)\ . (3.5)

This term, however, breaks the Pauli-Gursey SU(4)SU(4) symmetry explicitly due to the presence of a symplectic matrix in the flavor space

E=(0𝟏𝟏0),\displaystyle E=\left(\begin{array}[]{cc}0&{\bm{1}}\\ -{\bm{1}}&0\\ \end{array}\right)\ , (3.8)

in between the two quark fields. For this reason, the systematic treatment based on the viewpoint of the SU(4)SU(4) symmetry is spoiled by the quark masses. To recover the systematics, we introduce a spurion field ζsp\zeta_{\rm sp} which transforms as

ζspgζspgT.\displaystyle\zeta_{\rm sp}\to g\,\zeta_{\rm sp}\,g^{T}\ . (3.9)

To construct the SU(4)SU(4)-invariant Lagrangian, the quark mass term is promoted to the spurion term

QC2D(sp)=ΨTσ2τc2ζspΨΨσ2τc2ζspΨ.\displaystyle{\cal L}_{\rm QC_{2}D}^{\rm(sp)}=-\Psi^{T}\sigma^{2}\tau^{2}_{c}\zeta^{\dagger}_{\rm sp}\Psi-\Psi^{\dagger}\sigma^{2}\tau^{2}_{c}\zeta_{\rm sp}\Psi^{*}\ . (3.10)

In fact, one can show that the quark mass term (3.5) is appropriately reproduced by taking the vacuum expectation value (VEV) of the spurion field as

ζsp=ml2E.\displaystyle\langle\zeta_{\rm sp}\rangle=\frac{m_{l}}{2}E\ . (3.11)

In what follows, we construct the linear sigma model to describe hadrons at the low-energy regime of QC2D, based on the symmetries explained above. The fundamental building block of the linear sigma model in QC2D is a 4×44\times 4 matrix field Σij\Sigma_{ij} whose symmetry properties are the same as those of a quark bilinear field ΨjTσ2τc2Ψi\Psi_{j}^{T}\sigma^{2}\tau_{c}^{2}\Psi_{i}. That is, Σ\Sigma transforms as

ΣgΣgT,\displaystyle\Sigma\to g\Sigma g^{T}\ , (3.12)

under the SU(4)SU(4) transformation. As explained in Ref. [65] in detail, the Σ\Sigma can be parameterized by low-lying hadrons in QC2D as

Σ=(SaiPa)XaE+(BiiBi)XiE,\displaystyle\Sigma=(S^{a}-iP^{a})X^{a}E+(B^{\prime i}-iB^{i})X^{i}E\ , (3.13)

where SaS^{a}, PaP^{a}, BiB^{i} and BiB^{\prime i} represent scalar mesons, pseudoscalar mesons, positive-parity diquark baryons and negative-parity diquark baryons, respectively. The 4×44\times 4 matrices XaX^{a} and XiX^{i} are generators of U(4)U(4) defined by

Xa\displaystyle X^{a} =\displaystyle= 122(τfa00(τfa)T)(a=0,1,2,3),\displaystyle\frac{1}{2\sqrt{2}}\begin{pmatrix}\tau^{a}_{f}&0\\ 0&(\tau^{a}_{f})^{T}\end{pmatrix}\;\;\;(a=0,1,2,3)\ ,
Xi\displaystyle X^{i} =\displaystyle= 122(0Dfi(Dfi)0)(i=4,5),\displaystyle\frac{1}{2\sqrt{2}}\begin{pmatrix}0&D_{f}^{i}\\ (D_{f}^{i})^{\dagger}&0\end{pmatrix}\;\;\;(i=4,5)\ , (3.14)

where τf0=𝟏2×2\tau_{f}^{0}={\bm{1}}_{2\times 2} in the flavor space, and DfiD_{f}^{i} represent Df4=τf2D_{f}^{4}=\tau_{f}^{2} and Df5=iτf2D_{f}^{5}=i\tau_{f}^{2}. Following the parametrization given in Ref. [65], we employ the following hadron assignment for Σ\Sigma:

Σ=12(0B+iBσiη+a0iπ02a+iπ+BiB0aiπσiηa0+iπ02σiη+a0iπ02a+iπ0B¯+iB¯a++iπ+σiηa0+iπ02B¯iB¯0),\displaystyle\Sigma=\frac{1}{2}\begin{pmatrix}0&-B^{\prime}+iB&\frac{\sigma-i\eta+a^{0}-i\pi^{0}}{\sqrt{2}}&a^{+}-i\pi^{+}\\ B^{\prime}-iB&0&a^{-}-i\pi^{-}&\frac{\sigma-i\eta-a^{0}+i\pi^{0}}{\sqrt{2}}\\ -\frac{\sigma-i\eta+a^{0}-i\pi^{0}}{\sqrt{2}}&-a^{-}+i\pi^{-}&0&-\bar{B}^{\prime}+i\bar{B}\\ -a^{+}+i\pi^{+}&-\frac{\sigma-i\eta-a^{0}+i\pi^{0}}{\sqrt{2}}&\bar{B}^{\prime}-i\bar{B}&0\end{pmatrix}\ , (3.15)

where π0=P3\pi^{0}=P^{3} and π±=(P1iP2)/2\pi^{\pm}=(P^{1}\mp iP^{2})/\sqrt{2} are the pions, η=P0\eta=P^{0} is the η\eta meson, σ=S0\sigma=S^{0} is the iso-singlet scalar meson (σ\sigma meson), a00=S3a_{0}^{0}=S^{3} and a0±=(S1iS2)/2a_{0}^{\pm}=(S^{1}\mp iS^{2})/\sqrt{2} are the iso-triplet scalar mesons (a0a_{0} mesons), B=(B5iB4)/2B=(B^{5}-iB^{4})/\sqrt{2} [B¯=(B5+iB4)/2\bar{B}=(B^{5}+iB^{4})/\sqrt{2}] is the positive-parity diquark baryon (the antidiquark baryon), and B=(B5iB4)/2B^{\prime}=(B^{\prime 5}-iB^{\prime 4})/\sqrt{2} [B¯=(B5+iB4)/2\bar{B}^{\prime}=(B^{\prime 5}+iB^{\prime 4})/\sqrt{2}] is the negative-parity diquark baryon (the antidiquark baryon).

With the matrix Σ\Sigma given in Eq. (3.15), our linear sigma model in QC2D which respects the Pauli-Gursey SU(4)SU(4) symmetry is obtained as

LSM\displaystyle{\cal L}_{\rm LSM} =\displaystyle= tr[DμΣDμΣ]V,\displaystyle{\rm tr}[D_{\mu}\Sigma^{\dagger}D^{\mu}\Sigma]-V\ , (3.16)

where the covariant derivative for Σ\Sigma is defined by

DμΣ=μΣiμqδμ0(JΣ+ΣJT)withJ=(𝟏00𝟏).\displaystyle D_{\mu}\Sigma=\partial_{\mu}\Sigma-i\mu_{q}\delta_{\mu 0}(J\Sigma+\Sigma J^{T})\;\;\;\mbox{with}\;\;\;J=\begin{pmatrix}{\bm{1}}&0\\ 0&-{\bm{1}}\end{pmatrix}. (3.17)

Here, the quark chemical potential μq\mu_{q} is incorporated in the covariant derivative through gauging the U(1)BU(1)_{B} baryon-number symmetry. Besides, VV represents potential terms describing interactions among the hadrons, which is separated into three parts as

V=V0+Vsp+Vanom.\displaystyle V=V_{0}+V_{\rm sp}+V_{\rm anom}\ . (3.18)

V0V_{0} represents an invariant part under the Pauli-Gursey SU(4)SU(4) symmetry. When we include contributions up to the fourth order of Σ\Sigma as widely done in the linear sigma model for three-color QCD, it takes the form of

V0=m02tr[ΣΣ]+λ1(tr[ΣΣ])2+λ2tr[(ΣΣ)2],\displaystyle V_{0}=m_{0}^{2}{\rm tr}[\Sigma^{\dagger}\Sigma]+\lambda_{1}({\rm tr}[\Sigma^{\dagger}\Sigma])^{2}+\lambda_{2}{\rm tr}[(\Sigma^{\dagger}\Sigma)^{2}]\ , (3.19)

where m02m_{0}^{2} is a mass parameter, and λ1\lambda_{1} and λ2\lambda_{2} are coupling constants controlling the strength of four point interactions. The second piece of Eq. (3.18), VspV_{\rm sp}, is the spurion term corresponding to QC2D(sp){\cal L}_{\rm QC_{2}D}^{\rm(sp)} in Eq. (3.10), which is given by

Vsp=c¯tr[ζspΣ+Σζsp],\displaystyle V_{\rm sp}=-\bar{c}\,{\rm tr}[\zeta^{\dagger}_{\rm sp}\Sigma+\Sigma^{\dagger}\zeta_{\rm sp}]\ , (3.20)

where the parameter c¯\bar{c} is real, and has the mass dimension two. Although the VspV_{\rm sp} is invariant under the SU(4)SU(4) transformation thanks to Eq. (3.9), the spurion field χsp\chi_{\rm sp} must be replaced by its VEV in Eq. (3.11) so as to incorporate the effect of the finite quark mass in a final step for evaluating physical observables.

The last piece in the potential (3.18), Vanom.V_{\rm anom.}, includes U(1)AU(1)_{A} anomalous contributions which is responsible for the gluonic part in the non-conservation law of the axial current: the second term of the right-hand side (RHS) of Eq. (2.11). Within our present model, the U(1)AU(1)_{A} anomalous term is expressed by the Kobayashi-Maskawa-’t Hooft (KMT)-type interaction, [66, 67, 68, 69]

Vanom=c(detΣ+detΣ).\displaystyle V_{\rm anom}=-c({\rm det}\Sigma+\det\Sigma^{\dagger})\ . (3.21)

As demonstrated below, this anomalous term generates a mass difference between the pion and η\eta meson in the vacuum [65], and plays an important role in driving a finite topological susceptibility. It should be noted that the KMT-type interaction is described by four-incoming and four-outgoing quarks owing to the quark bilinear field ΣijΨ¯jTσ2τc2Ψi\Sigma_{ij}\sim\bar{\Psi}^{T}_{j}\sigma^{2}\tau_{c}^{2}\Psi_{i} based on the Pauli-Gursey SU(4)SU(4) symmetry. Thus, in hadronic-level diagrams, VanomV_{\rm anom} represents four-point interactions.

4 Topological susceptibility at low energy

General expressions and characteristics of the topological susceptibility in QC2D have been reviewed in Sec. 2, and the linear sigma model, which describes hadrons in the low-energy regime of QC2D, has been invented in Sec. 3. In this section, we explain our strategy to evaluate the topological susceptibility within our linear sigma model through matching with the underlying QC2D theory.

4.1 Matching between low-energy effective model and underlying QC2D

In Sec. 3 we have constructed the linear sigma model in order to describe the hadrons as low-energy excitations of underlying QC2D. On the basis of the concept of the low-energy effective theory, the linear sigma model is equivalent to QC2D in the low-energy regime through the generating functional:

ZQC2D=ZLSM=[dΣ]exp(id4xLSM).\displaystyle Z_{\rm QC_{2}D}=Z_{\rm LSM}=\int[d\Sigma]\exp\left(i\int d^{4}x{\cal L}_{\rm LSM}\right)\ . (4.1)

In this subsection, we discuss the matching of the physical quantities between the linear sigma model and underlying QC2D based on Eq. (4.1). Note that we neglect spin-11 hadronic excitations such as the ρ\rho meson in the low-energy theory ZLSMZ_{\rm LSM}, even though the mass spectrums of spin-11 mesons coexist with that of spin-0 mesons in the low-energy regime [74]. This is because the topological susceptibility is evaluated by only the susceptibility functions χπ\chi_{\pi} and χη\chi_{\eta} as in Eq. (2.14), which do not include spin-11 operators. The spin-11 hadronic excitations would hardly contribute to the following results.

In a similar way to Eq. (2.3), the effective action of the linear sigma model is given by

ΓLSM=ilnZLSM.\displaystyle\Gamma_{{\rm LSM}}=-i\ln Z_{{\rm LSM}}\ . (4.2)

From the equivalence in Eq. (4.1), we have the following matching condition in terms of Γ\Gamma’s:

ΓQC2D=ΓLSM,\displaystyle\Gamma_{{\rm QC_{2}D}}=\Gamma_{{\rm LSM}}\ , (4.3)

Here, we emphasize that both the effective actions ΓQC2D\Gamma_{{\rm QC_{2}D}} and ΓLSM\Gamma_{{\rm LSM}} depend on the spurion field ζsp\zeta_{\rm sp} commonly to maintain the systematics of SU(4)SU(4) symmetry. In general, ζsp\zeta_{\rm sp} takes the form of

ζsp=(ζSaiζPa)XaE+(ζBiiζBi)XiE,\displaystyle\zeta_{\rm sp}=(\zeta_{S}^{a}-i\zeta_{P}^{a})X^{a}E+(\zeta_{{B}^{\prime}}^{i}-i\zeta_{B}^{i})X^{i}E\ , (4.4)

where ζSa\zeta_{S}^{a} (ζPa\zeta_{P}^{a}) are scalar (pseudoscalar) source fields, and ζBi\zeta_{B}^{i} (ζBi\zeta_{B^{\prime}}^{i}) are source fields associated with the positive-parity (negative-parity) diquark baryons.

Taking functional derivatives with respect to the source fields in both sides of Eq. (4.3), the matching between the linear sigma model and underlying QC2D can be done. For instance, functional derivatives with respect to the scalar source field ζS0\zeta_{S}^{0} yield#2#2#2The quark mass term in Eq. (2.1) is now replaced by the spurion term (3.10).

ψ¯ψ\displaystyle\langle\bar{\psi}\psi\rangle =\displaystyle= 2δΓQC2DδζS0(x)|ζsp=ζsp\displaystyle-\sqrt{2}\frac{\delta\Gamma_{{\rm QC_{2}D}}}{\delta\zeta_{S}^{0}(x)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle} (4.5)
=\displaystyle= 2δΓLSMδζS0(x)|ζsp=ζsp=2c¯σ.\displaystyle-\sqrt{2}\frac{\delta\Gamma_{{\rm LSM}}}{\delta\zeta_{S}^{0}(x)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle}=-\sqrt{2}\bar{c}\langle\sigma\rangle\ .

This implies that the chiral condensate ψ¯ψ\langle\bar{\psi}\psi\rangle serving as an order parameter of the spontaneous breakdown of chiral symmetry is evaluated by a VEV of σ\sigma meson within the linear sigma model. Moreover, one can see that the chiral condensate is rewritten as

ψ¯ψ=(12ΨTσ2τc2EΨ+h.c.).\displaystyle\langle\bar{\psi}\psi\rangle=\left\langle\left(-\frac{1}{2}\Psi^{T}\sigma^{2}\tau_{c}^{2}E\Psi+{\rm h.c.}\right)\right\rangle\ . (4.6)

This shows that the chiral condensate is invariant under a transformation with hh which is an element of Sp(4)Sp(4) belonging to a subgroup of SU(4)SU(4),

hTEh=E.\displaystyle h^{T}Eh=E\ . (4.7)

Hence, the symmetry-breaking pattern caused by the chiral condensate is SU(4)Sp(4)SU(4)\to Sp(4).

Likewise, when we take functional derivatives of Eq. (4.3) with respect to ζB5\zeta_{B}^{5}, the following equivalence is obtained:

(i2ψTCγ5τc2τf2ψ+h.c.)\displaystyle\left\langle\left(-\frac{i}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)\right\rangle =\displaystyle= 2δΓQC2DδζB5(x)|ζsp=ζsp\displaystyle-\sqrt{2}\frac{\delta\Gamma_{{\rm QC_{2}D}}}{\delta\zeta_{B}^{5}(x)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle} (4.8)
=\displaystyle= 2δΓLSMδζB5(x)|ζsp=ζsp=2c¯B5,\displaystyle-\sqrt{2}\frac{\delta\Gamma_{{\rm LSM}}}{\delta\zeta_{B}^{5}(x)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle}=-\sqrt{2}\bar{c}\langle B^{5}\rangle\ ,

with the charge-conjugation operator C=iγ2γ0C=i\gamma^{2}\gamma^{0}. This equation indicates that the diquark condensate ψTCγ5τc2τf2ψ\langle\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi\rangle, which plays a role of the order parameter for the emergence of the baryon superfluid phases, is mimicked by a VEV of the diquark baryon field B5B^{5} in the linear sigma model. Since B5B^{5} carries a finite quark number, the quark-number conservation no longer holds in the superfluid phase. It should be noted that the common coefficient c¯\bar{c} in Eqs. (4.5) and (4.8) is the result of the Pauli-Gursey SU(4)SU(4) symmetry which combines mesons and diquark baryons into the single multiplet.

At zero temperature, low-energy effective theories such as the linear sigma model undergo the baryon superfluid phase transition at the half value of the vacuum pion mass: μqcr=mπvac/2\mu^{\rm cr}_{q}=m^{\rm vac}_{\pi}/2 [51, 53, 65]. Below this critical chemical potential, only the hadronic phase, where no diquark condensates emerge, is realized, and all thermodynamic quantities do not change against increment of μq\mu_{q}. This stable behavior is often referred to as the Silver-Braze property, and lattice simulations also support it [49]. Above the critical chemical potential μqcr\mu_{q}^{\rm cr}, the baryon superfluid phase transition occurs and accordingly, the baryonic density also arises there. Meanwhile, in the baryon superfluid phase, the chiral condensate begins to decrease with increasing the baryonic density, resulting in the (partial) restoration of chiral symmetry [51, 53, 65].

In what follows, we use

σ0σ,ΔB5,\displaystyle\sigma_{0}\equiv\langle\sigma\rangle\ ,\ \ \Delta\equiv\langle B^{5}\rangle\ , (4.9)

to refer to the VEVs, where the phase of B5\langle B^{5}\rangle has been chosen to make Δ\Delta real.

In Eqs. (4.5) and (4.8), we have demonstrated how the QCD observables for VEVs of single local operators are matched with physical quantities of the linear sigma model: the chiral condensate and diquark condensate. The matching can be also done for two-point correlation functions by taking second functional derivatives in Eq. (4.3) with respect to the source fields. In fact, by performing functional derivatives appropriately, one can find that the η\eta-meson and pion susceptibility functions, χη\chi_{\eta} and χπ\chi_{\pi} defined in Eqs. (2.10) and (2.13), are related to two-point functions of the η\eta meson and pion in the linear sigma model, respectively, as

χη\displaystyle\chi_{\eta} =\displaystyle= 2id4xδ2ΓQC2DδζP0(x)δζP0(0)|ζsp=ζsp\displaystyle-2i\int d^{4}x\frac{\delta^{2}\Gamma_{{\rm QC_{2}D}}}{\delta\zeta_{P}^{0}(x)\delta\zeta_{P}^{0}(0)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle} (4.10)
=\displaystyle= 2id4xδ2ΓLSMδζP0(x)δζP0(0)|ζsp=ζsp=2c¯2d4x0|Tη(x)η(0)|0,\displaystyle-2i\int d^{4}x\frac{\delta^{2}\Gamma_{{\rm LSM}}}{\delta\zeta_{P}^{0}(x)\delta\zeta_{P}^{0}(0)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle}=2\bar{c}^{2}\int d^{4}x\langle 0|T\eta(x)\eta(0)|0\rangle\ ,

and

χπδab\displaystyle\chi_{\pi}\delta^{ab} =\displaystyle= 2id4xδ2ΓQC2DδζPa(x)δζPb(0)|ζsp=ζsp\displaystyle-2i\int d^{4}x\frac{\delta^{2}\Gamma_{{\rm QC_{2}D}}}{\delta\zeta_{P}^{a}(x)\delta\zeta_{P}^{b}(0)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle}
=\displaystyle= 2id4xδ2ΓLSMδζPa(x)δζPb(0)|ζsp=ζsp=2c¯2d4x0|Tπa(x)πb(0)|0(for a,b=1,2,3).\displaystyle-2i\int d^{4}x\frac{\delta^{2}\Gamma_{{\rm LSM}}}{\delta\zeta_{P}^{a}(x)\delta\zeta_{P}^{b}(0)}\Biggl{|}_{\zeta_{\rm sp}=\langle\zeta_{\rm sp}\rangle}=2\bar{c}^{2}\int d^{4}x\langle 0|T\pi^{a}(x)\pi^{b}(0)|0\rangle\;\;\;(\mbox{for }a,b=1,2,3)\ .

Using these matching equations, we will present analytic expressions of the meson susceptibility functions within our linear sigma model.

4.2 Topological susceptibility across baryon superfluid phase transition

In this subsection, we proceed with analytic evaluation of the topological susceptibility from the linear sigma model.

In this work, we employ the mean-field approximation where loop corrections of hadronic fluctuations have not been taken into account. The effective potential of the linear sigma model at the tree level is evaluated as

Vmean=2μq2Δ2+m022(σ02+Δ2)+8λ1+2λ2c32(σ02+Δ2)22mlc¯σ0.\displaystyle V_{\rm mean}=-2\mu_{q}^{2}\Delta^{2}+\frac{m_{0}^{2}}{2}(\sigma_{0}^{2}+\Delta^{2})+\frac{8\lambda_{1}+2\lambda_{2}-c}{32}(\sigma_{0}^{2}+\Delta^{2})^{2}-\sqrt{2}m_{l}\bar{c}\sigma_{0}\ . (4.12)

In this potential the VEV of the spurion field (3.11) as well as the mean fields (4.9) are inserted. In Eq. (4.12) the quark chemical potential μq\mu_{q} appears in a quadratic term of Δ\Delta with a negative sign, indicating that the larger value of μq\mu_{q} yields nonzero Δ\Delta leading to the baryon superfluid phase as mentioned in Sec. 4.1. The vacuum configurations are determined by stationary conditinos of VmeanV_{\rm mean} with respect to σ0\sigma_{0} and Δ\Delta:

Vmeanσ0=0,VmeanΔ=0,\displaystyle\frac{\partial V_{\rm mean}}{\partial\sigma_{0}}=0\ ,\ \ \frac{\partial V_{\rm mean}}{\partial\Delta}=0\ , (4.13)

and hadrons appear as fluctuation modes upon the vacuum characterized by the conditions (4.13). In this description, hadron masses are evaluated by quadratic terms of the fluctuations in the Lagrangian (3.16) with σ0\sigma_{0} and Δ\Delta included. For instance, the pion mass reads

mπ2=m02+8λ1+2λ2c8(σ02+Δ2)=2mlc¯σ0.\displaystyle m_{\pi}^{2}=m_{0}^{2}+\frac{8\lambda_{1}+2\lambda_{2}-c}{8}(\sigma_{0}^{2}+\Delta^{2})=\frac{\sqrt{2}m_{l}\bar{c}}{\sigma_{0}}\ . (4.14)

We note that the second equality in Eq. (4.14) is obtained by considering the stationary condition of σ0\sigma_{0} in Eq. (4.13).

When we approximate the pion two-point function 0|Tπa(x)πb(0)|0\langle 0|T\pi^{a}(x)\pi^{b}(0)|0\rangle at the tree level in the linear sigma model, the pion susceptibility function in Eq. (LABEL:pi_sus_LSM) is evaluated to be

χπ=2ic¯21mπ2.\displaystyle\chi_{\pi}=-2i\bar{c}^{2}\frac{1}{m_{\pi}^{2}}\ . (4.15)

Similarly, we employ the tree-level approximation for the η\eta-meson two-point function 0|Tη(x)η(0)|0\langle 0|T\eta(x)\eta(0)|0\rangle. However, since the violation of U(1)BU(1)_{B} baryon-number symmetry in the baryon superfluid phase causes the mixing among η\eta meson, the negative-parity diquark BB^{\prime} and antidiquark B¯\bar{B}^{\prime} (or equivalently η\eta, B4B^{\prime}_{4} and B5B^{\prime}_{5}), the two-point function of the η\eta meson is not simply given by i/mη2-i/m_{\eta}^{2} where the η\eta mass is read from η2\eta^{2} term of the η\eta fluctuation from the vacuum. By taking into account the mixing structure, the inverse propagator matrix for the η\eta - B4B^{\prime}_{4} - B5B^{\prime}_{5} sector in the momentum space at the rest frame 𝒑=𝟎{\bm{p}}={\bm{0}} is obtained as

i𝑫1=i(DηηDηB4DηB5DB4ηDB4B4DB4B5DB5ηDB5B4DB5B5)1=(p02mη20mB5η20p02mB424iμqp0mB5η24iμqp0p02mB52),\displaystyle i{\bm{D}}^{-1}=i\left(\begin{array}[]{ccc}D_{\eta\eta}&D_{\eta B_{4}^{\prime}}&D_{\eta B_{5}^{\prime}}\\ D_{B_{4}^{\prime}\eta}&D_{B_{4}^{\prime}B_{4}^{\prime}}&D_{B_{4}^{\prime}B_{5}^{\prime}}\\ D_{B_{5}^{\prime}\eta}&D_{B_{5}^{\prime}B_{4}^{\prime}}&D_{B_{5}^{\prime}B_{5}^{\prime}}\\ \end{array}\right)^{-1}=\left(\begin{array}[]{ccc}p_{0}^{2}-m_{\eta}^{2}&0&-m_{B_{5}^{\prime}\eta}^{2}\\ 0&p_{0}^{2}-m_{B_{4}^{\prime}}^{2}&4i\mu_{q}p_{0}\\ -m_{B_{5}^{\prime}\eta}^{2}&-4i\mu_{q}p_{0}&p_{0}^{2}-m_{B_{5}^{\prime}}^{2}\\ \end{array}\right)\ , (4.22)

where we have defined the two-point functions DXYD_{XY} by DXY=F.T.0|TX(x)Y(0)|0D_{XY}={\rm F.T.}\langle 0|TX(x)Y(0)|0\rangle with X,Y=η,B4X,Y=\eta,B_{4}^{\prime} and B5B_{5}^{\prime}, and the mass parameters read

mη2\displaystyle m_{\eta}^{2} =\displaystyle= mπ2+λ22Δ2+c4(2σ02+Δ2),\displaystyle m_{\pi}^{2}+\frac{\lambda_{2}}{2}\Delta^{2}+\frac{c}{4}(2\sigma_{0}^{2}+\Delta^{2})\ ,
mB52\displaystyle m_{B_{5}^{\prime}}^{2} =\displaystyle= mπ24μq2+λ22σ02+c4(σ02+2Δ2),\displaystyle m_{\pi}^{2}-4\mu_{q}^{2}+\frac{\lambda_{2}}{2}\sigma_{0}^{2}+\frac{c}{4}(\sigma_{0}^{2}+2\Delta^{2})\ ,
mB5η2\displaystyle m^{2}_{B_{5}^{\prime}\eta} =\displaystyle= 2λ2c4σ0Δ.\displaystyle\frac{2\lambda_{2}-c}{4}\sigma_{0}\Delta\ . (4.23)

Thus, inverting the matrix (4.22), one can find that DηηD_{\eta\eta} which is of interest now takes the form of

Dηη(p0)=i=1,2,3iZϕip02mϕi2.\displaystyle D_{\eta\eta}(p_{0})=\sum_{i=1,2,3}\frac{iZ_{\phi_{i}}}{p_{0}^{2}-m_{\phi_{i}}^{2}}\ . (4.24)

In this expression, mϕ1m_{\phi_{1}}, mϕ2m_{\phi_{2}} and mϕ3m_{\phi_{3}} represent mass eigenvalues of the η\eta - B4B^{\prime}_{4} - B5B^{\prime}_{5} sector, where the subscripts ϕ1\phi_{1}, ϕ2\phi_{2} and ϕ3\phi_{3} stand for the corresponding eigenstates with which the masses satisfy mϕ1>mϕ2>mϕ3m_{\phi_{1}}>m_{\phi_{2}}>m_{\phi_{3}}. The renormalization constants ZϕiZ_{\phi_{i}} in Eq. (4.24) are evaluated by

Zϕ1\displaystyle Z_{\phi_{1}} =\displaystyle= 𝒩ηη(mϕ1)(mϕ12mϕ22)(mϕ12mϕ32),\displaystyle\frac{{\cal N}_{\eta\eta}(m_{\phi_{1}})}{(m_{\phi_{1}}^{2}-m_{\phi_{2}}^{2})(m_{\phi_{1}}^{2}-m_{\phi_{3}}^{2})}\ ,
Zϕ2\displaystyle Z_{\phi_{2}} =\displaystyle= 𝒩ηη(mϕ2)(mϕ22mϕ12)(mϕ22mϕ32),\displaystyle\frac{{\cal N}_{\eta\eta}(m_{\phi_{2}})}{(m_{\phi_{2}}^{2}-m_{\phi_{1}}^{2})(m_{\phi_{2}}^{2}-m_{\phi_{3}}^{2})}\ ,
Zϕ3\displaystyle Z_{\phi_{3}} =\displaystyle= 𝒩ηη(mϕ3)(mϕ32mϕ12)(mϕ32mϕ22),\displaystyle\frac{{\cal N}_{\eta\eta}(m_{\phi_{3}})}{(m_{\phi_{3}}^{2}-m_{\phi_{1}}^{2})(m_{\phi_{3}}^{2}-m_{\phi_{2}}^{2})}\ , (4.25)

with

𝒩ηη(p0)=p0416μq2p02(mB42+mB52)p02+mB42mB52.\displaystyle{\cal N}_{\eta\eta}(p_{0})=p_{0}^{4}-16\mu_{q}^{2}p_{0}^{2}-(m_{B_{4}^{\prime}}^{2}+m_{B_{5}^{\prime}}^{2})p_{0}^{2}+m_{B_{4}^{\prime}}^{2}m_{B_{5}^{\prime}}^{2}\ . (4.26)

We note that the constants satisfy a condition Zϕ1+Zϕ2+Zϕ3=1Z_{\phi_{1}}+Z_{\phi_{2}}+Z_{\phi_{3}}=1 reflecting the fraction conservation. Therefore, ZϕiZ_{\phi_{i}} correspond to the proportion of the mass eigenstates ϕi\phi_{i} in the two-point function DηηD_{\eta\eta} while the information on the respective pole positions is read from 1/(p02mϕi2)1/(p_{0}^{2}-m_{\phi_{i}}^{2}) in Eq. (4.24). #3#3#3In Eq. (4.24) we have expressed Dηη(p0)D_{\eta\eta}(p_{0}) in terms of three contributions of ϕ1\phi_{1}, ϕ2\phi_{2} and ϕ3\phi_{3} so as to see roles of the mass eigenstates clearly. In the low-energy limit p0=0p_{0}=0, Dηη(0)D_{\eta\eta}(0) is of course equivalent to the simple form of Dηη(0)=imB52mη2mB52mB5η4,\displaystyle D_{\eta\eta}(0)=-i\frac{m_{B_{5}^{\prime}}^{2}}{m_{\eta}^{2}m_{B_{5}^{\prime}}^{2}-m_{B_{5}^{\prime}\eta}^{4}}\ , (4.27) which can be straightforwardly derived by evaluating the inverse matrix of η\eta - B5B^{\prime}_{5} sector of Eq (4.22).

By using the η\eta-meson propagator based on the mass eigenstates ϕi\phi_{i} in Eq. (4.24), the η\eta-meson susceptibility function is evaluated as

χη=2c¯2Dηη(0)=i=1,2,3χϕi,\displaystyle\chi_{\eta}=2\bar{c}^{2}D_{\eta\eta}(0)=\sum_{i=1,2,3}\chi_{\phi_{i}}\ , (4.28)

with

χϕi2ic¯2Zϕimϕi2.\displaystyle\chi_{\phi_{i}}\equiv-2i\bar{c}^{2}\frac{Z_{\phi_{i}}}{m_{\phi_{i}}^{2}}\ . (4.29)

Since the susceptibility is defined at the low-energy limit: p0=0p_{0}=0, the susceptibilities χϕi\chi_{\phi_{i}} in Eq. (4.29) are written by Zϕi2/mϕi2Z^{2}_{\phi_{i}}/m^{2}_{\phi_{i}} with the constant 2ic¯2-2i\bar{c}^{2}. Therefore, the strength of χη\chi_{\eta} is controlled by the combination of the renormalization constants ZϕiZ_{\phi_{i}} and the mass eigenvalues mϕim_{\phi_{i}}.

5 Fate of topological susceptibility in dense QC2D

With the help of the susceptibility functions χπ\chi_{\pi} and χη\chi_{\eta} obtained in Eqs. (4.15) and (4.28), the topological susceptibility χtop\chi_{\rm top} is evaluated within our linear sigma model from Eq. (2.14):

χtop=iml24(χπi=1,2,3χϕi).\displaystyle\chi_{\rm top}=\frac{im_{l}^{2}}{4}\left(\chi_{\pi}-\sum_{i=1,2,3}\chi_{\phi_{i}}\right)\ . (5.1)

In this section, based on it, we show the numerical results of χtop\chi_{\rm top} at finite μq\mu_{q}.

5.1 U(1)AU(1)_{A} anomaly contribution and μq\mu_{q} dependence of topological susceptibility

As explained at the end of Sec. 2, the topological susceptibility is substantially controlled by the U(1)AU(1)_{A} axial anomaly, i.e., the mass difference between the η\eta meson and pion in low-energy QC2D. For this reason, we particularly investigate χtop\chi_{\rm top} at finite μq\mu_{q} with the two cases of mηvac/mπvac=1m^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}=1 and mηvac/mπvac=1.5m^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}=1.5. The former corresponds to vanishing anomaly effects for the hadron spectrum, while the latter implies the substantial anomaly effects.

Mass ratio cc λ1\lambda_{1} λ2\lambda_{2} m02m_{0}^{2} mlc¯/2m_{l}\bar{c}/2
mηvac/mπvac=1m^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}=1 0 0 65.665.6 (693MeV)2-(693\,{\rm MeV})^{2} (364MeV)3(364\,{\rm MeV})^{3}
mηvac/mπvac=1.5m^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}=1.5 21.821.8 0 54.754.7 (373MeV)2-(373\,{\rm MeV})^{2} (364MeV)3(364\,{\rm MeV})^{3}
Table 1: Fixed parameters for mηvac/mπvac=1m^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}=1 and mηvac/mπvac=1.5m^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}=1.5.

When we fix the mass ratio mηvac/mπvacm^{\rm vac}_{\eta}/m^{\rm vac}_{\pi}, there remain four parameters to be determined. As inputs, we employ mπvac=738m_{\pi}^{\rm vac}=738 MeV and mB(B¯)vac=1611m_{B^{\prime}(\bar{B}^{\prime})}^{\rm vac}=1611 MeV from the recent lattice data [74]. Besides, based on the previous work [65], σ0vac=250\sigma^{\rm vac}_{0}=250 MeV is used as another input as a typical value. For the last constraint, we take λ1=0\lambda_{1}=0 corresponding to the large NcN_{c} limit since λ1\lambda_{1} term includes a double trace in the flavor space. With those inputs, the model parameters are fixed as in Table. 1. The table indicates that mηvacm^{\rm vac}_{\eta} becomes larger than mπvacm^{\rm vac}_{\pi} only when c0c\neq 0. In other words, the KMT-type interaction mimicking the U(1)AU(1)_{A} anomaly effects in the linear sigma model generates the mass difference between the η\eta meson and pion, as expected from underlying QC2D.

In order to demonstrate typical phase structures at zero temperature and finite chemical potential described by the present linear sigma model, we depict μq\mu_{q} dependences of the chiral condensate σ0\sigma_{0} and diquark condensate Δ\Delta in the panel (a) of Fig. 1 for the two parameter sets of Table 1. This figure clearly shows that the baryon superfluid phase emerges from μqcr=mπvac/2\mu_{q}^{\rm cr}=m_{\pi}^{\rm vac}/2, and accordingly chiral symmetry begins to be restored. The mean field σ0\sigma_{0} decreases in the superfluid phase independently of the strength of the U(1)AU(1)_{A} anomaly effects, whereas the anomaly accelerates the increment of Δ\Delta there. We note that the smooth reduction of σ0\sigma_{0} in the superfluid phase is analytically evaluated as

σ0=mlc¯22μq2,\displaystyle\sigma_{0}=\frac{m_{l}\bar{c}}{2\sqrt{2}}\mu_{q}^{-2}\ , (5.2)

from the stationary conditions in Eq. (4.13). We also note that in the case of the nonlinear representation of the Nambu-Goldstone (NG) bosons, the vacuum manifold of the SU(4)SU(4) symmetry breaking is constrained as “σ02+Δ2=(constant)\sigma_{0}^{2}+\Delta^{2}=({\rm constant})” at the tree level which was found in Refs. [50, 51]. However, this is not the case in the linear representation [65].

Incidentally, Fig. 1 also depicts the μq\mu_{q} dependence of the baryon number density (ρ=4Δ2μq\rho=4\Delta^{2}\mu_{q}) normalized by 16fπ2mπvac16f_{\pi}^{2}m_{\pi}^{\rm vac} in panel (b). The baryon number density is generated after reaching the baryon superfluid phase. Owing to the increment of Δ\Delta, the baryon number density is enhanced by the U(1)AU(1)_{A} anomalous contribution.

Refer to caption (a) Refer to caption (b)
Figure 1: Chemical potential μq\mu_{q} dependences of chiral condensate σ0\sigma_{0} and diquark condensate Δ\Delta (a) and that of the baryon density ρ\rho normalized by 16fπ2mπvac16f_{\pi}^{2}m_{\pi}^{\rm vac} (b).

For later convenience, here we comment on the masses mϕim_{\phi_{i}} and renormalization constants ZϕiZ_{\phi_{i}} across the phase transition. Displayed in Fig. 2 is μq\mu_{q} dependences of the masses of η\eta - BB^{\prime} - B¯\bar{B}^{\prime} sector for mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (a) and mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 (b). In the baryon superfluid phase, mϕ1m_{\phi_{1}}, mϕ2m_{\phi_{2}} and mϕ3m_{\phi_{3}} correspond to the green curves from above: mϕim_{\phi_{i}} are ordered from the largest mass, mϕ1>mϕ2>mϕ3m_{\phi_{1}}>m_{\phi_{2}}>m_{\phi_{3}}. Panel (b) in Fig. 2 indicates that the mass ordering of BB^{\prime} and η\eta is interchanged below the critical chemical potential μqcr\mu_{q}^{\rm cr} for mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 whereas such a level crossing does not take place for mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1. For this reason, the masses of ϕ1\phi_{1}, ϕ2\phi_{2} and ϕ3\phi_{3} are connected to

(mϕ1,mϕ2,mϕ3)(mB¯,mB,mη)\displaystyle(m_{\phi_{1}},m_{\phi_{2}},m_{\phi_{3}})\to(m_{\bar{B}^{\prime}},m_{B^{\prime}},m_{\eta})\ \ formηvac/mπvac=1,\displaystyle{\rm for}\ \ m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1\ ,
(mϕ1,mϕ2,mϕ3)(mB¯,mη,mB)\displaystyle(m_{\phi_{1}},m_{\phi_{2}},m_{\phi_{3}})\to(m_{\bar{B}^{\prime}},m_{\eta},m_{B^{\prime}})\ \ formηvac/mπvac=1.5,\displaystyle{\rm for}\ \ m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5\ , (5.3)

at μqcr\mu_{q}^{\rm cr}. These correspondences are also reflected in the renormalization constants ZϕiZ_{\phi_{i}}, as depicted in Fig. 3. Indeed, the figure indicates that in the hadronic phase the ZϕiZ_{\phi_{i}} are reduced to

(Zϕ1,Zϕ2,Zϕ3)=(0,0,1)\displaystyle(Z_{\phi_{1}},Z_{\phi_{2}},Z_{\phi_{3}})=(0,0,1)\ \ formηvac/mπvac=1,\displaystyle{\rm for}\ \ m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1\ ,
(Zϕ1,Zϕ2,Zϕ3)=(0,1,0)\displaystyle(Z_{\phi_{1}},Z_{\phi_{2}},Z_{\phi_{3}})=(0,1,0)\ \ formηvac/mπvac=1.5.\displaystyle{\rm for}\ \ m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5\ . (5.4)

Thus, ϕ3\phi_{3} (ϕ2\phi_{2}) state is connected to the η\eta-meson state in this phase when mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5). Meanwhile, in a limit of μq\mu_{q}\to\infty, both the parameter sets read Zϕ21Z_{\phi_{2}}\to 1 while Zϕ1,Zϕ30Z_{\phi_{1}},Z_{\phi_{3}}\to 0, reflecting a fact that the state of ηψ¯iγ5ψ\eta\sim\bar{\psi}i\gamma_{5}\psi is dominated by ϕ2\phi_{2} solely at sufficiently large μq\mu_{q} where σ0\sigma_{0} is negligible [65]. It should be noted that the ϕ1\phi_{1} component in the η\eta state is suppressed at any chemical potential.

Refer to caption (a) Refer to caption (b)
Figure 2: μq\mu_{q} dependence of the masses of η\eta - BB^{\prime} - B¯\bar{B}^{\prime} sector for mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (a) and mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 (b). In the baryon superfluid phase, mϕ1m_{\phi_{1}}, mϕ2m_{\phi_{2}} and mϕ3m_{\phi_{3}} correspond to the green curves from above. In this figure the hadron masses are scaled by mπvacm_{\pi}^{\rm vac}.
Refer to caption (a) Refer to caption (b)
Figure 3: μq\mu_{q} dependence of the renormalization constants ZϕiZ_{\phi_{i}} for mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (a) and mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 (b).

Keeping the above properties in mind, we depict μq\mu_{q} dependences of the topological susceptibility χtop\chi_{\rm top} in Fig. 4. From panel (a) one can see the topological susceptibility is always zero in the absence of the U(1)AU(1)_{A} anomaly effects. In the hadronic phase, such a trend is easily understood by a fact that mπvacm_{\pi}^{\rm vac} coincides with mηvacm_{\eta}^{\rm vac} together with Eq. (2.14). The null topological susceptibility in the superfluid phase is rather surprising, but it is also understood as follows. Within our linear sigma model, the KMT-type interaction is introduced to mimic the gluonic anomalous part in the non-conservation law of the U(1)AU(1)_{A} axial current in Eq. (2.11). This structure is irrespective of changes of dynamical symmetry-breaking properties such as the emergence of the baryon superfluidity. Hence, even in the superfluid phase where the η\eta meson mixes with B4B^{\prime}_{4} and B5B^{\prime}_{5}, the θ\theta dependence in the quark mass term in Eq. (LABEL:gene_func_mass_theta) would be rotated away under U(1)AU(1)_{A} transformation when the KMT-type interaction, i.e., the U(1)AU(1)_{A} anomaly effect, is turned off. For this reason, the topological susceptibility defined by a second derivative with respect to θ\theta always vanishes as long as c=0c=0 is taken.

Refer to caption (a) Refer to caption (b)
Figure 4: μq\mu_{q} dependence of the topological susceptibility and the meson susceptibility functions for mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (a) and mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 (b). In this figure, the susceptibilities are scaled by (mπvac)4(m_{\pi}^{\rm vac})^{4}.

Here, we comment on behaviors of the each contribution from χπ\chi_{\pi} and χϕi\chi_{\phi_{i}} in panel (a) of Fig. 4. First, since the pion mass in the baryon superfluid phase is expressed as

mπ2=4μq2,\displaystyle m_{\pi}^{2}=4\mu_{q}^{2}\ , (5.5)

the pion susceptibility function χπ\chi_{\pi} decreases in this phase with a power of μq2\mu_{q}^{-2}. In contrast to the baryon superfluid phase, χπ\chi_{\pi} does not change in the hadronic phase. Next, the figure indicates that, in the hadronic phase, the η\eta-meson susceptibility function is completely dominated by χϕ3\chi_{\phi_{3}}, while χϕ1\chi_{\phi_{1}} and χϕ2\chi_{\phi_{2}} vanish there. Hence, χϕ3\chi_{\phi_{3}} coincides with χπ\chi_{\pi} to yield χtop=0\chi_{\rm top}=0. This behavior is understood by panel (a) of Fig. 3; the η\eta state in the hadronic phase is connected to the ϕ3\phi_{3} one solely. Moving on to the baryon superfluid phase, we find that χϕ2\chi_{\phi_{2}} grows from zero and χϕ3\chi_{\phi_{3}} becomes smaller than χπ\chi_{\pi} to compensate the growth. Although Fig. 3 exhibits the significant interchange of Zϕ2Z_{\phi_{2}} with Zϕ3Z_{\phi_{3}} above μq0.53mπvac\mu_{q}\sim 0.53m_{\pi}^{\rm vac}, χϕ2\chi_{\phi_{2}} is smaller than χϕ3\chi_{\phi_{3}} at any chemical potential due to the comparably strong suppression stemming from the mϕ22m_{\phi_{2}}^{-2} dependence in Eq. (4.29). Meanwhile, χϕ1\chi_{\phi_{1}} is always negligible because of the large mass suppression of 1/mϕ121/m_{\phi_{1}}^{2} and the small value of Zϕ1Z_{\phi_{1}} (see in Figs. 2 and Fig. 3).

The U(1)AU(1)_{A} anomaly effect represented by a nonzero cc in our linear sigma model can be seen from panel (b) of Fig. 4, where mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 is taken. In the hadronic phase mπvac<mηvacm_{\pi}^{\rm vac}<m_{\eta}^{\rm vac} holds, and thus, χη\chi_{\eta} becomes smaller than χπ\chi_{\pi}, resulting the nonzero topological susceptibility. In the baryon superfluid phase, χtop\chi_{\rm top} decreases monotonically and approaches zero. The detailed analysis on this asymptotic behavior is provided in Sec. 5.2. Before taking a closer look at the smooth suppression of χtop\chi_{\rm top} at larger μq\mu_{q}, we explain behaviors of the respective meson susceptibility functions in the presence of the KMT-type interaction. First, the μq\mu_{q} dependence of χπ\chi_{\pi} remains the same as one without the U(1)AU(1)_{A} anomaly effects: the μq\mu_{q} scaling of mπm_{\pi} in the superfluid phase, mπ2=4μq2m_{\pi}^{2}=4\mu_{q}^{2}, holds even when the anomaly is included. Second, in the hadronic phase only χϕ2\chi_{\phi_{2}} contributes to the topological susceptibility while χϕ1\chi_{\phi_{1}} and χϕ3\chi_{\phi_{3}} do not, as seen from panel (b) of Fig. 3. Third, in the superfluid phase, the finite χϕ3\chi_{\phi_{3}} is induced above μqcr\mu_{q}^{\rm cr} owing to the bump structure of Zϕ3Z_{\phi_{3}} shown in panel (b) of Fig. 3. But soon it begins to decrease and becomes negligible around μq0.8mπvac\mu_{q}\sim 0.8m_{\pi}^{\rm vac} accompanied by the suppression of Zϕ3Z_{\phi_{3}}. Meanwhile, the abrupt suppression of χϕ2\chi_{\phi_{2}} occurs above μqcr\mu_{q}^{\rm cr} to compensate the enhancement of χϕ3\chi_{\phi_{3}}, and at larger μq\mu_{q}, χϕ2\chi_{\phi_{2}} gradually approaches zero in accordance with the increment of mϕ2m_{\phi_{2}}. We note that χϕ1\chi_{\phi_{1}} is almost zero at any chemical potential from the same reason explained for mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.

5.2 Asymptotic behavior of topological susceptibility in dense baryonic matter

Here, we focus on the cases for mηvac/mπvac>1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}>1, in which the finite χtop\chi_{\rm top} is provided, in order to delineate the asymptotic behavior of χtop\chi_{\rm top} at larger μq\mu_{q}.

The smooth reduction of χtop\chi_{\rm top} at larger μq\mu_{q} in panel (b) of Fig. 4 can be explained by the continuous reduction of ψ¯ψ\langle\bar{\psi}\psi\rangle, as inferred from Eq. (2.17).#4#4#4A similar smooth decrease of χtop\chi_{\rm top} associated with the continuous chiral phase transition was observed in hot three-color QCD matter based on chiral model analyses [7, 21, 8], and lattice simulations at physical quark masses support such a behavior [10, 11, 12]. In order to see this behavior, we rewrite the topological susceptibility to χtop=(mlc¯)2δm/(8μq2)\chi_{\rm top}=(m_{l}\bar{c})^{2}\delta_{m}/(8\mu_{q}^{2}) with the help of Eqs. (4.5) and (5.2). The dimensionless quantity δm\delta_{m} is easily evaluated at sufficiently large μq\mu_{q} with an assumption that χη\chi_{\eta} is solely controlled by ϕ2\phi_{2} state. Indeed, the asymptotic behavior of mϕ2m_{\phi_{2}} is known to be mϕ2212μq2m_{\phi_{2}}^{2}\sim 12\mu_{q}^{2} [65], so that the quantity δm\delta_{m} is approximated to be δm2/3\delta_{m}\sim 2/3 with Eq. (5.5). Therefore, the asymptotic behavior of χtop\chi_{\rm top} would be analytically fitted by

χtop(mπvac)4(fπvac)212μq2,\displaystyle\frac{\chi_{\rm top}}{(m_{\pi}^{\rm vac})^{4}}\sim\frac{(f_{\pi}^{\rm vac})^{2}}{12}\mu_{q}^{-2}\ , (5.6)

where (mπvac)2=2mlc¯/σ0vac=mlc¯/fπvac(m_{\pi}^{\rm vac})^{2}=\sqrt{2}m_{l}\bar{c}/\sigma_{0}^{\rm vac}=m_{l}\bar{c}/f_{\pi}^{\rm vac} is used. The μq\mu_{q} scaling of χtop\chi_{\rm top} coincides with that of the chiral condensate in the superfluid phase [see Eq. (5.2)]. In Fig. 5, we plot the topological susceptibility in the baryon superfluid phase with several values of mηvac/mπvacm_{\eta}^{\rm vac}/m_{\pi}^{\rm vac} with keeping input values: mπvac=738MeVm_{\pi}^{\rm vac}=738{\rm MeV}, mB(B¯)vac=1611MeVm_{B^{\prime}(\bar{B}^{\prime})}^{\rm vac}=1611{\rm MeV} and σ0vac=250MeV\sigma_{0}^{\rm vac}=250{\rm MeV}. The figure shows that the asymptotic behavior of χtop\chi_{\rm top} is fitted by the analytic expression in Eq. (5.6) well, regardless of the value of mηvac/mπvacm_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}.

Refer to caption
Figure 5: Topological susceptibility in the baryon superfluid phase with several values of mηvac/mπvacm_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}.

The lattice simulation performed in Ref. [41] at T=0.45TcT=0.45T_{c} indicates that the topological susceptibility of QC2D in the hadronic phase has a finite value with the error bars and is not influenced by the μq\mu_{q} effect. #5#5#5Here, TcT_{c} denotes the pseudocritical temperature for the chiral phase transition at vanishing μq\mu_{q}, which are fixed to be Tc=200T_{c}=200 MeV [44]. Moreover, the lattice result shows that such an approximately μq\mu_{q}-independent behavior is further extended to the baryon superfluid phase, which obviously contradicts our model estimations. One possible scenario explaining this discrepancy is discussed in Sec. 6. In contrast to Ref. [41], the other lattice result reported in Ref. [45] would suggest that the topological susceptibility in the baryon superfluid phase is suppressed, as estimated by our present study.

6 Contamination by diquark source field

In the evaluations in Sec. 5, we have included only the VEV of a scalar source field from the spurion field ζsp\zeta_{\rm sp} which turns into the current quark mass as shown in Eq. (3.11). On the other hand, in lattice simulations diquark source effects incorporated from a VEV of ζB5\zeta_{B}^{5} in Eq. (4.4) would remain additionally, particularly in the baryon superfluid phase. Then, in this section we discuss the diquark source effects to the topological susceptibility at finite μq\mu_{q} within our linear sigma model.

6.1 Diquark source effect on topological susceptibility

To analytically find out contributions of the diquark source field jj to the topological susceptibility, we first incorporate jj in underlying QC2D by adding the VEV of ζB5=2j\langle\zeta_{B}^{5}\rangle=\sqrt{2}j from the spurion field (4.4). Now the VEV of ζsp\zeta_{\rm sp} reads

ζsp=ml2Ei2jX5E.\displaystyle\langle\zeta_{\rm sp}\rangle=\frac{m_{l}}{2}E-i\sqrt{2}jX^{5}E\ . (6.1)

With this VEV, a diquark operator tagged with the diquark source jj shows up as a new ingredient in the quark mass term:

QC2D(mass)\displaystyle{\cal L}_{\rm QC_{2}D}^{\rm(mass)} =\displaystyle= mlψ¯ψj(i2ψTCγ5τc2τf2ψ+h.c.).\displaystyle-m_{l}\bar{\psi}\psi-j\left(-\frac{i}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)\ . (6.2)

This mass term implies that the extra term characterized by jj explicitly breaks the U(1)AU(1)_{A} symmetry as well as the U(1)BU(1)_{B} baryon-number symmetry. In fact, under the U(1)AU(1)_{A} axial transformation with an angle satisfying αA=2θ\alpha_{A}=2\theta, the generating functional of QC2D with the modified mass term (6.2) is rotated to

ZQC2D\displaystyle Z_{\rm QC_{2}D} =\displaystyle= [dψ¯dψ][dA]exp[id4x(ψ¯iγμDμψmlψ¯exp(iθ/2γ5)ψ\displaystyle\int[d\bar{\psi}d\psi][dA]\exp\Biggl{[}i\int d^{4}x\Biggl{(}\bar{\psi}i\gamma^{\mu}D_{\mu}\psi-m_{l}\bar{\psi}\exp\left(i\theta/2\,\gamma_{5}\right)\psi (6.3)
j(i2ψTCγ5τc2τf2eiθ/2γ5ψ+h.c.)14GμνaGμν,a)],\displaystyle-j\left(-\frac{i}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}{\rm e}^{i\theta/2\gamma_{5}}\psi+{\rm h.c.}\right)-\frac{1}{4}G_{\mu\nu}^{a}G^{\mu\nu,a}\Biggl{)}\Biggl{]}\ ,

and hence, from Eq. (2.4) via Eq. (2.3) the net topological susceptibility χtopw/j\chi_{\rm top}^{\rm w/j} is evaluated to be

χtopw/j=χtop(M)+δχtop.\displaystyle\chi_{\rm top}^{\rm w/j}=\chi_{\rm top}^{\rm(M)}+\delta\chi_{\rm top}\ . (6.4)

In this expression, χtop(M)\chi_{\rm top}^{\rm(M)} is identical to χtop\chi_{\rm top} given by Eq. (2.14), but here the superscript “(M)” has been attached in order to emphasize that only contributions from the meson susceptibility functions are included:

χtop(M)=i4ml2(χπχη).\displaystyle\chi_{\rm top}^{\rm(M)}=\frac{i}{4}m_{l}^{2}(\chi_{\pi}-\chi_{\eta})\ . (6.5)

δχtop\delta\chi_{\rm top} denotes additional contributions from jj which is of the form

δχtop\displaystyle\delta\chi_{\rm top} =\displaystyle= 14[j(i2ψTCγ5τc2τf2ψ+h.c.)+ij2χB52ijmlχB5η],\displaystyle-\frac{1}{4}\left[j\left(-\frac{i}{2}\langle\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi\rangle+{\rm h.c.}\right)+ij^{2}\chi_{B_{5}^{\prime}}-2ijm_{l}\chi_{B_{5}^{\prime}\eta}\right]\ , (6.6)

where χB5\chi_{B_{5}^{\prime}} and χB5η\chi_{B_{5}^{\prime}\eta} represent a susceptibility function for the B5B_{5}^{\prime} channel and a mixed one between the η\eta and B5B_{5}^{\prime} channels, respectively. Those contributions are defined by

χB5\displaystyle\chi_{B_{5}^{\prime}} =\displaystyle= d4x0|T(12ψTCτc2τf2ψ+h.c.)(x)(12ψTCτc2τf2ψ+h.c.)(0)|0,\displaystyle\int d^{4}x\Big{\langle}0\Big{|}T\left(-\frac{1}{2}\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)(x)\left(-\frac{1}{2}\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)(0)\Big{|}0\Big{\rangle}\ ,
χB5η\displaystyle\chi_{B_{5}^{\prime}\eta} =\displaystyle= d4x0|T(ψ¯iγ5ψ)(x)(12ψTCτc2τf2ψ+h.c.)(0)|0.\displaystyle\int d^{4}x\Big{\langle}0\Big{|}T\Big{(}\bar{\psi}i\gamma_{5}\psi\Big{)}(x)\left(-\frac{1}{2}\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)(0)\Big{|}0\Big{\rangle}\ . (6.7)

The additional contributions (6.6) can be further reduced. That is, using an identity

i2ψTCγ5τc2τf2ψ+h.c.=ijχB4,\displaystyle-\frac{i}{2}\langle\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi\rangle+{\rm h.c.}=-ij\chi_{B_{4}}\ , (6.8)

with

χB4=d4x0|T(12ψTCγ5τc2τf2ψ+h.c.)(x)(12ψTCγ5τc2τf2ψ+h.c.)(0)|0,\displaystyle\chi_{B_{4}}=\int d^{4}x\langle 0|T\left(\frac{1}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)(x)\left(\frac{1}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)(0)\Big{|}0\Big{\rangle}\ , (6.9)

which is derived in Appendix B, the corrections δχtop\delta\chi_{\rm top} in Eq. (6.6) are rewritten in terms of the hadron susceptibility functions as

δχtop=χtop(mix)+χtop(B),\displaystyle\delta\chi_{\rm top}=\chi_{\rm top}^{(\rm mix)}+\chi_{\rm top}^{(\rm B)}\ , (6.10)

where#6#6#6Utilizing the matching condition (4.8) and the stationary condition for Δ\Delta in the presence of jj, one can show that the GOR-like relation with respect to the breakdown of U(1)BU(1)_{B} baryon-number symmetry reads fB2mB42=jψψ/2f_{B}^{2}m_{B_{4}}^{2}=-j\langle\psi\psi\rangle/2, with fB=Δ/2f_{B}=\Delta/\sqrt{2} being the corresponding decay constant and mB42=mπ24μq2m_{B_{4}}^{2}=m_{\pi}^{2}-4\mu_{q}^{2}. Here, ψψ\langle\psi\psi\rangle is identical to the LHS of Eq. (4.8). From this relation, χtop(B)\chi_{\rm top}^{(\rm B)} can be rewritten into χtop(B)=fB2mB422δm(B),\displaystyle\chi_{\rm top}^{(\rm B)}=\frac{f_{B}^{2}m_{B_{4}}^{2}}{2}\delta_{m}^{(\rm B)}\ , (6.11) with δm(B)=1χB5/χB4\delta_{m}^{(\rm B)}=1-\chi_{B_{5}^{\prime}}/\chi_{B_{4}}, analogous to the expression for the meson sector in Eq. (2.19).

χtop(mix)=i2mljχB5η,χtop(B)=i4j2(χB4χB5).\displaystyle\chi_{\rm top}^{(\rm mix)}=\frac{i}{2}m_{l}j\chi_{B_{5}^{\prime}\eta}\ ,\ \ \chi_{\rm top}^{(\rm B)}=\frac{i}{4}j^{2}(\chi_{B_{4}}-\chi_{B_{5}^{\prime}})\ . (6.12)

It is interesting to note that the baryonic contribution χtop(B)\chi_{\rm top}^{(\rm B)} is proportional to the difference of χB4\chi_{B_{4}} and χB5\chi_{B_{5}^{\prime}}, which takes a partner structure similarly to χtop(M)\chi_{\rm top}^{(\rm M)} argued in Sec. 2; the baryon susceptibility functions χB4\chi_{B_{4}} and χB5\chi_{B_{5}^{\prime}} are also transformed to each other under the U(1)BU(1)_{B} and U(1)AU(1)_{A} transformations. In the baryon sector, the partner structure reads

χB4{\chi_{B_{4}}}χB5{\chi_{B_{5}}}χB4{\chi_{B_{4}^{\prime}}}χB5{\chi_{B_{5}^{\prime}}}U(1)AU(1)_{A}U(1)BU(1)_{B}U(1)BU(1)_{B}U(1)AU(1)_{A}

as explicitly derived in Appendix A.

The susceptibility functions χB4\chi_{B_{4}}, χB5\chi_{B_{5}^{\prime}} and χB5η\chi_{B_{5}^{\prime}\eta} in Eq. (6.10) are evaluated within our linear sigma model by tracing a similar procedure in obtaining χη\chi_{\eta} and χπ\chi_{\pi} in Sec. 4.2. The B4B_{4} mode does not mix with other hadrons in the low-energy limit, so χB4\chi_{B_{4}} is simply expressed by

χB4=2ic¯21mB42,\displaystyle\chi_{B_{4}}=-2i\bar{c}^{2}\frac{1}{m_{B_{4}}^{2}}\ , (6.13)

where mB42=mπ24μq2m_{B_{4}}^{2}=m_{\pi}^{2}-4\mu_{q}^{2} [65]. Meanwhile, χB5\chi_{B_{5}^{\prime}} and χB5η\chi_{B_{5}^{\prime}\eta} are evaluated as

χB5=DB5B5(0),χB5η\displaystyle\chi_{B_{5}^{\prime}}=D_{B_{5}^{\prime}B_{5}^{\prime}}(0)\ ,\ \ \chi_{B_{5}^{\prime}\eta} =\displaystyle= DB5η(0),\displaystyle D_{B_{5}^{\prime}\eta}(0)\ , (6.14)

by inverting the matrix (4.22), which may be expressed in terms of three contributions of ϕ1\phi_{1}, ϕ2\phi_{2} and ϕ3\phi_{3} as done for Dηη(0)D_{\eta\eta}(0) in Eq. (4.24). Based on these expressions, we numerically investigate μq\mu_{q} dependences of χtopw/j\chi_{\rm top}^{\rm w/j} for several jj in the next subsection.

It should be noted that the non-conservation law of the U(1)AU(1)_{A} current in Eq. (2.11) is now modified as

μjAμ=2mlψ¯iγ5ψ+(jψTCτc2τf2ψ+h.c.)+g216π2ϵμνρσGμνaGρσa,\displaystyle\partial_{\mu}j_{A}^{\mu}=2m_{l}\bar{\psi}i\gamma_{5}\psi+\left(j\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\right)+\frac{g^{2}}{16\pi^{2}}\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G^{a}_{\rho\sigma}\ , (6.15)

where corrections of the diquark operator ψTCτc2τf2ψ\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi accompanied by the diquark source jj are present.

6.2 Diquark source effect on μq\mu_{q} dependence of topological susceptibility

In the presence of the diquark source jj, U(1)BU(1)_{B} baryon-number symmetry is explicitly broken even in the vacuum as understood from the modified quark mass term  in Eq. (6.2). In other words, Δ\Delta would be always nonzero in our linear sigma model, so that the phase structures are expected to be modified from the ones for j=0j=0. Then, before showing numerical results of the topological susceptibility (6.4), first we explore μq\mu_{q} dependences of σ0\sigma_{0} and Δ\Delta corresponding to the chiral condensate and the diquark condensate, to clarify the phase structures in the presence of jj.

In Fig. 6, we show the μq\mu_{q} dependence of the mean fields for j/ml=0.05j/m_{l}=0.05 and 0.180.18 with mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (a) and mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 (b). The figure indicates that the definite phase transition with respect to the baryon superfluidity disappears and the value of Δ\Delta continuously increases for j0j\neq 0, whereas the second-order phase transition has certainly occurred for j=0j=0 as seen from Fig. 1. Accompanied by such a continuous change of Δ\Delta, σ0\sigma_{0} also shows a similar smooth change. Besides, Fig. 6 indicates that σ0\sigma_{0} at μq0\mu_{q}\sim 0 is not significantly affected by the size of jj while Δ\Delta is significantly affected. This is because the diquark source field jj induces an additional tadpole term of Δ\Delta in the effective potential in Eq. (4.12) which only contributes to the stationary condition of Δ\Delta directly. Meanwhile, in the high-density region, jj contributions become negligible due to large μq\mu_{q}, so that the behavior of σ0\sigma_{0} (Δ\Delta) including the jj effect merges into the one for j=0j=0 there.

Refer to caption (a) Refer to caption (b)
Figure 6: μq\mu_{q} dependences of σ0\sigma_{0} and Δ\Delta for j/ml=0.05j/m_{l}=0.05 and 0.180.18 with mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1 (a) and mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5 (b).

Next, we show the diquark source effects on the topological susceptibility in Figs. 7 and 8. Figure 7 exhibits the μq\mu_{q} dependence of the topological susceptibility including diquark source effects in the absence of the KMT-type interaction: mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1. As depicted in panel (a), the net topological susceptibility χtopw/j\chi_{\rm top}^{{\rm w/j}} is null at any μq\mu_{q} regardless of the value of jj. This is because the gluonic U(1)AU(1)_{A} anomaly is mimicked by only the KMT-type interaction in the linear sigma model even with the diquark source field jj. We also analyze each component of the net topological susceptibility, χtop(M)\chi_{\rm top}^{(\rm M)}, χtop(mix)\chi_{\rm top}^{(\rm mix)} and χtop(B)\chi_{\rm top}^{(\rm B)} defined in Eqs. (6.5) and (6.12) for j/ml=0.18j/m_{l}=0.18, in panel (b). This panel clearly shows that the susceptibilities satisfy the relation χtop(M)=χtop(B)=12χtop(mix)\chi_{\rm top}^{(\rm M)}=\chi_{\rm top}^{(\rm B)}=-\frac{1}{2}\chi_{\rm top}^{(\rm mix)} to result in the null χtopw/j\chi_{\rm top}^{\rm w/j}. This notable relation can be analytically derived from the anomalous WTI associated with the U(1)AU(1)_{A} transformation as shown in Appendix C.

Refer to caption (a) Refer to caption (b)
Figure 7: μq\mu_{q} dependences of the topological susceptibility including the diquark source effects in the absence of the KMT-type interaction: mηvac/mπvac=1m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.

By taking the KMT-type interaction into account, the net topological susceptibility χtopw/j\chi_{\rm top}^{\rm w/j} becomes sensitive to the diquark source field especially above μqmπvac/2\mu_{q}\approx m_{\pi}^{\rm vac}/2 as depicted in Fig. 8. Panel (a) shows that the decreasing trend of χtopw/j\chi_{\rm top}^{\rm w/j} at higher μq\mu_{q} is hindered as we take the larger value of jj. Notably, when taking j/ml=0.18j/m_{l}=0.18, the net topological susceptibility approximately holds the vacuum value at any μq\mu_{q}. To grasp this behavior, we show separate contributions of χtop(M)\chi_{\rm top}^{(\rm M)}, χtop(B)\chi_{\rm top}^{(\rm B)} and χtop(mix)\chi_{\rm top}^{(\rm mix)} with j/ml=0.18j/m_{l}=0.18 in panel (b) of Fig. 8. This figure indicates that χtop(M)\chi_{\rm top}^{(\rm M)} is not substantially influenced by the diquark source and its decreasing behavior is governed by the smooth chiral restoration as explained in Sec. 5.2 in detail. In contrast, χtop(B)\chi_{\rm top}^{(\rm B)} is enhanced above μqmπvac/2\mu_{q}\approx m_{\pi}^{\rm vac}/2, which is understood by the increment of Δ\Delta. In fact, from the matching condition (4.8) and the stationary condition for Δ\Delta in the presence of jj, one can easily show χtop(B)=(jc¯)Δδm(B)/(22)\chi_{\rm top}^{(\rm B)}=(j\bar{c})\Delta\delta_{m}^{(\rm B)}/(2\sqrt{2}) with δm(B)=1χB5/χB4\delta_{m}^{(\rm B)}=1-\chi_{B_{5}^{\prime}}/\chi_{B_{4}}. Here, similarly to the discussion for the meson sector in Sec. 5.2, δm(B)\delta_{m}^{(\rm B)} approaches a constant value asymptotically in the high-density region. Therefore, we can prove that the growth of χtop(B)\chi_{\rm top}^{(\rm B)} can be determined by Δ\Delta at larger μq\mu_{q}. Hence, when the source contribution is sufficiently large, the net topological susceptibility can grow with increasing μq\mu_{q}. The last contribution, χtop(mix)\chi_{\rm top}^{(\rm mix)}, represents a mixing susceptibility between the mesonic and baryonic sectors, and this is suppressed compared to χtop(M)\chi_{\rm top}^{(\rm M)} and χtop(B)\chi_{\rm top}^{(\rm B)}, as long as jj is small, as shown in the figure. We note that, the mixing strength of B5B_{5}^{\prime} and η\eta becomes weak for larger value of cc as seen from mB5η2m_{B_{5}^{\prime}\eta}^{2} in Eq. (4.23), and hence, the larger cc we take, the smaller χtop(mix)\chi_{\rm top}^{(\rm mix)} we obtain.

To summarize, from the demonstration in this subsection, we have revealed that the diquark source jj contaminates the fate of the net topological susceptibility linked with the chiral restoration. Therefore, one can infer that the approximately μq\mu_{q}-independent behavior of the topological susceptibility exhibited by the lattice data [41] would be understood by the finite diquark source effects. Note that although the approximately μq\mu_{q}-independent behavior was found on the lattice at T=0.45TcT=0.45T_{c}, the temperature effects are expected to be insignificant. This is because the phase structure at T=0.45TcT=0.45T_{c} does not significantly differ from one at T=0T=0.

Refer to caption (a) Refer to caption (b)
Figure 8: μq\mu_{q} dependences of the topological susceptibility including the diquark source effects with the substantial anomaly effect of KMT-type interaction: mηvac/mπvac=1.5m_{\eta}^{\rm vac}/m_{\pi}^{\rm vac}=1.5.

7 Summary and discussion

In this paper, we have explored the topological susceptibility in QC2D with two flavors at finite quark chemical potential μq\mu_{q}, to clarify the U(1)AU(1)_{A} anomaly properties in cold and dense matter. With the help of the WTIs, we have found that the topological susceptibility is analytically expressed by a difference of the pion and η\eta-meson susceptibility functions with the current quark mass. We have also argued that, in the low-energy regime, this expression is understood as a generalization of the one invented by Leutwyler and Smilga based on the ChPT in three-color QCD [72].

In order to investigate the topological susceptibility at finite μq\mu_{q}, we have employed the linear sigma model in which the U(1)AU(1)_{A} anomaly effects are captured by the KMT-type determinant term, as a suitable low-energy effective theory of QC2[65]. This model successfully not only describes the emergence of the baryon superfluid phase but also reproduce the hadron mixings originated from the breakdown of U(1)BU(1)_{B} baryon-number symmetry there, which is indeed suggested by the lattice data [74]. Based on a mean-field treatment, we have found that the topological susceptibility is always zero at any μq\mu_{q} in both the hadronic and superfluid phases in the absence of the U(1)AU(1)_{A} anomaly effects, where the vacuum mass of pion coincides with one of η\eta meson. When the U(1)AU(1)_{A} anomaly effect is switched on, the nonzero and constant topological susceptibility is induced in the hadronic phase. Moving on to the superfluid phase, we have found that it begins to smoothly decrease with increasing μq\mu_{q}. We have analytically clarified that the latter smooth decrement is fitted by μq2\mu_{q}^{-2} at larger μq\mu_{q}, reflecting the continuous restoration of chiral symmetry. This property is qualitatively the same as in hot three-color QCD matter [10, 11, 12, 7, 21, 8]. From those examinations, we can conclude that, in cold and dense QC2D, roles of the topological susceptibility as an indicator for measuring the strength of U(1)AU(1)_{A} anomaly effects do not differ from those in hot three-color QCD, despite the complexity of phase structure due to the presence of the superfluidity.

In lattice simulations, effects from the diquark source would remain sizable. For this reason, we have further investigated the topological susceptibility in the presence of the diquark source. From this examination, we have revealed that the source effects enhance the topological susceptibility in accordance with the growth of the diquark condensate as μq\mu_{q} increases, such that the reduction of the topological susceptibility found in the presence of the U(1)AU(1)_{A} anomaly effects can be hindered. Hence, when the source contribution is sufficiently large, the topological susceptibility can grow with increasing μq\mu_{q}. On the other hand, when the U(1)AU(1)_{A} anomaly effects are absent, the topological susceptibility vanishes at any value of μq\mu_{q} consistently regardless of the size of the diquark source.

In closing, we give a list of some comments on our findings and its implications.

  • As argued in the later part of Sec. 2 in detail, the topological susceptibility in the vacuum is determined by only three basic observables: the pion decay constant, pion mass and η\eta mass, in the low-energy regime of QC2D. Thus, in order to pursue a consistent understanding of U(1)AU(1)_{A} anomaly effects in low-energy QC2D, we expect precise determination of both the decay constant and the η\eta mass as well as that of the topological susceptibility itself from lattice simulations. Those determinations would be regarded as a foundation toward more quantitative description of the topological susceptibility at finite μq\mu_{q}.

  • In the present analysis, we have used the linear sigma model based on a mean-field approach, and hence, all coupling constants in the model do not change at any μq\mu_{q}. On the basis of the functional renormalization group (FRG) method in three-color QCD, it was suggested that the coupling strength of the KMT-type interaction can be enhanced in medium, leading to the effective enhancement of the U(1)AU(1)_{A} anomaly effects [75, 76, 77, 78]. If this is the case, then the topological susceptibility can also be enhanced at finite μq\mu_{q}. Hence, analyses from effective models beyond the mean-field level such as the FRG method, in which fluctuations of the hadrons are non-perturbatively incorporated, are worth studying.

  • In our present study, we have focused on the topological susceptibility at finite μq\mu_{q} but with zero temperature. Currently, the μq\mu_{q} dependence on the topological susceptibility around the critical temperature has been also evaluated in the lattice QCD [41]. Thus, it would be worth investigating finite temperature effects to the topological susceptibility to fit the lattice data, to pursue more comprehensive description of the U(1)AU(1)_{A} anomaly effects on the phase diagram of QC2D.

  • In this study, we have clarified that the asymptotic behavior of the topological susceptibility at larger μq\mu_{q} is mostly determined by the smooth reduction of the chiral condenesate, despite the presence of the diquark condensate in dense QC2D. This structure is essentially understood by a fact that the WTIs used to express the topological susceptibility in terms of the meson susceptibility functions are not altered by the diquak condensate, unless the diquark source contributions remain finite. In ordinary three-color QCD, on the other hand, the WTI associated with the pion would be modified in the color-flavor locking (CFL) phase, since the CFL configuration changes the chiral-symmetry breaking pattern due to correlations from SU(3)cSU(3)_{c} color symmetry [79]. For this reason, it is not clear whether a similar asymptotic behavior is derived in cold and dense three-color QCD matter, and we leave this issue for a future study. Meanwhile, the two-flavor color superconductivity (2SC) is singlet under chiral symmetry, and hence, at intermediate density regime one can expect that a qualitatively similar behavior of the topological susceptibility follows even in the presence of the 2SC phase.

Acknowledgment

This work of M.K. is supported in part by the National Natural Science Foundation of China (NSFC) Grant Nos: 12235016, and the Strategic Priority Research Program of Chinese Academy of Sciences under Grant No XDB34030000. D.S. is supported by the RIKEN special postdoctoral researcher program. The authors thank K. Iida and E. Itou for fruitful discussion and useful information on their lattice results of the topological susceptibility.

Appendix A Partner structures of the susceptibility functions

In this appendix, we derive partner structures of the meson and diquark-baryon susceptibility functions with respect to appropriate transformations.

Here, we define the following composite operators

𝒪σψ¯ψ,𝒪a0aψ¯τfaψ,𝒪ηψ¯iγ5ψ,𝒪πaψ¯iγ5τfaψ,\displaystyle{\cal O}_{\sigma}\equiv\bar{\psi}\psi\ ,\ \ {\cal O}^{a}_{a_{0}}\equiv\bar{\psi}\tau_{f}^{a}\psi\ ,\ \ {\cal O}_{\eta}\equiv\bar{\psi}i\gamma_{5}\psi\ ,\ \ {\cal O}_{\pi}^{a}\equiv\bar{\psi}i\gamma_{5}\tau_{f}^{a}\psi\ ,
𝒪B412ψTCγ5τc2τf2ψ+h.c.,𝒪B5i2ψTCγ5τc2τf2ψ+h.c.,\displaystyle{\cal O}_{B_{4}}\equiv\frac{1}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\ ,\ \ {\cal O}_{B_{5}}\equiv-\frac{i}{2}\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi+{\rm h.c.}\ ,
𝒪B4=i2ψTCτc2τf2ψi2ψCτc2τf2ψ,𝒪B5=12ψTCτc2τf2ψ+12ψCτc2τf2ψ.\displaystyle{\cal O}_{B_{4}^{\prime}}=-\frac{i}{2}\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi-\frac{i}{2}\psi^{\dagger}C\tau_{c}^{2}\tau_{f}^{2}\psi^{*}\ ,\ \ {\cal O}_{B_{5}^{\prime}}=-\frac{1}{2}\psi^{T}C\tau_{c}^{2}\tau_{f}^{2}\psi+\frac{1}{2}\psi^{\dagger}C\tau_{c}^{2}\tau_{f}^{2}\psi^{*}\ .
(A.1)

The U(1)BU(1)_{B}, SU(2)VSU(2)_{V}, U(1)AU(1)_{A} and SU(2)ASU(2)_{A} rotations are generated by (a=1,2,3a=1,2,3)

ψU(1)BeiϵBψ,ψSU(2)VeiϵVaTfaψ,ψU(1)AeiϵAγ5ψ,ψSU(2)AeiϵAaTfaψ,\displaystyle\psi\overset{U(1)_{B}}{\to}{\rm e}^{-i\epsilon_{B}}\psi\ ,\ \ \psi\overset{SU(2)_{V}}{\to}{\rm e}^{-i\epsilon_{V}^{a}T_{f}^{a}}\psi\ ,\ \ \psi\overset{U(1)_{A}}{\to}{\rm e}^{-i\epsilon_{A}\gamma_{5}}\psi\ ,\ \ \psi\overset{SU(2)_{A}}{\to}{\rm e}^{-i\epsilon_{A}^{a}T_{f}^{a}}\psi\ , (A.2)

respectively. Then, the meson operators 𝒪σ{\cal O}_{\sigma}, 𝒪a0a{\cal O}^{a}_{a_{0}}, 𝒪η{\cal O}_{\eta} and 𝒪πa{\cal O}^{a}_{\pi} are invariant under the U(1)BU(1)_{B} and SU(2)VSU(2)_{V} rotations, while under the infinitesimal U(1)AU(1)_{A} and SU(2)ASU(2)_{A} ones they transform as

𝒪σU(1)A𝒪σ2ϵA𝒪η,𝒪a0aU(1)A𝒪a0a2ϵA𝒪πa,\displaystyle{\cal O}_{\sigma}\overset{U(1)_{A}}{\to}{\cal O}_{\sigma}-2\epsilon_{A}{\cal O}_{\eta}\ ,\ \ {\cal O}^{a}_{a_{0}}\overset{U(1)_{A}}{\to}{\cal O}^{a}_{a_{0}}-2\epsilon_{A}{\cal O}^{a}_{\pi}\ ,
𝒪ηU(1)A𝒪η+2ϵA𝒪σ,𝒪πaU(1)A𝒪πa+2ϵA𝒪a0a,\displaystyle{\cal O}_{\eta}\overset{U(1)_{A}}{\to}{\cal O}_{\eta}+2\epsilon_{A}{\cal O}_{\sigma}\ ,\ \ {\cal O}^{a}_{\pi}\overset{U(1)_{A}}{\to}{\cal O}^{a}_{\pi}+2\epsilon_{A}{\cal O}^{a}_{a_{0}}\ , (A.3)

and

𝒪σSU(2)A𝒪σϵAa𝒪πa,𝒪a0aSU(2)A𝒪a0aϵAa𝒪η,\displaystyle{\cal O}_{\sigma}\overset{SU(2)_{A}}{\to}{\cal O}_{\sigma}-\epsilon^{a}_{A}{\cal O}^{a}_{\pi}\ ,\ \ {\cal O}^{a}_{a_{0}}\overset{SU(2)_{A}}{\to}{\cal O}^{a}_{a_{0}}-\epsilon_{A}^{a}{\cal O}_{\eta}\ ,
𝒪ηSU(2)A𝒪η+ϵAa𝒪a0a,𝒪πaSU(2)A𝒪πa+ϵAa𝒪σ.\displaystyle{\cal O}_{\eta}\overset{SU(2)_{A}}{\to}{\cal O}_{\eta}+\epsilon_{A}^{a}{\cal O}^{a}_{a_{0}}\ ,\ \ {\cal O}^{a}_{\pi}\overset{SU(2)_{A}}{\to}{\cal O}^{a}_{\pi}+\epsilon_{A}^{a}{\cal O}_{\sigma}\ . (A.4)

Hence, one can find the following partner structure

χπ{\chi_{\pi}}χσ{\chi_{\sigma}}χa0{\chi_{a_{0}}}χη{\chi_{\eta}}U(1)AU(1)_{A}SU(2)SU(2)U(1)AU(1)_{A}

where the susceptibility functions are defined by

χσ=d4x0|T𝒪σ(x)𝒪σ(0)|0,χa0δab=d4x0|T𝒪a0a(x)𝒪a0b(0)|0,\displaystyle\chi_{\sigma}=\int d^{4}x\langle 0|T{\cal O}_{\sigma}(x){\cal O}_{\sigma}(0)|0\rangle\ ,\ \ \chi_{a_{0}}\delta^{ab}=\int d^{4}x\langle 0|T{\cal O}^{a}_{a_{0}}(x){\cal O}^{b}_{a_{0}}(0)|0\rangle\ ,
χη=d4x0|T𝒪η(x)𝒪η(0)|0,χπδab=d4x0|T𝒪πa(x)𝒪πb(0)|0.\displaystyle\chi_{\eta}=\int d^{4}x\langle 0|T{\cal O}_{\eta}(x){\cal O}_{\eta}(0)|0\rangle\ ,\ \ \chi_{\pi}\delta^{ab}=\int d^{4}x\langle 0|T{\cal O}^{a}_{\pi}(x){\cal O}^{b}_{\pi}(0)|0\rangle\ . (A.5)

Meanwhile, the diquark-baryon operators 𝒪B4{\cal O}_{B_{4}}, 𝒪B5{\cal O}_{B_{5}}, 𝒪B4{\cal O}_{B_{4}^{\prime}} and 𝒪B5{\cal O}_{B_{5}^{\prime}} are invariant under the SU(2)VSU(2)_{V} and SU(2)ASU(2)_{A} rotations, while under the U(1)BU(1)_{B} and U(1)AU(1)_{A} ones they transform as

𝒪B4U(1)B𝒪B4+2ϵB𝒪B5,𝒪B5U(1)B𝒪B52ϵB𝒪B4,\displaystyle{\cal O}_{B_{4}}\overset{U(1)_{B}}{\to}{\cal O}_{B_{4}}+2\epsilon_{B}{\cal O}_{B_{5}}\ ,\ \ {\cal O}_{B_{5}}\overset{U(1)_{B}}{\to}{\cal O}_{B_{5}}-2\epsilon_{B}{\cal O}_{B_{4}}\ ,
𝒪B4U(1)B𝒪B4+2ϵB𝒪B5,𝒪B5U(1)B𝒪B52ϵB𝒪B4,\displaystyle{\cal O}_{B_{4}^{\prime}}\overset{U(1)_{B}}{\to}{\cal O}_{B_{4}^{\prime}}+2\epsilon_{B}{\cal O}_{B_{5}^{\prime}}\ ,\ \ {\cal O}_{B_{5}^{\prime}}\overset{U(1)_{B}}{\to}{\cal O}_{B_{5}^{\prime}}-2\epsilon_{B}{\cal O}_{B_{4}^{\prime}}\ , (A.6)

and

𝒪B4U(1)A𝒪B4+2ϵA𝒪B4,𝒪B5U(1)A𝒪B5+2ϵA𝒪B5,\displaystyle{\cal O}_{B_{4}}\overset{U(1)_{A}}{\to}{\cal O}_{B_{4}}+2\epsilon_{A}{\cal O}_{B_{4}^{\prime}}\ ,\ \ {\cal O}_{B_{5}}\overset{U(1)_{A}}{\to}{\cal O}_{B_{5}}+2\epsilon_{A}{\cal O}_{B_{5}^{\prime}}\ ,
𝒪B4U(1)A𝒪B42ϵA𝒪B4,𝒪B5U(1)A𝒪B52ϵA𝒪B5.\displaystyle{\cal O}_{B_{4}^{\prime}}\overset{U(1)_{A}}{\to}{\cal O}_{B_{4}^{\prime}}-2\epsilon_{A}{\cal O}_{B_{4}}\ ,\ \ {\cal O}_{B_{5}^{\prime}}\overset{U(1)_{A}}{\to}{\cal O}_{B_{5}^{\prime}}-2\epsilon_{A}{\cal O}_{B_{5}}\ . (A.7)

Hence, similarly to the meson sector, one can find the following partner structure

χB4{\chi_{B_{4}}}χB5{\chi_{B_{5}}}χB4{\chi_{B_{4}^{\prime}}}χB5{\chi_{B_{5}^{\prime}}}U(1)AU(1)_{A}U(1)BU(1)_{B}U(1)BU(1)_{B}U(1)AU(1)_{A}

where the suseptibility functions are defined by

χB4=d4x0|T𝒪B4(x)𝒪B4(0)|0,χB5=d4x0|T𝒪B5(x)𝒪B5(0)|0,\displaystyle\chi_{B_{4}}=\int d^{4}x\langle 0|T{\cal O}_{B_{4}}(x){\cal O}_{B_{4}}(0)|0\rangle\ ,\ \ \chi_{B_{5}}=\int d^{4}x\langle 0|T{\cal O}_{B_{5}}(x){\cal O}_{B_{5}}(0)|0\rangle\ ,
χB4=d4x0|T𝒪B4(x)𝒪B4(0)|0,χB5=d4x0|T𝒪B5(x)𝒪B5(0)|0.\displaystyle\chi_{B_{4}^{\prime}}=\int d^{4}x\langle 0|T{\cal O}_{B_{4}^{\prime}}(x){\cal O}_{B_{4}^{\prime}}(0)|0\rangle\ ,\ \ \chi_{B_{5}^{\prime}}=\int d^{4}x\langle 0|T{\cal O}_{B_{5}^{\prime}}(x){\cal O}_{B_{5}^{\prime}}(0)|0\rangle\ . (A.8)

The infinitesimal transformation laws obtained in this appendix play important roles in deriving the WTIs employed in the present paper.

Appendix B Derivation of WTIs (2.12) and (6.8)

In this appendix, we derive the WTIs in Eqs. (2.12) and (6.8) which allow us to rewrite the topological susceptibility in terms of only the hadron susceptibility functions.

Toward derivation of Eq. (2.12), we try to perform the SU(2)ASU(2)_{A} rotation in the following path integral:

πa[dψ¯dψ][dA]𝒪πa(y)eid4xQC2D,\displaystyle{\cal I}^{a}_{\pi}\equiv\int[d\bar{\psi}d\psi][dA]\,{\cal O}_{\pi}^{a}(y)\,{\rm e}^{i\int d^{4}x{\cal L}_{\rm QC_{2}D}}\ , (B.1)

where the QC2D Lagrangian of interest here includes the diquark source term in addition to the mass term as

QC2D=ψ¯i/Dψml𝒪σj𝒪B514GμνaGμν,a+θg264π2ϵμνρσGμνaGρσa.\displaystyle{\cal L}_{\rm QC_{2}D}=\bar{\psi}i{\ooalign{\hfil/\hfil\crcr$D$}}\psi-m_{l}{\cal O}_{\sigma}-j{\cal O}_{B_{5}}-\frac{1}{4}G_{\mu\nu}^{a}G^{\mu\nu,a}+\theta\frac{g^{2}}{64\pi^{2}}\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a}\ . (B.4)

In this Lagrangian, the covariant derivative Dμψ=(μiμqδμ0igAμaTca)ψD_{\mu}\psi=(\partial_{\mu}-i\mu_{q}\delta_{\mu 0}-igA_{\mu}^{a}T_{c}^{a})\psi describes contributions from the quark chemical potential μq\mu_{q} and couplings with the gluons AμaA_{\mu}^{a}, and Gμνa=μAνaνAμa+gϵabcAμbAνbG_{\mu\nu}^{a}=\partial_{\mu}A_{\nu}^{a}-\partial_{\nu}A_{\mu}^{a}+g\epsilon^{abc}A_{\mu}^{b}A_{\nu}^{b} is the gluon field strength. Under the infinitesimal local SU(2)ASU(2)_{A} rotation, 𝒪πa{\cal O}^{a}_{\pi} transforms as shown in Eq. (A.4), while the QC2D Lagrangian exhibits the following transformation law:

QC2DSU(2)AQC2D+12(μϵAa)jAμ,a+mlϵAa𝒪πa,\displaystyle{\cal L}_{\rm QC_{2}D}\overset{SU(2)_{A}}{\to}{\cal L}_{\rm QC_{2}D}+\frac{1}{2}(\partial_{\mu}\epsilon_{A}^{a})j_{A}^{\mu,a}+m_{l}\epsilon_{A}^{a}{\cal O}_{\pi}^{a}\ , (B.5)

where jAμ,aψ¯γμγ5τfaψj_{A}^{\mu,a}\equiv\bar{\psi}\gamma^{\mu}\gamma_{5}\tau_{f}^{a}\psi represents the axial current. Thus, under the same rotation, Eq. (B.1) transforms as

πa\displaystyle{\cal I}_{\pi}^{a} SU(2)A\displaystyle\overset{SU(2)_{A}}{\to} πa+[dψ¯dψ][dA]{ϵAa(y)𝒪σ(y)\displaystyle{\cal I}_{\pi}^{a}+\int[d\bar{\psi}d\psi][dA]\Bigg{\{}\epsilon_{A}^{a}(y){\cal O}_{\sigma}(y) (B.6)
+id4xϵAb(x)[12μxjAμ,b(x)𝒪πa(y)+ml𝒪πb(x)𝒪πa(y)]}eid4xQC2D,\displaystyle+i\int d^{4}x\epsilon_{A}^{b}(x)\left[-\frac{1}{2}\partial_{\mu}^{x}j_{A}^{\mu,b}(x){\cal O}_{\pi}^{a}(y)+m_{l}{\cal O}_{\pi}^{b}(x){\cal O}_{\pi}^{a}(y)\right]\Bigg{\}}\,{\rm e}^{i\int d^{4}x{\cal L}_{\rm QC_{2}D}}\ ,

and imposing the invariance of πa{\cal I}_{\pi}^{a} under the SU(2)ASU(2)_{A} transformation, one can obtain the following WTI

𝒪σδab=id4x[12μx0|TjAμ,b(x)𝒪πa(y)|0ml0|T𝒪πb(x)𝒪πa(y)|0].\displaystyle\langle{\cal O}_{\sigma}\rangle\delta^{ab}=i\int d^{4}x\left[\frac{1}{2}\partial_{\mu}^{x}\langle 0|Tj_{A}^{\mu,b}(x){\cal O}_{\pi}^{a}(y)|0\rangle-m_{l}\langle 0|T{\cal O}_{\pi}^{b}(x){\cal O}_{\pi}^{a}(y)|0\rangle\right]\ . (B.7)

Here, since the QC2D Lagrangian explicitly breaks SU(2)L×SU(2)RSU(2)_{L}\times SU(2)_{R} chiral symmetry, there is no room for massless modes coupled to the axial current jAμ,bj_{A}^{\mu,b}, so that the first term of the RHS in Eq. (B.7) trivially vanishes from the surface integral. Therefore, we arrive at

ψ¯ψ=imlχπ,\displaystyle\langle\bar{\psi}\psi\rangle=-im_{l}\chi_{\pi}\ , (B.8)

with Eq. (A.5).

Similarly to the above derivation, the identity (6.8) is also derived by focusing on the U(1)BU(1)_{B} transformation of

B4[dψ¯dψ][dA]𝒪B4(y)eid4xQC2D.\displaystyle{\cal I}_{B_{4}}\equiv\int[d\bar{\psi}d\psi][dA]\,{\cal O}_{B_{4}}(y)\,{\rm e}^{i\int d^{4}x{\cal L}_{\rm QC_{2}D}}\ . (B.9)

In fact, under the infinitesimal local U(1)BU(1)_{B} rotation, 𝒪B4{\cal O}_{B_{4}} transforms as in Eq. (A.6) and the QC2D Lagrangian shows the following transformation law:

QC2DU(1)BQC2D+(μϵB)jBμ+2ϵBj𝒪B4,\displaystyle{\cal L}_{\rm QC_{2}D}\overset{U(1)_{B}}{\to}{\cal L}_{\rm QC_{2}D}+(\partial_{\mu}\epsilon_{B})j_{B}^{\mu}+2\epsilon_{B}j{\cal O}_{B_{4}}\ , (B.10)

where we have defined the vector current by jBμψ¯γμψj_{B}^{\mu}\equiv\bar{\psi}\gamma^{\mu}\psi. Thus, the U(1)BU(1)_{B} invariance of Eq. (B.9) yields

2𝒪B5=id4x[μx0|TjBμ(x)𝒪B4(y)|02j0|T𝒪B4(x)𝒪B4(y)|0].\displaystyle 2\langle{\cal O}_{B_{5}}\rangle=i\int d^{4}x\left[\partial_{\mu}^{x}\langle 0|Tj_{B}^{\mu}(x){\cal O}_{B_{4}}(y)|0\rangle-2j\langle 0|T{\cal O}_{B_{4}}(x){\cal O}_{B_{4}}(y)|0\rangle\right]\ . (B.11)

Here, the first term of the RHS vanishes since no massless modes couple to the vector current jBμj_{B}^{\mu} owing to the presence of jj term which violates U(1)BU(1)_{B} symmetry explicitly, and thus, one can obtain

i2ψTCγ5τc2τf2ψ+h.c.=ijχB4\displaystyle-\frac{i}{2}\langle\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi\rangle+{\rm h.c.}=-ij\chi_{B_{4}} (B.12)

by defining

χB4=d4x0|T𝒪B4(x)𝒪B4(0)|0,\displaystyle\chi_{B_{4}}=\int d^{4}x\langle 0|T{\cal O}_{B_{4}}(x){\cal O}_{B_{4}}(0)|0\rangle\ , (B.13)

with Eq. (A.8).

Appendix C Alternative expression of the topological susceptibility

In this appendix, we present an alternative expression of the topological susceptibility χtop\chi_{\rm top}.

For this purpose, here, we consider the U(1)AU(1)_{A} transformations in the following two path integrals:

η\displaystyle{\cal I}_{\eta} \displaystyle\equiv [dψ¯dψ][dA]𝒪η(y)eid4xQC2D,\displaystyle\int[d\bar{\psi}d\psi][dA]\,{\cal O}_{\eta}(y)\,{\rm e}^{i\int d^{4}x{\cal L}_{\rm QC_{2}D}}\ ,
B5\displaystyle{\cal I}_{B_{5}^{\prime}} \displaystyle\equiv [dψ¯dψ][dA]𝒪B5(y)eid4xQC2D.\displaystyle\int[d\bar{\psi}d\psi][dA]\,{\cal O}_{B_{5}^{\prime}}(y)\,{\rm e}^{i\int d^{4}x{\cal L}_{\rm QC_{2}D}}\ . (C.1)

Under the infinitesimal local U(1)AU(1)_{A} rotation, 𝒪η{\cal O}_{\eta} and 𝒪B5{\cal O}_{B_{5}^{\prime}} transform as in Eqs. (A.3) and (A.7) while the QC2D Lagrangian shows the followng transformation law:

QC2DU(1)AQC2D(μϵA)jAμ+ϵA(2ml𝒪η2j𝒪B5),\displaystyle{\cal L}_{\rm QC_{2}D}\overset{U(1)_{A}}{\to}{\cal L}_{\rm QC_{2}D}-(\partial_{\mu}\epsilon_{A})j_{A}^{\mu}+\epsilon_{A}(2m_{l}{\cal O}_{\eta}-2j{\cal O}_{B_{5}^{\prime}})\ , (C.2)

with the flavor-singlet U(1)AU(1)_{A} axial current defined by jAμψ¯γμγ5ψj_{A}^{\mu}\equiv\bar{\psi}\gamma^{\mu}\gamma_{5}\psi. Thus, tracing a similar procedure in deriving the WTIs (B.8) or (B.13), one can find

ψ¯ψ=i(mlχηjχB5η+2χQη),\displaystyle\langle\bar{\psi}\psi\rangle=-i\left(m_{l}\chi_{\eta}-j\chi_{B_{5}^{\prime}\eta}+2\chi_{Q\eta}\right)\ ,
i2ψTCγ5τc2τf2ψ+h.c.=i(mlχB5ηjχB5+2χQB5),\displaystyle-\frac{i}{2}\langle\psi^{T}C\gamma_{5}\tau_{c}^{2}\tau_{f}^{2}\psi\rangle+{\rm h.c.}=i\left(m_{l}\chi_{B_{5}^{\prime}\eta}-j\chi_{B_{5}^{\prime}}+2\chi_{QB_{5}^{\prime}}\right)\ , (C.3)

where we have defined mixed susceptibility functions by

χB5η=d4x0|T𝒪B5(x)𝒪η(0)|0,\displaystyle\chi_{B_{5}^{\prime}\eta}=\int d^{4}x\langle 0|T{\cal O}_{B_{5}^{\prime}}(x){\cal O}_{\eta}(0)|0\rangle\ ,
χQη=d4x0|TQ(x)𝒪η(0)|0,\displaystyle\chi_{Q\eta}=\int d^{4}x\langle 0|TQ(x){\cal O}_{\eta}(0)|0\rangle\ ,
χQB5=d4x0|TQ(x)𝒪B5(0)|0,\displaystyle\chi_{QB_{5}^{\prime}}=\int d^{4}x\langle 0|TQ(x){\cal O}_{B_{5}^{\prime}}(0)|0\rangle\ , (C.4)

with the gluonic topological operator Q=(g2/64π2)ϵμνρσGμνaGρσaQ=(g^{2}/64\pi^{2})\epsilon^{\mu\nu\rho\sigma}G_{\mu\nu}^{a}G_{\rho\sigma}^{a}. It should be noted that the U(1)AU(1)_{A} anomaly contributions have been properly incorporated when performing the U(1)AU(1)_{A} axial transformation in η{\cal I}_{\eta} and B5{\cal I}_{B_{5}^{\prime}} in Eq (C.1).

Here, using Eqs. (B.8) and (B.13), the anomalous WTIs in Eq. (C.3) are rewritten into

χtop(M)=12χtop(mix)+i2mlχQη,χtop(B)=12χtop(mix)i2jχQB5.\displaystyle\chi_{\rm top}^{({\rm M})}=-\frac{1}{2}\chi_{\rm top}^{({\rm mix})}+\frac{i}{2}m_{l}\chi_{Q\eta}\ ,\ \ \chi_{\rm top}^{({\rm B})}=-\frac{1}{2}\chi_{\rm top}^{({\rm mix})}-\frac{i}{2}j\chi_{QB_{5}^{\prime}}\ . (C.5)

Therefore, when we suppose that the U(1)AU(1)_{A} anomaly effects can be neglected for some reason, the operator QQ vanishes and

χtop(M)=χtop(B)=12χtop(mix)\displaystyle\chi_{\rm top}^{({\rm M})}=\chi_{\rm top}^{({\rm B})}=-\frac{1}{2}\chi_{\rm top}^{({\rm mix})} (C.6)

is satisfied. Indeed, this relation is consistent with the numerical results in Fig. 7 (a) where the U(1)AU(1)_{A} anomaly effects are absent.

References

  • [1] Steven Weinberg. The U(1) Problem. Phys. Rev. D, 11:3583–3593, 1975.
  • [2] Michael Buballa. NJL model analysis of quark matter at large density. Phys. Rept., 407:205–376, 2005.
  • [3] Gerard ’t Hooft. How Instantons Solve the U(1) Problem. Phys. Rept., 142:357–387, 1986.
  • [4] Tetsuo Hatsuda and Teiji Kunihiro. QCD phenomenology based on a chiral effective Lagrangian. Phys. Rept., 247:221–367, 1994.
  • [5] A. Gómez Nicola and J. Ruiz de Elvira. Pseudoscalar susceptibilities and quark condensates: chiral restoration and lattice screening masses. JHEP, 03:186, 2016.
  • [6] A. Gomez Nicola and J. Ruiz de Elvira. Patterns and partners for chiral symmetry restoration. Phys. Rev. D, 97(7):074016, 2018.
  • [7] Mamiya Kawaguchi, Shinya Matsuzaki, and Akio Tomiya. Analysis of nonperturbative flavor violation at chiral crossover criticality in QCD. Phys. Rev. D, 103(5):054034, 2021.
  • [8] Chuan-Xin Cui, Jin-Yang Li, Shinya Matsuzaki, Mamiya Kawaguchi, and Akio Tomiya. New interpretation of the chiral phase transition: Violation of the trilemma in QCD. Phys. Rev. D, 105(11):114031, 2022.
  • [9] Chuan-Xin Cui, Mamiya Kawaguchi, Jin-Yang Li, Shinya Matsuzaki, and Akio Tomiya. New aspect of chiral and axial breaking in QCD. 5 2022.
  • [10] Peter Petreczky, Hans-Peter Schadler, and Sayantan Sharma. The topological susceptibility in finite temperature QCD and axion cosmology. Phys. Lett. B, 762:498–505, 2016.
  • [11] Sz. Borsanyi et al. Calculation of the axion mass based on high-temperature lattice quantum chromodynamics. Nature, 539(7627):69–71, 2016.
  • [12] Claudio Bonati, Massimo D’Elia, Guido Martinelli, Francesco Negro, Francesco Sanfilippo, and Antonino Todaro. Topology in full QCD at high temperature: a multicanonical approach. JHEP, 11:170, 2018.
  • [13] Thomas D. Cohen. The High temperature phase of QCD and U(1)-A symmetry. Phys. Rev. D, 54:R1867–R1870, 1996.
  • [14] Sinya Aoki, Hidenori Fukaya, and Yusuke Taniguchi. Chiral symmetry restoration, eigenvalue density of Dirac operator and axial U(1) anomaly at finite temperature. Phys. Rev. D, 86:114512, 2012.
  • [15] A. Tomiya, G. Cossu, S. Aoki, H. Fukaya, S. Hashimoto, T. Kaneko, and J. Noaki. Evidence of effective axial U(1) symmetry restoration at high temperature QCD. Phys. Rev. D, 96(3):034509, 2017. [Addendum: Phys.Rev.D 96, 079902 (2017)].
  • [16] S. Aoki, Y. Aoki, G. Cossu, H. Fukaya, S. Hashimoto, T. Kaneko, C. Rohrhofer, and K. Suzuki. Study of the axial U(1)U(1) anomaly at high temperature with lattice chiral fermions. Phys. Rev. D, 103(7):074506, 2021.
  • [17] Kei Suzuki, Sinya Aoki, Yasumichi Aoki, Guido Cossu, Hidenori Fukaya, Shoji Hashimoto, and Christian Rohrhofer. Axial U(1) symmetry and mesonic correlators at high temperature in Nf=2N_{f}=2 lattice QCD. PoS, LATTICE2019:178, 2020.
  • [18] A. Bazavov et al. The chiral transition and U(1)AU(1)_{A} symmetry restoration from lattice QCD using Domain Wall Fermions. Phys. Rev. D, 86:094503, 2012.
  • [19] Michael I. Buchoff et al. QCD chiral transition, U(1)A symmetry and the dirac spectrum using domain wall fermions. Phys. Rev. D, 89(5):054514, 2014.
  • [20] Tanmoy Bhattacharya et al. QCD Phase Transition with Chiral Quarks and Physical Quark Masses. Phys. Rev. Lett., 113(8):082001, 2014.
  • [21] Mamiya Kawaguchi, Shinya Matsuzaki, and Akio Tomiya. Nonperturbative flavor breaking in topological susceptibility at chiral crossover. Phys. Lett. B, 813:136044, 2021.
  • [22] Gert Aarts. Introductory lectures on lattice QCD at nonzero baryon number. J. Phys. Conf. Ser., 706(2):022004, 2016.
  • [23] Shin Muroya, Atsushi Nakamura, Chiho Nonaka, and Tetsuya Takaishi. Lattice QCD at finite density: An Introductory review. Prog. Theor. Phys., 110:615–668, 2003.
  • [24] Simon Hands, John B. Kogut, Maria-Paola Lombardo, and Susan E. Morrison. Symmetries and spectrum of SU(2) lattice gauge theory at finite chemical potential. Nucl. Phys. B, 558:327–346, 1999.
  • [25] J. B. Kogut, D. K. Sinclair, S. J. Hands, and S. E. Morrison. Two color QCD at nonzero quark number density. Phys. Rev. D, 64:094505, 2001.
  • [26] Simon Hands, Istvan Montvay, Luigi Scorzato, and Jonivar Skullerud. Diquark condensation in dense adjoint matter. Eur. Phys. J. C, 22:451–461, 2001.
  • [27] Shin Muroya, Atsushi Nakamura, and Chiho Nonaka. Behavior of hadrons at finite density: Lattice study of color SU(2) QCD. Phys. Lett. B, 551:305–310, 2003.
  • [28] Shailesh Chandrasekharan and Fu-Jiun Jiang. Phase-diagram of two-color lattice QCD in the chiral limit. Phys. Rev. D, 74:014506, 2006.
  • [29] Simon Hands, Seyong Kim, and Jon-Ivar Skullerud. Deconfinement in dense 2-color QCD. Eur. Phys. J. C, 48:193, 2006.
  • [30] B. Alles, Massimo D’Elia, and M. P. Lombardo. Behaviour of the topological susceptibility in two colour QCD across the finite density transition. Nucl. Phys. B, 752:124–139, 2006.
  • [31] Simon Hands, Peter Sitch, and Jon-Ivar Skullerud. Hadron Spectrum in a Two-Colour Baryon-Rich Medium. Phys. Lett. B, 662:405–412, 2008.
  • [32] Simon Hands, Seyong Kim, and Jon-Ivar Skullerud. A Quarkyonic Phase in Dense Two Color Matter? Phys. Rev. D, 81:091502, 2010.
  • [33] Simon Hands and Philip Kenny. Topological Fluctuations in Dense Matter with Two Colors. Phys. Lett. B, 701:373–377, 2011.
  • [34] Seamus Cotter, Pietro Giudice, Simon Hands, and Jon-Ivar Skullerud. Towards the phase diagram of dense two-color matter. Phys. Rev. D, 87(3):034507, 2013.
  • [35] Simon Hands, Seyong Kim, and Jon-Ivar Skullerud. Non-relativistic spectrum of two-color QCD at non-zero baryon density. Phys. Lett. B, 711:199–204, 2012.
  • [36] Tamer Boz, Seamus Cotter, Leonard Fister, Dhagash Mehta, and Jon-Ivar Skullerud. Phase transitions and gluodynamics in 2-colour matter at high density. Eur. Phys. J. A, 49:87, 2013.
  • [37] V. V. Braguta, E. M. Ilgenfritz, A. Yu. Kotov, A. V. Molochkov, and A. A. Nikolaev. Study of the phase diagram of dense two-color QCD within lattice simulation. Phys. Rev. D, 94(11):114510, 2016.
  • [38] M. Puhr and P. V. Buividovich. Numerical Study of Nonperturbative Corrections to the Chiral Separation Effect in Quenched Finite-Density QCD. Phys. Rev. Lett., 118(19):192003, 2017.
  • [39] Tamer Boz, Ouraman Hajizadeh, Axel Maas, and Jon-Ivar Skullerud. Finite-density gauge correlation functions in QC2D. Phys. Rev. D, 99(7):074514, 2019.
  • [40] N. Yu. Astrakhantsev, V. G. Bornyakov, V. V. Braguta, E. M. Ilgenfritz, A. Yu. Kotov, A. A. Nikolaev, and A. Rothkopf. Lattice study of static quark-antiquark interactions in dense quark matter. JHEP, 05:171, 2019.
  • [41] Kei Iida, Etsuko Itou, and Tong-Gyu Lee. Two-colour QCD phases and the topology at low temperature and high density. JHEP, 01:181, 2020.
  • [42] Jonas Wilhelm, Lukas Holicki, Dominik Smith, Björn Wellegehausen, and Lorenz von Smekal. Continuum Goldstone spectrum of two-color QCD at finite density with staggered quarks. Phys. Rev. D, 100(11):114507, 2019.
  • [43] P. V. Buividovich, D. Smith, and L. von Smekal. Numerical study of the chiral separation effect in two-color QCD at finite density. Phys. Rev. D, 104(1):014511, 2021.
  • [44] Kei Iida, Etsuko Itou, and Tong-Gyu Lee. Relative scale setting for two-color QCD with NfN_{f}=2 Wilson fermions. PTEP, 2021(1):013B05, 2021.
  • [45] N. Astrakhantsev, V. V. Braguta, E. M. Ilgenfritz, A. Yu. Kotov, and A. A. Nikolaev. Lattice study of thermodynamic properties of dense QC2D. Phys. Rev. D, 102(7):074507, 2020.
  • [46] V. G. Bornyakov, V. V. Braguta, A. A. Nikolaev, and R. N. Rogalyov. Effects of Dense Quark Matter on Gluon Propagators in Lattice QC2D. Phys. Rev. D, 102:114511, 2020.
  • [47] P. V. Buividovich, D. Smith, and L. von Smekal. Electric conductivity in finite-density SU(2)SU(2) lattice gauge theory with dynamical fermions. Phys. Rev. D, 102(9):094510, 2020.
  • [48] P. V. Buividovich, D. Smith, and L. von Smekal. Static magnetic susceptibility in finite-density SU(2)SU\left(2\right) lattice gauge theory. Eur. Phys. J. A, 57(10):293, 2021.
  • [49] Kei Iida and Etsuko Itou. Velocity of sound beyond the high-density relativistic limit from lattice simulation of dense two-color QCD. PTEP, 2022(11):111B01, 2022.
  • [50] J. B. Kogut, Misha A. Stephanov, and D. Toublan. On two color QCD with baryon chemical potential. Phys. Lett. B, 464:183–191, 1999.
  • [51] J. B. Kogut, Misha A. Stephanov, D. Toublan, J. J. M. Verbaarschot, and A. Zhitnitsky. QCD - like theories at finite baryon density. Nucl. Phys. B, 582:477–513, 2000.
  • [52] J. T. Lenaghan, F. Sannino, and K. Splittorff. The Superfluid and conformal phase transitions of two color QCD. Phys. Rev. D, 65:054002, 2002.
  • [53] Claudia Ratti and Wolfram Weise. Thermodynamics of two-colour QCD and the Nambu Jona-Lasinio model. Phys. Rev. D, 70:054013, 2004.
  • [54] Gao-feng Sun, Lianyi He, and Pengfei Zhuang. BEC-BCS crossover in the Nambu-Jona-Lasinio model of QCD. Phys. Rev. D, 75:096004, 2007.
  • [55] Tomas Brauner, Kenji Fukushima, and Yoshimasa Hidaka. Two-color quark matter: U(1)(A) restoration, superfluidity, and quarkyonic phase. Phys. Rev. D, 80:074035, 2009. [Erratum: Phys.Rev.D 81, 119904 (2010)].
  • [56] Takuya Kanazawa, Tilo Wettig, and Naoki Yamamoto. Chiral Lagrangian and spectral sum rules for dense two-color QCD. JHEP, 08:003, 2009.
  • [57] Masayasu Harada, Chiho Nonaka, and Tetsuro Yamaoka. Masses of vector bosons in two-color dense QCD based on the hidden local symmetry. Phys. Rev. D, 81:096003, 2010.
  • [58] Nils Strodthoff, Bernd-Jochen Schaefer, and Lorenz von Smekal. Quark-meson-diquark model for two-color QCD. Phys. Rev. D, 85:074007, 2012.
  • [59] Daiki Suenaga and Toru Kojo. Gluon propagator in two-color dense QCD: Massive Yang-Mills approach at one-loop. Phys. Rev. D, 100(7):076017, 2019.
  • [60] Romain Contant and Markus Q. Huber. Dense two-color QCD from Dyson-Schwinger equations. Phys. Rev. D, 101(1):014016, 2020.
  • [61] T. G. Khunjua, K. G. Klimenko, and R. N. Zhokhov. The dual properties of chiral and isospin asymmetric dense quark matter formed of two-color quarks. JHEP, 06:148, 2020.
  • [62] Toru Kojo and Daiki Suenaga. Thermal quarks and gluon propagators in two-color dense QCD. Phys. Rev. D, 103(9):094008, 2021.
  • [63] Daiki Suenaga and Toru Kojo. Delineating chiral separation effect in two-color dense QCD. Phys. Rev. D, 104(3):034038, 2021.
  • [64] Toru Kojo and Daiki Suenaga. Peaks of sound velocity in two color dense QCD: Quark saturation effects and semishort range correlations. Phys. Rev. D, 105(7):076001, 2022.
  • [65] Daiki Suenaga, Kotaro Murakami, Etsuko Itou, and Kei Iida. Probing the hadron mass spectrum in dense two-color QCD with the linear sigma model. Phys. Rev. D, 107(5):054001, 2023.
  • [66] M. Kobayashi and T. Maskawa. Chiral symmetry and eta-x mixing. Prog. Theor. Phys., 44:1422–1424, 1970.
  • [67] M. Kobayashi, H. Kondo, and T. Maskawa. Symmetry breaking of the chiral u(3) x u(3) and the quark model. Prog. Theor. Phys., 45:1955–1959, 1971.
  • [68] Gerard ’t Hooft. Computation of the Quantum Effects Due to a Four-Dimensional Pseudoparticle. Phys. Rev. D, 14:3432–3450, 1976. [Erratum: Phys.Rev.D 18, 2199 (1978)].
  • [69] Gerard ’t Hooft. Symmetry Breaking Through Bell-Jackiw Anomalies. Phys. Rev. Lett., 37:8–11, 1976.
  • [70] Kazuo Fujikawa. Path Integral Measure for Gauge Invariant Fermion Theories. Phys. Rev. Lett., 42:1195–1198, 1979.
  • [71] T.P. Cheng and L.F. Li. Gauge Theory of Elementary Particle Physics. OUP Oxford, 1994.
  • [72] H. Leutwyler and Andrei V. Smilga. Spectrum of Dirac operator and role of winding number in QCD. Phys. Rev. D, 46:5607–5632, 1992.
  • [73] Max A. Metlitski and Ariel R. Zhitnitsky. Theta-parameter in 2 color QCD at finite baryon and isospin density. Nucl. Phys. B, 731:309–334, 2005.
  • [74] Kotaro Murakami, Daiki Suenaga, Kei Iida, and Etsuko Itou. Measurement of hadron masses in 2-color finite density QCD. PoS, LATTICE2022:154, 2023.
  • [75] G. Fejos and A. Hosaka. Thermal properties and evolution of the UA(1)U_{A}(1) factor for 2+1 flavors. Phys. Rev. D, 94(3):036005, 2016.
  • [76] G. Fejős and A. Hosaka. Mesonic and nucleon fluctuation effects at finite baryon density. Phys. Rev. D, 95:116011, 2017.
  • [77] G. Fejos and A. Hosaka. Axial anomaly and hadronic properties in a nuclear medium. Phys. Rev. D, 98(3):036009, 2018.
  • [78] G. Fejös and A. Patkos. Backreaction of mesonic fluctuations on the axial anomaly at finite temperature. Phys. Rev. D, 105(9):096007, 2022.
  • [79] Mark G. Alford, Andreas Schmitt, Krishna Rajagopal, and Thomas Schäfer. Color superconductivity in dense quark matter. Rev. Mod. Phys., 80:1455–1515, 2008.