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

Phase Diagram of the Two-Flavor Schwinger Model at Zero Temperature

Ross Dempsey,1 Igor R. Klebanov,1,2 Silviu S. Pufu,1,2 Benjamin T. Søgaard,1 and Bernardo Zan1 1Joseph Henry Laboratories, Princeton University, Princeton, NJ 08544, USA 2Princeton Center for Theoretical Science, Princeton University, Princeton, NJ 08544, USA
Abstract

We examine the phase structure of the two-flavor Schwinger model as a function of the θ\theta-angle and the two masses, m1m_{1} and m2m_{2}. In particular, we find interesting effects at θ=π\theta=\pi: along the SU(2)\mathrm{SU}(2)-invariant line m1=m2=mm_{1}=m_{2}=m, in the regime where mm is much smaller than the charge gg, the theory undergoes logarithmic RG flow of the Berezinskii-Kosterlitz-Thouless type. As a result, dimensional transmutation takes place, leading to a non-perturbatively small mass gap eAg2/m2\sim e^{-Ag^{2}/m^{2}}. The SU(2)\mathrm{SU}(2)-invariant line lies within a region of the phase diagram where the charge conjugation symmetry is spontaneously broken and whose boundaries we determine numerically. Our numerical results are obtained using the Hamiltonian lattice gauge formulation that includes the mass shift mlat=mg2a/4m_{\rm lat}=m-g^{2}a/4 dictated by the discrete chiral symmetry.

pacs:
11.15.Ha, 71.10.Fd, 11.15.-q, 75.30.Kz
preprint: PUPT-2640

I Introduction

Quantum Electrodynamics (QED) in 1+11+1 dimensions, also known as the Schwinger model Schwinger:1962tp , is a famous model of Quantum Field Theory (QFT) that has played an important role for over 60 years Lowenstein:1971fc ; Casher:1974vf ; Coleman:1975pw . It is a useful theoretical laboratory for various important phenomena, including QFT anomalies and confinement of charge. Its lattice Hamiltonian implementations Kogut:1974ag ; Banks:1975gq have connections with condensed matter and atomic physics, and, in recent years, there have been efforts to construct experimental setups for its quantum simulations (for a review, see Banuls:2019bmf ).

The model with one massless Dirac fermion of charge gg is exactly solvable, reducing to the non-interacting Schwinger boson of mass MS=g/πM_{S}=g/\sqrt{\pi}; this can be concisely demonstrated via the bosonization of the fermion Coleman:1975pw . The U(1)\mathrm{U}(1) chiral symmetry of the massless action is broken by the Schwinger anomaly. The massive model, in addition to containing the obvious dimensionless parameter m/gm/g, depends on the θ\theta angle related to the introduction of a background electric field Coleman:1975pw . This parameter, which has periodicity 2π2\pi, is somewhat analogous to the θ\theta angle of the 3+13+1 dimensional gauge theory.

Generalizations of the Schwinger model to Nf>1N_{f}>1 flavors of fermions of charge gg exhibit a richer set of phenomena Coleman:1976uz . When the fermions are massless, the Schwinger model has SU(Nf)×SU(Nf)\mathrm{SU}(N_{f})\times\mathrm{SU}(N_{f}) chiral symmetry. Its low-energy limit is described Gepner:1984au ; Affleck:1985wa by the SU(Nf)1\mathrm{SU}(N_{f})_{1} Wess-Zumino-Witten (WZW) model, which is a Conformal Field Theory (CFT) of central charge Nf1N_{f}-1. The Nf>1N_{f}>1 Schwinger model also contains a massive sector that includes the Schwinger boson. Therefore, it was hoped that the multiflavor Schwinger models may provide simple realizations of the “unparticle physics” idea Georgi:2007ek , and this motivated the papers Georgi:2019tch ; Georgi:2020jik ; Georgi:2022sdu . As in these papers, we will focus on Nf=2N_{f}=2, where for m=0m=0 the IR CFT is described by a compact scalar at the self-dual radius. While investigations of this model have a long history including Coleman:1976uz ; Steinhardt:1977tx ; Smilga:1992hx ; Hetrick:1995wq ; Smilga:1998dh ; Hosotani:1998kd ; Berruto:1999ga ; Hip:2021jgp ; Albergo:2022qfi ; Funcke:2023lli , we will present a number of new results: 1) Even in the limit of small masses, we can have spontaneous symmetry breaking of the charge conjugation symmetry, or critical behavior, or an IR trivial phase. 2) For θ=π\theta=\pi and m/g1m/g\ll 1, there is an effective field theory description in terms of the sine-Gordon model with β8π\beta\approx\sqrt{8\pi} Coleman:1976uz ; Smilga:1998dh . We describe the SU(2)\mathrm{SU}(2)-invariant RG trajectory, which flows from asymptotic freedom in the UV, and in the IR it produces an exponentially small mass gap eAg2/m2\sim e^{-Ag^{2}/m^{2}}, with A0.111A\approx 0.111 as we show below. Therefore, the Nf=2N_{f}=2 Schwinger model with θ=π\theta=\pi has some qualitative similarities with QCD because it can exhibit dimensional transmutation.

We discuss the zero-temperature phase diagram as a function of θ\theta and the masses m1m_{1} and m2m_{2} of the two fermion flavors, which we can restrict to be positive (some aspects of the phase structure were discussed in the past Coleman:1976uz ; Smilga:1998dh ; Georgi:2022sdu ). Our proposal is that, while for all θπ\theta\neq\pi this model has a non-degenerate vacuum, for θ=π\theta=\pi the phase diagram is as in Figure 1. It contains two critical curves that pass through the origin, along which the low-energy physics is governed by the 2D Ising CFT of central charge c=1/2c=1/2. In the shaded region of Figure 1, the charge conjugation symmetry CC, defined below, is spontaneously broken, leading to two degenerate vacua. This phenomenon, which was recently studied in Funcke:2023lli , is reminiscent of the spontaneous breaking of CPCP symmetry in 4D Yang-Mills theory at θ=π\theta=\pi Dashen:1970et ; Creutz:1995wf ; Smilga:1998dh ; Creutz:2010ts ; Creutz:2018vgl ; Gaiotto:2017yup , and there are analogous phenomena in 2D scalar QED Komargodski:2017dmc ; Sulejmanpasic:2020lyq . We present both analytical and numerical evidence for the phase diagram in Figure 1. On the numerical side, our calculations using the Hamiltonian lattice approach are in excellent agreement with the continuum analysis. The convergence of the numerical calculations is significantly improved by including the mass shift (4) derived in Dempsey:2022nys .

Refer to caption
Figure 1: The schematic phase diagram for the 2-flavor Schwinger model at θ=π\theta=\pi (it is similar to the phase diagram of the 2-flavor QCD, exhibited in Creutz:2010ts ; Creutz:2018vgl ). In the shaded region, charge conjugation symmetry is broken, and there are two degenerate vacua. This region is bounded by two critical curves of Ising CFTs. These two curves meet at the origin, where the low-energy description is provided by the SU(2)1\mathrm{SU}(2)_{1} WZW model. For m1=m2gm_{1}=m_{2}\ll g, the model exhibits an exponentially large correlation length due to dimensional transmutation.

II The setup

Let us consider the Schwinger model with NfN_{f} fermion flavors of masses mαm_{\alpha}, with α=1,,Nf\alpha=1,\ldots,N_{f}. While the mαm_{\alpha} are, in general, complex parameters, the U(1)Nf\mathrm{U}(1)^{N_{f}} axial transformations can be used to set all mαm_{\alpha} real with mα0m_{\alpha}\geq 0. Then, the Lagrangian density is

=14g2Fμν2θ4πϵμνFμν+α=1NfΨ¯α(imα)Ψα.\begin{split}{\cal L}&=-\frac{1}{4g^{2}}F_{\mu\nu}^{2}-\frac{\theta}{4\pi}\epsilon^{\mu\nu}F_{\mu\nu}+\sum_{\alpha=1}^{N_{f}}\overline{\Psi}_{\alpha}(i\not{D}-m_{\alpha})\Psi_{\alpha}.\end{split} (1)

Here, ϵ01=1\epsilon^{01}=1, =γμ(μ+iAμ)\not{D}=\gamma^{\mu}(\partial_{\mu}+iA_{\mu}), and (γ0,γ1)=(σ3,iσ2)(\gamma^{0},\gamma^{1})=(\sigma_{3},i\sigma_{2}) obey {γμ,γν}=2ημν=2diag{1,1}\{\gamma^{\mu},\gamma^{\nu}\}=2\eta^{\mu\nu}=2\mathop{\rm diag}\nolimits\{1,-1\}.

To study this model numerically, we use the Hamiltonian lattice formulation of Kogut:1974ag ; Banks:1975gq , where the spatial direction is discretized into NN sites, with NN even, while the time direction remains continuous. The two-component Dirac fermions of each flavor are staggered, with the γ0\gamma^{0} eigenstates of eigenvalue +1+1 and 1-1 being placed on even and odd sites, respectively. The lattice Hamiltonian is

H=g2a2n=0N1(Ln+θ2π)2+α=1Nfmlat,αn=0N1(1)ncn,αcn,αi2an=0N1α=1Nf(cn,αUncn+1,αcn+1,αUncn,α).\begin{split}H&=\frac{g^{2}a}{2}\sum_{n=0}^{N-1}\left(L_{n}+\frac{\theta}{2\pi}\right)^{2}+\sum_{\alpha=1}^{N_{f}}m_{\text{lat},\alpha}\sum_{n=0}^{N-1}(-1)^{n}c^{\dagger}_{n,\alpha}c_{n,\alpha}\\ &{}-\frac{i}{2a}\sum_{n=0}^{N-1}\sum_{\alpha=1}^{N_{f}}\left(c^{\dagger}_{n,\alpha}U_{n}c_{n+1,\alpha}-c^{\dagger}_{n+1,\alpha}U_{n}^{\dagger}c_{n,\alpha}\right)\,.\end{split} (2)

Here, aa is the lattice spacing, cn,αc_{n,\alpha} and cn,αc_{n,\alpha}^{\dagger} are the annihilation and creation operators for a fermion of flavor α\alpha on site nn, and Un=eiϕnU_{n}=e^{i\phi_{n}} is a unitary operator living on the link between sites nn and n+1n+1. The electric field strengths Ln=iϕnL_{n}=-i\frac{\partial}{\partial\phi_{n}} are integer-valued, while θ[0,2π)\theta\in[0,2\pi) comes from the θ\theta-term in the action, and θ2π\frac{\theta}{2\pi} acts as a fractional background electric field. The Hamiltonian should be supplemented by the Gauss law constraint

LnLn1=α(cn,αcn,α1(1)n2).\begin{split}L_{n}-L_{n-1}&=\sum_{\alpha}\left(c_{n,\alpha}^{\dagger}c_{n,\alpha}-\frac{1-(-1)^{n}}{2}\right)\,.\end{split} (3)

The parameters gg and θ\theta of the lattice model should be identified with the analogous parameters of the continuum model (1). As argued in Dempsey:2022nys , one should take

mlat,α=mαNfg2a8.\begin{split}m_{\text{lat},\alpha}=m_{\alpha}-\frac{N_{f}g^{2}a}{8}\,.\end{split} (4)

In Dempsey:2022nys , it was also shown that when NfN_{f} is even and mα=0m_{\alpha}=0, the lattice theory is invariant under translation by one site, which corresponds to a discrete chiral symmetry in the continuum. In the leading strong coupling limit, where the hopping term is ignored, the ground state can be highly degenerate. For Nf=2N_{f}=2 and m1=m2=0m_{1}=m_{2}=0, we find that the strong coupling degeneracy is 3N+13^{N}+1 for θ=0\theta=0, while it is 2N2^{N} for θ=π\theta=\pi. The latter fact provides a starting point for the correspondence between the Nf=2N_{f}=2 Schwinger model at θ=π\theta=\pi and the Heisenberg antiferromagnet Hosotani:1998kd ; Berruto:1999ga .

The integrated fermion bilinear operator 𝑑xΨ¯Ψ\int dx\,\bar{\Psi}\Psi translates into the lattice operator n=0N1(1)ncn,αcn,α\sum_{n=0}^{N-1}(-1)^{n}c^{\dagger}_{n,\alpha}c_{n,\alpha}, which is odd under the unit shift. The uniqueness of the ground state away from the strong coupling limit for mα=0m_{\alpha}=0, and the symmetry under the unit translation, imply that the VEV of the mass operator vanishes on a periodic lattice with an even number of sites.

When θ=0\theta=0 or π\pi, for any mαm_{\alpha}, the models (1) and (2) are invariant under a charge conjugation symmetry CC. In the continuum, CC acts as

C:AμAμ,Ψαγ5Ψα,\begin{split}\text{$C$}:\qquad A_{\mu}\to-A_{\mu}\,,\qquad\Psi_{\alpha}\to\gamma^{5}\Psi_{\alpha}^{*}\,,\end{split} (5)

where γ5=γ0γ1=σ1\gamma^{5}=\gamma^{0}\gamma^{1}=\sigma_{1}, and on the lattice it acts as Berruto:1999ga

C:LnLn+1θπ,UnUn+1,cn,αcn+1,α,cn,αcn+1,α.\begin{split}\text{$C$}:\qquad L_{n}&\to-L_{n+1}-\frac{\theta}{\pi}\,,\qquad U_{n}\to U_{n+1}^{\dagger}\,,\\ c_{n,\alpha}&\to c_{n+1,\alpha}^{\dagger}\,,\qquad c_{n,\alpha}^{\dagger}\to c_{n+1,\alpha}\,.\end{split} (6)

It is this symmetry CC that is broken whenever there are two degenerate vacua in the phase diagram in Figure 1.

For any θ\theta and mα=mm_{\alpha}=m for all α\alpha, the models (1) and (2) are invariant under an SU(Nf)\mathrm{SU}(N_{f}) symmetry under which the fermions transform in the fundamental representation. When m=0m=0, in the continuum model (1), this SU(Nf)\mathrm{SU}(N_{f}) symmetry is enhanced to SU(Nf)×SU(Nf)\mathrm{SU}(N_{f})\times\mathrm{SU}(N_{f}).

III Continuum Treatment of the Two-flavor Schwinger model

Let us now take Nf=2N_{f}=2 and present continuum field theory arguments in support of the phase diagram of the 2-flavor Schwinger model in Figure 1.

III.1 One very massive fermion

When one of the fermions is very massive, it can be integrated out, leaving us with the Nf=1N_{f}=1 model. For a fermion mass mm, the phase diagram of the Nf=1N_{f}=1 model exhibits a line of first order phase transitions at θ=π\theta=\pi that extends over the interval (mcr,)(m_{\text{cr}},\infty) Coleman:1976uz , with mcr0.33gm_{\text{cr}}\approx 0.33g. At m=mcrm=m_{\text{cr}}, there is evidence Byrnes:2002gj ; Byrnes:2002nv that the second order phase transition is in the 2D Ising universality class. For m>mcrm>m_{\text{cr}} and θ=π\theta=\pi, there are two degenerate vacua, each of which breaks CC spontaneously. Everywhere else on the phase diagram there is a non-degenerate vacuum and a non-zero gap.

Without loss of generality, suppose we take m2/g1m_{2}/g\gg 1. Integrating out Ψ2\Psi_{2} in (1) yields the effective Lagrangian

Nf=1+12d2yAμ(x)Aν(y)Π2μν(xy)+O(A4),\begin{split}{\cal L}^{N_{f}=1}+\frac{1}{2}\int d^{2}y\,A_{\mu}(x)A_{\nu}(y)\Pi_{2}^{\mu\nu}(x-y)+O(A^{4})\,,\end{split} (7)

where Π2μν=(2ημν+μν)Π2(x)\Pi_{2}^{\mu\nu}=(-\partial^{2}\eta^{\mu\nu}+\partial^{\mu}\partial^{\nu})\Pi_{2}(x) is the one-loop vacuum polarization. The Fourier transform Π2(q)=d2xΠ2(x)eiqx\Pi_{2}(q)=\int d^{2}x\,\Pi_{2}(x)e^{iq\cdot x} is (see (7.90) of Peskin:1995ev ) 111The result in (7.90) of Peskin:1995ev should be divided by a factor of 2e22e^{2}. The division by e2e^{2} is due to our normalization of the gauge kinetic term, and the division by 22 is due to the fact that for two-component spinors tr𝟏=2\mathop{\rm tr}\nolimits{\bf 1}=2 as opposed to tr𝟏=4\mathop{\rm tr}\nolimits{\bf 1}=4 as was assumed in Peskin:1995ev .

Π2(q)=1π01𝑑ξξ(1ξ)m22ξ(1ξ)q216πm22\begin{split}\Pi_{2}(q)=-\frac{1}{\pi}\int_{0}^{1}d\xi\,\frac{\xi(1-\xi)}{m_{2}^{2}-\xi(1-\xi)q^{2}}\approx-\frac{1}{6\pi m_{2}^{2}}\end{split} (8)

at large m2m_{2}. Thus, Π2(x)16πm22δ(2)(x)\Pi_{2}(x)\approx-\frac{1}{6\pi m_{2}^{2}}\delta^{(2)}(x), and the effective Lagrangian (7) becomes, approximately, that of the one-flavor model with an effective gauge coupling:

geff2=g2(1+16πg2m22).\begin{split}g^{-2}_{\text{eff}}=g^{-2}\left(1+\frac{1}{6\pi}\frac{g^{2}}{m_{2}^{2}}\right)\,.\end{split} (9)

Since the one-flavor Schwinger model exhibits an Ising second-order phase transition at mcr0.33gm_{\text{cr}}\approx 0.33g at θ=π\theta=\pi, it follows that the 2-flavor Schwinger model with m2/g1m_{2}/g\gg 1 also exhibits an Ising phase transition at θ=π\theta=\pi for mcr0.33geffm_{\text{cr}}\approx 0.33g_{\text{eff}}. Expanding this we get

m1,cr(m2)0.33g(1112πg2m22+O(g4m24)).\begin{split}m_{1,\text{cr}}(m_{2})\approx 0.33g\left(1-\frac{1}{12\pi}\frac{g^{2}}{m_{2}^{2}}+O\left(\frac{g^{4}}{m_{2}^{4}}\right)\right)\,.\end{split} (10)

The phase diagram should of course be invariant under interchanging m1m2m_{1}\leftrightarrow m_{2} so, at θ=π\theta=\pi, there should also be an Ising transition at m2,cr(m1)m_{2,\text{cr}}(m_{1}) given by the RHS of (10) with m2m1m_{2}\to m_{1}. The expression (10) and the one obtained after interchanging m1m2m_{1}\leftrightarrow m_{2} represent the asymptotic behaviors of the blue curves in Figure 1. The large mass analysis also shows that in the wedge between the two curves we expect two degenerate ground states, while outside of this wedge we expect a non-degenerate ground state, just as in the Nf=1N_{f}=1 model at θ=π\theta=\pi.

This argument also shows that when θπ\theta\neq\pi and one of the fermions is very massive, the ground state is non-degenerate because this is also the case in the one-flavor model. In fact, for θ0,π\theta\neq 0,\pi we must have a non-degenerate ground state because there is no charge conjugation symmetry that can be spontaneously broken.

III.2 Small mass regime

Near m1=m2=0m_{1}=m_{2}=0, a useful equivalent description is obtained using Abelian bosonization Coleman:1976uz . (One can also use non-Abelian bosonization, as in Gepner:1984au .) Following Coleman:1976uz , we bosonize the fermions Ψ1,2\Psi_{1,2} to scalar fields ϕ1,2\phi_{1,2}, and reparameterize them via ϕ+=21/2(ϕ1+ϕ2+12π1/2θ)\phi_{+}=2^{-1/2}(\phi_{1}+\phi_{2}+{\textstyle\frac{1}{2}}\pi^{-1/2}\theta) and ϕ=21/2(ϕ1ϕ2)\phi_{-}=2^{-1/2}(\phi_{1}-\phi_{2}).

Let us restrict our attention to m1=m2=mm_{1}=m_{2}=m. The bosonized Lagrangian is

bos=14g2Fμν2ϕ+2πϵμνFμν+12(μϕ+)2+12(μϕ)2+eγπmμ+μNμ+cos[2πϕ+θ2]Nμcos[2πϕ],\begin{split}{\cal L}_{\text{bos}}&=-\frac{1}{4g^{2}}F_{\mu\nu}^{2}-\frac{\phi_{+}}{\sqrt{2\pi}}\epsilon^{\mu\nu}F_{\mu\nu}+\frac{1}{2}(\partial_{\mu}\phi_{+})^{2}+\frac{1}{2}(\partial_{\mu}\phi_{-})^{2}\\ &{}+\frac{e^{\gamma}}{\pi}m\sqrt{\mu_{+}\mu_{-}}N_{\mu_{+}}\cos\left[\sqrt{2\pi}\phi_{+}-\frac{\theta}{2}\right]N_{\mu_{-}}\cos\left[\sqrt{2\pi}\phi_{-}\right]\,,\end{split} (11)

where NN_{\cal M} means that the expression that follows is normal ordered by subtracting the two-point functions of a scalar field of mass {\cal M}. A convenient choice is μ+=μ\mu_{+}=\mu, where μ\mu is defined below, and μ/g0\mu_{-}/g\to 0.

For m=0m=0, integrating out the gauge field shows that ϕ+\phi_{+} has mass μ=2πg\mu=\sqrt{\frac{2}{\pi}}g, while ϕ\phi_{-} remains massless. The field ϕ\phi_{-} obeys the identification ϕϕ+2π\phi_{-}\sim\phi_{-}+\sqrt{2\pi}, which corresponds to the self-dual radius of the compact scalar. Thus, for m=0m=0 we have a massive sector described by ϕ+\phi_{+}, and a sector consisting of the c=1c=1 self-dual scalar CFT, which has SU(2)×SU(2)\mathrm{SU}(2)\times\mathrm{SU}(2) symmetry. At low energies, the massive sector can also be integrated out, and we are left with the self-dual scalar CFT.

After integrating out the gauge field, we can integrate out ϕ+\phi_{+} order by order in mm:

bos=12(μϕ)2+mμμeγπ𝒪+(x)𝒪(x)+ie2γ2π2m2μμd2y𝒪+(x)𝒪+(y)𝒪(x)𝒪(y)+O(m3),\begin{split}{\cal L}_{\text{bos}}&=\frac{1}{2}(\partial_{\mu}\phi_{-})^{2}+m\sqrt{\mu\mu_{-}}\frac{e^{\gamma}}{\pi}\left\langle{\cal O}_{+}(x)\right\rangle{\cal O}_{-}(x)\\ &{}+i\frac{e^{2\gamma}}{2\pi^{2}}m^{2}\mu\mu_{-}\int d^{2}y\,\left\langle{\cal O}_{+}(x){\cal O}_{+}(y)\right\rangle{\cal O}_{-}(x){\cal O}_{-}(y)\\ &{}+O(m^{3})\,,\end{split} (12)

with 𝒪+Nμcos[2πϕ+θ2]{\cal O}_{+}\equiv N_{\mu}\cos\left[\sqrt{2\pi}\phi_{+}-\frac{\theta}{2}\right], 𝒪Nμcos[2πϕ]{\cal O}_{-}\equiv N_{\mu_{-}}\cos\left[\sqrt{2\pi}\phi_{-}\right], and the expectation values taken in the theory of a free massive scalar field ϕ+\phi_{+} of mass μ\mu. For θπ\theta\neq\pi, 𝒪+=cosθ2\left\langle{\cal O}_{+}\right\rangle=\cos\frac{\theta}{2} gives the effective theory

bos=12(μϕ)2+mμμeγπcosθ2Nμcos[2πϕ]+O(m2).\begin{split}{\cal L}_{\text{bos}}&=\frac{1}{2}(\partial_{\mu}\phi_{-})^{2}+m\sqrt{\mu\mu_{-}}\frac{e^{\gamma}}{\pi}\cos\frac{\theta}{2}N_{\mu_{-}}\cos\left[\sqrt{2\pi}\phi_{-}\right]\\ &{}+O(m^{2})\,.\end{split} (13)

This is the self-dual scalar CFT deformed by the operator cos[2πϕ]\cos\left[\sqrt{2\pi}\phi_{-}\right] of dimension 1/21/2, which triggers a RG flow to a gapped phase. One can then show using RG scaling arguments or re-normal ordering Coleman:1976uz that the mass gap is of order |mcos(θ/2)|2/3g1/3\sim|m\cos(\theta/2)|^{2/3}g^{1/3}.

When θ=π\theta=\pi, the coefficient of the relevant operator in (13) vanishes. Nevertheless, the mass deformation is not exactly marginal, because the only marginal deformation of the self-dual compact scalar CFT is the change in radius of the scalar, which breaks the symmetry to U(1)×U(1)\mathrm{U}(1)\times\mathrm{U}(1). This would be in contradiction with the SU(2)\mathrm{SU}(2) symmetry of the equal-mass Schwinger model. We can evaluate the O(m2)O(m^{2}) term in (12) using the propagator GM(x)=ϕ(x)ϕ(0)=12πK0(Mx2)G_{M}(x)=\langle\phi(x)\phi(0)\rangle=\frac{1}{2\pi}K_{0}(M\sqrt{-x^{2}}) of a free scalar field ϕ\phi of mass MM, which implies that 𝒪+(x)𝒪+(y)=sinh[2πGμ(xy)]\langle{\cal O}_{+}(x){\cal O}_{+}(y)\rangle=\sinh\left[2\pi G_{\mu}(x-y)\right]. We also have

𝒪(x)𝒪(y)=e2πGμ(xy)2Nμcos[2π(ϕ(x)+ϕ(y))]+e2πGμ(xy)2Nμcos[2π(ϕ(x)ϕ(y))].\begin{split}&{\cal O}_{-}(x){\cal O}_{-}(y)\\ &{}=\frac{e^{-2\pi G_{\mu_{-}}(x-y)}}{2}N_{\mu_{-}}\cos\left[\sqrt{2\pi}(\phi_{-}(x)+\phi_{-}(y))\right]\\ &{}+\frac{e^{2\pi G_{\mu_{-}}(x-y)}}{2}N_{\mu_{-}}\cos\left[\sqrt{2\pi}(\phi_{-}(x)-\phi_{-}(y))\right]\,.\end{split} (14)

Plugging these results into (12), changing variables to z=μ(yx)z=\mu(y-x) and passing to Euclidean signature, we see that the integral receives contributions only from small |z|\left\lvert z\right\rvert. Expanding in |z|\left\lvert z\right\rvert and evaluating the integral gives

bos=12(μϕ)2+e3γIsm28π2μ2(2πe2γ(μϕ)2+μ2Nμcos(8πϕ))+O(m4),\begin{split}\mathcal{L}_{\text{bos}}&=\frac{1}{2}(\partial_{\mu}\phi_{-})^{2}+\frac{e^{3\gamma}I_{s}m^{2}}{8\pi^{2}\mu^{2}}\biggl{(}2\pi e^{-2\gamma}(\partial_{\mu}\phi_{-})^{2}\\ &{}+\mu_{-}^{2}N_{\mu_{-}}\cos(\sqrt{8\pi}\phi_{-})\biggr{)}+O(m^{4})\,,\end{split} (15)

where Is=2π0𝑑ξξ2sinhK0(ξ)10.08I_{s}=2\pi\int_{0}^{\infty}d\xi\,\xi^{2}\sinh K_{0}(\xi)\approx 10.08 (see also Smilga:1998dh ). The Lagrangian (15) is that of the sine-Gordon model, a two-dimensional boson with interaction term cos(βϕ)\sim\cos(\beta\phi) with β>0\beta>0. By rescaling the boson to have canonical normalization, we have β2<8π\beta^{2}<8\pi. For mgm\ll g, β28π\beta^{2}\rightarrow 8\pi, and the scaling dimension of the cosine operator approaches 22. In this limit, the model is closely related to the continuum description of the Heisenberg antiferromagnet Cheng:2022sgb .

The RG flow of the sine-Gordon model near β2=8π\beta^{2}=8\pi was computed in Kosterlitz:1974sm ; Amit:1979ab and shown to describe the BKT transition. Generically, both the coefficient of the cosine and radius of the scalar will flow. Up to first order in the bare parameters α\alpha and δ=β28π1\delta=\frac{\beta^{2}}{8\pi}-1, the sine-Gordon model is defined by Amit:1979ab

=1δ2(μϕ)2+αe2γ32πμ2Nμcos(8πϕ).\begin{split}{\cal L}=\frac{1-\delta}{2}(\partial_{\mu}\phi)^{2}+\frac{\alpha e^{2\gamma}}{32\pi}\mu_{-}^{2}N_{\mu_{-}}\cos(\sqrt{8\pi}\phi)\,.\end{split} (16)

The one-loop beta functions for the running couplings α¯\overline{\alpha} and δ¯\overline{\delta} are Kosterlitz:1974sm ; Amit:1979ab

βα¯=2α¯δ¯,βδ¯=132α¯2.\begin{split}\beta_{\overline{\alpha}}=2\overline{\alpha}\overline{\delta}\,,\qquad\beta_{\overline{\delta}}=\frac{1}{32}\overline{\alpha}^{2}\,.\end{split} (17)

The effective theory (15) may be restricted to have the SU(2)\mathrm{SU}(2) symmetry that arises from the SU(2)\mathrm{SU}(2) symmetry of the Schwinger model with equal fermion masses. Then, in the two-dimensional parameter space (α¯,δ¯)(\overline{\alpha},\overline{\delta}), only the SU(2)\mathrm{SU}(2)-invariant RG trajectory can be accessed. This trajectory is the line α¯=8δ¯\overline{\alpha}=-8\overline{\delta} that passes through the origin, as can be seen from the fact that (15) and (16) imply

α=8eγIs4m2g2=8δ,\begin{split}\alpha=\frac{8e^{\gamma}I_{s}}{4}\frac{m^{2}}{g^{2}}=-8\delta\,,\end{split} (18)

or from analyzing the SU(2)\mathrm{SU}(2)-invariant operators in the model (16). On this locus with SU(2)\mathrm{SU}(2) symmetry, the sine-Gordon model (16) is related via bosonization to the SU(2)\mathrm{SU}(2) Thirring model Amit:1979ab ; Banks:1975xs , which contains two massless Dirac fermions ψa\psi^{a}. Their interaction is i=13JiJi\sim\sum_{i=1}^{3}J^{i}J^{i} where the SU(2)\mathrm{SU}(2) currents are Ji=12ψ¯aσabiψbJ^{i}=\frac{1}{2}\bar{\psi}^{a}\sigma^{i}_{ab}\psi^{b}.

The β\beta-function for the running mass parameter m¯\overline{m} can be inferred from (18) and (17):

βm¯=Mdm¯dM=eγIs4g2m¯3,\displaystyle\beta_{\overline{m}}=M\frac{d\overline{m}}{dM}=-\frac{e^{\gamma}I_{s}}{4g^{2}}\overline{m}^{3}\,, (19)

where MM is the RG scale. Thus, the interaction strength in the effective sine-Gordon model, and equivalently in the SU(2)\mathrm{SU}(2) Thirring model, is asymptotically free. The interaction strength formally diverges far in the IR, at the scale comparable to the mass gap (this scale is analogous to ΛQCD\Lambda_{\rm QCD}):

EgapeAg2m2,A=2eγIs0.111.\displaystyle E_{\rm gap}\sim e^{-A\frac{g^{2}}{m^{2}}}\,,\qquad A=\frac{2e^{-\gamma}}{I_{s}}\approx 0.111\,. (20)

This exponentially small mass gap implies that, for small mm, the correlation length diverges as ξ1EgapeAg2m2\xi\sim\frac{1}{E_{\text{gap}}}\sim e^{A\frac{g^{2}}{m^{2}}}. Similarly, at small mm all observables can be expressed in terms of the energy scale EgapE_{\text{gap}}. For instance, since Ψ¯αΨα\overline{\Psi}_{\alpha}\Psi_{\alpha} flows to an operator of dimension Δ+=2\Delta_{+}=2 in the c=1c=1 theory at m=0m=0, we must have Ψ¯αΨαEgapΔ+e2Ag2m2\langle\overline{\Psi}_{\alpha}\Psi_{\alpha}\rangle\sim E_{\text{gap}}^{\Delta_{+}}\sim e^{-2A\frac{g^{2}}{m^{2}}}. Likewise, the operator Ψ1Ψ1Ψ2Ψ2\Psi_{1}\Psi_{1}-\Psi_{2}\Psi_{2} that takes us away from the m1=m2m_{1}=m_{2} line in Figure (1) flows to an operator of dimension Δ=1/2\Delta_{-}=1/2. This allows us to estimate that the width of the symmetry breaking region is ΔmEgap2Δ=e3A2g2m2\Delta m\sim E_{\text{gap}}^{2-\Delta_{-}}=e^{-\frac{3A}{2}\frac{g^{2}}{m^{2}}}.

Refer to caption
Figure 2: The heat map depicts the entanglement entropy S1/2(N=216,a=0.3)S_{1/2}(N=216,a=0.3) with open boundary conditions as a function of the fermion masses at θ=π\theta=\pi. The black points are estimates of the location of the c=12c=\frac{1}{2} critical curve in the continuum limit a0a\to 0. The asymptotic shape of the curves agrees with (10). For mgm\ll g, the two c=12c=\frac{1}{2} critical curves become exponentially close to each other.

That charge conjugation symmetry is spontaneously broken for any m>0m>0 can can be seen from (15). Indeed, in the bosonized description at θ=π\theta=\pi, the charge conjugation symmetry CC acts as

AμAμ,ϕ+ϕ+,ϕϕ+π2.\begin{split}A_{\mu}\to-A_{\mu}\,,\quad\phi_{+}\to-\phi_{+}\,,\quad\phi_{-}\to-\phi_{-}+\sqrt{\frac{\pi}{2}}\,.\end{split} (21)

This is clearly a symmetry of the Lagrangian (11) and also of the effective Lagrangian (15). However, over the range of one period ϕ[0,2π]\phi_{-}\in[0,\sqrt{2\pi}], the potential cos(8πϕ)\sim-\cos(\sqrt{8\pi}\phi_{-}) in (15) has two minima, one at ϕ=0\phi_{-}=0 and one ϕ=π2\phi_{-}=\sqrt{\frac{\pi}{2}}. These minima are exchanged by the symmetry CC in (21). Semi-classically, we thus have two vacua in which CC is broken spontaneously.

The spontaneous breaking of CC in the two-flavor Schwinger model provides a nice analogy to the breaking of CPCP and presence of two degenerate vacua in 4D QCD with θ=π\theta=\pi and two light flavors Creutz:1995wf . The height of the barrier separating the two symmetry breaking vacua is of order m2m^{2}, just as in QCD Creutz:1995wf ; Smilga:1998dh . The CPCP violation Dashen:1970et can be seen using the chiral Lagrangian for QCD, and the zero-temperature phase diagram as a function of light quark masses mum_{u} and mdm_{d} has a similar structure Creutz:2010ts ; Creutz:2018vgl to our Figure 1. The boundaries of the region where CPCP is spontaneosuly broken can be found from the condition that the mass of the neutral pion vanishes there. The width of the symmetry broken region is found to behave as (mu+md)2/fπ(m_{u}+m_{d})^{2}/f_{\pi}, which is parametrically much bigger than the exponentially small width that we find in the Schwinger model.

IV Numerical results

We study the Nf=2N_{f}=2 Schwinger model numerically using the lattice Hamiltonian (2) (see also Funcke:2023lli ).

Refer to caption
Figure 3: The estimated correlation length (up to a constant factor β\beta) for m1=m2mm_{1}=m_{2}\equiv m as a function of m2m^{-2}, showing a scaling of the form ξeAg2m2\xi\sim e^{A\frac{g^{2}}{m^{2}}}, as inferred from (20). The dashed lines have slope A0.111A\approx 0.111. Extrapolating the slope from the lattice calculation to a0a\to 0 gives 0.11(1), in agreement with the theoretical value.

While the one-flavor model can be studied efficiently via exact diagonalization Hamer:1982mx ; Dempsey:2022nys , with two flavors the number of states grows so quickly with the number NN of lattice sites that this becomes impractical. Instead, we employ tensor network methods, using a matrix product state (MPS) ansatz to approximate the ground state Banuls:2013jaa ; Funcke:2023lli . To optimize the MPS ansatz, we use ITensors.jl ITensor-r0.3 ; ITensor . We use open boundary conditions, since this allows us to study the behavior of much larger lattices.

The MPS form of the ground state makes it especially simple to calculate the entanglement entropy for a left-right bipartition of the open chain. Let Sx(N,a)S_{x}(N,a) denote the entanglement entropy for a subsystem of the leftmost xNxN sites in a chain of NN sites with lattice spacing aa. Then, at a critical point with central charge cc, the entropy is expected to grow like Calabrese:2009qy

Sx(N,a)=c6log(2Nπsinπx)+const.\begin{split}S_{x}(N,a)=\frac{c}{6}\log\left(\frac{2N}{\pi}\sin\pi x\right)+\text{const}.\end{split} (22)

At any other point, this logarithmic growth of the entropy will plateau when NξaN\sim\frac{\xi}{a}, where ξ\xi is the correlation length. By combining this result with a finite-size scaling analysis, one can derive very precise estimates for the locations of critical points in the continuum theory from values of the entanglement entropy on a finite lattice Campostrini:2014lta .

In Figure 2, we show the behavior of the entanglement entropy for a fixed lattice, along with the precise estimate of the critical curve obtained via the intersection method outlined in Campostrini:2014lta . The finite-size scaling analysis confirms that this curve has c=12c=\frac{1}{2}. By fitting the leading large-mass behavior of this curve, we find

m2,cr(m1)=0.335(4)0.0097(17)(gm1)2,\begin{split}m_{2,\text{cr}}(m_{1})=0.335(4)-0.0097(17)\left(\frac{g}{m_{1}}\right)^{2},\end{split} (23)

and the coefficient of g2/m12g^{2}/m_{1}^{2} is in good agreement with the value 0.333512π0.0088\frac{0.3335}{12\pi}\approx 0.0088 predicted from (10).

We can also use lattice calculations of the entanglement entropy to estimate the growth of the correlation length for small m1=m2=mm_{1}=m_{2}=m. For a fixed lattice, we can compare the dependence of the entanglement entropy on the subsystem size with (22) to obtain an estimate cestc_{\text{est}} for the central charge. Anywhere away from a critical point, this estimate will tend to zero around NξaN\sim\frac{\xi}{a}. We can thus take a fiducial cutoff for cestc_{\text{est}}, and define βξa\frac{\beta\xi}{a} as the lattice size when cestc_{\text{est}} crosses below this cutoff, where β\beta is an unknown constant.

Figure 3 shows this estimate of the logarithm of the correlation length along the SU(2)\mathrm{SU}(2)-invariant line as a function of m2m^{-2}. The linear behavior suggests a scaling of the form ξeAg2m2\xi\sim e^{A\frac{g^{2}}{m^{2}}} at small m/gm/g, as explained after (20). Furthermore, extrapolating the slope as m0m\to 0 to the continuum limit a0a\to 0 gives A=0.11(1)A=0.11(1), in good agreement with the theoretical value in (20).

V Discussion

In this paper, we presented analytical and numerical evidence for the phase diagram of the Nf=2N_{f}=2 Schwinger model at θ=π\theta=\pi shown in Figure 1. The behavior we find is quite different from that at θπ\theta\neq\pi: along the SU(2)\mathrm{SU}(2)-invariant line the theory contains a nearly marginal operator which leads to logarithmic RG flow of BKT type. As a result, for mgm\ll g the mass gap is exponentially small, eAg2/m2\sim e^{-Ag^{2}/m^{2}}. Along this SU(2)\mathrm{SU}(2)-symmetric line, Georgi Georgi:2022sdu calculated the anomalous dimensions of operators perturbatively in powers of (m/g)2(m/g)^{2}. The fact that the mass gap is exponentially small makes the theory for mgm\ll g “nearly conformal” in a large range of energies, so that perturbative anomalous dimension calculations should be parametrically reliable. We thus hope that the calculations of Georgi:2022sdu can be checked numerically using the lattice Hamiltonian setup, but we leave this question for future work.

We find that the 2\mathbb{Z}_{2} charge conjugation symmetry is spontaneously broken in the entire shaded region of the phase diagram in Figure 1. This region becomes exponentially narrow near m=0m=0 and is bounded by 2D Ising CFTs. It is interesting to ask how the addition to the action of 4-fermion operators may change this phase diagram. We also leave this question for future work.

Acknowledgments

We are grateful to Michael Creutz, Howard Georgi, Etsuko Itou, Andrei Katsevich, Zohar Komargodski, Lenny Susskind, Yuya Tanizaki, Grisha Tarnopolsky, and Edward Witten for very useful discussions. We thank a referee for important comments and pointing out very useful references. This work was supported in part by the US National Science Foundation under Grants No. PHY-2111977 and PHY-2209997, and by the Simons Foundation Grants No. 488653 and 917464. RD was also supported in part by an NSF Graduate Research Fellowship.

References

  • (1) J. S. Schwinger, “Gauge Invariance and Mass. 2.,” Phys. Rev. 128 (1962) 2425–2429.
  • (2) J. H. Lowenstein and J. A. Swieca, “Quantum electrodynamics in two-dimensions,” Annals Phys. 68 (1971) 172–195.
  • (3) A. Casher, J. B. Kogut, and L. Susskind, “Vacuum polarization and the absence of free quarks,” Phys. Rev. D 10 (1974) 732–745.
  • (4) S. R. Coleman, R. Jackiw, and L. Susskind, “Charge Shielding and Quark Confinement in the Massive Schwinger Model,” Annals Phys. 93 (1975) 267.
  • (5) J. B. Kogut and L. Susskind, “Hamiltonian Formulation of Wilson’s Lattice Gauge Theories,” Phys. Rev. D 11 (1975) 395–408.
  • (6) T. Banks, L. Susskind, and J. B. Kogut, “Strong Coupling Calculations of Lattice Gauge Theories: (1+1)-Dimensional Exercises,” Phys. Rev. D 13 (1976) 1043.
  • (7) M. C. Bañuls et. al., “Simulating Lattice Gauge Theories within Quantum Technologies,” Eur. Phys. J. D 74 (2020), no. 8 165, 1911.00003.
  • (8) S. R. Coleman, “More About the Massive Schwinger Model,” Annals Phys. 101 (1976) 239.
  • (9) D. Gepner, “Nonabelian Bosonization and Multiflavor QED and QCD in Two-dimensions,” Nucl. Phys. B 252 (1985) 481–507.
  • (10) I. Affleck, “On the Realization of Chiral Symmetry in (1+1)-dimensions,” Nucl. Phys. B 265 (1986) 448–468.
  • (11) H. Georgi, “Unparticle physics,” Phys. Rev. Lett. 98 (2007) 221601, hep-ph/0703260.
  • (12) H. Georgi and B. Noether, “Non-perturbative Effects and Unparticle Physics in Generalized Schwinger Models,” 1908.03279.
  • (13) H. Georgi, “Automatic Fine-Tuning in the Two-Flavor Schwinger Model,” Phys. Rev. Lett. 125 (2020), no. 18 181601, 2007.15965.
  • (14) H. Georgi, “Mass perturbation theory in the 2-flavor Schwinger model with opposite masses with a review of the background,” JHEP 10 (2022) 119, 2206.14691.
  • (15) P. J. Steinhardt, “SU(2) Flavor Schwinger Model on the Lattice,” Phys. Rev. D 16 (1977) 1782.
  • (16) A. V. Smilga, “On the fermion condensate in Schwinger model,” Phys. Lett. B 278 (1992) 371–376.
  • (17) J. E. Hetrick, Y. Hosotani, and S. Iso, “The Massive multi - flavor Schwinger model,” Phys. Lett. B 350 (1995) 92–102, hep-th/9502113.
  • (18) A. V. Smilga, “QCD at theta similar to pi,” Phys. Rev. D 59 (1999) 114021, hep-ph/9805214.
  • (19) Y. Hosotani, “Antiferromagnetic S = 1/2 Heisenberg chain and the two flavor massless Schwinger model,” Phys. Rev. B 60 (1999) 6198–6199, hep-th/9809066.
  • (20) F. Berruto, G. Grignani, G. W. Semenoff, and P. Sodano, “On the correspondence between the strongly coupled two flavor lattice Schwinger model and the Heisenberg antiferromagnetic chain,” Annals Phys. 275 (1999) 254–296, hep-th/9901142.
  • (21) I. Hip, J. F. N. Castellanos, and W. Bietenholz, “Finite temperature and δ\delta-regime in the 2-flavor Schwinger model,” PoS LATTICE2021 (2022) 279, 2109.13468.
  • (22) M. S. Albergo, D. Boyda, K. Cranmer, D. C. Hackett, G. Kanwar, S. Racanière, D. J. Rezende, F. Romero-López, P. E. Shanahan, and J. M. Urban, “Flow-based sampling in the lattice Schwinger model at criticality,” 2202.11712.
  • (23) L. Funcke, K. Jansen, and S. Kühn, “Exploring the CP-Violating Dashen Phase in the Schwinger Model with Tensor Networks,” 2303.03799.
  • (24) R. F. Dashen, “Some features of chiral symmetry breaking,” Phys. Rev. D 3 (1971) 1879–1889.
  • (25) M. Creutz, “Quark masses and chiral symmetry,” Phys. Rev. D 52 (1995) 2951–2959, hep-th/9505112.
  • (26) M. Creutz, “Quark mass dependence of two-flavor QCD,” Phys. Rev. D 83 (2011) 016005, 1010.4467.
  • (27) M. Creutz, “CP violation in QCD,” PoS Confinement2018 (2018) 171, 1810.03543.
  • (28) D. Gaiotto, A. Kapustin, Z. Komargodski, and N. Seiberg, “Theta, Time Reversal, and Temperature,” JHEP 05 (2017) 091, 1703.00501.
  • (29) Z. Komargodski, A. Sharon, R. Thorngren, and X. Zhou, “Comments on Abelian Higgs Models and Persistent Order,” SciPost Phys. 6 (2019), no. 1 003, 1705.04786.
  • (30) T. Sulejmanpasic, D. Göschl, and C. Gattringer, “First-Principles Simulations of 1+1D Quantum Field Theories at θ=π\theta=\pi and Spin Chains,” Phys. Rev. Lett. 125 (2020), no. 20 201602, 2007.06323.
  • (31) R. Dempsey, I. R. Klebanov, S. S. Pufu, and B. Zan, “Discrete chiral symmetry and mass shift in the lattice Hamiltonian approach to the Schwinger model,” Phys. Rev. Res. 4 (2022), no. 4 043133, 2206.05308.
  • (32) T. Byrnes, P. Sriganesh, R. J. Bursill, and C. J. Hamer, “Density matrix renormalization group approach to the massive Schwinger model,” Nucl. Phys. B Proc. Suppl. 109 (2002) 202–206, hep-lat/0201007.
  • (33) T. Byrnes, P. Sriganesh, R. J. Bursill, and C. J. Hamer, “Density matrix renormalization group approach to the massive Schwinger model,” Phys. Rev. D 66 (2002) 013002, hep-lat/0202014.
  • (34) M. E. Peskin and D. V. Schroeder, An Introduction to quantum field theory. Addison-Wesley, Reading, USA, 1995.
  • (35) M. Cheng and N. Seiberg, “Lieb-Schultz-Mattis, Luttinger, and ’t Hooft – anomaly matching in lattice systems,” 2211.12543.
  • (36) J. M. Kosterlitz, “The Critical properties of the two-dimensional x y model,” J. Phys. C 7 (1974) 1046–1060.
  • (37) D. J. Amit, Y. Y. Goldschmidt, and G. Grinstein, “Renormalization Group Analysis of the Phase Transition in the 2D Coulomb Gas, Sine-Gordon Theory and xy Model,” J. Phys. A 13 (1980) 585.
  • (38) T. Banks, D. Horn, and H. Neuberger, “Bosonization of the SU(N) Thirring Models,” Nucl. Phys. B 108 (1976) 119.
  • (39) C. J. Hamer, J. B. Kogut, D. P. Crewther, and M. M. Mazzolini, “The Massive Schwinger Model on a Lattice: Background Field, Chiral Symmetry and the String Tension,” Nucl. Phys. B 208 (1982) 413–438.
  • (40) M. C. Bañuls, K. Cichy, K. Jansen, and J. I. Cirac, “The mass spectrum of the Schwinger model with Matrix Product States,” JHEP 11 (2013) 158, 1305.3765.
  • (41) M. Fishman, S. R. White, and E. M. Stoudenmire, “Codebase release 0.3 for ITensor,” SciPost Phys. Codebases (2022) 4–r0.3. Publisher: SciPost.
  • (42) M. Fishman, S. R. White, and E. M. Stoudenmire, “The ITensor software library for tensor network calculations,” SciPost Phys. Codebases (2022) 4. Publisher: SciPost.
  • (43) P. Calabrese and J. Cardy, “Entanglement entropy and conformal field theory,” J. Phys. A 42 (2009) 504005, 0905.4013.
  • (44) M. Campostrini, A. Pelissetto, and E. Vicari, “Finite-size scaling at quantum transitions,” Phys. Rev. B 89 (2014), no. 9 094516, 1401.0788.