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

Can negative bare couplings make sense? The ϕ4\vec{\phi}^{4} theory at large NN

Ryan D. Weller [email protected] Department of Physics, University of Colorado, Boulder, CO 80309, USA
Abstract

Scalar λϕ4\lambda\phi^{4} theory in 3+1D, for a positive coupling constant λ>0\lambda>0, is known to have no interacting continuum limit, which is referred to as quantum triviality. However, it has been recently argued that the theory in 3+1D with an NN-component scalar ϕ\vec{\phi} and a (ϕϕ)2=ϕ4(\vec{\phi}\cdot\vec{\phi})^{2}=\vec{\phi}^{4} interaction term does have an interacting continuum limit at large NN. It has been suggested that this continuum limit has a negative (bare) coupling constant and exhibits asymptotic freedom, similar to the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric gϕ4-g\phi^{4} field theory. In this paper I study the ϕ4\vec{\phi}^{4} theory in 3+1D at large NN with a negative coupling constant g<0-g<0, and with the scalar field taking values in a 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric complex domain. The theory is non-trivial, has asymptotic freedom, and has a Landau pole in the IR, and I demonstrate that the thermal partition function matches that of the positive-coupling λ>0\lambda>0 theory when the Landau poles of the two theories (in the λ>0\lambda>0 case a pole in the UV) are identified with one another. Thus the ϕ4\vec{\phi}^{4} theory at large NN appears to have a negative bare coupling constant; the coupling only becomes positive in the IR, which in the context of other 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric and large-NN quantum field theories I argue is perfectly acceptable.

I Introduction

Recently, Lagrangians with negative or even complex couplings have attracted interest in the context of renormalizable large-NN quantum field theories [1, 2, 3, 4, 5]. As an example, ϕ4\vec{\phi}^{4} theory111Here, ϕ4\vec{\phi}^{4} is a shorthand (in rather abusive notation) for (ϕϕ)2=(ϕ2)2(\vec{\phi}\cdot\vec{\phi})^{2}=(\vec{\phi}^{2})^{2}, which is the form of the interaction term in this theory. with NN-component scalars in 3+1D has been investigated by Romatschke non-perturbatively222Here, I mean non-perturbatively in the coupling. in the large-NN limit. The theory has a Landau pole (even non-perturbatively in the coupling), and it has been argued that the interacting continuum limit of the theory must have a negative bare coupling constant and exhibits asymptotic freedom [2, 4, 5], similarly to a proposal by Symanzik in the 70s for a ϕ4\phi^{4} theory with negative coupling that is asymptotically free [6, 7, 8].

To see if this is the case, one should explicitly study the negative-coupling ϕ4\vec{\phi}^{4} theory in 3+1D, as will be done in this work. I will use the theory’s path integral; however, a negative coupling constant in the Euclidean action for the ϕ4\vec{\phi}^{4} theory leads to an unbounded path integral over real-valued scalar fields ϕ(x)N\vec{\phi}(x)\in\mathbb{R}^{N}. Therefore, for the negative-coupling theory, the path integral must be made convergent by integrating over an appropriate half-dimensional subspace of all complex-valued scalar fields ϕ(x)N\vec{\phi}(x)\in\mathbb{C}^{N}.

As an additional complication, not every complex domain of path integration is “appropriate”; that is, not every domain necessarily corresponds to a physical theory. For example, the path integral for the negative-coupling ϕ4\vec{\phi}^{4} theory in 1D converges on the domain ϕ(τ)=s(τ)eiπ/4\vec{\phi}(\tau)=\vec{s}(\tau)e^{\textrm{i}\pi/4} for s(τ)N\vec{s}(\tau)\in\mathbb{R}^{N}, but this domain yields a complex-valued partition function333This corresponds to analytically continuing the eigenvalues of the positive-coupling Hamiltonian En(λ)=En(1)λ1/3E_{n}(\lambda)=E_{n}(1)\lambda^{1/3} to negative values of the coupling λ<0\lambda<0, i.e. rotating the En(λ)E_{n}(\lambda) by a complex phase factor eiπ/3e^{\textrm{i}\pi/3}.. Certain complex domains of path integration do, nonetheless, correspond to a physical partition function. Such domains appear, for example, in 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric quantum theories (see e.g. the construction of the path integral in [9]).

𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theories have non-Hermitian Hamiltonians with negative or complex couplings and yet can exhibit spectra that are real and bounded from below, as was first discovered by Bender and Boettcher [10]. With the additional construction of an appropriate 𝒞\mathcal{C} operator and a 𝒞𝒫𝒯\mathcal{C}\mathcal{P}\mathcal{T} inner product of states [11], probabilities in 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theories can be defined and calculated. Here it should be noted that ϕ4\phi^{4} field theory (for N=1N=1) with a negative coupling constant can be given meaning as a 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory and in 3+1D it is asymptotically free [12, 13].

It is interesting to note that, similar to the NN-component ϕ4\vec{\phi}^{4} theory in 3+1D, the Lee model [14] from the 1950s also exhibits a divergence in the coupling at some critical scale, above which the squared coupling g2g^{2} becomes negative. In this case the Lee model can be interpreted as a quantum field theory with a non-Hermitian Hamiltonian and a negative squared coupling constant, as pointed out by Kleefeld [15]. In fact, the Lee model can be interpreted as a 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory, and the negative-norm ghost states that appear from the renormalization can be reinterpreted as physical (positive-norm) states when the 𝒞𝒫𝒯\mathcal{C}\mathcal{P}\mathcal{T} inner product is introduced [16, 17]. Thus, Hermiticity can be replaced, for example by (unbroken) 𝒫𝒯\mathcal{P}\mathcal{T} symmetry, or in general an antilinear symmetry [18], as a requirement for a physical theory. Moreover, the example of the Lee model suggests that a theory can have a negative bare coupling constant.

On the path integral level, a 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory simply corresponds to integrating over any complex domain that asymptotically terminates within certain regions called Stoke’s wedges444See equation (61) in Bender et al. [9] and the surrounding discussion. The path integral of the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory is simply defined as an integral over an appropriate complex domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}}. The 𝒞\mathcal{C} operator does not show up explicitly in the path integral picture. See also [19] for this last point.. For the gϕ4-g\phi^{4} theory these lie at angles in the lower half of the complex plane [10]. Therefore, even with a negative coupling constant, there exist domains of path integration that give physical theories. Because such domains do exist, a negative-coupling theory can have meaning; that is, the theory can be predictive. Making sense of the negative-coupling theory simply amounts to finding such a domain.

In [2, 4] that studied the NN-component ϕ4\vec{\phi}^{4} model, the partition function for the negative-coupling theory was not calculated directly. Rather, it was calculated assuming the unproven recent conjecture by Ai, Bender, and Sarkar in [20] for the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric gϕ4-g\phi^{4} theory. Therefore, there has not yet been an explicit demonstration that the negative-coupling theory (path-integrated on some appropriate domain) is really a continuum limit of the positive-coupling theory. Such a demonstration is the aim of this paper.

That is to say, in this paper I show that there exists at least some complex domain of path integration for which the negative-coupling ϕ4\vec{\phi}^{4} theory at large NN has the same partition function as the positive-coupling theory. This domain respects the 𝒫𝒯\mathcal{P}\mathcal{T} symmetry ϕϕ\vec{\phi}\to-\vec{\phi}, ii\textrm{i}\to-\textrm{i} (although it is not the usual domain one might consider for the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory, which I will define).

This paper is organized as follows. First, I review the result at large NN for the partition function for the positive-coupling ϕ4\vec{\phi}^{4} theory (i.e. the standard O(N)\textrm{O}(N) model) in 3+1D. Then, I introduce the negative-coupling theory. I define a domain on which the partition function of the gϕ4/N-g\vec{\phi}^{4}/N theory can be calculated at large NN. Next, I renormalize the coupling gg, and demonstrate that the partition function of the negative-coupling theory on this domain is equivalent to that of the positive-coupling theory after renormalization. We will see that the negative-coupling theory is asymptotically free and has no Landau pole in the UV, indicating that the ϕ4\vec{\phi}^{4} theory at large NN really is non-trivial and has an interacting continuum limit, as was argued by Romatschke [5]. Lastly, I’ll conclude with remarks about future directions of this work.

II Calculation

In the large-NN expansion, the thermal partition function of the theory can be written as

ZeNβVp(β),Z\propto e^{N\beta Vp(\beta)}, (1)

where VV\to\infty is the volume of space and p(β)p(\beta) is the pressure per component as a function of inverse temperature β=1/T\beta=1/T. At leading order in large NN, p(β)p(\beta) will not depend on NN.

In this section I first review the leading-order result for the partition function Z+Z_{+} of the positive-coupling theory, and then I calculate the partition function ZZ_{-} for the negative-coupling theory defined on some complex domain 𝒞\mathcal{C}_{-} of path integration.

II.1 Review of the positive-coupling theory

The thermal partition function of the positive-coupling theory is formally given by

Z+𝒟NϕeS+[ϕ]Z_{+}\propto\int\mathcal{D}^{N}\vec{\phi}\,e^{-\int S_{+}[\vec{\phi}]} (2)

where the path integral is over real-valued fields ϕ(x)N\vec{\phi}(x)\in\mathbb{R}^{N} with periodic boundary conditions in Euclidean time τ\tau: ϕ(τ+β,𝐱)=ϕ(τ,𝐱)\vec{\phi}(\tau+\beta,\mathbf{x})=\vec{\phi}(\tau,\mathbf{x}). The Euclidean action S+[ϕ]S_{+}[\vec{\phi}] is

S+[ϕ]=β,Vd4x(12(μϕ)2+λNϕ4),S_{+}[\vec{\phi}]=\int_{\beta,V}\textrm{d}^{4}x\,\bigg{(}\frac{1}{2}(\partial_{\mu}\vec{\phi})^{2}+\frac{\lambda}{N}\vec{\phi}^{4}\bigg{)},

where β,Vd4x=0βdτVd3𝐱\int_{\beta,V}\textrm{d}^{4}x=\int_{0}^{\beta}\textrm{d}\tau\int_{V}\textrm{d}^{3}\mathbf{x}. In this work I only consider the massless case, but it is easy to generalize to the case with a mass term Mϕ2M\vec{\phi}^{2} as done in [4, 5].

For this theory, the expression for p(β)p(\beta) in (1) has already been calculated and can be found for example in [4]. In dimensional regularization in (3–2ϵ2\epsilon)+1D, the expression is

p(β)=m416λ+m464π2(1ϵ+ln(μ¯2m2)+32)+m22π2β2n=1K2(nβm)n2.\begin{split}p(\beta)=&\frac{m^{4}}{16\lambda}+\frac{m^{4}}{64\pi^{2}}\bigg{(}\frac{1}{\epsilon}+\ln\bigg{(}\frac{\bar{\mu}^{2}}{m^{2}}\bigg{)}+\frac{3}{2}\bigg{)}\\ &+\frac{m^{2}}{2\pi^{2}\beta^{2}}\sum_{n=1}^{\infty}\frac{K_{2}(n\beta m)}{n^{2}}.\end{split} (3)

Here, μ¯\bar{\mu} is the MS¯\overline{\textrm{MS}} scale and K2(x)K_{2}(x) is a modified Bessel function of the second kind. mm is a gap parameter that is fixed by solving the gap equation p(β)/m2=0\partial p(\beta)/\partial m^{2}=0. Only the solution to the gap equation which gives the larger pressure contributes in the large-volume limit.

The expression (3) for p(β)p(\beta) has a 1/ϵ1/\epsilon divergence from the dimensional regularization which can be absorbed into the renormalized coupling λr(μ¯)\lambda_{\textsc{r}}(\bar{\mu}) by defining

1λr(μ¯)=1λ(μ¯)+14π2ϵ.\frac{1}{\lambda_{\textsc{r}}(\bar{\mu})}=\frac{1}{\lambda(\bar{\mu})}+\frac{1}{4\pi^{2}\epsilon}. (4)

The renormalized coupling depends on μ¯\bar{\mu} such that the pressure per component p(β)p(\beta) does not depend on μ¯\bar{\mu}. This requirement, p(β)/μ¯=0\partial p(\beta)/\partial\bar{\mu}=0, which gives the beta function, fixes the running coupling to be

λr(μ¯)=4π2ln(Λlp2/μ¯2),\lambda_{\textsc{r}}(\bar{\mu})=\frac{4\pi^{2}}{\ln(\Lambda_{\textsc{lp}}^{2}/\bar{\mu}^{2})}, (5)

with Λlp\Lambda_{\textsc{lp}} a scale that appears from the constant in the integration of the beta function. Λlp\Lambda_{\textsc{lp}} is determined for example by fixing the value λ0\lambda_{0} of the coupling λr(Λ0)=λ0\lambda_{\textsc{r}}(\Lambda_{0})=\lambda_{0} at some UV scale Λ0\Lambda_{0}. The running of the coupling is visualized in figure 1.

Refer to caption
Figure 1: A plot of the renormalized running coupling λr(μ¯)\lambda_{\textsc{r}}(\bar{\mu}) from equation (5) in the λϕ4/N\lambda\vec{\phi}^{4}/N theory as a function of MS¯\overline{\textrm{MS}} scale μ¯\bar{\mu}. One sees that at the scale μ¯=Λlp\bar{\mu}=\Lambda_{\textsc{lp}} of the Landau pole, the coupling diverges, and above that scale, the coupling becomes negative.

One immediate objection can be made. If the coupling has some non-zero, positive value at some finite scale μ¯\bar{\mu}, then the scale Λlp>μ¯\Lambda_{\textsc{lp}}>\bar{\mu} will also be finite. Above the scale Λlp\Lambda_{\textsc{lp}}, the coupling in (5) becomes negative. Therefore, there is no way to take the UV scale Λ0\Lambda_{0} to infinity and keep the UV coupling λ0\lambda_{0} positive. Since a positive coupling was assumed in order to make the calculation (3) of p(β)p(\beta) and of the partition function Z+Z_{+}, this is problematic555However, I am being a little heuristic. It is really the coupling λ\lambda, and not λr\lambda_{\textsc{r}}, that was assumed to be positive here. In dimensional regularization in (3–2ϵ\epsilon)+1D this means that λ(μ¯)\lambda(\bar{\mu}) becomes negative at a scale μ¯=Λlpe1/2ϵ\bar{\mu}=\Lambda_{\textsc{lp}}\,e^{-1/2\epsilon}. The fact remains that the coupling diverges and becomes negative at some scale. In cutoff regularization with cutoff Λ\Lambda, there is a similar scale Λ=Λlp\Lambda=\Lambda_{\textsc{lp}} above which the coupling λ(Λ)\lambda(\Lambda) becomes negative [5], and λ(Λ)\lambda(\Lambda) is identical to λr(μ¯)\lambda_{\textsc{r}}(\bar{\mu}) in (5) if μ¯\bar{\mu} is identified with Λ\Lambda.. The interacting theory appears to have no continuum limit Λ0\Lambda_{0}\to\infty, and if one tries to fix this problem by taking Λlp\Lambda_{\textsc{lp}} to infinity, the coupling at any finite scale μ¯<\bar{\mu}<\infty will be zero, such that the theory is trivial (non-interacting). This is a feature of scalar ϕ4\phi^{4} field theory, which is known to be quantum trivial in 3+1D for N=1,2N=1,2 components [21, 22].

As an additional note, the coupling λr(μ¯)\lambda_{\textsc{r}}(\bar{\mu}) will diverge at the scale μ¯=Λlp\bar{\mu}=\Lambda_{\textsc{lp}}. Therefore the ϕ4\vec{\phi}^{4} theory at large NN has a Landau pole even non-perturbatively (as opposed to only perturbatively).

Despite these potential issues, assuming that μ¯<Λlp\bar{\mu}<\Lambda_{\textsc{lp}} and that the ϕ4\vec{\phi}^{4} theory is simply an effective (cut-off) theory, the calculation of the pressure p(β)p(\beta) in (3) can proceed, and has been done by Romatschke, and the result plotted, in [2, 4]. The renormalized pressure is given by

p(β)=m464π2(ln(Λlp2m2)+32)+m22π2β2n=1K2(nβm)n2.\begin{split}p(\beta)=&\frac{m^{4}}{64\pi^{2}}\bigg{(}\ln\bigg{(}\frac{\Lambda_{\textsc{lp}}^{2}}{m^{2}}\bigg{)}+\frac{3}{2}\bigg{)}\\ &+\frac{m^{2}}{2\pi^{2}\beta^{2}}\sum_{n=1}^{\infty}\frac{K_{2}(n\beta m)}{n^{2}}.\end{split} (6)

At low temperatures TTc0.616ΛlpT\leq T_{\textsc{c}}\approx 0.616\,\Lambda_{\textsc{lp}}, one finds two solutions mm to the gap equation p(β)/m2=0\partial p(\beta)/\partial m^{2}=0. The dominant solution has m2=eΛlp2m^{2}=e\Lambda_{\textsc{lp}}^{2} at T=0T=0 and determines the pressure p(β)p(\beta) via (6). At higher temperatures T>TcT>T_{\textsc{c}} there are also two solutions to the gap equation, but now they are a complex-conjugate pair of solutions m+,m=m+m_{+},m_{-}=m_{+}^{*} corresponding to complex-conjugate pressures p+(β),p(β)=p+(β)p_{+}(\beta),p_{-}(\beta)=p_{+}^{*}(\beta). Putting the partition function Z+Z_{+} into the large-NN form of (1) leaves one with

Z+eNβVRep+(β)+lncos(NβVImp+(β))eNβVRep+(β),β<βc,\begin{split}Z_{+}&\propto e^{N\beta V\operatorname{Re}p_{+}(\beta)+\ln\cos\big{(}N\beta V\operatorname{Im}p_{+}(\beta)\big{)}}\\ &\approx e^{N\beta V\operatorname{Re}p_{+}(\beta)},\,\,\,\,\beta<\beta_{\textsc{c}},\end{split} (7)

which follows formally at large NN from only keeping the part of lnZ+\ln Z_{+} that scales like NN 666Note that the argument for setting p(β)=Rep+(β)p(\beta)=\operatorname{Re}p_{+}(\beta) in [2, 4] was based on the unproven conjecture in [20] for the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric gϕ4-g\phi^{4} theory. Here instead I’ve alternatively based it on the formal expansion of lnZ\ln Z in large NN. Later, in subsection III.2, I discuss the consequences of not neglecting the imaginary parts of both saddles, and how it suggests 𝒫𝒯\mathcal{P}\mathcal{T} symmetry breaking at high TT.. One finds that p(β<βc)=Rep+(β)p(\beta<\beta_{\textsc{c}})=\operatorname{Re}p_{+}(\beta) is physically well-behaved (e.g. increases with temperature) and p(β)p(\beta) has a continuous first derivative at β=βc\beta=\beta_{\textsc{c}}.

Interestingly enough, the pressure per component p(β)p(\beta) of the λϕ4/N\lambda\vec{\phi}^{4}/N theory, calculated in this way, asymptotes toward the Stefan–Boltzmann limit for a single free boson at high temperatures β0\beta\to 0 [2], suggesting that the ϕ4\vec{\phi}^{4} theory at large NN is asymptotically free. This should not be true based on the argument that the Landau pole in (5) prevented the positive-coupling theory from having an interacting continuum limit. Moreover, the beta function λr/lnμ¯\partial\lambda_{\textsc{r}}/\partial\ln\bar{\mu} has the same sign as the coupling assuming μ¯<Λlp\bar{\mu}<\Lambda_{\textsc{lp}}. Asymptotically free theories, on the other hand, are interacting, do have a continuum limit, and have a beta function whose sign is opposite that of the coupling as μ¯\bar{\mu}\to\infty.

The solution to this this apparent contradiction, and to the problem of triviality, is to allow the bare coupling to be negative. Non-triviality and a negative bare coupling were argued in [5]. But if the bare coupling is negative, the calculation of p(β)p(\beta) must be done in some other way that assumes negative coupling from the start.

That is the main new result and next part of this work.

II.2 Defining the negative-coupling theory

The thermal partition function for the negative-coupling theory is given by

Z𝒞𝒟NϕeS[ϕ]Z_{-}\propto\int_{\mathcal{C}_{-}}\mathcal{D}^{N}\vec{\phi}\,e^{-\int S_{-}[\vec{\phi}]} (8)

where the fields still have periodic boundary conditions in Euclidean time but now the path integral is over complexified fields that live on the domain 𝒞\mathcal{C}_{-} (which will be specified shortly). The Euclidean action S[ϕ]S_{-}[\vec{\phi}] is

S[ϕ]=β,Vd4x(12(μϕ)2gNϕ4),S_{-}[\vec{\phi}]=\int_{\beta,V}\textrm{d}^{4}x\,\bigg{(}\frac{1}{2}(\partial_{\mu}\vec{\phi})^{2}-\frac{g}{N}\vec{\phi}^{4}\bigg{)}, (9)

where g<0-g<0 is now the coupling.

A choice for the domain 𝒞\mathcal{C}_{-} that one might consider for the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric gϕ4/N-g\vec{\phi}^{4}/N theory is the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} parametrized by NN real-valued fields χ(x)\chi(x)\in\mathbb{R}, η(x)N1\vec{\eta}(x)\in\mathbb{R}^{N\textrm{--}1} as

ϕ(x)e^=χ(x)f(χ(x))ϕ(x)(ϕ(x)e^)e^=η(x)f(χ(x)),\begin{split}\vec{\phi}(x)&\cdot\hat{e}=\chi(x)f\big{(}\chi(x)\big{)}\\ \vec{\phi}(x)-&\big{(}\vec{\phi}(x)\cdot\hat{e}\big{)}\hat{e}=\vec{\eta}(x)f\big{(}\chi(x)\big{)},\end{split} (10)

where e^\hat{e} is some unit vector in N\mathbb{R}^{N} and the function ff is defined as

f(χ)θ(χ)eiπ/4+θ(χ)eiπ/4,f(\chi)\equiv\theta\big{(}\chi)e^{-\textrm{i}\pi/4}+\theta(\textrm{--}\chi)e^{\textrm{i}\pi/4}, (11)

where θ\theta is the Heaviside step function. This forces one component ϕ(x)e^\vec{\phi}(x)\cdot\hat{e} of the field ϕ\vec{\phi} to live on the union of two half-lines at angles π/4-\pi/4, 3π/4-3\pi/4 in the lower half of the complex plane at every point xx in Euclidean spacetime. If every component lived independently on the same union of half-lines, the path integral would be unbounded, as was pointed out in [23]. This domain respects the 𝒫𝒯\mathcal{P}\mathcal{T} symmetry ϕϕ\vec{\phi}\to-\vec{\phi}, ii\textrm{i}\to-\textrm{i}, which is also a symmetry of the action S[ϕ]S_{-}[\vec{\phi}], and any deformation of this domain with the same Lefschetz thimble decomposition (e.g. terminating within the associated Stoke’s wedges) gives the same partition function Z=Z𝒫𝒯Z_{-}=Z_{\mathcal{P}\mathcal{T}}.

Refer to caption
Figure 2: A visualization of the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric domains of path integration 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10) and 𝒞\mathcal{C}_{-} in (12)\eqref{eq:domain-neg-coup}. The axes are expressed in some unit of mass (it does not matter which). For the case of 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}}, ϕ(x)e^\vec{\phi}(x)\cdot\hat{e} lives on either the solid red line or the solid blue line, and every component of ϕ(x)(ϕ(x)e^)e^\vec{\phi}(x)-\big{(}\vec{\phi}(x)\cdot\hat{e}\big{)}\hat{e} lives on the corresponding dashed line of the same color (red or blue). For the case of 𝒞\mathcal{C}_{-}, the zero mode ϕ0e^\vec{\phi}_{0}\cdot\hat{e} lives on either the solid red or the solid blue line, while the non-zero modes ϕ(x)e^\vec{\phi}^{\prime}(x)\cdot\hat{e} and every component of ϕ(x)(ϕ(x)e^)e^\vec{\phi}(x)-\big{(}\vec{\phi}(x)\cdot\hat{e}\big{)}\hat{e} live on the corresponding dashed line of the same color.

The path integral on the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10), however, is difficult to solve analytically using standard large-NN techniques, because there are an infinite number of sharp kinks in the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}}, with one kink at χ(x)=0\chi(x)=0 for every point xx in spacetime. These kinks can be managed straightforwardly in a lattice calculation (and the N=1N=1 lattice calculation in 3+1D was done in [5] and in 1D in [23]), but on the lattice the path integral (8) with 𝒞=𝒞𝒫𝒯\mathcal{C}_{-}=\mathcal{C}_{\mathcal{P}\mathcal{T}} from (10) has a sign problem, so the calculation is difficult to scale to NN components.

One could consider a domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} without kinks by generalizing the hyperbola ϕ(x)=2i1+iχ(x)\phi(x)=-2\textrm{i}\sqrt{1+\textrm{i}\chi(x)} for χ(x)\chi(x)\in\mathbb{R} in [9, 24] to NN-component scalar fields, but it is unclear how to do this. One such generalization to NN components for the 1D theory was done in [25] by letting the radial coordinate live on this complex hyperbola, i.e. ϕ 2(x)=4(1+iχ(x))\vec{\phi}^{\,2}(x)=-4\big{(}1+\textrm{i}\chi(x)\big{)} for χ(x)\chi(x)\in\mathbb{R}, but this generalization gives an unbounded Hamiltonian spectrum and thus an unbounded partition function in 1D once angular momentum is considered777The interested reader can check this by including the angular momentum quantum number \ell in the radial Schrödinger equation and finding the Hermitian Hamiltonian which is equivalent to the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric non-Hermitian Hamiltonian along the lines of [24, 9, 25].; it likewise gives an unbounded path integral in 3+1D. Another parametrization in the literature, ϕi(x)=2ici+iχi(x)\phi_{i}(x)=-2\textrm{i}\sqrt{c_{i}+\textrm{i}\chi_{i}(x)} [26], also gives an unbounded path integral upon closer inspection because there is a flat direction χiχj\chi_{i}-\chi_{j} for iji\neq j in the potential.

There is an alternative choice for the domain 𝒞\mathcal{C}_{-} that gets rid of most of these kinks by only forcing one component of the Fourier zero mode of the scalar field to live on the half-lines at π/4-\pi/4, 3π/4-3\pi/4 in the lower half of the complex plane. Let ϕ(x)=ϕ0+ϕ(x)\vec{\phi}(x)=\vec{\phi}_{0}+\vec{\phi}^{\prime}(x), where ϕ0\vec{\phi}_{0} is the zero mode of the scalar field and ϕ\vec{\phi}^{\prime} contains all the non-zero modes. Then the domain 𝒞\mathcal{C}_{-} can be parametrized in terms of the real-valued fields χ(x)=χ0+χ(x)\chi(x)=\chi_{0}+\chi^{\prime}(x)\in\mathbb{R} and η(x)N1\vec{\eta}(x)\in\mathbb{R}^{N\textrm{--}1} as

ϕ0e^=χ0f(χ0),ϕ(x)e^=χ(x)f(χ0),ϕ(x)(ϕ(x)e^)e^=η(x)f(χ0),\begin{split}\vec{\phi}_{0}\cdot\hat{e}&=\chi_{0}f(\chi_{0}),\\ \vec{\phi}^{\prime}(x)\cdot&\hat{e}=\chi^{\prime}(x)f(\chi_{0}),\\ \vec{\phi}(x)-&\big{(}\vec{\phi}(x)\cdot\hat{e}\big{)}\hat{e}=\vec{\eta}(x)f(\chi_{0}),\end{split} (12)

where χ0\chi_{0} and χ\chi^{\prime} denote the zero mode and non-zero modes of the field χ\chi, respectively, and where ff is the same function defined in (11). See figure 2 for a visualization of this domain and the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10)\eqref{eq:domain-PT}. This domain still gives a bounded path integral and respects the 𝒫𝒯\mathcal{P}\mathcal{T} symmetry ϕϕ\vec{\phi}\to-\vec{\phi}, ii\textrm{i}\to-\textrm{i}, but now it makes the path integral amenable to large-NN techniques, as we will see. Moreover, it gives a physical partition function. However, this domain does not (as far as I can tell) correspond to a Hamiltonian picture where quantum states and an inner product like the 𝒞𝒫𝒯\mathcal{C}\mathcal{P}\mathcal{T} inner product can be defined. The domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in equation (10) however will have a Hamiltonian, inner product, states, and an associated notion of unitary. For ease of calculation, the domain 𝒞\mathcal{C}_{-} specified by (12) is the choice that will be considered in this paper for the negative-coupling theory.

One more remark should be made before continuing on to the calculation of the partition function ZZ_{-}. It is perhaps unsatisfactory that the domain 𝒞\mathcal{C}_{-} in (12), and also the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10), breaks the usual O(N)\textrm{O}(N) symmetry of the theory (which is only a symmetry of the action) down to O(N1)\textrm{O}(N-1). However, it is not apparent how to keep the O(N)\textrm{O}(N) symmetry and the 𝒫𝒯\mathcal{P}\mathcal{T} symmetry while keeping the path integral bounded. As was already mentioned, letting the radial coordinate live on the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric hyperbola ϕ 2(x)=4(1+iχ(x))\vec{\phi}^{\,2}(x)=-4\big{(}1+\textrm{i}\chi(x)\big{)}, as was done for the 1D NN-component theory in [25], gives an unbounded path integral, even though it keeps the O(N)\textrm{O}(N) symmetry.

It should be pointed out that the calculation of the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory’s partition function using the constraint method (also applied in this paper in subsection II.3) was done by Ogilvie and Meisinger in [26] without explicitly breaking the O(N)\textrm{O}(N) symmetry. These authors acquired a result equivalent to (23) in this work and also to the Hermitian theory. However in using the constraint calculation they do not specify a domain of integration for the field ϕ\vec{\phi} or for the auxiliary fields, nor do they renormalize the theory or do the Lefschetz thimble analysis that is done later in this work. The only domain they specify, earlier in the work, gives an unbounded path integral and also breaks the O(N)\textrm{O}(N) symmetry, and is not used explicitly in the calculation, unlike what will be done with the domain 𝒞\mathcal{C}_{-} here.

Alternatively, there are composite 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric domains that keep the O(N)\textrm{O}(N) symmetry, for example by extending the half-lines in the lower half of the complex plane in figure 2 to full lines, with the sacrifice that the domain of integration intersects itself and is no longer parametrized simply by NN real fields. I note that if this is done for the domain 𝒞\mathcal{C}_{-}, the O(N)\textrm{O}(N) symmetry can be maintained, and one still gets the same result as in this work. One can also restore the O(N)\textrm{O}(N) symmetry, allowing it only to be spontaneously broken, by integrating the unit vector e^\hat{e} in (10) and (12) over the sphere SN1S^{N-1}, and one will again get the same result as this work.

Here, I will stick to the domain 𝒞\mathcal{C}_{-} in (12) with the intent of demonstrating there is at least some domain 𝒞\mathcal{C}_{-} for which the negative-coupling theory “renormalizes into” the positive-coupling theory.

II.3 Calculating the partition function ZZ_{-}

The partition function ZZ_{-} will now be calculated. With the choice in (12) for the domain 𝒞\mathcal{C}_{-}, and in terms of the real fields χ\chi, η\vec{\eta}, the path integral (8) for the negative-coupling theory becomes

Z+,±eiNπVΣ 

/4
(0±dχ0
𝒟χ𝒟N1ηeS±[χ,η]).
\begin{split}Z_{-}\propto\sum_{+,-}\pm&e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}\bigg{(}\int_{0}^{\pm\infty}\textrm{d}\chi_{0}\\ &\int\mathcal{D}\chi^{\prime}\,\mathcal{D}^{N\textrm{--}1}\vec{\eta}\,\,e^{-S_{-}^{\pm}[\chi,\vec{\eta}]}\bigg{)}.\end{split}
(13)

Here, the symbol Σ 

\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}
is shorthand for Σ 

nd3𝐤/(2π)3
\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}\equiv\sum_{n}\int\textrm{d}^{3}\mathbf{k}/(2\pi)^{3}
(a sum over Matsubara frequencies ωn\omega_{n} and a integral over 3-momenta 𝐤\mathbf{k}), and S±[χ,η]S_{-}^{\pm}[\chi,\vec{\eta}] is given by

S±[χ,η]=β,Vd4x(±12i((μχ)2+(μη)2)+gN(χ2+η 2)2).\begin{split}S_{-}^{\pm}[\chi,\vec{\eta}]=\int_{\beta,V}\textrm{d}^{4}x\,\bigg{(}\pm\frac{1}{2\textrm{i}}&\big{(}(\partial_{\mu}\chi)^{2}+(\partial_{\mu}\vec{\eta})^{2}\big{)}\\ &+\frac{g}{N}(\chi^{2}+\vec{\eta}^{\,2})^{2}\bigg{)}.\end{split} (14)

The fields χ\chi, η\vec{\eta} have periodic boundary conditions in Euclidean time. We see the path integral becomes a sum of two pieces, each corresponding to one of the half-lines at angles π/4-\pi/4, 3π/4-3\pi/4 on which the zero mode ϕ0e^\vec{\phi}_{0}\cdot\hat{e} lives.

In order to solve this integral at large NN, it is convenient to perform a Hubbard–Stratonovich transformation by introducing the auxiliary fields ζ\zeta and σ=χ2+η 2\sigma=\chi^{2}+\vec{\eta}^{\,2}, as follows:

Z+,±eiNπVΣ 

/4
(0±dχ0
𝒟ζ𝒟σ𝒟χ𝒟N1ηeS±[χ,η,ζ,σ]),
\begin{split}Z_{-}\propto\sum_{+,-}&\pm e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}\bigg{(}\int_{0}^{\pm\infty}\textrm{d}\chi_{0}\\ &\int\mathcal{D}\zeta\,\mathcal{D}\sigma\,\mathcal{D}\chi^{\prime}\,\mathcal{D}^{N\textrm{--}1}\vec{\eta}\,\,e^{-S_{-}^{\pm}[\chi,\vec{\eta},\zeta,\sigma]}\bigg{)},\end{split}
(15)

where S±[χ,η,ζ,σ]S_{-}^{\pm}[\chi,\vec{\eta},\zeta,\sigma] is given by

S±[χ,η,ζ,σ]=β,Vd4x(±12i((μχ)2+(μη)2)+gNσ2+i2ζ(χ2+η 2σ)).\begin{split}S_{-}^{\pm}[\chi,\vec{\eta},\zeta,\sigma]=\int_{\beta,V}&\textrm{d}^{4}x\,\bigg{(}\pm\frac{1}{2\textrm{i}}\big{(}(\partial_{\mu}\chi)^{2}+(\partial_{\mu}\vec{\eta})^{2}\big{)}\\ &+\frac{g}{N}\sigma^{2}+\frac{\textrm{i}}{2}\zeta(\chi^{2}+\vec{\eta}^{\,2}-\sigma)\bigg{)}.\end{split} (16)

This amounts to the insertion of a delta function into the path integral that enforces σ=χ2+η 2\sigma=\chi^{2}+\vec{\eta}^{\,2}. The action (16) is quadratic in the auxiliary field σ\sigma, so σ\sigma can be integrated out to give

Z+,±eiNπVΣ 

/4
(0±dχ0
𝒟ζ𝒟χ𝒟N1ηeS±[χ,η,ζ]),
\begin{split}Z_{-}\propto\sum_{+,-}&\pm e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}\bigg{(}\int_{0}^{\pm\infty}\textrm{d}\chi_{0}\\ &\int\mathcal{D}\zeta\,\mathcal{D}\chi^{\prime}\,\mathcal{D}^{N\textrm{--}1}\vec{\eta}\,\,e^{-S_{-}^{\pm}[\chi,\vec{\eta},\zeta]}\bigg{)},\end{split}
(17)

where S±[χ,η,ζ]S_{-}^{\pm}[\chi,\vec{\eta},\zeta] is now given by

S±[χ,η,ζ]=β,Vd4x(±12i((μχ)2+(μη)2)+i2ζ(χ2+η 2)+Nζ216g).\begin{split}S_{-}^{\pm}[\chi,\vec{\eta},\zeta]=\int_{\beta,V}&\textrm{d}^{4}x\,\bigg{(}\pm\frac{1}{2\textrm{i}}\big{(}(\partial_{\mu}\chi)^{2}+(\partial_{\mu}\vec{\eta})^{2}\big{)}\\ &+\frac{\textrm{i}}{2}\zeta(\chi^{2}+\vec{\eta}^{\,2})+\frac{N\zeta^{2}}{16g}\bigg{)}.\end{split} (18)

The auxiliary field ζ\zeta can be split into a zero mode ζ0\zeta_{0} and non-zero modes ζ\zeta^{\prime} as ζ(x)=ζ0+ζ(x)\zeta(x)=\zeta_{0}+\zeta^{\prime}(x). At leading order in large NN, only the zero mode contributes to the pressure888Including the non-zero modes ζ\zeta^{\prime} amounts to including 1/N1/N corrections, and can be done via Rnn resummation methods [27, 28]., so the integral over the non-zero modes can be neglected [29] and the partition function becomes

Z+,±eiNπVΣ 

/4
(0±dχ0
dζ0𝒟χ𝒟N1ηeS±[χ,η,ζ0]),
\begin{split}Z_{-}\propto\sum_{+,-}&\pm e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}\bigg{(}\int_{0}^{\pm\infty}\textrm{d}\chi_{0}\\ &\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\int\mathcal{D}\chi^{\prime}\mathcal{D}^{N\textrm{--}1}\vec{\eta}\,\,e^{-S_{-}^{\pm}[\chi,\vec{\eta},\zeta_{0}]}\bigg{)},\end{split}
(19)

with ζ\zeta in expression (18) now replaced by only the zero mode ζ0\zeta_{0}.

There is a term iζ0(χ2+η2)/2\textrm{i}\zeta_{0}(\chi^{2}+\eta^{2})/2 in the action S±[χ,η,ζ0]S_{-}^{\pm}[\chi,\vec{\eta},\zeta_{0}] in (19) which contains a cross-term iζ0χ0χ\textrm{i}\zeta_{0}\chi_{0}\chi^{\prime} between the zero and non-zero modes of χ\chi. Because ζ0χ0\zeta_{0}\chi_{0} does not vary with xx and χ\chi^{\prime} contains only non-zero modes, this cross-term vanishes under the integral β,Vdx4\int_{\beta,V}\textrm{d}x^{4}, leaving one with a term iζ0(χ02+χ2+η2)/2\textrm{i}\zeta_{0}(\chi_{0}^{2}+\chi^{\prime 2}+\eta^{2})/2 in the action. The action S±[χ,η,ζ0]S_{-}^{\pm}[\chi,\vec{\eta},\zeta_{0}] is thus an even function of χ0\chi_{0}, and so the bounds of the integral over χ0\chi_{0} can be extended from [0,±)[0,\pm\infty) to (,)(-\infty,\infty) without consequence, which gives

Z+,eiNπVΣ 

/4
dζ0𝒟NχeS±[χ,ζ0]
,
\begin{split}Z_{-}\propto\sum_{+,-}&e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\int\mathcal{D}^{N}\vec{\chi}\,\,e^{-S_{-}^{\pm}[\vec{\chi},\zeta_{0}]},\end{split}
(20)

where χ(x)=(χ(x),η(x))N\vec{\chi}(x)=\big{(}\chi(x),\vec{\eta}(x)\big{)}\in\mathbb{R}^{N} has been introduced simply to group together the fields χ(x)\chi(x) and η(x)\vec{\eta}(x), and S±[χ,ζ0]S_{-}^{\pm}[\vec{\chi},\zeta_{0}] is

S±[χ,ζ0]=βVNζ0216g+β,Vdx4(±12i(μχ)2+i2ζ0χ 2).\begin{split}S_{-}^{\pm}[\vec{\chi},\zeta_{0}]=&\beta V\frac{N\zeta_{0}^{2}}{16g}\\ &+\int_{\beta,V}\textrm{d}x^{4}\bigg{(}\pm\frac{1}{2\textrm{i}}(\partial_{\mu}\vec{\chi})^{2}+\frac{\textrm{i}}{2}\zeta_{0}\vec{\chi}^{\,2}\bigg{)}.\end{split} (21)

The action (21) is quadratic in χ\vec{\chi}, and so χ\vec{\chi} can be integrated out, yielding

Z+,eiNπVΣ 

/4
(dζ0
eNlndet(±iμ2+iζ0)/2βVNζ02/16g)+,dζ0eNtrln(μ2ζ0)/2βVNζ02/16g=+,dζ0eNβVp(β,ζ0).
\begin{split}Z_{-}\propto\,\,&\sum_{+,-}e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}\bigg{(}\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\\ &\,\,\,\,\,\,\,\,e^{-N\ln\det(\pm\textrm{i}\partial_{\mu}^{2}+\textrm{i}\zeta_{0})/2-\beta VN\zeta_{0}^{2}/16g}\bigg{)}\\ \propto\,\,&\sum_{+,-}\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\,e^{-N\textrm{tr}\ln(-\partial_{\mu}^{2}\mp\zeta_{0})/2-\beta VN\zeta_{0}^{2}/16g}\\ &=\,\,\sum_{+,-}\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\,e^{N\beta Vp(\beta,\mp\zeta_{0})}.\end{split}
(22)

Note that the complex Jacobian term eiNπVΣ 

/4
e^{\mp\textrm{i}N\pi V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}/4}
has been canceled by pulling out a eNtrln(i)/2e^{-N\textrm{tr}\ln(\mp\textrm{i})/2} from inside the path integral, where the trace of the operator on a single field component brings out a phase space factor VΣ 

V\mathord{\mathchoice{\ooalign{\hfil$\displaystyle\Sigma$\hfil\cr\hfil$\displaystyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\textstyle\Sigma$\hfil\cr\hfil$\textstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptstyle\Sigma$\hfil\cr\hfil$\scriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}{\ooalign{\hfil$\scriptscriptstyle\Sigma$\hfil\cr\hfil$\scriptscriptstyle\textrm{\hskip 1.0pt\scalebox{1.2}{$\int$}}$\hfil\cr}}}
. Here, also, I’ve introduced the pressures per component p(β,ζ0)p(\beta,\mp\zeta_{0}), which are given more explicitly by

p(β,ζ0)=ζ0216g12βn=μ2ϵd32ϵ𝐤(2π)3ln(ωn2+𝐤2ζ0),\begin{split}p(\beta,\mp\zeta_{0})=&-\frac{\zeta_{0}^{2}}{16g}\\ -\frac{1}{2\beta}\sum_{n=-\infty}^{\infty}&\int\mu^{2\epsilon}\frac{\textrm{d}^{3-2\epsilon}\mathbf{k}}{(2\pi)^{3}}\ln(\omega_{n}^{2}+\mathbf{k}^{2}\mp\zeta_{0}),\end{split} (23)

where ωn=2πn/β\omega_{n}=2\pi n/\beta is the nnth Matsubara frequency and the integral over 3-momenta 𝐤\mathbf{k} can be done using dimensional regularization in 32ϵ3-2\epsilon dimensions. μ\mu is the scale of dimensional regularization, related to the MS¯\overline{\textrm{MS}} scale μ¯\bar{\mu}.

The thermal sum-integral in (23) has been done, for example, in [30], and the resulting pressures per component are

p(β,ζ0)=ζ0216g+ζ0264π2(1ϵ+ln(μ¯2ζ0)+32)ζ02π2βn=1K2(nβζ0)n2.\begin{split}p(\beta,\mp\zeta_{0})=&-\frac{\zeta_{0}^{2}}{16g}+\frac{\zeta_{0}^{2}}{64\pi^{2}}\bigg{(}\frac{1}{\epsilon}+\ln\bigg{(}\frac{\bar{\mu}^{2}}{\mp\zeta_{0}}\bigg{)}+\frac{3}{2}\bigg{)}\\ &\mp\frac{\zeta_{0}}{2\pi^{2}\beta}\sum_{n=1}^{\infty}\frac{K_{2}(n\beta\sqrt{\mp\zeta_{0}})}{n^{2}}.\end{split} (24)

This is the same expression as (3) from the positive-coupling theory, with m2m^{2} replaced by m2=ζ0m^{2}=\mp\zeta_{0} and λ\lambda replaced by g-g.

As with the positive-coupling theory, there is a 1/ϵ1/\epsilon divergence from the dimensional regularization that can be absorbed by defining the renormalized coupling gr(μ¯)g_{\textsc{r}}(\bar{\mu}) as

1gr(μ¯)=1g(μ¯)14π2ϵ.\frac{1}{g_{\textsc{r}}(\bar{\mu})}=\frac{1}{g(\bar{\mu})}-\frac{1}{4\pi^{2}\epsilon}. (25)

Then, requiring the pressures to be independent of the MS¯\overline{\textrm{MS}} scale μ¯\bar{\mu}, such that p(β,ζ0)/μ¯=0\partial p(\beta,\mp\zeta_{0})/\partial\bar{\mu}=0, yields the beta function and fixes the running coupling to be

gr(μ¯)=4π2ln(μ¯2/Λlp2),g_{\textsc{r}}(\bar{\mu})=\frac{4\pi^{2}}{\ln(\bar{\mu}^{2}/\Lambda_{\textsc{lp}}^{2})}, (26)

where Λlp\Lambda_{\textsc{lp}} is again a scale introduced from the integration of the beta function. The renormalized pressures p(β,ζ0)p(\beta,\mp\zeta_{0}) are then given by

p(β,ζ0)=ζ0264π2(ln(Λlp2ζ0)+32)ζ02π2β2n=1K2(nβζ0)n2.\begin{split}p(\beta,\mp\zeta_{0})=&\frac{\zeta_{0}^{2}}{64\pi^{2}}\bigg{(}\ln\bigg{(}\frac{\Lambda_{\textsc{lp}}^{2}}{\mp\zeta_{0}}\bigg{)}+\frac{3}{2}\bigg{)}\\ &\mp\frac{\zeta_{0}}{2\pi^{2}\beta^{2}}\sum_{n=1}^{\infty}\frac{K_{2}(n\beta\sqrt{\mp\zeta_{0}})}{n^{2}}.\end{split} (27)
Refer to caption
Figure 3: The imaginary part of the pressure p(β,ζ0)p(\beta,\zeta_{0}) in (27) at a temperature T=1/β=0.65ΛlpT=1/\beta=0.65\,\Lambda_{\textsc{lp}}, showing that the function is multi-sheeted. The horizontal axes are expressed in units of Λlp2\Lambda_{\textsc{lp}}^{2} and the vertical axis in units of Λlp4\Lambda_{\textsc{lp}}^{4}. There is a branch point at ζ0=0\zeta_{0}=0 and a branch cut can be made along the negative half of the real axis. In order to avoid the branch cut, one has to integrate eNβVp(β,ζ0)e^{N\beta Vp(\beta,\zeta_{0})} along a contour ζ0+i0+\zeta_{0}+\textrm{i}0^{+}\in\mathbb{R} slightly shifted in the imaginary direction, which is shown in red.

Note that the running coupling in (26) is exactly the same expression as the running coupling in (5) for the positive-coupling theory up to a minus sign. That is, gr(μ¯)=λr(μ¯)g_{\textsc{r}}(\bar{\mu})=-\lambda_{\textsc{r}}(\bar{\mu}) if the scale Λlp\Lambda_{\textsc{lp}} in both theories is identified as the same scale.

At large NN, the sum of integrals in (22) can be replaced by

Z+,nj±jeNβVp(β,ζ0j),Z_{-}\approx\sum_{+,-}n_{j}^{\pm}\sum_{j}e^{N\beta Vp(\beta,\zeta_{0j})}, (28)

where ζ0j\zeta_{0j} denotes the jjth saddle of p(β,ζ0)p(\beta,\zeta_{0}) and nj±n^{\pm}_{j} denotes the intersection number of the contour of integration with the Lefschetz anti-thimble passing through the jjth saddle. For an introduction to Lefschetz thimbles, see [31]. Only the relevant saddles with nj0n_{j}\neq 0 contribute to the partition function, and only the saddle with the largest pressure dominates in the large-volume limit. In order to determine the saddles, one has to apply the saddle condition (or gap equation) p(β,ζ0)/ζ0=0\partial p(\beta,\mp\zeta_{0})/\partial\zeta_{0}=0 to fix the gap m=ζ0m=\sqrt{\mp\zeta_{0}}.

However, the function for the pressures in (27) has a branch point at ζ0=0\zeta_{0}=0 and is multi-sheeted. See figure 3 for an illustration of this feature. A branch cut can be made on the real axis for ζ0<0\mp\zeta_{0}<0, defining the principle sheet, and the integrals of eNβVp(β,ζ0)e^{N\beta Vp(\beta,\mp\zeta_{0})} in (22) can be done slightly above or below the real axis in order to avoid the branch point. Here there is an apparent ambiguity whether the integration should occur below or above the branch point, but the requirement of 𝒫𝒯\mathcal{P}\mathcal{T} symmetry settles the ambiguity. Since 𝒫𝒯\mathcal{P}\mathcal{T} symmetry interchanges ζ0ζ0\zeta_{0}\to-\zeta_{0} and ii\textrm{i}\to-\textrm{i} 999This is because 𝒫𝒯\mathcal{P}\mathcal{T} exchanges the two half lines at angles π/4-\pi/4, 3π/4-3\pi/4 in figure 2, or equivalently switches the two integrals over χ0>0\chi_{0}>0 and χ0<0\chi_{0}<0 in (13), which in (22) amounts to swapping p(β,ζ0)p(\beta,-\zeta_{0}) with p(β,ζ0)p(\beta,\zeta_{0}), i.e. ζ0ζ0\zeta_{0}\to-\zeta_{0}., the integrals in (22) should be

Zdζ0eNβVp(β,ζ0+i0+)+dζ0eNβVp(β,ζ0i0+)=dζ0eNβVp(β,ζ0+i0+)+dζ0eNβVp(β,ζ0i0+)\begin{split}Z_{-}\propto&\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\,e^{N\beta Vp(\beta,-\zeta_{0}+\textrm{i}0^{+})}\\ &+\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\,e^{N\beta Vp(\beta,\zeta_{0}-\textrm{i}0^{+})}\\ =&\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\,e^{N\beta Vp(\beta,\zeta_{0}+\textrm{i}0^{+})}\\ &+\int_{-\infty}^{\infty}\textrm{d}\zeta_{0}\,e^{N\beta Vp(\beta,\zeta_{0}-\textrm{i}0^{+})}\end{split} (29)

where here I have used the fact that aadxf(x)=aadxf(x)\int_{-a}^{a}\textrm{d}x\,f(-x)=\int_{-a}^{a}\textrm{d}x\,f(x) for any function ff.

The relevant saddles of p(β,ζ0)p(\beta,\zeta_{0}) in (29) can be determined from the downward flow of the contours ζ0±i0+\zeta_{0}\pm\textrm{i}0^{+}\in\mathbb{R} onto the Lefschetz thimbles (contours of steepest descent) passing through the saddles. The downward flow is defined as

ζ0s=(p(β,ζ0)ζ0),\frac{\partial\zeta_{0}}{\partial s}=-\bigg{(}\frac{\partial p(\beta,\zeta_{0})}{\partial\zeta_{0}}\bigg{)}^{*}, (30)

where ss is “flow time” and the the lines following the downward flow toward the saddle are the Lefschetz anti-thimbles. A visualization of this downward flow at one selected temperature is in figure 4.

Refer to caption
Figure 4: The downward flow (p(β,ζ0)/ζ0)-\big{(}\partial p(\beta,\zeta_{0})/\partial\zeta_{0}\big{)}^{*} of the pressure p(β,ζ0)p(\beta,\zeta_{0}) at a temperature T=1/β=0.65ΛlpT=1/\beta=0.65\,\Lambda_{\textsc{lp}}. Both axes are expressed in units of Λlp2\Lambda_{\textsc{lp}}^{2}. The line of integration of eNβVp(β,ζ0)e^{N\beta Vp(\beta,\zeta_{0})} over ζ0+i0+\zeta_{0}+\textrm{i}0^{+}\in\mathbb{R} is indicated in red. Note that it picks up a contribution only from the lower saddle indicated in red. The line of integration ζ0i0+\zeta_{0}-\textrm{i}0^{+}\in\mathbb{R} in blue corresponds to the integral of eNβVp(β,ζ0)e^{N\beta Vp(\beta,-\zeta_{0})}, and it picks up a contribution only from the upper saddle indicated in blue. The total partition function is a sum of both integrals, so it gets a contribution from both complex conjugate saddles for T>TcT>T_{\textsc{c}}.

For T<Tc0.616ΛlpT<T_{\textsc{c}}\approx 0.616\,\Lambda_{\textsc{lp}} there are two real saddles which are relevant for both contours ζ0±i0+\zeta_{0}\pm\textrm{i}0^{+}\in\mathbb{R}, and the saddle with the larger value of ζ0>0\zeta_{0}>0 gives the dominant pressure per component p(β)p(\beta) in (1). At zero temperature this saddle has ζ0=eΛlp2\zeta_{0}=e\Lambda_{\textsc{lp}}^{2}, which corresponds to a non-vanishing mass gap m=ζ0m=\sqrt{\zeta_{0}} for the scalar field ϕ\vec{\phi} in the vacuum. The saddle of the non-preferred phase is trivial at T=0T=0. When the scale Λlp\Lambda_{\textsc{lp}} is equated to the scale Λlp\Lambda_{\textsc{lp}} in the positive-coupling theory, these saddles correspond exactly to those for the positive-coupling theory and those found by Romatschke in [2], and the pressures of these saddles are the same in both theories. Above the critical temperature T>TcT>T_{\textsc{c}}, there are, just as in the positive-coupling theory, a complex conjugate pair of saddles. The second integral in (29), over the contour slightly below the branch cut (shown in figures 3 and 4), picks up a contribution only from the saddle with Im(ζ0)<0\operatorname{Im}(\zeta_{0})<0. Meanwhile, the first integral, on the contour above the branch cut in figure 4, picks up a contribution from the saddle with Im(ζ0)>0\operatorname{Im}(\zeta_{0})>0. The saddles have the same pressures p+(β)p_{+}(\beta), p(β)=p+(β)p_{-}(\beta)=p^{*}_{+}(\beta) as the complex conjugate pair of saddles in the positive-coupling theory. Just as in (7), keeping only the part of lnZ\ln Z_{-} that scales like NN gives a pressure per component p(β)p(\beta) in (1) that is equal to Rep+(β)\operatorname{Re}p_{+}(\beta).

The pressure per component p(β)p(\beta) as a function of temperature is plotted in figure 5. I discuss results in the next section.

III Results

The primary thing to observe from the preceding calculation of ZZ_{-} is the following: At large NN, when the scale Λlp\Lambda_{\textsc{lp}} appearing in both the positive- and negative-coupling ϕ4\vec{\phi}^{4} theories is considered to be the same scale, the thermal partition functions of both theories are equal, Z+=ZZ_{+}=Z_{-}, and the renormalized couplings λr(μ¯)\lambda_{\textsc{r}}(\bar{\mu}) and gr(μ¯)g_{\textsc{r}}(\bar{\mu}) are simply related by a minus sign. In other words, there exists a complex domain of path integration 𝒞\mathcal{C}_{-} for the gϕ4/N-g\vec{\phi}^{4}/N theory such that

Z=𝒞𝒟NϕeS[ϕ]=𝒟NϕeS+[ϕ]=Z+Z_{-}=\int_{\mathcal{C}_{-}}\mathcal{D}^{N}\vec{\phi}\,e^{-S_{-}[\vec{\phi}]}=\int\mathcal{D}^{N}\vec{\phi}\,e^{-S_{+}[\vec{\phi}]}=Z_{+} (31)

when the renormalized couplings in both theories are identified as λr(μ¯)=gr(μ¯)\lambda_{\textsc{r}}(\bar{\mu})=-g_{\textsc{r}}(\bar{\mu}). The two theories are only different in the range of choices for the MS¯\overline{\textrm{MS}} renormalization scale μ¯\bar{\mu}. For the λϕ4/N\lambda\vec{\phi}^{4}/N theory, μ¯\bar{\mu} is taken to be less than Λlp\Lambda_{\textsc{lp}}, whereas for the gϕ4/N-g\vec{\phi}^{4}/N theory, μ¯\bar{\mu} is taken to be greater than Λlp\Lambda_{\textsc{lp}}. However, the result is independent of the choice of μ¯\bar{\mu}, and therefore independent of the sign of the coupling. This indicates that the negative-coupling theory is simply the UV completion of the positive-coupling theory, and they are one and the same theory. The ϕ4\vec{\phi}^{4} theory at large NN has a negative bare coupling constant, as argued in [2, 4, 5], and the sign of the coupling changes under renormalization.

Note that the 𝒫𝒯\mathcal{P}\mathcal{T} symmetry of the domain 𝒞\mathcal{C}_{-} was important in ensuring this equivalence between partition functions, at least above the temperature TcT_{\textsc{c}}. (Namely, the 𝒫𝒯\mathcal{P}\mathcal{T} symmetry played a role in the Lefschetz thimble analysis for the integral (22) over the auxiliary field ζ0\zeta_{0}.)

Refer to caption
Figure 5: A plot of the pressure per component p(β)p(\beta) as a function of temperature T=1/βT=1/\beta for the gϕ4/N-g\vec{\phi}^{4}/N theory. The vacuum pressure is subtracted and p(β)p(\beta) is scaled by T4T^{4}. This agrees with the result of Romatschke in [2]. The dashed line indicates the Stefan–Boltzmann limit p(β)=π2T4/90p(\beta)=\pi^{2}T^{4}/90 for a free scalar boson. One can see that the pressure approaches the free theory limit at high temperatures, which is a consequence of asymptotic freedom.

The pressure per component p(β)p(\beta) of the ϕ4\vec{\phi}^{4} theory with negative bare coupling is plotted in figure 5 and is identical to that found by Romatschke in [2]. At T=TcT=T_{\textsc{c}} the volumetric heat capacity cV(T)=NT2p(β)/T2c_{V}(T)=NT\partial^{2}p(\beta)/\partial T^{2} is discontinuous and so there is a 2nd-order phase transition.

III.1 Non-triviality and asymptotic freedom

A notable feature of the pressure in figure 5 is that it approaches the Stefan–Boltzmann limit p(β)=π2T4/90p(\beta)=\pi^{2}T^{4}/90 for a free scalar boson at high temperatures. This is a characteristic feature of asymptotically free theories. This makes perfect sense from having a negative bare coupling λ0=g0<0\lambda_{0}=-g_{0}<0 as Λ0\Lambda_{0}\to\infty as seen in figure 1; the sign of the beta function gr/ln(μ¯)\partial g_{\textsc{r}}/\partial\ln(\bar{\mu}) is opposite the sign of the coupling as μ¯\bar{\mu}\to\infty and the coupling (26) approaches zero in the UV, so that the theory is almost free at high energies.

In doing the calculation of the partition function ZZ_{-} with negative coupling, the only trouble we ran into was that the scale μ¯\bar{\mu} could not be taken below the scale Λlp\Lambda_{\textsc{lp}} or else the assumption of negative coupling would be violated101010If the coupling were taken to be positive, as happens for μ¯<Λlp\bar{\mu}<\Lambda_{\textsc{lp}}, the path integral (8) on the domain 𝒞\mathcal{C}_{-} would be unbounded.. Thus the negative-coupling theory has a Landau pole, albeit a Landau pole111111Here, by Landau pole, I mean any place where the coupling diverges and becomes infinite. in the IR and not in the UV. A Landau pole in the IR, unlike in the positive-coupling case, does not prevent there from being an interacting continuum limit of the theory. Indeed, the negative-coupling ϕ4\vec{\phi}^{4} theory at large NN is not quantum trivial.

But we have seen here that the negative-coupling theory simply extends the positive-coupling theory to scales μ¯\bar{\mu} greater than that of the Landau pole Λlp\Lambda_{\textsc{lp}}, giving an equivalent partition function Z+=ZZ_{+}=Z_{-} at any scale μ¯\bar{\mu}. This means that an interacting continuum limit of the positive-coupling theory can be taken while keeping Λlp\Lambda_{\textsc{lp}} finite. The continuum limit μ¯>Λlp\bar{\mu}>\Lambda_{\textsc{lp}} is simply the negative-coupling theory. Therefore, since the negative-coupling theory is non-trivial, the standard ϕ4\vec{\phi}^{4} theory at large NN is non-trivial, as pointed out by Romatschke [5].

III.2 High temperatures T>TcT>T_{\textsc{c}}

One last result should be mentioned. The appearance of two saddles with complex-conjugate pressures p+(β),p+(β)p_{+}(\beta),p_{+}^{*}(\beta) for temperatures T>TcT>T_{\textsc{c}} above the critical temperature TcT_{\textsc{c}} is reminiscent of a phase of spontaneously broken 𝒫𝒯\mathcal{P}\mathcal{T} symmetry, which can occur in Hamiltonian systems when energy eigenvalues appear in complex-conjugate pairs [32], causing the partition function to become oscillatory as a function of temperature. In the formal large-NN limit a physical pressure per component p(β)p(\beta) in (1) can still be extracted, but the appearance of complex saddles signals an instability. This instability might affect real-time dynamics (for example, in calculations of viscosities or other transport coefficients along the lines of [33, 34, 35]), since a complex conjugate pair of masses m2m^{2} for the field ϕ\vec{\phi} leads to an unstable pole of the propagator in the region Im(ω)>0\operatorname{Im}(\omega)>0 when Wick-rotated into real time.

However, it is too soon to conclude that the theory is unstable or somehow sick at high temperatures (just as it was prematurely concluded in the 70s that the positive-coupling theory was “sick” at zero temperature [36, 37]). This is because there may be non-constant saddles ζ(x)\zeta(x) of the action for the auxiliary field (once ϕ\vec{\phi} is integrated out) that are preferred at high temperatures, so that this apparent 𝒫𝒯\mathcal{P}\mathcal{T}-broken phase is suppressed in the large-volume limit.

If this is the case, at leading order in large NN there is still a reason to expect that the partition functions Z+Z_{+} and ZZ_{-} of the positive- and negative-coupling theories will be identical after renormalization. This can be shown straightforwardly. The partition function Z+Z_{+} allowing for a non-constant gap parameter m(x)=iζ(x)m(x)=\sqrt{\textrm{i}\zeta(x)} is

Z+𝒟ζeNtrln(μ2+iζ)/2β,VNζ2/16λ=𝒟ζeNs[iζ],\begin{split}Z_{+}\propto&\int\mathcal{D}\zeta\,e^{-N\textrm{tr}\ln(-\partial_{\mu}^{2}+\textrm{i}\zeta)/2-\int_{\beta,V}N\zeta^{2}/16\lambda}\\ &=\int\mathcal{D}\zeta\,e^{-Ns[\textrm{i}\zeta]},\end{split} (32)

where β,V\int_{\beta,V} is shorthand for β,Vd4x\int_{\beta,V}\textrm{d}^{4}x and I’ve introduced the action per component

s[m2]=12trln(μ2+m2(x))+β,Vd4xm416λ.s[m^{2}]=\frac{1}{2}\textrm{tr}\ln\big{(}-\partial_{\mu}^{2}+m^{2}(x)\big{)}+\int_{\beta,V}\textrm{d}^{4}x\frac{m^{4}}{16\lambda}.

At leading order in large NN this will be replaced by a sum over the relevant saddles ζ(x)=im2(x)\zeta(x)=-\textrm{i}m^{2}(x) which satisfy the condition δs[m2]/δm2(x)=0\delta s[m^{2}]/\delta m^{2}(x)=0, i.e.

Z+njjeNs[mj2],Z_{+}\approx n_{j}\sum_{j}e^{-Ns[m^{2}_{j}]}, (33)

with mj2m^{2}_{j} denoting the jjth saddle, and njn_{j}, as in (28), denoting the intersection number of the domain of path integration with the Lefschetz anti-thimble passing through the jjth saddle. Meanwhile the partition function of the negative-coupling theory is given by

Z+,(𝒟ζeNtrln(μ2ζ)/2β,VNζ2/16g+ΔlnZ±[ζ])+,±𝒟ζeNs[ζ]+ΔlnZ±[ζ],\begin{split}Z_{-}\propto&\sum_{+,-}\bigg{(}\int\mathcal{D}\zeta\\ &e^{-N\textrm{tr}\ln(-\partial_{\mu}^{2}\mp\zeta)/2-\int_{\beta,V}N\zeta^{2}/16g+\Delta\ln Z_{-}^{\pm}[\zeta]}\bigg{)}\\ &\propto\sum_{+,-}\pm\int\mathcal{D}\zeta\,e^{-Ns[\mp\zeta]+\Delta\ln Z_{-}^{\pm}[\zeta]},\end{split} (34)

where there is a 𝒪(N0)\mathcal{O}(N^{0}) term ΔlnZ±[ζ]\Delta\ln Z_{-}^{\pm}[\zeta] which is given by

eΔlnZ±[ζ]±0±dχ0𝒟χeΔS±[χ,ζ]\begin{split}e^{\Delta\ln Z^{\pm}_{-}[\zeta]}\propto\pm\int_{0}^{\pm\infty}\textrm{d}\chi_{0}\int\mathcal{D}\chi^{\prime}\,\,e^{-\Delta S_{-}^{\pm}[\chi,\zeta]}\end{split} (35)

with

ΔS±[χ,ζ]=12trln(±iμ2+iζ(x))+β,Vd4x(±12i(μχ)2+i2ζχ2).\begin{split}\Delta S_{-}^{\pm}[\chi,\zeta]=&-\frac{1}{2}\textrm{tr}\ln\big{(}\pm\textrm{i}\partial_{\mu}^{2}+\textrm{i}\zeta(x)\big{)}\\ &+\int_{\beta,V}\textrm{d}^{4}x\,\bigg{(}\pm\frac{1}{2\textrm{i}}(\partial_{\mu}\chi)^{2}+\frac{\textrm{i}}{2}\zeta\chi^{2}\bigg{)}.\end{split} (36)

At leading order in large NN, this term does not contribute and the path integral can also be replaced by a sum over saddles of s[m2]s[m^{2}]:

Z+,nj±jeNs[mj2],Z_{-}\approx\sum_{+,-}n^{\pm}_{j}\sum_{j}e^{-Ns[m^{2}_{j}]}, (37)

with nj±n^{\pm}_{j} and mj2=ζjm^{2}_{j}=\mp\zeta_{j} defined similarly as in (28). In (34) and (37), the functional s[m2]s[m^{2}] is the same as for the positive-coupling theory, with λ\lambda replaced by g-g. Comparing (33) and (37), one can see that the partition functions at positive and negative coupling will be equivalent at large NN (when the scale Λlp\Lambda_{\textsc{lp}} is equivalent) as long as there are non-zero intersection numbers njn_{j} or nj±n^{\pm}_{j} for the same dominant saddles mj2m^{2}_{j} in both the positive- and negative-coupling theories.

If the non-constant saddles ζ(x)\zeta(x) dominate at high temperatures, however, I might expect that the 2nd-order phase transition would become a 1st-order phase transition. This should be investigated.

IV Conclusion

The ϕ4\vec{\phi}^{4} theory is just one recent example of large-NN theories with negative or complex bare couplings. Berges, Gurau, and Preis studied an O(N)3\textrm{O}(N)^{3} model with one coupling ig\textrm{i}g that is imaginary in the UV and possesses a Landau pole in the IR, which also exhibits asymptotic freedom [1]. Grable and Weiner studied a Gross–Neveu-like model of NN interacting fermions in 3+1D, and found that the coupling diverges at a scale Λc\Lambda_{\textsc{c}} and becomes negative in the UV [3]; like the scalar ϕ4\vec{\phi}^{4} and O(N)3\textrm{O}(N)^{3} theories, this fermionic theory is also asymptotically free at high energies. These models are interesting since calculations can be done analytically in the non-perturbative regime, and they realize non-trivial features like asymptotic freedom and phase transitions, and (at least in the case of the ϕ4\vec{\phi}^{4} model) bound states [2, 5].

Negative couplings can be meaningful and useful in a quantum field theory. The theory simply has to be path-integrated on some complex domain; the fields (e.g. ϕ\vec{\phi}) can no longer be real. However, 𝒫𝒯\mathcal{P}\mathcal{T} (or some antilinear) symmetry of the Lagrangian and of the asymptotic Stokes wedges in which the domain terminates appears to be important for ensuring a physical result for quantities like the partition function. The existence of 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theories with complex or negative couplings, like the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric gϕ4-g\phi^{4} theory, indicates that a coupling constant need not be positive in order to have a predictive interacting theory.

The implication of this present work is that the coupling constant of the ϕ4\vec{\phi}^{4} theory at large NN is only positive in the IR, while the bare coupling is negative. A negative bare coupling from the renormalization was already discussed in the 70s and 80s by Kobayashi and Kugo [38], by Bardeen and Moshe [39], and by Stevenson, Allès, and Tarrach [40, 41], but the theory explicitly at negative coupling was never studied. As mentioned before, a similar feature appears in the Lee model, which has a purely imaginary bare coupling [15, 16]. Moreover, it has been pointed out that under Wilsonian renormalization, the effective action of a non-Hermitian 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory can emerge from a Hermitian theory [42]. Thus a sign change of the coupling under renormalization is expected to be a feature of other theories, too.

Life simplifies at large NN and it may be true that the proper 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric gϕ4/N-g\vec{\phi}^{4}/N theory at large NN, for example defined on the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10), extends the positive-coupling (Hermitian) theory above the scale Λlp\Lambda_{\textsc{lp}}, as was first suggested by Romatschke in [2]. My use here of the domain 𝒞\mathcal{C}_{-} in (12) is motivated by the domain 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10) on which the path integral of the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theory could be defined.

Moreover, although the domain 𝒞\mathcal{C}_{-} considered in this paper does not correspond to a Hamiltonian theory with a 𝒞𝒫𝒯\mathcal{C}\mathcal{P}\mathcal{T} inner product of states, it still yields a partition function that agrees with that of the physical positive-coupling theory, and the theory on the domain 𝒞\mathcal{C}_{-} may still be predictive beyond thermodynamic quantities like the pressure p(β)p(\beta).

One might object to the presence of asymptotic freedom in large-NN scalar and fermionic theories and in 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric theories. Coleman and Gross in the 70s argued that the only asymptotically free theories are non-Abelian gauge theories. However, their argument assumed that scalar field theory with a negative coupling constant led to an unbounded potential [43]. It is no longer true that the potential is unbounded from below if the scalar fields live on an appropriate complex domain, such as 𝒞𝒫𝒯\mathcal{C}_{\mathcal{P}\mathcal{T}} in (10) or 𝒞\mathcal{C}_{-} in (12), or the 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric domain considered in [9, 24] for the gϕ4-g\phi^{4} theory. Thus scalar field theories can in fact be asymptotically free.

An objection can also be raised specifically about the ϕ4\vec{\phi}^{4} model at large NN. In the 70s, Coleman, Jackiw, and Politzer studied the O(N)\textrm{O}(N) model at large NN in 3+1D and found it possesses a tachyon, and on this basis they argued that the theory is “sick” [36]. However, it must be pointed out that this tachyon appears only in the dispreferred vacuum in which the field ϕ\vec{\phi} has no mass gap. This instability goes away when one realizes that the correct vacuum is the one where the field has a mass gap m=eΛlpm=\sqrt{e}\Lambda_{\textsc{lp}}, as pointed out by Abbott, Kang, and Schnitzer [37].

Here I have non-perturbatively calculated the partition function of the ϕ4\vec{\phi}^{4} theory at large NN assuming negative coupling from the start. With the proper choice for the complex domain 𝒞\mathcal{C}_{-} of path integration, involving 𝒫𝒯\mathcal{P}\mathcal{T} symmetry, the resulting thermodynamics are the same as those in the positive-coupling O(N)\textrm{O}(N) theory. The form of the running coupling in both theories indicates that the coupling, in fact, changes sign under renormalization, and that the negative-coupling theory simply extends the positive-coupling theory to scales above that of the Landau pole. Because of this, and because of other examples in large-NN and 𝒫𝒯\mathcal{P}\mathcal{T}-symmetric field theory, I point out that a negative bare coupling can be sensible. In addition, the ϕ4\vec{\phi}^{4} theory at large NN gives an example of a scalar field theory with asymptotic freedom and which is not quantum trivial.

A next step would be to see if thermodynamics in the ϕ4\vec{\phi}^{4} theory (and the issue of 𝒫𝒯\mathcal{P}\mathcal{T} symmetry breaking) is modified by the dominance of pressures associated with non-constant saddle configurations m2(x)m^{2}(x) of the auxiliary field. It would also be fruitful to calculate viscosities and other transport coefficients in the ϕ4\vec{\phi}^{4} theory in 3+1D along the lines of [33, 34, 35]. One way to do this involves the use of Rnn resummation methods [28, 27]. Lastly, it would be interesting to see if the results of this work continue at sub-leading order in the large-NN expansion.

Acknowledgements.
I would like to thank Paul Romatschke for many useful discussions. In addition, I would like to thank Seth Grable, Chun-Wei Su, and Sebastian Vazquez-Carson, as well as the heavy-ion theory group at CERN, including Urs Wiedemann and Jasmine Brewer, for conversations and comments on this subject. This work was supported by the Department of Energy, DOE award No. DE-SC0017905.

References