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

𝒫𝒯\mathcal{PT} Symmetry and Renormalisation in Quantum Field Theory

Carl M Bender1    Alexander Felski2    S P Klevansky3    and Sarben Sarkar4 1Department of Physics, Washington University, St Louis, Missouri 63130, USA cmb@wustl.edu 2Institut für Theoretische Physik, Universität Heidelberg, 69120 Heidelberg, Germany felski@thphys.uni-heidelberg.de 3Institut für Theoretische Physik, Universität Heidelberg, 69120 Heidelberg, Germany spk@physik.uni-heidelberg.de 4Department of Physics, King’s College London, London WC2R 2LS, UK sarben.sarkar@kcl.ac.uk
Abstract

Quantum systems governed by non-Hermitian Hamiltonians with 𝒫𝒯\mathcal{PT} symmetry are special in having real energy eigenvalues bounded below and unitary time evolution. We argue that 𝒫𝒯\mathcal{PT} symmetry may also be important and present at the level of Hermitian quantum field theories because of the process of renormalisation. In some quantum field theories renormalisation leads to 𝒫𝒯\mathcal{PT}-symmetric effective Lagrangians. We show how 𝒫𝒯\mathcal{PT} symmetry may allow interpretations that evade ghosts and instabilities present in an interpretation of the theory within a Hermitian framework. From the study of examples 𝒫𝒯\mathcal{PT}-symmetric interpretation is naturally built into a path integral formulation of quantum field theory; there is no requirement to calculate explicitly the 𝒫𝒯\mathcal{PT} norm that occurs in Hamiltonian quantum theory. We discuss examples where 𝒫𝒯\mathcal{PT}-symmetric field theories emerge from Hermitian field theories due to effects of renormalization. We also consider the effects of renormalization on field theories that are non-Hermitian but 𝒫𝒯\mathcal{PT}-symmetric from the start.

1 Introduction

Non-Hermitian Hamiltonians govern systems that in general receive energy from and/or dissipate energy into their environment and so they are typically not in equilibrium. Their energy is not conserved and their energy levels are complex. However, in 1998 [1] non-Hermitian 𝒫𝒯\mathcal{PT} systems were shown to offer new possibilities for unitary time evolution in quantum mechanics. Since then there has been extensive research activity [2], particularly in material science and optics, which has implemented the ideas of 𝒫𝒯\mathcal{PT}-symmetric quantum mechanics.

𝒫𝒯\mathcal{PT}-symmetric quantum mechanics can be considered to be a one-dimensional quantum field theory. To date there have been no clear examples of 𝒫𝒯\mathcal{PT}-symmetric quantum field theory in higher dimensions. We examine the possibility that 𝒫𝒯\mathcal{PT} symmetry may emerge when considering the effect of quantum fluctuations in a higher-dimensional field theory. We also examine the converse process where we consider the effect of quantum fluctuations on an initially 𝒫𝒯\mathcal{PT}-symmetric quantum field theory.

Using numerical techniques the work of Bender and Boettcher [1] showed that quantum-mechanical Hamiltonians of the form

H=12p2+x2(ix)ϵH=\frac{1}{2}p^{2}+x^{2}\left(ix\right)^{\epsilon} (1)

have a real positive spectrum. Using a correspondence between ordinary differential equations and integrable field theory models, Dorey et al. [3] provided an analytical proof of the result of Bender and Boettcher.

Mostafazadeh [4] considered the framework of pseudo-Hermiticity which contained 𝒫𝒯\mathcal{PT} symmetry as a special case. A Hamiltonian HH is pseudo-Hermitian if HH is not selfadjoint but

H=ηHη1,H^{\dagger}=\eta H\eta^{-1}, (2)

where η\eta is a positive-definite Hermitian operator. In terms of η\eta the Hilbert space of states has a positive-definite inner product given by

φ,ηχ\left\langle\varphi,\eta\chi\right\rangle (3)

where ,\left\langle,\right\rangle is a conventional inner product on the Hilbert space of states for the Hermitian part of the Hamiltonian. There is not a universal expression for η\eta, which is difficult to calculate in general. This would seem to imply an impediment to calculating correlation functions in 𝒫𝒯\mathcal{PT}-symmetric field theories. Within the context of quantum mechanics, formulated in terms of functional integrals, Jones and Rivers [5] showed through examples that it was not necessary to calculate η\eta. Hence, we shall take the path-integral representation as the definition of a quantum field theory and, except for the Lee model, not consider η\eta explicitly. It is a convenient formulation for both perturbative and nonperturbative calculations.

We consider four examples of field theories to illustrate three different aspects of 𝒫𝒯\mathcal{PT} symmetry:

  1. 1.

    the Lee model and the elimination of ghosts [6];

  2. 2.

    the role of the top quark and the Higgs instability [7] in the standard model of particle physics [8];

  3. 3.

    the new epsilon expansion for 𝒫𝒯\mathcal{PT}-symmetric scalar field theory [9, 10];

  4. 4.

    the functional renormalisation group and the preservation of 𝒫𝒯\mathcal{PT} symmetry in the infrared after renormalisation [11].

2 Role of 𝒫𝒯\mathcal{PT} symmetry in the Lee model

Hermitian quantum field theory (QFT) has been studied for many decades both perturbatively and nonperturbatively. QFT is considered to be a viable and successful description of quantum electrodynamics, the weak, and the strong interactions. The use of dispersion relations in QFT, provides a way to bypass some of the limitations of perturbation theory and was developed decades ago (starting with the work of Lehmann, Symanzik, and Zimmerman [12, 13]). The theory of dispersion relations is based on one particularly important assumption: There exists an interpolating (i.e. a renormalised) field ϕ(x)\phi(x), where ϕ(x)\phi(x) denotes an arbitrary field of the theory, such that we have the weak convergence relation

limt±α|ϕ(x)ϕout,in(x)|β=0,\lim_{t\to\pm\infty}\matrixelement{\alpha}{\phi(x)-\phi_{\rm out,\,in}(x)}{\beta}=0, (4)

where |α\ket{\alpha} and |β\ket{\beta} are experimentally accessible states and ϕout\phi_{\rm out} and ϕin\phi_{\rm in} are free field operators associated with noninteracting particles. The Green’s functions in dispersion relations are assumed to have been made finite by some form of renormalisation (which need not be specified). In this framework it is possible, within the context of any specified polynomial field-theory Lagrangian, to establish coupled integral equations among 1-particle irreducible Green’s functions. For example, in a Yukawa theory of a pseudoscalar field interacting with fermions, 2-point functions are coupled with the 3-point (vertex) function. One conclusion is that such a theory may contain ghosts unless the vertex function vanishes as the momentum transfer tends to infinity [13]. So it is quite possible that a Hermitian quantum field theory may contain ghosts. The Lee model [14], which we discuss next, is just such an example where the vertex function and in particular coupling-constant renormalisation is related to such unusual behaviour. The role of 𝒫𝒯\mathcal{PT} symmetry is crucial in banishing the ghosts and making the field theory viable [6].

2.1 Lee model

The Lee model is very different from the Standard Model of particle physics but it has the advantage of being soluble and so provides a testing ground for new methods in field theory. Indeed, the model was devised as a field theory whose coupling-constant and wave-function renormalisation could be computed exactly in principle. One variant of the model contains fields for infinitely heavy spinless fermions, which have two internal states VV and NN, as well as a neutral scalar field θ\theta. Antifermions are not in the theory and so there is no crossing symmetry. The permitted reaction channel is

VN+θ,V\rightleftharpoons N+\theta, (5)

but the channel

NV+θN\rightleftharpoons V+\theta (6)

is not allowed.

In momentum space the Hamiltonian of the Lee model (in a space of finite volume 𝒱\mathcal{V}) is

H=H0+Hint,H=H_{0}+H_{\rm int}, (7)

where

H0=mVψ¯VψV+mNψ¯NψN+𝐤ω𝐤a𝐤a𝐤H_{0}=m_{V}\overline{\psi}_{V}{\psi}_{V}+m_{N}\overline{\psi}_{N}{\psi}_{N}+\sum_{\bf k}\omega_{\bf k}a^{{\dagger}}_{\bf k}a_{\bf k} (8)

and

Hint=δmVψ¯VψVg0𝒱12𝐤u(ω𝐤)(2ω𝐤)12(ψ¯VψNa𝐤+ψ¯NψVa𝐤).H_{\rm int}=\delta m_{V}\overline{\psi}_{V}{\psi}_{V}-g_{0}{\mathcal{V}}^{-\frac{1}{2}}\sum_{\bf k}\frac{u(\omega_{\bf k})}{{{(2\omega_{\bf k})}}^{\frac{1}{2}}}(\overline{\psi}_{V}{\psi}_{N}a_{\bf k}+\overline{\psi}_{N}{\psi}_{V}a^{{\dagger}}_{\bf k}). (9)

u(ω𝐤)u(\omega_{\bf k}) is a dimensional cut-off function which is chosen to tend to 0 for large ω𝐤\omega_{\bf k} where ω𝐤=𝐤2+μ2\omega_{\bf k}=\sqrt{{\bf k}^{2}+\mu^{2}}. Standard commutation and anti-commutation rules are assumed:

{ψ¯V,ψV}\displaystyle\{\overline{\psi}_{V},{\psi}_{V}\} ={ψ¯N,ψN}=1\displaystyle=\{\overline{\psi}_{N},{\psi}_{N}\}=1 (10)
[a𝐤,a𝐤]\displaystyle\left[a_{\bf k},a^{{\dagger}}_{\bf k^{\prime}}\right] =δ𝐤,𝐤.\displaystyle=\delta_{\bf k,\bf k^{\prime}}. (11)

All other commutators and anticommutators vanish. Bare operators and coupling constant appear in HH. However mV,mN,and,μm_{V},\,m_{N},\,{\rm and},\,\mu are renormalised parameters and so are determined by experiment; δmV\delta m_{V} is the mass counterterm for the VV particle and is a function of g0g_{0}. No further mass renormalisation is actually needed. The coupling g0g_{0} in principle is determined from the scattering cross-section of NN and θ\theta although there are interesting complications of non-Hermiticity if the coupling is not small.

However, let us for the moment persist in the view that g0g_{0} is real and so HH is Hermitian. Let us choose as a basis for the Hilbert space states of the form

|=|nV,nN,{n𝐤}.\ket{}=\ket{n_{V},n_{N},\{{n_{\bf{k}}}\}}. (12)

From HH it is clear that there are two conserved quantities BB and QQ:

B=nV+nN,Q=Nθ+nN,B=n_{V}+n_{N},\quad Q=-N_{\theta}+n_{N}, (13)

where Nθ=𝐤n𝐤N_{\theta}=\sum_{\bf k}n_{\bf k}. Hence the Hilbert space is partitioned into an infinite number of independent sectors with fixed BB and QQ. Although this makes the model soluble, the analysis of sectors with large BB and QQ is complicated. Since our main point is to show how renormalisation can lead to a non-Hermitian Hamiltonian, we simplify our analysis further by considering (i) a nontrivial sector B=1B=1 and Q=0Q=0 and (ii) modifying the model so that there is no 𝐤\bf k dependence. This simplified model is a quantum-mechanical one since the quantum fields in (8) and (9) are replaced by quantum operators. Hence, infinities that arise from summing over an infinite set of modes in quantum field theory are absent but some features of coupling-constant renormalisation persist. The resulting Hamiltonian is =0+1\mathcal{H}={\cal{H}}_{0}+{\cal{H}}_{1}, where

0=mVVV+mNNN+μaa{\cal{H}}_{0}={m_{V}}V^{{\dagger}}V+m_{N}N^{{\dagger}}N+\mu a^{{\dagger}}a (14)

and

1=g0(VNa+aNV)+δmVVV.{\cal{H}}_{1}=g_{0}(V^{{\dagger}}Na+a^{{\dagger}}N^{{\dagger}}V)+\delta m_{V}V^{{\dagger}}V. (15)

We will look at the 2-dimensional Hilbert space associated with this sector and consider the eigenstates, which we will denote by

|V\displaystyle\ket{V} =c11|1,0,0+c12|0,1,1,\displaystyle=c_{11}\ket{1,0,0}+c_{12}\ket{0,1,1}, (16)
|Nθ\displaystyle\ket{N\theta} =c21|1,0,0+c22|0,1,1,\displaystyle=c_{21}\ket{1,0,0}+c_{22}\ket{0,1,1}, (17)

with eigenvalues mVm_{V} and ENθE_{N\theta}. Let us denote (mV+δmV)(m_{V}+\delta m_{V}) by mV0m_{V_{0}}, the bare mass of VV. The eigenvalues can be shown to satisfy the equations

mV=\displaystyle m_{V}= 12(mN+mθ+mV0M02+4g02),\displaystyle\frac{1}{2}(m_{N}+m_{\theta}+m_{V_{0}}-\sqrt{M_{0}^{2}+4g_{0}^{2}}), (18)
ENθ=\displaystyle E_{N\theta}= 12(mN+mθ+mV0+M02+4g02),\displaystyle\frac{1}{2}(m_{N}+m_{\theta}+m_{V_{0}}+\sqrt{M_{0}^{2}+4g_{0}^{2}}), (19)

where M0mN+mθmV0.M_{0}\equiv m_{N}+m_{\theta}-m_{V_{0}}. Following field theory, we define the wave-function renormalisation constant ZVZ_{V} through the relation

1=0|1ZVV|V,1=\bra{0}\frac{1}{\sqrt{Z_{V}}}V\ket{V}, (20)

which gives

ZV=2g02M02+4g02(M02+4g02M0),Z_{V}=\frac{2g_{0}^{2}}{\sqrt{M_{0}^{2}+4g_{0}^{2}}\,(\sqrt{M_{0}^{2}+4g_{0}^{2}}-M_{0})}, (21)

and the coupling constant renormalisation through

g2g02=ZV.\frac{g^{2}}{g_{0}^{2}}=Z_{V}. (22)

We deduce that

g02=g21g2M2,g_{0}^{2}=\frac{g^{2}}{1-\frac{g^{2}}{M^{2}}}, (23)

where MmN+mθmVM\equiv m_{N}+m_{\theta}-m_{V}. If gg, the experimentally determined value, exceeds MM then g0g_{0} becomes pure imaginary and the Hamiltonian becomes non-Hermitian. Although the Hamiltonian has become non-Hermitian, it is 𝒫𝒯\mathcal{PT}-symmetric. Explicitly, the transformations due to 𝒫\mathcal{P} are [6]

𝒫V𝒫\displaystyle\mathcal{P}V\mathcal{P} =V,\displaystyle=-V, 𝒫N𝒫\displaystyle\mathcal{P}N\mathcal{P} =N,\displaystyle=-N, 𝒫a𝒫\displaystyle\mathcal{P}a\mathcal{P} =a,\displaystyle=-a, (24)
𝒫V𝒫\displaystyle\mathcal{P}V^{{\dagger}}\mathcal{P} =V,\displaystyle=-V^{{\dagger}}, 𝒫N𝒫\displaystyle\mathcal{P}N^{{\dagger}}\mathcal{P} =N,\displaystyle=-N^{{\dagger}}, 𝒫a𝒫\displaystyle\mathcal{P}a^{{\dagger}}\mathcal{P} =a.\displaystyle=-a^{{\dagger}}. (25)

The transformations due to 𝒯\mathcal{T} are

𝒯V𝒯\displaystyle\mathcal{T}V\mathcal{T} =V,\displaystyle=V, 𝒯N𝒯\displaystyle\mathcal{T}N\mathcal{T} =N,\displaystyle=N, 𝒯a𝒯\displaystyle\mathcal{T}a\mathcal{T} =a\displaystyle=a (26)
𝒯V𝒯\displaystyle\mathcal{T}V^{{\dagger}}\mathcal{T} =V\displaystyle=V^{{\dagger}} 𝒯N𝒯\displaystyle\mathcal{T}N^{{\dagger}}\mathcal{T} =N\displaystyle=N^{{\dagger}} 𝒯a𝒯\displaystyle\mathcal{T}a^{{\dagger}}\mathcal{T} =a.\displaystyle=a^{{\dagger}}. (27)

It is now straightforward to check that i|g0|(VNa+aNV)i|g_{0}|(V^{{\dagger}}Na+a^{{\dagger}}N^{{\dagger}}V) is 𝒫𝒯\mathcal{PT}-symmetric. This is our first example of a 𝒫𝒯\mathcal{PT}-symmetric field theory that arises from a Hermitian field theory after renormalisation. On introducing a 𝒫𝒯\mathcal{PT}-symmetric inner product in the ghost regime of the Lee model, the so-called ghost state (identified within the Hermitian frame work) turns out to have a positive norm[6]. The Lee model then can be interpreted as an acceptable quantum field theory.

3 Higgs instability and 𝒫𝒯\mathcal{PT} symmetry

The conventional approach to renormalisation involves the regularisation of loop integrals in Feynman diagrams. This led to scale dependence of the parameters of the theory which can be understood in a different way using the approach [15] of Wilson to renormalisation. Wilson’s approach leads naturally to the concept of effective actions, and the form of the effective actions for some theories in high-energy physics is non-Hermitian and 𝒫𝒯\mathcal{PT}-symmetric.

The change from the conventional to the Wilson approach can be illustrated by considering a scalar field Φ\Phi whose (conventional) action is given by

SΛ[Φ]=dDx(12μΦμΦ+UΛ(Φ))S_{\Lambda}\left[\Phi\right]=\int d^{D}x(\frac{1}{2}\partial_{\mu}\Phi\partial^{\mu}\Phi+U_{\Lambda}(\Phi)) (28)

where the UV cut-off is Λ\Lambda. If we integrate over Fourier modes with momenta of magnitude pp in the shell kpΛk\leq p\leq\Lambda we arrive at an action which we will denote by Sk[ϕ]S_{k}\left[\phi\right] where ϕ\phi has Fourier modes with pkp\leq k (a sharp infrared cut-off).111Wetterich [16] introduced a smooth infrared cut-off function Rk(p2)R_{k}(p^{2}) rather than a sharp cut-off so that Sk[ϕ]=dDx[12μϕμ+Vk(ϕ)]+Δk[ϕ]S_{k}\left[\phi\right]=\int d^{D}x[\frac{1}{2}\partial_{\mu}\phi\partial^{\mu}+V_{k}(\phi)]+\Delta_{k}[\phi], where Δk[ϕ]=12dDpϕpRk(p2)ϕp\Delta_{k}[\phi]=\frac{1}{2}\int d^{D}p\phi_{p}R_{k}(p^{2})\phi_{-p}. A common choice for Rk(p2)R_{k}(p^{2}) is Rk(p2)=(k2p2)Θ(k2p2)R_{k}(p^{2})=(k^{2}-p^{2})\Theta(k^{2}-p^{2}). The associated partition function is

Zk[j]exp(Wk[j])Dϕexp(Sk[ϕ]pjpϕp)Z_{k}[j]\equiv\exp(-W_{k}[j])\equiv\int D\phi\exp(-S_{k}[\phi]-\int_{p}j_{p}\phi_{-p}) (29)

and a Legendre transformation on Wk[j]W_{k}[j] leads to the effective action Γk[ϕc]\Gamma_{k}[\phi_{c}]

Γk[ϕc]=Wk[j]dDxjϕcΔk[ϕc].\Gamma_{k}[\phi_{c}]=W_{k}\left[j\right]-\int d^{D}xj\phi_{c}-\Delta_{k}[\phi_{c}]. (30)

Within systematic (derivative) approximation schemes [16, 17] the effective action can be represented (at the lowest level of approximation) by the ansatz

Γk[ϕc]=dDx(12μϕcμϕc+Uk(ϕc)).\Gamma_{k}[\phi_{c}]=\int d^{D}x(\frac{1}{2}\partial_{\mu}\phi_{c}\partial^{\mu}\phi_{c}+U_{k}(\phi_{c})). (31)

The effective potential, Uk(ϕc)U_{k}(\phi_{c}) which results from renormalisation, will show the emergence of 𝒫𝒯\mathcal{PT} symmetry in the next models that we will discuss. This ansatz represents quite a severe approximation since, for an xx-independent ϕc\phi_{c}, we have a one-dimensional approximation to an infinite-dimensional field.

We consider two theories of current interest using this formalism:

  1. 1.

    a theory of dynamical breaking of gravity via a graviton condensate field [18] φc\varphi_{c} with U(φc)=φc4log(iφc)U(\varphi_{c})=-{\varphi_{c}}^{4}\log(i\varphi_{c}) for large φc\varphi_{c}r,

    and

  2. 2.

    the Standard Model of particle physics for which φc\varphi_{c} is the Higgs field and U(φc)=φc4log(φc2)U(\varphi_{c})=-{\varphi_{c}}^{4}\log(\varphi_{c}^{2}) for large φc\varphi_{c} [7].

Both examples, in conventional quantum mechanics, would show unstable behaviour for large φc\varphi_{c}. Under 𝒫\mathcal{P} we have φcφc\varphi_{c}\rightarrow{-\varphi_{c}} and under 𝒯\mathcal{T} we have iii\rightarrow{-i}. So U(φc)U(\varphi_{c}) is 𝒫𝒯\mathcal{PT} symmetric in both cases.

We study such effective potentials by considering three related quantum mechanical Hamiltonians HiH_{i}, i=1,2,3i=1,2,3:

H1=\displaystyle H_{1}= p2+x4log(ix),\displaystyle p^{2}+{x^{4}}\log(ix), (32)
H2=\displaystyle H_{2}= p2x4log(ix),\displaystyle p^{2}-{x^{4}}\log(ix), (33)
H3=\displaystyle H_{3}= p2x4log(x2),\displaystyle p^{2}-x^{4}\log(x^{2}), (34)

which are non-Hermitian but 𝒫𝒯\mathcal{PT}-symmetric. We will find that the the first two Hamiltonians will show unbroken 𝒫𝒯\mathcal{PT} symmetry and the Hamiltonian H3H_{3} will show broken 𝒫𝒯\mathcal{PT} symmetry.

3.1 The spectra for H1H_{1}

We use the WKB method of semiclassical quantum mechanics to determine the energy spectrum of H1H_{1}:

  • Locate turning points.

  • Examine the complex classical trajectories on an infinite-sheeted Riemann surface.

  • Determine the open or closed nature of the trajectories.

The turning points satisfy the equation

E=x4log(ix).E=x^{4}\log(ix). (35)

We take E=1.24909E=1.24909 because this is the numerical value of the ground-state energy obtained separately by solving the Schrödinger equation. (See the table below.)

One turning point lies on the negative imaginary-xx axis. To find this point we set x=irx=-ir (r>0r>0) and obtain the algebraic equation E=r4logrE=r^{4}\log r. Solving this equation by using Newton’s method, we find that the turning point lies at x=1.39316ix=-1.39316i. To find the other turning points we seek solutions to (35) in polar form x=reiθx=re^{i\theta} (r>0,θrealr>0,\,\theta\,{\rm real}). Substituting for xx in (35) and taking the imaginary part, we obtain

logr=(2kπ+θ+π/2)cos(4θ)/sin(4θ),\log r=-(2k\pi+\theta+\pi/2)\cos(4\theta)/\sin(4\theta), (36)

where kk is the sheet number in the Riemann surface of the logarithm. (We choose the branch cut to lie on the positive-imaginary axis.) Using (36), we simplify the real part of (35) to

E=r4(2kπ+θ+π/2)/sin(4θ).E=-r^{4}(2k\pi+\theta+\pi/2)/\sin(4\theta). (37)

We then use (36) to eliminate rr from (37) and use Newton’s method to determine θ\theta. For k=0k=0 and E=1.24909E=1.24909, two 𝒫𝒯\mathcal{PT}-symmetric (left-right symmetric) pairs of turning points lie at ±0.938030.38530i\pm 0.93803-0.38530i and at ±0.32807+0.75353i\pm 0.32807+0.75353i. For k=1k=1 and E=1.24909E=1.24909 there is a turning point at 0.53838+0.23100i-0.53838+0.23100i; the 𝒫𝒯\mathcal{PT}-symmetric image of this turning point lies on sheet k=1k=-1 at 0.53838+0.23100i0.53838+0.23100i.

The turning points determine the shape of the classical trajectories. Two topologically different kinds of classical paths are shown in Figs. 1 and 2. All classical trajectories are closed and left-right symmetric, and this implies that the quantum energies are all real [19].

The WKB quantization condition is a complex path integral on the principal sheet of the logarithm (k=0k=0). On this sheet a branch cut runs from the origin to +i+i\infty on the imaginary axis; this choice of branch cut respects the 𝒫𝒯\mathcal{PT} symmetry of the configuration. The integration path goes from the left turning point xLx_{{\rm L}} to the right turning point xRx_{{\rm R}} [20]:

(n+12)πxLxR𝑑xEV(x)(n>>1).\left(n+\frac{1}{2}\right)\pi\sim\int_{x_{{\rm L}}}^{x_{{\rm R}}}dx\sqrt{E-V(x)}\quad(n>>1). (38)
Refer to caption
Figure 1: Three nested closed classical paths.
Refer to caption
Figure 2: Closed classical path for energy E=1.24909E=1.24909.

If the energy is large (En1)\left(E_{n}\gg 1\right), then from (36) with k=0k=0 we find that the turning points lie slightly below the real axis at xR=reiθx_{\rm R}=re^{i\theta} and at xL=reπiiθx_{{\rm L}}=re^{-\pi i-i\theta} with

θπ/(8logr)andr4logrE.\theta\sim-\pi/(8\log r)\quad{\rm and}\quad r^{4}\log r\sim E. (39)

We choose the path of integration in (38) to have a constant imaginary part so that the path is a horizontal line from xLx_{{\rm L}} to xRx_{{\rm R}}. Since EE is large, rr is large and thus θ\theta is small. We obtain the simplified approximate quantization condition

(n+12)πr3logr11𝑑t1t4,\left(n+\frac{1}{2}\right)\pi\sim r^{3}\log r\int_{-1}^{1}dt\sqrt{1-t^{4}}, (40)

which leads to the WKB approximation for n1n\gg 1:

En[log(En)]1/3[Γ(7/4)(n+1/2)πΓ(5/4)2]4/3.\frac{E_{n}}{[\log(E_{n})]^{1/3}}\sim\left[\frac{\Gamma(7/4)(n+1/2)\sqrt{\pi}}{\Gamma(5/4)\sqrt{2}}\right]^{4/3}. (41)
Calculation of values of energy
Energy level nn EnE_{n} En[log(En)]1/3\frac{E_{n}}{[\log(E_{n})]^{1/3}} WKB % error
0 1.24909 2.06161 0.54627 73.5028
3 13.7383 9.96525 7.31480 26.5969
6 31.6658 20.9458 16.6979 20.2804
9 52.9939 33.4674 27.6956 17.2463
12 76.9748 47.1776 39.9324 15.3573
Refer to caption
Figure 3: Three nested classical trajectories for H2H_{2} with E=2.07734E=2.07734.
Refer to caption
Figure 4: Complex classical trajectory for H2H_{2} with E=2.07734E=2.07734.

Analysis of the supergravity model Hamiltonian H2H_{2}: The classical trajectories for the Hamiltonian H2H_{2} are plotted in Figs. 3 and 4. Like the classical trajectories for the Hamiltonian H1H_{1}), these trajectories are closed, which implies that all the eigenvalues for H2H_{2} are real.

Analysis of the Higgs model Hamiltonian H3H_{3}: To make sense of H3H_{3} we again introduce a parameter ϵ\epsilon and we define H3H_{3} as the limit of H=p2+x2(ix)ϵlog(x2)H=p^{2}+x^{2}(ix)^{\epsilon}\log\left(x^{2}\right) as ϵ: 02\epsilon:\,0\to 2. This case is distinctly different from that for H2H_{2}. Figure 5 shows that the 𝒫𝒯\mathcal{PT} symmetry is broken for all ϵ0\epsilon\neq 0. When ϵ=2\epsilon=2, there are only four real eigenvalues: E0=1.1054311E_{0}=1.1054311, E1=4.577736E_{1}=4.577736, E2=10.318036E_{2}=10.318036, and E3=16.06707E_{3}=16.06707. To confirm this result we plot a classical trajectory for ϵ=2\epsilon=2 in Fig. 6. In contrast with Fig. 4, the trajectory is open and not left-right symmetric.

Refer to caption
Figure 5: Energies of the Hamiltonian H=p2+x2(ix)ϵlog(ix)H=p^{2}+x^{2}(ix)^{\epsilon}\log(ix) plotted versus ϵ\epsilon.
Refer to caption
Figure 6: Classical path for the Hamiltonian H=p2x4log(x2)H=p^{2}-x^{4}\log\left(x^{2}\right).

4 Renormalisation and canonical scalar field theory

We have shown examples of 𝒫𝒯\mathcal{PT}-symmetric quantum field theories arising from the process of renormalisation in field theories. The properties of 𝒫𝒯\mathcal{PT}-symmetric field theories have yet to be established. In quantum mechanics much progress in understanding 𝒫𝒯\mathcal{PT} symmetry has been made through the study of the Hamiltonian in (1). A natural generalisation to an Euclidean field theory Hamiltonian in DD-dimensions is to consider the Lagrangian

L=12(ϕ)2+12μ2φ2+12gμ02ϕ2(iμ0δ2ϕ)ϵL=\frac{1}{2}(\nabla\phi)^{2}+\frac{1}{2}\mu^{2}\varphi^{2}+\frac{1}{2}g\mu_{0}^{2}\phi^{2}(i\mu_{0}^{\frac{\delta}{2}}\phi)^{\epsilon} (42)

where ϕ\phi is a dimensional pseudoscalar field, δ=2D\delta=2-D and ϵ0\epsilon\geq 0. It is natural to study this class of field theories since we can then compare our findings in certain limits with those found in quantum mechanical systems. For noninteger ϵ\epsilon the interaction is nonpolynomial and the methods that are normally used for calculating Greens functions in quantum field theory do not apply. The method we will use to set up a quantum field theory involves

  1. 1.

    rewriting of the interaction in terms of a formal series of polynomial interactions;

  2. 2.

    creating modified Feynman rules in a nonperturbative expansion;

  3. 3.

    consideration of infinite contributions and renormalisation.

The first part of the method, used in (i), was introduced decades ago and has been recently developed in (ii) and (iii) within the context of 𝒫𝒯\mathcal{PT}-symmetric field theory. The procedure of (i) allows us to use Wick’s theorem and the linked-cluster theorem in conjunction with (ii).

In a conventional field theory described by a Lagrangian \mathcal{L} within polynomial interactions in a field φ\varphi, we calculate connected Green’s functions using a partition functional Z(J)Z(J) defined through a path integral as

Z(J)=𝒟φexp(dDx(+Jφ)),Z(J)=\int{\mathcal{D}}\varphi\exp(-\int d^{D}x({\mathcal{L}}+J\varphi)), (43)

where JJ is a source. It is well known that the normalised partition functional Z(J)Z(0)\frac{Z(J)}{Z(0)} satisfies

lnZ(J)Z(0)=exp(Connectedsourcetosourcediagrams).\ln\frac{Z(J)}{Z(0)}=\exp(\sum{\rm Connected\,source-to-source\,diagrams}). (44)

This result simplifies our calculations in (ii) and (iii). First, we note that formally, on writing ψ=μ0δ2φ\psi=\mu_{0}^{\frac{\delta}{2}}\varphi (and suppressing the argument of ψ\psi)

ψ2(iψ)ϵ=n=0ϵnn!r=0n(nr)(iπ|ψ|ψ)rψ2(12lnψ2)nr=n=1ϵnn+ψ2\psi^{2}(i\psi)^{\epsilon}=\sum_{n=0}^{\infty}\frac{\epsilon^{n}}{n!}\sum_{r=0}^{n}\left(\begin{array}[]{c}n\\ r\end{array}\right)\Big{(}\frac{i\pi|\psi|}{\psi}\Big{)}^{r}\psi^{2}\Big{(}\frac{1}{2}\ln\psi^{2}\Big{)}^{n-r}=\sum_{n=1}^{\infty}\epsilon^{n}\mathcal{L}_{n}+\psi^{2} (45)

because

ln(iψ)=12iπ|ψ|/ψ+12lnψ2.\ln(i\psi)=\frac{1}{2}i\pi|\psi|/\psi+\frac{1}{2}\ln\psi^{2}. (46)

The series in (45) can be expressed in terms of powers of ψ\psi on noting that

  1. 1.
    (|ψ|ψ)r=w=0wψ(2w+1)r,\Big{(}\frac{|\psi|}{\psi}\Big{)}^{r}=\sum_{w=0}^{\infty}\mathcal{I}_{w}\psi^{(2w+1)r}, (47)

    where w=2π0𝑑t(t2)w(2w+1)!\mathcal{I}_{w}=\frac{2}{\pi}\int_{0}^{\infty}dt\frac{(-t^{2})^{w}}{(2w+1)!}.

  2. 2.
    (lnψ2)s=N,s0ψ2N,(\ln\psi^{2})^{s}=\mathcal{F}^{0}_{N,s}\psi^{2N}, (48)

    where N,s0=limN0(ddN)s\mathcal{F}^{0}_{N,s}=\lim_{N\to 0}(\frac{d}{dN})^{s}.

Thus, we can rewrite the partition function in (43) as

Z[J]=𝒟φexp(dDx(n=0ϵnn+Jφ)),Z[J]=\int{\mathcal{D}}\varphi\exp(-\int d^{D}x(\sum_{n=0}^{\infty}\epsilon^{n}\mathcal{L}_{n}+J\varphi)), (49)

where 0\mathcal{L}_{0} is the free (that is, ϵ\epsilon-independent) part of the Lagrangian. All nn-point Green’s functions can be evaluated as

δnZ[J]δJ(x1)δJ(x2)δJ(xn)|J=0=𝒟φexp(dDxn=0ϵnn)φ(x1)φ(xn).\frac{\delta^{n}Z[J]}{\delta J(x_{1})\delta J(x_{2})\cdots\delta J(x_{n})}\Big{|}_{J=0}=\int{\mathcal{D}}\varphi\exp(-\int d^{D}x\sum_{n=0}^{\infty}\epsilon^{n}\mathcal{L}_{n})\varphi(x_{1})\cdots\varphi(x_{n}).

We are only interested in the connected Green’s functions Gc(x1,x2,,xn)G_{c}(x_{1},x_{2},\cdots,x_{n}). Expanding the integrand in the path integral in powers of ϵ\epsilon we obtain products of n\mathcal{L}_{n} which consist of integrals of powers of ϕ\phi and the operations denoted by w\mathcal{I}_{w} and N,s0\mathcal{F}^{0}_{N,s}. On passing the operators w\mathcal{I}_{w} and N,s0\mathcal{F}^{0}_{N,s} from inside the path integral to outside the path integral, we are left with a path integral which can be evaluated using Wick’s theorem. This is a well-defined procedure (explored in two recent papers [9, 10]) which has the advantage that the path integral is performed along the real ϕ\phi-axis. Thus, we need not be concerned with the hopelessly complicated (infinite-dimensional) integration paths that terminate in complex Stokes sectors.

4.1 Results to O(ϵ){\rm O}(\epsilon)

In general Z=exp(E0V)Z=\exp(-E_{0}V) where E0E_{0} is the ground-state energy density and VV is the spacetime volume. Using the above method to O(ϵ)O(\epsilon) we find for general DD that

ΔE=14ϵ(4π)D/2Γ(112D){ln[2(4π)D/2Γ(112D)]+ψ(3/2)}.\Delta E=\frac{1}{4}\epsilon(4\pi)^{-D/2}\Gamma(1-\frac{1}{2}D)\Big{\{}\ln[2(4\pi)^{-D/2}\Gamma(1-\frac{1}{2}D)]+\psi(3/2)\Big{\}}. (50)

As a check, for D=0D=0 and first order in ϵ\epsilon, we have

ΔE=ϵ22π𝑑ϕexp(12ϕ2)ϕ2ln(iϕ)=ϵ4[ψ(3/2)+ln(2)],\Delta E=\frac{\epsilon}{2\sqrt{2\pi}}\int_{-\infty}^{\infty}d\phi\exp(-\frac{1}{2}\phi^{2})\phi^{2}\ln(i\phi)=\frac{\epsilon}{4}[\psi(3/2)+\ln(2)], (51)

which agrees with (50). For D=1D=1, ΔE\Delta E is the expectation of the interaction Hamiltonian to O(ϵ)O(\epsilon) in the unperturbed (Gaussian) ground state

ΔE=ϵ2𝑑xexp(x2)x2ln(ix)/𝑑xexp(x2)=ϵ8ψ(3/2).\Delta E=\frac{\epsilon}{2}\int_{-\infty}^{\infty}dx\exp(-x^{2})x^{2}\ln(ix)\Big{/}\int_{-\infty}^{\infty}dx\exp(-x^{2})=\frac{\epsilon}{8}\psi(3/2). (52)

The calculated higher order connected Green’s function to O(ϵ)O(\epsilon) is

Gc(y1,,yn)=12ϵ(i)nΓ(n21)[12(4π)D/2Γ(112D)]1n/2dDxk=1n1(ykx),G_{c}(y_{1},\cdots,y_{n})=-\frac{1}{2}\epsilon(-i)^{n}\Gamma(\frac{n}{2}-1)[\frac{1}{2}(4\pi)^{-D/2}\Gamma(1-\frac{1}{2}D)]^{1-n/2}\int d^{D}x\prod_{k=1}^{n}\triangle_{1}(y_{k}-x), (53)

where the free propagator λ(x)\triangle_{\lambda}(x) is associated with L0=12(ϕ)2+12λ2ϕ2L_{0}=\frac{1}{2}(\nabla\phi)^{2}+\frac{1}{2}\lambda^{2}\phi^{2} and obeys the equation

(2+λ2)λ(x)=δ(D)(x).(-\nabla^{2}+\lambda^{2})\triangle_{\lambda}(x)=\delta^{(D)}(x). (54)

The solution of (54)(\ref{e58}) is

λ(x)=λD/21|x|1D/2(2π)D/2K1D/2(λ|x|)\triangle_{\lambda}(x)=\lambda^{D/2-1}|x|^{1-D/2}(2\pi)^{-D/2}K_{1-D/2}(\lambda|x|) (55)

and

λ(0)=λD2(4π)D/2Γ(1D2)12πδ(δ0).\triangle_{\lambda}(0)=\lambda^{D-2}(4\pi)^{-D/2}\Gamma(1-\frac{D}{2})\sim\frac{1}{2\pi\delta}\,\quad(\delta\to 0). (56)

From (53) it is clear that as δ0+\delta\to 0+, the connected Green’s functions Gc(y1,,yn)0G_{c}(y_{1},\cdots,y_{n})\to 0 for n3n\geq 3. This indicates that at least to O(ϵ){\rm O}(\epsilon) the theory is noninteracting at D=2D=2. When we consider O(ϵ2){\rm O}(\epsilon^{2}) contributions, we will reexamine this issue.

Turning to n=1n=1 and n=2n=2 we have

G1=iϵ12π(4π)D/2Γ(1D2)G_{1}=-i\epsilon\sqrt{\frac{1}{2}\pi(4\pi)^{-D/2}\Gamma(1-\frac{D}{2})} (57)

and the two-point function G~2(p){\widetilde{G}}_{2}(p) in momentum space is

G~2(p)=1/[p2+1+ϵK+O(ϵ2)]{\widetilde{G}}_{2}(p)=1/[p^{2}+1+\epsilon K+O(\epsilon^{2})] (58)

where K=3212γ+12ln[12(4π)D/2Γ(1D2)]K=\frac{3}{2}-\frac{1}{2}\gamma+\frac{1}{2}\ln[\frac{1}{2}(4\pi)^{-D/2}\Gamma(1-\frac{D}{2})]. Thus the renormalised mass to O(ϵ)O(\epsilon) is

MR2=1+Kϵ+O(ϵ2).M_{R}^{2}=1+K\epsilon+O(\epsilon^{2}). (59)

So near D=2D=2

G1iϵ12δG_{1}\sim-i\epsilon\frac{1}{2\sqrt{\delta}} (60)

and

MR212ϵlnδ+A,M_{R}^{2}\sim-\frac{1}{2}\epsilon\ln\delta+A, (61)

where A=1+ϵ[32γ212ln(4π)]A=1+\epsilon[\frac{3}{2}-\frac{\gamma}{2}-\frac{1}{2}\ln(4\pi)]. Because of the singularities in G1G_{1} and MR2M_{R}^{2} as δ0\delta\to 0 some renormalisation is needed to remove these infinities. The question is whether perturbative renormalisation can be performed in the context of the novel ϵ\epsilon-expansion method. We can remove the divergence in G1G_{1} by introducing in the Lagrangian a linear counter term ivϕiv\phi where vv has dimension (mass)1+D/2(\rm{mass})^{1+D/2} and v=v1ϵ+v2ϵ2+v3ϵ3+v=v_{1}\epsilon+v_{2}\epsilon^{2}+v_{3}\epsilon^{3}+\cdots. Since vv is real such a term is compatible with 𝒫𝒯\mathcal{PT}-symmetry. Adding also a mass counter term μ\mu we can consider the Lagrangian density

L=12(ϕ)2+12μ2ϕ2+12gμ02ϕ2(iμ01D/2ϕ)ϵ,L=\frac{1}{2}(\nabla\phi)^{2}+\frac{1}{2}\mu^{2}\phi^{2}+\frac{1}{2}g\mu_{0}^{2}\phi^{2}(i\mu_{0}^{1-D/2}\phi)^{\epsilon}, (62)

where ϕ\phi is dimensionful. Using this Lagrangian we obtain

G1=iϵgm2μ0D/2112πmD21(0)+iϵv1μ02m2G_{1}=-\frac{i\epsilon g}{m^{2}}\mu_{0}^{D/2-1}\sqrt{\frac{1}{2}\pi m^{D-2}\triangle_{1}(0)}+\frac{i\epsilon v_{1}}{\mu_{0}^{2}m^{2}} (63)

and for the renormalised mass

MR2=(mμ0)2+12ϵgμ02{3γ+ln[12mD21(0)]}.M_{R}^{2}=(m\mu_{0})^{2}+\frac{1}{2}\epsilon g\mu_{0}^{2}\{3-\gamma+\ln[\frac{1}{2}m^{D-2}\triangle_{1}(0)]\}. (64)

In both expressions we have introduced the dimensionless quantity m2=g+μ2/μ02m^{2}=g+\mu^{2}/\mu_{0}^{2} and as δ0\delta\to 0

G1iϵgμ02+μ2(v1gμ022δ)G_{1}\sim\frac{i\epsilon}{g\mu_{0}^{2}+\mu^{2}}(v_{1}-\frac{g\mu_{0}^{2}}{2\sqrt{\delta}}) (65)

and

MR2μ212ϵgμ02lnδ+A,M_{R}^{2}\sim\mu^{2}-\frac{1}{2}\epsilon g\mu_{0}^{2}\ln\delta+A, (66)

where A=gμ02{1+ϵ[32γ212ln(4π)]}A=g\mu_{0}^{2}\{1+\epsilon[\frac{3}{2}-\frac{\gamma}{2}-\frac{1}{2}\ln(4\pi)]\} is a finite quantity. By setting v1=gμ022δv_{1}=\frac{g\mu_{0}^{2}}{2\sqrt{\delta}} we have a finite G1=0G_{1}=0. MRM_{R} is logarithmically divergent in δ\delta. We absorb this divergence into μ\mu by setting

μ2=B+ϵ2gμ02lnδ,\mu^{2}=B+\frac{\epsilon}{2}g\mu_{0}^{2}\ln\delta, (67)

so that MR2=A+BM_{R}^{2}=A+B and BB is a finite quantity determined in principle from experiment.

4.2 Calculations to second order in ϵ\epsilon

In second order the calculations become much more involved. The O(ϵ2){\rm O}(\epsilon^{2}) contribution to the connected part of G1G_{1}, which we denote by G1,2G_{1,2}, can be shown to be

G1,2=12igm212π(0){[ln[2μ02D(0)]+ψ(2)}+18ig2m412π(0){(ln[2μ02D(0)]+ψ(32))(6D)+2D4+4(0)μ02m2dDx(1+(x)(0))2ln[1+(x)(0)]},G_{1,2}=-\frac{1}{2}igm^{-2}\sqrt{\frac{1}{2}\pi\triangle(0)}\Big{\{}[\ln[2\mu_{0}^{2-D}\triangle(0)]+\psi(2)\Big{\}}+\frac{1}{8}ig^{2}m^{-4}\sqrt{\frac{1}{2}\pi\triangle(0)}\Big{\{}(\ln[2\mu_{0}^{2-D}\triangle(0)]\\ +\psi(\frac{3}{2}))(6-D)+2D-4+4\triangle(0)\mu_{0}^{2}m^{2}\int d^{D}x(1+\frac{\triangle(x)}{\triangle(0)})^{2}\ln[1+\frac{\triangle(x)}{\triangle(0)}]\Big{\}}, (68)

where (0)\triangle(0) denotes for brevity μ0m(0)\triangle_{\mu_{0}m}(0). As δ0\delta\to 0,

G1,2i4gm2δ1/2[ψ(2)ln(πδ)]+i4g2m4δ1/2[1+ψ(3/2)lnπlnδ]+O(δ).G_{1,2}\sim-\frac{i}{4}gm^{-2}\delta^{-1/2}[\psi(2)-\ln(\pi\delta)]+\frac{i}{4}g^{2}m^{-4}\delta^{-1/2}[1+\psi(3/2)-\ln\pi-\ln\delta]+O(\delta). (69)

So the algebraic divergence δ1/2\delta^{-1/2} persists to O(ϵ2)O(\epsilon^{2}). The divergence can be removed through v2v_{2}. Similarly the lnδ\ln\delta divergence persists for G2G_{2} at second order. Interestingly, the higher-order Green’s functions continue to vanish for D=2D=2. Avoidance of a noninteracting theory remains an open question in this approach [10].

5 Renormalisation group flows of 𝒫𝒯\mathcal{PT}-symmetric theories

So far we have considered whether a Hermitian theory can lead to a non-Hermitian 𝒫𝒯\mathcal{PT}-symmetric theory due to the effect of renormalisation. We ask the opposite question in this section: Can a 𝒫𝒯\mathcal{PT}-symmetric field theory retain its 𝒫𝒯\mathcal{PT} symmetry as the Lagrangian flows due to renormalisation?

We review some preliminary work on this question. In its full generality this is an intractable problem. We turn to the framework of the functional renormalization group [15], which combines the functional formulation of quantum field theory with the Wilsonian renormalization group. It is possible to make some progress in solving the functional equations using the simpler approach developed by Wetterich [16] and Morris [17] which, in the local potential approximation (31), leads to a nonlinear partial differential equation (PDE) rather than a functional equation. This simplification enables substantial progress in understanding the effective potential in arbitrary spacetime dimensions. In DD dimensions this PDE is

kUk(ϕc)=1πDkD+1k2+Uk′′(ϕc),\partial_{k}U_{k}\left(\phi_{c}\right)=\frac{1}{\pi_{D}}\frac{k^{D+1}}{k^{2}+U_{k}^{\prime\prime}\left(\phi_{c}\right)}, (70)

where πD=D(2π)DSD1\pi_{D}=\frac{D(2\pi)^{D}}{S_{D-1}} and SD1=2πD/2Γ(D/2)S_{D-1}=\frac{2\pi^{D/2}}{\Gamma(D/2)} is the surface area of a unit DD-dimensional sphere. We may assume that the equations for Uk(ϕc)U_{k}\left(\phi_{c}\right) involve dimensionless quantities. If this were not so, we could achieve dimensionlessness by introducing a mass scale MM. For example, for D=1D=1 the dimensionless variables, denoted by a tilde, are ϕ~=M1/2ϕ\tilde{\phi}=M^{1/2}\phi, g~=M3g\tilde{g}=M^{-3}g, k~=M1k\tilde{k}=M^{-1}k, and μ~=M1μ\tilde{\mu}=M^{-1}\mu. The Wetterich equation (70) can be thought of as being in terms of such dimensionless variables.

To avoid difficulties in numerical analysis associated with boundary conditions, in the past the PDE (70) has been analyzed by approximating Uk(ϕc)U_{k}\left(\phi_{c}\right) as a finite series in powers of the field ϕc\phi_{c}. This ansatz leads to a sequence of coupled nonlinear ordinary differential equations. The consistency of such a procedure has not been established.

For D=1D=1 (the quantum-mechanical case), (70) becomes

kUk(ϕc)=132π2k2k2+Uk′′(ϕc).\partial_{k}U_{k}\left(\phi_{c}\right)=\frac{1}{32\pi^{2}}\frac{k^{2}}{k^{2}+U_{k}^{\prime\prime}\left(\phi_{c}\right)}. (71)

Even in this one-dimensional setting, no exact solution to this nonlinear PDE is known and only numerical solutions have been discussed.

We depart from the above treatment by performing an asymptotic analysis for large values of the cut-off kk [11]. The novelty of the approach used here is that it avoids the appearance of coupled nonlinear ordinary differential equations. To leading order the results of this analysis are qualitatively different depending on whether the space-time dimension DD is greater or less than 22.

Letting z=k2+Dz=k^{2+D} and Uk(ϕc)=U(z,ϕ)U_{k}(\phi_{c})=U(z,\phi), we can rewrite (70) as

Uz(z,ϕ)=1(2+D)πD1z22+D+Uϕϕ(z,ϕ),U_{z}(z,\phi)=\frac{1}{(2+D)\pi_{D}}\frac{1}{z^{\frac{2}{2+D}}+U_{\phi\phi}(z,\phi)}, (72)

where the subscripts on UU indicate partial derivatives. We now assume that for large zz we can neglect the UϕϕU_{\phi\phi} term in the denominator. (The consistency of this assumption is easy to verify when D<2D<2.) Then, for large zz we have to leading order in our approximation scheme

Uz(z,ϕ)1(2+D)πDz22+D(z1).U_{z}(z,\phi)\sim\frac{1}{(2+D)\pi_{D}}z^{-\frac{2}{2+D}}\quad(z\gg 1). (73)

On incorporating a correction ϵ\epsilon to this leading behavior

U(z,ϕ)=1DπDzD2+D+ϵ(z,ϕ),U(z,\phi)=\frac{1}{D\pi_{D}}z^{\frac{D}{2+D}}+\epsilon(z,\phi), (74)

we get to order O(ϵ2)O(\epsilon^{2})

ϵz(z,ϕ)=1(D+2)πDz42+Dϵϕϕ(z,ϕ).\epsilon_{z}(z,\phi)=-\frac{1}{(D+2)\pi_{D}}z^{-\frac{4}{2+D}}\epsilon_{\phi\phi}(z,\phi). (75)

On making the further change of variable t=D+22DzD2D+2t=\frac{D+2}{2-D}z^{\frac{D-2}{D+2}}, (75) becomes

ϵt(t,ϕ)=1(D+2)πDϵϕϕ(t,ϕ).\epsilon_{t}(t,\phi)=\frac{1}{(D+2)\pi_{D}}\epsilon_{\phi\phi}(t,\phi). (76)

The variable tt is positive for D<2D<2 and negative for D>2D>2 and is not defined at D=2D=2. Thus, (76) is a conventional diffusion equation for D<2D<2 but is a backward diffusion equation for D>2D>2. The backward diffusion equation is an inverse problem that is ill-posed. The problems associated with this ill-posedness may be connected with difficulties in solving (70) numerically when D=4D=4.

In the preliminary study these issues were not addressed further. The case D=1D=1 was considered and some simple 𝒫𝒯\mathcal{PT}-symmetric theories were studied [11]. From (70), on defining U^kUkkDDπD\widehat{U}_{k}\equiv U_{k}-\frac{k^{D}}{D\pi_{D}}, we can deduce that

kU^k=kD1U^k′′πD(k2+U^k′′).\partial_{k}\widehat{U}_{k}=-\frac{k^{D-1}\widehat{U}_{k}^{\prime\prime}}{\pi_{D}\left(k^{2}+\widehat{U}_{k}^{\prime\prime}\right)}. (77)

From (77) it is consistent to assume that U^V(ϕ)\widehat{U}\to V(\phi) and U^′′k20\frac{\widehat{U}^{{}^{\prime\prime}}}{k^{2}}\to 0 as kk\to\infty. (For simplicity of notation we have dropped the suffix kk in U^k\widehat{U}_{k}.) Let us write the correction as V(ϕ)+1kU1(ϕ)V(\phi)+\frac{1}{k}U_{1}(\phi). On substituting in (77), we obtain

U1(ϕ)=1πV′′(ϕ).U_{1}(\phi)=\frac{1}{\pi}V^{\prime\prime}(\phi). (78)

We can proceed in this way and write the next correction as U^(ϕ)=V(ϕ)+1kπV′′(ϕ)+1k2U2(ϕ)\widehat{U}(\phi)=V(\phi)+\frac{1}{k\pi}V^{\prime\prime}(\phi)+\frac{1}{k^{2}}U_{2}(\phi). This leads to U2(ϕ)=12π2V(4)U_{2}(\phi)=\frac{1}{2\pi^{2}}V^{(4)} where V(4)(ϕ)d4dϕ4V(ϕ)V^{(4)}(\phi)\equiv\frac{d^{4}}{d\phi^{4}}V(\phi). Repeating this procedure, we get U3(ϕ)=13(12π3V(6)(ϕ)V(2)(ϕ)2)U_{3}(\phi)=\frac{1}{3}\left(\frac{1}{2\pi^{3}}V^{(6)}(\phi)-{V^{(2)}(\phi)}^{2}\right). This procedure can be formalized: On writing δ=π/k\delta=\pi/k (not to be confused with δ\delta in the last section) and x=πϕx=\pi\phi, (77) becomes

δU^=2x2U^1+δ22x2U^\frac{\partial}{\partial\delta}\widehat{U}=\frac{\frac{\partial^{2}}{\partial x^{2}}\widehat{U}}{1+\delta^{2}\frac{\partial^{2}}{\partial x^{2}}\widehat{U}} (79)

and U^=n=0δnUn(ϕ)\widehat{U}=\sum_{n=0}^{\infty}\delta^{n}U_{n}(\phi).

Refer to caption
Refer to caption
Figure 7: Effective potential flow for massive iϕ3i\phi^{3} with the mass parameter μ=1\mu=1. Shown is a plot of the real and imaginary parts of the P55P_{5}^{5} approximant plotted as functions of real ϕ\phi. Observe that there are no poles. The real part of the potential is right-side-up, so there is no instability. Furthermore, apart from small fluctuations near the origin, the imaginary part of the potential behaves like iϕ3i\phi^{3} for large |ϕ||\phi|.

When δ\delta is small, the scale kk is large and we are probing the microscopic potential. As δ\delta\to\infty we probe the infrared limit of the effective potential. However, the analysis outlined above was based on a perturbation theory in δ\delta. There is a parallel to the theory of phase transitions, where a physical entity is expanded in inverse powers of the temperature TT, and then an extrapolation procedure is used to find the critical behavior at lower TT. This often-used extrapolation technique is based on Padé approximation.

So far, our treatment has been for a general potential; we will now specialize V(x)V(x) to two cases V(x)=gx3V(x)=gx^{3} and V(x)=gx4V(x)=gx^{4}, which are 𝒫𝒯\mathcal{PT}-symmetric for g=ig=i and gg real. We also include a mass term. Our purpose is to consider the series n=0NδnUn(x)\sum_{n=0}^{N}\delta^{n}U_{n}(x) generated from the series relevant to these two cases and to consider the δ0\delta\to 0 limit.

5.1 Cubic potentials

We consider both massless and massive cubic potentials. In the former case U0(ϕ)=gϕ3U_{0}(\phi)=g\phi^{3} and in the latter case U0(ϕ)=μ2ϕ2+gϕ3U_{0}(\phi)=\mu^{2}\phi^{2}+g\phi^{3}. For the massless case the solution to (79) is

U1(ϕ)=6gϕ,U10(ϕ)=1621728175g5ϕ3,U2(ϕ)=0,U11(ϕ)=178822081925g5ϕ4665611g6ϕ6,U3(ϕ)=12g2ϕ2,U12(ϕ)=30608064385g6ϕ4,U4(ϕ)=6g2,U13(ϕ)=27993613g7ϕ7453796732825025g6ϕ2,U5(ϕ)=2165g3ϕ3,U14(ϕ)=45127549447007g7ϕ55900745888175175g6,U6(ϕ)=4565g3ϕ,U15(ϕ)=1439488599552525525g7ϕ3167961615g8ϕ8,U7(ϕ)=12967g4ϕ4,U16(ϕ)=295324339824175175g7ϕ250637431685005g8ϕ6,U8(ϕ)=3466835g4ϕ2,U17(ϕ)=1007769617g9ϕ91051037069898242977975g8ϕ4,U9(ϕ)=77769g5ϕ589496315g4,U18(ϕ)=321231088742485085g9ϕ71010146205387522127125g8ϕ2.\begin{array}[]{ll}U_{1}(\phi)=6g\phi,&\qquad\quad U_{10}(\phi)=\textstyle{\frac{1621728}{175}}g^{5}\phi^{3},\\ U_{2}(\phi)=0,&\qquad\quad U_{11}(\phi)=\textstyle{\frac{17882208}{1925}}g^{5}\phi-\textstyle{\frac{46656}{11}}g^{6}\phi^{6},\\ U_{3}(\phi)=-12g^{2}\phi^{2},&\qquad\quad U_{12}(\phi)=-\textstyle{\frac{30608064}{385}}g^{6}\phi^{4},\\ U_{4}(\phi)=-6g^{2},&\qquad\quad U_{13}(\phi)=\textstyle{\frac{279936}{13}}g^{7}\phi^{7}-\textstyle{\frac{4537967328}{25025}}g^{6}\phi^{2},\\ U_{5}(\phi)=\textstyle{\frac{216}{5}}g^{3}\phi^{3},&\qquad\quad U_{14}(\phi)=\textstyle{\frac{4512754944}{7007}}g^{7}\phi^{5}-\textstyle{\frac{5900745888}{175175}}g^{6},\\ U_{6}(\phi)=\textstyle{\frac{456}{5}}g^{3}\phi,&\qquad\quad U_{15}(\phi)=\textstyle{\frac{1439488599552}{525525}}g^{7}\phi^{3}-\textstyle{\frac{1679616}{15}}g^{8}\phi^{8},\\ U_{7}(\phi)=-\textstyle{\frac{1296}{7}}g^{4}\phi^{4},&\qquad\quad U_{16}(\phi)=\textstyle{\frac{295324339824}{175175}}g^{7}\phi-\textstyle{\frac{25063743168}{5005}}g^{8}\phi^{6},\\ U_{8}(\phi)=-\textstyle{\frac{34668}{35}}g^{4}\phi^{2},&\qquad\quad U_{17}(\phi)=\textstyle{\frac{10077696}{17}}g^{9}\phi^{9}-\textstyle{\frac{105103706989824}{2977975}}g^{8}\phi^{4},\\ U_{9}(\phi)=\textstyle{\frac{7776}{9}}g^{5}\phi^{5}-\textstyle{\frac{89496}{315}}g^{4},&\qquad\quad U_{18}(\phi)=\textstyle{\frac{3212310887424}{85085}}g^{9}\phi^{7}-\textstyle{\frac{101014620538752}{2127125}}g^{8}\phi^{2}.\end{array} (80)

These expressions, which are valid for pure imaginary gg as well as for real gg, are cumbersome but manageable. We stress that there has been no truncation of the function space on which U^(ϕ)\widehat{U}(\phi) has support. This contrasts with the usual approach which requires a truncation at the onset of the calculation of the renormalization group flow. This absence of truncation continues to be a feature for the massive case.

The iterative solution to (79) for the massive case is similar to that for the massless case, but the expressions for Un(ϕ)U_{n}(\phi) have many more terms. For example, the coefficient of δ9\delta^{9} is

U9(ϕ)\displaystyle U_{9}(\phi) =\displaystyle= 8315(34020g5ϕ5+56700g4μ2ϕ411187g4+37800g3μ4ϕ3\displaystyle\textstyle{\frac{8}{315}}\big{(}34020g^{5}\phi^{5}+56700g^{4}\mu^{2}\phi^{4}-11187g^{4}+37800g^{3}\mu^{4}\phi^{3}
+12600g2μ6ϕ2+2100gμ8ϕ+140μ10),\displaystyle~{}~{}+12600g^{2}\mu^{6}\phi^{2}+2100g\mu^{8}\phi+140\mu^{10}\big{)},

which in the massless limit μ0\mu\to 0 reduces to the two-term expression in (80). We refrain from listing the coefficients explicitly and instead proceed directly to the large-δ\delta behavior of the diagonal Padé approximants. We denote the diagonal Padé approximants in the massless case by PN0,N(δ)P_{N}^{0,N}(\delta) and in the massive case by PNN(δ)P_{N}^{N}(\delta). Let us examine some low-order Padé approximants in the limit δ\delta\to\infty. For example,

limδP22(δ)=18g3ϕ5+3g2(10μ2ϕ49)+14gμ4ϕ3+2μ6ϕ22(3gϕ+μ2)2.\lim_{\delta\to\infty}P_{2}^{2}(\delta)=\frac{18g^{3}\phi^{5}+3g^{2}\left(10\mu^{2}\phi^{4}-9\right)+14g\mu^{4}\phi^{3}+2\mu^{6}\phi^{2}}{2\left(3g\phi+\mu^{2}\right)^{2}}.

[As a check, when μ=0\mu=0 this expression agrees with the corresponding expression gϕ332ϕ2g\phi^{3}-\frac{3}{2\phi^{2}} for P20,2(δ)P_{2}^{0,2}(\delta).] For N=3N=3 we obtain

limδP30,3(ϕ)=gϕ3(736gϕ5+1705)25800gϕ5andP33(ϕ)=u3/d3,\lim_{\delta\to\infty}P_{3}^{0,3}(\phi)=\frac{g\phi^{3}\left(736g\phi^{5}+1705\right)}{25-800g\phi^{5}}\qquad{\rm and}\qquad P_{3}^{3}(\phi)=-u_{3}/d_{3},

where

u3\displaystyle u_{3} =\displaystyle= 1609632g8ϕ8+729g7ϕ3(5888μ2ϕ4+5115)+243g6μ2ϕ2(23008μ2ϕ4+15345)\displaystyle 1609632g^{8}\phi^{8}+729g^{7}\phi^{3}\left(5888\mu^{2}\phi^{4}+5115\right)+243g^{6}\mu^{2}\phi^{2}\left(23008\mu^{2}\phi^{4}+15345\right)
+1296g5μ4ϕ(3376μ2ϕ4+945)+15120g4μ6(142μ2ϕ4+9)+658944g3μ10ϕ3\displaystyle~{}~{}+1296g^{5}\mu^{4}\phi\left(3376\mu^{2}\phi^{4}+945\right)+15120g^{4}\mu^{6}\left(142\mu^{2}\phi^{4}+9\right)+658944g^{3}\mu^{10}\phi^{3}
+121824g2μ12ϕ2+12288gμ14ϕ+512μ16,\displaystyle~{}~{}+121824g^{2}\mu^{12}\phi^{2}+12288g\mu^{14}\phi+512\mu^{16},
d3=225g2[7776g5ϕ5+81g4(160μ2ϕ43)+8640g3μ4ϕ3+2880g2μ6ϕ2+480gμ8ϕ+32μ10].d_{3}=225g^{2}[7776g^{5}\phi^{5}+81g^{4}(160\mu^{2}\phi^{4}-3)+8640g^{3}\mu^{4}\phi^{3}+2880g^{2}\mu^{6}\phi^{2}+480g\mu^{8}\phi+32\mu^{10}].

[As a check, we have limμ0P33(ϕ)=P30,3(ϕ)\lim_{\mu\to 0}P_{3}^{3}(\phi)=P_{3}^{0,3}(\phi).]

Note that for the massless case P20,2P_{2}^{0,2} has a double pole for both real and imaginary gg. In fact, the same is true for Pn0,nP_{n}^{0,n} for even nn. This pole is an artifact of the Padé approximation and is not present when nn is odd, so we will only consider the behavior of these approximants for odd nn. In general, the odd-nn diagonal Padé approximants have no singularities at all on the real-ϕ\phi axis when gg is imaginary but singularities occur for the case of real gg . These findings indicate that the 𝒫𝒯\mathcal{PT}-symmetric effective potential is well behaved in the infrared limit . From the expressions for the diagonal Padé approximants we see that for large |ϕ||\phi|, the leading behavior of the imaginary part of the effective potential is exactly iϕ3i\phi^{3}. Consequently the 𝒫𝒯\mathcal{PT} nature of the interaction is preserved under renormalization.

A similar investigation of quartic potentials also shows that 𝒫𝒯\mathcal{PT}-symmetry is maintained under renormalization [11].

6 Conclusions

The study of 𝒫𝒯\mathcal{PT}-symmetric quantum field theory is still in its infancy. We have shown that there is a link between renormalization of Hermitian field theories and 𝒫𝒯\mathcal{PT}-symmetric field theories. This provides a motivation for the study of 𝒫𝒯\mathcal{PT}-symmetric field theories. Often, the usual tools of quantum field theory cannot be applied directly. We have reviewed some interesting and promising approaches to studying these new field theories. Clearly, there remain some unresolved issues with these approaches, and this will lead to many opportunities for future research.

Acknowledgments

CMB thanks the Alexander von Humboldt and Simons Foundations for financial support. SS and CMB thank the UK Engineering and Physical Sciences Research Council for financial support.

References

References

  • [1] C. M. Bender and S. Boettcher, Phys. Rev. Lett. 80, 5243 (1998)
  • [2] C. M. Bender et al., 𝒫𝒯\mathcal{PT} Symmetry in Quantum and Classical Physics (World Scientific, Singapore, 2019)
  • [3] P. E. Dorey, C. Dunning, and R. Tateo, J. Phys. A: Math. Gen. 34, 5679 (2001)
  • [4] A. Mostafazadeh, J. Math. Phys. 43, 205 (2002); J. Phys. A 36, 7081 (2003)
  • [5] H. F. Jones and R. J. Rivers, Phys. Rev. D 75, 025023 (2007)
  • [6] C. M. Bender, S. F. Brandt, J.-H. Chen, and Q. Wang, Phys. Rev. D 71, 025014 (2005)
  • [7] M. Sher, Phys. Rept. 179, 273 (1989)
  • [8] C. M. Bender, D. Hook, N. E. Mavromatos, and S. Sarkar, J. Phys. A: Math. Gen. 49 , 45LT01 (2016)
  • [9] C. M. Bender, N. Hassanpour, S. P. Klevansky, and S. Sarkar, Phys. Rev. D 98, 125003 (2018)
  • [10] A. Felski, C. M. Bender, S. P. Klevansky, and S. Sarkar, arXiv:2103.07577 [hep-th]
  • [11] C. M. Bender and S. Sarkar, J. Phys. A: Math. Gen. 51, 225202 (2018)
  • [12] H. Lehmann, K. Symanzik, and W. Zimmermann, Nuovo Cimento 1, 205 (1955)
  • [13] G. Barton, Introduction to Advanced Field Theory (Wiley & Sons, New York, 1963), Chap. 12
  • [14] T. D. Lee, Phys. Rev. 95, 1329 (1954)
  • [15] H. Gies, Lect. Notes Phys. 852, 287 (2012)
  • [16] C. Wetterich, Int. J. Mod. Phys. A 16, 1951 (2001)
  • [17] T. R. Morris, Phys. Lett. B 329, 241 (1994)
  • [18] J. Alexandre, N. Houston and N. E. Mavromatos, Phys. Rev. D88 125017 (2013); J. Alexandre, N. Houston and N. E. Mavromatos, Intnl. J. Mod. Phys. D 24 1541004 (2015)
  • [19] C. M. Bender, D. D. Holm and D. W. Hook, J. Phys. A: Math. Theor. 40 F81 (2007); D. C. Brody and D. W. Hook, J. Phys. A: Math. Theor. 41 352003 (2008); A. Fring and B. Bagchi, J. Phys. A: Math. Theor. 44 325201 (2011)
  • [20] C. M. Bender and S. A. Orszag, Advanced Mathematical Methods for Scientists and Engineers (McGraw Hill, New York, 1978)