Symmetry and Renormalisation in Quantum Field Theory
Abstract
Quantum systems governed by non-Hermitian Hamiltonians with symmetry are special in having real energy eigenvalues bounded below and unitary time evolution. We argue that 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 -symmetric effective Lagrangians. We show how 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 -symmetric interpretation is naturally built into a path integral formulation of quantum field theory; there is no requirement to calculate explicitly the norm that occurs in Hamiltonian quantum theory. We discuss examples where -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 -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 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 -symmetric quantum mechanics.
-symmetric quantum mechanics can be considered to be a one-dimensional quantum field theory. To date there have been no clear examples of -symmetric quantum field theory in higher dimensions. We examine the possibility that 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 -symmetric quantum field theory.
Using numerical techniques the work of Bender and Boettcher [1] showed that quantum-mechanical Hamiltonians of the form
(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 symmetry as a special case. A Hamiltonian is pseudo-Hermitian if is not selfadjoint but
(2) |
where is a positive-definite Hermitian operator. In terms of the Hilbert space of states has a positive-definite inner product given by
(3) |
where 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 , which is difficult to calculate in general. This would seem to imply an impediment to calculating correlation functions in -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 . Hence, we shall take the path-integral representation as the definition of a quantum field theory and, except for the Lee model, not consider 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 symmetry:
2 Role of 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 , where denotes an arbitrary field of the theory, such that we have the weak convergence relation
(4) |
where and are experimentally accessible states and and 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 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 and , as well as a neutral scalar field . Antifermions are not in the theory and so there is no crossing symmetry. The permitted reaction channel is
(5) |
but the channel
(6) |
is not allowed.
In momentum space the Hamiltonian of the Lee model (in a space of finite volume ) is
(7) |
where
(8) |
and
(9) |
is a dimensional cut-off function which is chosen to tend to for large where . Standard commutation and anti-commutation rules are assumed:
(10) | ||||
(11) |
All other commutators and anticommutators vanish. Bare operators and coupling constant appear in . However are renormalised parameters and so are determined by experiment; is the mass counterterm for the particle and is a function of . No further mass renormalisation is actually needed. The coupling in principle is determined from the scattering cross-section of and 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 is real and so is Hermitian. Let us choose as a basis for the Hilbert space states of the form
(12) |
From it is clear that there are two conserved quantities and :
(13) |
where . Hence the Hilbert space is partitioned into an infinite number of independent sectors with fixed and . Although this makes the model soluble, the analysis of sectors with large and 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 and and (ii) modifying the model so that there is no 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 , where
(14) |
and
(15) |
We will look at the 2-dimensional Hilbert space associated with this sector and consider the eigenstates, which we will denote by
(16) | ||||
(17) |
with eigenvalues and . Let us denote by , the bare mass of . The eigenvalues can be shown to satisfy the equations
(18) | ||||
(19) |
where Following field theory, we define the wave-function renormalisation constant through the relation
(20) |
which gives
(21) |
and the coupling constant renormalisation through
(22) |
We deduce that
(23) |
where . If , the experimentally determined value, exceeds then becomes pure imaginary and the Hamiltonian becomes non-Hermitian. Although the Hamiltonian has become non-Hermitian, it is -symmetric. Explicitly, the transformations due to are [6]
(24) | ||||||||
(25) |
The transformations due to are
(26) | ||||||||
(27) |
It is now straightforward to check that is -symmetric. This is our first example of a -symmetric field theory that arises from a Hermitian field theory after renormalisation. On introducing a -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 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 -symmetric.
The change from the conventional to the Wilson approach can be illustrated by considering a scalar field whose (conventional) action is given by
(28) |
where the UV cut-off is . If we integrate over Fourier modes with momenta of magnitude in the shell we arrive at an action which we will denote by where has Fourier modes with (a sharp infrared cut-off).111Wetterich [16] introduced a smooth infrared cut-off function rather than a sharp cut-off so that , where . A common choice for is . The associated partition function is
(29) |
and a Legendre transformation on leads to the effective action
(30) |
Within systematic (derivative) approximation schemes [16, 17] the effective action can be represented (at the lowest level of approximation) by the ansatz
(31) |
The effective potential, which results from renormalisation, will show the emergence of symmetry in the next models that we will discuss. This ansatz represents quite a severe approximation since, for an -independent , we have a one-dimensional approximation to an infinite-dimensional field.
We consider two theories of current interest using this formalism:
Both examples, in conventional quantum mechanics, would show unstable behaviour for large . Under we have and under we have . So is symmetric in both cases.
We study such effective potentials by considering three related quantum mechanical Hamiltonians , :
(32) | ||||
(33) | ||||
(34) |
which are non-Hermitian but -symmetric. We will find that the the first two Hamiltonians will show unbroken symmetry and the Hamiltonian will show broken symmetry.
3.1 The spectra for
We use the WKB method of semiclassical quantum mechanics to determine the energy spectrum of :
-
•
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
(35) |
We take 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- axis. To find this point we set () and obtain the algebraic equation . Solving this equation by using Newton’s method, we find that the turning point lies at . To find the other turning points we seek solutions to (35) in polar form (). Substituting for in (35) and taking the imaginary part, we obtain
(36) |
where 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
(37) |
We then use (36) to eliminate from (37) and use Newton’s method to determine . For and , two -symmetric (left-right symmetric) pairs of turning points lie at and at . For and there is a turning point at ; the -symmetric image of this turning point lies on sheet at .
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 (). On this sheet a branch cut runs from the origin to on the imaginary axis; this choice of branch cut respects the symmetry of the configuration. The integration path goes from the left turning point to the right turning point [20]:
(38) |


If the energy is large , then from (36) with we find that the turning points lie slightly below the real axis at and at with
(39) |
We choose the path of integration in (38) to have a constant imaginary part so that the path is a horizontal line from to . Since is large, is large and thus is small. We obtain the simplified approximate quantization condition
(40) |
which leads to the WKB approximation for :
(41) |
Calculation of values of energy | ||||
Energy level | 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 |


Analysis of the supergravity model Hamiltonian : The classical trajectories for the Hamiltonian are plotted in Figs. 3 and 4. Like the classical trajectories for the Hamiltonian ), these trajectories are closed, which implies that all the eigenvalues for are real.
Analysis of the Higgs model Hamiltonian : To make sense of we again introduce a parameter and we define as the limit of as . This case is distinctly different from that for . Figure 5 shows that the symmetry is broken for all . When , there are only four real eigenvalues: , , , and . To confirm this result we plot a classical trajectory for in Fig. 6. In contrast with Fig. 4, the trajectory is open and not left-right symmetric.


4 Renormalisation and canonical scalar field theory
We have shown examples of -symmetric quantum field theories arising from the process of renormalisation in field theories. The properties of -symmetric field theories have yet to be established. In quantum mechanics much progress in understanding symmetry has been made through the study of the Hamiltonian in (1). A natural generalisation to an Euclidean field theory Hamiltonian in -dimensions is to consider the Lagrangian
(42) |
where is a dimensional pseudoscalar field, and . 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 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.
rewriting of the interaction in terms of a formal series of polynomial interactions;
-
2.
creating modified Feynman rules in a nonperturbative expansion;
-
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 -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 within polynomial interactions in a field , we calculate connected Green’s functions using a partition functional defined through a path integral as
(43) |
where is a source. It is well known that the normalised partition functional satisfies
(44) |
This result simplifies our calculations in (ii) and (iii). First, we note that formally, on writing (and suppressing the argument of )
(45) |
because
(46) |
The series in (45) can be expressed in terms of powers of on noting that
-
1.
(47) where .
-
2.
(48) where .
Thus, we can rewrite the partition function in (43) as
(49) |
where is the free (that is, -independent) part of the Lagrangian. All -point Green’s functions can be evaluated as
We are only interested in the connected Green’s functions . Expanding the integrand in the path integral in powers of we obtain products of which consist of integrals of powers of and the operations denoted by and . On passing the operators and 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 -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
In general where is the ground-state energy density and is the spacetime volume. Using the above method to we find for general that
(50) |
As a check, for and first order in , we have
(51) |
which agrees with (50). For , is the expectation of the interaction Hamiltonian to in the unperturbed (Gaussian) ground state
(52) |
The calculated higher order connected Green’s function to is
(53) |
where the free propagator is associated with and obeys the equation
(54) |
The solution of is
(55) |
and
(56) |
From (53) it is clear that as , the connected Green’s functions for . This indicates that at least to the theory is noninteracting at . When we consider contributions, we will reexamine this issue.
Turning to and we have
(57) |
and the two-point function in momentum space is
(58) |
where . Thus the renormalised mass to is
(59) |
So near
(60) |
and
(61) |
where . Because of the singularities in and as some renormalisation is needed to remove these infinities. The question is whether perturbative renormalisation can be performed in the context of the novel -expansion method. We can remove the divergence in by introducing in the Lagrangian a linear counter term where has dimension and . Since is real such a term is compatible with -symmetry. Adding also a mass counter term we can consider the Lagrangian density
(62) |
where is dimensionful. Using this Lagrangian we obtain
(63) |
and for the renormalised mass
(64) |
In both expressions we have introduced the dimensionless quantity and as
(65) |
and
(66) |
where is a finite quantity. By setting we have a finite . is logarithmically divergent in . We absorb this divergence into by setting
(67) |
so that and is a finite quantity determined in principle from experiment.
4.2 Calculations to second order in
In second order the calculations become much more involved. The contribution to the connected part of , which we denote by , can be shown to be
(68) |
where denotes for brevity . As ,
(69) |
So the algebraic divergence persists to . The divergence can be removed through . Similarly the divergence persists for at second order. Interestingly, the higher-order Green’s functions continue to vanish for . Avoidance of a noninteracting theory remains an open question in this approach [10].
5 Renormalisation group flows of -symmetric theories
So far we have considered whether a Hermitian theory can lead to a non-Hermitian -symmetric theory due to the effect of renormalisation. We ask the opposite question in this section: Can a -symmetric field theory retain its 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 dimensions this PDE is
(70) |
where and is the surface area of a unit -dimensional sphere. We may assume that the equations for involve dimensionless quantities. If this were not so, we could achieve dimensionlessness by introducing a mass scale . For example, for the dimensionless variables, denoted by a tilde, are , , , and . 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 as a finite series in powers of the field . This ansatz leads to a sequence of coupled nonlinear ordinary differential equations. The consistency of such a procedure has not been established.
For (the quantum-mechanical case), (70) becomes
(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 [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 is greater or less than .
Letting and , we can rewrite (70) as
(72) |
where the subscripts on indicate partial derivatives. We now assume that for large we can neglect the term in the denominator. (The consistency of this assumption is easy to verify when .) Then, for large we have to leading order in our approximation scheme
(73) |
On incorporating a correction to this leading behavior
(74) |
we get to order
(75) |
On making the further change of variable , (75) becomes
(76) |
The variable is positive for and negative for and is not defined at . Thus, (76) is a conventional diffusion equation for but is a backward diffusion equation for . 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 .
In the preliminary study these issues were not addressed further. The case was considered and some simple -symmetric theories were studied [11]. From (70), on defining , we can deduce that
(77) |
From (77) it is consistent to assume that and as . (For simplicity of notation we have dropped the suffix in .) Let us write the correction as . On substituting in (77), we obtain
(78) |
We can proceed in this way and write the next correction as . This leads to where . Repeating this procedure, we get . This procedure can be formalized: On writing (not to be confused with in the last section) and , (77) becomes
(79) |
and .


When is small, the scale is large and we are probing the microscopic potential. As we probe the infrared limit of the effective potential. However, the analysis outlined above was based on a perturbation theory in . There is a parallel to the theory of phase transitions, where a physical entity is expanded in inverse powers of the temperature , and then an extrapolation procedure is used to find the critical behavior at lower . This often-used extrapolation technique is based on Padé approximation.
So far, our treatment has been for a general potential; we will now specialize to two cases and , which are -symmetric for and real. We also include a mass term. Our purpose is to consider the series generated from the series relevant to these two cases and to consider the limit.
5.1 Cubic potentials
We consider both massless and massive cubic potentials. In the former case and in the latter case . For the massless case the solution to (79) is
(80) |
These expressions, which are valid for pure imaginary as well as for real , are cumbersome but manageable. We stress that there has been no truncation of the function space on which 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 have many more terms. For example, the coefficient of is
which in the massless limit reduces to the two-term expression in (80). We refrain from listing the coefficients explicitly and instead proceed directly to the large- behavior of the diagonal Padé approximants. We denote the diagonal Padé approximants in the massless case by and in the massive case by . Let us examine some low-order Padé approximants in the limit . For example,
[As a check, when this expression agrees with the corresponding expression for .] For we obtain
where
[As a check, we have .]
Note that for the massless case has a double pole for both real and imaginary . In fact, the same is true for for even . This pole is an artifact of the Padé approximation and is not present when is odd, so we will only consider the behavior of these approximants for odd . In general, the odd- diagonal Padé approximants have no singularities at all on the real- axis when is imaginary but singularities occur for the case of real . These findings indicate that the -symmetric effective potential is well behaved in the infrared limit . From the expressions for the diagonal Padé approximants we see that for large , the leading behavior of the imaginary part of the effective potential is exactly . Consequently the nature of the interaction is preserved under renormalization.
A similar investigation of quartic potentials also shows that -symmetry is maintained under renormalization [11].
6 Conclusions
The study of -symmetric quantum field theory is still in its infancy. We have shown that there is a link between renormalization of Hermitian field theories and -symmetric field theories. This provides a motivation for the study of -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., 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)