arxiv: v1 [hep-th] 30 Oct 2018

Similar documents
Heuristic Explanation of the Reality of Energy in PT -Symmetric Quantum Field Theory

arxiv: v1 [quant-ph] 22 Jun 2012

PT-symmetric quantum mechanics

PT-symmetric quantum systems

PT-symmetric interpretation of double scaling in QFT

Department of Physics, Washington University, St. Louis, MO 63130, USA (Dated: November 27, 2005)

arxiv: v2 [math-ph] 26 Feb 2017

Tunneling in classical mechanics

arxiv: v1 [hep-th] 28 May 2009

Summary of free theory: one particle state: vacuum state is annihilated by all a s: then, one particle state has normalization:

REVIEW REVIEW. Quantum Field Theory II

Quantum Field Theory II

Latest results on PT quantum theory. Carl Bender Washington University

arxiv: v1 [hep-th] 5 Jun 2015

Review of scalar field theory. Srednicki 5, 9, 10

arxiv: v1 [quant-ph] 19 Mar 2015

Complex WKB analysis of energy-level degeneracies of non-hermitian Hamiltonians

Recursive calculation of effective potential and variational resummation

Euclidean path integral formalism: from quantum mechanics to quantum field theory

arxiv:physics/ v3 [math-ph] 6 May 1998

PT-symmetric quantum theory, nonlinear eigenvalue problems, and the Painlevé transcendents

arxiv: v1 [physics.optics] 8 Nov 2011

Real Spectra in Non-Hermitian Hamiltonians Having PT Symmetry

arxiv:quant-ph/ v1 29 Mar 2003

Quantum Field Theory II

The C operator in PT -symmetric quantum field theory transforms as a Lorentz scalar

Light-Cone Quantization of Electrodynamics

1 The Quantum Anharmonic Oscillator

The 3 dimensional Schrödinger Equation

Path integral in quantum mechanics based on S-6 Consider nonrelativistic quantum mechanics of one particle in one dimension with the hamiltonian:

Regularization Physics 230A, Spring 2007, Hitoshi Murayama

arxiv:math-ph/ v1 10 May 2000

Loop corrections in Yukawa theory based on S-51

Graviton contributions to the graviton self-energy at one loop order during inflation

PT-symmetric quantum theory

Section 9 Variational Method. Page 492

Geometric Aspects of Space-Time Reflection Symmetry in Quantum Mechanics

THE IMAGINARY CUBIC PERTURBATION: A NUMERICAL AND ANALYTIC STUDY JEAN ZINN-JUSTIN

Carl M. Bender. Theoretical Physics, Blackett Laboratory, Imperial College, London SW7 2BZ, UK

Comment on On the Lagrangian and Hamiltonian description of the damped linear harmonic oscillator

arxiv: v1 [nlin.ps] 5 Aug 2014

Entire functions, PT-symmetry and Voros s quantization scheme

Attempts at relativistic QM

arxiv: v2 [cond-mat.dis-nn] 23 Jun 2017

Renormalization-group study of the replica action for the random field Ising model

4 4 and perturbation theory

4 Power Series Solutions: Frobenius Method

Maxwell s equations. electric field charge density. current density

Page 404. Lecture 22: Simple Harmonic Oscillator: Energy Basis Date Given: 2008/11/19 Date Revised: 2008/11/19

Generalized PT symmetry and real spectra

Summation Techniques, Padé Approximants, and Continued Fractions

Week 5-6: Lectures The Charged Scalar Field

Quantum Field Theory Homework 3 Solution

arxiv: v1 [hep-ph] 30 Dec 2015

Physics 221A Fall 2017 Notes 27 The Variational Method

d 1 µ 2 Θ = 0. (4.1) consider first the case of m = 0 where there is no azimuthal dependence on the angle φ.

Harmonic Oscillator I

arxiv: v1 [physics.optics] 21 Dec 2015

Computation of the scattering amplitude in the spheroidal coordinates

LSZ reduction for spin-1/2 particles

REVIEW REVIEW. A guess for a suitable initial state: Similarly, let s consider a final state: Summary of free theory:

Classical Particles Having Complex Energy Exhibit Quantum-Like Behavior

Quantum dynamics with non-hermitian PT -symmetric operators: Models

Supporting Information

arxiv: v2 [math-ph] 13 Dec 2012

Donoghue, Golowich, Holstein Chapter 4, 6

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

Harmonic Oscillator with raising and lowering operators. We write the Schrödinger equation for the harmonic oscillator in one dimension as follows:

Total Angular Momentum for Hydrogen

Physics 70007, Fall 2009 Answers to Final Exam

Vacuum Energy and Effective Potentials

PT -symmetric quantum electrodynamics

Second quantization: where quantization and particles come from?

Evaluation of Triangle Diagrams

1 Quantum fields in Minkowski spacetime

Final reading report: BPHZ Theorem

Lecture notes for QFT I (662)

Physics 127b: Statistical Mechanics. Renormalization Group: 1d Ising Model. Perturbation expansion

PY 351 Modern Physics - Lecture notes, 3

Symmetry Factors of Feynman Diagrams for Scalar Fields

PT-SYMMETRIC INTERPRETATION of OPEN QUANTUM and CLASSICAL SYSTEMS

Alex Eremenko, Andrei Gabrielov. Purdue University. eremenko. agabriel

Application of Resurgence Theory to Approximate Inverse Square Potential in Quantum Mechanics

UNIVERSITY OF SURREY FACULTY OF ENGINEERING AND PHYSICAL SCIENCES DEPARTMENT OF PHYSICS. BSc and MPhys Undergraduate Programmes in Physics LEVEL HE2

arxiv: v1 [quant-ph] 28 May 2018

Physics 250 Green s functions for ordinary differential equations

Ginzburg-Landau Theory of Type II Superconductors

I. Perturbation Theory and the Problem of Degeneracy[?,?,?]

Time Domain Modeling of All-Optical Switch based on PT-Symmetric Bragg Grating

2 Quantization of the scalar field

Analogous comments can be made for the regions where E < V, wherein the solution to the Schrödinger equation for constant V is

arxiv: v1 [quant-ph] 11 Jun 2009

arxiv:math-ph/ v1 30 Sep 2003

1 Mathematical preliminaries

1 Solutions in cylindrical coordinates: Bessel functions

1 Equal-time and Time-ordered Green Functions

arxiv: v2 [quant-ph] 18 May 2018

Lecture-05 Perturbation Theory and Feynman Diagrams

TWISTOR DIAGRAMS for Yang-Mills scattering amplitudes

221A Lecture Notes Convergence of Perturbation Theory

Transcription:

PT -symmetric quantum field theory in D dimensions Carl M. Bender a,c, Nima Hassanpour a, S. P. Klevansky b, and Sarben Sarkar c a Department of Physics, Washington University, St. Louis, Missouri 633, USA b Institut für Theoretische Physik, Universität Heidelberg, 69 Heidelberg, Germany c Department of Physics, King s College London, London WCR LS, UK arxiv:8.479v [hep-th] 3 Oct 8 PT -symmetric quantum mechanics began with a study of the Hamiltonian H = p + x (ix). A surprising feature of this non-hermitian Hamiltonian is that its eigenvalues are discrete, real, and positive when. This paper examines the corresponding quantum-field-theoretic Hamiltonian H = ( φ) + φ (iφ) in D-dimensional spacetime, where φ is a pseudoscalar field. It is shown how to calculate the Green s functions as series in powers of directly from the Euclidean partition function. Exact finite expressions for the vacuum energy density, all of the connected n-point Green s functions, and the renormalized mass to order are derived for D <. For D the one-point Green s function and the renormalized mass are divergent, but perturbative renormalization can be performed. The remarkable spectral properties of PT -symmetric quantum mechanics appear to persist in PT -symmetric quantum field theory. I. INTRODUCTION The study of PT -symmetric quantum theory may be traced back to a series of papers that proposed a new perturbative approach to scalar quantum field theory. Instead of a conventional perturbation expansion in powers of a coupling constant, it was proposed that a parameter δ that measures the nonlinearity of the theory could be used as a perturbation parameter [, ]. Thus, to solve a gφ 4 field theory we studied a gφ (φ ) δ theory and treated the parameter δ as small. The procedure was to obtain a perturbation expansion in powers of δ and then to set δ = to obtain the results for the gφ 4 theory. Detailed investigation showed that this perturbative calculation is numerically accurate and does not require the coupling constant g to be small [, ]. An important feature of this approach was that φ and not φ had to be raised to the power δ in order to avoid raising a negative quantity to a noninteger power, thereby generating complex numbers as an artifact of the procedure. This δ expansion was also used to solve nonlinear classical differential equations of physics [3]: the Thomas-Fermi equation (nuclear charge density) y (x) = [y(x)] 3/ / x becomes y (x) = y(x)[y(x)/x] δ ; the Lane- Emdon equation (stellar structure) y (x) + y (x)/x + [y(x)] n = becomes y (x) + y (x)/x + [y(x)] +δ ; the Blasius equation (fluid dynamics) y (x) + y (x)y(x) = becomes y (x)+y (x)[y(x)] δ = ; the Korteweg-de Vries equation (nonlinear waves) u t + uu x + u xxx = becomes u t + u δ u x + u xxx =. In each case the quantity raised to the power δ is positive and when δ = the equation becomes linear. Also, these δ expansions have a nonzero radius of convergence and are numerically accurate. PT -symmetric quantum mechanics began with the surprising discovery that spurious complex numbers do not appear if the quantity raised to the power δ is PT symmetric (invariant under combined space and time reflection) [4, 5]. This fact is highly nontrivial and was totally unexpected. Indeed, the eigenvalues of the non- Hermitian PT -symmetric Hamiltonian H = p + x (ix) ( ) () are entirely real, positive, and discrete when because ix is PT invariant. A proof that the spectrum is real when > was given by Dorey, Dunning, and Tateo [6, 7]. Numerous PT -symmetric model Hamiltonians have been studied at a theoretical level [8] and many laboratory experiments have been performed on PT -symmetric physical systems [9 ]. The purpose of this paper is to introduce powerful new tools and techniques that can be used to investigate PT -symmetric quantum field theories. We illustrate these tools by studying the quantum-field-theoretic analog of () whose D-dimensional Euclidean-space Lagrangian density is L = ( φ) + φ (iφ) ( ), () where φ is a pseudoscalar field so that L is PT invariant. We treat as small and show how to calculate the vacuum energy density E, the connected n-point Green s functions G n, and the renormalized mass M R as series in powers of. In this paper we assume that D < to avoid the appearance of renormalization infinities and then we comment briefly on the perturbative renormalization procedure for the case D. To first order in ( << ), the unusual Lagrangian density L in () has a logarithmic self-interaction term: Electronic address: cmb@wustl.edu Electronic address: nimahassanpourghady@wustl.edu Electronic address: spk@physik.uni-heidelberg.de Electronic address: sarben.sarkar@kcl.ac.uk L = ( φ) + φ + φ log(iφ) + O( ). (3) For a quantum field theory having a complex logarithmic interaction term it is not obvious whether one can find

Feynman rules for performing perturbative diagramatic calculations. We will show how to construct such Feynman rules. We begin by replacing the complex logarithm with a real logarithm and we do so in such a way as to preserve PT symmetry; to wit, we define and we define log(iφ) i + log(φ) (if φ > ) log(iφ) i + log( φ) (if φ < ). Combining these two equations, we make the replacement log(iφ) = i φ /φ + log(φ ). (4) Note that in (4) the imaginary part is odd in φ and the real part is even in φ. Thus, (4) enforces the PT symmetry because the pseudoscalar field φ changes sign under parity P and i changes sign under time reversal T. [To derive (4) we must assume that φ is real. The reality of φ is explained in Sec. II.] Graphical techniques were developed in Ref. [] to handle real logarithmic interaction terms. These techniques are generalizations of the replica trick [], which has been used in the study of spin glasses. The idea of the replica trick is that a logarithmic term log( A can be reformulated as the limit log A = lim N N A N ), or equivalently and slightly more simply as the limit d log A = lim N dn AN. (5) One then regards N as an integer and identifies A N as an N-point vertex in a graphical expansion. Of course, this procedure is not rigorous because it requires taking the continuous limit N. The validity of this approach has not been proved, but when it is possible to compare with exactly known results in low-dimensional theories, the replica trick gives the correct answer. In this paper we verify our field-theoretic results by comparing them with the exact answers for D = (where the functional integral becomes an ordinary integral) and for D = (quantum mechanics). The graphical calculations in this paper are done in coordinate space. Once the vertices have been identified, all that one needs is the free propagator in D-dimensional Euclidean space (x y), which satisfies the differential equation ( x + ) (x y) = δ (D) (x y). (6) Taking the Fourier transform of this equation gives the amplitude for the free propagator of a particle of mass in momentum space: (p) = /(p + ). The D-dimensional inverse Fourier transform of this expression then gives the D-dimensional coordinate-space propagator in terms of an associated Bessel function: (x x ) = () D x x D K D ( x x ). (7) If we let x x, we obtain the amplitude () for a self loop, which is the amplitude for a line to originate from and return to the same point: () = (4) D/ ( D/). (8) This expression is finite and nonsingular for D <. This paper is organized as follows. We calculate the ground-state energy density E in Sec. II, the one-point Green s function G in Sec. III, the two-point Green s function G and the renormalized mass M R in Sec. IV, and the general connected n-point Green s function G n in Sec. V, all to first order in. These quantities are finite when D <, but the three quantities E, G, and M R diverge when D so it is necessary to introduce a renormalization procedure. In Sec. VI we discuss the issues of renormalization. We show that a redefinition of the energy scale, an additive shift in the field, and a mass counterterm eliminate these infinities. In this section we also discuss our future calculational objectives, namely, calculating the Green s functions to higher order in. II. FIRST-ORDER CALCULATION OF THE GROUND-STATE ENERGY DENSITY If we expand the partition function Z() = Dφ e d D x L (9) for the Lagrangian density L to first order in and use (4), the functional integral (9) becomes ( Z() = Dφ e d D x L d D y { iφ(y) φ(y) 4 +φ (y) log [ φ (y) ]} ) + O( ), () where the free Lagrangian density L = ( φ) + φ () is obtained by setting = in L. Note that the imaginary part of the functional integrand in () is odd in φ, so Z() is real. We emphasize that the functional integration in () is performed along the real-φ axis and not in the complex-φ domain. This justifies the use of (4). We are not concerned here with complex functional integration paths that terminate in complex Stokes sectors because the functional integral () converges term-by-term in powers of. In PT -symmetric quantum mechanics the boundary conditions on the Schrödinger equation associated with the Hamiltonian () [see (4)] are imposed in complex Stokes sectors [5]. However, in the context of quantum field theory it would be hopelessly unwieldy to consider functional Stokes sectors. This is why we treat as small. This paper is concerned with calculating the coefficients

3 in the series and we do not consider here the mathematical issues involved with the summation of such a series for large values of. In general, a partition function is the exponential of the ground-state energy density E multiplied by the volume of spacetime V : Z = e EV. Thus, the shift in the ground-state energy density E to order is given by E = 4Z()V Dφ e d D x L d D y φ (y) log [ φ (y) ]. Hence, from (5) we obtain d E = lim Dφ e d D x L d D y φ N (y). N 4Z()V dn This expression has a graphical interpretation as the product of N self loops from the spacetime point y back to y. There are exactly (N )!! ways to construct these self loops, so the expression for E simplifies to E = lim N 4V d dn d D y [ ()] N (N )!!. Next, we note that the integral d D y is the volume of spacetime V, so this formula simplifies further: E = lim N 4 d dn [ ()]N (N )!!. Finally, we use the duplication formula for the gamma function [3] to write (N )!! = N ( N + ) / and then take the derivative with respect to N to get E = 4 () { log[ ()] + ( 3 )/ ( 3 )}, () where ( 3 )/ ( 3 ) = γ log. The result in () may be verified for the special cases of D = and D =. Special case D = : For D = the normalized partition function becomes an ordinary integral, which we can expand to first order in : Z = dφ e φ (iφ) Z dφ e φ[ Z φ log(iφ) ], (3) where Z =. If we take the negative logarithm of this result, we obtain the first-order shift in the groundstate energy density E = dφ e φ φ log(iφ). We then integrate separately from to and from to and combine the two integrals to obtain a single real integral that we evaluate as follows: E = dφ e φ φ log φ = 4[ ( 3 )/ ( 3 ) + log ]. (4) Taking D = in (8) gives () =, so () reduces exactly to (4) in zero-dimensional spacetime. Special case D = : In quantum mechanics, E to leading order in is the expectation value of the interaction Hamiltonian H I = x log(ix) [see (3)] in the unperturbed ground-state eigenfunction ψ (x) = exp ( x) : E = / dx e x x log(ix) dx e x = 8 ( 3 )/ ( 3 ). (5) Taking D = in (8) gives () =, so the general result in () reduces exactly to (5) in one-dimensional spacetime. Note that in PT -symmetric quantum mechanics the calculation of expectation values requires the C operator [8]. However, the C operator is not needed for any of the calculations in this paper because a functional integral involves vacuum expectation values. The vacuum state is an eigenstate of the C operator with eigenvalue, C =, so all reference to C disappears. This simplification was first pointed out in Ref. [4]. III. FIRST-ORDER CALCULATION OF THE ONE-POINT GREEN S FUNCTION The one-point Green s function G is nonperturbative in character but it can be calculated by following the approach used above to calculate E. Keeping terms that do not vanish under φ φ, we evaluate directly the functional-integral representation G (a) = Dφ φ(a)e d D x L 4Z() d D y iφ(y) φ(y), (6) where L is the free Euclidean Lagrangian in (). We use the integral identity φ φ = φ dt t sin(tφ) (7) to replace φ(y) φ(y) in the functional integral (6) and then we replace sin(tφ) by its Taylor series: sin(tφ) = n= ( ) n t n+ φ n+. (8) (n + )! This converts (6) to the product of an infinite sum in n, a one-dimensional integral in t, a D-dimensional integral in y, and a functional integral in φ: G (a) = i ( ) n t n dt d D y t= (n + )! n= Dφ e d D x L φ(a)[φ(y)] n+3. (9) Z()

4 The sum and multiple integrals in (9) may appear to be difficult, but like the calculation of the ground-state energy density, the functional integral in the second line of (9) also has a graphical interpretation; it is merely the product of the free propagator (a y) representing a line from a to y multiplied by n + self loops from y to y, and this product is accompanied by the combinatorial factor (n + 3)!!. Thus, the second line in (9) reduces to (n + 3)!! (y a) n+ (). This result simplifies further because, as we can see from (6), the D-dimensional integral is trivial: d D y (y a) =. This establishes the translation invariance of G. The rest is straightforward: G = i = i t= t= dt n= ( ) n t n n+ ()(n + 3)!! (n + )! dt () [ 3 ()t ] e ()t = i ()/. () This expression for G is exact to order. Observe that the expression for G is a negative imaginary number. This is precisely what we would expect based on previous studies of classical PT -symmetric systems. The classical trajectories in complex coordinate space of a particle described by the Hamiltonian () are left-right symmetric but they lie mostly in the lower-half plane [8]. Thus, the average value of the classical orbits is a negative-imaginary number. Special case D = : The one-point Green s function in D = is given by the second line in (3) with an additional extra factor of φ: G Z() dφ φe φ[ φ log(iφ) ], where Z() =. The integration over the first term in the square brackets vanishes by oddness. We evaluate the contribution of the second term by integrating first from to and then from to. Combining these two integrals, we obtain G = i /. This result is in exact agreement with the general result in () because () = when D =. Special case D = : When D =, the expression for G in () reduces to G = i. () In the Appendix we derive G in quantum mechanics and verify (). IV. FIRST-ORDER CALCULATION OF THE TWO-POINT GREEN S FUNCTION To obtain the connected two-point Green s function G (a, b), one must subtract G from the vacuum expectation value of φ(a)φ(b). However, we have seen that G is of order. Therefore, to first order in we need only evaluate Z() in () with φ(a)φ(b) inserted after Dφ and then divide this integral by Z(). We may neglect the imaginary terms in these integrals because they are odd under φ φ. Thus, the expression that we must evaluate for G (a, b) is Dφ φ(a)φ(b)e { d D x L 4 d D y φ (y) log [ φ (y) ]} Dφ e { d D x L 4 dd y φ (y) log [ φ (y) ]}. We expand this expression to first order in as a sum of three functional integrals: G (a, b) = A + B + C, () where A = Dφ φ(a)φ(b)e d D x L, Z() B = Dφ φ(a)φ(b)e d D x L 4Z() d D y φ (y) log [ φ (y) ], C = Dφ φ(a)φ(b)e d D x L 4Z() Dφ e d D x L d D y φ (y) log [ φ (y) ]. (3) Z() We must now evaluate the three contributions A, B, and C. The functional integral A in (3) is simply the free propagator (a b). This result verifies that if we set = in the Lagrangian (), we obtain a free field theory; the two-point Green s function for such a field theory is (a b). The double functional integral C presents a complication. The first line of C is proportional to A and evaluates to 4 (a b). However, as we showed in the calculation of the ground-state energy, the next two lines of C evaluate to 4 V E, where E is the first-order shift in the ground-state energy density () and V is the volume of Euclidean spacetime. Thus, C = (a b)v E, and this quantity is divergent because V is infinite. We resolve this divergence problem by calculating B. Using (5) we express the B integral as B = 4Z() lim d Dφ e d D x L N dn d D y φ(a)φ(b)φ N (y). (4) This functional integral requires that we connect in a pairwise fashion the set of N + points consisting of a, b, and the N points y with the free propagator in (7). There are two cases to consider. In the first case a is connected to b and the remaining N points at y are connected in pairs. Note that this reproduces the result above for C except with the opposite sign. Thus, the volume divergences exactly cancel.

5 In the second case a is not connected to b. Instead, a connects to a point y (there are N ways to do this) and b connects to one of the remaining N points y (there are N ways to do this). The rest of the N points y are joined in pairs [there are (N 3)!! ways to do this]. The amplitude for this case is d D d y (a y) (y b) lim N dn N(N )!! N (), which simplifies to K d D y (a y) (y b), where K = + ( ( 3 ) / 3 ) + log[ ()] = 3 γ + log [ ()]. (5) Thus, our final result for the coordinate-space twopoint Green s function to order is G (a b) = (a b) K d D y (a y) (y b). (6) In momentum space this is G (p) = p + K (p + ) + O( ). (7) From (7) we construct the (, ) Padé approximant, which is just the geometric sum of a chain of bubbles: G (p) = p + + K + O ( ). (8) We then read off the square of the renormalized mass to first order in : M R = + K + O ( ). (9) Special case D = : To verify (8) and (9) for the case D = we evaluate the ordinary one-dimensional integrals in G = dx x e x (ix) / dx e x (ix) / dx x e [ x / 4 x log(x ) ] = dx e x / [ 4 x log(x ) ]. These integrals are not difficult and to order we get G = + ( 3 γ (3) log ). This result agrees exactly with that in (8) for D =. Special case D = : We verify (9) for the case D = by calculating the energy level of the first excited state of the Hamiltonian H in (). When =, H becomes the harmonic-oscillator Hamiltonian, so the first excited state eigenfunction is xe x / and the associated eigenvalue is 3. We solve the time-independent Schrödinger equation Hψ = Eψ perturbatively by substituting ψ(x) = xe x / + ψ (x) + O ( ), E = 3 + E + O ( ), (3) and collecting powers of. To first order in the function ψ (x) satisfies ψ (x) + x ψ (x) 3 ψ (x) = [ x log(ix) + E ] xe x /. To solve this equation we use the technique of reduction of order and let ψ (x) = xe x / f(x). The equation for f(x) is then xf (x) + ( x ) f(x) = E x + x 3 log(ix). Multiplying this equation by the integrating factor x exp ( x ) gives d [ x f (x)e x] = [ E x + x 4 log(ix) ] e x. dx Therefore, if we integrate this equation from to, we obtain an equation for E : E = 4 dx x 4 log ( x ) e x/ dx x e x, where we have replaced log(ix) by log ( x ). These integrals are easy to evaluate and we get ( E = 3 8 8 3 γ log ). Thus, the first excited eigenvalue of H to order is 3 + 3 8 ( 8 3 γ log ). (3) We have already calculated the ground-state energy in Sec. II: + 8( γ log ). (33) The renormalized mass M R is the first excitation above the ground state so M R is the difference of these two energies: M R = + 4(3 γ log ). (34) If we square this result and keep terms of order, we get M R = + ( 3 γ log ), (35) which exactly reproduces (9) for the case D =. V. HIGHER-ORDER GREEN S FUNCTIONS The connected three-point Green s function is given by the cumulant G 3 (x, y, z) = φ(x)φ(y)φ(z) Z φ(x)φ(z) φ(y) Z + φ(x) φ(y) φ(z) Z 3. φ(x)φ(y) φ(z) Z φ(y)φ(z) φ(x) Z

6 However, to order VI. DISCUSSION AND FUTURE WORK G 3 (x, y, z) = φ(x)φ(y)φ(z) (x y)g Z (x z)g (y z)g. The calculation of G 3 is somewhat tedious, but the procedure follows exactly the calculation of the two-point Green s function. The final result after the disconnected terms have canceled is G 3 (x, y, z) = i () d D u (x u) (y u) (z u). (36) The connected four-point function is defined by the cumulant G 4 (x, y, z, w) = φ(x)φ(y)φ(z)φ(w) Z φ(x)φ(y)φ(z) φ(w) Z (three permutations) φ(x)φ(y) φ(z)φ(w) Z (two permutations) + φ(x)φ(y) φ(z) φ(w) Z 3 + (five permutations) 6 φ(x) φ(y) φ(z) φ(w) Z 4. Again, calculating G 4 is tedious but the result is simply G 4 (x, y, z, w) = d D u (x u) () (y u) (z u) (w u). (37) The pattern is now evident and with some effort we can calculate the connected n-point Green s function to order and obtain a general formula valid for all n : G n (x, x,... x n ) = ( i)n ( n ) [ ()] n/ n d D u ( x k u ). (38) k= Thus, the connected Green s functions are all of order except for G in (6), which is of order. Observe that (38) reduces to (), (36), and (37) for n =, 3, and 4. We emphasize that (38) holds for both odd and even n. Our calculation of the odd-n Green s functions uses the first term on the right side of (4) and our calculation proceeds by introducing the integral representation in (7) followed by using the Taylor series in (8). On the other hand, our calculation of the even-n Green s functions uses the second term on the right side of (4) and the calculation proceeds by applying the derivative identity in (5). It is satisfying that these two strikingly different techniques lead to the single universal formula in (38) for G n. The principal advance reported in this paper is that we have developed all of the machinery necessary to calculate the Green s functions of a PT -symmetric quantum field theory in () to any order in. Thus, this paper opens a vast area for future study and investigation; one can investigate the masses of the theory (the poles of the Green s functions), scattering amplitudes, critical indices, and so on. The Green s-function calculations done in Secs. II-V are exact to first order in. However, the procedures presented here immediately generalize to all orders in. Furthermore, as we show below, even in low orders the perturbation series in powers of is highly accurate and it continues to be accurate for large. To illustrate, we calculate the one-point Green s function G in D = to second order in. In D = this Green s function is a ratio of two ordinary integrals: G = dx x exp [ ( x + L + L )] dx exp [ x ( + L) ], (39) where L = log(ix) = i x /x + log ( x ). Evaluating these integrals is straightforward and the result is [ G = i + 4 (γ 3 log ) + O ( )] [ ( = i.8756 + O )]. (4) To check of the accuracy of (4) we calculate the onepoint Green s function for a cubic theory ( = ). We convert the expansion in (4) to a [, ] Padé approximant, G = i +.8756, and then set = to obtain the approximate result G =.668i. The exact value of G for the zero-dimensional cubic theory = is given by the ratio of integrals G = dx x exp ( ix3) dx exp ( ix3) = i (/3) 3 (/3) =.6369i. Thus, the two-term expansion (4) has an accuracy of 5%, which is impressive for such a large value of. This good result is consistent with the results found in previous studies of the accuracy of the expansion for various classical equations (see Ref. [3]). The third-order version of (4) is G = i = i [ + 4(γ 3 log ) + ( 9 54 log + 44 log 36γ log + 6γ 48γ + 48 ) + O ( 3)] [.8756 +.6447 + O ( 3)].

7 Converting the expansion above to a [, ] Padé approximant and setting =, we obtain the result that G =.63i, which now differs from the exact result by only %. These numerical results strongly motivate us to extend our studies of the expansion of PT - symmetric quantum field theories to higher order in. We will publish the higher-order calculations in a future paper. A second issue that that needs to be examined in depth is that of renormalization. Because () becomes singular when the dimension D of Euclidean spacetime reaches, the one-point Green s function G and the renormalized mass M R become singular. (To first order in the higher Green s functions do not become infinite when D =.) Thus, for D we must undertake a perturbative renormalization procedure. For simplicity, in this paper we have worked entirely with dimensionless quantities. However, to carry out a perturbative renormalization one must work with the Lagrangian L = ( φ) + µ φ + gµ φ ( iµ D/ φ ) for which the dimensional parameters are explicit: µ is the unrenormalized mass µ, µ is a fixed parameter having dimensions of mass, and g is a dimensionless unrenormalized coupling constant. The mass renormalization procedure consists of expressing the renormalized mass M R in terms of these Lagrangian parameters and absorbing the divergence that arises when D into parameter µ. The coupling-constant renormalization procedure is similar; we define the renormalized coupling constant G R as the value of the three-point or four-point Green s functions at particular values of the external momentum and again absorb the divergence that arises into the Lagrangian parameter g. One must then verify that all higher Green s functions are finite when expressed in terms of M R and G R. This program will be carried out explicitly in a future paper. Acknowledgments CMB thanks the Heidelberg Graduate School of Fundamental Physics for its hospitality. VII. APPENDIX In this Appendix we calculate the one-point Green s function G for the case D = (quantum mechanics). In quantum mechanics G is the expectation value of the operator x in the ground state. Thus, if ψ(x) is the (unnormalized) ground-state eigenfunction in coordinate space, we can express G as the ratio of integrals G x = dx xψ (x)/ dx ψ (x). (4) For the quantum-mechanical Hamiltonian () the ground-state eigenfunction obeys the time-independent Schrödinger equation ψ (x) + x (ix) ψ(x) = Eψ(x), (4) where E is the ground-state energy. To first-order in this differential equation becomes ψ (x) + x ψ(x) + x log(ix)ψ(x) = Eψ(x). (43) We solve this equation perturbatively by substituting ψ(x) = ψ (x) + ψ (x) + O ( ), E = E + E + O ( ), (44) where ψ (x) = exp ( x), E =, and as derived in (5), E = 8 ( 3 )/ ( 3 ). Collecting powers in, we see that ψ (x) and ψ (x) satisfy the differential equations ψ (x) + x ψ (x) ψ (x) =, ψ (x) + x ψ (x) ψ (x) = x log(ix)ψ (x) + E ψ (x). (45) The equation for ψ (x) is an inhomogenous version of the equation for ψ (x). Thus, we use reduction of order to solve the ψ (x) equation; we substitute and obtain ψ (x) = ψ f(x) = e x f(x) (46) f (x) xf (x) = x log(ix) E. We then multiply this equation by the integrating factor exp( x ) and integrate from to x: x f (x)e x = C + ds [ s ] log(is) E e s. The integration constant C vanishes because E = Thus, we find that ds s log(is)e s/ ds e s. ψ (x) = e x t x dt e t ds [ s ] log(is) E e s. (47) We can now evaluate the integrals in (4) to first order in. This expression simplifies considerably because of the factor of x: G = t x dx xe x dt e t ds [ s log(is) E ] e s. (48)

8 To evaluate this integral we integrate from x = to and then from x = to and combine the two integrals. The resulting integral simplifies further because the logarithm in the integrand collapses to log( ) = i: The r integral can now be done, and we obtain a sum of two double integrals: G = i x dx xe x dt e t ds s e s. (49) Evaluating this triple integral is not trivial, but by using some tricks it can be calculated exactly and in closed form. We begin by interchanging the order of the s and t integrals: t G = i [ /4 dθ cos θ sin θ + / dθ cos θ sin θ /4 cos θ sin θ dz ( z ) 3 ] dz ( z ) 3. (5) G = i min(s,x) dx xe x ds s e s dt e t. (5) Next, we introduce polar coordinates x = r cos θ and s = r sin θ and make the change of variable t = rz: G = i / dr r 5 dθ cos θ sin θ min(sin θ,cos θ) dz e r z r. (5) These double integrals may be evaluated by using any algebraic manipulation code such as Mathematica. The final result is G = i. (53) This verifies the general result in () for the case D =. [] C. M. Bender, K. A. Milton, M. Moshe, S. S. Pinsky, and L. M. Simmons, Jr., Phys. Rev. Lett. 58, 65 (987). [] C. M. Bender, K. A. Milton, M. Moshe, S. S. Pinsky, and L. M. Simmons, Jr., Phys. Rev. D 37, 47 (988). [3] C. M. Bender, K. A. Milton, S. S. Pinsky, and L. M. Simmons, Jr., J. Math. Phys. 3, 447 (989). [4] C. M. Bender and S. Boettcher, Phys. Rev. Lett. 8, 543 (998). [5] C. M. Bender, S. Boettcher, and P. N. Meisinger, J. Math. Phys. 4, (999). [6] P. E. Dorey, C. Dunning, and R. Tateo, J. Phys. A: Math. Gen. 34, 5679 (). [7] P. E. Dorey, C. Dunning, and R. Tateo, J. Phys. A: Math. Theor. 4, R5 (7). [8] For a review of some of the early work see C. M. Bender, Rep. Prog. Phys. 7, 947 (7). [9] J. Rubinstein, P. Sternberg, and Q. Ma, Phys. Rev. Lett. 99, 673 (7). [] A. Guo, G. J. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. A. Siviloglou, and D. N. Christodoulides, Phys. Rev. Lett. 3, 939 (9). [] C. E. Rüter, K. G. Makris, R. El-Ganainy, D. N. Christodoulides, M. Segev, and D. Kip, Nat. Phys. 6, 9-95 (). [] K. F. Zhao, M. Schaden, and Z. Wu, Phys. Rev. A 8, 493 (). [3] Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao, and D. N. Christodoulides, Phys. Rev. Lett. 6, 39 (). [4] L. Feng, M. Ayache, J. Huang, Y.-L. Xu, M. H. Lu, Y. F. Chen, Y. Fainman, and A. Scherer, Science 333, 79 (). [5] J. Schindler, A. Li, M. C. Zheng, F. M. Ellis, and T. Kottos, Phys. Rev. A 84, 4(R) (). [6] S. Bittner, B. Dietz, U. Günther, H. L. Harney, M. Miski- Oglu, A. Richter, and F. Schäfer, Phys. Rev. Lett. 8, 4 (). [7] N. Chtchelkatchev, A. Golubov, T. Baturina, and V. Vinokur, Phys. Rev. Lett. 9, 545 (). [8] C. Zheng, L. Hao, and G. L. Long, Phil. Trans. R. Soc. A 37, 53 (3). [9] C. M. Bender, B. Berntson, D. Parker, and E. Samuel, Am. J. Phys. 8, 73 (3). [] B. Peng, Ş. K. Özdemir, F. Lei, F. Monifi, M. Gianfreda, G. L. Long, S. Fan, F. Nori, C. M. Bender, and L. Yang, Nat. Phys., 394 (4). [] S. Assawaworrarit, X. Yu, and S. Fan, Nature 546, 387 (7). [] M. Mezard, G. Parisi, and M. Virasoro, Spin Glass Theory and Beyond (World Scientific, Singapore, 987). [3] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions (Dover, New York, 97). [4] H. F. Jones and R. J. Rivers, Phys. Rev. D 75, 53 (7).