Next Article in Journal
Application of Relative Entropy and Gradient Boosting Decision Tree to Fault Prognosis in Electronic Circuits
Previous Article in Journal
Generalized Preinvex Functions and Their Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Geometrical Basis of 𝒫𝒯 Symmetry

by
Luis L. Sánchez-Soto
1,2,* and
Juan J. Monzón
1
1
Departamento de Óptica, Facultad de Física, Universidad Complutense, 28040 Madrid, Spain
2
Max-Planck-Institut für die Physik des Lichts, Staudtstraße 2, 91058 Erlangen, Germany
*
Author to whom correspondence should be addressed.
Symmetry 2018, 10(10), 494; https://doi.org/10.3390/sym10100494
Submission received: 24 September 2018 / Revised: 6 October 2018 / Accepted: 10 October 2018 / Published: 14 October 2018

Abstract

:
We reelaborate on the basic properties of PT symmetry from a geometrical perspective. The transfer matrix associated with these systems induces a Möbius transformation in the complex plane. The trace of this matrix classifies the actions into three types that represent rotations, translations, and parallel displacements. We find that a PT invariant system can be pictured as a complex conjugation followed by an inversion in a circle. We elucidate the physical meaning of these geometrical operations and link them with measurable properties of the system.

1. Introduction

In quantum theory, the Hamiltonian—and any other real-world observable—is represented by a Hermitian operator. This assumption ensures that the measurement of the energy always yields a real number. In addition, the time-evolution operator generated by a Hermitian Hamiltonian is unitary, which enforces the conservation of probabilities during the evolution of a quantum state.
Nonetheless, non-Hermitian potentials have been used to phenomenologically describe losses [1]. This is the case in open systems, with injecting sources and absorbing sinks, or in systems that decay, e.g., by spontaneous emission of photons from an excited level. The whole system still obeys conventional quantum mechanics; the non-Hermitian Hamiltonian only comes out as an effective subsystem within a projective subspace.
A crucial twist in this field occurred a few years ago, when Bender and coworkers [2,3,4,5,6,7] proposed the use of complex potentials that have neither parity ( P ) nor time-reversal symmetry ( T ), but retain combined PT invariance. They can display real energy eigenvalues, thus suggesting a credible generalization of quantum theory. Actually, by redefining the inner product, the time-evolution operator generated by such potentials could be unitary [8,9,10]. Moreover, they can also exhibit a spontaneous PT symmetry-breaking, at which the reality of the eigenvalues is lost [11,12,13,14].
This concept has prompted a continuing debate in several forefronts, including self-trapped modes [15], quantum field theories [16,17], Anderson localization [18,19,20], complex crystals [21,22,23,24], Lie algebras [25,26,27] and open quantum systems [28], to mention but a few.
The possibility of realizing PT -symmetric potentials led to a flurry of activity in optics, for the paraxial wave equation is fully analogous to Scrödinger equation. The complex refractive index plays here the role of the potential, so they can be accomplished through a sensible inclusion of gain and loss regions. This has been experimentally observed [29]. Besides, PT synthetic materials can exhibit enthralling features; such as power oscillations [30], nonreciprocity of light propagation [31], Bloch oscillations [32], coherent perfect absorbers [33,34], nonlinear switching structures [35], or unidirectional invisibility [36,37,38].
All these issues are classical and, in a broad sense, they are effective models. However, in the quantum regime, Bender proposed two interesting applications related to quantum computation: ultrafast quantum state transformation [39] and quantum state discrimination with single-shot measurement [40]. This also inspired ideas on shortcuts to adiabaticity [41,42], despite some subtleties [43].
Interesting as they are, these developments have one potential criticism: the physical interpretation of PT symmetry remains vague [44,45]. Our purpose here is to put forth a simple but fundamental geometrical characterization of the action of these systems.
To this end we resort to the time-honored transfer-matrix method [46]. Via the Möbius transformation [47,48,49], this matrix induces a rich geometry in the complex plane C ; its trace classifies the actions into three types that represent rotations, translations, or parallel displacements. For the case of PT symmetry, we argue that the transfer matrix may be understood as a point in the de Sitter space [50,51,52] and it may be pictured as a complex conjugation followed by an inversion in a circle. We also explore the physical meaning of these fundamental building blocks.
Our approach does not provide any benefit in terms of efficiency in solving practical problems. Rather, it provides a unifying framework to analyze the behavior of complex potentials, which, in our opinion, is of relevance for the field.

2. Basic Concepts on the Transfer Matrix

Let us first set the stage for our analysis. We consider the scattering of a particle of mass m by a local complex potential V ( x ) defined on R [53,54,55]. The dynamics is determined by the time-independent Schrödinger equation
H Ψ ( x ) = d 2 d x 2 + U ( x ) Ψ ( x ) = ε Ψ ( x ) ,
where ε = 2 m E / 2 and U ( x ) = 2 m V ( x ) / 2 , with E the energy of the particle. We assume that U ( x ) 0 fast enough as x ± , so there exist independent solutions of (1) behaving asymptotically as
Ψ ( x ) = A + e + i k x + A e i k x x , B + e + i k x + B e i k x x + .
Here, k = ε is the wavenumber and the subscripts + and − discriminate right-moving modes exp ( + i k x ) from left-moving modes exp ( i k x ) , respectively.
The exact determination of A ± and B ± amounts to solving (1) with the suitable boundary conditions. This leads to two linear relations among the coefficients A ± and B ± , which can be solved for any amplitude pair in terms of the other two. The transfer matrix [46] corresponds to specifying the relation between the wave amplitudes on both sides of the scatterer, viz,
A + A = M B + B .
Obviously, M depends in a complicated way on the potential U ( x ) . Fortunately, we still can gain insights into the asymptotic behavior without explicitly calculating it. First, we apply (3) successively to a right-moving [ ( A + = 1 , B = 0 ) ] and to a left-moving wave [ ( A + = 0 , B = 1 ) ], both of unit amplitude:
1 r = M t 0 , 0 t r = M r r 1 .
Here, t , r and r , r are the transmission and reflection coefficients for a wave incoming at the potential from the left and from the right, respectively [56].
Equation (4) can also be understood as the superposition of the two independent solutions
Ψ k ( x ) = e + i k x + r ( k ) e i k x x , t ( k ) e + i k x x + , Ψ k r ( x ) = t r ( k ) e i k x x , e i k x + r r ( k ) e + i k x x + .
The wave function Ψ k ( x ) describes a wave incident from [ exp ( + i k x ) ] and the interaction with the potential produces a reflected wave [ r ( k ) exp ( i k x ) ] that escapes to and a transmitted wave [ t ( k ) exp ( + i k x ) ] that moves off to + . The solution Ψ k r ( x ) can be interpreted in a similar fashion.
The Wronskian of the solutions (5) is independent of x. Therefore, we can compute W ( Ψ k , Ψ k r ) = Ψ k Ψ k r Ψ k Ψ k r first for x and then for x . The final results reads
i 2 k W ( Ψ k , Ψ k r ) = t r ( k ) = t ( k ) t ( k ) ,
which ensures that the transmission coefficient is always independent of the direction of the incident wave.
We can now get back to (4) and write the solution for M as (in what follows, to lighten notation, we omit the k dependence of the coefficients)
M = 1 t 1 r r r t 2 r r r .
It is direct to confirm that det M = + 1 , so M SL(2, C ) , the group of 2 × 2 complex matrices with unit determinant [57].

3. Geometry of Transfer Matrices

Let us consider a generic matrix M SL(2, C ), we shall denote by
M = a b c d ,
with a , b , c , d C and det M = a d b c = 1 . The matrix M induces a mapping in C via a Möbius (or linear fractional) transformation [47,48,49]
z = a z + b c z + d .
From a physical perspective, (9) arises because the relevant variables are quotients of amplitudes (such as, e.g., A + / A ) rather than the amplitudes themselves. For each value of z, Equation (9) gives one and only one value of z . There is no exception to this statement if we introduce the point at infinity: if c 0 , z = d / c is transformed into z = and z = into z = a / c . If c = 0 , z = is transformed into z = .
Equation (9) can be inverted, obtaining
z = d z + b c z a ,
which is another Möbius transformation.
The fixed points are those that remain invariant under M ; i.e., z = z . They are given by
z ± = ( a d ) ± [ Tr ( M ) ] 2 4 2 c ,
and determined solely by the trace. We will be mostly interested in the case when Tr ( M ) is a real number, as it will happen for PT invariance. The Möbius transformations are then called hyperbolic, elliptic, or parabolic, according [ Tr ( M ) ] 2 is greater than, lesser than, or equal to 4, respectively.
Equation (9) can be rewritten in a very suggestive form using the fixed points; namely,
z z + z z = K z z + z z ,
where we have taken c 0 (the case c = 0 can be treated similarly). The factor K is called the multiplier and its value is determined by
K + 1 K = [ Tr ( M ) ] 2 2 .
If the quotients appearing in both sides of (12) are designated by Z and Z, respectively, (9) can be compactly recast as Z = K Z , so K determines all its properties. Writing this factor as K = A e i θ (with A > 0 ), it turns out that the aforementioned hyperbolic, elliptic, and parabolic actions appear as
Z = A Z , Z = e i θ Z , Z = Z + c ,
whose interpretation is clear. The hyperbolic action (with fixed points 0 and ) performs a stretching by a factor A 1 . Every straight line through the origin is transformed into itself and any circle with center at the origin is transformed into some other circle with center at the origin. The elliptic action is a rotation about the origin of angle θ . The straight lines and circles of the hyperbolic action have their roles interchanged. Finally, for the parabolic transformation (with only one fixed point), K = 1 and consists in parallel displacements. Every straight line parallel to the translation vector c is invariant. The only fixed point now is at infinity.
We next introduce the concept of the isometric circle [58], which has been used in several applications [59,60] and will play a fundamental role in our later developments. It is defined as the locus of points in the neighborhood of which lengths are unaltered in magnitude by the Möbius transformation. Since
d z d z = 1 ( c z + d ) 2 ,
the isometric circle is ( c 0 )
| c z + d | = 1 .
Similarly, the inverse transformation (10) has the isometric circle
| c z a | = 1 .
The isometric circle of the direct transformation, C d , has its center at O d = d / c and radius R c = 1 / | c | ; the isometric circle of the inverse transformation C i , has its center at O i = a / c and the same radius. Actually, the Möbius transformation carries the isometric circle C d into C i . As a point moves counterclockwise around C d , the corresponding point moves clockwise around C i .
The central result for our purposes is that any Möbius transformation with real trace is equivalent to an inversion in C d followed by a reflection in L [58], which is the perpendicular bisector of the line segment joining the centers of C d and C i . This can be easily intuited from the sketch in Figure 1. To prove it in a quantitative way, we explicitly decompose (9) in two successive steps [59]
z d * = d c * z + | d | 2 1 | c | 2 z + d c , z = a + d c z d * + | a | 2 | d | 2 | c | 2 a + d c * .
The first one, z z d , corresponds to the inversion in the circle C d , whereas the second one, z d z is the reflection in the line L. The same result holds if, instead of inverting in C d and then reflecting in L, we reflect in L and then invert in C i , as it is also indicated in Figure 1.
We recall that if C is a circle with center w and radius r, an inversion in the circle C maps the point ζ into the point ζ along the same radius in such a way that the product of distances from the center w satisfies
| ζ w | | ζ w | = r 2
so that
ζ * = w * + r 2 ζ w .
For the hyperbolic transformations one fixed point is within C d , the other is outside; for the elliptic transformation both fixed points, and for the parabolic transformation the single fixed point, are on C d , as can be seen in Figure 2. Identical statements are true for C i for similar reasons. In the elliptic transformation C d and C i intersect and L is the common chord. The points of intersection are fixed points. In the parabolic transformation, L is the common tangent to C d and C i at their point of tangency, which is precisely the fixed point.
Each one of the transformations we are considering has a one-parameter family of fixed circles, including, as we have discussed, the line joining the centers of C d and C i . These circles are relevant because once we know the initial point z is in one of them, the transformed point z is also in that same circle. The family is easily constructed [58]: it consists of circles with centers on L orthogonal to C d . For, being orthogonal to C d , such a circle is transformed into itself by an inversion in C d ; and a reflection in L, a diameter, transforms it again into itself. Each fixed circle is also orthogonal to C i from symmetry. These families are shown in Figure 1 and Figure 2 for the three kind of transformations.
To interpret the physical significance of the inversion, let us interchange incoming and outgoing fields, which is just reversing the time arrow. Since for a given forward-traveling field A + , the conjugate A + * represents a backward phase-conjugate wave of the original field, the time-reversal operation can be viewed in this context as the transformation
z 1 z * ,
that is, an inversion in the unit circle.

4. Geometry of PT -Invariant Transfer Matrices

We impose now the additional conditions of PT -symmetry on the transfer matrix. We remind that the parity transformation is a space reflection, so that x x and p p . The action is
Ψ ( x ) P Ψ ( x ) .
The time reversal inverts the sense of time, so that x x , p p and i i . The operator T implementing this transformation is antiunitary:
Ψ ( x ) T Ψ * ( x ) .
Therefore, under a combined PT transformation, we have
Ψ ( x ) PT Ψ * ( x ) .
The transfer matrix of the PT -transformed system, we denote by M ( PT ) , has to fulfil
B + * B * = M ( PT ) A + * A * .
By comparing with (3), we get that M ( PT ) = ( M 1 ) * , which implies
M 11 PT M 22 * , M 12 PT M 12 * , M 21 PT M 21 * , M 22 PT M 11 * .
We shall denote henceforth the PT -invariant transfer matrix by M . It must hold then
M 1 = M * ,
which imposes the constraints
Re r t = Re r r t = 0 .
In consequence, the general form of a PT -invariant transfer matrix is
M = x 1 + i x 2 i ( x 3 + x 0 ) i ( x 3 x 0 ) x 1 i x 2 ,
with
x 1 = Re 1 t , x 2 = Im 1 t , x 3 = r r r 2 i t , x 0 = r r + r 2 i t .
The condition det M = 1 gives
x 1 2 + x 2 2 + x 3 2 x 0 2 = 1 ,
so that we can look at M as defining a point in a single-sheeted unit hyperboloid, which is known as the de Sitter space d S 3 [50,51,52].
The fixed points of M are
z ± = x 2 i x 1 2 1 x 3 x 0 .
For hyperbolic actions they are conjugate, whereas for the elliptic case, both points are real.
Let us focus for the time being on this elliptic case, as the other two can be worked out much in the same way. The reflection line L is now the real axis and the associated reflection is then a complex conjugation. To be specific, we shall resort to the simple model of a single one-dimensional PT -symmetric slab of total length h with balanced refractive index n = 3 ± 0.005 i in each half [34]. By adjusting h we can get a typical case of transfer matrix such as, e.g.,
M = 0.7779 0.1263 i 0.2507 i 1.5116 i 0.7779 + 0.1263 i .
The fixed points are z + = 0.3322 and z = 0.4993 , whereas the isometric circles are centered at O d = 0.0836 0.5146 i and O i = 0.0836 + 0.5146 i , with radii R d = R i = 0.6616 .
As one can appreciate in Figure 3, the action of M , mapping the point z onto z , can be equivalently decomposed either in the sequence z z * z (inversion in C d and complex conjugation) or z z * z (complex conjugation and inversion in C i ). Please note that C d and C i are complex conjugate each other.
Let us conclude by treating a particularly interesting example. We take x 0 = x 3 , so we have a lower triangular matrix
M = exp ( i φ ) 0 i 2 x 3 exp ( i φ ) .
Both isometric circles C d and C i pass through the origin, with centers at O d = e i φ / ( i 2 x 3 ) and O i = e i φ / ( i 2 x 3 ) and radius 1 / ( 2 x 3 ) . Again, the mapping z z induced by M can be decomposed as before. When φ = 0 , according to Equation (30), this potential satisfies r = 0 and t = 1 , so it is left invisible.

5. Concluding Remarks

Geometry constitutes the natural arena to formulate numerous physical ideas. In this paper, we have worked out a geometric scenario of PT -invariant systems. More specifically, we have reduced the action of any of these systems to a complex conjugation and an inversion. We have given an interpretation of these actions and expressed them in terms of the physical parameters of the system. The behavior analyzed here is relevant not only in optics, but in all those fields in which the transfer matrix is the method of choice.

Author Contributions

Both authors contributed equally to all aspects of preparing this manuscript.

Funding

Financial support from the Spanish MINECO (Grant No. FIS2015-67963-P) is gratefully acknowledged.

Acknowledgments

We acknowledge illuminating discussions with José María Montesinos.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Muga, J.G.; Palao, J.P.; Navarro, B.; Egusquiza, I.L. Complex absorbing potentials. Phys. Rep. 2004, 395, 357–426. [Google Scholar] [CrossRef]
  2. Bender, C.M.; Boettcher, S. Real spectra in non-Hermitian Hamiltonians having 𝒫𝒯 symmetry. Phys. Rev. Lett. 1998, 80, 5243–5246. [Google Scholar] [CrossRef]
  3. Bender, C.M.; Boettcher, S.; Meisinger, P.N. 𝒫𝒯-symmetric quantum mechanics. J. Math. Phys. 1999, 40, 2201–2229. [Google Scholar] [CrossRef]
  4. Bender, C.M.; Brody, D.C.; Jones, H.F. Complex extension of quantum mechanics. Phys. Rev. Lett. 2002, 89, 270401. [Google Scholar] [CrossRef] [PubMed]
  5. Bender, C.M.; Brody, D.C.; Jones, H.F. Must a Hamiltonian be Hermitian? Am. J. Phys. 2003, 71, 1095–1102. [Google Scholar] [CrossRef]
  6. Bender, C.M. Making sense of non-Hermitian Hamiltonians. Rep. Prog. Phys. 2007, 70, 947–1018. [Google Scholar] [CrossRef] [Green Version]
  7. Bender, C.M.; Mannheim, P.D. 𝒫𝒯 symmetry and necessary and sufficient conditions for the reality of energy eigenvalues. Phys. Lett. A 2010, 374, 1616–1620. [Google Scholar] [CrossRef]
  8. Mostafazadeh, A. Exact 𝒫𝒯-symmetry is equivalent to Hermiticity. J. Phys. A 2003, 36, 7081. [Google Scholar] [CrossRef]
  9. Mostafazadeh, A. Pseudo-Hermiticity and generalized 𝒫𝒯- and 𝒞𝒫𝒯-symmetries. J. Math. Phys. 2003, 44, 974–989. [Google Scholar] [CrossRef]
  10. Wang, Q.H.; Chia, S.Z.; Zhang, J.H. PT-symmetry as a generalization of Hermiticity. J. Phys. A 2010, 43, 295301. [Google Scholar] [CrossRef]
  11. Delabaere, E.; Pham, F. Eigenvalues of complex Hamiltonians with 𝒫𝒯-symmetry. Phys. Lett. A 1998, 250, 25–28. [Google Scholar] [CrossRef]
  12. Klaiman, S.; Günther, U.; Moiseyev, N. Visualization of branch points in 𝒫𝒯-symmetric waveguides. Phys. Rev. Lett. 2008, 101, 080402. [Google Scholar] [CrossRef] [PubMed]
  13. Guo, A.; Salamo, G.J.; Duchesne, D.; Morandotti, R.; Volatier-Ravat, M.; Aimez, V.; Siviloglou, G.A.; Christodoulides, D.N. Observation of 𝒫𝒯-symmetry breaking in complex optical potentials. Phys. Rev. Lett. 2009, 103, 093902. [Google Scholar] [CrossRef] [PubMed]
  14. Levai, G. Spontaneous breakdown of 𝒫𝒯 symmetry in the complex Coulomb potential. Pramana 2009, 73, 329–335. [Google Scholar]
  15. Musslimani, Z.H.; Makris, K.G.; El-Ganainy, R.; Christodoulides, D.N. Optical solitons in 𝒫𝒯 periodic potentials. Phys. Rev. Lett. 100, 030402. [Google Scholar]
  16. Bender, C.M.; Brody, D.C.; Jones, H.F. Extension of 𝒫𝒯-symmetric quantum mechanics to quantum field theory with cubic interaction. Phys. Rev. D 2004, 70, 025001. [Google Scholar] [CrossRef]
  17. Jones, H.F. Equivalent Hamiltonians for 𝒫𝒯-symmetric versions of dual 2D field theories. J. Phys. A 2006, 39, 10123–10132. [Google Scholar] [CrossRef]
  18. Goldsheid, I.Y.; Khoruzhenko, B.A. Distribution of eigenvalues in non-Hermitian Anderson models. Phys. Rev. Lett. 1998, 80, 2897–2900. [Google Scholar] [CrossRef]
  19. Heinrichs, J. Eigenvalues in the non-Hermitian Anderson model. Phys. Rev. B 2001, 63, 165108. [Google Scholar] [CrossRef]
  20. Molinari, L.G. Non-Hermitian spectra and Anderson localization. J. Phys. A 2009, 42, 265204. [Google Scholar] [CrossRef] [Green Version]
  21. Bender, C.M.; Dunne, G.V.; Meisinger, P.N. Complex periodic potentials with real band spectra. Phys. Lett. A 1999, 252, 272–276. [Google Scholar] [CrossRef] [Green Version]
  22. Jones, H.F. The energy spectrum of complex periodic potentials of Kronig–Penney type. Phys. Lett. A 1999, 262, 242–264. [Google Scholar] [CrossRef]
  23. Znojil, M. 𝒫𝒯-symmetric square well. Phys. Lett. A 2001, 285, 7–10. [Google Scholar] [CrossRef]
  24. Ahmed, Z. Energy band structure due to a complex, periodic, 𝒫𝒯-invariant potential. Phys. Lett. A 2001, 286, 231–235. [Google Scholar] [CrossRef]
  25. Bagchia, B.; Quesne, C. sl(2, C) as a complex Lie algebra and the associated non-Hermitian Hamiltonians with real eigenvalues. Phys. Lett. A 2000, 273, 285–292. [Google Scholar] [CrossRef]
  26. Bender, C.M.; Klevansky, S.P. 𝒫𝒯-symmetric representations of fermionic algebras. Phys. Rev. A 2011, 84, 024102. [Google Scholar] [CrossRef]
  27. Cherbal, O.; Trifonov, D.A. Extended 𝒫𝒯- and 𝒞𝒫𝒯-symmetric representations of fermionic algebras. Phys. Rev. A 2012, 85, 05212. [Google Scholar] [CrossRef]
  28. Rotter, I. A non-Hermitian Hamilton operator and the physics of open quantum systems. J. Phys. A 2009, 42, 153001. [Google Scholar] [CrossRef]
  29. Ruter, C.E.; Makris, K.G.; El-Ganainy, R.; Christodoulides, D.N.; Segev, M.; Kip, D. Observation of parity-time symmetry in optics. Nat. Phys. 2010, 6, 192–195. [Google Scholar] [CrossRef]
  30. Makris, K.G.; El-Ganainy, R.; Christodoulides, D.N.; Musslimani, Z.H. Beam dynamics in 𝒫𝒯 symmetric optical lattices. Phys. Rev. Lett. 2008, 100, 103904. [Google Scholar] [CrossRef] [PubMed]
  31. Zheng, M.C.; Christodoulides, D.N.; Fleischmann, R.; Kottos, T. 𝒫𝒯 optical lattices and universality in beam dynamics. Phys. Rev. A 2010, 82, 010103. [Google Scholar] [CrossRef]
  32. Longhi, S. Bloch oscillations in complex crystals with 𝒫𝒯 symmetry. Phys. Rev. Lett. 2009, 103, 123601. [Google Scholar] [CrossRef] [PubMed]
  33. Longhi, S. 𝒫𝒯-symmetric laser absorber. Phys. Rev. A 2010, 82, 031801. [Google Scholar] [CrossRef]
  34. Chong, Y.D.; Ge, L.; Stone, A.D. 𝒫𝒯-symmetry breaking and laser-absorber modes in optical scattering systems. Phys. Rev. Lett. 2011, 106, 093902. [Google Scholar] [CrossRef] [PubMed]
  35. Sukhorukov, A.A.; Xu, Z.; Kivshar, Y.S. Nonlinear suppression of time reversals in 𝒫𝒯-symmetric optical couplers. Phys. Rev. A 2010, 82, 043818. [Google Scholar] [CrossRef]
  36. Ahmed, Z.; Bender, C.M.; Berry, M.V. Reflectionless potentials and 𝒫𝒯 symmetry. J. Phys. A 2005, 38, L627–L630. [Google Scholar] [CrossRef]
  37. Lin, Z.; Ramezani, H.; Eichelkraut, T.; Kottos, T.; Cao, H.; Christodoulides, D.N. Unidirectional invisibility dnduced by 𝒫𝒯-symmetric periodic structures. Phys. Rev. Lett. 2011, 106, 213901. [Google Scholar] [CrossRef] [PubMed]
  38. Longhi, S. Invisibility in 𝒫𝒯-symmetric complex crystals. J. Phys. A 2011, 44, 485302. [Google Scholar] [CrossRef]
  39. Bender, C.M.; Brody, D.C.; Jones, H.F.; Meister, B.K. Faster than Hermitian quantum mechanics. Phys. Rev. Lett. 2007, 98, 040403. [Google Scholar] [CrossRef] [PubMed]
  40. Bender, C.M.; Brody, D.C.; Caldeira, J.; Günther, U.; Meister, B.K.; Samsonov, B.F. 𝒫𝒯-symmetric quantum state discrimination. Philos. Transact. A Math. Phys. Eng. Sci. 2013, 371, 20120160. [Google Scholar] [CrossRef] [PubMed]
  41. Ibáñez, S.; Martínez-Garaot, S.; Chen, X.; Torrontegui, E.; Muga, J.G. Shortcuts to adiabaticity for non-Hermitian systems. Phys. Rev. A 2011, 84, 023415. [Google Scholar] [CrossRef]
  42. Torosov, B.T.; Della Valle, G.; Longhi, S. Non-Hermitian shortcut to adiabaticity. Phys. Rev. A 2013, 87, 052502. [Google Scholar] [CrossRef] [Green Version]
  43. Lee, Y.C.; Hsieh, M.H.; Flammia, S.T.; Lee, R.K. Local 𝒫𝒯 symmetry violates the no-signaling principle. Phys. Rev. Lett. 2014, 112, 130404. [Google Scholar] [CrossRef] [PubMed]
  44. Weigert, S. The physical interpretation of 𝒫𝒯-invariant systems. Czech J. Phys. 2004, 54, 1139–1142. [Google Scholar] [CrossRef]
  45. Jin, L.; Song, Z. A physical interpretation for the non-Hermitian Hamiltonian. J. Phys. A 2011, 44, 375304. [Google Scholar] [CrossRef]
  46. Sánchez-Soto, L.L.; Monzón, J.J.; Barriuso, A.G.; Cariñena, J. The transfer matrix: A geometrical perspective. Phys. Rep. 2012, 513, 191–227. [Google Scholar] [CrossRef] [Green Version]
  47. Needham, T. Visual Complex Analysis; Oxford University Press: Oxford, UK, 1997. [Google Scholar]
  48. Anderson, J.W. Hyperbolic Geometry; Springer: New York, NY, USA, 1999. [Google Scholar]
  49. Ratcliffe, J.G. Foundations of Hyperbolic Manifolds; Springer: Berlin, Germany, 2006. [Google Scholar]
  50. O’Neill, B. Semi-Riemannian Geometry with Applications to Relativity; Academic Press: London, UK, 1983. [Google Scholar]
  51. Moschella, U. The de Sitter and anti-de Sitter sightseeing tour. In Proceedings of the 20th Seminaire Poincaré, Paris, France, 21 November 2005; pp. 1–12. [Google Scholar]
  52. Monzón, J.J.; Barriuso, A.G.; Montesinos-Amilibia, J.M.; Sánchez-Soto, L.L. Geometrical aspects of 𝒫𝒯-invariant transfer matrices. Phys. Rev. A 2013, 87, 012111. [Google Scholar] [CrossRef]
  53. Mostafazadeh, A. Spectral Singularities of Complex Scattering Potentials and Infinite Reflection and Transmission Coefficients at Real Energies. Phys. Rev. Lett. 2009, 102, 220402. [Google Scholar] [CrossRef] [PubMed]
  54. Cannata, F.; Dedonder, J.P.; Ventura, A. Scattering in 𝒫𝒯-symmetric quantum mechanics. Ann. Phys. 2007, 322, 397–433. [Google Scholar] [CrossRef]
  55. Ahmed, Z. New features of scattering from a one-dimensional non-Hermitian (complex) potential. J. Phys. A 2012, 45, 032004. [Google Scholar] [CrossRef]
  56. Boonserm, P.; Visser, M. One dimensional scattering problems: A pedagogical presentation of the relationship between reflection and transmission amplitudes. Thai J. Math. 2010, 8, 83–97. [Google Scholar]
  57. Mostafazadeh, A.; Mehri-Dehnavi, H. Spectral singularities, biorthonormal systems and a two-parameter family of complex point interactions. J. Phys. A 2009, 42, 125303. [Google Scholar] [CrossRef] [Green Version]
  58. Ford, L.R. Automorphic Functions; AMS Chelsea Publishing: New York, NY, USA, 1972. [Google Scholar]
  59. Bolinder, E.F. Impedance and Power Transformations by the Isometric Circle Method and Non-Euclidean Hyperbolic Geometry; Technical Report; MIT: Cambridge, MA, USA, 1957. [Google Scholar]
  60. Rudolph, J.G.; Cheng, D.K. Isometric-circle interpretation of bilinear transformation and its application to VSWR minimization. Radio Sci. 1965, 69D, 1271–1283. [Google Scholar] [CrossRef]
Figure 1. (Left) The isometric method for a Möbius transformation of the hyperbolic type. We have plotted the isometric circles for the direct ( C d ) and the inverse ( C i ) transformations, with centers in O d and O i , respectively. The fixed points are marked in green and L is the reflection line. z and z are the points related by M , whereas z d is the point obtained by an inversion of z respect to C d and z d the reflected of z by L. The inversion of z d respect to C i gives z . (Right) Two typical families of fixed lines for the same transformation.
Figure 1. (Left) The isometric method for a Möbius transformation of the hyperbolic type. We have plotted the isometric circles for the direct ( C d ) and the inverse ( C i ) transformations, with centers in O d and O i , respectively. The fixed points are marked in green and L is the reflection line. z and z are the points related by M , whereas z d is the point obtained by an inversion of z respect to C d and z d the reflected of z by L. The inversion of z d respect to C i gives z . (Right) Two typical families of fixed lines for the same transformation.
Symmetry 10 00494 g001
Figure 2. Families of fixed lines for elliptic (left) and parabolic (right) Möbius transformations. We show the isometric circles C d and C i and their centers O d and O i . The fixed points are marked in green.
Figure 2. Families of fixed lines for elliptic (left) and parabolic (right) Möbius transformations. We show the isometric circles C d and C i and their centers O d and O i . The fixed points are marked in green.
Symmetry 10 00494 g002
Figure 3. The isometric method as in Figure 2, but now applied to an elliptic PT -invariant transfer matrix M given in (33). We show again the isometric circle for the direct ( C d ) and the inverse ( C i ) transformations, with centers in O d and O i , respectively. The symmetry line L coincides with the real axis, so the corresponding reflection reduces to a complex conjugation.
Figure 3. The isometric method as in Figure 2, but now applied to an elliptic PT -invariant transfer matrix M given in (33). We show again the isometric circle for the direct ( C d ) and the inverse ( C i ) transformations, with centers in O d and O i , respectively. The symmetry line L coincides with the real axis, so the corresponding reflection reduces to a complex conjugation.
Symmetry 10 00494 g003

Share and Cite

MDPI and ACS Style

Sánchez-Soto, L.L.; Monzón, J.J. The Geometrical Basis of 𝒫𝒯 Symmetry. Symmetry 2018, 10, 494. https://doi.org/10.3390/sym10100494

AMA Style

Sánchez-Soto LL, Monzón JJ. The Geometrical Basis of 𝒫𝒯 Symmetry. Symmetry. 2018; 10(10):494. https://doi.org/10.3390/sym10100494

Chicago/Turabian Style

Sánchez-Soto, Luis L., and Juan J. Monzón. 2018. "The Geometrical Basis of 𝒫𝒯 Symmetry" Symmetry 10, no. 10: 494. https://doi.org/10.3390/sym10100494

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop