ArticlePDF Available

Multiply Degenerate Exceptional Points and Quantum Phase Transitions

Authors:

Abstract and Figures

The realization of a genuine phase transition in quantum mechanics requires that at least one of the Kato's exceptional-point parameters becomes real. A new family of finite-dimensional and time-parametrized quantum-lattice models with such a property is proposed and studied. All of them exhibit, at a real exceptional-point time t=0, the Jordan-block spectral degeneracy structure of some of their observables sampled by Hamiltonian H(t) and site-position Q(t). The passes through the critical instant t=0 are interpreted as schematic simulations of non-equivalent versions of the Big-Bang-like quantum catastrophes.
Content may be subject to copyright.
arXiv:1412.6634v1 [quant-ph] 20 Dec 2014
Multiply Degenerate Exceptional Points
and Quantum Phase Transitions
Denis I. Borisov,
Institute of Mathematics CS USC RAS, Chernyshevskii str., 112, Ufa, Russia, 450008
and
Bashkir State Pedagogical University, October Rev. st., 3a, Ufa, Russia, 450000
e-mail: BorisovDI@yandex.ru
and
Frantiˇsek Ruˇziˇcka and Miloslav Znojil
Nuclear Physics Institute ASCR, Hlavn´ı 130, 250 68 ˇ
Reˇz, Czech Republic
e-mail: fruzicka@gmail.com and znojil@ujf.cas.cz
1
Abstract
The realization of a genuine phase transition in quantum mechanics requires that at least one of
the Kato’s exceptional-point parameters becomes real. A new family of finite-dimensional and
time-parametrized quantum-lattice models with such a property is proposed and studied. All
of them exhibit, at a real exceptional-point time t= 0, the Jordan-block spectral degeneracy
structure of some of their observables sampled by Hamiltonian H(t) and site-position Q(t). The
passes through the critical instant t= 0 are interpreted as schematic simulations of non-equivalent
versions of the Big-Bang-like quantum catastrophes.
2
1 Introduction
Although the concept of phase transition originates from classical thermodynamics, it recently
found a new area of applicability in the context of quantum mechanics where one may decide
to work with the pseudo-Hermitian (the term used in mathematics, see reviews [1, 2]) alias
PT symmetric (using the terminology of physicists, see review [3]), manifestly non-Hermitian
Hamiltonians which are still capable of generating a stable, unitary evolution of the quantum
system in question.
A key phenomenological novelty is that whenever our quantum Hamiltonian H6=Hvaries
with a real parameter (say, H=H(λ)), the “physical intervals” of the acceptability of λare,
typically, different from the whole real axis. In the mathematical language of Kato [4] one can
also say that for the analytic operator functions H(λ) of the parameter, the whole interval (or
rather a union of intervals) lies in an open complex set Dand that the (in general, complex) points
of its boundary Dcoincide with the Kato’s exceptional points (EPs) as introduced in loc. cit..
The main mathematical feature of the EPs λ(EP )
jDis that the limiting operators H(EP )=
limλλ(EP )
j
H(λ) cease to be tractable as physical Hamiltonians because even if the spectrum of
the energies happens to remain real these operators cease to be diagonalizable. In a suitable basis
they acquire a Jordan-blocks-containing canonical form [1, 5, 6]. Thus, even if one of values of
λ(EP )
jis real, the limiting transition λλ(E P )
jmust still be perceived as a process during which
certain eigenvectors of H(λ) “parallelize” and become linearly dependent, resulting in the loss of
the usual probabilistic tractability of the quantum system in question [7] - [13].
It is worth adding that in the vast majority of the standard applications of quantum mechanics
the EP parameters λ(EP )
jare not real so that one cannot “cross” them when a parameter moves
just along the real line. In such a case, as a rule, the boundaries of the domain of unitarity coincide
with the boundaries of the Riemann surface Rof the analyticity of the operator function H(λ)
so that one must speak about an end of the applicability of the Hamiltonian rather than about a
quantum phase transition as realized within the same physical model.
In our present paper we intend to pay attention to the scenarios in which the EP values
λ(EP )
jDare real and do not lie on any boundary of R. In such an arrangement (observed
already, in [14], in 1998) one may treat the EPs as the points of a true analytic realization of
quantum phase transitions.
2 Pure quantum states in parallel representations
2.1 Three Hilbert space pattern
Whenever one reveals that the spectrum of a non-Hermitian operator of an observable of a quantum
system (be it a Hamiltonian H6=Hor a position operator Q6=Q, etc) is all real and discrete, one
feels tempted to expect that a smooth theoretical translation of all predictions into the language
3
of the textbook quantum mechanics may be performed [15].
One of the realizations of such a conceptually appealing project has been proposed in nuclear
physics (cf. [16]). According to our recent summaries [17, 18] of this possibility one simply
has to make use of the simultaneous representation of the quantum states in three alternative
Hilbert spaces H(F,S,P ). Out of the triplet the first space (viz., H(F), often chosen in the form of
L2(R)) is most friendly. Naturally, whenever this space leaves our observables non-Hermitian, it
must be declared unphysical. Fortunately, in the second, standard and physical space H(S)the
Hermitization of the same operators of observables may be comparatively easily achieved via the
mere introduction of an amended, metric-mediated (i.e., often, non-local) inner product.
In the whole scheme, the latter space is finally (and, usually, constructively) declared unitarily
equivalent to the third Hilbert space H(P)with trivial metric which, as a rule, appears prohibitively
complicated and inaccessible to any constructive fructification [16].
2.2 Five Hilbert space pattern
The generic technical nontriviality of the whole three-Hilbert-space (THS) parallel-representation
recipe may force its users to apply the trick twice. More details may be found in [19]. In essence,
the first, preparatory application yields just a tentative, simplified inner-product metric ΘT, the
study of which may be motivated, e.g., by the generic difficulty of the verification of the reality
of the spectrum or of the closed-form construction of any less elementary amendments of the
tentative positive-definite metric. Subsequently, with ΘT= Θ
T6=Iat our disposal, the final
specification of the second, “sophisticated” metric ΘSis expected to be guided by the redirection
of emphasis from mathematics to phenomenology.
The following diagram characterizes the resulting five-Hilbert-space representation of a given
4
quantum system,
(initial stage)
friendly space H(F)
friendly observable F6=F
false metric Θ(F)=I
(1)the left map,preparatory
(aim :simplicity)ւ ց
(2)the right,correct Dysons map
(aim :real world )
(intermediate result)
test space H(T)
F=F= Θ1
TFΘT
trial ΘT=
TT6=I
spectral reality proof
6=
(final result)
standard space H(S)
F=F= Θ1
SFΘS
correct ΘS=
SS6=I
experimental predictions
ւր respective unitary equivalences ցտ
(mathematical reference)
auxiliary space H(A)
(math.)
f(math.)= TF1
T=f
(math.)
(trivial metric)
nonequivalent outcomes
(physical reference)
prohibited space H(P)
(phys.)
f(phys.)=f
(phys.)of textbooks,
isospectral to F .
(1)
Typically, one starts from an observable Hamiltonian F=Hor position F=Q(etc) which are
all defined in H(F). One then moves towards the first nontrivial (though still unphysical) positive
definite artificial metric candidate ΘT(say, to prove the reality of the spectrum). In the second
step of construction one proceeds to the ultimate analysis of the standard representation of the
system using a realistic, physical ΘSwhich is often known solely in an approximate form [20].
3 Exceptional points in a schematic model
In the context of our present paper the key purpose of the start of analysis from an operator
of observable Fwhich is defined in an unphysical Hilbert space H(F)(i.e., which is manifestly
non-Hermitian there and which will be mostly sampled by the operator of position Qin what
follows) is that such operators often possess the real exceptional points [21] - [23].
3.1 Traditional studies starting from a Hamiltonian
In the traditional considerations one usually assumes that in the vicinity of a real Kato’s EP value
of parameter λ(EP )
jD Rthe basis in the Hilbert space H(F)(where the superscript (F)stands
5
for “friendly” [18]) is such that the whole diagonalizable part of a pre-determined Hamiltonian
H(λ) is diagonalized. Thus, at λ(E P )
jone may only pay attention to the non-diagonalizable part
of the Hamiltonian. The latter operator may be also assumed to have acquired the standard
canonical form of a direct sum of Jordan blocks. Naturally, one could further restrict attention
just to one of the Jordan blocks (of matrix dimension N), considering its small vicinity in the
following special form of the multiparametric finite- and tridiagonal-matrix toy model with an
additional up-down symmetry of its matrix elements,
H(toy)=
0 1 α0 0 0 0 ... 0 0 0
α0 1 β0 0 0 ... 0 0 0
0β0 1 γ0 0 ... 0 0 0
0 0 γ0 1 δ0... 0 0 0
0 0 0 δ..........
.
..
.
..
.
.
.
.
..
.
..
.
.......0 1 δ0 0 0
0 0 0 ... 0δ0 1 γ0 0
0 0 0 ... 0 0 γ0 1 β0
0 0 0 ... 000 β0 1 α
0 0 0 ... 0 0 0 0 α0
.(2)
During our preparatory numerical experiments with the spectra of various Hamiltonian matrices
of the generic tridiagonal form (2) (cf. also Refs. [24, 25]) it became clear that the transitions
of quantum systems through their Jordan-block alias multiple-EP (or degenerate-EP) quantum-
phase-transition points may prove to be a phenomenologically relevant and interesting process.
In fact, there emerged no true surprises in the context of mathematics where one simply
observed differing patterns of behavior (and, in particular, of the complexification) of certain
eigenvalues before and after the EP singularity. Unfortunately, once we started thinking about
time tas parameter (with its EP value, say, at t= 0), we immediately imagined that certain
purely formal nontriviality of the three-Hilbert-space (THS) formalism would force us to accept
the restriction to adiabatically slow changes of the system in general and of the Hamiltonian
operator H(t) in particular (cf. [18] for details).
3.2 An amended strategy of analysis based on a given site operator ˜
Q
The unpleasant necessity of the slowness of changes of H(t) seems hardly compatible with the
abrupt nature of the changes of the system near an EP singularity at t= 0. For methodical reasons
it seem reasonable, therefore, to replace the study of the time-dependent Hamiltonians H(t) (which
combine the role of the operators of an observable energy with a partially independent role of the
6
generators of time evolution) by the formally similar but conceptually less confusing study of
some other operators sharing the same perturbed-Jordan-block matrix form but representing, say,
a non-Hermitian spin Σ(t) [26] - [28] or, better, the observable Q(t) of discretized position alias
site in a finite quantum lattice (cf. [29] - [31]).
Our present strategy of the build-up of the theory initiated by the site operator ˜
Qmight have
been supported not only by the above-mentioned circumvention of the problems with adiabatic
approximation (which are basically technical) but also by a few other, physics-oriented arguments.
One of them is based on the observation [32] that whenever one interprets the one-parametric
family of eigenvalues qn(t) of a site operator ˜
Qchosen in virtually any form of a perturbed
Ndimensional Jordan block, then the unfolding pattern of these eigenvalues (very well sampled
by their particular special case qn(t)cntas derived, in Refs. [33] - [35], for an exactly solvable
discrete and PT symmetric anharmonic oscillator) resembles the cosmological phenomenon of
Big Bang. Indeed, in loc. cit., all of these eigenvalues stayed complex (i.e., unobservable) before
the Big-Bang instant t= 0 while all of them became observable (i.e., real and non-degenerate and
even, in absolute value, growing quickly) immediately after the Big-Bang time t= 0.
4 The phase-transition interpretation of the real excep-
tional points
In our present paper let us finalize the definition of our family of models (with Hreplaced by Q
in Eq. (2)) in such a way that all of the components of the input Jplet of variable parameters
~
ξ={α,...,ω}={ξ1, ξ2, . . . , ξJ}of Eq. (2) become either proportional to an absolute value of
time (thus, we shall have the even-function-superscripted components ξj=ξ(e)
j(t) = |t|defined at
all real tR) or proportional to the plain time (for the remaining, odd-function-superscripted
components ξk=ξ(o)
k(t) = t,tR). This means that every eligible quantum model of our family
(with J=entier(N/2) in general) may be naturally characterized by the respective superscripts
in vector ξ, i.e., by a word of length Jin the two-letter alphabet {o, e}(i.e., one has two eligible
words 0= (o) and 1= (e) at J= 1, four candidates 0= (oo), 1= (oe), 2= (eo) and
3= (ee) at J= 2, etc).
Let us now add that in accord with the specific implementations of the three-Hilbert-space
pattern of Refs. [16, 18] the key technical task of its users should be seen to lie in the reconstruction
of a suitable metric ΘSfrom a given site-position matrix Q(N)
()(t). At the not too large and positive
times t > 0 such a construction is, due to our choice of tridiagonal Q(N)
()(t), recurrent and entirely
routine [36].
4.1 Spectra at times close to the Big Bang instant
At an illustrative matrix dimension N= 10, the first nontrivial sample of our present lattice-site
operator (2) will contain five free parameters α,... alias ξ1, ξ2,...,ξ5which we decided to
7
–0.6
–0.4
–0.2
0
0.2
0.4
0.6
0.1
0.05
0
–0.05
–0.1 t
q(t)
Figure 1: The time-dependence of the real part of the spectrum of the perturbed-Jordan-block
ten-by-ten matrix (3) with index 1= (ooooe).
choose in the first nontrivial concrete form of quintuplet t, t, t, t, |t|. As we indicated above, this
choice is encoded in the word 1= (ooooe) yielding the matrix
Q(10)
(1)(t) =
0 1 t0 0 0 0 0 0 0 0
t0 1 t0 0 0 0 0 0 0
0t0 1 t0 0 0 0 0 0
0 0 t0 1 t0 0000
0 0 0 t0 1 |t|0000
0 0 0 0 |t|0 1 t000
0 0 0 0 0 t0 1 t0 0
0 0 0 0 0 0 t0 1 t0
0 0 0 0 0 0 0 t0 1 t
0 0 0 0 0 0 0 0 t0
.(3)
This matrix represents one of many possible t6= 0 perturbations of the 10 times 10 Jordan block
to which it degenerates at the EP time t= 0.
At the small but non-vanishing times t6= 0 the tdependence of the real eigenvalues of matrix
(3) is displayed in Fig. 1. Once we accept the physical interpretation of such a matrix as a site-
position operator of a dynamical ten-point quantum lattice near its phase transition instant t= 0
of the Big-Bang type, we may conclude that at the positive times t > 0, i.e., after the Big Bang
instant this operator has the whole spectrum real and, hence, it may be perceived as representing
an observable quantity.
We should emphasize that in contrast to an analogous model with 0= (ooooo) (possessing just
an empty real spectrum to the left from t= 0), there now exists a pair of eigenvalues q(t) which
still remain real also before the Big Bang. Thus, Fig. 1 may be perceived as a schematic sample
of an EP-interrupted evolution in which just a simpler, two-site part of our ten-point quantum
8
–0.6
–0.4
–0.2
0
0.2
0.4
0.6
–0.1
–0.05
0
0.05
0.1 t
q(t)
Figure 2: The time-dependence of the real part of the spectrum of the perturbed-Jordan-block
ten-by-ten matrix with index 7= (ooeee).
lattice may be interpreted as observable at t < 0. In other words, one could read the message
delivered by Fig. 1 as opening the possibility of an innovative, quantum cosmology resembling
scenario in which a subspace-based, simpler, two-level observable Universe evolved from the left,
i.e., during the previous Eon and towards its own t= 0 Big Crunch collapse.
4.2 Alternative changes of physics during the pass of our quantum
lattices through the EP time t= 0
In spite of a truly elementary form of our family of the Big-Bang-simulating toy models Q(N)
()(t),
their capability of simulation and variability of alternative Big-Crunch/Big-Bang quantum phase-
transition scenarios seems truly inspiring. In particular, the N= 10 Fig. 1 obtained at =1=
(ooooe) may be complemented by its analogue of Fig. 2 with 7= (ooeee) where the six levels
remain real before the Big Crunch instant, etc.
Naturally, a fully left-right symmetric picture would be obtained at 31 = (eeeee), representing
a schematic ten-dimensional-Universe counterpart to the well known (though very differently sup-
ported and constructed) Penrose’s cyclic-cosmology pattern in which the structure of the Universe
before and after the Big Bang singular time is not expected to be too different [37]. In contrast,
our present family of models may be understood as supporting a rather non-standard hypothesis
of an “evolutionary” cosmology in which the complexity of the structure of the Universe may
change during the EP singularity and, in principle, grow with time from some more elementary
structures existing during the previous Eons.
Needless to add that within the similar “evolutionary cosmology” phenomenological specu-
lations, the structure of the Universe during the newer Eons (including also, e.g., its spatial
dimensionality, etc) could have been also enriched by certain newly emergent qualities, the ob-
servability (i.e., the reality of the corresponding eigenvalues) of which had been suppressed earlier,
emerging only on a sufficiently sophisticated level of the iterative global cosmic evolution.
Once we now return back to the elementary mathematical level of the concrete study of our
9
schematic quantum models, we still have to address a number of multiple less ambitious questions
like, e.g, the purely technical problem of the necessity of the modification of the definition of the
“old” physical Hilbert spaces which had to exist and describe our quantum lattice before t= 0.
In fact, at the finite dimensions N < such a modification remains mathematically more or
less trivial. Indeed, at t < 0 one merely has to preserve (i.e., project out) just the vector space
which is spanned by the eigenvectors of Q(t) with the real eigenvalues. The resulting algorithmic
pattern will partially resemble Eq. (1) in having the very similar five-Hilbert space form of the
diagram in which the left and right triplet of windows represents our quantum lattice alias discrete
Universe before and after the Big Bang, respectively:
Big Bang instant plus its vicinity :
the same friendly space H(F)
Q(0) not diagonalizable
beforeBigBang map
t < 0ւ ց afterBigBang map
t > 0
less,N<N observable sites
reduced space H(R)
reduced QR=Q
R= Θ1
RQ
RΘR
old timer ΘR=
RR6=I
=NNghosts projected out
6=
all N sites
standard space H(S)
Q=Q= Θ1
SQΘS
current metric ΘS=
SS6=I
our Eon
ւր respective unitary equivalences ցտ
Hermitian reference
previous Eon space H(P)
(old)
q(old)= RQR1
R=q
(old)
extinct physics
quantum phase transition
Hermitian reference
this Eon space H(P)
(now)
“our′′ coordinates q(now)=q
(now)
contemporary physics .
(4)
Naturally, the extension of such a recipe and diagram to the infinte-dimensional Hilbert space
limit N=will lead to multiple further open questions in functional analysis. As long as such
a step already lies fairly beyond the scope of our present paper, let us merely conjecture that in
the infinite-dimensional Hilbert-space limit and in the previous Eon” with t < 0, the necessary
elimination of the “ghost-supporting” spurious subspace (in which the eigenvalues of Q(t) are not
yet real) might proceed in an analogy with the Mostafazadeh’s elimination trick of Ref. [38].
10
5 Discussion
Our present paper was inspired by the recent enormous popularity of the building of quantum mod-
els in which the unitary evolution is guaranteed and controlled, paradoxically, by non-Hermitian
Hamiltonians H6=H. On this background our attention was shifted from the traditional
PT symmetric alias pseudo-Hermitian Hamiltonians (such that one has H=PHP16=H
in terms of an indeterminate pseudometric P) to the other, less prominent operators of observ-
ables. We emphasized that the choice of some other observables might enrich, first of all, our
insight in some less understood processes in which the quantum system in question is forced to
move through an exceptional-point singularity.
5.1 Considerations inspired by the open problems in physics
For the sake of definiteness we paid attention to certain specific time-dependent site-position
operators Q(t) which were chosen, for the sake of mathematical simplicity, in the form of finite
matrices. In parallel, in a way emphasizing their phenomenological appeal we choose our Nby N
toy-model matrices Q(N)(t) in such a form that at the EP value of time t= 0 they degenerated to an
Ndimensional and quantum-phase-transiton inducing Jordan-block non-diagonalizable matrix
Q(N)(0).
We emphasized that any operator Qrepresenting an observable quantity admits in fact the
same mathematical treatment as the most often considered energy operator alias Hamiltonian. In
particular, once we wish to declare any operator observable in H(S), the equivalent requirement
of the Hermiticity of its image in H(P)must be imposed, say, in the form
q= SQ1
S=q.
This condition may be also equivalently reformulated, inside the second physical space H(S), as
follows,
Q(tS(t) = ΘS(t)Q(t).(5)
In full analogy with the standard THS recipe, one can start, therefore, from any family of the
benchmark-model forms of the operator Qand perform the reconstruction of the admissible metrics
ΘSvia Eq. (5).
In the context of physics we revealed that after an appropriate further sophistication and de-
velopment our initial idea of studying the phase-transition pass through the Big-Bang-resembling
EP singularity of Q(N)(t) at t= 0 could find multiple future applications even in quantum cos-
mology. We conjectured that the traditional alternative scenarios of “nothing before Big Bang”
and of the “cyclic repetition of Big Bangs” (e.g., in the form as proposed by Penrose [37]) could
be, in this light, complemented by the slightly subtler possibilities and speculations about having
“Darwin-like evolution” and various “physics-structure jumps” at the subsequent Big Bangs.
11
Figure 3: The norm of the resolvent R(z) = [zQ(10)
(t)]1with = (eooee) after Big Bang, at
positive t= 0.1.
5.2 Considerations inspired by the open problems in mathematics
Naturally, all of the above-sampled phenomenological speculations suffer from the insufficiently
realistic and oversimplified finite-dimensional alias discrete-lattice nature of our toy-model observ-
ables Q(N)
()(t). In this sense, the main opening mathematical challenge (not to be addressed here
at all) may be now seen in the study of N extensions of the discrete-lattice models.
Another set of the open mathematical problems emerges even at the finite matrix dimensions
N < . One of the most important ones concerns the problem of the influence of perturbations
on any toy-model based qualitative result. In the conclusion of this paper let us now pay more
attention to this fairly important subproblem.
Our basic inspiration has been provided by the experience covered by the Trefethen’s and Em-
bree’s book [39] with its extreme emphasis on the constructive considerations and with its rather
persuasive recommendation that the influence of the perturbations should be always characterized
in the language of the so called pseudospectra [40].
An easy and compact introduction in the concept of the pseudospectrum may be provided here
either by the reference to loc. cit. or to Fig. 3. In the latter picture one sees that the purely real
spectrum (i.e., the real-eigenvalue positions znRof the infinitely high peaks of the norm |R(z)|
of the resolvent) may be perceived as very well approximated by the 0 < ε 1 pseudospectra
(which are defined, according to loc. cit., as the - in general, multiply connected - open complex
domains Jεof zin which |R(z)|>1).
12
Figure 4: The norm of the resolvent of Fig. 3 before Big Bang, at negative t=0.1.
In contrast to the after Big Bang picture as provided by Fig. 3, a conceptually much more
challenging problem emerges at the negative, pre-Big-Bang times at which some of the eigenvalues
of matrix Q(N)
()(t) remain non-real. As we already mentioned above, we are not going to address
this challenge directly but rather we intend to recall the Trefethen’s and Embree’s advice.
Although, in their book, the pseudospectra are predominantly recommended for an estimate
of the influence of a generic perturbation, the role of their study in our present context will be
different. Indeed, in the light of the diagram of Eq. (4) one of the key tasks of the constructive
approach to the quantitative description of our present versions of the quantum phase transitions
lies in the possibility of a clear separation of the “quantum observables before Big Bang“ (i.e., of
a subspace which is spanned by the real-eigenvalue eigenvectors of Q(t)) from the perpendicular
subspace of the unobservable, non-real-eigenvalue “ghosts”.
In practice, naturally, the separation of this type (which, incidentally, resembles strongly the
Gupta-Bleuler trick known from textbooks on quantum electrodynamics) must be performed nu-
merically. This implies that one must specify the domains of applicability of a feasible separation
of this type.
In this context, our present final recommendation is to use the inspection of the pseudospectra
for the purpose. For illustration, let us first recall Figs. 4 (with its equivalent form 5) in which
one sees an entirely distinct numerical separation of the states with the real eigenvalues from the
ghosts. In such a dynamical scenario (with =19 = (eooee)) one may expect that the projector-
operator elimination of the ghosts from the physical before-the-Big-Bang reduced Hilbert space
13
−0.4 0 Re z
−0.4
0
Im z
Figure 5: The real function of Fig. 4 in its two-dimensional representation. The lines marking the
constant norm are the boundaries of the complex domains called “pseudospectra” [39].
−0.4 Re z 0.4
−0.5
Im z
0.5
Figure 6: A rearrangement of the pseudospectra of Fig. 5 at another index, = (eoooe).
14
−0.5 Re z 0.5
−0.3
Im z
0.3
Figure 7: Another set of the before-the-Big-Bang pseudospectra, with = (eoeee).
H(R)will be numerically feasible and straightforward.
In contrast, the inspection of another, =17 = (eoooe) toy model may be expected to
indicate an emergence of the technical and numerical difficulties because the related Fig. 6 shows
that up to the very small εs, the pseudospectral domains Jεof zare shared by more peaks and do
not sufficiently clearly separate each of the two real eigenvalues from its two complex conjugate
neighbors. At the same time, the four “remote” complex eigenvalues still seem to be separated in
a sufficiently clear manner.
From the perturbation-influence-estimate point of view, by far the worst situation is encoun-
tered at = (eoeee) where our last Fig. 7 indicates that and why the separation of the eight-
dimensional ghost subspace may be expected to be truly difficult.
Acknowledgements
D.B. was partially supported by grant of RFBR, grant of President of Russia for young scientists-
doctors of sciences (MD-183.2014.1) and Dynasty fellowship for young Russian mathematicians.
15
References
[1] A. Mostafazadeh, Int. J. Geom. Meth. Mod. Phys. 7 (2010) 1191.
[2] E. Caliceti and S. Graffi, in “Nonselfadjoint operators in quantum physics: mathematical
aspects.” (F. Bagarello et al, Eds., Wiley, Hoboken, 2015, p. 183), in print.
[3] C. M. Bender, Rep. Prog. Phys. 70 (2007) 947.
[4] T. Kato, Perturbation theory for linear operators. (Springer-Verlag, Berlin, 1966).
[5] M. V. Keldysh, Russian Math. Surveys 26 (1971) 15.
[6] D. Borisov, Acta Polytechnica 54 (2014) 93.
[7] W. D. Heiss, J. Phys. A: Math. Theor. 45 (2012) 444016.
[8] Y. N. Joglekar, C. Thompson, D. D. Scott and G. Vemuri, Eur. Phys. J. Appl. Phys. 63
(2013) 30001.
[9] A. Tanaka, S. W. Kim and T. Cheon, Phys. Rev. E 89 (2014) 042904.
[10] J. Fuchs, J. Main, H. Cartarius and G. Wunner, J. Phys. A: Math. Theor. 47 (2014) 125304.
[11] F. Bagarello and F. Gargano, Phys. Rev. A (in print), arXiv:1402.6201.
[12] I. Rotter, J. Phys. A: Math. Theor. 42 (2009) 153001.
[13] N. Moiseyev, Non-Hermitian Quantum Mechanics. (Cambridge University Press, Cambridge,
2011).
[14] C. M. Bender and S. Boettcher, Phys. Rev. Lett. 80 (1998) 5243.
[15] A. Turbiner, private communication (May 2000).
[16] F. G. Scholtz, H. B. Geyer and F. J. W. Hahne, Ann. Phys. (NY) 213 (1992) 74.
[17] M. Znojil, Phys. Rev. D 78 (2008) 085003.
[18] M. Znojil, SIGMA 5 (2009) 001 (arXiv overlay: 0901.0700).
[19] M. Znojil and H. B. Geyer, Fort. d. Physik 61 (2013) 111.
[20] A. Mostafazadeh, J. Phys. A: Math. Gen. 39 (2006) 10171.
[21] M. Znojil, Ann. Phys. (NY) 336 (2013) 98.
[22] S. Longhi, Europhys. Lett. 106 (2014) 3400.
16
[23] A. V. Smilga, Int. J. Theor. Phys. (in print), arXiv:1409.8450.
[24] M. Znojil, J. Phys. A: Math. Theor. 45 (2012) 444036.
[25] G. evai, F. Ruzicka and M. Znojil, Int. J. Theor. Phys. 53 (2014) 2875.
[26] M. Znojil, J. Phys. A: Math. Gen. 39 (2006) 441.
[27] M. Znojil, Phys. Lett. A 353 (2006) 463.
[28] M. Znojil, J. Phys. A: Math. Gen. 39 (2006) 4047.
[29] M. Znojil, Phys. Rev. D. 80 (2009) 045022.
[30] M. Znojil, Phys. Rev. D. 80 (2009) 045009.
[31] M. Znojil, Phys. Lett. A 375 (2011) 3176.
[32] M. Znojil, J. Phys.: Conf. Ser. 343 (2012) 012136.
[33] M. Znojil, J. Phys. A: Math. Theor. 40 (2007) 4863.
[34] M. Znojil, J. Phys. A: Math. Theor. 40 (2007) 13131.
[35] M. Znojil, J. Phys. A: Math. Theor. 41 (2008) 244027.
[36] M. Znojil, J. Phys. A: Math. Theor. 45 (2012) 085302.
[37] V. G. Gurzadyan and R. Penrose, Eur. Phys. J. Plus 128 (2013) 22.
[38] A. Mostafazadeh, Phil. Trans. R. Soc. A 371 (2013) 20120050.
[39] L. N. Trefethen and M. Embree, Spectra and pseudospectra - the behavior of nonnormal
matrices. (Princeton University Press, Princeton, 2005).
[40] D. Krejcirik, P. Siegl, M. Tater and J. Viola, Pseudospectra in non-Hermitian quantum
mechanics. arXiv:1402.1082.
17
... In the limit U/t → 0, the two EPs converge at λ EP = 0 to create a conical intersection with a gradient discontinuity on the real axis. This gradient discontinuity defines a critical point in the ground-state energy, where a sudden change occurs in the electronic wave function, and can be considered as a zero-temperature quantum phase transition (QPT) [29,[69][70][71][72][73][74][75]. ...
... When a Hamiltonian is parametrised by a variable such as λ, the existence of abrupt changes in the eigenstates as a function of λ indicate the presence of a zero-temperature QPT [29,[69][70][71][72][73][74][75]. Meanwhile, as an avoided crossing becomes increasingly sharp, the corresponding EPs move increasingly close to the real axis. ...
Article
Full-text available
We explore the non-Hermitian extension of quantum chemistry in the complex plane and its link with perturbation theory. We observe that the physics of a quantum system is intimately connected to the position of complex-valued energy singularities, known as exceptional points. After presenting the fundamental concepts of non-Hermitian quantum chemistry in the complex plane, including the mean-field Hartree-Fock approximation and Rayleigh-Schrödinger perturbation theory, we provide a historical overview of the various research activities that have been performed on the physics of singularities. In particular, we highlight seminal work on the convergence behaviour of perturbative series obtained within Møller-Plesset perturbation theory, and its links with quantum phase transitions. We also discuss several resummation techniques (such as Padé and quadratic approximants) that can improve the overall accuracy of the Møller-Plesset perturbative series in both convergent and divergent cases. Each of these points is illustrated using the Hubbard dimer at half filling, which proves to be a versatile model for understanding the subtlety of analytically-continued perturbation theory in the complex plane.
... This gradient discontinuity defines a critical point in the ground-state energy, where a sudden change occurs in the electronic wave function, and can be considered as a zero-temperature quantum phase transition. 29,[69][70][71][72][73][74][75] Remarkably, the existence of these square-root singularities means that following a complex contour around an EP in the complex λ plane will interconvert the closed-shell ground and excited states (see Fig. 1b). This behaviour can be seen by expanding the radicand in Eq. (4a) as a Taylor series around λ EP to give ...
... When a Hamiltonian is parametrised by a variable such as λ, the existence of abrupt changes in the eigenstates as a function of λ indicate the presence of a zero-temperature quantum phase transition (QPT). 29,[69][70][71][72][73][74][75] Meanwhile, as an avoided crossing becomes increasingly sharp, the corresponding EPs move in-creasingly close to the real axis. When these points converge on the real axis, they effectively "annihilate" each other and no longer behave as EPs. ...
Preprint
Full-text available
We explore the non-Hermitian extension of quantum chemistry in the complex plane and its link with perturbation theory. We observe that the physics of a quantum system is intimately connected to the position of complex-valued energy singularities, known as exceptional points. After presenting the fundamental concepts of non-Hermitian quantum chemistry in the complex plane, including the mean-field Hartree--Fock approximation and Rayleigh--Schr\"odinger perturbation theory, we provide a historical overview of the various research activities that have been performed on the physics of singularities. In particular, we highlight seminal work on the convergence behaviour of perturbative series obtained within M{\o}ller--Plesset perturbation theory, and its links with quantum phase transitions. We also discuss several resummation techniques (such as Pad\'e and quadratic approximants) that can improve the overall accuracy of the M{\o}ller--Plesset perturbative series in both convergent and divergent cases. Each of these points is illustrated using the Hubbard dimer at half filling, which proves to be a versatile model for understanding the subtlety of analytically-continued perturbation theory in the complex plane.
... short reviews [8][9][10] and also . Moreover, relevance of EPs in the context of quantum chaos and quantum phase transitions has been demonstrated theoretically [12,[38][39][40][41][42][43][44][45][46][47][48]. The role of EPs in the superradiance phenomenon has also been recognized [49,50]. ...
Article
Full-text available
We study exceptional points (EPs) of a nonhermitian Hamiltonian \hat{H}\n(\lambda,\delta) whose parameters λC\lambda \in {\mathbb C} and δR\delta \in {\mathbb R}. As the real control parameter δ\delta is varied, the k-th EP (or k-th cluster of simultaneously existing EPs) of \hat{H}\n(\lambda,\delta) moves in the complex plane of λ\lambda along a continuous trajectory, λk(δ)\lambda_k(\delta). Using an appropriate non-hermitian formalism (based upon the c-product and not upon the conventional Dirac product), we derive a self contained set of equations of motion (EOM) for the trajectory λk(δ)\lambda_k(\delta), while interpreting δ\delta as the propagation time. Such EOM become of interest whenever one wishes to study the response of EPs to external perturbations or continuous parametric changes of the pertinent Hamiltonian. This is e.g.~the case of EPs emanating from hermitian curve crossings/degeneracies (which turn into avoided crossings/near-degeneracies when the Hamiltonian parameters are continuously varied). The presented EOM for EPs have not only their theoretical merits, they possess also a substantial practical relevance. Namely, the just presented approach can be regarded even as an efficient numerical method, useful for generating EPs for a broad class of complex quantum systems encountered in atomic, nuclear and condensed matter physics. Performance of such a method is tested here numerically on a simple yet nontrivial toy model.
... In the language of mathematics such a "critical value" of the coupling is to be identified with the Kato's EP parameter λ (EP ) . Recently, the concept acquired an immediate experimental meaning in several phenomenological applications ranging from the relativistic quantum mechanics [33,34] and quantum cosmology [35,36] up to the efficient toy-model simulations of the various forms of quantum phase transitions [37,38,39]. Many years ago, the use of the EPs already caused a change of the paradigm in the mathematical foundations of the perturbation theory of linear operators [18,40]. ...
Preprint
A family of non-Hermitian but PT{\cal PT}-symmetric 2J by 2J toy-model tridiagonal-matrix Hamiltonians H(2J)=H(2J)(t)H^{(2J)}=H^{(2J)}(t) with J=K+M=1,2,J=K+M=1,2,\ldots and t<J2t<J^2 is studied, for which a real but non-Hermitian 2K by 2K tridiagonal-submatrix component C(t) of the Hamiltonian is assumed coupled to its other two complex but Hermitian M by M tridiagonal-submatrix components A(t) and B(t). By construction, (i) all of the submatrices get decoupled at t=tM=M(2JM)t=t_M=M\,(2J-M) with M=1,2,,JM=1,2,\ldots,J; (ii) at all of the parameters t=tMt=t_M with M=JK=0,1,,J1M=J-K=0,1,\ldots,J-1 the Hamiltonian ceases to be diagonalizable exhibiting the Kato's exceptional-point degeneracy of order 2K; (iv) the system's PT{\cal PT}-symmetry gets spontaneously broken when ttJ1=J21t\leq t_{J-1}=J^2-1.
... The latter studies were often motivated by the physics of systems exhibiting a genuine quantum phase transition 28,29 . According to our most recent commentary 30 , most of these systems have been considered "open", interacting with a certain not too well specified "environment". ...
Article
Full-text available
In the problem of classification of the parameter-controlled quantum phase transitions, attention is turned from the conventional manipulations with the energy-level mergers at exceptional points to the control of mergers of the exceptional points themselves. What is obtained is an exhaustive classification which characterizes every phase transition by the algebraic and geometric multiplicity of the underlying confluent exceptional point. Typical qualitative characteristics of non-equivalent phase transitions are illustrated via a few elementary toy models.
... The latter studies were often motivated by the physics of systems exhibiting a genuine quantum phase transition [28,29]. According to our most recent commentary [30], most of these systems have been considered "open", interacting with a certain not too well specified "environment". ...
Preprint
Full-text available
Specific quantum phase transitions of our interest are assumed associated with the fall of a closed, unitary quantum system into its exceptional-point (EP) singularity. The physical realization of such a "quantum catastrophe" (connected, typically, with an instantaneous loss of the diagonalizability of the corresponding parameter-dependent Hamiltonian H(g)) depends, naturally, on the formal mathematical characteristics of the EP, i.e., in essence, on its so called algebraic multiplicity N and geometric multiplicity K. In our paper we assume that both of them are finite, and we illustrate and discuss, using several solvable toy models, some of the most elementary mechanisms of the EP-merger realization of the process of the transition gg(EP)g \to g^{(EP)}.
... In spite of the manifest non-Hermiticity of the P T −symmetric candidates H for Hamiltonians, these operators were shown eligible as generators of unitary evolution [12,14]. In this context, one of the basic methodical assumptions accepted in the current literature on BH models [19,20,[34][35][36][37] was that the infinite-dimensional matrix (5) as well as all of its separate submatrices (6) had to be complex symmetric, tridiagonal and P T −symmetric, with P equal to an antidiagonal unit matrix, and with symbol T representing an antilinear operation of Hermitian conjugation (i.e., transposition plus complex conjugation). This led to the conclusion (or rather conjecture) that in the EPN limit (i.e., at the instant of the loss of diagonalizability), the canonical representation of every N by N submatrix H (N) (γ (EP) ) can be given the form of the N by N Jordan matrix (16) with, due to P T −symmetry, η = 0. ...
Article
Full-text available
It is well known that, using the conventional non-Hermitian but PT−symmetric Bose–Hubbard Hamiltonian with real spectrum, one can realize the Bose–Einstein condensation (BEC) process in an exceptional-point limit of order N. Such an exactly solvable simulation of the BEC-type phase transition is, unfortunately, incomplete because the standard version of the model only offers an extreme form of the limit, characterized by a minimal geometric multiplicity K = 1. In our paper, we describe a rescaled and partitioned direct-sum modification of the linear version of the Bose–Hubbard model, which remains exactly solvable while admitting any value of K≥1. It offers a complete menu of benchmark models numbered by a specific combinatorial scheme. In this manner, an exhaustive classification of the general BEC patterns with any geometric multiplicity is obtained and realized in terms of an exactly solvable generalized Bose–Hubbard model.
... improvement of our understanding of many other EP-related models [15][16][17][18]. At present, one can observe that this direction of research did already branch into a large number of various separate experimental as well as theoretical subdirections [19][20][21][22][23][24][25][26][27][28][29][30][31]. ...
Article
Full-text available
The description of unitary evolution using non-Hermitian but ‘hermitizable’ Hamiltonians H is feasible via an ad hoc metric Θ = Θ ( H ) and a (non-unique) amendment 〈 ψ 1 | ψ 2 〉 → 〈 ψ 1 | Θ | ψ 2 〉 of the inner product in Hilbert space. Via a proper fine-tuning of Θ ( H ) this opens the possibility of reaching the boundaries of stability (i.e. exceptional points) in many quantum systems sampled here by the fairly realistic Bose–Hubbard (BH) and discrete anharmonic oscillator (AO) models. In such a setting, it is conjectured that the EP singularity can play the role of a quantum phase-transition interface between different dynamical regimes. Three alternative ‘AO ↔ BH’ implementations of such an EP-mediated dynamical transmutation scenario are proposed and shown, at an arbitrary finite Hilbert-space dimension N , exact and non-numerical.
Article
In paper I (Znojil, 2020) it has been demonstrated that besides the well known use of the Arnold’s one-dimensional polynomial potentials V(k)(x)=xk+1+c1xk−1+… in the classical Thom’s catastrophe theory, some of these potentials (viz., the confining ones, with k=2N+1) could also play an analogous role of genuine benchmark models in quantum mechanics, especially in the dynamical regime in which N+1 valleys are separated by N barriers. For technical reasons, just the ground states in the spatially symmetric subset of V(k)(x)=V(k)(−x) have been considered. In the present paper II we will show that and how both of these constraints can be relaxed. Thus, even the knowledge of the trivial leading-order form of the excited states will be shown sufficient to provide a new, truly rich level-avoiding spectral pattern. Secondly, the fully general asymmetric-potential scenarios will be shown tractable perturbatively.
Article
Phenomenological quantum Hamiltonians H(N)(λ)=J(N)+λV(N)(λ) representing a general real N2-parametric perturbation of an exceptional-point-related unperturbed Jordan-block Hamiltonian J(N) are considered. Tractable as non-Hermitian (in a preselected, unphysical Hilbert space) as well as, simultaneously, Hermitian (in another, “physical” Hilbert space), these matrices may represent a unitary, closed quantum system if and only if the spectrum is real. At small λ we show that the parameters are then confined to a “stability corridor” S of the access to the extreme dynamical exceptional-point λ→0 regime. The corridors are narrow and N-dependent: they are formed by multiscale perturbations which are small in physical Hilbert space, i.e., which are such that λVj+k,j(N)(λ)=O(λ(k+1)/2) at k=1,2,...,N−1 and all j.
Article
Full-text available
The correspondence between exotic quantum holonomy, which occurs in families of Hermitian cycles, and exceptional points (EPs) for non-Hermitian quantum theory is examined in quantum kicked tops. Under a suitable condition, an explicit expression of the adiabatic parameter dependencies of quasienergies and stationary states, which exhibit anholonomies, is obtained. It is also shown that the quantum kicked tops with the complexified adiabatic parameter have a higher-order EP, which is broken into lower-order EPs with the application of small perturbations. The stability of exotic holonomy against such bifurcation is demonstrated.
Article
Full-text available
The spectral and transport properties of a non-Hermitian tight-binding lattice with unidirectional hopping are theoretically investigated in three different geometrical settings. It is shown that, while for the infinitely-extended (open) and for the ring lattice geometries the spectrum is complex, lattice truncation makes the spectrum real. However, an exceptional point of order equal to the number of lattice sites emerges. When a homogeneous dc force is applied to the lattice, in all cases an equally-spaced real Wannier-Stark ladder spectrum is obtained, corresponding to periodic oscillatory dynamics in real space. Possible physical realizations of non-Hermitian lattices with unidirectional hopping are briefly discussed.
Article
Full-text available
Three classes of finite-dimensional models of quantum systems exhibiting spectral degeneracies called quantum catastrophes are described in detail. Computer-assisted symbolic manipulation techniques are shown unexpectedly efficient for the purpose.
Article
Full-text available
We discuss the role of pseudo-fermions in the analysis of some two-dimensional models, recently introduced in connection with non self-adjoint hamiltonians. Among other aspects, we discuss the appearance of exceptional points in connection with the validity of the extended anti-commutation rules which define the pseudo-fermionic structure.
Article
Full-text available
The spectra of, e.g. open quantum systems are typically given as the superposition of resonances with a Lorentzian line shape, where each resonance is related to a simple pole in the complex energy domain. However, at exceptional points two or more resonances are degenerate and the resulting non-Lorentzian line shapes are related to higher order poles in the complex energy domain. In the Fourier-transform time domain an n-th order exceptional point is characterised by a non-exponentially decaying time signal given as the product of an exponential function and a polynomial of degree n1n-1. The complex positions and amplitudes of the non-degenerate resonances can be determined with high accuracy by application of the nonlinear harmonic inversion method to the real-valued resonance spectra. We extend the harmonic inversion method to include the analysis of exceptional points. The technique yields, in the energy domain, the amplitudes of the higher order poles contributing to the spectra, and, in the time domain, the coefficients of the polynomial characterising the non-exponential decay of the time signal. The extended harmonic inversion method is demonstrated on two examples, viz. the analysis of exceptional points in resonance spectra of the hydrogen atom in crossed magnetic and electric fields, and an exceptional point occurring in the dynamics of a single particle in a time-dependent harmonic trap.
Article
The statement in the title discussed earlier in association with the Pais-Uhlenbeck oscillator with equal frequencies is illustrated for an elementary matrix model. In the limit when the order of the exceptional point N tends to infinity, an infinity of nontrivial states that do not change their norm during evolution appear. These states have real energies lying in a continuous interval. The norm of the "precursors" of these states at large finite N is not conserved, but the characteristic time scale where this nonconservation shows up grows linearly with N.
Article
A diagonalizable non-Hermitian Hamiltonian having a real spectrum may be used to define a unitary quantum system, if one modifies the inner product of the Hilbert space properly. We give a comprehensive and essentially self-contained review of the basic ideas and techniques responsible for the recent developments in this subject. We provide a critical assessment of the role of the geometry of the Hilbert space in conventional quantum mechanics to reveal the basic physical principle motivating our study. We then offer a survey of the necessary mathematical tools, present their utility in establishing a lucid and precise formulation of a unitary quantum theory based on a non-Hermitian Hamiltonian, and elaborate on a number of relevant issues of fundamental importance. In particular, we discuss the role of the antilinear symmetries such as , the true meaning and significance of the so-called charge operators and the -inner products, the nature of the physical observables, the equivalent description of such models using ordinary Hermitian quantum mechanics, the pertaining duality between local-non-Hermitian versus nonlocal-Hermitian descriptions of their dynamics, the corresponding classical systems, the pseudo-Hermitian canonical quantization scheme, various methods of calculating the (pseudo-) metric operators, subtleties of dealing with time-dependent quasi-Hermitian Hamiltonians and the path-integral formulation of the theory, and the structure of the state space and its ramifications for the quantum Brachistochrone problem. We also explore some concrete physical applications and manifestations of the abstract concepts and tools that have been developed in the course of this investigation. These include applications in nuclear physics, condensed matter physics, relativistic quantum mechanics and quantum field theory, quantum cosmology, electromagnetic wave propagation, open quantum systems, magnetohydrodynamics, quantum chaos and biophysics.
Article
We propose giving the mathematical concept of the pseudospectrum a central role in quantum mechanics with non-Hermitian operators. We relate pseudospectral properties to quasi-Hermiticity, similarity to self-adjoint operators, and basis properties of eigenfunctions. The abstract results are illustrated by unexpected wild properties of operators familiar from PT-symmetric quantum mechanics.