Sie sind auf Seite 1von 17

doi: 10.1098/rspa.2008.

0278
, 71-86 465 2009 Proc. R. Soc. A

Eric A Galapon

collapse on the appearance of particle


Theory of quantum arrival and spatial wave function

References
ml#ref-list-1
http://rspa.royalsocietypublishing.org/content/465/2101/71.full.ht
This article cites 37 articles, 4 of which can be accessed free
Subject collections
(35 articles) quantum physics

Articles on similar topics can be found in the following collections


Email alerting service
here the box at the top right-hand corner of the article or click
Receive free email alerts when new articles cite this article - sign up in
http://rspa.royalsocietypublishing.org/subscriptions go to: Proc. R. Soc. A To subscribe to
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
Theory of quantum arrival and spatial wave
function collapse on the appearance of particle
BY ERIC A. GALAPON*
Theoretical Physics Group, National Institute of Physics, University of the
Philippines Diliman, Quezon City 1101, The Philippines
The unexpected connection is unravelled between the collapse of the wave function on
the appearance of particle and the quantum time-of-arrival problem in one dimension.
To do so, a theory of quantum rst time of arrival is developed in the interacting case for
arbitrary arrival point in one dimension based on a self-adjoint and canonical coarse
graining of a time-of-arrival operator that derives the classical time-of-arrival observable.
The appearance of particle in quantum mechanics is then considered in the light of this
theory. It is found that the appearance of particle arises as a combination of the collapse
of the initial wave function into one of the eigenfunctions of the time-of-arrival operator,
followed by a unitary Schrodinger evolution of the eigenfunction.
Keywords: time of arrival; quantum measurement; wave function collapse; particles
1. Introduction
There are two dynamics in standard quantum mechanics (SQM): the well-
understood and uncontroversial continuous, unitary evolution of quantum states
according to the Schro dinger equation; and the ill-understood and controversial
discontinuous, non-unitary evolution of the same states during quantum measure-
ments (Bhom1952a,b; Everett 1957; Grifths 1984; Ghirardi et al. 1986; Zurek 2003;
Schlosshauer 2004). The abrupt (discontinuous) and irreversible (non-unitary)
evolutionduring quantummeasurements is knownas wave functioncollapse. It is the
consensus that the collapse occurs at the moment the measurement is made, and the
wave function collapses randomly into one of the eigenfunctions of the observable
being measured. Now one distinct aspect of the wave function collapse is the
appearance of particles fromthewave descriptionof matter. The particles are distinct
among manifestations of the collapse because particles are reidentiable, localized
entities with permanent properties. These properties of particles are at odds with
their quantummechanical description as spread-out waves evolving according to the
Schro dinger equation, and their appearance as localized objects presents a problem.
Schrodinger recognizes this problem early on and points out that emerging
particle from decaying nuclei is described as a spherical wave that impinges
continuously on a surrounding luminescent screen over its full expanse[;] the
screen, however, does not show a more or less constant uniform surface glow, but
rather lights up at one instant at one spot (Schro dinger 1980b). Schro dinger
Proc. R. Soc. A (2009) 465, 7186
doi:10.1098/rspa.2008.0278
Published online 29 August 2008
*eric.galapon@up.edu.ph
Received 7 July 2008
Accepted 31 July 2008 71 This journal is q 2008 The Royal Society
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
nds this idea of abrupt, instantaneous and spatially non-local collapsing of the
entire spread-out wave function into a function of point support at the moment
of the appearance of the particle ridiculous (Schro dinger 1952) and altogether
denies the existence of particles. However, such collapse of the wave function
seems inevitable if we need to reconcile quantum description with our experience
with measuring instruments. And out of this inevitability of the collapse a
question arises: Is the collapse of the wave function on the appearance of particle
fundamental? That is, does the collapse occur at the moment of the appearance
of particle and hence cannot be broken down into a series of casually separated
processes? The consensus within SQM is that the collapse is fundamental. In this
paper, we consider this question in the light of the quantum arrival problem
within SQM (Bohm 1952a; Kijowski 1974; Holland 1993; Grot et al. 1996;
Delgado & Muga 1997; Leavens 1998; Kochanski & Wo dkiewicz 1999; Leo n et al.
2000; Muga & Leavens 2000; Muga et al. 2002; Galapon et al. 2002a,b, 2004,
2005a,b; Galapon 2004; Hegerfeldt et al. 2004) and nd a different answer.
The appearance of particle, as posed by Schro dinger, is evidently a quantum
arrival problem: nding a quantum mechanical mechanism for the localization of
a unitarily evolving wave function at a denite point in space at a denite time at
the registration of the particle. As the problem stands, the quantum time-
of-arrival problem (QTOAP)the problem of nding the time-of-arrival
distribution at some given arrival points in space for a given initial stateis a
promising framework to meet the problem head on. That is so because the
QTOAP deals with the appearance of the particle at the arrival point at the
arrival time. Any solution then to the QTOAP should not only be able to predict
the time-of-arrival distributions but must also explain how the particle appears
at the moment of arrival. The QTOAP, however, is wrought with controversy
and there are as many solutions as there are independent researchers in the eld
(Muga & Leavens 2000), not to mention non-trivial assertions that the problem
is not meaningful or possesses no solution at all (Allcock 1969ac; Yamada &
Takagi 1991a,b, 1992; Halliwell & Zaris 1998). Moreover, available QTOA
theories have only addressed the time-of-arrival distribution aspect of the
problem (Bohm 1952a,b; Kijowski 1974; Holland 1993; Grot et al. 1996;
Delgado & Muga 1997; Leavens 1998; Kochanski & Wo dkiewicz 1999; Baute
et al. 2000; Leon et al. 2000; Muga & Leavens 2000; Muga et al. 2002; Hegerfeldt
et al. 2004), so that connection to wave function collapse at the appearance
of particle has never been made until now.
In the following, we make the connection between the QTOAP and the
problem of particles in one dimension within the connes of SQM. We do so by a
generalization of our earlier solution to the one-dimensional free QTOAP
(Galapon et al. 2005b) for arbitrary arrival point, for arbitrary interaction
potential via spatial connement followed by a limiting procedure for arbitrarily
large conning lengths. We will nd that the resulting quantum time-of-arrival
theory suggests that the collapse of the wave function on the appearance of the
particle is not fundamental: the collapse occurs much earlier than the appearance
of the particle and that the subsequent localization of the wave function on the
appearance of the particle arises from the unitary Schro dinger equation.
The paper is organized as follows. Section 2 summarizes the construction of
time-of-arrival operator for arbitrary interaction potential. Section 3 discusses
the coarse graining of the time-of-arrival operators by spatial connement and
E. A. Galapon 72
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
the dynamics of their coarse-grained versions. Section 4 investigates the
behaviour of the eigenfunctions of the conned time of arrival (CTOA) operators
in the limit of arbitrarily large conning lengths. Section 5 discusses the
emergent picture of the appearance of particles in the light of quantum arrival
theory. Section 6 lays down the complete formalism in the computation of the
time-of-arrival distribution for arbitrary arrival point for arbitrary interaction
potential in one dimension. Section 7 provides the conclusion. Appendix A
details the numerical implementation of the algorithm used in computing the
time-of-arrival distribution.
2. Quantum time-of-arrival operator in the interacting case
Generally in SQM, to determine the distribution of a given observable one needs
only the spectral resolution of an operator representation of the observable in
question. Then, naturally to address QTOAP within SQM one needs to construct
the appropriate time-of-arrival operator for the given system. However, the
consensus is that no such operator can be constructed in the most general case
of arbitrary arrival point and of arbitrary interaction potential. In one
dimension, the most quoted reason is that the classical time of arrival at point
xgiven by T
x
q; pZKsgnp

m=2
_ _
q
x
Hq; pKVq
0

K1=2
dq
0
, where H is
the Hamiltonian, V is the interaction potential, m is the mass of the particle
and (q, p) are, respectively, the position and the momentum at tZ0does
not admit a sensible quantization because it is generally not everywhere real
and single valued in the entire phase space (Peres 1995; Leo n et al. 2000;
Muga & Leavens 2000). For this reason, it is believed that if a theory of quantum
arrival existed it could not rest on the spectral resolution of a time-
of-arrival operator.
But if we go by the standard formulation, we must insist on nding an
operator Ta time-of-arrival operatorthat must, foremost, reduce to the
classical time of arrival T
x
(q, p). This is the minimum requirement that T must
satisfy to be identiable as a time-of-arrival operator. Motivated to breaking the
circularity of quantization when invoking the correspondence principle and to
sidestepping the well-known existence of obstruction to quantization in
important spaces such as the Euclidean space, we introduced in Galapon
(2004) the idea of supraquantizationthe derivation of a quantum observable
corresponding to a classical observable without explicit reference of quantiza-
tionand found such an operator T appropriately reducing to T
x
(q, p) in the
classical limit.
In coordinate representation, the operator T is the integral operator
T4qZ
_
N
KN
qjTjq
0
h i4q
0
dq
0
, with the kernel given by
qjTjq
0
_
Z
m
iZ
Tq; q
0
sgnqKq
0
; 2:1
in which we have referred to T(q, q
0
) as the kernel factor and is determined by the
interaction potential V(q). To simplify our discussion, let us for the moment
consider arrival at the origin, xZ0, and consider later arbitrary arrival points.
73 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
For such a case, the kernel factor is the solution to
K
v
2
Tq; q
0

vq
2
C
v
2
Tq; q
0

vq
0 2
C
2m
Z
2
VqKVq
0
Tq; q
0
Z0; 2:2
subject to the conditions T(q, q)Zq/2 and T(q, Kq)Z0. The operator T is the
solution to timeenergy canonical commutation relation in the formhfjT; Hj4iZ
iZhfj4i, where f, 4 are innitely differentiable functions with compact supports,
subject to the condition that T must reduce to T
xZ0
q; pZT
0
q; p in the
limit Z/0.
The classical time of arrival derives from T via the WeylWigner transform
of its kernel: T
Z
q; pZ
_
N
KN
hqCv=2jTjqKv=2iexpKivp=Zdv. For linear
systems, we have T
Z
q; pZt
0
q; p; while for nonlinear systems, T
Z
q; pZt
0
q; pCOZ
2
, where t
0
(q, p) is the expansion of T
0
(q, p) about the classical free
time of arrival, and referred to as the local time of arrival in Galapon (2004).
Generally, t
0
(q, p) converges only in some neighbourhood of the arrival point. In
its region of convergence, t
0
(q, p) is real and single valued and is equal to
T
0
(q, p). That is, t
0
(q, p) is the classical rst time of arrival for initial states lying
in the neighbourhood of the arrival point. This indicates that our operator T is a
quantum rst time-of-arrival operator and derives T
0
(q, p) via the unique
extension of t
0
(q, p) in the larger classical phase space.
We now devote the rest of the paper to studying the physical content of T,
and show that it comprises a solution to the QTOAP and Schro dinger problem in
one dimension.
3. Coarse graining of the time-of-arrival operator
A major obstacle in understanding the physics of T is its current inaccessibility to
analysis; for example, solving for the eigenvalue problem of T may be intractable
for arbitrary potential. We then approach the unravelling of its physical content
by successive coarse grainingsuccessive approximation of T with discrete
observables
1
. To do that, we need rst to explicitly write T in terms of the
position and momentum operators, q and p, respectively. T must be written in
them such that, in coordinate representation, the kernel of T is given by equation
(2.1). It turns out that this can be accomplished by Weyl quantizing T
Z
q; p. To
proceed, let us, in the mean time, consider everywhere analytic potentials. For
such cases, T
Z
q; p is an expansion in q
n
p
Km
for positive integers n and m. The
explicit operator form of T is then obtained by replacing q
n
p
Km
with
T
Km;n
Z2
Kn

n
jZ0
n
j
_ _
q
j
$p
Km
q
nKj
in T
Z
q; p. In this form, the canonical relation hfj H; Tj4iZiZhfj4i translates
to T; HZiZI. Moreover, in this form, it is clear that T is just the Weyl
quantization of the local time of arrival t
0
(q, p) for linear systems; while, for
nonlinear systems, it is the quantization of t
0
(q, p) plus quantum corrections
1
Coarse graining by discrete observables is usually necessary in quantum measurements of
observables with continuous spectrum (Busch et al. 1995).
E. A. Galapon 74
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
required to satisfy the timeenergy commutation relation. We emphasize that,
due to the existence of obstruction to quantization in Euclidean space, T cannot
be constructed, except for linear systems, via direct quantization of the classical
time of arrival.
Now a coarse graining of T is constructed by conning the system in a large
box of length 2l centred at the arrival point. The coarse-grained version of T is
then obtained by projecting its explicit operator form in the Hilbert space
H
l
ZL
2
Kl; l , under the condition that the Hamiltonian is purely kinetic for
vanishing interaction potential. This condition projects the momentum operator
p into the ring of momentum operators fp
g
ZKi Zv
q
; jgj%p=2g, with p
g
having
the domain consisting of absolutely continuous functions f(q) in H
l
ZKl; l with
square integrable rst derivatives, which further satisfy the boundary condition
fKl Ze
K2ig
fl . Since T depends on the momentum operator, the coarse
graining of T is then the set of operators {T
g
} with each T
g
corresponding to the
momentum p
g
.
In coordinate representation, each T
g
in the Hilbert space H
l
ZL
2
Kl; l is
the integral operator
T
g
4q Z
_
l
Kl
hqjT
g
jq
0
i4q
0
dq
0
:
Using the coordinate representation of the operators T
m,n
in H
l
for a given g, the
kernel of T
g
can be shown to be given by
hqjT
gs0
jq
0
i ZKm
Tq; q
0

Z sin g
e
ig
HqKq
0
Ce
Kig
Hq
0
Kq ; 3:1
hqjT
gZ0
jq
0
i Z
m
iZ
Tq; q
0
sgnqKq
0
K
m
ilZ
Bq; q
0
; 3:2
with T(q, q
0
) the kernel factor in equation (2.1), and Bq; q
0
Z
_
z
0
Th; z dz,
Th; zZTq; q
0
in which zZ(qKq
0
) and hZ(qCq
0
)/2, and H(x) is the
Heaviside step function (Galapon 2006). Using equations (3.1) and (3.2) together
with equation (2.2), our earlier restriction on the potential can already be lifted
to include a more general interaction potential. Observe that equation (3.2)
explicitly reduces to equation (2.1) as l approaches innity, which must be the
case. Owing to this we will exclusively use T
0
in our investigation in the limit of
arbitrarily large conning lengths.
Now comparing kernels (3.1) and (3.2) with those constructed in Galapon
(2006), we nd that they are just the CTOA operators for a given interaction
potential. The CTOA operators for a given conning length then constitute the
coarse graining of T. For continuous potentials, the CTOA operators are
compact non-degenerate self-adjoint operators. Their compactness implies that
they have pure discrete spectrum, with corresponding complete square inte-
grable eigenfunctions. It is their compactness that makes them a coarse graining
of the operator T. The eigenfunctions can be written as 4
G
n
, where the sign
indicates the sign of the corresponding eigenvalue, with t
C
n
ZKt
K
n
, nZ1,2, .;
moreover, they are ordered according to t
G
1

O t
G
2

O/. We have referred to an


eigenfunction 4
G
n
as non-nodal (nodal) when it does not vanish (it vanishes) in
the interval [Kl,l ] (Galapon 2004; Galapon et al. 2004).
75 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
The spectral properties of the CTOA operators have the surprising property
that they are intimately tied with the internal unitary dynamics of the system:
the eigenfunctions evolve according to Schro dingers equation such that the event
of the position expectation value assuming the arrival point, and the event of the
position uncertainty being minimum occur at the same instant of time equal to
their corresponding eigenvalues, with the uncertainty decreasing with the
magnitude of the eigenvalue. We referred to this property as unitary arrival
of the eigenfunctions at the arrival point (Galapon et al. 2004; Galapon 2006).
The unitary arrival of the eigenfunctions at their respective eigenvalues is
consistent with the interpretation that T is a rst time-of-arrival operator.
However, the arrival is not sharp in the sense that the evolving eigenfunctions
may have considerable spatial supports even at their eigenvalues; this is true
for the large eigenvalueeigenfunctions. Nevertheless, the property already hints
that T may account for the localization of the wave function in space and time
on the appearance of particle. Indeed that is what we will nd as the coarse
graining gets more rened: as the conning length increases indenitely. We
proceed numerically.
4. The limit for arbitrarily large conning lengths
Let us consider specically the behaviour of the harmonic oscillator CTOA
operator for the arrival point xZ0 in the limit of arbitrarily large conning
lengths. The corresponding kernel factor is Tq; q
0
ZZ sinhmuq
2
Kq
0 2
=2Z=
2muqKq
0
; this is found by solving equation (2.2) with V(q)Zu
2
m/2 (Galapon
2006). (Throughout all our computations we set ZZmZuZ1.) Figure 1 shows
the representative behaviour of the evolving eigenfunctions of the harmonic
oscillator CTOA operators for nite l. Now let us investigate the behaviour of T
0
for large values of l. For any given time t we can nd an l
0
O0 large enough such
that jtj is less than the norm or the absolute value of the largest eigenvalue of
the CTOA operator; that is, there exists an eigenvalue of T
0
greater than t. For
any given length lOl
0
of spatial connement, we can nd a pair of nodal and
15
10
5
0
1
0
1 0
0.05
0.10
tim
e
p
o
s
i
t
i
o
n
p
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y
(a)
6
4
2
0
1
0
1 0
0.05
0.10
tim
e
p
o
s
i
t
i
o
n
(b)
Figure 1. The evolved eigenprobability densities, j4
n
q; tj
2
, for the (a) ninth non-nodal (gZ0) and
(b) eighth nodal (gZp/2) positive eigenvalueeigenfunctions of the harmonic oscillator CTOA
operator for lZ1. The eigendensities take their maximum values at the arrival point at their
respective eigenvalues. The eigendensities become more localized at the arrival point at the
eigenvalue as n increases. Adapted from Galapon (2006) with permission from World Scientic.
E. A. Galapon 76
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
non-nodal eigenfunctions P
k
such that their eigenvalues t
k
are closest to t.
Consider a sequence of monotonically increasing l values, l
1
, l
2
, l
3
, ., with
l
0
!l
1
!l
2
!l
3
!/. Then, there will be a k
1
corresponding to l
1
such that t
k
1
is closest to t, and a k
2
corresponding to l
2
such that t
k
2
is closest to t; and so
on. Our computation indicates that as l/N, t
G
2kC1
/t
G
2kC2
. That is, the pair of
eigenfunctions P
k
becomes degenerate.
Our computation likewise indicates that the eigenvalues in the neighbourhood
of t get denser as the conning length tends towards innity, so that, for a given
t, t
k
r
for a given l
r
converges to t. This indicates that the spectrum of the
CTAO operator tends to the continuum, which implies that the eigenfunctions
will eventually be thrown out of the Hilbert spacethey become non-square
integrable. Now, we know from our earlier results (Galapon et al. 2004) that the
width of j4
n
k
q; tj
2
is minimum at the eigenvalue t
n
k
of 4
n
k
q. What is the
behaviour of j4
n
k
q; t
n
k
j
2
as l
k
increases indenitely? Figure 2 shows that for a
given xed time t, j4
n
r
q; t
k
r
j
2
tends to the Dirac delta with support at the
arrival point as l
r
tends to innity. Then, the CTOA eigenfunctions evolve to a
singular support at the arrival point at their respective eigenvalues in the limit;
in particular, j4
n
r
q; t
k
r
j
2
tends to the position eigenfunction at the arrival
point! These results have been shown earlier to be true for the free particle as
well (Galapon et al. 2005b). Moreover, computations on other potentials,
including the quartic oscillator, show the same dynamical behaviour for
arbitrarily large l values. These all together suggest that the behaviours of the
CTOA operators in the limit are the same.
5. The appearance of particles
If indeed this is the case, then their limiting dynamical behaviour gives us the
picture of quantum arrival within the SQM as that of unitarily evolving CTOA
eigenfunction to a localized support at the arrival point. If we accept the dogma
that particles are wave packets of singular support, then the CTOA
1.0 0.5 0 0.5 1.0
0
2
4
6
8
position
p
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y
a
t

e
i
g
e
n
v
a
l
u
e
(a)
1.0 0.5 0 0.5 1.0
0
1
2
3
4
5
6
7
position
p
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y
(b)
Figure 2. (a) The evolved eigenprobability densities at their respective eigenvalues corresponding
to the non-nodal eigenfunctions with eigenvalues nearest to tZ0.11 for lengths lZ1 (solid line),
lZ2 (dashed line) and lZ3 (dotted line), for the harmonic oscillator CTOA operator T
0
. (b) The
corresponding nodal eigenfunctions for lengths lZ2 (solid line), lZ3 (dashed line) and lZ4 (dotted
line). Peak increases with l.
77 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
eigenfunctions for the arbitrarily large values of l are particles at their respective
eigenvalues. This endorses a mechanism for the localization of the wave function
in space and in time at the registration of the particle: consider a quantum
particle prepared in some initial state j
0
. Without loss of generality, we can
assume that j
0
has a compact support. We can then enclose the system by a box
of very large length (for all practical purposes innite), with the support of j
0
laying completely in the box. Since the CTOA eigenfunctions are complete we
can decompose j
0
in terms of these eigenfunctions. Now if we presuppose that
particle detectors somehow respond only to a localized wave packet or localized
energy, then registration or arrival of the particle at the arrival point at time t
can be interpreted as the detection of the component eigenfunction whose
eigenvalue is t that is unitarily arriving (essentially collapsing for arbitrarily
large l ) at the arrival point. This implies that the appearance or arrival of
the particle is a combination of a collapse of the initial wave function into one
of the eigenfunctions of the time-of-arrival operator right after the preparation of
the initial state followed by a unitary evolution of the eigenfunction. That is, the
collapse of the wave function on the appearance of particle is not fundamental
but decomposable into a series of casually separated processes.
Clearly, our interpretation contrasts with the standard interpretation of the
collapse of the spatial wave function on the appearance of the particle. In SQM,
when a quantum object is prepared in some initial state j
0
and when an
observable of the object is measured at a later time T, then the state at the
moment of measurement, which is jTZU
T
j
0
, where U
t
is the time-evolution
operator, collapses randomly into one of the eigenfunctions of the observable.
Now the consensus is that the appearance of particle is a position measurement,
so that the appearance at point q
0
at some time T is the projection of the evolved
wave function jq; TZU
T
j
0
q to the eigenfunction d(qKq
0
) of the position
operator. But according to our quantum arrival description, the collapse occurs
much earlier than the appearance of the particle, with the initial state collapsing
to one of the eigenfunctions of the time-of-arrival operator right after the
preparation and with the particle appearing later at the moment the
eigenfunction has evolved to a state of localized support at the arrival point.
The appearance of particles then (at least within the context of quantum arrival)
does not arise out of position measurement but rather out of time measurement.
One may, however, question the validity of our interpretation when there was
no initial intention to observe the arrival of the particle. If from the very start the
instrument has been set-up to detect the arrival of the particle, it can be argued
that the set-up is already a measurement that has caused the initial state to
collapse into one of the CTOA eigenfunctions, then evolve until observed. But
this reasoning appears untenable when the decision to observe arrival is deferred,
because the initial state has been evolving according to the Schro dinger equation,
assuming that no other observation is made. But not quite. Quantum mechanics
is inherently non-local in time (Wheeler 1978; Scully & Druhl 1982; Kwiat et al.
1992; Aharonov & Zubairy 2005); and that means the description of the past
must bear actions of the present (Greene 2004).
This temporal non-locality of quantum mechanics is aptly illustrated by
Wheelers delayed-choice gedanken experiment, which is depicted in gure 3:
a photon is incident on a 5050 beam splitter BS1, and a choice is available on
whether to insert the beam splitter BS2 or not once the photon has already
E. A. Galapon 78
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
passed through BS1. Now when BS2 is not present, either detector D1 or D2
clicks, indicating that the photon has taken either path P1 or P2, respectively;
since BS1 is 5050, we expect that D1 and D2 have equal probabilities of
registering the arrival of the photon. On the other hand, when BS2 is present, the
phase shift between the two paths can be xed such that, say, D1 has 100% of
detecting the photon; this requires the photon taking the two paths
simultaneously and interfering with itself destructively at D2 and constructively
at D1. The absence of BS2 elicits the particle property of the photon; while its
presence elicits the wave property of the same photon. These properties are
mutually exclusive descriptions of the photon.
Now if from the very start BS2 is already absent, we already know that the
photon will have to take either one of the paths, and manifest particle behaviour;
on the other hand, if from the very start BS2 is already present, we already know
that the photon will take both paths, and manifest wave behaviour. In either
case, we have an unequivocal description of the history of the photon before
registering in one of the detectors. But what can be said of the photons history
when the decision to insert or not the second beam splitter is delayed even after
the photon has passed through the rst beam splitter? SQM predicts that
the description of the photon depends on whether BS2 is present or not,
independently of when BS2 is introducedeven after the photon has long passed
through BS1. The recent experiment by Jacques et al. (2007) has realized
Wheelers gedanken experiment and has shown that the behaviour of the photon
is consistent with this prediction of quantum mechanics. In Wheelers words,
We, now, by moving the [beam splitter] in or out have an unavoidable effect on
what we have a right to say about the already past history of that photon
(Wheeler 1984). That is the description of the past is not complete without
regard to present actions.
Thus, the collapse right after the preparation (when arrival measurement is to
be made) and the Schro dinger evolution right after the preparation (when some
other measurement is to be made) are two mutually exclusive potentialities that
are simultaneously true for the system, which one is realized depends on the
mirror
mirror
P1
BS1
P2
BS2
D1
D2
Figure 3. Wheelers delayed-choice gedanken experiment.
79 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
decision what to do with the system at the moment. Temporal non-locality then
replaces the spatial non-locality inherent in the spontaneous localization of the
wave function at the appearance of the particle in the standard interpretation.
6. Quantum time-of-arrival distribution
(a ) The formalism
With the above interpretation, we can now naturally use without ambiguity the
standard rules of quantum mechanics in computing the probabilities and
densities of quantum arrivals at arbitrary arrival point x for any given initial
state j
0
(q) under an interaction potential V(q). In principle, we should use T to
compute for the probability, but to facilitate the proper interpretation of the
resulting probability we replace the mean time T with its coarse graining T
g
Z0
for some very large conning length l. Meanwhile, let us assume that j
0
(q) has a
compact support. We then construct the operator T
0
for the given l, with its
kernel given by equation (3.2) and kernel factor solved through equation (2.2). In
general, solving for the kernel factor T(q, q
0
) is non-trivial. Fortunately, it can be
expanded in the form Tq; q
0
ZT
0
q; q
0
CT
1
q; q
0
CT
2
q; q
0
C/, where the
leading term is given by
T
0
q; q
0
Z
1
2
_
h
0
ds
0
F
1
; 1;
m
2Z
2
z
2
fVhKVsg
_ _
; 6:1
inwhich
0
F
1
is a specic hypergeometric function, withzZ(qKq
0
) and hZ(qCq
0
)/2.
For linear systems, only the leading term contributes. But for nonlinear systems, all
terms contribute; however, in the classical limit the higher order terms have the
leading OZ
2
contribution so that T
0
(q, q
0
) may sufciently approximate T(q, q
0
)
(Galapon 2004, 2006).
Once T
0
is constructed, compute P
j
0
x; tZ

t
l
0;s
%t
jh4
l
0;s
jj
0
ij
2
, where 4
l
0;s
and t
l
0;s
are, respectively, the eigenfunctions and eigenvalues of T
0
. The overlap
jh4
l
0;s
jj
0
ij
2
is the probability that the initial state will collapse into the sth
eigenfunction right after the preparation. With this interpretation of the overlap,
P
j
0
x; t, in the limit of innite l, can be naturally interpreted as the probability
that one of the components of the eigenfunctions with corresponding eigenvalues
less than or equal to t shall have unitarily evolved to a localized wave function
at the arrival point x. If our detector is what we have presupposed above, then
P
j
0
x; t is the probability of detection or arrival at x after some time t. Given
P
j
0
x; t, the time-of-arrival probability density is found by differentiation with
respect to time, P
j
0
x; tZv
t
P
j
0
x; t. The peaks of P
j
0
x; t determine the most
likely times of arrival at the given arrival point. Now, if the initial state has innite
tails, we can always approximate it with arbitrary accuracy by a function j
l
whose support lies entirely in the interval [Kl,l ] such that j
l
/j
0
as l/N. Then,
P
j
0
x; t is computed as above. The whole process can be implemented numerically
by choosing the conning length to be large enough. The probability density can
then be obtained by numerical interpolation and differentiation (see appendix A
for details).
E. A. Galapon 80
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
Now, when the arrival point is different from the origin, our results above can
be carried over by a mere translation of the origin to the arrival point. This
affects a change from the original potential V(q) to
~
V~ qZV~ qCx, and from the
original initial state j
0
(q) to
~
j
0
~ qZj
0
~ qCx. We then conne the system with
a large box with length 2l centred at x. The box must contain the support of
~
j
0
~ q. Then, we proceed as described above.
In principle, if we can solve the eigenvalue problem for T, we do not need to go
through the above limiting procedure. If E
t
is the spectral decomposition of T,
not necessarily projection valued, then the probability of arrival at time t is given
by P
j
0
x; tZhj
0
jE
t
jj
0
i. However, it may only be for the free particle that E
t
can
be solved explicitly (Galapon et al. 2005b), so that we are forced to go about the
above coarse graining of T to compute for the arrival probability in the
interaction case. But more than a calculational tool, the coarse graining has
enabled us to give an unambiguous interpretation for P
j
0
x; t.
Note that the theory allows us to compute for the time-of-arrival distribution
anywhere in the conguration spaceeven at classically forbidden regions. Of
course, that is the essence of quantum tunnelling. But does our formulation give
insight as to how a quantum particle initially prepared without sufcient energy
to surmount a potential barrier surmounts it nevertheless? Since the detection of
the particle at some point in the conguration space is, according to our
interpretation, a measurement of the quantum observable T and since T is
conjugate to the system Hamiltonian, the detection inevitably perturbs the
energy of the quantum particle. That is, by the uncertainty principle, precise
measurement of the arrival of the particle translates to a large uncertainty in its
energy. This resulting broad distribution of energy makes available sufcient
energy to the particle to surmount the potential hill. This follows naturally if we
insist on the conservation of energy.
(b ) Example: the harmonic oscillator
As an example, let us consider the still untouched harmonic oscillator time-
of-arrival problem. It is sufcient for us to consider cases where we can compare
with the classical case, in particular arrival at the origin. We choose our initial
states to be particular Gaussians of the form 4qZNe
KqKq
0

2
=4dq
2
Cip
0
q=Z
. We
compare the time-of-arrival distribution P
4
xZ0; t computed using our
algorithm above with the classical time of arrival Tq
0
; p
0
ZKu
K1
tan
K1
q
0
=up
0
. Figure 4a shows the computed time-of-arrival distribution for
a xed average position q
0
and for varying average momentum p
0
. Evidently, the
time-of-arrival distribution becomes localized with increasing average momen-
tum. Figure 4b shows the logarithm of the difference between the classical time of
arrival and the most likely quantum time of arrival against the average
momentum of the Gaussian state. The difference decreases with increasing
average momentum, or the quantum time-of-arrival distribution becomes
increasingly localized at the classical time of arrival. This implies that the
quantum rst time-of-arrival distribution approaches the classical distribution
for arbitrarily large momenta or for high energy oscillators. For small momenta
or small energies, the most likely time of arrival is shorter than the classical time
of arrival, so that quantum oscillators are, on the average, faster than their
classical counterparts.
81 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
One desirable property of the time-of-arrival distribution is covariance; that
is, translation in time should not affect the distribution. Covariance has been a
primary requirement on time operators, and it is the lack of covariance of
the CTOA operators (they being compact) that their introduction has been
initially doubted upon. Figure 5 gives evidence of covariance of the distribution
for times smaller than the period of the harmonic oscillator time-evolution
operator. The given initial state is evolved through different times. These evolved
states are used as the initial states in the computation for the TOA distribution.
If the distribution is covariant, then the resulting distributions must be
translations of each other. This is evident in the gure. Thus, while the CTOA
operators are non-covariant, covariance may naturally emerge in the limit of
innite conning length.
0.05 0 0.05 0.10 0.15
0
10
20
30
40
time
t
i
m
e
-
o
f
-
a
r
r
i
v
a
l
p
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y
(a)
0 0.05 0.10 0.15
10
20
30
40
50
time
(b)
Figure 5. Evidence for covariance in the limit of innite l. (a) The probability densities corresponding
to times tZ0 (solid line), tZ0.02 (dashed line) and tZ0.08 (dotted line) for p
0
ZK30, q
0
Z2.25 and
ZZmZsZ1. (b) The probability densities for times tZ0 (solid line) and tZK0.03 (dotted line) for
two coherent superpositions of Gaussians with p
0
(1)ZK30, q
0
(2)Z2.25 and p
0
(2)ZK20, q
0
(2)Z1.25,
respectively. In all cases, the distributions can be obtained from each other via translations in time.
0.1 0.2 0.3 0.4 0
20
40
60
80
100
time
t
i
m
e
-
o
f
-
a
r
r
i
v
a
l
p
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y
(a)
0 10 20 30 40 50 60 70 80
8
7
6
5
4
3
2
momentum
l
o
g
D
(b)
Figure 4. (a) The probability density for q
0
Z2.25 and p
0
Z10, 20, 30 and 50. (b) The logarithm of
the difference between the classical time of arrival and the time at which the time-of-arrival
distribution takes the maximum value.
E. A. Galapon 82
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
7. Conclusion
We have developed a theory of quantum arrival in one dimension for arbitrary
arrival point, for arbitrary interaction potential. It is already the most standard
theory that we can conjure within SQM: It is a theory based on a self-adjoint and
canonical coarse graining of a time-of-arrival operator that derives the classical
time of arrival observable in the classical limit; moreover, it is a theory based on
the collapse-supplemented Schro dinger equation, with probabilities computed
in the standard way. It then provides a new opportunity of studying quantum
time of arrival, which may give us new insights into other areas involving time in
quantum mechanics, such as dynamical aspects of quantum tunnelling. But more
than this opportunity is the novel insight the theory provides on the problem of
particles. It suggests that the appearance of particle arises as a combination
of the collapse of the initial wave function into one of the eigenfunctions of the
time-of-arrival operator, followed by the unitary Schro dinger evolution of the
eigenfunction. This implies that particles do not arise out of position measure-
ments but out of time-of-arrival measurements, and that the collapse of the wave
function on the appearance of particle is not fundamental but decomposable
into a series of casually separated processes.
Schrodinger, in his 1953 Geneva lecture, concludes, Well, what are these
corpuscles, really? [It] may be permissible to say that one can think of particles
as more or less temporary entities within the wave eld whose form and
general behavior are nevertheless so clearly and sharply determined by the laws
of waves (Schro dinger 1980a). The emergent description of the appearance of
particle out of our time-operator-based theory of quantum arrival is consistent
with this expectation of Schrodingerthe particle appears out of an evolving
eigenfunction via Schro dingers wave equation, and appears temporarily at the
moment the eigenfunction assumes a well-localized support. Schro dinger might
have been happy to learn that such a description existed; but, at the same time,
appalled at the thought that such a description still appealed to collapse of the
wave function. Nevertheless, he might still have found the collapse at least
consolatory because it has brought the idea of particles nearer to his
uncompromising preconceptions.
This work was supported by UP-OVCRD Outright Research Project No. 070703 PNSE and UP
System Grant, and also beneted from a recent collaboration with F. Delgado, I. Egusquiza and
Prof. J.G. Muga. The author especially acknowledges the numerous discussions with Prof. J. G.
Muga, which have contributed to the development of the theory.
Appendix A
Let us describe our numerical implementation of the above limiting procedure in
the calculation of the time-of-arrival distribution. For sufciently large conning
length, we have approximately P
j
0
x; t
s
z
^
P
j
0
x; t
s
for every eigenvalue t
s
. We
divide the eigenvalues into nodal {t
i
} and non-nodal {t
j
} eigenvalues (the
eigenvalues corresponding to the nodal and non-nodal eigenfunctions, respect-
ively), and dene the pairs ft
j
; P
j
ZP
j
0
x; t
j
g and ft
i
; P
i
ZP
j
0
x; t
i
g. Since
P
j
0
x; t
s
z
^
P
j
0
x; t
s
for any eigenvalue t
s
, we have P
i
z
^
P
j
0
x; t
i
and
83 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
P
j
z
^
P
j
0
x; t
j
. We then expect that interpolating P
i
and P
j
gives the
approximate relation P
I
zP
J
z
^
P, where P
I
and P
J
are the respective
interpolants of P
i
and P
j
. Either P
I
or P
J
can then approximate the true
accumulated probability. In this paper, we have taken the average of P
I
and P
J
to approximate
^
P. The time-of-arrival distribution is then obtained by
numerically differentiating this average. This implementation has been found
to hold for the non-interacting case where the exact distribution is known
(Galapon et al. 2005b), and we do not know just yet any another implementation
that correctly reproduces the known exact result. Finally, the CTOA-operator
eigenvalue problem has been solved numerically using Nystroms method
(Delves & Mohamed 1985).
References
Aharonov, Y. & Zubairy, M. S. 2005 Time and the quantum: erasing the past and impacting the
future. Science 307, 875879. (doi:10.1126/science.1107787)
Allcock, G. R. 1969a The time of arrival in quantum mechanics. I. Formal considerations. Ann.
Phys. (N.Y.) 53, 253285. (doi:10.1016/0003-4916(69)90251-6)
Allcock, G. R. 1969b The time of arrival in quantum mechanics. II. The individual measurement.
Ann. Phys. (N.Y.) 53, 286310. (doi:10.1016/0003-4916(69)90252-8)
Allcock, G. R. 1969c The time of arrival in quantum mechanics. III. The measurement ensemble.
Ann. Phys. (N.Y.) 53, 311348. (doi:10.1016/0003-4916(69)90253-X)
Baute, A. D., Mayato, R. S., Palao, J. P., Muga, J. G. & Egusquiza, I. L. 2000 Time-of-arrival
distribution for arbitrary potentials and Wigners time-energy uncertainty relation. Phys.
Rev. A 61, 022 118. (doi:10.1103/PhysRevA.61.022118)
Bhom, D. 1952a A suggested interpretation of the quantum theory in terms of hidden variables I.
Phys. Rev. 85, 166179. (doi:10.1103/PhysRev.85.166)
Bhom, D. 1952b A suggested interpretation of the quantum theory in terms of hidden variables.
II. Phys. Rev. 85, 180193. (doi:10.1103/PhysRev.85.180)
Busch, P., Grabowski, M. & Lahti, P. 1995 Operational quantum physics. Berlin, Germany:
Springer.
Delgado, V. & Muga, J. G. 1997 Arrival time in quantum mechanics. Phys. Rev. A 56, 34253435.
(doi:10.1103/PhysRevA.56.3425)
Delves, L. M. & Mohamed, J. L. 1985 Computational methods for integral equations. Cambridge,
UK: Cambridge University Press.
Everett III, H. 1957 Relative state formulation of quantum mechanics. Rev. Mod. Phys. 29,
454462. (doi:10.1103/RevModPhys.29.454)
Galapon, E. A. 2002a Paulis theorem and quantum canonical pairs: the consistency of a bounded,
self-adjoint time operator canonically conjugate to a Hamiltonian with non-empty point
spectrum. Proc. R. Soc. A 458, 451472. (doi:10.1098/rspa.2001.0874)
Galapon, E. A. 2002b Self-adjoint time operator is the rule for discrete semi-bounded Hamiltonians.
Proc. R. Soc. A 458, 26712689. (doi:10.1098/rspa.2002.0992)
Galapon, E. A. 2004 Shouldnt there be an antithesis to quantization? J. Math. Phys. 45, 3180.
(doi:10.1063/1.1767297)
Galapon, E. A. 2006 Theory of quantum rst time of arrival via spatial connement I: conned
time of arrival operators for continuous potentials. Int. J. Mod. Phys. A. 21, 6351. (doi:10.1142/
S0217751X06034215)
Galapon, E. A., Caballar, R. F. & Bahague Jr, R. T. 2004 Conned quantum time of arrivals. Phys.
Rev. Lett. 93, 180 406. (doi:10.1103/PhysRevLett.93.180406)
Galapon, E. A., Caballar, R. & Bahague, R. T. 2005a Conned quantum time of arrival for
vanishing potential. Phys. Rev. A 72, 062 107. (doi:10.1103/PhysRevA.72.062107)
E. A. Galapon 84
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
Galapon, E. A., Delgado, F., Muga, J. G. & Egusquiza, I. 2005b Transition from discrete to
continuous time-of-arrival distribution for a quantum particle. Phys. Rev. A 72, 042 107.
(doi:10.1103/PhysRevA.72.042107)
Ghirardi, G. C., Ramini, A. & Weber, T. 1986 Unied dynamics for microscopic and macroscopic
systems. Phys. Rev. D 34, 470491. (doi:10.1103/PhysRevD.34.470)
Greene, B. 2004 The fabric of the cosmos. New York, NY: Vintage Books.
Grifths, R. B. 1984 Consistent histories and the interpretation of quantum mechanics. J. Stat.
Phys. 36, 219272. (doi:10.1007/BF01015734)
Grot, N., Rovelli, C. & Tate, R. S. 1996 Time of arrival in quantum mechanics. Phys. Rev. A 54,
46764690. (doi:10.1103/PhysRevA.54.4676)
Halliwell, J. J. & Zaris, E. 1998 Decoherent histories approach to the arrival time problem. Phys.
Rev. D 57, 33513364. (doi:10.1103/PhysRevD.57.3351)
Hegerfeldt, G. C., Seidel, D., Muga, J. G. & Navarro, B. 2004 Operator-normalized quantum
arrival times in the presence of interactions. Phys. Rev. A 70, 012 110. (doi:10.1103/PhysRevA.
70.012110)
Holland, P. R. 1993 The quantum theory of motion. Cambrige, UK: Cambridge University Press.
Jacques, V., Wu, E., Grosshans, F., Treussart, F., Grangier, P., Aspect, A. & Roch, J.-F. 2007
Experimental realization of Wheelers delayed-choice Gedanken experiment. Science 315,
966968. (doi:10.1126/science.1136303)
Kijowski, J. 1974 On the time operator in quantum mechanics and the Heisenberg uncertainty
relation for energy and time. Rep. Math. Phys. 6, 362.
Kochanski, P. & Wodkiewicz, K. 1999 Operational time of arrival in quantum phase space. Phys.
Rev. A 60, 26892699. (doi:10.1103/PhysRevA.60.2689)
Kwiat, P. G., Steinberg, A. M. & Chiao, R. Y. 1992 Observation of a quantum eraser: a revival of
coherence in a two-photon interference experiment. Phys. Rev. A 45, 77297739. (doi:10.1103/
PhysRevA.45.7729)
Leavens, C. R. 1998 Time of arrival in quantum and Bohmian mechanics. Phys. Rev. A 58,
840847. (doi:10.1103/PhysRevA.58.840)
Leo n, J., Julve, J., Pitanga, P. & de Urr es, F. J. 2000 Time of arrival in the presence of
interactions. Phys. Rev. A 61, 062 101. (doi:10.1103/PhysRevA.61.062101)
Muga, J. G. & Leavens, C. R. 2000 Arrival time in quantum mechanics. Phys. Rep. 338, 353438.
(doi:10.1016/S0370-1573(00)00047-8)
Muga, J. G., Sala Mayato, R. & Egusquiza, I. L. (eds) 2002 Time in quantum mechanics. Berlin,
Germany: Springer.
Peres, A. 1995 Quantum theory: concepts and methods. Dordrecht, The Netherlands: Kluwer
Academic Publisher.
Schlosshauer, M. 2004 Decoherence, the measurement problem, and interpretations of quantum
mechanics. Rev. Mod. Phys. 76, 12671305. (doi:10.1103/RevModPhys.76.1267)
Schrodinger, E. 1952 Are there quantum jumps? Br. J. Philos. Sci. 3, 233242. (doi:10.1093/bjps/
III.11.233)
Schrodinger, E. 1980a What is matter? Sci. Am. Particles Fields 11, 1116.
Schrodinger, E. 1980b The present situation in quantum mechanics: a translation of Schrodingers
cat paradox (transl. J. D. Trimmer). Proc. Am. Phil. Soc. 124, 323338.
Scully, M. O. & Druhl, K. 1982 Quantum eraser: a proposed photon correlation experiment
concerning observation and delayed choice in quantum mechanics. Phys. Rev. A 25,
22082213. (doi:10.1103/PhysRevA.25.2208)
Wheeler, J. 1978 The past and the delayed choice double slit experiment. In Mathe-
matical foundations of quantum theory (ed. A. R. Marlow), pp. 948. New York, NY: Academic
Press.
Wheeler J. 1984 Delayed choice. In Quantum theory and measurement (eds J. Wheeler &
W. H. Zurek), pp. 182203. Princeton, NJ: Princeton University Press.
85 Time and particles
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from
Yamada, N. & Takagi, S. 1991a Quantum mechanical probabilities on a general spacetime-surface.
Prog. Theor. Phys. 85, 9851012. (doi:10.1143/PTP.85.985)
Yamada, N. & Takagi, S. 1991b Quantum mechanical probabilities on a general spacetime-surface.
IInontrivial example of non-interfering alternatives in quantum mechanics. Prog. Theor.
Phys. 86, 599615. (doi:10.1143/PTP.86.599)
Yamada, N. & Takagi, S. 1992 Spacetime probabilities in nonrelativistic quantum mechanics.
Prog. Theor. Phys. 87, 7791. (doi:10.1143/PTP.87.77)
Zurek, W. H. 2003 Decoherence, einselection, and the quantum origins of the classical. Rev. Mod.
Phys. 75, 715775. (doi:10.1103/RevModPhys.75.715)
E. A. Galapon 86
Proc. R. Soc. A (2009)
on December 20, 2012 rspa.royalsocietypublishing.org Downloaded from

Das könnte Ihnen auch gefallen