ArticlePDF Available

Feynman Paths and Weak Values

Authors:

Abstract and Figures

There has been a recent revival of interest in the notion of a 'trajectory' of a quantum particle. In this paper, we detail the relationship between Dirac's ideas, Feynman paths and the Bohm approach. The key to the relationship is the weak value of the momentum which Feynman calls a transition probability amplitude. With this identification we are able to conclude that a Bohm 'trajectory' is the average of an ensemble of actual individual stochastic Feynman paths. This implies that they can be interpreted as the mean momentum flow of a set of individual quantum processes and not the path of an individual particle. This enables us to give a clearer account of the experimental two-slit results of Kocsis et al.
Content may be subject to copyright.
entropy
Article
Feynman Paths and Weak Values
Robert Flack and Basil J. Hiley *
Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK;
r.flack@ucl.ac.uk
*Correspondence: b.hiley@bbk.ac.uk
Received: 16 April 2018; Accepted: 9 May 2018; Published: 14 May 2018


Abstract:
There has been a recent revival of interest in the notion of a ‘trajectory’ of a quantum
particle. In this paper, we detail the relationship between Dirac’s ideas, Feynman paths and the
Bohm approach. The key to the relationship is the weak value of the momentum which Feynman
calls a transition probability amplitude. With this identification we are able to conclude that a Bohm
‘trajectory’ is the average of an ensemble of actual individual stochastic Feynman paths. This implies
that they can be interpreted as the mean momentum flow of a set of individual quantum processes
and not the path of an individual particle. This enables us to give a clearer account of the experimental
two-slit results of Kocsis et al.
Keywords: Feynman paths; weak values; Bohm theory
1. Introduction
One of the basic tenets of quantum mechanics is that the notion of a particle trajectory has
no meaning. The established view has been unambiguously defined by Landau and Lifshitz [
1
]:
“In quantum mechanics there is no such concept as the path of a particle”. This position was not
arrived at without an extensive discussion going back to the early debates of Bohr and Einstein [
2
],
the pioneering work of Heisenberg [3] and many others [4].
Yet Kocsis et al. [
5
] have experimentally determined an ensemble of what they call ‘photon
trajectories’ for individual photons traversing a two-slit interference experiment. The set of trajectories,
or what we will call flow-lines, they construct is very similar in appearance to the ensemble of Bohmian
trajectories calculated by Philippidis et al. [
6
]. Mahler et al. [
7
] have gone further and claimed that
their new experimental results provide evidence in support of Bohmian mechanics. However such a
claim cannot be correct because Bohmian mechanics is based on the Schrödinger equation which holds
only for non-relativistic particles with non-zero rest mass, whereas photons are relativistic, having
zero rest mass.
The flow-lines are calculated from experimentally determined weak values of the momentum
operator, a notion that was introduced originally by Aharonov et al. [
8
] for the spin operator. When
examined closely, the momentum weak value is the Feynman transition probability amplitude (TPA) [
9
].
In fact, Schwinger [
10
] explicitly writes the TPA of the momentum in exactly the same form as the weak
value. Recall that the TPA involving the momentum operator plays a central role in the discussion of
the path integral method, an approach that was inspired by an earlier paper of Dirac [
11
] who was
interested in developing the notion of a ‘quantum trajectory’.
Weak values are in general complex numbers, as are TPAs. The real part of the momentum
weak value is the local momentum, sometimes known as the Bohm momentum. The imaginary part
turns out to be the osmotic momentum introduced by Nelson [
12
] in his stochastic derivation of the
Schrödinger equation. In this paper, we will show how the weak value of momentum, Feynman
paths and the Bohm trajectories are related enabling us to give a different meaning to the flow-lines
constructed in experiments of the type carried out by Kocsis et al. [5] and Mahler et al. [7].
Entropy 2018,20, 367; doi:10.3390/e20050367 www.mdpi.com/journal/entropy
Entropy 2018,20, 367 2 of 11
Feynman [
9
] also shows that in his approach the usual expression for the kinetic energy becomes
infinite unless one introduces a small fluctuation in the mass of the particle. We will show that this is
equivalent to introducing the quantum potential, a new quality of energy that appears in the real part
of the Schrödinger equation under polar decomposition of the wave function [13].
2. Dirac’s Notion of a Quantum Trajectory
2.1. Dirac Trajectories
To make the context of our discussion clear, we will begin by drawing attention to an early paper
by Dirac [
11
] who attempted to generalise the Heisenberg algebraic approach through his unique
bra-ket notation, not as elements in a Hilbert space, but as elements of a non-commutative algebra.
In this approach the operators of the algebra are functions of time. Dirac argued that to get round
the difficulties presented by a non-commutative quantum algebra, strict attention must be paid to the
time-order of the appearance of elements in a sequence of operators.
In the non-relativistic limit, operators at different times always commute. (In this paper, we will,
for simplicity, only consider the non-relativistic domain. Dirac himself shows how the ideas can be
extended to the relativistic domain.) This means that a time ordered sequence of position operators
can be written in the form,
hxt|xt0i=Z···Zhxt|xtjidxjhxtj|xtj1i. . . hxt2|xt1idx1hxt1|xt0i. (1)
This breaks the TPA,
hxt|xt0i
, into a sequence of adjacent points, each pair connected by an
infinitesimal TPA. Dirac writes “. . . one can regard this as a trajectory .. . and thus makes quantum
mechanics more closely resemble classical mechanics”.
In order to analyse the sequence in Equation (1) further, Dirac assumed that for a small time
interval t=e, we can write
hx|x0ie=exp[iSe(x,x0)/¯h](2)
where we will take Se(x,x0)to be a real function in the first instance. Then Dirac [14] shows that
p0
e(x,x0) = hx|ˆ
P0|x0ie=i¯hx0hx|x0ie=−∇x0Se(x,x0)hx|x0ie(3)
and
pe(x,x0) = hx|ˆ
P|x0ie=i¯hxhx|x0ie=xSe(x,x0)hx|x0ie. (4)
Here
ˆ
P
is the momentum operator. The remarkable similarity of these objects to the canonical
momentum appearing in the classical Hamilton-Jacobi theory should be noted, a fact of which Dirac
was well aware. They are also the canonical momenta appearing in the real part of the Schrödinger
equation under polar decomposition of the wave function exploited by Bohm [
13
] who identified the
momentum with the gradient of the phase of the wave function.
In an earlier paper, Dirac [
11
] did not specify how
Se(x
,
x0)
could be determined. It was
Feynman [
9
] who later identified its relation to the classical Lagrangian
L(˙
x
,
x
,
t)
through the relation
Stt0(x,x0) = Min Zt
t0L(˙
x,x,t)dt. (5)
However, this Lagrangian determines the classical path, so using the exponent of the classical
action seems puzzling. Is there a mathematical explanation for such a choice? The answer is ‘yes’
and is discussed in Guillemin and Sternberg [
15
]. The essential reason for this lies in the relation
between the symmetry group, in this case the symplectic group, and its covering group. Exploiting
this structure, de Gosson and Hiley [
16
] have shown in detail how it is possible to mathematically
Entropy 2018,20, 367 3 of 11
‘lift’ classical trajectories onto this covering space. It is from this structure that the wave properties
emerge. The lift is achieved by exponentiating the classical action, namely using
exp[iSe(x
,
x0)]
. It is
the existence of this structure that the close relation between the Dirac quantum ‘trajectories’ and the
de Broglie-Bohm ‘trajectories’ first calculated by Philippidis et al. [
6
] emerges. We will bring out this
relationship in the rest of this paper.
2.2. The Feynman Propagator
Equation (5) allows us to write the propagator in the well known form
K(x,x0) = Zx
x0eiS(x,x0)Dx0
where the integral is taken over all paths connecting
x0
to
x
. We have written
Dx0
for
dx0
A
,
. . .
,
dxj1
A
where
(x0
,
x1
,
. . .
,
xj1)
are points on the path and
A
is the normalising factor introduced by Feynman.
Clearly here S(x,x0)is real.
For a free particle with mass m, we have L=m˙
x2/2 and one can show that
Ktt0(x,x0) = 1
Aexp im(xx0)2
h(tt0)(6)
where
A=2πi(tt0)
m1/2
. With this propagator, Feynman was able to derive the Schrödinger equation
by assuming the underlying paths were continuous and differentiable.
However if we examine the terms
hx|x0ie
for
e
0, we find the curves, although continuous,
are non-differentiable. To show this let us introduce the TPA of a function F(x,t)defined by
hφt|F|ψt0iS=Lime0Z···Zφ(x,t)F(x0,x1, . . . , xj)
×exp "i
¯h
j1
k=0
S(xk+1,xk)#ψ(x0,t0)Dx(t).
Here Dis now written as Dx(t) = dx0
A. . . dxj1
Adxj.
These TPAs can be evaluated by using functional derivatives. In fact, the average of the functional
derivative of a function F(x,t)is given by
δF
δx(s)S
=i
¯hFδS
δx(s)S
(7)
at the point x(s)on the path x(t). In the case of the specific integral
ZF
xk
exp[(i/¯h)S(x(t))]Dx(t),
Equation (7) can be written in the form
F
xkS
=i
¯hFS
xkS
.
Feynman notes that the quantities in this expression need not be observables, nevertheless the
equivalence is true [17].
Entropy 2018,20, 367 4 of 11
Let us now consider three adjacent points
xk1
,
xk
,
xk+1
, each separated by a small time
difference e, we have
¯h
iF
xkS
=FS(xk+1,xk)
xk
+S(xk,xk1)
xkS
.
This equation is correct to zero and first order in
e
. If we choose the action for a particle moving
in a potential V, we have
S(x,x0) = m(xx0)2
2eeV(x,x0).
Then at the point xkthis gives us
¯h
iF
xkS
=Fmxk+1xk
exkxk1
eeV
xk
(xk).
If Fis unity and we divide by ewe get
0=1
emxk+1xk
exkxk1
eV
xk
(xk). (8)
If we follow Feynman and call
(xk+1xk)/e
a ‘velocity’, then this equation gives the ‘average’
over an ensemble of individual velocities. It is the quantum equivalent of Newton’s second law of motion;
the potential
V
at
xk
gives rise to a force which changes the incoming momentum
m(xkxk1)/e
to the outgoing momentum
m(xk+1xk)/e
. Notice to order
e
, no extra term corresponding to the
quantum potential appears. de Gosson and Hiley [
18
] have shown in a detailed analysis that this is to
be expected.
These paths are reminiscent of Brownian motion, a characteristic feature of which is the appearance
of two ‘derivatives’ at
xk
, a ‘forward’ and a ‘backward’ derivative, illustrating the non-differentiable
nature of the path. In this paper, we need not discuss the precise nature of these paths to arrive at
our conclusion. It is sufficient for us to note that the substructure of a quantum process is certainly
not classical. In passing we should also note that the ‘velocities’, being of order
(¯h/me)1/2
, diverge as
e0 and therefore, in Feynman’s terms, are not observables.
2.3. TPAs Involving the Momentum
In 1974 Hirschfelder [
19
,
20
] introduced a quantity
ψ(x
,
t)1ˆ
pψ(x
,
t)
, which he called a
‘sub-observable’ as he could see no way of measuring it directly, although integrating it over the
whole of configuration space gave the measurable expectation value. Using the polar form of the wave
function,
ψ(x
,
t) = R(x
,
t)exp[iS(x
,
t)h]
, this ‘sub-observable’ is the weak value of the momentum
operator which can be written in the form
ψ(x,t)1ˆ
pψ(x,t) = hx|ˆ
p|ψ(t)i
hx|ψ(t)i=m[vB(x,t)ivO(x,t)], (9)
where explicitly
vB(x
,
t) = S(x
,
t)/m
is the local Bohm velocity and
vO(x
,
t) = R(x
,
t)/mR(x
,
t)
is
the localising osmotic velocity, originally introduced by Nelson [
12
] in a stochastic theory. The meaning
of these velocities is discussed in more detail in Bohm and Hiley [
21
]. Much later Hiley [
22
] showed
exactly how these expressions emerged directly from the weak value of the momentum operator.
It should be noted that weak values are essentially TPAs of the type considered by Feynman [
9
] and
Schwinger [23].
In the spirit of Schwinger [
10
], where he argues that “the quantum dynamical laws will find
their proper expression in terms of the transformation functions” that is TPAs, we can introduce two
Entropy 2018,20, 367 5 of 11
momentum TPAs,
hx|
P|ψ(t)i
and
hψ(t)|
P|xi
where
P=i¯h
and
P=i¯h
. Notice by placing
the arrows over the momentum operators, we are emphasising the distinction between left and right
multiplication and it is this distinction that is equivalent to the forward and backward derivatives.
In fact we may identify
hX|
P|ψ(t0)i=hX|
P|x0iψ(x0,t0) = ilim
(x0X)
ψ(X)ψ(x0)
(Xx0)
with the forward derivative at X, a point that lies between x0and x.
hψ(t)|
P|Xi=ψ(x,t)hx|
P|Xi=ilim
(Xx)
ψ(x)ψ(X)
(xX)
corresponds to the backward derivative. Note that the words ‘forward’ and ‘backward’ here have
nothing to do with time order.
If we again evaluate these TPAs using ψ=Rexp(iS/¯h), we find
1
2
hx|
ˆ
P|ψ(t)i
hx|ψ(t)i+hψ(t)|
ˆ
P|xi
hψ(t)|xi
=S(x,t) = PB(x,t), (10)
and
1
2i
hx|
ˆ
P|ψ(t)i
hx|ψ(t)ihψ(t)|
ˆ
P|xi
hψ(t)|xi
=R(x,t)
R(x,t)=PO(x,t). (11)
Notice how the sums and differences of the left/right operators produce real values.
We can immediately connect these results with those of Dirac [
11
] if, in Equations (3) and (4),
we replace the real value of
Se(x
,
x0)
by a complex value which we will write as
S0
e(x
,
x0) = Se(x
,
x0)
iln Re(x,x0). In this case, we find
p0
e(x,x0) = −∇x0Se(x,x0)ix0Re(x,x0)
Re(x,x0)(12)
and
pe(x,x0) = xSe(x,x0)ixRe(x,x0)
Re(x,x0). (13)
Notice also the connection with the classical relations obtained in Equations (3) and (4).
2.4. The Relation between Weak Values and TPAs
In the previous two sections, we have shown how TPAs of the form
hφt|ˆ
F|ψt0i
arise from some
underlying non-differentiable process. The original assumption was that these quantities could not be
investigated experimentally. However starting from a different perspective, the notion of a weak value,
introduced by Aharonov, Albert and Vaidman [
8
], allows us to experimentally measure these quantities.
A weak value of an operator ˆ
Fis defined by
hˆ
Fiw=hφt|ˆ
F|ψt0i
hφt|ψt0i.
Clearly these weak values are Feynman TPAs. Using the suggestions of Leavens [
24
] and
Wiseman [
25
], Kocsis et al. [
5
] have actually measured the weak value of the transverse momentum in
an optical two-slit experiment and as a result have constructed what they called photon ‘trajectories’.
We refer to their paper to explain the details of how this is done.
Entropy 2018,20, 367 6 of 11
Unfortunately photons cannot be treated as particles that satisfy the Schrödinger equation.
They have zero rest mass and are excitations of the electromagnetic field. Nevertheless this does
not invalidate the notion of a momentum flow line; the question remains “How are we to understand
these flow lines?” Flack and Hiley [
26
] have shown that if we generalise the Bohm approach to include
the electromagnetic field [27], each flow line emerges as the locus of a weak Poynting vector.
To connect with the non-relativistic approach we are discussing in this paper, we need to use atoms.
Indeed experiments are being developed at UCL to measure weak values of spin and momentum,
hˆ
piw
, for helium atoms [
28
] and argon atoms [
29
] respectively. The experimental details can be found
in these references. In this paper, we will clarify further the relation between the Feynman paths and
weak values.
3. Weak Values Are Weighted TPAs
3.1. Flow Lines Constructed from Weak Values
In quantum mechanics, the uncertainty principle does not allow us to give meaning to the
‘trajectory’ of a single particle so we are left with the question: “How does a particle get from
A
to
B
?”.
Rather than taking two points, consider two small volumes,
V0(x0)
surrounding the point
A=x0
and
V(x)
surrounding
B=x
. We assume these volumes are initially large enough to avoid problems
with the uncertainty principle.
Now imagine a sequence of particles emanating from
V0(x0)
, each with a different momentum.
Over time we will have a spray of possible momenta emerging from the volume
V0(x0)
, the nature
of this spray depending on the size of
V0(x0)
. Similarly there will be a spray of momenta over time
arriving at the small volume V(x)surrounding the point x.
Better still let us consider a small volume surrounding the midpoint
X
. At this point there is
a spray arriving and a spray leaving a volume
V(X)
as shown in Figure 1. To see how the local
momenta behave at the midpoint X, we will use the real part of S0
e(x,x0)defined by
Se(x,x0) = m
2
(xx0)2
e. (14)
X
(x',t')
Figure 1. Behaviour of the momenta sprays at the midpoint of hx,t|x0,t0ie.
Let us define a quantity
PX(x,x0) = Se(x,x0)
X=Se(X,x0)
X+Se(x,X)
X, (15)
then using the action (Equation (14)), we find
PX(x,x0) = m(Xx0)
e(xX)
e=p0
X(x,x0) + pX(x,x0). (16)
Not surprisingly, this is exactly what Feynman [
9
] is averaging over at the point
X
, agreeing with
the term between the brace [. . . ].
Entropy 2018,20, 367 7 of 11
What is more important is the relation of Equation (16) to Equation (10) which is the real part of
the weak value of the momentum operator. Thus, the mean momentum of a set of Feynman paths at
X
is clearly the real part of this weak value. However, this weak value is just the Bohm momentum.
Thus the Bohm ‘trajectories’ are simply an ensemble of the average of the ensemble of individual
Feynman paths.
To see how this unexpected result also emerges from a different perspective, let us consider
the process in Figure 1which we regard as an image of an ensemble of actual individual quantum
processes. We are interested in finding the average behaviour of the momentum,
PX
, at the
point X
.
However, we have two contributions to consider, one coming from the point
x0
and one leaving for
the
point x
. We must determine the distribution of momenta in each spray to produce a result that
is consistent with the wave function
ψ(X)
at
X
. Feynman suggests [
9
] that we can think of
ψ(X)
as ‘information coming from the past’ and
ψ(X)
as ‘potential information appearing in the future’.
This suggests that we can write
lim
x0Xψ(x0) = Zφ(p0)eip0Xdp0and lim
Xxψ(x) = Zφ(p)eip X dp.
The
φ(p0)
contains information regarding the probability distribution of the incoming momentum
spray, while
φ(p)
contains information about the probability distribution in the outgoing momentum
spray. These wave functions must be such that in the limit
e
0 they are consistent with the wave
function ψ(X).
Thus, we can define the mean momentum, P(X), at the point Xas
ρ(X)P(X) = Z Z Pφ(p)eipX φ(p0)eip0Xδ(P(p0+p)/2)dPdpdp0(17)
where
ρ(X)
is the probability density at
X
. We have added the restriction
δ(P(p0+p)/
2
)
since
momentum is conserved at X. We can rewrite Equation (17) and form
ρ(X)P(X) = 1
2πZ Z Pφ(p+θ/2)eiXθφ(pθ/2)dθdP
or equivalently taking Fourier transforms
ρ(X)P(X) = 1
2πZ Z Pψ(Xσ/2)eiPσψ(X+σ/2)dσdP
which means that
P(X)
is the conditional expectation value of the momentum weighted by the Wigner
function. Equation (17) can be put in the form
ρ(X)P(X) = 1
2i[(x1x2)ψ(x1)ψ(x2)]x1=x2=X(18)
an equation that appears in the Moyal approach [
30
], which is based on a different non-commutative
algebra. If we evaluate this expression for the wave function written in polar form
ψ(x) = R(x)exp[iS(x)]
, we find
P(X) = S(X)
which is identical to the expression for the local
(Bohm) momentum used in the Bohm interpretation.
This then confirms the conclusion we reached above, namely, that the set of Bohm ‘trajectories’ is
an ensemble of the average ensemble of individual paths. Notice, once again, that this gives a very
different picture of the Bohm momentum from the usual one used in Bohmian mechanics [
31
]. It is
not the momentum of a single ‘particle’ passing the point
X
, but the mean momentum flow at the point
in question.
This conclusion is supported by the experiments of Kocsis et al. [
5
]. They construct the flow lines
from an average made over many individual input photons. Thus, the so-called ‘photon’ flow-lines
Entropy 2018,20, 367 8 of 11
are constructed statistically from an ensemble of individual events. As was shown in Flack and
Hiley [
26
], these flow lines are an average of the momentum flow as described by the weak value of the
Poynting vector. This agrees with what one would expect from standard quantum electrodynamics,
where the notion of a ‘photon trajectory’ has no meaning, but the notion of a ‘momentum flow’ does
have meaning.
Bliokh et al. [
32
] have presented a beautiful illustration showing the results of a two-slit
interference experiment. Figure 2a shows the real part of the momentum flow lines in the
electromagnetic field, while the imaginary component (osmotic) momentum flow lines are shown in
Figure 2b. It is then clear that we can regard
vB(x
,
t) = pB(x
,
t)/m
as a local velocity, while the
osmotic velocity
vO(x
,
t) = pO(x
,
t)/m
can be regarded as a localising velocity as discussed in
Bohm and Hiley [
33
]. The osmotic velocity behaves in such a way as to maintain the form of the
probability distribution.
Figure 2. (a) Local field momentum; (b) Localising field momentum.
3.2. Where Is the Quantum Potential?
One of the features that many find ‘mysterious’ [
34
] is the appearance of the ‘quantum potential’
in the Bohm approach. Is there any trace of it in the Feynman paper [
9
]? To answer this question,
we must first refer to de Gosson and Hiley [
18
] where it is shown that this energy term is absent in
quantum processes when taken only to O(t=e)so we must consider terms to O(t=e2).
Feynman shows that the kinetic energy is of
O(e2)
when written in the form
K.E. = [(xk+1xk)/e]2
, and diverges as
e
0. Feynman points out that this quantity is not an
observable functional. However, let us now define the kinetic energy to be
K.E.0=m
2xk+1xk
exkxk1
e.
This function is finite to
O(e)
and therefore is an observable functional. Feynman then shows that
if we allow “the mass to change by a small amount to
m(
1
+δ)
for a short time, say
e
around
tk
” we
can obtain the relation
m
2xk+1xk
exkxk1
e=m
2xk+1xk
e2
+¯h
2ie, (19)
the extra term arising from the normalising function
A
. Thus, we must add a ‘correction’ term to the
K.E. in order for the total energy to be finite to O(e2).
This is the forerunner of mass renormalisation used in quantum electrodynamics. In that case the
charged particle is subjected to electromagnetic vacuum fluctuations. The particle we are considering
here is not charged and so the fluctuation must arise from a different source, but however it arises,
it changes the TPA by δ.
Entropy 2018,20, 367 9 of 11
Later in the same paper, Feynman shows that any random fluctuation in the phase function
will produce the same effect. A random fluctuation at the point
xk
implies we must replace
S(xk+1,tk+1;xk,tk)by Sδ(xk+1,tk+1;xk,tkδ). Thus, to the first order in δwe have
hξ|1|ψiShξ|1|ψiSδ=iδ
¯hhξ|Hk|ψiS
where Hkis the Hamiltonian functional
Hk=S(xk+1,tk+1;xk,tk)
tk+1
+¯h
2i(tk+1tk). (20)
Apart from the minus sign, the last term is identical to the last term in Equation (19).
Thus Feynman required extra energy to appear from somewhere. A more detailed discussion of
this feature appears in Feynman and Hibbs [
35
]. The Bohm approach indicates that some ‘extra’ energy
appears in the form of the quantum potential energy at the expense of the kinetic energy. Could it be
that the source of the energy is the same?
To explore this possibility, let us use the method explained in Section 2.3 to obtain a more general
result for the K.E. The real part of the weak value of the momentum operator squared is obtained from
hψ(t)|ˆ
p2|xi+hx|ˆ
p2|ψ(t)i/
2. Under polar decomposition of the wave function, we find the real part
of the weak value of the kinetic energy is
1
2mhˆ
p2iw=1
2m(S)22R
R. (21)
With the identification
Sm(xk+1xk)/e
, we see that the quantum potential is playing a
similar role as the mass/energy fluctuation in Feynman’s approach. In fact, de Broglie’s original
suggestion was that the quantum potential could be associated with a change of the rest mass [36].
Notice that the quantum potential appears essentially as a derivative of the osmotic velocity,
which in turn is obtained from the imaginary part of
S0(x
,
x0)
. Any fluctuating term added to the real
part of
Se(x
,
x0)
should also be added to the imaginary part. This would also introduce some change
in the energy relation shown in Equation (20). This interplay between the real components of the
complex
Se(x
,
x0)
is clearly presented as an average over fluctuations arising from some background.
Here we can recall Bohr insisting that quantum phenomena must include a description of the whole
experimental arrangement. More details will be found in Smolin [37] and in Hiley [38].
4. Conclusions
Our explorations of the weak values of the momentum operator [
22
] have led us to reconsider the
basis on which Feynman [
9
] developed his path integral approach. We have shown that there is an
unexpected close connection between the Feynman propagator, the weak values of the momentum
and the original Bohm approach [13].
Feynman had already noticed that to prevent the kinetic energy tending to infinity as the time
interval between steps tends to zero, it was necessary to introduce a ‘fluctuation’ in the mass/energy
of the particle. This extra energy can be thought of as arising in a way similar to the way the quantum
potential energy appears as an effect of some background field. Indeed, as we have remarked above,
de Broglie [
36
] had already proposed that the quantum potential could be included in the mass term
M=[m2+ (¯h2/c2)R/R]
,
R
being the amplitude of the wave function. Hiley [
38
] has shown a
similar conclusion arises for the Dirac equation.
The approach outlined in this paper shows that the basic assumption made in Bohmian mechanics,
namely, that each particle follows one of the ensemble of ‘trajectories’ calculated by
Philippidis et al. [6]
from
PB(x
,
t)
cannot be maintained. Rather the trajectories should be interpreted as a statistical average
of the momentum flow of a basic underlying stochastic process.
Entropy 2018,20, 367 10 of 11
It is now possible to experimentally explore weak values, perhaps clarifying the nature of this
stochastic process. In the case of the electromagnetic field this has already been done by
Kocsis et al. [5]
,
but as we have seen the notion of a ‘photon trajectory’ has no meaning. However, the average
momentum flow does have meaning [
26
]. As mentioned above, new experiments using argon and
helium atoms are now being carried out at UCL by Morley et al. [
29
] and by Monachello, Flack, and
Hiley [
28
]. It is hoped that these future experiments will throw more light on the nature of individual
quantum processes.
Author Contributions: Both authors contributed equally to this manuscript.
Funding: This research was funded in part by the Fetzer Franklin Memorial Trust.
Acknowledgments:
Special thanks to Bob Callaghan, Glen Dennis and Lindon Neil for their helpful discussions.
Thanks also to the Franklin Fetzer Foundation for their financial support.
Conflicts of Interest:
The founding sponsors had no role in the design of the study; in the collection, analyses,
or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.
References
1.
Landau, L.D.; Lifshitz, E.M. Quantum Mechanics: Non-Relativistic Theory; Pergamon Press: Oxford, UK, 1977.
2.
Einstein, A. Albert Einstein: Philosopher-Scientist; Schilpp, A.P., Ed.; Library of the Living Philosophers:
Evanston, IL, USA, 1949; pp. 665–676.
3.
Heisenberg, W. Physics and Philosophy: The Revolution in Modern Science; George Allen and Unwin: London,
UK, 1958.
4. Jammer, M. The Philosophy of Quantum Mechanics; Wiley: New York, NY, USA, 1974.
5.
Kocsis, S.; Braverman, B.; Ravets, S.; Stevens, M.J.; Mirin, R.P.; Shalm, L.K.;
Steinberg, A.M.
Observing the
average trajectories of single photons in a two-slit interferometer. Science
2011
,332, 1170–1173. [CrossRef]
[PubMed]
6.
Philippidis, C.; Dewdney, C.; Hiley, B.J. Quantum interference and the quantum potential. Il Nuovo Cimento B
1979,52, 15–28. [CrossRef]
7.
Mahler, D.H.; Rozema, L.A.; Fisher, K.; Vermeyden, L.; Resch, K.J.; Braverman, B.; Wiseman, H.M.;
Steinberg, A.M.
Measuring bohm trajectories of entangled photons. In Proceedings of the 2014 Conference
on Lasers and Electro-Optics (CLEO)—Laser Science to Photonic Applications, San Jose, CA, USA,
8–13 June 2014.
8.
Aharonov, Y.; Albert, D.Z.; Vaidman, L. How the result of a measurement of a component of the spin of a
spin-1/2 particle can turn out to be 100. Phys. Rev. Lett. 1988,60, 1351–1354. [CrossRef] [PubMed]
9.
Feynman, R.P. Space-time approach to non-relativistic quantum mechanics. Rev. Mod. Phys.
1948
,20,
367–387. [CrossRef]
10. Schwinger, J. The theory of quantised fields I. Phys. Rev. 1951,82, 914–927. [CrossRef]
11.
Dirac, P.A.M. On the analogy between Classical and Quantum Mechanics. Rev. Mod. Phys.
1945
,17, 195–199.
[CrossRef]
12.
Nelson, E. Derivation of schrödinger’s equation from newtonian mechanics. Phys. Rev.
1966
,150, 1079–1085.
[CrossRef]
13.
Bohm, D. A suggested interpretation of the quantum theory in terms of hidden variables, I. Phys. Rev.
1952
,
85, 180–193. [CrossRef]
14. Dirac, P.A.M. The Principles of Quantum Mechanics; Oxford University Press: Oxford, UK, 1947.
15.
Guillemin, V.W.; Sternberg, S. Symplectic Techniques in Physics; Cambridge University Press: Cambridge,
UK, 1984.
16.
De Gosson, M.; Hiley, B.J. Imprints of the quantum world in classical mechanics. Found. Phys.
2011
,41,
1415–1436. [CrossRef]
17. Brown, L. Feynman’s Thesis: A New Approach to Quantum Mechanics; World Scientific Press: Singapore, 2005.
18.
De Gosson, M.; Hiley, B.J. Short-time quantum propagator and bohmian trajectories. Phys. Lett.
2013
,377,
3005–3008. [CrossRef] [PubMed]
19.
Hirschfelder, J.O.; Christoph, A.C.; Palke, W.E. Quantum mechanical streamlines I. square potential barrier.
J. Chem. Phys. 1974,61, 5435–5455. [CrossRef]
Entropy 2018,20, 367 11 of 11
20.
Hirschfelder, J.O. Quantum mechanical equations of change. I. J.Chem. Phys.
1978
,68, 5151–5162. [CrossRef]
21.
Bohm, D.; Hiley, B.J. Non-locality and locality in the stochastic interpretation of quantum mechanics.
Phys. Rep. 1989,172, 93–122. [CrossRef]
22.
Hiley, B.J. Weak values: Approach through the Clifford and Moyal algebras. J. Phys. Conf. Ser.
2012
,
361, 012014. [CrossRef]
23. Schwinger, J. The theory of quantum fields III. Phys. Rev. 1953,91, 728–740. [CrossRef]
24.
Leavens, C.R. Weak measurements from the point of view of bohmian mechanics. Found. Phys.
2005
,35,
469–491. [CrossRef]
25.
Wiseman, H.M. Grounding bohmian mechanics in weak values and bayesianism. New J. Phys.
2007
,9,
165–177. [CrossRef]
26.
Flack, R.; Hiley, B.J. Weak values of momentum of the electromagnetic field: Average momentum flow lines,
not photon trajectories. arXiv 2016, arXiv:1611.06510.
27.
Bohm, D.; Hiley, B.J.; Kaloyerou, P.N. An ontological basis for the quantum theory: II—A causal interpretation
of quantum fields. Phys. Rep. 1987,144, 349–375. [CrossRef]
28.
Monachello, V.; Flack, R.; Hiley, B.J. A method for measuring the real part of the weak value of spin using
non-zero mass particles. arXiv 2017, arXiv:1701.04808.
29.
Morley, J.; Edmunds, P.D.; Barker, P.F. Measuring the weak value of the momentum in a double slit
interferometer. J. Phys. Conf. Ser. 2016,701, 012030. [CrossRef]
30. Moyal, J.E. Quantum mechanics as a statistical theory. Proc. Camb. Phil. Soc.1949,45, 99–123. [CrossRef]
31. Dürr, D.; Teufel, S. Bohmian Mechanics; Springer: Berlin/Heidelberg, Germany, 2009.
32.
Bliokh, K.Y.; Bekshaev, A.Y.; Kofman, A.G.; Nori, F. Photon trajectories, anomalous velocities and weak
measurements: A classical interpretation. New J. Phys. 2013,15, 073022. [CrossRef]
33.
Bohm, D.; Hiley, B.J. The Undivided Universe: An Ontological Interpretation of Quantum Theory; Routledge:
London, UK, 1993.
34.
Feynman, R.P.; Leighton, R.B.; Sands, M. The Feynman Lectures on Physics; III, Sec 21-8; Addison-Wesley:
Boston, MA, USA, 1965.
35. Feynman, R.P.; Hibbs, A.R. Quantum Mechanics and Path Integrals; McGraw-Hill: New York, NY, USA, 1965.
36.
De Broglie, L. Non-Linear Wave Mechanics: A Causal Interpretation; Elsevier: Amsterdam, The Netherlands,
1960; p. 116.
37.
Smolin, L. Quantum mechanics and the principle of maximal variety. Found. Phys.
2016
,46, 736–758.
[CrossRef]
38. Hiley, B.J. On the Nature of a Quantum Particle. 2018, in press.
c
2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... In this work, we consider momentum conditionally averaged on positions following Moyal's [26] and Sonego's [38] arguments. Also, it was previously realized by many researchers thatp are the real, measurable part of the weak value of the momenta [32,76,77]. Therefore, the conditionally averaged momenta act as an effective momenta of a system. ...
... It being equal to the conditionally averaged momentum also shows that it is an effective object. This was also realized in certain other studies [38] and its relation to the weak measurements of Aharonov et al. [72] has been established before [32,76,77]. ...
Article
Full-text available
The hydrodynamic interpretation of quantum mechanics treats a system of particles in an effective manner which allows one to study the system in a statistical fashion. In this work, we investigate squeezed coherent states within the hydrodynamic interpretation. The Hamiltonian operator in question is time dependent, n-dimensional and in quadratic order. We start by deriving a phase space Wigner probability distribution and an associated equilibrium entropy for the squeezed coherent states. Then, we decompose the joint phase space distribution into two portions: A marginal position distribution and a momentum distribution that is conditioned on the post-selection of positions. Our conditionally averaged momenta are shown to be equal to the Bohm’s momenta whose connection to the weak measurements is already known. We also keep track of the corresponding classical system evolution by identifying shear, magnification and rotation components of the symplectic phase space dynamics. This allows us to pinpoint which portion of the underlying classical motion appears in which quantum statistical concept. We show that our probability distributions satisfy the Fokker–Planck equations exactly and they can be used to decompose the equilibrium entropy into the missing information in positions and in momenta as in the Sackur–Tetrode entropy of the classical kinetic theory. Eventually, we define a quantum pressure, a quantum temperature and a quantum internal energy which are related to each other in the same fashion as in the classical kinetic theory. We show that the quantum potential incorporates the kinetic part of the internal energy and the fluctuations around it. This allows us to suggest a quantum conditional virial relation. In the end, we show that the kinetic internal energy is linked to the fractional Fourier transformer part of the underlying classical dynamics similar to the case where the energy of a quantum oscillator is linked to its Maslov index.
... cf. Flack and Hiley [21]. Note that the minimum (10) can also be interpreted as E ψ (p 2 O,j ). ...
Preprint
Full-text available
Given a normalized state-vector ψ\psi , we define the conditional expectation Eψ(AB)\mathbb{E }_{\psi } (A | B ) of a Hermitian operator A with respect to a strongly commuting family of self-adjoint operators B as the best approximation, in the operator mean square norm associated to ψ\psi , of A by a real-valued function of B.B . A fundamental example is the conditional expectation of the momentum operator P given the position operator X , which is found to be the Bohm momentum. After developing the Bohm theory from this point of view we treat conditional expectations with respect to general B , which we apply to non-relativistic spin 1/2-particles. We derive the dynamics of the conditional expectations of momentum and spin with respect to position and the third spin component. These dynamics can be interpreted in terms of classical continuum mechanics as a two-component fluid whose components carry intrinsic angular momentum. Interpreting the joint spectrum of the conditioning operators as a space of beables, we can introduce a classical-stochastic particle dynamics on this space which is compatible with the time-evolution of the Born probability, by combining the de Broglie-Bohm guidance condition with a Markov jump process, following an idea of J. Bell. This results in a new Bohm-type model for particles with spin. A basic problem is that such auxiliary particle dynamics are far from unique. We finally examine the relation of our conditional expectations with the conditional expectations of the theory of CC^* -algebras and, as an application, derive a general evolution equation for conditional expectations for operators acting on finite dimensional Hilbert spaces. Two appendices re-interpret the classical Bohm model as an integrable constrained Hamiltonian system, and provide the details of the two-fluid interpretation.
... In this way, the continuous "trajectories" calculated by Philippidis, Dewdney and Hiley [13] appear as energy flow lines analogous to those shown in Berry [11,12] for optical flow lines. Furthermore, it can be shown that the quantum "trajectories" are an average of an ensemble of Feynman paths [50]. This then provides us with an intuitive image of the Heisenberg equation of motion (Equation (11)). ...
Article
Full-text available
What is striking about de Broglie’s foundational work on wave–particle dualism is the role played by pseudo-Riemannian geometry in his early thinking. While exploring a fully covariant description of the Klein–Gordon equation, he was led to the revolutionary idea that a variable rest mass was essential. DeWitt later explained that in order to obtain a covariant quantum Hamiltonian, one must supplement the classical Hamiltonian with an additional energy ℏ2Q from which the quantum potential emerges, a potential that Berry has recently shown also arises in classical wave optics. In this paper, we show how these ideas emerge from an essentially geometric structure in which the information normally carried by the wave function is contained within the algebraic description of the geometry itself, within an element of a minimal left ideal. We establish the fundamental importance of conformal symmetry, in which rescaling of the rest mass plays a vital role. Thus, we have the basis for a radically new theory of quantum phenomena based on the process of mass-energy flow.
... The connection between p W, j (x, t ) and v j (x, t ) is approached elsewhere [52][53][54][55][56][57][58][59], while more attention has recently been paid to the meaning of u j (x, t ) [60][61][62][63][64][65][66][67][68]. We distinguish in this paper three types of expectation values: (i) â , with the "hat," for the operatorâ; (ii) a W , with subscript "W," for 033168-3 the weak value a W ; and (iii) a for the value a obtained by postprocessing the weak value. ...
Article
Full-text available
According to both Bohmian and stochastic quantum mechanics, the standard quantum mechanical kinetic energy can be understood as consisting of two hidden-variable components. One component is associated with the current (or Bohmian) velocity, while the other is associated with the osmotic velocity (or quantum potential), and they are identified with the phase and the amplitude, respectively, of the wave function. These two components are experimentally accessible through the real and imaginary parts of the weak value of the momentum postselected in position. In this paper, a kinetic energy equipartition is presented as a signature of quantum thermalization in closed systems. This means that the expectation value of the standard kinetic energy is equally shared between the expectation values of the squares of these two hidden-variable components. Such components cannot be reached from expectation values linked to typical Hermitian operators. To illustrate these concepts, numerical results for the nonequilibrium dynamics of a few-particle harmonic trap under random disorder are presented. Furthermore, the advantages of using the center-of-mass frame of reference for dealing with systems containing many indistinguishable particles are also discussed.
... In other words, while p W,j (x, t) is linked to the hermitian operatorp j , no hermitian operators can be linked to v j (x, t) and u j (x, t). The connection among p W,j (x, t) and v j (x, t) is approached elsewhere [49][50][51][52][53][54][55][56], while more attention has recently been paid to the meaning of u j (x, t) [57][58][59][60][61][62][63][64][65]. We distinguish in the paper three types of expectation values: (i) â with "hat" for the operatorâ; (ii) a W with subindex "W" for the weak value a W ; (iii) a for the value a obtained by post-processing the weak value. ...
Preprint
Full-text available
The Orthodox kinetic energy has, in fact, two hidden-variable components: one linked to the current (or Bohmian) velocity, and another linked to the osmotic velocity (or quantum potential), and which are respectively identified with phase and amplitude of the wavefunction. Inspired by Bohmian and Stochastic quantum mechanics, we address what happens to each of these two velocity components when the Orthodox kinetic energy thermalizes in closed systems, and how the pertinent weak values yield experimental information about them. We show that, after thermalization, the expectation values of both the (squared) current and osmotic velocities approach the same stationary value, that is, each of the Bohmian kinetic and quantum potential energies approaches half of the Orthodox kinetic energy. Such a `kinetic energy equipartition' is a novel signature of quantum thermalization that can empirically be tested in the laboratory, following a well-defined operational protocol as given by the expectation values of (squared) real and imaginary parts of the local-in-position weak value of the momentum, which are respectively related to the current and osmotic velocities. Thus, the kinetic energy equipartion presented here is independent on any ontological status given to these hidden variables, and it could be used as a novel element to characterize quantum thermalization in the laboratory, beyond the traditional use of expectation values linked to Hermitian operators. Numerical results for the nonequilibrium dynamics of a few-particle harmonic trap under random disorder are presented as illustration. And the advantages in using the center-of-mass frame of reference for dealing with systems with many indistinguishable particles are also discussed.
... So the idea that a single quantum object moves along a trajectory ought to be seen as a hypothesis which has not been empirically verified. However, by making use of measurements of weak values it is possible to measure average trajectories (see Flack and Hiley 2018). ...
Chapter
Full-text available
Consciousness and quantum mechanics are two mysteries in our times. A careful and thorough examination of possible connections between them may help unravel these two mysteries. On the one hand, an analysis of the conscious mind and psychophysical connection seems indispensable in understanding quantum mechanics and solving the notorious measurement problem. On the other hand, it seems that in the end quantum mechanics, the most fundamental theory of the physical world, will be relevant to understanding consciousness and even solving the mind-body problem when assuming a naturalist view. This book is the first volume which provides a comprehensive review and thorough analysis of intriguing conjectures about the connection between consciousness and quantum mechanics. Written by leading experts in this research field, this book will be of value to students and researchers working on the foundations of quantum mechanics and philosophy of mind.
... So the idea that a single quantum object moves along a trajectory ought to be seen as a hypothesis which has not been empirically verified. However, by making use of measurements of weak values it is possible to measure average trajectories (see Flack and Hiley 2018). ...
Chapter
Full-text available
Researchers have suggested since the early days of quantum theory that there are strong analogies between quantum phenomena and mental phenomena and these have developed into a vibrant new field of quantum cognition during recent decades. After revisiting some early analogies by Niels Bohr and David Bohm, this paper focuses upon Bohm and Hiley’s ontological interpretation of quantum theory which suggests further analogies between quantum phenomena and biological and psychological phenomena, including the proposal that the human brain operates in some ways like a quantum measuring apparatus. After discussing these analogies I will also consider, from a quantum perspective, Hintikka’s suggestion that Kant’s notion of things in themselves can be better understood by making an analogy between our knowledge-seeking activities and an elaborate measuring apparatus.
Article
Understanding how the interference pattern produced by a quantum particle in Young’s double-slit setup builds up—the “only mystery” of quantum mechanics according to Feynman—is still a matter of discussion and speculation. Recent works have revisited the possibility of acquiring which-way information based on weak measurements. Weak measurements preserve the interference pattern due to their minimally perturbing character while still leading to a final position detection. Here, we investigate a simplified double-slit setup by including weakly coupled pointers. We examine how the information provided by the weak pointers can be interpreted to infer the dynamics within a local picture through “weak trajectories”. We contrast our approach with non-local dynamical accounts, such as the modular momentum approach to weak values and the trajectories defined by the de Broglie–Bohm picture.
Article
The interference pattern produced by a quantum particle in Young's double-slit setup is attributed to the particle's wave function having gone through both slits. In the path integral formulation, this interference involves a superposition of paths, going through either slit, linking the source to the detection point. We show how these paths superpositions can in principle be observed by implementing a series of minimally perturbing weak measurements between the slits and the detection plane. We further propose a simplified protocol in order to observe these “weak trajectories” with single photons.
Article
Full-text available
A method for measuring the real part of the weak (local) value of spin is presented using a variant on the original Stern-Gerlach apparatus. The experiment utilises metastable helium in the 23S1\rm 2^{3}S_{1} state. A full simulation using the impulsive approximation has been carried out and it predicts a displacement of the beam by Δw=1733μm\rm \Delta_{w} = 17 - 33\,\mu m. This is on the limit of our detector resolution and we will discuss ways of increasing Δw\rm \Delta_{w}. The simulation also indicates how we might observe the imaginary part of the weak value.
Article
Full-text available
In a recent paper Mahler {\em et al.} have argued that the experiments of Kocsis {\em et al.} provide experimental evidence for Bohmian mechanics. Unfortunately these experiments used relativistic, zero rest mass photons whereas Bohmian mechanics is based on non-relativistic Schr\"{o}dinger particles having non-zero rest mass. The experimental results can be correctly understood in terms of a different approach based on the electromagnetic field that was already outlined by Bohm in an appendix of the second of his 1952 papers. A subsequent development of this approach by Bohm, Hiley, Kaloyerou and Holland, show in detail how this theory accounts for the experimental results. We are led to the conclusion that the experiments have constructed mean momentum flow lines by measuring the real part of the weak Poynting vector. These results support and clarify the analysis of Bliokh {\em et al}. The experimental flow lines can be constructed independently of the number of photons in the beam leading to the conclusion that flow lines cannot be interpreted as `photon trajectories'. We discuss exactly how the notion of a photon arises in the field approach.
Article
Full-text available
We describe the development of an experiment to measure the weak value of the transverse momentum operator (local momentum [1]) of cold atoms passing through a matter- wave interferometer. The results will be used to reconstruct the atom's average trajectories. We describe our progress towards this goal using laser cooled argon atoms.
Article
Full-text available
Quantum mechanics is derived from the principle that the universe contain as much variety as possible, in the sense of maximizing the distinctiveness of each subsystem. The quantum state of a microscopic system is defined to correspond to an ensemble of subsystems of the universe with identical constituents and similar preparations and environments. A new kind of interaction is posited amongst such similar subsystems which acts to increase their distinctiveness, by extremizing the variety. In the limit of large numbers of similar subsystems this interaction is shown to give rise to Bohm's quantum potential. As a result the probability distribution for the ensemble is governed by the Schroedinger equation. The measurement problem is naturally and simply solved. Microscopic systems appear statistical because they are members of large ensembles of similar systems which interact non-locally. Macroscopic systems are unique, and are not members of any ensembles of similar systems. Consequently their collective coordinates may evolve deterministically. This proposal could be tested by constructing quantum devices from entangled states of a modest number of quits which, by its combinatorial complexity, can be expected to have no natural copies.
Book
Richard Feynman's never previously published doctoral thesis formed the heart of much of his brilliant and profound work in theoretical physics. Entitled “The Principle of Least Action in Quantum Mechanics," its original motive was to quantize the classical action-at-a-distance electrodynamics. Because that theory adopted an overall space–time viewpoint, the classical Hamiltonian approach used in the conventional formulations of quantum theory could not be used, so Feynman turned to the Lagrangian function and the principle of least action as his points of departure. The result was the path integral approach, which satisfied — and transcended — its original motivation, and has enjoyed great success in renormalized quantum field theory, including the derivation of the ubiquitous Feynman diagrams for elementary particles. Path integrals have many other applications, including atomic, molecular, and nuclear scattering, statistical mechanics, quantum liquids and solids, Brownian motion, and noise theory. It also sheds new light on fundamental issues like the interpretation of quantum theory because of its new overall space–time viewpoint. The present volume includes Feynman's Princeton thesis, the related review article “Space–Time Approach to Non-Relativistic Quantum Mechanics” [Reviews of Modern Physics 20 (1948), 367–387], Paul Dirac's seminal paper “The Lagrangian in Quantum Mechanics'' [Physikalische Zeitschrift der Sowjetunion, Band 3, Heft 1 (1933)], and an introduction by Laurie M Brown. © 2005 by World Scientific Publishing Co. Pte. Ltd. All rights reserved.
Chapter
Every modern century has its paragons, its culture-heroes, who fill the place once taken by saints and messiahs. For the first half of the twentieth century, outside the totalitarian states, the most celebrated paragons have been Gandhi, Schweitzer, and Einstein; and of these the most influential as a thinker was clearly Albert Einstein.
Article
Exact numerical calculations are made for the scattering of quantum mechanical particles from a square two‐dimensional potential barrier. This treatment is an exact analog of both frustrated total reflection of perpendicularly polarized light and the longitudinal Goos‐Hänchen shift. Quantum mechanical streamlines (which are analogous to either classical mechanical trajectories or optical rays) are plotted. These streamlines are smooth, continuous, and have continuous first derivatives even through the classically forbidden region. The streamlines form quantized vortices surrounding each of the nodal points (which result from interference between the incident and reflected waves). Similar vortices occur in reactive collisions of H + H2 (McCullough and Wyatt; Kuppermann, Adams, and Truhlar) and undoubtedly play an important role in molecular collisiondynamics. The theory for these vortices is given in a companion paper. A comparison is given between our numerical calculations and the stationary phase approximation (SPA). Although our incident wave packet has a half‐width of more than one de Broglie wavelength in contrast to the SPA which replaces the diffuse beams by rays with delta function profiles, the agreement was surprisingly good for both the Goos‐Hänchen shifts and for the reflection coefficients. However, we found that the Goos‐Hänchen shift for the transmitted beam is significantly smaller than for the reflected beam, although in the stationary phase approximation these two shifts are equal. Furthermore, we found that the scattering from the potential barrier has very little effect on the shape of the wave packets. The power series expansion of the incident Debye‐Picht wave packet ψ I has an extremely small radius of convergence, whereas the power series for ψ I *ψ I has a radius of convergence of more than two de Broglie wavelengths. The imaginary velocity v i is introduced into the Madelung‐Landau‐London hydrodynamical formulation of quantum mechanics. The corresponding imaginary streamlines will be considered in a forthcoming paper. The time‐independent Schrödinger equation for real wavefunctions is reduced to solving the nonlinear first order partial differential equation: ℏ▿·v i = 2(E − V) + v i 2 . Here, v i is irrotational. This equation may lead to interesting new methods of solving the Schrödinger equation. It does lead to a generalization of the Prager‐Hirschfelder perturbation scheme which invokes an electrostatic analogy. The philosophical implications of the hydrodynamical formulation of quantum mechanics are discussed. Penetration of a one‐dimensional square potential barrier is used to demonstrate exactly how tunneling occurs by particles ``riding over the barrier'' à la Bohm. Cases are cited where quantum and classical mechanical motions are identical.
Article
Scitation is the online home of leading journals and conference proceedings from AIP Publishing and AIP Member Societies