You are on page 1of 12

MATTER AS STRINGS OF OSCILLATING NODES IN A

SPATIAL CONTINUUM.
THE SCHRÖDINGER WAVE EQUATION.

BJØRN URSIN KARLSEN

To Elizabeth

Abstract. This paper is one of a number of papers backing up a mechanical


model of light, matter, and the basic forces of nature outlined in my booklet
”The Great Puzzle”. It is a followup of another paper, ”Standing waves be-
tween singularities in an elastic continuum of infinite extension”, where I shall
propose how the presented ideas in that paper can be sewn together to form a
model of photons and material particles. As a precondition there has got to be
an enormous concentration of potential and kinetic energy released in a very
limited area of the spatial continuum, where initially various constellations of
disturbances can occur. Finally I will show how the outlined model can be
brought into accordance with The Schrödinger wave equation.

1. The Point-like Nature of Matter

Space is taken to be an elastic continuum of infinite extension in which there may be


singularities in the form of oscillating nodes, which oscillate between compression
and inflation along an infinitely long chain of nodes [3]. Compression and inflation
are located to the area around each node, but for convenience I will speak of the
spatial mass as if it was going into or out of the nodes themselves. Normally the
spatial mass that is ejected from one node goes as a compression into its adjacent
nodes such that the net displacement is zero, but there may be a single node in
the chain that is inflated to a higher degree than what is taken up by the other
nodes in the chain. A node of this kind is – also for convenience – said to be a
bubble. The bubble is of course not stable, but will immediately be filled up with
an inward stream of spatial mass from say the next node in the chain. In this way
the bubble is moving in a stepwise manner from node to node along the chain.
When the domain around the node representing the bubble is exactly filled up, the
inward stream of spatial mass has its maximum speed and will continue for a while
until compression stops the movement and reverses the stream. Therefore nodes
that are left behind have got to oscillate with decreasing amplitudes as the bubble
recedes. On the other hand the next node in the chain cannot suddenly be inflated
to the exited level, but has got to start oscillating with increasing amplitude in
good time before the bubble arrives. Hence the arrival of the exited node has got
to be preceded by a pilot chain of nodes that oscillate with increasing strength as
the inflated node approaches. The entire chain of nodes may extend far out in

Date: 15-09-10.
1
2 BJØRN URSIN KARLSEN

both directions. This picture conforms well with the findings in the cited paper [3]
where I found that oscillating nodes only are possible in an infinitely long chain
of oscillating nodes. Here the assumption has got to be modified to incorporate
very long, steadily increasing strings of nodes (which I will dub the preamble) that
culminate in a strongly inflated node (dubbed the bubble) whereupon the trailing
nodes (which I will dub the postamble) gradually decrease towards zero. In this
picture all matter has got to be viewed as composed of long, almost infinitely long
strings of oscillating nodes with a strongly excited node at its core.

A delta function of the type sin(ωt)/t seems to cover this situation well for the
oscillation of each node along the chain of nodes when t varies between − ∞ and
+ ∞. This is a calculated guess and I cannot prove it, but I will try to see what
implications it will have. The limit of the function as t approaches zero is given by

sin(ωt)
lim = ω,
t→0 t

where ω is the angular velocity of the oscillation. (See the leftmost blue graph in
Figure 1.) Let us give each node along the chain (represented by a black dot in the
graph) an integer number n such that the next node in the chain has the number
n + 1, and let the displacement from the domain of node #n at the time t be given
by:

sin[ω(t − nπ/ω)]
Dn (n, t) = k
t − nπ/ω

where k is some constant. First we notice that according to the above assumption,
when rotation is disregarded, the displacement from node #0 at the time t = 0 is
given by:

sin(ωt)
D0 (0, 0) = k lim = kω
t→0 t

while the displacements from all the other nodes in the chain are zero, i.e. the
displacement from the entire chain of nodes is kω. Half a period later node #1 is
fully inflated while the displacements from all the other nodes again are zero. In
fact this is the case for all times t = nπ/ω

sin[ω(t − nπ/ω)]
(1.1) Dn (n, nπ/ω) = lim k = kω.
t→nπ/ω t − nπ/ω

In the intermediate phases, the inflation gradually decreases while the new displace-
ment builds up in the next node. In this way the inflated node moves from node to
node with a frequency like 2f while all the other nodes along the chain oscillate be-
tween fully inflated state and fully compressed state with the same frequency. The
blue columns in Figure 1 show the the displacement from each node in successive
steps in the course of half a period.

If this shall be a viable model, however, then at least the net displacement has got
to be the same during the intermediate phases, so we write down the sum of the
1 3

Displacement
Flow

Figure 1. Model of a photon in 5 successive steps separated by one


eighth of a period. The blue columns represents the displacement
and the red columns the flow from each node in the string.

displacements from all the nodes in the chain at any time


n
X
D(t) = lim Dm (t)
n→∞
m=−n
n
X sin(ωt − mπ)
= lim kω , ωt − πm 6= 0.
n→∞
m=−n
ωt − mπ

First we notice that sin(ωt−mπ) = sin(ωt)·(−1)m for any m ∈ {· · ·−2, −1, 0, 1, 2, · · · }.


Hence
n
X (−1)m
(1.2) D(t) = kω sin(ωt) lim .
n→∞
m=−n
ωt − mπ

From http://functions.wolfram.com/01.10.09.0001.01 we have


n
X (−1)m z
csc(z) == lim /; ∈
/ Z,
n→∞
m=−n
z − πm π

where csc z = 1/ sin z. Hence


1
(1.3) D(t) = kω sin(ωt) = kω.
sin(ωt)
We can conclude that the displacement from the whole chain of oscillating nodes
is the same at any time during an oscillating cycle. A chain of nodes meeting
4 BJØRN URSIN KARLSEN

these requirements will represent a net displacement, which we could call a bubble,
moving with some speed v through space.

Since the nodes can be organized along the x-axis of a coordinate system at a
distance of λ/2 from each other, then the displacement from each node along the
x-axis can be represented by applying the Dirac’s delta function1
Z ∞
sin[2π(x − c̃t)/λ]
D(n, t) = k δ(x − n · λ/2) dx.
−∞ 2π(x − c̃t)/λ

In another paper, ”Confined energy in an expanding elastic continuum compared


with gravity”[4], I showed that confined wave energy in the spatial continuum will
displace an amount of spatial mass given by the relation
E
D= .
3(λ + 2µ)
If we assume that this is a general principle also the other way around such that to
a given displacement there is a net amount of energy given by
(1.4) E = 3D(λ + 2µ) = 3k(λ + 2µ)ω = ~ω,
where h and ~ is a couple of new constants given by ~ = h/2π = 3k(λ + 2µ). Hence
~ h
(1.5) k= = .
3(λ + 2µ) 6π(λ + 2µ)

The angular velocity ω in Equation (1.4) can be replaced by the frequency ν by the
identity ω = 2πν which yields
(1.6) E =hν,

and
hν 2π~ν
(1.7) D= = .
3(λ + 2µ) 3(λ + 2µ)
This establish the remarkable condition that the energy in any chain of oscillating
nodes is proportional to the frequency alone.

We now turn back to the amplitudes of the oscillation by each node, which by the
way are situated at a distance of half a period from each other. Between maximum
compression and inflation there has got to be a phase of inward and outward flow
of spatial mass to and from the nodes. The rate of change of the displacement from
each node will naturally be a measure of the flow, so by taking the time derivative
of the displacement we will get an expression for the flow, i.e. the flow through
a surface around the node where div u = 0. It turns out that the flow may be

1Dirac’s delta function has the fundamental property that


Z ∞
f (x)δ(x − a) = f (a).
−∞

(Eric W. Weisstein. ”Delta Function.” From MathWorld–A Wolfram Web Resource.


http://mathworld.wolfram.com/DeltaFunction.html).
1 5

represented by a function of the type ωt [cos(ωt) − sin(ωt)


ωt ], or for the entire chain of
nodes the flow, F (n, t), from node #n at the time t is given by
hω h sin(ωt − nπ) i
Fn (n, t) = nπ cos(ωt − nπ) − .
t− ω ωt − nπ

We can draw a graph of the displacement from each node along the y-axis (blue
column) and the flow out of each node along the z-axis (red column). Just think of
the nodes as not moving entities and the graph as moving forward along the x-axis
just showing the displacement and flow from each node as a function of time. The
first graph in Figure 1 shows a moment when all the displacement is from just one
of the nodes, while the next tree graphs show the situation 1/8, 2/8, and 3/8 of a
period later, and the last graph shows the displacement when the next node is fully
inflated 4/8 of a period later. We notice that the excitation reaches its maximum
value two times in the course of one cycle of oscillation.

This idealized model can possibly explain some of the properties of material par-
ticles if the path is curled up in some way, and even the photon if the path is a
stretched out line, for instance the quantum and point-like property of the photon,
but it does not say anything about the electromagnetic and spin properties and why
the photon moves with the speed of light. On the contrary; an irrotational stand-
ing wave can be composed of two oppositely moving progressive waves that moves
with about the double of the speed of solenoidal waves. It is therefore necessary to
include some sorts of coupling to solenoidal oscillations into the model.

2. Spin and polarization

In the paper ”Standing waves between singularities in an elastic continuum of in-


finite extension” [3], I pointed out the possibility that there might be coupled
oscillations between irrotational and solenoidal oscillations in the same string of
nodes. The coupling has got to take place during the inward stream of spatial
mass towards the singularity when energy is transferred from the radial movement
to a circular movement, and conversely from circulation to an outwards directed
movement when the node is inflated. Hence the rotational component will follow a
similar delta function as the displacement, i.e. sin(ωt)/t, and reach its maximum
when the bubble is inflated. It is important to realize that the coupling can only
take place in the immediate neighborhood of the singularity where the displacement
gradients are so great that the linear theory of elasticity is not applicable and we
have got to resort to the Navier/Stokes equation2. Everywhere else in space out-
side this tiny area the two waveforms move literally independent of each other, and
the displacement components originated from each of the nodes can be superposed
linearly.

2With great deformation gradients the velocity of spatial mass points, v, is different from the
partial derivative of u with respect on time as it is taken to be in the Navier/Cauchy equation.
In fact v = ∂u/∂t + v(∇v) [1, Chapter 4] (i.e. ’The velocity of a mass element in an elastic
continuum is like the partial time derivative of u plus the change of v in the direction of v times
the norm of v’) which is accounted for in the Navier/Stokes equation
6 BJØRN URSIN KARLSEN

If a point-like mass, m, is moving with velocity v along the circumference of a circle


with radius r, then according to classical physics it has an angular momentum, also
called spin, given by L = mrv. The spin axis is normal to the circular plane and
goes through the center of the circle with a direction according to the righthand
rule. The system’s kinetic energy is of course 1/2mv 2 . Accordingly a rotating body
has an intrinsic spin given by L = Iω and a kinetic energy E = 1/2Iω 2 , where I is
the momentum of inertia and ω the angular velocity.

In order to get an impression of the relation between energy an spin in a chain


of oscillating nodes, let us think of the chain of nodes as pulsating beads on an
elastic string where each bead rotates in the opposite direction of its two adjacent
beads. Further let the pulsating and rotational energy be of the same magnitude,
which amounts to the same as saying that the rotational energy is half of the total
energy in the entire chain of nodes. At the exact moment when the rotation of the
inflated bead has its maximum value, all the other beads are coming to a complete
standstill. Hence the rotational energy is half of the total energy, i.e. E = 1/2 Iω 2 ,
and the net spin of the system is given by L = Iω. In the above section we found
that the total energy of the system is given by E[tot] = ~ω. By assuming that ω is
the same in both cases we can combine these properties and obtain
1 2
2 Iω = 12 ~ω,
Iω = ~,

L = ~.

This simplified thought experiment indicates that the spin of the system is inde-
pendent of the energy and therefore becomes a natural constant.

In this thought experiment the spin axes were taken to be directed along the string
of nodes, but that need not be the case. A more probable assumption is that the
axes are directed in the right and left direction along the string. Say that that is
the case and that the bubble is moving forwards with the speed of c. Then the
right and left rotation in the preamble may well generate a transversal progressive
wave. The same mechanism will also take place in the postamble, but there is a
problem: The plane polarized wave in the postamble is half a period out of phase
with the corresponding wave in the preamble.

Let us assume that a progressive transversal wave is generated in the preamble.


When the spin around the bubble-node itself builds up, however, the buildup of
rotation is delayed by a quarter of a period and it gets increasingly out of phase
with the field in the associated wave. The rotational axes will therefore get a
precession towards the forward or backward direction of the string. The precession
will continue during the build-down phase until the rotation axis has made a 180
degree pirouette while the node enters the postamble section of the string. With this
correction in place, the phase of the waves generated in the preamble and postamble
will be in tune with each other, and the progressive wave that follows the bubble
as it moves along, will - at least near the string - resemble a free transversal wave
moving along with the bubble. A lot of wave-trains with the same frequencies and
polarization directions may well be thought to move in step, and the superposition
1 7

of the fields will approximate a plane polarized wave like a laser beam. This should
be a plausible explanation of the electromagnetic property of the photon model,
but it also indicate why a photon has spin: As the rotation around the bubble
reaches its maximum value, it will point in the forward or backward direction of
the movement vector just as the spin of a real photon.

So let us put this model of a material particle to a test to see how it fits into a
realistic picture of material particles.

3. The Schrödinger wave equation

As suggested above a material particle is taken to be a bubble moving with speed c


along a curled up sting of oscillating nodes all oscillating with the same frequency,
ν = E/h, and with rotational components around axes that all point along the
same spatial direction. The spin in the forward or backward direction does not
contribute to creating the solenoidal wave and can be neglected. The whole system
is advancing with a group velocity v ¿ c. The field around each node is composed
of concentric outwards and inwards progressive waves that interfere with each other
forming standing waves, and the complete pattern is made up of a superposition of
the waves from each node in the wave packet. The pressure fields in longitudinal
waves are pure scalar fields and can readily be added, and so can the displacement
vector components in different solenoidal fields, for instance by the three component
in a cartesian representation. In the latter case, however, let all the rotation axes be
in the z direction. Then all displacement components are in the x, y plane and the
number of components to add reduces to two. This makes it possible to represent
the solenoidal fields as a superposition of the real and imaginary parts of the fields
around each node in the vicinity, which adds up to the complex field ψ(r, t). The
squared value of ψ, i.e. ψ 2 = ψ ∗ ψ, is a real property, which is like the square of the
displacement vector, u, provided that a suitable choice of constants are chosen.

The displacements outside the singularities can be described by the Navier-Cauchy


equation [2] without any external forces. It is given by
∂2u
(λs + 2µs ) grad div u − µs curl curl u = ρs
∂t2
which by the identity curl curl A = grad div A − ∇2 A yields
∂2u
(λs + µs ) grad div u + µs ∇2 u = ρs .
∂t2
The solenoidal field is described by setting div u = 0 and we acquire
ρs ∂ 2 u
∇2 u = .
µs ∂t2
The divergence-free displacement field around an oscillation node can be repre-
sented by a complex number ψ(r, t) as stated above, and by inserting the wave
speed c2 = µs /ρs we obtain the wave equation
1 ∂ 2 ψ(r, t)
(3.1) ∇2 ψ(r, t) = ,
c2 ∂t2
8 BJØRN URSIN KARLSEN

which can be solved by the product method. Notice that I only seek solutions that
is a superposition of standing waves around a plethora of oscillating singularities
with rotating axes all pointing in the same direction as described above.

Let ψ(r, t) = g(r) · f (t). Then


1 ∂2
∇2 [g(r) · f (t)] = [g(r) · f (t)]
c2 ∂t2
g ∂2f
f ∇2 g = 2 2
c ∂t
∇2 g 1 ∂2f
= 2
g c f ∂t2
The left and right side of this equation are only functions of r and t respectively,
and can therefore be set to the same constant, say −q 2 , and we obtain the two
equations
∇2 g = −q 2 g,
∂2f
= −c2 q 2 f,
∂t2
which have soultions of the form [5, page 16]
g ∝ exp(−i q · r), q 2 ≡ |q|2 ,
f = exp(−ickt),
from which we obtain
ψ = exp(−i q · r) · exp(−icqt),
£ ¤
ψ = exp − i(q · r + cqt) .
This is an expression for a plane wave moving in the direction of q and oscillating
with a frequency ν = cq/2π. As presumed above, the energy of the system is a
function of the frequency, and we obtain
E E
cq = 2πν = 2π = , ~ = h/2π,
h ~
E
q= .
~c
The displacement function, ψ, takes the form
h −iE i
(3.2) ψ = exp − i(q · r) + t .
~
I will use a double-line font to distinguish the internal energy in a wave packet from
the classical kinetic and potential energies of the particle.

With these properties inserted the wave equation


1 ∂2ψ
(3.3) ∇2 ψ = ,
c2 ∂t2
can be developed further into
(−i)2 E2
(3.4) ∇2 ψ = ψ,
~2 c2
1 9

which is the equation for the internal (not observable) oscillation of a wave packet
with energy E.

Now let us consider the particle as seen by an observer moving along with the
speed v ¿ c, and let the oscillation in the fixed coordinate system be represented
by primed coordinates
ψ(x0 , y 0 , z 0 , t0 ) = g(x0 , y 0 , z 0 )f (t0 ),
£ E0 ¤
= g(x0 , y 0 , z 0 ) exp(−i t0 )
~
and
1 ∂2f
f ∇2 g =
g .
c2 ∂t2
Next transform the movements over to the unprimed coordinate coordinate system
of the observer, who is moving along with the speed −v. From the observers
standpoint the particle seems to be moving in the positive direction. It is only
in a Lorenz coordinate system that the divergence-free properties are invariant by
transformations3, so we apply the Lorenz transformations
x − vt
x0 = p ,
1 − v 2 /c2
y 0 = y,
z 0 = z,
t − vx/c2
t0 = p ,
1 − v 2 /c2
and consider the particle at the position x = 0 and time t = 0.

From the observer’s point of view the oscillation takes the form
³ −iE0 ´
f (x, y, z, t) = exp p ·t ,
~ 1 − v 2 /c2
h −i i
f (x, y, z, t) ≈ exp (E0 + 12 E0 v 2 /c2 ) · t
~
If we define m = E0 /c2 then we recognize the term 12 E0 v 2 /c2 = 21 mv 2 as the kinetic
energy, T , of the wave packet and we obtain
h −i i
f (x, y, z, t) = exp (E0 + T ) · t .
~
E0 is supposed to be the internal energy before the particle is set into motion, and
m is defined as the rest mass of the particle.

Accordingly we can transform g into


£ −i ¤
g(x, y, z, t) = exp (q0 + k) ,
~
where q0 is representing the particle’s internal and k its progressive component of
motion. Hence k is pointing in the direction of the particle’s motion while q0 is
3Discussed in my paper ”A comparison between the Linear Theory of Elasticity and the Clas-
sical Theory of Electromagnetism” [2].
10 BJØRN URSIN KARLSEN

pointing in the direction of the curled-up path along the string of nodes. For each
positive q0 there is an equally great negative value somewhere else in space so the
mean value of q0 · k is zero. This amounts to the same as stating that q0 and k
are orthogonal to each other.

When we split up the energy, E, into a stationary part E0 and a kinetic part T ,
and accordingly q into q0 + k, the wave function (3.2) takes the form

h −iE0 −iT i
ψ = exp − i(q0 · r) − i(k · r) + t+ t ,
~ ~
h iE0 i h iT i
ψ = exp − i(q0 · r) − t · exp − i(k · r) − t
~ ~

Let ψ = ψ1 ψ2 . Then generally

∇2 ψ = ψ1 ∇2 ψ2 + ψ2 ∇2 ψ1 + 2∇ψ1 ∇ψ2 .
∂2ψ
= ψ1 ψ¨2 + 2ψ˙1 ψ˙2 + ψ2 ψ¨1 .
∂t2

With
h iE0 i
ψ1 = exp − i(q0 · r) − t ,
~
h iT i
ψ2 = exp − i(k · r) − t ,
~

and q0 orthogonal to k we have

∇ψ1 ∇ψ2 = (−iq0 ψ1 ) · (−ikψ2 ),


= −ψ1 ψ2 q0 · k
= 0.

With these terms inserted into the wave equation (3.1) we obtain

1
ψ1 ∇2 ψ2 + ψ2 ∇2 ψ1 = (ψ1 ψ¨2 + 2ψ˙1 ψ˙2 + ψ2 ψ¨1 ).
c2

By Equation (3.3) underlined terms cansel out and the above equation reduces to

1
ψ1 ∇ 2 ψ2 = (ψ1 ψ¨2 + 2ψ˙1 ψ˙2 ).
c2
−iE0 −iT (−i)2 T 2
c2 ψ1 ∇2 ψ2 = 2 ψ1 ψ2 + ψ1 ψ2 ,
~ ~ ~2
E0 T T
c2 ∇2 ψ2 = −2 2 (1 + )ψ2 .
~ 2E0
1 11

Since T ¿ E0 , the last term is negligible and we obtain


~2 c2 2
− ∇ ψ2 = T ψ2 , (E0 = mc2 ),
2E0
~2 2
− ∇ ψ2 = T ψ2 ,
2m
~2 2 h iT i
− ∇ ψ2 = T exp − i(k · r) − t ,
2m ~
~2 2 ~ ∂ h iT i
− ∇ ψ2 = T exp − i(k · r) − t ,
2m −iT ∂t ~
~2 2 ∂ψ2
− ∇ ψ2 = i~ ,
2m ∂t
We can now replace ψ2 -old with ψ-new and acquire

~2 2 ∂ψ
∇ ψ = −i~
2m ∂t
This is the free Schrödinger wave equation for a particle moving with constant
speed.

If the particle is moving in a potential field, then E0 is varying with position, but it
is possible to introduce a small energy component V such that E0 − V is constant.
To keep the total energy constant we have got to add the same component to T . We
obtain two new properties: V , which can be interpreted as the potential in which
the particle is moving, and E = V + T , which is the total kinetic and potential
energy of the particle. The wave function takes the form
h i(E0 − V ) i h iE i
ψ = exp − i(q0 · r) − t · exp − i(k · r) − t .
~ ~
If we neglect terms with E 2 and V 2 we can proceed to
(−i)2 E02 −iE0 iV −iE0 −iE
c2 ψ1 ∇2 ψ2 + c2 ψ2 ∇2 ψ1 = ψ2 2
ψ1 + 2 ψ1 ψ2 + 2 ψ1 ψ2 .
~ ~ ~ ~ ~
By Equation (3.4) the underlined terms cansel out, and we obtain
~2 c2 2
− ∇ ψ2 + V ψ2 = Eψ2 ,
2E0
or by the same procedure as above we obtain
~2 c2 2 ∂ψ2
− ∇ ψ2 + V ψ2 = i~ .
2E0 ∂t
With m = E0 /c2 and ψ2 -old replaced by ψ-new this equation takes the form

~2 2 ∂ψ
− ∇ ψ + V ψ = i~
2m ∂t

which is the the Schrödinger wave equation [5, Eq. 1.12], and like the Schrödinger
equation it is only valid for small velocities when v ¿ c. It tells us about two
12 BJØRN URSIN KARLSEN

basic properties of the material particle: Its frequency and wave number. From
these properties one can find its energy and momentum, but not the position of
a free particle. Only if the particle is restricted to be inside a given volume will
it be possible to tell something about the whereabouts of the naked particle: The
property ψ is a measure of displacement, and its squared value becomes a measure
of the intensity of the deformation field. An elementary material particle in this
picture is an evacuated bubble that may be found anywhere in the wave packet
with an increasing probability along with the intensity of the field. Since it has
got to be somewhere in the restricted volume – and it surely has got to visit every
single node along the chain – the integral of ψ 2 dv all over the volume element can
be normalized to unity, and ψ 2 can thus be taken to mean the probability density
of finding the bubble in a given volume element inside the considered volume.

A considerate amount of interactions between material particles that meet the re-
quirements above, can be solved by applying this Schrödinger-like wave equation.
For example can different particles exchange momentum by superposition of the
fields around the individual particles that in turn leads to scattering by the inter-
action. Likewise will the progressive waves around particles, which are restricted
to move in closed volumes or orbits, tend to form standing waves that only can
occur at certain energy levels. By a double slit experiment the generated wave in
the surroundings of the string of nodes passes partly through both slits causing an
interference with varying field intensity of ψ 2 while the bubble itself has got to pass
through one of the slits tending to follow paths with stronger probable densities be-
yond the slits. A complete description of all possible interactions, however, is only
possible when spin is taken into account, and the equations are developed without
the restriction that v ¿ c.

References
1. Yavuz Başar and Dieter Weichert, Nonlinear continuum mechanics of solids, Springer, 1999.
2. Bjørn Ursin Karlsen, A comparison between the Linear Theory of Elasticity and the Classical
Theory of Electromagnetism, http://home.online.no/˜ukarlsen.
3. , Standing waves between singularities in an elastic continuum of infinite extension,
http://home.online.no/˜ukarlsen.
4. , Confined energy in an expanding elastic continuum compared with gravity,
http://home.online.no/˜ukarlsen, 2002.
5. Paul C.W. Davis and David S. Betts, Quantum mechanics, 2 ed., Stanley Thornes (Publishers)
Ltd, 1999.

E-mail address: ukarlsen@online.no

You might also like