You are on page 1of 216

Coverpicture

The giant nebula NGC 3603


NASA’s Hubble Space Telescope
The Great Puzzle
A test to see if matter can fit into a
mechanical model of nature

Bjørn Ursin Karlsen

June 1, 2009
To Elizabeth
3

Preface

This book is an attempt to figure out what space and matter


has got to be like in order that all known phenomena shall fit
in. It turns out that an elastic spatial continuum with certain
simple properties has the capacity to explain all electromagnetic
properties in empty space as well as the nature of Special Rel-
ativity. Standing waves between singularities along a string of
nodes in the spatial continuum may serve as a foundation to
see how luminous energy and matter is built up, basically as
disturbance energy in the elastic continuum. Material particles
and photons will manifest themselves as evacuated ”bubbles”
moving from node to node along the string with the propagat-
ing speed of transversal waves. Confined disturbance energy
will displace some of the elastic continuum from where it is lo-
cated, and the pressure gradient that it creates will act as a
wall around the energy concentration and constitute the strong
forces needed to keep it at bay. A huge implosion, similar to
what we see when a gas bubble implodes to emit light in an ex-
periment with sonoluminecence, may have compressed space to
an extent where the symmetry was broken and brought space
itself to the boiling point when all matter was created in the
Big Bang. From then on space would have been expanding,
and all confined energy in such an expanding space, will create
around itself an extremely weak pressure gradient that can be
compared with the gravitational potential. The varying density
thus created around heavy bodies will influence the speed of lon-
gitudinal and transversal waves in the continuum causing rays
4

of radiation to bend towards such bodies. Since material parti-


cles are bubbles that always move along strings with the speed
of light along some curled up paths in the spatial continuum,
they will always be drawn towards a gravitating body. It turns
out that any measuring system that we possibly can manage
to define, will render the speed of light unaffected of position.
Therefore it will be necessary to calculate wave movement in
space with time and length units that varies throughout space,
which will lead to defining space as being curved. Hence matter
and light will move along geodesics in space-time according to
General Relativity.
Contents

1 Basic Ideas 9
1.1 The nature of space . . . . . . . . . . . . . . . . 10
1.2 Electromagnetism compared with elastodynamics 13
1.3 Maxwell’s and The Navier-Cauchy Equations . . 19
1.4 Confined waves in the Spatial continuum . . . . . 24
1.5 The Big Bang . . . . . . . . . . . . . . . . . . . . 33
1.6 Waves in the spatial continuum . . . . . . . . . . 41
1.7 The photon . . . . . . . . . . . . . . . . . . . . . 49
1.8 The electron . . . . . . . . . . . . . . . . . . . . . 52
1.9 Systems of many particles . . . . . . . . . . . . . 58
1.10 Atomic nuclei and the strong forces . . . . . . . . 62
1.11 Gravitation . . . . . . . . . . . . . . . . . . . . . 69
6 CONTENTS

1.12 The General Theory of Relativity . . . . . . . . . 72

2 The Linear Theory of Elasticity 79


2.1 Displacement fields . . . . . . . . . . . . . . . . . 80
2.2 System of forces . . . . . . . . . . . . . . . . . . 82
2.3 The stress-strain relation . . . . . . . . . . . . . . 85
2.4 The Navier-Cauchy equation . . . . . . . . . . . 86
2.5 Field energy and energy transport . . . . . . . . 89
2.6 Solenoidal deformations and Electrodynamics . . 95
2.7 Reformulation of the Navier-Cauchy’s Equation . 96
2.8 The stress energy tensor . . . . . . . . . . . . . . 99
2.9 The vector potential in the elastic continuum . . 106
2.10 Energy flow and momentum . . . . . . . . . . . . 109
2.11 Summing up . . . . . . . . . . . . . . . . . . . . 113

3 Standing Waves between Singularities 115


3.1 The Navier-Cauchy Equation . . . . . . . . . . . 116
3.2 Scalar longitudinal waves in the spatial continuum 120
3.3 Irrotational standing waves . . . . . . . . . . . . 122
3.4 Solenoidal standing waves . . . . . . . . . . . . . 140
3.5 Standing waves between singularities . . . . . . . 143
CONTENTS 7

3.6 Chains of irrotational and solenoidal oscillating


nodes . . . . . . . . . . . . . . . . . . . . . . . . 149
3.7 The Navier-Stokes equation and Coupled oscilla-
tions . . . . . . . . . . . . . . . . . . . . . . . . . 151

4 Spatial Continuum Mechanics 159


4.1 Strain . . . . . . . . . . . . . . . . . . . . . . . . 160
4.2 Stress-strain relations . . . . . . . . . . . . . . . 167
4.3 Velocity and acceleration . . . . . . . . . . . . . . 174
4.4 The strain-stress relation for small deformations 175
4.5 Inertia and the speed of waves as a function of
the compression of space . . . . . . . . . . . . . . 179

5 Confined Energy and Gravity 187


5.1 Confined energy in the spatial continuum . . . . 188
5.2 The expanding spatial continuum . . . . . . . . . 192
5.3 Confined energy in expanding space . . . . . . . 195
5.4 Movements of energy packets in a space with
varying wave speeds . . . . . . . . . . . . . . . . 200
5.5 Newton’s Gravitational law . . . . . . . . . . . . 204
5.6 A numerical comparison . . . . . . . . . . . . . . 207
8 CONTENTS
Chapter 1

Basic Ideas

In the nineteenth century the scientific community of that time


had discovered that light was waves and, since it was thought
that waves have to travel through a medium of some kind, it
seemed obvious that space was something that could mediate
those waves. Since it also is obvious that the earth, as well as
all other bodies, has got to move around in this space, it would
be only a matter of creating the right apparatus to measure the
speed with which the earth actually travels through space. Such
an apparatus was created by Michelson and Moreley in the
1880th, and the world kept its breath when the apparatus was
set to work. Even the traffic in Chicago was halted - at least
so I have been told - to make the conditions perfect for the
measurements. It came as a complete shock to all who saw the
10 Basic Ideas

significance of the experiment, that no such movements through


space could be detected. Since then the experiment has been
carried out so many times and in so many ways that it would be
foolish to believe that the negative result should be caused by
some errors in the setup, so either is there no ether out there,
or else there is an underlying principle that it cannot be felt by
our apparatuses. In either case an obvious conclusion lay near
at hand: Why bother? If it is impossible to detect an ether,
then it cannot disturb our observations, so just let us forget all
about it. On this grounds the twentieth century’s scientists have
created a mathematical tool to describe almost all observable
physical phenomenon to an almost unbelievable accuracy, but
it has come at a price, namely that the human mind has had
serious problems in keeping pace. So, is there any future in
supposing that space has mechanical properties, and above all,
is there any point in searching for such a space even if it should
exist? It is questions like these I would like to find an answer
to in this booklet, and - as I see it - the only way to sort it out,
is to go ahead and try it out. I am quite prepared to stumble
along and almost certainly make mistakes, and I have no belief
that I will find all the answers, but as long as I honestly can’t
see that the road is blocked, I’ll keep on trying.

1.1 The nature of space

I will take up the thread from the nineteenth century that space
has got to have mechanical properties, but it is almost cer-
The nature of space 11

tain that it cannot be like earthly materials like steel or water.


These materials consist of atoms and molecules bound together
by forces that on an atomic scale can be linked back to elec-
tromagnetic properties. What I have in mind is a true spatial
continuum with some elastic properties. As a first approach it
turns out that The Linear Theory of Elasticity is general enough
to describe deformations and movements in such a continuum,
and in a later chapter I will use this theory to explain how all
the properties in the classical theory of electrodynamics can be
described in such a space if we make one distinct assumption
about the nature of electric charges. Moreover this description
will fit into and explain virtually all aspects of the Special The-
ory of Relativity. The topic is discussed in some more detail in
Chapter 2, Section 2.6 ff.

Since The Linear Theory of Elasticity is developed in order to


describe only very small deformations – in fact only infinitesimal
deformations – I will suggest some more fundamental proper-
ties of space that might describe also the greatest deformations.
This I have done to some depth in Chapter 4, but here I will
only outline the basic ideas. First and foremost space is a true
continuum with no inner structure. The continuum, however,
will exert resistance against being deformed and compressed,
and as a result of such enforcements, the continuum will be able
to contain huge amounts of deformation energy, for example if
it is compressed to a fraction of its initial volume. Furthermore
all deformations are reversible so that no strain can bring about
any permanent deformations in the continuum, and all the en-
ergy that goes into a deformation will be recovered when the
12 Basic Ideas

strain subsides, i.e. there are no hysteresis involved in the de-


formations. It shows up that such a continuum will obey all the
static rules encountered in The Linear Theory of Elasticity (see
e.g. section 4.4) so everything one can deduce form that theory
apply equally well to the spatial continuum.

However, the equations are not complete until it also can deal
with motion. Thus space has got to have some inertial prop-
erties, and a crucial question will be to figure out how inertia
enters the picture. This is not an entirely trivial question. In
Chapter 4 I have looked into two possibilities. First what I have
called the classical approach; that space has an intrinsic iner-
tia in its undeformed state that only changes as space is being
compressed or inflated; and second the spatial approach that
space initially has no inertia at all and only gets its inertia from
the deformation energy by being compressed or perhaps even
inflated. Note that the spatial approach actually is more in line
with how ordinary matter gets its inertia (m = e/c2 ).

This completes the preliminary definition of space, and on this


grounds I will try to show that all known properties of matter
and space just might be explained without bringing in any ad-
ditional assumption about the nature of space. I must stress
that I don’t pretend to be able to prove that it has to be so –
that would require me to rewrite the whole of physics – so I will
just take some of the pieces of the Great Puzzle that mother
Nature presents to us and try to see how they might fit into the
outlined picture of space.
Electromagnetism compared with elastodynamics 13

1.2 Electromagnetism compared with


elastodynamics

This section is a short historical review of the development of


a couple of the most significant equations that were developed
in the nineteenth century, and a couple of interpretations that
were launched at the dawn of the twentieth century. A major
breakthrough came in the 1860’s when Maxwell presented his
electromagnetic equations which have been valid in describing
literally all non-quantum electromagnetic properties to this day.
They led directly to Heinrik Hertz’ demonstration of the electro-
magnetic radiation followed by threadless communication, and
opened for a deeper understanding of the building blocks of
matter through Quantum Electro Dynamics (QED).

Maxwell presented his finding as eight equations in 1861, but


they have later been reformulated into a modern form of four
equations by Oliver Heaviside (1850-1925). The two first equa-
tions (see Figure 1.1), are Gauss’s law for magnetism and elec-
tric charge. The next one is Faraday’s law of induction, and
the final equation is Ampère’s circuital law. The first equa-
tion tells us that the magnetic field is solenoidal, i.e. the field
lines do not have any beginning or end (there are no magnetic
monopoles). The next two equations states that electric fields
have their sources in electric charges (ii), or may be generated
by a change in the magnetic field vector (iii). The last equa-
tion, (iv), shows us that the sum of the electric current and
a change in the electric field vector (Maxwell’s correction to
14 Basic Ideas

The Heaviside version of Maxwell’s equations are a set of four equations


to describe all electromagnetic phenomena of nature. In SI units they
take the form:

i) div B = 0,
ρ
ii) div E = ,
ε0
∂B
iii) curl E = − ,
∂t
∂E
iv) curl B = µ0 ε0 + µ0 j.
∂t

We can add the field energy equation that follows naturally from the
above equations

ε0 2 1
v) u= E + B2 .
2 2µ0

Figure 1.1: The Maxwell equations.

Ampère’s law) make up a solenoidal field that create a mag-


netic field around the field lines.
The equations were published in approximately this form by
James Clerk Maxwell (1831–1879) in 1861 when he was profes-
sor in London. Fundamental to the thoughts of the time was
the discussion between several scientists around the properties
of the ether, which then was believed to fill all space. The dis-
cussion went on especially between Maxwell and William Thom-
son (Lord Kelvin) (1824–1907) about the nature of electricity
and magnetism. Thomson attributed as early as 1847 a linear
character to electric force and electric current, and a rotatory
Electromagnetism compared with elastodynamics 15

character to magnetism, while Maxwell at first in 1855 regarded


magnetic force as a linear, and electric current as a rotatory
phenomenon, but he later adopted Thomson’s view: “The trans-
ference of electrolytes in fixed directions by the electric current,
and the rotation of polarized light in fixed directions by mag-
netic force” he wrote in a couple of memoirs in 1861-62, “are
the facts the consideration of which has induced me to regard
magnetism as a phenomenon of rotation, and electric current
as phenomena of translation.” [10, page 247]. Kelvin fulfilled
his view in his Math. and Phys. Papers, iii (1890), p. 436,
where he showed that in his model a linear current could be “
. . . represented by a piece of endless cord, of the same quality as
the solid and embedded in it, if a tangential force were applied
to the cord uniformly all round the circuit. The force so applied
tangentially produce a tangential drag on the surrounding solid;
and the rotary displacement thus caused is everywhere propor-
tional to the magnetic vector ” (see [10, page 279-280] and Figure
1.2).
Maxwell’s equations are not invariant by transformations be-
tween Cartesian coordinate systems in rectilinear movement rel-
ative to each other, e.g. a stationary charge density pattern in
one coordinate system will represent a current when viewed from
within a moving frame. The equations can, however, be rewrit-
ten in such a form that they are invariant by transformations
between different Lorentz1 coordinate systems. In this rather
modern nomenclature the electric and magnetic field vectors
are merged together into a single tensor F, and the charge den-
1 Hendrik Antoon Lorentz (1853–1928)
16 Basic Ideas

The Navier-Cauchy equation for an elastic continuum takes the form

(λs + 2µs ) grad div u − µs curl curl u − ρs ü = −b.

In a continuum of infinite extension Kelvin’s theorem states that the


energy density is given by

λs + 2µs µs ρs 2
e= ( div u)2 + ( curl u)2 + u̇ .
2 2 2

According to Helmhotz’ decompositions theorem the N-C equation can


be split into two literally independent equations

(λs + 2µs ) grad div u1 − ρs ü1 = −b1 ,


−µs curl curl u2 − ρs ü2 = −b2 ,

where u = u1 + u2 = grad Ψ + curl A, and div A = 0.

Figure 1.3: The Navier-Cauchy equation.

sity and current into a vector J in a four-dimensional manifold.


Maxwell’s equations can then be written in a very compact form,
and Hermann Minkowski (1864–1909) showed in 1908 that elec-
tromagnetic forces, the energy density, the energy transport and
momentum can be expressed by the 16 components of a single
stress energy tensor T. Written in this nomenclature the equa-
tions are invariant by transformation between different Lorentz
frames that are in uniform, rectilinear movement relative to each
other. On the other hand, if our measuring devises are exclu-
sively based on electromagnetic properties, then all our reference
frames would be strictly Lorentzian and it would be impossible
to distinguish one Lorentzian frame from another, i.e it would
Electromagnetism compared with elastodynamics 17

be impossible to see any movements through space by consid-


ering electromagnetic phenomena. This is not a far-fetched as-
sumption since we know that all forces, except the extremely
weak gravitational forces and the extremely short ranged strong
forces, are of electromagnetic origin. On this grounds it would
be quite legible to launch the principle of relativity to all move-
ments in space. Einstein, however, in 1905 elevated the principle
of relativity to be a fundamental law of nature, and in the time
that has elapsed since then, it is his view that has gained more
and more momentum, even if it still is a controversial question.
In a memoir of 1821
published in 1827,
Claude Louis-Marie-
Henry Navier (1875–
1836) presented an
equation for an elastic
solid. The equation
was later developed
further - mainly by
Sir George Gabriel
Stokes (1819–1903) -
to also include fluids,
Figure 1.2: Lord Kelvin’s cord. but in this paper I
will mostly consider
the Linear Theory of Elasticity, which is a discipline in its
own right. For a homogeneous and isotropic continuum the
Navier-Cauchy2 equation takes the form shown in Figure 1.3.
2 Augustin Louis Cauchy(1789-1857)
18 Basic Ideas

The vector field, u, represents the displacement of any points


of the continuum from their original positions, b represents a
(hypothetical) body force presumably from the outside word
(but it can be shown to have other implications), ρs is the mass
density, and λs and µs are Lamé’s elastic constants3 .
The energy in a deformation field is given by Kelvin’s theorem
which implies that in a body with such a great extension that
there are no surface forces acting, the energy density can be ex-
pressed solely by the properties u̇, div u, and curl u squared; i.e.
Kelvin’s theorem is valid in a continuum of infinite extension:
In this paper I will only consider a spatial continuum of infinite
extension (or nearly so), so Kelvin’s theorem will be valid.
Hermann von Helmholtz (1821–94) has shown that a vector field
can be divided into an irrotational (curl-free) and an solenoidal
(divergence-free) field. These two fields give rise to different
wave equations describing longitudinal and transversal waves
respectively, the former moving with about the double speed
of the latter. These two waveforms are literally independent of
each other, and even if they are initiated at the same point (as
they might be for instance by an earthquake) they live their own
lives and spread independently (e.g. as P- and S-waves). Thus
in all dynamic connections the Navier-Cauchy equation can be
divided into one irrotational and one solenoidal part.
So much for history. In the next section I will try to compare the
N-C equation and some mathematical identities in an isotropic
and homogeneous elastic continuum of (nearly) infinite exten-
sion with Maxwell’s equations. I will do a more comprehensive
3 Note that ρ is the spatial mass density while ρ will be used to express
s
the sink-source density. Similarly the Lamé constants λs and µs are used
with indices to avoid confusion with electrodynamic constants.
Maxwell’s and The Navier-Cauchy Equations 19

comparison in Chapter 2 where also the relativistic properties


will be discussed.

1.3 Maxwell’s and The Navier-


Cauchy Equations
In this section I will make a comparison between Maxwell’s
equations and the solenoidal part of the Navier-Cauchy equation
for an elastic continuum of infinite extension. By redefining the
terms in the N-C equation slightly, we get a new set of equations
that formally are like Maxwell’s equations except from equation
ii) that needs a closer examination (see Figure 1.4). In the dis-
cipline of hydrodynamics it is common to use hypothetical sinks
and sources as means to mathematically describe real moving
bodies of different shapes in a perfect fluid. Here, however, I will
assume that there may be free sinks and sources as real point-like
entities in the spatial continuum. How such entities can be real-
ized will be discussed elsewhere, but here it is only necessary to
state that a source will be seen as a negative sink, and that they
can only be created in pairs, one equally strong source for every
sink. Let the strength of a sink be defined as the inflow of spa-
tial mass per time unit, and let a number of sinks be smoothly
distributed in space. We can then define a sink density, ρ, as
the sum of all sinks in a small volume element – that still is
great enough to contain many sinks – divided by the volume of
the volume element. We can then put up the last equation that
completes the comparison between the Navier-Cauchy equation
20 Basic Ideas

By reformulating the terms in the N-C equation

1
b = j, u̇ = −E, curl u = B, ρs = ε0 , µs = ,
µ0

it transforms into

∂E
iv) curl B = ε0 µ0 + µ0 j.
∂t

Generally we have that the divergence of a curl is zero, so

i) div B = 0,

∂( curl u)
and by the identity ∂t
= curl ( ∂u
∂t
) we have

∂B
iii) = − curl E.
∂t

By Kelvin’s theorem the field energy density becomes

ε0 2 1
v) e= E + B2 .
2 2µ0

Figure 1.4: Reformulating the N-C equation

and Maxwell’s equations (see Figure 1.4). We also find that


there is a hidden dependency between the properties ρ and j.
If our initial condition holds that there may be freely movable
sinks and sources in the spatial continuum, and that they can
only be created by pair production, then Equation vi) can only
Maxwell’s and The Navier-Cauchy Equations 21

Let qτ be the sum of all sinks (sources are negative sinks) inside a
volume element τ , and let τ shrink towards a little volume ǫ that still
contains many sources. Then we can define a sink density given by

qτ 1
I
ρ = lim = −ε0 lim u̇ndf = −ε0 div u̇,
τ →ǫ τ τ →ǫ τ τ
ii) ρ = ε0 div E.

There is a dependency between ρ and j.


Take the divergence of Equation iv):

div curl B = ε0 µ0 div Ė + µ0 div j,


ε0 div Ė = − div j,

and take the partial derivative of Equation ii) with respect on t:

ε0 div Ė = ρ̇.

By subtracting the two equations from each other we obtain the conti-
nuity equation

vi) ρ̇ + div j = 0.

Figure 1.5: Density of sinks

be interpreted as a continuity equation, meaning that a change


of sink density in an area can only be accomplished by an in-
or outflow of sinks. Hence j has got to represent a flow of sinks,
and further, since j originally was set like the body force b, that
a flow of sinks will create a drag in the spatial continuum just
as lord Kelvin’s proposed in 1890 (see Figure 1.2).
22 Basic Ideas

The assumption that div u̇ 6= 0 while div u ≡ 0 can only be re-


alized by assuming that the equivalents to positive and negative
electric charges are point-like entities. Thus everywhere in be-
tween the sinks and sources, and hence all over space since the
singularities themselves do not contribute to the mean density of
the spatial continuum, we have that div u = 0 as required. We
know from Quantum Electro Dynamics (QED) that it is pos-
sible to explain forces between electric charges by an exchange
of photons between them, so a further discussion about how
sinks and sources can be formed in a continuum, has got to be
discussed in connection with the nature of electric charges, but
once sinks and sources are realized as viable entities, they will
behave exactly like electric charges and exert the drag on the
spatial continuum that lord Kelvin predicted.
The above equations were all developed in a Euclidian coordi-
nate system, and they are not invariant by transformation be-
tween such coordinate systems in relative motion to each other.
For example can a static pattern of sinks and sources in one
frame be seen as a flow of the entities in another. It is, how-
ever, possible to go a step further and show that the fields may
be developed in frame independent form, which make them in-
variant by transformations between different Lorentz coordinate
systems. Hence, even if the initial comparison between elasto-
dynamic and electromagnetic fields were performed in a fixed
frame, the result would be equally valid in any Lorenz frame in
uniform rectilinear motion.
Naturally the question of which frame the observer measures
the phenomena in, will immediately arise. The observer has
Maxwell’s and The Navier-Cauchy Equations 23

got to rely on measuring rods and clocks that he brings with


him, and if they are subject to changes when they move along
with him as he performs his observations, the result will be
affected. Moreover, if the devices for measuring length and time
is changed in agreement with the Lorenz contraction and time
dilatation as G.F. FitzGerald (1851-1901) has proposed, then
the observer’s frame would be strictly Lorentzian and he would
have no means whatsoever to find out how he is moving through
the spatial continuum.4
This leads us to a philosophical question. If we cannot detect
any motion through space with any of our means even if there
really is a fixed reference frame, then why bother about it?
We should use Occam’s razor and dispose of the whole con-
cept: There is no ether out there! This is what was done at
the passage from the 19th to the 20th century, and the result
was Special Relativity. Since then there has never been dis-
covered any phenomenon that enables us to measure a speed
through space so SR has proved very well fitted to describe
all kinds of movements explicitly in relation to other bodies,
because a discrepancy would mean that an Ether would be de-
tectable. The principle, which has been ascribed to William
of Occam (c. 1295−1349), is usually translated from Latin to
4 From Larmor, Joseph (1900), Aether and matter, Cambridge, [Eng-

land]: Cambridge University Press : n.p. I quote: ”... if the internal forces
of a material system arise wholly from electromagnetic actions between the
system of electrons which constitute the atoms, then the effect of imparting
to a steady material system a uniform velocity of translation is to produce
a
puniform contraction of the system in the direction of motion, of amount
1 − v2 /c2 .”
24 Basic Ideas

mean that entities should not be multiplied beyond necessity. It


states that one should not take into account more than is neces-
sary to describe a phenomenon, hence exit of the Ether. Strong
forces, however, cannot be described by electromagnetic laws
of physic, so no wonder it has proved very difficult to wrestle
the laws of the smallest entities under the law of SR. It should
be sufficient to mention the phenomenon of non locality; two
particles may be entangled and influence each other instantly
no matter how great the distance between them is. We can de-
scribe gravity with General Relativity, but we cannot (as yet)
describe the forces of gravity along the same line as the force of
electromagnetism and the Standard Model. Finally the Cosmic
Background radiation seems to be a fixed reference frame that
we can determine our speed in relation to, but for all practical
use, SR will always be a handy tool even if we should come
to accepting that there is a spatial continuum, which at least
locally could represent a fixed reference frame.

1.4 Confined waves in the Spatial con-


tinuum
We know from Einstein’s5 famous equation, e = mc2 , that en-
ergy and mass are equivalent properties, a view that fits per-
fectly well into this model. In the defined spatial continuum
there are only two basic forms of energy, namely potential and
5 Albert Einstein (1879-1955)
Confined waves in the Spatial continuum 25

kinetic energy, so matter simply has got to be confined energy of


this kind. There are two known waveforms in a spatial contin-
uum that both may be described by the Navier-Cauchy equation
of motion, namely irrotational and solenoidal waves. Both kinds
of waves will carry energy and momentum and it is natural first
to try to identify matter as confined wave energy of this type.
Say that one of these waveforms, or a mixture of them, is con-
fined in a certain area of space from where they are not allowed
to escape. If that is so, it has got to be a property of space
itself that keeps them at bay and reflects them back to the area
in which they are confined. Since the waves also have momen-
tum, this reflection will mean that an outward directed force is
involved, and accordingly a certain amount of spatial mass will
be displaced from the area in question. The radiation pressure
in isotropic radiation is one third of the energy density in the
radiation [7, page 670], so to maintain a variable energy den-
sity the radiation itself has got to introduce a body forces into
the space in which the radiation exists. This body force can be
inserted into the Navier-Cauchy equation and it can be shown
that a total amount of radiation energy confined in this way will
displace a certain amount of spatial mass that is completely in-
dependent of how the energy is distributed in space Equation
(5.10).

A pattern of would meet some of the properties of matter. It


could move in space and would possess momentum. To acceler-
ate it we would have to add energy and thus increase the total
amount of energy and accordingly the mass. Great or small
masses, like the planets in the sky, or smaller bodies on the
26 Basic Ideas

d(mv) ds
Newton’s second law: F = , where v= .
dt dt
Einstein’s energy equation: E = mc .2
From these equations we have:

d(mv) d E  1 
dE = ds = v ds = 2 v · dE + E · dv dv,
dt dt c2 c
dE 1 v · dv
= 2 ,
E c 1 − v2 /c2
1 C
ln E = ln p + ln C = ln p ,
1 − v2 /c2 1 − v2 /c2
E0
E= p ,
1 − v2 /c2
m0
m= p .
1 − v2 /c2

NEWTON’S SECOND LAW is relativistic if we combine it with the


knowledge of the 20th century that matter and energy are equivalent
properties (E = mc2 ). The change in energy is like the force times the
distance over which the force is acting so that the resulting energy of
the system is the rest energy plus the energy added by the accelerating
force.

Figure 1.6: Newton’s second law.

earth, can with a high degree of accuracy be described by New-


ton’s6 laws of motion, but at great velocities compared to the
speed of light, his laws seems to be insufficient because mass
increases with speed. Newton, however, formulated his second
law of motion by stating that the change of motion (not the

6 Isaac Newton (1643-1727)


Confined waves in the Spatial continuum 27

change of velocity as we tend to formulate it) is proportional


to the force that acts upon a body. It is impressive that he in-
tentional or unintentional formulated his law this way, because
if we interpret motion as momentum, and combine it with the
knowledge of the 20th century that matter and energy are equiv-
alent properties so that the energy that accelerates a body goes
as an addition to the bodies rest energy, Newton’s second law
of motion would in fact be relativistic in this sense of the word
(see Figure 1.6).
However, even if the idea that matter entirely consists of con-
fined wave energy meets some of the properties of matter, it
immediately encounters several serious objections. How can
wave energy be confined inside small volumes? That is forbid-
den by Huygen’s principle, which states that every point on a
wave front may be considered to be a new source of disturbance
from which spherical wavelets issue. Thus ordinary waves in a
continuum will always spread in space and die out. Elementary
particles like electrons are known to be point-like, or very nearly
so. How then can a wave packet be point-like? An even more
basic difficulty arises. Waves in an elastic continuum is as we
have seen either solenoidal or irrotational the former propagat-
ing with approximately the double of the speed of the latter.
These waves are literally independent of each other, but both
effects are needed to create a model of matter.
By first taking the divergence and then the curl of Navier-
Cauchy equation, we get two wave equation that represent lon-
gitudinal and transversal waves in the spatial continuum (see
Figure 1.7), the first propagating with a speed that is about the
28 Basic Ideas

By first taking the divergence and then the curl of the Navier-Cauchy
equation we derive

ρs ∂ 2 ( div u)
∇2 ( div u) = ,
λs + 2µs ∂t2
ρs ∂ 2 ( curl u)
∇2 ( curl u) = .
µs ∂t2

Next by defining two new properties, Ψ = div u, and, B = curl u, they


transform into the wave equations

1 ∂2Ψ
− + ∇2 Ψ = 0,
c12 ∂t2
1 ∂2B
− + ∇2 B = 0.
c02 ∂t2

The compression/rarefication scalar, Ψ,prepresents a longitudinal wave


that propagates with the speed c1 = ρs /(λs + 2µs ), and the rota-
tionalpvector, B, a transversal wave that propagates with the speed
c0 = ρs /µs .

Figure 1.7: Wave equations.

double of the latter. Such waves would spread in space accord-


ing to Huygens principle 7 and inevitably die out. Besides they
could in no way be able to represent material particles that is
known to be very small and perhaps even point-like. It is there-
for necessary to seek other waveforms than these waves, which
in connection with earthquakes also are termed P- and S-waves

7 Christiaan Huygens (1629-95)


Confined waves in the Spatial continuum 29

respectively.

Standing longitudinal waves bouncing back and forth between a center


node and a hypothetical rigid shell embedded in the spatial continuum.

div u

u
A>0

A=0
0

A<0

r
0 h sin(pr) cos(pr) i R1
u(r, t) = A cos(c1 pt) · − r̂,
p3 r 2 p2 r
A
div u = cos(c1 pt) sin(pr)
pr
A  
= sin(pr + c1 pt) + sin(pr − c1 pt) ,
2pr

Here p is a constant, c1 the propagating speed for longitudinal waves, A


the amplitude (A in the figure is an integration constant), r the radius
and R1 the radius from the center to the first possible rigid shell.

Figure 1.8: Standing waves.

In Chapter 3 I have considered a central symmetrical longitu-


dinal waves bouncing back and forth between the center and
the inner surface of a hypothetical rigid sphere. The setting
30 Basic Ideas

is strictly mathematical, but it opens for a study of a possible


pattern of standing waves in the spatial continuum. By solving
the Navier-Cauchy equation for a spherical symmetric case, we
find that such standing waves really are possible. One solution
is shown in Figure 1.8. When the integration constant, A, is
set to zero, the required border condition, u(0, t) = 0, at the
center is fulfilled. The crucial point in this thought experiment,
is that a singularity in the center of the shell can be a node in a
standing wave given the right initial conditions. By increasing
the frequency we can get several harmonic oscillations inside the
sphere, and the sphere itself can be enlarged ad infinitum with
the same oscillatory frequency if the other border condition,
u(Rn , t) = 0, at the surface of the shell is fulfilled.

It can be shown that standing waves can be decomposed into


two progressive shell-waves, one moving outwards and the other
converging inwards towards the center. Since progressive waves
will move literally independent of each other, it is possible to
describe even complex systems of oscillating nodes just by su-
perposing the amplitudes from all the nodes in the system. First
let us see if a single oscillating node is feasible. One necessary
condition has got to be that the energy in the system is finite.
It turns out that the total energy will be infinite even by the
smallest amplitudes, so one single oscillating node is not feasible.
Another pattern it could be worth to consider is an oscillating
dipole. So let two nodes in the vicinity of each other be oscillat-
ing in opposite phase with the same frequency and let us treat
the field between them by superposing the progressive waves
from each of the nodes. We get a dipole-like oscillation, but
Confined waves in the Spatial continuum 31

then the energy needed would still approach infinity, so even a


single oscillating dipole is not feasible. If we, however, organize
a long chain of identical oscillating nodes at a suitable distance
from each other, such that every node in the chain is oscillating
in opposite phase with its two neighbors, the field in the trans-
verse direction to the chain tends to cancel out with distance,
and we get a finite field energy per unit length. Hence if space at
all should be filled with oscillating nodes of the type indicated
above, they will tend to organize along strings.
Along the same guidelines it is possible to show that even
solenoidal standing waves are feasible. As with irrotational
waves they can be seen as a superposition of progressive waves
moving in opposite direction. The propagating speed is, how-
ever, only about half of of the speed of the other, namely c0 . In
this connection an interesting question arises. If solenoidal and
irrotational standing waves exist side by side in the spatial con-
tinuum, would it then be any coupling between them? Based
on the Navier-Cauchy the answer is no. The two waveforms are
literally independent of each other and cannot interact. The N-
C equation, however, is not complete because it only deals with
small (in fact only infinitesimal) deformations. This proves to
be sufficient for all large scale considerations, but in the area
around the singularities the deformation becomes significant.
Here we have to take into account that the acceleration of a vol-
ume element is given by Du̇/Dt = ∂ u̇/∂t + (u̇∇)u̇, that leads
to the Navier-Stokes equation which also is valid for a fluid.
It can be shown that in the area near the center node, there can
be a transfer of energy from the irrotational to the solenoidal
32 Basic Ideas

field and back again, so at least one necessary condition for


a coupling is fulfilled. Hence in a hypothetical setup with the
rigid sphere in the first node, there might be a standard mode of
coupled oscillation where the two waveforms oscillate with the
same frequency. The drain of energy from the irrotational field
could slow down the reflection from the center node so much
that it falls in step with the oscillation in the solenoidal field.
Finally the transfer of energy from displacement to rotation oc-
cur when the displacement is building up to a maximum and is
reversed when the displacement is returning to neutral. Hence
the rotation has got to be in phase with the displacement.

I have not been able to prove that this may be a feasible situ-
ation, but all coupled oscillations seem to have such standard
modes where they oscillate in step with each other with the same
frequency. I assume that this is a possible situation. Now let us
try to expand the oscillation to higher harmonics with several
wavelength between the center node and the rigid sphere. First
I think the coupling outside the first node will be to feeble to
accomplish any exchange of energy. Therefore the next node in
the solenoidal field will not coincide with any node in the irro-
tational field, but perhaps that might be so in the third node.
We know that the wave speed in the irrotational field is about
the double of that in the solenoidal field, and the possibility is
surely present that the relation is exactly 2 to 1. If that is so,
the third node in the solenoidal field will exactly coincide with
the second node in the irrotational field, and a new pattern of
standing wave is realized. In this way we could expand the sys-
tem ad infinitum. I will stretch my imagination one step further
The Big Bang 33

and assume that a string of such nodes is possible. The mecha-


nism might well be different from what I have anticipated above,
but much of what is needed to build a viable model of matter in
the spatial continuum depends on the assumption that strings
of coupled oscillating nodes are real entities.

1.5 The Big Bang


According to the standard cosmological theory of today, our
entire universe was created in the Big Bang. The theory tells
us nothing about what preceded the Big Bang, but it is very
well suited to describing what happened afterwards; even back
to the first fraction of a second. The theory not only predicts
how matter was created, but it also implies that space itself was
created in the same gigantic explosion. Naturally we have got
to try to explain what might have brought about an event of
this dimension in an elastic space.
The two familiar waveforms in an elastic continuum are longitu-
dinal and transversal waves, the former moving with about the
double speed of the latter. Consider longitudinal waves spread-
ing out from an imaginary pulsating sphere in the spatial contin-
uum. Such shell formed waves can be described by solving the
Navier-Cauchy equation, and there are always two solutions to
this problem; one for waves that spread out from the disturbance
and another for waves that travel inwards towards a focal point
in the middle of the sphere. By making the sphere sufficiently
large we could at the same time create a compression pulse mov-
34 Basic Ideas

ing outwards from the sphere by suddenly enlarging the radius


of the imaginary sphere by a certain amount, for instance ∆r,
and a depression pulse moving inwards towards the center of the
sphere. If we on the other hand shrink the radius by a similar
amount, the compression pulse will be moving inwards and the
depression pulse outwards (see Figure 1.9). Exactly what will
happen when the compression or depression pulse reaches the
center of the sphere cannot be read out of the N-C equation be-
cause the amplitudes, i.e. compression or depression, increase
proportional to 1/r and hence reaches a singularity at the focal
point. Certainly, however, the surplus amount of the spatial
continuum in the compression pulse, or the displaced amount
of the spatial continuum from the depression pulse have got in-
crease or decrease the density in a very little volume around
the center for a short time before the pulse eventually might be
reflected back into space.

A phenomenon that involves inward moving compression pulses


can be studied in a phenomenon called sonoluminescence. By
introducing a strong sound wave into a liquid containing small
cavitation bubbles, the liquid pressure will oscillate in step with
the pressure in the sound waves. When pressure in this way
falls, the bubbles will grow to a maximum, but when the pres-
sure again builds up, the bubbles will undergo a dramatic vol-
ume reduction. The radii will be reduced to a hundredth and the
volumes to a millionth of their normal value. This violent implo-
sion will continue until the pressure inside each bubble becomes
100,000 times the atmospheric pressure, and a temperature high
enough for light, with frequencies up to the ultraviolet part of
The Big Bang 35

the specter, to be emitted [9].

Spherical pulses in the spatial continuum.

f (c1 t − r) f (c1 t + r)
div u = + .
r r

A sudden reduction, ∆r, in the radius of an imaginary sphere with


radius r in the spatial continuum leads to an imploding compression
wave traveling towards the center of the sphere.

Figure 1.9: Imploding shock wave.

The implosion goes on in two steps: First the bubble shrinks


to a minimum size when it slams into the confined air in the
bubble at a speed more than four times the speed of sound
(Physics News 307, February 12, 1997). Then the sudden stop
of inward movement creates a shock wave in the trapped air that
36 Basic Ideas

converges towards a focal point in the middle of the bubble. How


light is being emitted is still a disputed question, but one of the
best propositions is that the high temperature results in the
production of relatively cold plasma that emits light by thermal
Bremsstrahlung, a process in which collisions between electrons,
and the accelerating charges associated with these collisions,
emits a broad spectrum. In this connection, however, I am
more interested in the fact that while the implosion is dramatic
the release of the pressure takes quite a while, approximately
half a sound cycle, and one can ask which forces can keep the
pressure at bay for such a long time, almost an eternity, in the
extremely small volume we are dealing with (see Figure 1.5).
Let us try to figure out what might happen if an imploding
shock wave containing a huge amount of energy, for some reason
or other, should occur in the spatial continuum. The imploding
shock wave would increase in amplitude as it approaches the
focal point of the implosion. At the same time the increased
density would lead to an increased inertial density and conse-
quently slowing down of the wave-speed. Just like a tsunami
slows down and raises to an immense tidal wave as it approaches
shallower waters near a shore, the backwards part of the com-
pression wave will catch up with the foremost part and utterly
increase the pressure. Only our imagination can possibly tell us
anything about what happens when the wave eventually slams
into the focal point where the mathematics just will collapse
into a singularity.
If the imploding wave is extremely perfect shell-formed it might
perhaps be reversed and disappear back into the depth of space
The Big Bang 37

leaving the area around the focal point as quiet as it was before,
but it is more probable that small irregularities will break down
the symmetry such that the implosion would end up not only in
one singularity, but in a plethora of singularities each of which
releasing their energy in tiny explosions. An ordinary earthly
explosion, especially that of an atomic bomb, sends particles
and gas molecules outwards with so great a speed that the site
of the explosion is evacuated. The same could happen to the
singularities resulting in evacuated bubbles in the spatial con-
tinuum. The bubbles, however, would not be stable, but would
immediately collapse and be filled up with the spatial mass from
a nearby still exploding singularity. In this way spatial mass can
come to oscillate between the nodes in a such way that the out-
flow from one node is taken up as an inflow to a couple of nearby
nodes leaving a residual of bubbles moving from one of the oscil-
lating nodes to the next. A ”bubble” in this context is taken to
mean a partly evacuated area with a singularity at its core. The
general idea is that a bubble that collapses can borrow spatial
mass from the next node in a long chain of oscillating nodes and
hence move over to this node and further to the next node and
so on in a never ending repetition. The energy that goes into
this combined oscillation and displacement makes up the total
amount of matter in the Universe, and the Universe itself will be
taken to mean the compressed space left over by the Big Bang.

Now the energy in the original imploding wave is divided in


three distinctly different energy forms: Vacuum bubbles that
displace a corresponding amount of spatial mass, the kinetic
and potential energy in the oscillating nodes, and a raised pres-
38 Basic Ideas

sure in the spatial continuum itself. The bubbles can probably


disintegrate into smaller bubbles or join with other bubbles into
bigger bubbles, but the sum of the displacement from the evacu-
ated bubbles might well be conserved when seen in a short time
span compared to the age of the universe. Likewise the number
of oscillating nodes may be a conserved property. The pressure
in the spatial continuum is raised because it, like a yeast dough,
requires more space than it did before it became populated with
evacuated bubbles. In addition the spatial continuum is com-
pressed by the residue of the original shock wave. Naturally
this compressed area of space will expand at the expense of the
surrounding not compressed area and the expanding universe is
established.

One question con-


cerning the structure
of space remains to
be answered: How
does the Universe
behave at its border?
It might perhaps
be comparable to
what can happen in
some funnel-shaped
estuaries when the
tide initiates a bore.
Figure 1.10: Sonoluminescence cycle. A tidal bore is a
wave that can travel
upstream in certain rivers with, the most famous of which is
The Big Bang 39

the Black (or Silver) Dragon of the Qiantang River in China,


but also several estuaries in England (e.g. Severn) among
others around the world. The bore itself has a very limited
extension in the direction of movement - almost like a single
wavelength with a marked increase in water level (up to 7.5
meter in the Black Dragon) - and it travels upstream with a
rather slow speed. In front the river is flowing downward in
a natural and undisturbed manner, but when the bore has
passed, the current is reversed into a rather smooth upstream
motion. Similar phenomena can occur in connection with
all kinds of wave movement when dispersive and nonlinear
effects are balanced causing a wave pulse to propagate without
distortion. Moreover since such wave pulses also can travel
through each other and regain their original waveform and else
behave more like particles than waves, Zabusky and Kruskal
at Princeton University named them solitons. At the border
of the Universe there will be a strong nonlinearity in the
spatial density as it ultimately falls down to zero and below,
and the initial wave, or from the Big Bang might well form a
solitary wave (or ”hydraulic jump”) that moves outwards with
more or less constant speed that might be several orders of
magnitude slower than the speed of transversal waves in the
continuum. Behind this wave-front the spatial continuum is
flowing outwards just like the current behind a tidal bore, but
inside the whole sphere I suppose that the pressure can be quite
uniform expanding uniformly throughout the whole Universe.

If space gets its inertia from the compression energy as I have


suggested in Chapter 4, then the mass density of space out-
40 Basic Ideas

side the compressed area approaches zero and the speed of any
wave movements will approach infinity. No waves can ever move
from an area with finite wave speed and into an area with infinite
wave speed, but will be reflected, so whatever waves there might
have been created by the explosion, are confined to the area and
have got to expand together with the expanding space. Matter
in the form of energetic bubbles may, however, move freely from
node to node within the space of the compressed volume, and
would probably distribute evenly throughout the allowed area
that henceforth will constitute the entire universe. I suspect
that this uniformity also will apply to the spatial mass itself,
and so we have a volume in space with relatively evenly dis-
tributed pressure surrounded by a zone of zero pressure, which
expands in all directions as the pressure makes space itself to ex-
pand. Along with this expansion all kinds of matter also expand
and loose energy, until matter in the form of leptons and hadrons
combine to form atoms, mainly hydrogen, helium and lithium.
At approximately the same time the universe gets transparent
to light, and the radiation takes form of thermal radiation at
about 3000 K. The universe, however, continue to expand, and
the temperature in the cosmic background radiation drops along
with the expansion until it today has reached a temperature of
ca 2.7 K, as was first observed by A. A. Penzias and R. W. Wil-
son in 1964. The dipole oscillations will also represent energy,
and like the microwave background they will also loose energy
as space expands. If they don’t decline in number, they will
have to decline in amplitude, so we must expect that all space
is filled with very faint oscillating dipoles or strings of oscillating
nodes.
Waves in the spatial continuum 41

If this picture is correct, we must assume that all matter and


thereby the stars and galaxies roughly will follow along with the
expansion of space, but that does not imply that matter, as we
now know it, is completely at rest in space. The earth moves
around the sun, the sun moves in the galaxy, and The Milky Way
itself moves in space. Recent studies of the background radiation
have revealed that the waves are coming in with remarkably
equal strength from all directions of the sky, but that is only
when one has corrected for a Doppler shift of about 380 km/s
relative to the radiation8. This feature is consistent with the
Earth moving at some 380 km/s towards the constellation Virgo.
By taking into account the relative movement of the sun in
the galaxy, one can deduce that The Milky Way as a whole
moves with a speed of approximately 600 km/s in relation to
the background radiation. If we assume that the background
radiation is isotropic in relation to the spatial continuum, it
would mean that The Milky Way moves with an absolute speed
through space that amounts to approximately 0.2% of the speed
of light.

1.6 Waves in the spatial continuum


After the Big Bang the spatial continuum is compressed behind
the hydraulic jump that brings the pressure down to a much
lower or perhaps even negative level on the outside. The wave-
8 Dipole anisotropy discovered by Conclin in 1969 and confirmed by the

COBE Project Team.


42 Basic Ideas

front moves outwards like a solitary wave with some speed, but
any objects on the inside cannot feel this velocity because they
follow the outwards movement like drifting log in a river. If an
observer on one drifting object observes other drifting objects,
he will find that they move away from him and faster the longer
away they are. From his point of view he lives in an expanding
space.

In the boiling inferno inside the expanding area there is a huge


amount of disturbance energy in all possible forms. The en-
ergy is trapped in this environment because the strong pressure
and density gradient at the border will reflect all waves that try
to escape. Among the occurring waveforms there will presum-
ably be a fraction of oscillating nodes, some of them will be in
the form of coupled solenoidal and irrotational standing waves.
They will all tend to organize along strings in the spatial contin-
uum in such a way that the spatial mass that goes into one node
is taken from the adjacent nodes in the chain. The net displace-
ment from such a chain should be zero. Next the implosion is
supposed to end up in a plethora of tiny explosions that result in
small evacuated areas around singularities. It should be all right
to term them evacuated bubbles as long as we keep in mind what
we mean by the term. Such evacuated bubbles cannot be stable,
but will immediately implode again taking spatial mass from a
nearby expanding node. We now have two elements from which
we can construct a model of matter. First evacuated bubbles
that can be interpreted as naked material particles, and next a
chain of oscillating nodes that is paving the way for the bubble.
These two elements are closely connected to each other and can-
Waves in the spatial continuum 43

not be separated. Now let us assume that one of the nodes at


the outset is inflated to a bubble. The bubble is of course not
stable, but will immediately be filled up by an inwards stream
of spatial mass taken from a nearby expanding node. When the
bubble is exactly filled up, the inwards stream of spatial mass
has its maximum speed and will continue for a while until com-
pression stops the movement and reverse the stream. In this
way nodes which are left behind the bubble has got to oscillate
with decreasing amplitudes until the oscillation eventually fades
down to the background noise. On the other hand a very faint
oscillating node at the bubbles new position cannot suddenly
deliver all the spatial mass needed to fill up the bubble, but has
got to start oscillating with increasing amplitude in good time
before the bubble arrives. In this way the bubble has got to be
preceded by a chain of nodes oscillating with increasing ampli-
tudes in the preamble, and followed by a chain of nodes with
decreasing amplitudes in what we could call the postamble 9 . In
this model absolutely every elementary particle including pho-
tons will be represented by an evacuated bubble preceded by a
preamble and followed by a postamble as sketched above. They
can err around in a cloud of oscillating nodes in a more or less
irregular way to make up a material particle, or move along in
a straight line as would be the case for the model of a photon.
In both cases the bubble will move along a chain of nodes that
either will be a straight line to make up a photon, or a more or
less interwoven line to make up a material particle.

9 Postamble is not a valid English word, but it suits the purpose, and I

have seen it used in other contexts.


44 Basic Ideas

Displacement
Flow

sin(ω(t − nπ/ω)
Dn (n, t) = h ,
t − nπ/ω
hω h sin(ωt − nπ) i
Fn (n, t) = cos(ωt − nπ) − .
t − nπ/ω ωt − nπ
X
Dn = hω.
n

5 successive steps in the progressive movement of a photon. Dn is the


displacement and Fn the flow from node #n at the time t. The sum of
the displacement from all the nodes in the chain taken together is hω.

Figure 1.11: Model of a photon.

A delta function of the type sin(ωt)/t seems to cover this sit-


uation well for the oscillation of each node along the string of
nodes when t varies between −∞ and +∞. This is a calculated
Waves in the spatial continuum 45

guess and I cannot prove it, but I will try to see what implica-
tions it will have. The δ-function above is mathematically well
documented, and it shows out that the limit of the function as
t approaches zero, is like the angular velocity ω, or 2πf where
f is the frequency of the oscillation. A period, T = 1/f , is
taken to be the time it takes for a single node in the chain to
pass from neutral to fully compressed state, back to neutral and
further to fully inflated state, and back to neutral again. In
the course of that time the bubble has been fully inflated two
times. The positive and negative amplitude of the oscillation
multiplied with a suitable constant is then measures of respec-
tively displacement from the nodes and compression in them.
The displacement from all the nodes taken together will at any
time be proportional with the frequency, and if the energy in the
particle model is proportional with the displaced spatial mass,
the energy too will be proportional with the frequency. If we
call the proportionality constant h, we arrive at the equation
E = hf , and we notice that if h happens to be like Planck’s
constant of action, this formula is identical with the well known
formula for the energy of a photon. We notice, however, that
the electromagnetic property and the speed by which the bubble
is moving along the string cannot be read out of these formu-
las, but I will address those questions along with an attempt to
understand polarization and spin in the next section.

We now turn back to the amplitudes of the oscillation by each


node. Between maximum compression and dilatation 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
46 Basic Ideas

each node will naturally be a measure of the flow to and from


the nodes, so by taking the time derivative of the displacement
we will get an expression for the flow. It turns out that the flow
may be represented by a function of the type ωt [cos(ωt)− sin(ωt)
ωt ].

We can now draw a graph of the displacement to and from


each node along the y-axis and the flow to and from the nodes
along the z-axis. Just think of the nodes as not moving entities
and the graph moving forward along the x-axis just showing
the displacement and flow to and from each of the nodes as a
function of time. Figure 1.11 shows the movement of the bubble
along the cain of nodes in 5 successive steps. The foremost graph
shows the situation when the displacement (blue graph) is from
just one of the nodes while the displacement from all of the other
nodes are all zero and the flow to and from the other nodes are at
their maxima (red graph). The bubble is gradually moving over
to the next node where it will be located at the next eighth of
a period, while the other nodes alternate between compression
and depression. The next 4 graphs show the situation 1/8, 2/8,
3/8, and 1/2 of a period later. At any time, however, the sum
of the displacement from all of the nodes has got to be exactly
the same and like the displacement from the bubble itself when
it is fully located in one of the nodes (see Figure 1.12). This
also confirms that the sum of the flow to and from all the nodes
taken together is zero at any time.
We notice that the bubble reaches its maximum value two times
in the course of one cycle of oscillation; so counting in cycles the
bubble hits its maximum at 0, π, 2π, 3π, and so forth. If we
consider a node in front of the approaching bubble, it alternates
Waves in the spatial continuum 47

The net displacement from all the nodes taken together is


X sin(ωt − nπ)
D(t) = hω , ωt − πn 6= 0.
n=−∞
ωt − nπ

First we notice that sin(ωt − nπ) = sin(ωt) · (−1)n for any n ∈


{0, 1, 2, 3, · · · }, and that (−1)−n = (−1)n . Hence


X (−1)n
D(t) = hω sin(ωt) . (1.1)
n=−∞
ωt − nπ

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

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

where csc z = 1/ sin z. Hence

1
D(t) = hω sin(ωt) = hω. (1.2)
sin(ωt)

Figure 1.12: Displacement from all the nodes.

between compression and evacuation with half a cycle’s inter-


vals, but when the bubble itself arrives, the outward stream goes
on for three quarters of a cycle to empty the node to its max-
imum level whereupon the node, now containing the bubble, is
filled up during the next three quarters of a cycle and then goes
back to alternating with half cycles intervals until it fades away
as the bubble recedes. Thus the oscillation in the trailing nodes
48 Basic Ideas

are displaced with half a period as compared with the oscillation


of the nodes in front.

This idealized model can possibly explain parts of the nature of


both material particles and photons. To serve as a model of a
material particle, the chain of nodes has got to curl up in some
way or other to form a simple helix or a more or less complicated
stretched out knot-formed line in space. The tighter the helix or
the knot is wound, the slower the bubble will advance in space,
and if the chain on the other hand is stretched out towards a
straight line the bubble will move with a speed that approaches
the characteristic speed for the movement of the bubble along
the chain, a velocity a material particle can never achieve with-
out giving up its general shape becoming a straight line and
loosing its material properties. The bubbles might even move
in a more or less random manner within a cloud of oscillating
nodes obscuring the more orderly knot-like movement. If the
chain on the other hand is a straight line from the beginning,
the bubble will always move with this characteristic speed and
thus serve as a model of a photon moving with the speed of
light. So far the model can explain properties like wavelength,
frequency and energy of a photon, and both the point-like and
wave-like properties of matter, but to serve as a complete model
it will be necessary to also include electromagnetic properties
like polarization and spin.
The photon 49

1.7 The photon

In order to bring in electromagnetic properties and spin, it is


necessary to consider coupled oscillations between irrotational
and solenoidal oscillating nodes. In Section 1.4 I discussed dif-
ferent waveforms that may occur in the spatial continuum, and
I found that a coupling between between solenoidal and irrota-
tional oscillations only can occur in the near vicinity of a dis-
continuity. I also found that the coupling has got to be in such a
phase that the maximum rotational velocity has got to coincide
with the maximum displacement such that the rotation will fol-
low a similar delta function as the displacement, i.e. ∼ sin ωt/t,
(see page 44). Hence a string of oscillation nodes will bring
with it, not only a net displacement, but also a certain amount
of intrinsic angular momentum, or spin.
Spin also has a direction in space, and since the rotation in each
node has got to alternate between two opposite directions and
such that the spin component in one node is opposite to the spin
components in its two adjacent nodes, it seems reasonable to in-
fer that the directions are traverse to the direction of the chain
of nodes, say to the right and left. In the model of a bubble
moving along a string of nodes there is hitherto no indication as
to what speed the bubble is advancing with, but it will be deter-
mined by the oscillatory frequency and the distance between the
nodes, hence v = 2f d. Let us therefore suppose that the bub-
ble is advancing with the speed of transversal waves, c. Then
the alternating left and right rotation and twist in the preamble
might well trigger a progressive transversal wave which accord-
50 Basic Ideas

ing to to this model would be identical with an electromagnetic


wave, and which – at least along the chain of nodes – would
approximate a plane wave. As the bubble itself builds up in the
course of three quarter of a cycle and decline again in the same
amount of time, the oscillations in the postamble come half a
period out of phase measured against whatever pseudo-plane
progressive wave that might have been triggered in the pream-
ble. Seemingly no progressive waves can be created under such
conditions, but let us all the same assume that there really is
created a progressive transversal wave in the preamble. It grows
stronger and stronger as the bubble approaches and reaches its
strongest level around the bubble itself. So when the spin in and
around the bubble builds up, it comes increasingly out of phase
with the progressive wave and is bound to get a precession to-
wards the chain axis and reaches its greatest value strictly along
this axis before it precesses further on until it fades out opposite
of its start-up direction. This pirouette will bring the oscillation
in the postamble in line with the traveling progressive wave and
will keep it up until it fades away towards zero as the bubble re-
cedes. Notice that the resulting spin of the entire chain of nodes
will be directed along the chain axis in the forward or backward
direction. In Figure 1.13 I have made some considerations as to
the magnitude of the spin, and it indicates that it may very well
have a fixed value of ~.

If a system like this really can occur in the spatial continuum,


it will have almost all the properties of a photon. There will
be some phenomena that needs to be investigated further, for
instance circular polarization, but the model covers every tan-
The photon 51

A strange coincident?
The energy in a rotational system is given by

E = 1/2 · Iω 2 ,

where I is the moment of inertia and ω the angular velocity of the


rotation. The spin angular momentum is given by:

L = Iω.

Say that the energy in the rotation by the principle of equipartition


is half of the total energy, ~ω, in the system, and that the angular
momentums are the same ωrot = ωtrans in both systems. Then we can
put up the relation:

2
1/2 · Iωrot = 1/2 · ~ωtrans ,
Iωrot = ~,
L = ~.

This is the correct value for the spin of a photon, but it is of course
no proof. At best it tells us that the spin is of the correct order of
magnitude.

Figure 1.13: Spin angular momentum.

gible property of the photon. It has got to move with the speed
of light to keep up with the associated electromagnetic wave,
it is polarized, it has spin directed along the movement axis,
it carries energy and momentum proportional to the displaced
spatial mass, and it has at the same time both particle and
wave nature. A single photon in this model is surrounded by a
transversal wave of considerable extension which will be subject
to ordinary interference laws, and since there is a mutual inter-
52 Basic Ideas

action between this wave and the components generated in each


node, the pilot wave will take the most favorable direction under
these circumstances and pave the way that the bubble has got
to follow. In this way the photon - even if it is a particle and can
be traced as such - will behave like a wave. If one, however, try
to trace the particle, the associated solenoidal wave will be dis-
rupted and the interference will vanish. Since the upcoming and
trailing nodes all oscillates in exact step with each other there is
a certain non-locality involved, what happens in the front may
immediately influence the nodes father back. If a lot of particles
like this move in the same direction and in step with each other,
they will produce a truly plane polarized coherent wave, i.e. a
laser beam. There is even a possibility that a photon like this
may be able to split up to make two electric charges, which will
be the topic of the next section.

1.8 The electron

In Section 1.3 I made a comparison between Maxwell’s electro-


dynamic equations and Navier-Caucy’s elastodynamic equation.
I found that electric charges can be associated with sinks and
sources in the spatial continuum. That seems at first to be an
impossible assumption, but there is a possibility. I have already
suggested that an elementary material particle in this model
might consist of a curled up string of nodes along which a bub-
ble is moving. It therefore has a preamble and a postamble that
in principle stretch far out both in the forward and backward
The electron 53

direction of the particle’s movements. The oscillations in the


nodes in the preamble build up to a displacement, which I have
dubbed a bubble, that moves one step forwards for each oscilla-
tion cycle, while the oscillations in the trailing nodes diminish
down towards nil. Say that the oscillations in the preamble of
one chain of nodes build up to a higher displacement than is
built down in the postamble, and that the situation in another
nearby chain of nodes is the opposite. The excess void in the
first chain has got to be filled up with spatial mass from the
insufficient displacement in the second chain, and we get a flow
of spatial mass between the doublets that corresponds to the
electric field between two charges. From there on it is straight-
forward to see that the farther the two entities are separated
from each other, the more energy goes into the velocity field
between them, and therefore it will be required a force like the
Coulomb force to increase the distance.

In Chapter 3 I discuss how coupled oscillations between irro-


tational and solenoidal waves could occur and which standard
modes of oscillation one could expect. I found that in the near
vicinity of a singularity there may be a transfer of energy be-
tween the two fields. Depending on the initial condition I pro-
posed that three standard modes of oscillations could be possi-
ble. If energy in the first half cycle is transferred from displace-
ment to rotation and back again in the next half cycle, then the
displacement is dampened and the rotation is enhanced. If on
the other hand the energy in the first half cycle is transferred
from rotation to displacement, then it is the rotation that is
dampened and the displacement that is enhanced. I also saw
54 Basic Ideas

the possibility that the shift between the direction of energy


transfer could happen to be in the middle of the two possibili-
ties such that no net transfer of energy would occur. We could
dub the two forms of asymmetric oscillation rotation enhanced
mode and displacement enhanced mode respectively, and the
symmetric oscillation we could dub neutral mode.

Model of two opposite charges moving in the right direction. Blue, thick
lines represent the displacements from each node, and red thick lines
represent rotation. Thin lines suggest the displacement and rotation at
the nodes in a photon. Black arrows illustrate the sink and a source
respectively.

Figure 1.14: Sink and source

If these assumptions are correct, the pre- and postamble of any


elementary particle or photon will be in one of these modes, so
let us try it out. The pre- and postamble of a photon should
be in neutral mode. The displacement builds up to the bub-
ble in the preamble and down to nil again in the postamble.
The electron 55

But say that the preamble of the first of two particles is in dis-
placement enhanced mode that shifts over to rotation enhanced
mode in the postamble, and the preamble of the second particle
is in rotation enhanced mode that shifts over to displacement
enhanced mode in the postamble. Then the displacement in the
first particle is built up to a higher displacement in the preamble
than is built down in the postamble, and the displacement in
the preamble of the second particle is built up to a lower level
than is built down in the postamble. This discrepancy can only
be equalized by a constant flow of spatial mass from the second
to the first entity like the velocity field around a source-sink
pair (see Figure 1.14). We could call the first entity e+ and the
second e− and notice that e− acts like a source and e+ as a
sink (the choice of the signs is somewhat arbitrary). We also
notice that if the distance between the two entities is increased,
then more energy will go into the velocity field between them,
and hence it requires an application of force to bring them far-
ther apart like the force between two opposite electrical charges.
What actually happens in the bubble itself is in the first case
that the inflow to the bubble boosts the rotation up from its
dampened mode to enhanced mode, and in the second case that
rotation veritably blows spatial mass away from the node on ac-
count of its own rotation. Note that this shift corresponds with
the delay in phase caused by the building up and down of the
bubble in the course of two half cycles.

From fluid dynamics we know that the spin in a perfect fluid


without viscosity is a conserved property. The spatial contin-
uum is supposed to be perfect in the sense that any deformation
56 Basic Ideas

is recovered without any loss when the stress is removed, but


unlike in a fluid or a gas, spin cannot be conserved as such.
But a spin will build up a torque that in turn is delivered back
to a spin in the opposite direction. Hence the combined effect
of spin and torque is conserved. So when a photon splits up
into two new entities, both of them can only acquire half the
integral spin each, and can therefore not be two new photons
moving along in a straight line in space. The alternative is that
the two bubbles move along two curled up strings, but still with
the speed c along the strings. Depending on how tight each of
the strings are wound up, they may move along with any group
speed v < c.
Sources and sinks according to this model can only be created
by pair production and disappear by being annihilated back to
what a photon model is like, just as is the case for electrical
charges. A surplus of sources or sinks in a volume can therefore
be changed only by a flow of sources or sinks into or out of
the volume. Thus it is possible to form a law quite similar to
the law of conservation of charge. The coulomb force between
two charges can be computed by computing the increase in the
field energy between the charges as the distance between them
is increased. It can readily be shown that the velocity field
between the source and the sink has got a kinetic energy that
exactly matches the field energy in an electric field, provided
that the amount of matter flowing between the two entities per
unit time is unchanged by distance, and hence the forces will
obey the same law, i.e. Coulomb’s law.
As we have already seen in Section Maxwell’s and The Navier-
The electron 57

Cauchy Equations on page 19: the only option for describing


electrical charges in this model is by the means of sinks and
sources in the spatial continuum. At first it seemed like an
impossible demand, but the above model shows us a way out
of the conundrum. The photon and indeed any possible parti-
cles in this model have got to have a preamble and postamble
with quite distinct properties. First and foremost the pream-
ble builds up to a displacement of spatial mass that also is a
measure of the energy of the particle. It resolves the appar-
ent paradox that a particle is more energetic the smaller it is.
Next the spin component is being built up in the preamble and
manifest itself by a resultant spin angular momentum of the
spatial mass itself, or possibly by a combination of that and a
real rotary motion of the bubble along a corkscrew-like string of
oscillating nodes. At the same time as the spin is being built up,
it triggers a transversal wave which perhaps can be fitted into
the Schrödinger equation; and the transversal solenoidal wave
redirects the spin from the transverse direction to a direction
along the movement axis of a photon. Finally an asymmetry
of the oscillation may put the preamble and postamble in three
different modes of oscillation that may describe the electrical
properties of matter. It also gives us a possible solution to what
a neutrino is like. It might be a particle where both the pre-
and postamble are in the neutral mode of oscillation.

One more question needs to be answered. If a photon splits


up into two new half spin particles oscillating with about half
the frequency of the photon, and the two particles starts to
circle around a common point in space because they do not
58 Basic Ideas

have energy enough to separate completely, they would together


make up a composite particle that is known as a positronium.
However, they cannot shear the same pre- and postamble, but
have got to make their own string of nodes to move along. These
two strings would not be allowed to interfere with each other,
so the only possible way is that the nodes, which exist in the
same environment, shall put their individual nodes in between
those of the other ones like the teeth of a zipper, and oscillate
a quarter of a period out of phase with each other. Then the
resulting oscillatory frequency of the system will be about the
same as the original photon, disregarding the energy that has
gone into the system in order to separate them.

1.9 Systems of many particles

In the previous section I assumed that the nodes of the two parti-
cles oscillate with half the frequency of the original photon, but
the frequencies are interlocked like the teeth of a zipper such
that the net frequency of the system is the double of the indi-
vidual particles, like the density of teethes in a locked zipper is
the double of the density in each parts of the open zipper. I will
follow up this thought and assume that in a composite system
of many half spin particles like the quarks and electrons in an
atom, a similar mechanism is at work. Each particle creates its
own set of nodes in such a way that the oscillation of the nodes’
peaks are in between each other. In this way the entangled set
of nodes are interlocked like the teeth of a ”multidimensional”
Systems of many particles 59

The simplest imaginable material particle is a string of nodes wound


up like the threads of a screw along which a bubble is moving with
velocity c. Its forwards advancing speed, v, however, is depending on
the pitch of the windings. Along the string of nodes there is generated
a progressive transversal wave, and the oscillation in adjacent nodes
has got to be in phase with each other. The progressive wave gen-
erates another wave, a so called phase wave, and simple geometrical
considerations show that this fictive wave, also called matter wave, is
moving with a great velocity, vp = c2 /v, through the wave packet.

Figure 1.15: Illustration of a material particle

zipper, and the frequency of the system as a whole is like the


sum of the frequencies of all the nodes. Each of the particle has
an energy given by its oscillatory frequency fn (En = hfn ), so
the system’s oscillatory frequency has got to increase propor-
tional with the energy, E, of all the half spin particles in the
system, f = E/h = E/2π~.
60 Basic Ideas

To get an impression of what the waves in connection with an


elementary particle might be like, I will consider a simple ideal-
ized case when the bubble follows a path like the threads on a
screw. The bubble itself moves along with the speed of c, but it
advances in space with a far lower group velocity vg depending
of the pitch of the windings. I think that the movement of the
bubble in a real elementary particle will be far more compli-
cated than this thought experiment should indicate, but in this
special ”particle” the bubble will move along on the surface of a
cylinder with speed c while the particle itself moves along with
the speed vg . I have tried to visualized this situation by cut-
ting open and unfolding the cylinder in Figure 1.15. The grey
lines represent the pilot wave. It has got to be in phase from
thread to thread. The bubble has got to follow the threads -
the slanted line in the figure - and the speed is dictated by the
pilot wave moving with the speed of transversal waves, c, while
the whole packet moves with the speed v. The horizontal wave
is the phase wave through the packet, and a simple geometri-
cal considerations shows that it moves with a speed far greater
than the transversal waves, namely vp = c2 /v. This oversimpli-
fied case serves only to illustrate what is going on, so in Figure
1.16 and 1.17 I have tried to investigate a more general case
that leads to the formula for the Louis de Broglie’s (1892–1987)
electron wave, and the Erwin Schrödinger’s (1887–1961) wave
equation that were discovered in the nineteen twenties and laid
the foundation for the introduction of quantum mechanics.

Once we have found the amplitude of the wave function through-


out space we know that the particle has got to be located some-
Systems of many particles 61

where in the cloud of nodes and hence we know where the parti-
cle might be. Max Born interpreted this amplitude as the prob-
ability amplitude for the particle, which squared value gives the
probability where to find the particle in space at any time. Since
the particle has got to be somewhere in space, we can normalize
this property by setting the integral of the probability all over
space to unity. According to the Copenhagen interpretation the
only fact that can be said about the position of a particle is the
probability amplitude. There is no hidden variable that could
possibly tell where the particle is situated at a given time. In
this model, however, there surely has got to be a hidden variable
where the naked particle in the form of an evacuated bubble is
located. This view was discussed by David Böhm who found
that a hidden variable could give meaning to the abstract for-
mulas of physics, but John Stewart Bell showed that then one
has to accept the principle of nonlocality10 . This is of course

10 From Wikipedia I cite: The Bohm interpretation of quantum mechan-


ics, sometimes called Bohmian mechanics, the ontological interpretation,
or the causal interpretation, is an interpretation postulated by David Bohm
(1917-92) in 1952 as an extension of Louis de Broglie’s pilot-wave theory
of 1927 . Consequently it is sometimes called the de Broglie-Bohm theory.
Bohm’s interpretation is an example of a hidden variables theory. It is
hoped that the hidden variables would provide a local deterministic objec-
tive description that would resolve or eliminate many of the paradoxes of
quantum mechanics, such as Schrödinger’s cat, the measurement problem,
the collapse of the wavefunction, and similar concerns. However, Bell’s
inequality complicates this hope, as it demonstrates that there is no local
hidden variable theory that is compatible with quantum mechanics. Thus,
one is left with choosing between the lesser of two evils: discarding local-
ity, or discarding realism. The Bohmian interpretation opts for keeping
realism and accepting nonlocality.
62 Basic Ideas

no problem in this model of matter since nonelocality is part


of the assumptions that has got to be made. The nodes in the
pre- and postamble has got to oscillate in step with each other
even if they are far apart, even in principle if they are indefi-
nitely distant from each other. With that fact established, this
interpretation fits well into the proposed model of matter. If
a particle can reach a target along two (or more) paths as in
a double slit experiment with light, the pilot wave fills up all
the possible pathways and rejoin at the target where they inter-
fere with each other making an interference pattern of weak and
strong spots, while the particle itself can follow only one of the
routes and most probably hit the target where the amplitude of
the nodes is greatest. But if we try to spot the particle in one of
the paths, we inevitably have got to disturb the pilot wave such
that the interference is lost. Indeed we have got to recalculate
the wave function with the disturbance point as a new origin
from which the particle starts anew, and now it can reach the
target only by one route and there is no interference.

1.10 Atomic nuclei and the strong


forces
The energy concentration in a nuclear particle is extremely high.
It amounts to the mass density we would have if we could
squeeze together the entire earth to the size of an orange. Keep-
ing this amount of energy in place demands an immense force,
and conversely, such an energy concentration would create a
Atomic nuclei and the strong forces 63

First consider a composite particle moving alon with speed v in a


primed coordinate system following the particle

E
f= ,
h

where E is the energy of all particles in the system, and f is the fre-
quency of the oscillations in the entire cloud of nodes. The oscillation
can be described by
′ ′
t
Ψ(x′ , y ′ , z ′ , t′ ) =A(x′ , y ′ , z ′ ) · e−i2πf
E′ ′
=A(x′ , y ′ , z ′ ) · e−i ~ t .

Next transform the movement over to the coordinate system of the


observer. It is only in Lorenz frames that the properties are invariant
by transformation, so we apply Lorenz transformations

x − vx t y − vy t z − vz t
x′ = p , y′ = p , z′ = p ,
1 − v2 /c2 1 − v2 /c2 1 − v2 /c2
t − (vx x + vy y + vz z)/c2
t′ = p .
1 − v2 /c2

From the observer’s point of view the oscillation takes the form

y−vy t
Ψ(x, y, z, t) =A √x−v2x t ,√ , √ z−v2z t

1−v /c2 1−v /c2
2 1−v /c2
√ −iE [t−(vx x+vy y+vz z)/c2 ]
1−v2 /c2
· e~ ,

i.e. it represents two waves. A group wave vg = v and a phase wave


vp = c2 /v.

Figure 1.16: Oscillation of a moving composite particle


64 Basic Ideas

We take the second derivative of Ψ with respect on t and get

∂2Ψ −i √ E [t−(vx x+vy y+vz z)/c2 ]  −iE 2


~ 1−v2 /c2
=Ae ,
∂t2
p
2
~ 1 − v /c 2

1 ∂2Ψ −E 2
2 2
=Ψ 2 2 .
c ∂t c ~ (1 − v2 /c2 )

Accordingly

∂2Ψ ∂2Ψ ∂2Ψ −E 2


+ + =Ψ 4 2 (vx 2 + vy 2 + vz 2 ),
∂x2 ∂y 2 ∂z 2 c ~ (1 − v2 /c2 )
−v2 E 2
∇2 Ψ = Ψ .
c4 ~2 (1 − v2 /c2 )

The two equations can be drawn together to

1 ∂2Ψ E 2 v2 E2
∇2 Ψ − =− 2 4 Ψ+ 2 2 Ψ
c2 ∂t2 ~ c (1 − v2 /c2 ) ~ c (1 − v2 /c2 )
E2
= Ψ.
~2 c2

Finally we define a new property m given by E = mc2 and the


equation above takes the form

1 ∂2Ψ m2 c2
∇2 Ψ − 2 2
= Ψ,
c ∂t ~2

which we recognize as de Broglie’s equation for the electron wave,


sometimes also called The Matter wave.

Figure 1.17: The Matter wave.


Atomic nuclei and the strong forces 65

colossal outward pressure that would partly evacuate the vol-


ume where the energy is confined. A nucleon would then consist
of a shell with a great radially directed grad div u-field inside
which there might be a fairly homogenous inflated area where
the elementary particles of the nucleus bounce around. It is in
an environment like that we find the quarks, and I suspects that
the difference between leptons like electrons and quarks are more
due to the differences in environment than to the nature of the
particles themselves. They are both fermions with half-integral
spin, and they are basically bubbles surrounded by clouds of
oscillating nodes, which the bubbles themselves move between.
Why the quarks do not break out of their confinement is not
easy to tell, but say that they cannot exist in any other envi-
ronment than the partly evacuated area in the nucleon. Then
their only option is to bounce back and forth inside the shell
making the shell exactly big enough to keeping up the right en-
vironment where they can exist. If their movements are mainly
along circular orbits, then there would be needed at least three
particles orbiting around three orthogonal axes to keep up a uni-
form pressure in all three spatial directions. That may be the
reason why three quarks are needed to create a stable nucleon.

Quarks come in pairs, a quark and an anti-quark, and such pairs


of quarks are the building blocks of a group of short-lived par-
ticles called meson. They disintegrate by various routs down to
photons, and thus in our picture we can say that they are sym-
metrical with a photon as their symmetry-axis, just as electrons
and a positrons. In much the same way as a positron and an
electron can form a short-lived positronium atom, a quark and
66 Basic Ideas

an anti-quark can make up a meson. Thus, according to what


I already have proposed, the two quarks have got to be out of
phase with each other existing in a cloud of oscillating nodes
with a frequency the double of what they have by themselves.
Fluctuations may, however, soon bring the converging points in
the two clouds to coincide, and then the two particles will in-
evitable annihilate each other and form a cascade of different
particles eventually ending up as photons. As the electric field
keeps the electron and positron together in a positronium, the
strong nuclear force keeps the quark and anti-quark together in
a meson. The electric field has as its counterpart a velocity field,
but such fields would be far too weak to prevent the two quarks
from flying apart. A huge energy density like that in a nucleon,
however, will displace some of the spatial continuum and create
an inflated area in which the quarks will reside. At the border
of this area, there has got to be a strong pressure gradient which
links the lower pressure in the partly evacuated area with the
normal pressure in the surroundings. It is this kind of fields
that have got to represent the strong nuclear forces, i.e. the
counterpart to the strong nuclear force is a comparable strong
grad div u-field.

A nucleon is composed of three quarks, the proton of two up


quarks and one down quark whereas a neutron has two down
quarks and one up quark. A baryon may even be built up of
three similar quarks, and this fact seems to disagree with the
Pauli exclusion principle, which states that no two particles in
an atom may be in the same quantum state. Therefore the
quarks in a baryon are ascribed different colors, which of course
Atomic nuclei and the strong forces 67

Figure 1.18: Three quarks forming a group.

are no real colors, but a kind of charge called color charge, which
distinguish them from each other. In a nucleon the quarks have
the colors red, blue, and green, which make the nucleon itself
white. The color of the quarks in a meson is always one of these
colors and its anti-color, also making the meson white. In fact,
to my knowledge, no one has ever seen a quark in isolation nor
a colored particle, which probably amounts to the same.
Consider a set of three strings, dubbed red, green, and blue, con-
fined in a closed area of the spatial continuum oscillating with
the same frequency and one sixth of a period out of phase with
each other. The net oscillating frequency of the entire group
68 Basic Ideas

is now three times that of the individual strings. Furthermore


say that the peak rotational directions in each string are suc-
cessively in the up and down direction. Since there are exactly
three different sets of nodes in the group, the net result is that
the entire group will also oscillate between the up and down
direction successively. When inflation peaks in the bubble, the
direction of the spin component gets a precession that rotates it
through the forward or backward direction in the course of one
half cycle, giving each of the entities a net spin angular momen-
tum, and flips the rotation components such that the oscillation
in the preamble and the postamble come in phase with each
other. See Figure 1.18.
The picture of these idealized particles is hardly the whole truth.
Probably will the internal life of a composite particle be more
chaotic than this model suggests, with virtual particles popping
up and disappearing all the time governed by some inherent or-
derliness that is not easy to spot. Any multiple of three single
particles, however, will add up to a smooth right/left oscilla-
tion mode, so by the same argument as above, any composite
group of such three-groups may be entangled to form a complete
atomic nucleus. If this picture is correct, it becomes a little more
understandable that particles are entangled in strange ways be-
cause they are all linked to the same orchestra of oscillating
nodes. Since bubbles always have got to move along some paths
in space it seems that some kind of string theory may be applied
to get a better understanding of nature.
Gravitation 69

1.11 Gravitation

Newton discovered the law that two heavenly bodies attract each
other with a force proportional to the product of their masses
divided by the distance between them squared. This is a force
that acts at a distance. A more modern view is to consider
gravitation as an effect of a gravitational potential, the gradi-
ent of which gives the gravitational acceleration at any point in
space. The potential is determined by an equation discovered by
the French mathematician Siméon-Denis Poisson (1781-1840):
∇2 Φ = −4πGρ where G is the gravitation constant and ρ the
mass density. Poisson’s equation can be solved under rather
general conditions yielding the gravitational potential all over
space. The tremendous advantage of this approach is that grav-
itation can be determined from any distribution of mass, and
not only from spherical bodies. A still more modern view is to
apply the General Theory of Relativity, GR, where bodies move
along geodesics in curved space-time.
It is almost obvious that the gravitational potential in this model
has got to be linked to a grad div u-field, i.e. a space with
varying density. However, from the Navier-Cauchy equation we
can immediately see that a static density gradient cannot exist
in the open spatial continuum, so we have got to seek a dynamic
solution. The key lies in the expansion of space in which a
body is represented by confined energy. Which effect it is that
keeps the energy confined is irrelevant in this connection, but it
is essential that the force needed to keep the energy at bay is
absorbed by the spatial continuum itself. Hence confined energy
70 Basic Ideas

exerts an outward directed force that displaces a certain amount


of spatial mass and is surrounded by a steep uphill pressure
gradient that keeps it at bay. The pressure in confined energy
of this kind is like one third of the energy density11 , and it can
be shown that it displaces the same amount of spatial mass
regardless of how it is distributed (see the paper chap:Confined
Energy and Gravity).
In Section 1.5 I found that the spatial continuum has got to
expand continuously, and if we place energy packets in such an
environment, they will loose energy all the time. The outwards
pressure against a receding wall exerts a work that is given by
the force times the receding speed, and the energy has got to
be taken from the confined energy. The lost energy has got
to be transported away as a wave, and the obvious choice is a
longitudinal wave. Think of a chunk of energy as a compres-
sion pulse that moves outwards with velocity c1 . By solving the
Navier-Cauchy equation we find that the amplitude of a shell-
formed compression wave will decline as a function of 1/r as it
moves outwards from the confined energy. The whole energy
transport will be in the form of an anharmonic wave, or a static
fall in pressure outwards from the area. The gravitational po-
11 This is by the way a general principle in physics. From
http://www.answers.com/topic/radiation-pressure-1 I cite: ”It may be shown
by electromagnetic theory, by quantum theory, or by thermodynamics, mak-
ing no assumptions as to the nature of the radiation, that the pressure
against a surface exposed in a space traversed by radiation uniformly in all
directions is equal to 1/3 the total radiant energy per unit volume within
that space”, so there is no reason that it should not apply to any disturbance
energy confined in an elastic continuum.
Gravitation 71

tential behaves in exactly the same way around a gravitating


body in space, thus there is a one to one correlation between
the pressure around a confined energy in an elastic continuum
and the gravitational potential around an energy concentration
in the form of mass in space. The pressure gradient is extremely
small, but then gravitational forces are only one part in 1036 of
electromagnetic forces.
An increase in pressure will lead to an increase in the mass den-
sity, ρs , in the spatial continuum and slow down the propagating
speed of all kinds of wave movement, so also the intern move-
ment of energy inside all energy packets. The speed of these
waves are assumed to be c under all conditions, so matter in the
form of wave packets will reside in a space with varying wave
speed. The waves will constantly be deflected towards areas
with lower wave speeds, and the whole energy packet will get
an acceleration towards these areas, i.e. towards another assem-
bly of energy packets, for example towards the earth. Since the
intern energy movements are the same in all bodies, heavy or
light, they will all get the same acceleration, namely the acceler-
ation of gravity. In Chapter 5 I have considered these properties
in more detail, and I found that the conformity with gravitation
is excellent.
72 Basic Ideas

1.12 The General Theory of Relativ-


ity

The assumption that an energy concentration should create


around itself a simple scalar Φ-field cannot be general enough. It
doesn’t include any dynamic properties of the confined energy.
A better approach would be that confined energy should intro-
duce a system of forces into space. A better candidate for intro-
ducing such a system of forces into space would be Minkowsy’s
symmetric stress energy tensor, T, in a four-dimensional space-
time. The trace of that tensor would give us the pressure that
displace spatial mass from the energy concentrations. T has 16
components whereof only 10 are different. The so-called Bianchi
identity reduces this 10 components to a minimum of six, and
that is the minimum of components the stress energy tensor can
be brought down to. That is still two more than the scalar field
Φ represents, even if it is expressed in the same four-dimensional
manifold as T is expressed in, and we can conclude that Φ can
only give us an approximate description of the true properties
of space.
The classic law for the conservation of energy states that the
divergence of the stress energy tensor has got to be zero (see
e.g. [8, page 132]), but that is not the case in this model when
viewed in an universal frame12 . Confined energy always has
got to loose energy due to the expansion of the universe, and
12 In a local frame, however, the divergence of the stress energy tensor

may well be like zero as I will try to verify below.


The General Theory of Relativity 73

therefore ∇ · T is different from zero, which represents the loss


of energy into deep space from the confined energy. It is the
energy flow that is supposed to generate the system of forces
that in turn results in a varying density throughout space. If
all kinds of matter surround itself with a gradually decreasing
density with distance, as indicated above, this property will in-
fluence the speed of transversal waves such that the speed of
light decreases when a photon gets into a gravitational field. In
1911 Einstein himself suggested that the velocity of light must
depend on the gravitational field, a view that he later aban-
doned, but let us investigate this possibility a little further. A
ray of light passing from point A to point B in a space with
varying c would follow an optical path from A to B that makes
the time shortest possible in accordance with Fermat’s principle
RB
of least time, δ A dsc = 0. It will pay off for the ray of light
to take a small detour through areas with greater light speed
to gain time on its way from A to B, and hence get deflected
towards areas with lower light speeds.

To execute the above calculation it is necessary to have at hand a


universal measuring system for time, distance and mass. When
we, who are living in this space, however, shall create our own
measuring devices, we have got to rely on the properties of mat-
ter. We could choose to use the simplest material particle readily
available, i.e. the hydrogen atom, as a reference to do that. The
mass of the hydrogen atom could be chosen as the unit of mass.
The wavelength of the photon that is emitted when the atom is
exited to its first energy level could be made the unit of length,
and finally we could define the time unit as the time the photon
74 Basic Ideas

uses to travel that distance. In this local frame the speed of


light would always be unity per definition. This would of course
not be the most effective way to define our basic units, but it
will serve as an example. Let us also define a universal frame
based on the same principle, but with a single hydrogen atom at
rest far away from any heavenly bodies, and let a modified kind
of Maxwell’s demon be able to see this frame from anywhere in
the universe in order to be able to compare it with the local
frame.

Now, let our demon calculate the time a ray of light takes to
pass from point A to point B past a heavy object in space where
the speed of light is known to be slower. He moves one universal
length unit for every tick of his universal clock and finds that the
shortest time has passed when he chooses a path a little farther
out from the body than the straight line. Next, let us do the
same calculation with our local length and time units. Say first
that our length unit is unaffected by the position. Then, in
order to keep the measured speed of light constant, our clock
has got to go a bit slower near the object than farther out, and
we too will find that the time is shortest when we move along
the same path as the demon found. If on the other hand the
time unit is unaffected by the position, then, for the same reason
as above, the length unit has got to be a bit shorter near the
heavy object than farther out, and again we get a bit faster past
the object by choosing the same route because of the slightly
longer length unit covered by each tick of the clock along this
route than along a path closer to the heavy object. The route
we found in this way, by applying the calculus of variation, is
The General Theory of Relativity 75

called a geodesic; in this case – since it is a ray of light that is


considered – it is called a null geodesic. Finally a more probable
choice of units would be that the time runs a bit slower and the
meter becomes a bit shorter than the universal units, and the
space-time in the area around the great mass could be said to
be curved. To mess around i such a space is no simple task, but
luckily for Albert Einstein (1879-1955) had Bernhard Riemann
(1826-1866) and others already developed a geometry that could
be used in a space-time like that. So when he – assisted by
Marcel Grossmann (1878-1936) – in 1916 developed his Special
Relativity (SR) the geometry was already in place. According to
the Riemannian geometry the measuring units at every point in
space are given by the ten different components13 of the metric
tensor, g.
How then will a material particle or a body like a feather or
even the earth move in space-time? In this model matter is
composed of elementary particles which all consist of bubbles
moving along strings of nodes with the same speed, namely the
speed of light. They will all be deviated by the geometry of
space-time and in a complicated way follow geodesics just like
a single ray of light. The resulting movement will therefore also
be geodesics though considerably more deviated ones, because
each of the intrinsic particles moves a far longer way than the
body itself, as the material body moves between two positions
in space.
In Figure 1.19 I have tried to illustrate what is going on. The
13 Strictly speaking, it is only six because of the Bianchi identity.
76 Basic Ideas

The stress energy tensor in a universal frame has got to be given a


correction in order to produce the stress energy tensor measured in
a local frame:

T[universal] − T[loss] = T[local] .

If we first define a new tensor

8πG
R=− T[universal] ,
c4

and make the construct

1 8πG
R · g = − 4 T[loss] ,
2 c

where R = Rγ γ , and finally put these results in the above equation

1 8πG
R− R · g = − 4 T,
2 c

we arrive at an equation that is similar to the Einstein field equation.

Figure 1.19: The Einstein field equation

universal measurement of the stress energy tensor has got to


be given a correction in order to find the local measurement.
By defining a couple of new properties, this relation is brought
over to a form that formally is identical with the Einstein field
equation. This may give us a clue to interpret the terms in the
The General Theory of Relativity 77

Einstein field equation. First we see that the Ricci tensor might
be associated with the real stress energy tensor measured in a
universal frame. The trace of that tensor produces the Ricci
scalar, which when corrected by the metric tensor should be a
measure of the loss of energy from the system.
According to my definition of the spatial continuum, there has
got to be a residual pressure leftover from the Big Bang. This
pressure is uniform throughout space – or perhaps only nearly
so – and will contribute to the stress energy tensor. Due to the
expansion of space, it will also loose energy and cause an effect,
Λ · gαβ , that may be significant over great distances in space. It
is of course an open question if the gαβ found here is the same
as the Riemannian metric cited above, but I will leave it at that.
In 1917 Willem de Sitter and Abbé George Lemaitre showed
on the basis of GR that the universe is expanding all the time.
The expansion, however, is by some believed to be on a cosmic
scale and does not apply to distances in the solar system or at
an atomic level. It has been likened with a balloon with coins
glued to it. If the balloon is inflated, the distances between the
coins increase, but the coins themselves are not changed (See
GRAVITATION by Misner, Thorn, and Wheeler (1973) p. 719)
[8]. In this model, however, all distances are increasing as the
universe expands. Without this expansion, however small and
undetectable on a local scale, there would be no gravitation, and
hence the expansion of the universe is not caused by gravitation,
but is the first principle from which gravitation has got to be
deduced. The most direct proof of the energy loss can be seen
in the background radiation discovered by A. A. Penzias and R.
78 Basic Ideas

W. Wilson in 1964. When the temperature in the Big Bang had


cooled down to about 3000◦ K, the space got transparent to the
pure temperature radiation, and it is this radiation we now can
observe as a background radiation with a temperature of about
2.7◦ K. So the cooling has got to be caused by the expansion
of the space itself because the photons are irrevocably loosing
energy as they move through the expanding space. In this model
the red shift is not caused by the Doppler effect alone, but rather
partly on the cooling effect of the expanding universe.
Chapter 2

The Linear Theory of


Elasticity

The Linear Theory of Elasticity is a discipline in its own right,


and this paper is only meant as an introduction to the topic.
The theory was probably originally intended to describe ordi-
nary elastic bodies consisting of particles bound together by
molecular forces, but it is through the centuries refined to be a
theory that can describe deformations in a true elastic contin-
uum. Here, I will focus on the part of the theory that describes
deformations of an elastic continuum of infinite extension or
nearly so. In its undeformed state it is supposed to be homoge-
neous and isotropic if anything else isn’t stated explicitly. For
80 The Linear Theory of Elasticity

a more thorough investigation I refer to [3]. Only a couple of


the equations, namely (3.1), (3.10), and (2.18) are used in later
developments so if those equations are familiar, this section may
be skipped.

2.1 Displacement fields


The space B under consideration, also called a Kelvin space, is
all filled up with an elastic continuum which in its undeformed
state is homogeneous and isotropic with mass density ρs that
obeys the deformation laws of The Linear Theory of Elasticity.
Notice that I already now put the index s on the mass density
in order to distinguish it from the charge density ρ of electro-
dynamics.
The displacement in this space is described by the displacement
field ; its value u(x) at a point x is the infinitesimal displacement
of x. The symmetric part
ǫ = 12 ∇u + ∇uT ,

(2.1)
1
ǫij = 2 (u i, j + uj, i ),
of the displacement gradient, ∇u, is the infinitesimal strain field,
and the above equation, relating ǫ to u, is called the strain-
displacement relation. In this context the (local) space is to
be understood as the whole of the deformed area, or at least
an area through which border no significant forces due to the
inside deformation are conveyed. In addition u has got to be
continuous and sufficiently smooth.
Displacement fields 81

We call

div u = tr ǫ,

the dilatation. The infinitesimal volume change δv(P ) of a part


P of space due to a continuous displacement of the field u is
defined by
I
δv (P ) = u · n da,
P

where n is the unit vector normal to the surface element da of


the surface of P , and we say that u is solenoidal if δv(P ) = 0
for every P . By the divergence theorem we have
Z Z
δv (P ) = div u dv = tr ǫ dv.
P P

Thus u is solenoidal if and only if

tr ǫ ≡ 0, or equivalently, div u ≡ 0,

all over space. Then there exist a vector field Ψ such that

u = curl Ψ.

A deformation field is said to be irrotational if it satisfies the


condition

curl u ≡ 0,
82 The Linear Theory of Elasticity

all over the field. Then there exist a scalar field φ in space such
that

u = ∇φ.

Let u be a vector field where [u]∞ = 0, then Helmholtz’s theorem


states that there exist a smooth scalar field φ and a vector field
Ψ on B such that

u = ∇φ + curl Ψ, where div Ψ = 0,

(see e.g. http://mathworld.wolfram.com/HelmholtzsTheorem.html).


In plain words this theorem states that an arbitrary deforma-
tion field, u, can be decomposed into two fields; an irrotational
field u1 = ∇φ and an solenoidal field u2 = curl Ψ, such that

u1 = grad φ, curl u1 = 0
u = u1 + u2 , where
u2 = curl Ψ, div u2 = 0.
(2.2)

which implies that the superposition of u1 and u2 gives a com-


plete description of any local deformation field in the spatial
continuum.

2.2 System of forces


In an elastic continuum there may be a system of forces acting
on a part, S, of space basically consisting of a surface force sn
System of forces 83

and a body force b. By the same right as u can be divided in an


irrotational and an solenoidal component, so can also the body
force b, making b = b1 + b2 with b1 and b2 belonging to the
irrotational and solenoidal field respectively.

b1 = grad ϕ
b = b1 + b2 (2.3)
b2 = curl A, div A = 0.

Since ϕ initially can be set to any level, it might as well be


associated with a possible uniform pressure in the continuum,
so an initial uniform pressure will not alter the equations in the
least.
We assume that there all over space is a strictly positive function
ρs called the density such that the mass of any part P of space
is given by
Z
ρs dv
P

The motion of the body is described by the (infinitesimal) dis-


placement field u(x, t) such that
∂u ∂2u
u̇ = and ü =
∂t ∂2t
are the velocity and acceleration respectively. The linear mo-
mentum l of P is
Z
l(P ) = ρs u̇ dv,
P
84 The Linear Theory of Elasticity

and the body counterforce b′ caused by acceleration is


Z
b′ (P ) = −l̇(P ) = − ρs ü dv.
P

In addition to this initial body force, I will keep the possibil-


ity open that there might be another hypothetical body force b
caused by the external world, just in order to see how such a
force would change the spatial continuum. The total force f (P )
on a part P of space is the total surface force from the stress
vector sn exerted across the surface ∂P plus the total body force
exerted on P by the external world
Z Z
f (P ) = sn da + b dv.
∂P P

The Cauchy-Poisson theorem [3, page 44] states that if u is an


admissible motion and f is a system of forces, then [u, f ] is a
dynamic process if and only if the following two conditions are
satisfied:

1. there exists a symmetric tensor field σ called the stress


field, such that for each unit vector n,
σn = σ n;

2. u, σ, and b satisfy the equation of motion


div σ + b = ρs ü. (2.4)

This theorem is one of the major results of continuum mechan-


ics.
The stress-strain relation 85

2.3 The stress-strain relation

In a linearly elastic continuum there exists a relation between


strain and the stress it causes, which can be expressed by the
relation
 
σ(x) = C ǫ(x) ,

where C is a fourth-order symmetric tensor that maps the space


of strain onto the space of stress according to

σij = Cijkl ǫkl .

C is called the elasticity tensor, and since the continuum un-


der consideration is assumed to be homogeneous over the actual
space, the 36 components of C is independent of the position
vector x. Since ǫ is the symmetric part of the deformation gra-
dient, ∇u, it may look like this relation rules out the possibility
that there might be a residual pressure in the continuum, but
that is not so if ǫ is interpreted as the actual stress minus the
residual stress, provided that we interpret the corresponding
surface trajectory accordingly [1, footnote 1 on page 68].
The spatial continuum under consideration is not only homo-
geneous, but also isotropic. This property immediately reduces
the 36 components of C such that C may be described by only
two different scalar constants. The stress-strain relation in a
homogeneous and isotropic continuum thus takes the relatively
86 The Linear Theory of Elasticity

simple form
σ = 2µs ǫ + λs (trǫ)I, (2.5)
σij = 2µs ǫij + λs ǫkk δij
= µs (u i, j + uj, i ) + λs uk, k δij ,
where µs and λs are Lamé’s elastic moduli 1 , which are constants
in a homogeneous elastic continuum, and δij is the Kronecker
delta

1 if i = j ,
δij =
0 if i 6= j .

2.4 The Navier-Cauchy equation


From the strain field (2.1), the Stress-strain relation (2.5) and
the Equation of motion (2.4) one can derive the Navier-Cauchy
equation [3, page 213]
µs ui, jj + µs uj,ij + (λs uk,k δij ), j + bi = ρs üi
µs ui, jj + µs uj, ji + λs uk,ki + bi = ρs üi ,
µs ∇2 u + (λs + µs )∇divu + b = ρs ü, (2.6)
or equivalently by the mathematical identity curl curl u =
∇ div u − ∇2 u
(λs + 2µs )∇divu − µs curl curl u + b = ρs ü. (2.7)
1 Note that I have put on the indices s to avoid mixing them up with

other properties in electrodynamics.


The Navier-Cauchy equation 87

At this point it may be appropriate to stress the point that the


Navier-Cauchy equation only treats the limit where deforma-
tions can be considered infinitesimal, and it must not be mixed
up with Navier-Stokes equation, which also incorporates viscos-
ity and takes into account the hydrodynamic property that v̇
may be different from ∂v/∂t [i.e. v̇ = ∂v/∂t + (v · ∇)v].
According to Helmholtz’s Theorem any vector field satisfying

[∇ · v]∞ = 0,
[∇ × v]∞ = 0,

(no velocities at infinite distance from considered area) may be


written as the sum of an irrotational part and a solenoidal part,

v = −∇φ + ∇ × A,

where
∇·v
Z
φ=− d3 r′ ,
V 4π|r − r|

∇×v 3 ′
Z
A= d r,
V 4π|r − r|

(see http://mathworld.wolfram.com/HelmholtzsTheorem.html).
As the spatial continuum is of infinite extension, or nearly so,
any deformations have to be confined to a finite part of space,
so this theorem will be applicable on all deformations. Hence
the displacement field can be decomposed into two properties

u = u1 + u2 ,
88 The Linear Theory of Elasticity

where
u1 = −∇φ = − grad φ,
u2 = ∇ × Ψ = curl Ψ.
Since curl grad φ ≡ 0, and div curl Ψ ≡ 0, the Navier-Cauchy
equation (3.1) can be divided into two independent equations,
one for an irrotational field
ρs b1
∇divu1 = ü1 − , (2.8)
(λs + 2µs ) λs + 2µs
and the other for a solenoidal field
ρs b2
−curl curl u2 = ü2 − . (2.9)
µs µs

By defining two new constants


s
λs + 2µs µs
r
c1 = , c2 = , (2.10)
ρs ρs

the N-C equation takes the form


b
c12 ∇divu − c22 curl curl u + = ü. (2.11)
ρs

Operating on Equation (3.1) with the div operator and on Equa-


tion (3.2) with the curl operator yields respectively
1 ∂ 2 (div u) div b
∇2 (div u) − 2 2
=− (2.12)
c1 ∂t λs + 2µs
Field energy and energy transport 89

1 ∂ 2 (curl u) curl b
∇2 (curl u) − 2 2
=− (2.13)
c2 ∂t µs

With b = 0 we have two wave equations where the dilatation,


divu, satisfies a wave moving with the speed c1 , while the rota-
tional component curlu, satisfies a wave moving with the speed
c2 . In fact the Propagation theorem for isotropic bodies states
that if a body is isotropic, then a wave is either longitudinal, in
which case c = c1 , or transversal, in which case c = c2 [3, page
256]. This diversion of the Navier-Cauchy equation into one ir-
rotational and one solenoidal part, allows us to examine these
two parts separately and thereby simplifies the strain-stress re-
lation immensely by reducing the elastic constants to only one
single constant (the wave speed) in each equation (c1 6= c2 ).
We see from Equation (3.5) that the two wave speeds are re-
lated pto each other with a fixed constant given by the relation
c1 = 2 + λs /µs · c2 . We notice that c1 is about the double of
c2 .
Finally we notice that all information of curlu is lost in (3.7)
and all information of divu in (3.8).

2.5 Field energy and energy transport


From the Navier-Cauchy equation one can find the internal field
energy in an admissible field in B by performing the following
thought experiment: Introduce a hypothetical body force, −b
(negative because b is a breaking force), from the outside world
90 The Linear Theory of Elasticity

such that it eradicates the entire field in B; i.e. u and all func-
tions of u become constant like zero all over B. In addition I
will assume that the entire field is confined inside B such that
u is zero on the surface of B and beyond. The energy released
by this operation, E, would then be like the total field energy
in B.

Z Z0
E=− dv bdu
B f (u)

Z f (u)
h Z  i
= dv ρs ü − (λs + 2µ)grad divu + µcurl curlu du
B 0

E can be separated into three integrals, i.e. E = E1 + E2 + E3 .


The first of these integrals is simply the kinetic energy of the
system

Zu̇
du̇ 
Z 
E1 = dv ρs du
dt
B 0
Z  Zu̇ 
= dv ρs du̇ · u̇ ,
B 0
Z
1 2
E1 = 2 ρs u̇ dv.
B
Field energy and energy transport 91

The next part can be integrated by using the mathematical iden-


tity (2.17) and inserting φ = div u and A = du

Z div
Z u h i
E2 = (λs + 2µ) dv div u · div(du) − div(du · div u)
B 0
Z div
Z u
 
= (λs + 2µ) dv div u · d(divu) −
B 0
Z div
Z u

(λs + 2µ) dv · div du · divu .
B 0

The first part of the integral can readily be integrated, and the
last part can be transformed into a surface integral over ∂B by
the Divergence theorem2 and disappear because u is constant
like zero on the border of B and beyond. Thus
Z
1 2
E2 = 2 (λs + 2µ)(div u) dv.
B

We can find E3 in much the same way by using the identity

div (A × B) = curl A · B − curl B · A, (2.14)


2
H R
(A · n) df = div A dv
∂B B
92 The Linear Theory of Elasticity

and inserting B = curl u and A = du


Z curl
Z u h i
E3 = −µ dv div(du × curl u) − curl u · curl(du)
B 0
Z curl
Z u
=µ dv curl u · d(curl u)−
B 0
Z  curl
Z u 
µ dv · div du × curl u .
B 0

Again the first part can be integrated and the last part disappear
by the same reason as above, and we get
Z
1 2
E3 = 2 µ(curl u) dv.
B

Finally we can write the total energy in the deformed area


Z h i
1 2 1 2 1 2
E= 2 ρ s u̇ + 2 (λs + 2µ)(div u) + 2 µ(curl u) dv.
B
(2.15)

The development may be a bit unorthodox, but the result is


already known as Kelvin’s theorem [3, page 208], and is a proven
theorem in the Linear Theory of Elasticity. The result can be
interpreted as the local energy density even if this development
Field energy and energy transport 93

does not prove where in the field the energy is to be found; only
that there to a curl u and a div u always corresponds an energy
given by the equation above, and no other energy is present as
long as we deal with infinitesimal deformations restricted to a
limited area of a homogeneous and isotropic continuum covered
by the Linear Theory of Elasticity.3 With this restriction in
mind the local energy density, e, in the spatial continuum is
given by

e= 1
2 ρs u̇2 + 1
2 (λs + 2µs )( div u)2 + 1
2 µs ( curl u)2 (2.16)

leaving the possibility open that there may be a residual pressure


and a corresponding homogeneous residual energy density in
addition to this field energy. It is noteworthy that the energy
density in any field of strain and motion is nonnegative even
if the space itself should happen to contain a huge amount of
uniformly distributed energy due to an initial pressure.

The energy transport in the deformation field can be found by


deriving the equation above with respect on time. We acquire

∂e
= ρs u̇ü + (λs + 2µs ) div u div u̇ + µs curl u curl u̇.
∂t

3 The corresponding expression for the energy density in an electromag-

netic field has the same limitation, but nonetheless it is usually interpreted
as the local energy density.
94 The Linear Theory of Elasticity

We substitute ρs ü from Equation (3.1) and get

∂e  
=u̇ (λs + 2µs ) grad div u − µs curl curl u + b
∂t
+ (λs + 2µs ) div u div u̇ + µs curl u curl u̇,
∂e  
bu̇ = + µs curl ( curl u) · u̇ − curl u̇( curl u)
∂t  
− (λs + 2µs ) ( div u) div u̇ + u̇ grad ( div u) .

By the mathematical identity (2.14) and the identity

div (φA) = φ div A + A grad φ, (2.17)

this equation develops into


∂e  
+ div (µs curl u × u̇) − div (λs + 2µs ) div u · u̇ = bu̇.
∂t
We define a new vector

S = µs curl u × u̇ − (λs + 2µs ) div u · u̇, (2.18)

and in the absence of external forces we acquire the compact


equation:
∂e
+ div S = 0.
∂t
Since the increase in energy density has got to be equal to the
inflow of energy per unit volume, S can be interpreted as the
energy flow vector.
Solenoidal deformations and Electrodynamics 95

2.6 Solenoidal deformations and Elec-


trodynamics

In this section I will redefine some of the terms used in elasto-


dynamics to terms that can be directly compared to those in
electrodynamics. I will stress that these redefinitions only are
intended to make the comparison simpler and will not change
the physics behind the original terms in any way. With the ad-
ditional assumption that there might be true and in the elastic
continuum, I will also show that they will influence the elas-
todynamic fields in exactly the same way as electric charges
influence the electromagnetic fields. How sinks and sources can
be more than pure mathematical entities will be discussed in
another paper4 . Some terms are used quite differently in me-
chanics and electrodynamics. For example the Greek letter ρ
is used for mass density in mechanics, but as charge density in
electrodynamics. To avoid confusion I will use an index s on the
mechanical terms whenever necessary. Hence ρs means spatial
mass density while ρ means the density of sinks – the spatial
counterpart to charge density.

4 See a proposition on how it might work in


http://www.scribd.com/doc/3014850/The-Great-Puzzle.
96 The Linear Theory of Elasticity

2.7 Reformulation of the Navier-


Cauchy’s Equation
First we define some new properties
def 1
µ0 = , [L2 F −1 ] = [LT 2 M −1 ], (2.19)
µs
def
ε 0 = ρs , [F T 2 L−4 ] = [M L−3 ], (2.20)
def ∂u2
E = − , [LT −1 ], (2.21)
∂t
def
B = curl u2 , [ ], (2.22)
def
j = b, [F L−3 ] = [M L−2 T −2 ], (2.23)
1
c2 = , [L2 T −2 ]. (2.24)
ε0 µ0

Note that in this section all the defined properties refer exclu-
sively to elastic properties. The notations inside the square
brackets are the dimensions of the properties in front, but in
this context I have found it convenient to alternatively replace
mass with force as a fundamental unit. Hence the mass unit is
converted to the force unit by the relation
[F ] = [M LT −2].

By the identities curl (∂(·)/∂t) = ∂ curl (·)/∂t and div curl (·) =
0 we immediately get the relation between B and E
curl E + Ḃ = 0, (2.25)
Reformulation of the Navier-Cauchy’s Equation 97

and

div B = 0. (2.26)

Navier’s Equation (3.4) takes the form

1
curl B − Ė = µ0 j. (2.27)
c2

Now I will make the assumption that there may be real sinks
and sources in the spatial continuum. How this is possible will
be discussed elsewhere, but here I take entities like that for
granted. I will take sinks as positive entities and sources as
negative sinks, and assume that they can only be created by
pair production; one sink for for one equally strong source. If
there are more sinks than sources in an area, the sink density
is positive, and if there are more sources than sinks, then the
sink density is negative. The strength of a spatial sink, Qs , can
be defined as the inflow of spatial mass through a closed surface
around the sink
I
def
Q = −ρs u̇ndf, [F T L−1] = [M T −1 ],
V

where n is an outwards pointing unit vector normal to df .

Now let sinks and sources with strength Q1 , Q2 , Q3 , · · · , Qn be


sufficiently smoothly distributed in space. Then the density of
98 The Linear Theory of Elasticity

sinks is given by
m
def 1 X
ρ = lim Qn
V →ǫ V
n=1
1
I
= −ρs lim u̇ndf
V →ǫ V V
= −ρs div u̇,

or with the new terms incerted

ρ = ε0 div E, [F T L−4 ] = [M L−3 T −1 ],


(2.28)

where m is the number of sinks in a volume V of space, and ǫ


is a small volume, but still great enough to contain many sinks.
(Note the difference between the spatial mass density ρs and the
density of sinks ρ.)
There is a hidden dependency between the sink density ρ and
and the volume force j. By taking the divergence of Equation
(2.27):

 Ė 
div curl B − div = div µ0 j,
c2
1
− div Ė = div j, (I)
ε0

and the partial derivative with respect on time of Equation


The stress energy tensor 99

(2.28):

∂ 1 ∂ρ
div E =
∂t ε0 ∂t
1 ∂ρ
div Ė = , (II)
ε0 ∂t

and evaluating the combination I + II, we acquire

ρ̇ + div j = 0. (2.29)

Since sinks and sources by definition can only be created or


disappear in pairs, the only way the density can change in a
volume is by out- or inflow, hence the vector j can be interpreted
as a flow of sinks or sources, i.e. a flow of sinks or sources will
create a force field (a drag) in the spatial continuum.

2.8 The stress energy tensor


According to (3.10) and the newly defined properties the elas-
todynamic field energy in a divergence-free field is

ε0 2 1 2
e= E + B , [F L−2 ] = [M L−1 T −2 ]. (2.30)
2 2µ0

Since this field may contain energy, we must also expect that it
can move around in space as the field changes. To examine this
property we can start by deriving the field energy (2.30) with
100 The Linear Theory of Elasticity

respect on time and get


∂e 1
= ε0 E · Ė + B · Ḃ.
∂t µ0
From this expression we can eliminate the time derivatives of E
and B by applying (2.27) and (2.25)

∂e  1  1
= ε0 E · c2 curl B − j − B · curl E,
∂t ε0 µ0
∂e 1 
= curl B · E − curl E · B − j · E,
∂t µ0

and further by the mathematical identity (2.14) it develops into

∂e 1  
+ div E × B = −j · E, {= u̇ · b}. (2.31)
∂t µ0
The right side of this equation is the rate of work done by exter-
nal forces per unit volume on the continuum, and the left side
can be interpreted as the rate of increase in energy density plus
the rate at which the energy is leaving per unit volume. Thus
the energy flow vector is

def 1
S = (E × B), [F L−1 T −1 ] = [M T −3 ]. (2.32)
µ0

With this property inserted, Equation (2.31) takes the form

1 ∂e 1 1
+ div S = − j · E. (2.33)
c ∂t c c
The stress energy tensor 101

To examine the forces involved by a change of momentum, we


can derive the momentum vector with respect on t and again
eliminate the time derivatives of E and B by applying (2.27)
and (2.25)
Ṡ 1 
2
= 2 E × Ḃ + Ė × B ,
c c µ0

= ε0 E × (− curl E) + (c2 curl B − c2 µ0 j) × B ,
 
c 2

Ṡ 1
2
= −ε0 E × curl E + curl B × B + B × j.
c µ0
By applying the mathematical identity
grad (A · A) = 2[A × curl A + (A · ∇)A], (2.34)
we obtain
Ṡ ε0  1 
2
+ grad E · E − ε0 (E · ∇)E + grad B·B
c 2 2µ0
1
− (B · ∇)B = (B × j).
µ0
(2.35)
We then write out the above equation in component form5 :
Ṡi ε0 1 1
+( E·E+ B · B),i − ε0 Ej Ei, j − Bj Bi, j
c2 2 2µ0 µ0
= ǫijk Bj jk ,
5 Note that Latin indices go from 1 to 3 while Greek indices go from 0

to 3.
102 The Linear Theory of Elasticity

and expand it by adding ±ε0 Ej,j Ei and ± µ10 Bj,j Bi

Ṡi ε0 2 1 2
+ E + B , i − ε0 Ej Ei, j − ε0 Ej, j Ei + ε0 Ej, j Ei
c2 2 2µ0
1 1 1
− Bj Bi, j − Bj, j Bi + Bj, j Bi = ǫijk Bj jk .
µ0 µ0 µ0

The term Bj, j is like zero by (2.26), Ej,j = ρ/ε0 by Equation


(2.28), and the rest can be manipulated into

∂Si
− + σij,j = ρEi − ǫijk Bj jk . (2.36)
c2 ∂t
where the new tensor σij is given by

def 1 1 1 2
σij = ε0 Ei Ej + Bi Bj − ε 0 E2 + B δij . (2.37)
µ0 2 µ0

Now we write out (2.33) and (2.36) in component form and


obtain the set of equations (the zeroes are inserted for clarity)

∂e ∂Sx ∂Sy ∂Sz Ex jx Ey jy Ez jz


+ + + =0 − − − ,
c∂t c∂x c∂y c∂z c c c
∂Sx ∂σxx ∂σxy ∂σxz
− − − = − Ex ρ + 0 − Bz jy + By jz ,
c2 ∂t ∂x ∂y ∂z
∂Sy ∂σyx ∂σyy ∂σyz
− − − = − Ey ρ + Bz jx + 0 − Bx jz ,
c2 ∂t ∂x ∂y ∂z
∂Sz ∂σzx ∂σxy ∂σzz
− − − = − Ez ρ − By jx + Bx jy + 0.
c2 ∂t ∂x ∂y ∂z
The stress energy tensor 103

The four differential equations can be written as one matrix


equation (note that BT · AT = A · B)
 ∂   
c∂t e Sx /c Sy /c Sz /c
 ∂   Sx /c −σxx −σxy −σxz 
∂x ·

 
   Sy /c −σyx −σyy −σyz 
∂y
∂ Sz /c −σzx −σzy −σzz
∂z
 
0 Ex /c Ey /c Ez /c
 −Ex /c 0 Bz −By 
= −Ey /c −Bz
 · [−cρ jx jy jz ]
0 Bx 
−Ez /c By −Bx 0
(2.38)

which formally can be written

Tαβ ,β = Fαβ Jβ ,

or in frame independent notation

∇ · T = F · J.

Here the second order tensor T, the stress energy tensor, is given
by
 
e Sx /c Sy /c Sz /c
def  S /c −σxx −σxy −σxz 
Tαβ =  x

, (2.39)
Sy /c −σyx −σyy −σyz 
Sz /c −σzx −σzy −σzz
104 The Linear Theory of Elasticity

the second order tensor F by


 
0 Ex /c Ey /c Ez /c
def −E x /c 0 Bz −By
Fαβ = 
 
, (2.40)
 −Ey /c −Bz 0 Bx 
−Ez /c By −Bx 0
and finally J by
Jα = (−cρ, jx , jy , jz ). (2.41)

In four-space we need some definitions. First the Minkowski


metric:
 
−1 0 0 0
0 1 0 0 
ηαβ = η αβ = 

.
 0 0 1 0 
0 0 0 1
We can use the metric tensor to raise or lower the indices. For
example
Jβ = Jα η αβ ,

hence

Jα = (cρ, jx , jy , jz ).
Coordinates in 4-space:
xα =(x0 , x1 , x2 , x3 ) = (ct, x, y, z),
xβ =xα ηαβ = (−ct, x, y, z),
The stress energy tensor 105

and the orthogonal base vectors are

{e0 , e1 , e2 , e3 }.

The del operator in four-space:

∂ 1 ∂ ∂ ∂ ∂ 
∇α =, α = ∂α = α
= , , , ,
∂x c ∂t ∂x ∂y ∂z
∂  1 ∂ ∂ ∂ ∂ 
∇α =, α = ∂ α = = − , , , ,
∂xα c ∂t ∂x ∂y ∂z
∂2  1 ∂2 ∂2 ∂2 ∂2 
∇2 =, α, α = ∂ α ∂α = = − , , , .
∂xα ∂xα c2 ∂t2 ∂x2 ∂y 2 ∂z 2

From the above definitions we get that the 4-spatial distance


between two points (events) A and B in four-space is given by

∆s = −c(t2 − t1 )e0 + (x2 − x1 )e1 + (y2 − y1 )e2 + (z2 − z1 )e3 ,

and

(∆s)2 = (∆sα ) · (∆sα ) = −c2 (∆t)2 + (∆x)2 + (∆y)2 + (∆z)2 .

By this extremely short introduction I have transformed some


elastodynamic formulas into the formalism of flat Minkowski’s
space. The advantage of this formalism is that all formulas
remain invariant by transformations between different Lorenz
frames.
106 The Linear Theory of Elasticity

2.9 The vector potential in the elastic


continuum
Equation (2.26) means that the field B can be derived from
some vector potential A

B = curl A, (2.42)

where div A is temporarily arbitrary, but can be given a fixed


meaning later without changing the term curl A.
By inserting (2.42) into (2.25) we get

curl (E + A,t ) = 0.

Therefore E + A,t may be represented as some gradient

E + A,t = −c grad φ,

hence

E = −(c grad φ + A,t ). (2.43)

Thus both E and B can be represented by some potentials A


and φ. For the choice of A and φ the Equations (2.25) and
(2.26) are fulfilled.
By adding and subtracting the same term c grad ψ,t into (2.43),
we acquire

E = −[c grad (φ − ψ,t ) + (A + c grad ψ),t ].


The vector potential in the elastic continuum 107

We also have that adding c grad ψ to A leaves B unchanged.


Hence the substitutions
φ → φ′ = φ + ψ,t , A → A′ = A + c grad ψ (2.44)
leave the properties E, B, j, and ρ unchanged for ar-
bitrary functions ψ. The substitutions (2.44) are called
Gauge transformations (see http://www.mathematik.tu-
darmstadt.de/ bruhn/Maxwell-Theory.html).
The most obvious gauge is to set div A = div u which means to
infer that the spatial continuum is uncompressed. It would work
equally well to set div u = const. This picture is complicated by
the assumption that there are true point-like sinks and sources
around, hence − div u̇ = ρ/ε0 (see Equation (2.28)), so we can
introduce a potential φ such that
ρ
−∇2 φ = .
ε0
This leads to the Coulomb gauge which works well if we consider
a fixed frame in the spatial continuum. What we need, however,
is a gauge that works equally well in a moving frame. This
requirement leads to the Lorenz gauge after the Danish physicist
Ludvig Valentin Lorenz (1829-1891):
1
div A + φ, t = 0. (2.45)
c
Inserting (2.42) and (2.43) into (2.27) yields
1
curl curl A + (A,tt + c grad φ,t ) = µ0 j,
c2
108 The Linear Theory of Elasticity

and by applying the mathematical identity

curl curl A = grad div A − ∇2 A, (2.46)

we obtain
1 1
A,tt − ∇2 A + grad ( div A + φ, t ) = µ0 j.
c2 c
Analogously by inserting the same properties into (2.28) we ob-
tain
ρ
−(c∇2 φ + div A,t ) = ,
ε0
or by adding and subtracting 1/cφ,tt we acquire
1 1 ρ
φ,tt − c∇2 φ − ( div A + φ, t ),t = ,
c c ε0
1 1 1 cρ
φ,tt − ∇2 φ − ( div A + φ, t ),t = µ0 · 2 .
c2 c c c ε0 µ0
By the Lorenz gauge and (2.24) the two potentials reduce to
1
− φ,tt + ∇2 φ = −µ0 · cρ, (2.47)
c2
1
− 2 A,tt + ∇2 A = −µ0 · j. (2.48)
c

These two equations can be expressed as one vector potential in


four-space

∂ν ∂ ν Aα = −µ0 · Jα ,
Energy flow and momentum 109

where

Aα = (φ, Ax , Ay , Az ),
Jα = (cρ, jx , jy , jz ),

or in frame independent notation

∇2 A = −µ0 · J. (2.49)

2.10 Energy flow and momentum


Think of disturbance energy as a substance that flows through
space with velocity c. Then the energy flow vector alternatively
can be expressed as

S = e · c,

and the energy density as

|S|
e= .
c
Next if we only consider disturbance energy without the pres-
ence of any sinks or sources, Equation (2.36) reduces to

∂Si
= σij,j ,
c2 ∂t
∂ S
= ∇ · σ.
∂t c2
110 The Linear Theory of Elasticity

Now, if σ is a stress tensor, then the divergence of it represents


a force, hence

∂ S
= f,
∂t c2
and we can define a new vector
def
p = S/c2 , (2.50)

to obtain
∂p
= f.
∂t

Let us imagine some elastodynamic radiation trapped inside an


imaginary box with reflecting walls, and let the box be subdi-
vided into m small cells containing small parts of the radiation
energy En . At a given time the sum of the energy flow is given
by
m
X
E·v = Sn , (2.51)
n=1

hence the box is moving with some velocity v in the direction


of S and it contains an amount of energy given by
m
X
E= En .
n=1
Energy flow and momentum 111

We divide Equation (2.51) by c2 and take the time derivative of


it. We obtain
m
∂  Ev  X ∂  Sn 
= ,
∂t c2 n=1
∂t c2
m
∂  Ev  X
= fn ,
∂t c2 n=1
∂  Ev 
= f.
∂t c2
Finally we define a new property m given by

def E
m = ,
c2
or

E = mc2 , (2.52)

and acquire


(mv) = f . (2.53)
∂t

We can interpret this equation such that if we have a box con-


taining a disturbance energy similar to m, then a force f is
needed to give it an acceleration a = v̇, provided that the prop-
erty m, which we could call the mass of radiation, is kept con-
stant.
112 The Linear Theory of Elasticity

When the box is accelerated from zero velocity6 , however, we


have got to add energy to it given by

dE = f · ds,
= f ds,

if ds is in the direction of f . By the equations above we can


develop this equation further into

d(mv)
dE = ds
dt
d E 

= v ds
dt c2
1 
= 2 v · dE + E · dv dv,
c
dE 1 v · dv
= 2 ,
E c 1 − v 2 /c2

6 The situation is considerably more complicated if the box and the ob-

server have an initial velocity, say v0 . To address that question, one first
has got to assume that the phenomenon is observed in a Lorentz frame
that makes the equations above invariant for the change of the observer’s
coordinate system, as Lorentz showed already in the fall of the nineteenth
century. That would make the observation fully relativistic, and v0 could
be set to zero from where the deduction could proceed as shown.
Summing up 113

which can be solved


1
ln E = ln p + ln C
1 − v 2 /c2
C
= ln p ,
1 − v 2 /c2
E0
E= p ,
1 − v 2 /c2
m0
m= p . (2.54)
1 − v 2 /c2
Equations (2.53) and (2.54) bring the dynamics of confined dis-
turbance energy in line with Newton’s second law of motion and
the relativistic mass increases with velocity. Equation (2.52) is
of course like the famous Einstein energy/mass relation.

2.11 Summing up
In this paper we have seen that the four equations (2.25) through
(2.28) correspond to James Clerk Maxwell’s (1831–1879) elec-
trodynamic equations. Provided that there are free moving sinks
and sources in the spatial continuum, Equation (2.29) demon-
strates that they will generate a ”drag” just like Lord Kelvin
postulated for moving electrons in 1890 [10, page 247]. The en-
ergy flow vector in Equation (2.32) is formally like Poynting’s
vector after John Henry Poynting (1852-1914). In a notation
introduced by Hermann Minkowski (1864–1909), the field ten-
sor Fαβ in Equation (2.40) is like the Electromagnetic tensor,
114 The Linear Theory of Elasticity

and the Elastodynamic stress-energy tensor, Tαβ , corresponds


exactly to the Electromagnetic stress-energy tensor. Note also
that the spatial stresses, σxy , correspond exactly to Maxwell’s
stress tensor that represent the mechanical stresses caused by
electromagnetic fields in space. Finally it is possible to describe
deformation fields as a vector potential in the spatial contin-
uum. In this notation the fields, like electromagnetic fields, are
invariant by transformations between different Lorentz frames,
after Hendrik Antoon Lorentz (1853–1928), in rectilinear mo-
tion relative to each other. A moving weightless box containing
an amount of disturbance energy will have a momentum corre-
sponding to the energy in a material body with the same energy
content. The force needed to change its velocity corresponds
to Newton’s second Law of motion, and moreover, to increase
the velocity of such a box towards the propagating speed c of
transversal waves will increase the trapped energy towards in-
finity.
Chapter 3

Standing Waves
between Singularities

Waves in an elastic continuum are mainly of two different types,


longitudinal and transversal waves, which are well documented
in various textbooks. In this paper I will look into the possibility
that there might be a third alternative, namely standing waves
between oscillating nodes in the form of singularities in an elastic
continuum of infinite extension. First I will show that a standing
wave can form between a hypothetical rigid sphere embedded
in the spatial continuum and the center node, thus establishing
that a singularity may form the one endpoint in such oscillations.
Next I will show that if the spatial continuum initially is agitated
116 Standing Waves between Singularities

to an extent that it contains a plethora of such oscillating nodes,


they will tend to organize along endless strings. Finally I will
try to look into the possibility that there might be a coupling
between the two field components in question, irrotational and
solenoidal fields.

3.1 The Navier-Cauchy Equation


In order to find how scalar waves propagate in the spatial con-
tinuum we start with recalling the Navier-Cauchy equation

(λs + 2µs )∇ div u − µs curl curl u + b = ρs ü. (3.1)

or equivalently by the mathematical identity curl curl u =


∇ div u − ∇2 u

µs ∇2 u + (λs + µs )∇ div u + b = ρs ü, (3.2)

At this point it may be appropriate to stress the point that the


Navier-Cauchy equation only treats the limit where deforma-
tions can be considered infinitesimal, and it must not be mixed
up with Navier-Stokes equation, which also incorporates viscos-
ity and takes into account the hydrodynamic property that v̇
may be different from ∂v/∂t [i.e. v̇ = ∂v/∂t + (v · ∇)v].
According to Helmholtz’s Theorem any vector field satisfying

[∇ · v]∞ = 0,
[∇ × v]∞ = 0,
The Navier-Cauchy Equation 117

(no velocities at infinite distance from considered area) may be


written as the sum of an irrotational part and a solenoidal part,

v = −∇φ + ∇ × A,

where
∇·v
Z
φ=− d3 r′ ,
V 4π|r′ − r|

∇×v 3 ′
Z
A= d r,
V 4π|r′ − r|

(see http://mathworld.wolfram.com/HelmholtzsTheorem.html).
As the spatial continuum is of infinite extension, or nearly so,
any deformations have to be confined to a finite part of space,
so this theorem will be applicable on all deformations. Hence
the displacement field can be decomposed into two properties

u = u1 + u2 ,

where

u1 = −∇φ = − grad φ,
u2 = ∇ × Ψ = curl Ψ, div Ψ = 0.

Since curl grad φ ≡ 0, and div curl Ψ ≡ 0, the Navier-Cauchy


equation (3.1) can be divided into two independent equations,
one for an irrotational field
ρs b1
∇ div u1 = ü1 − , (3.3)
(λs + 2µs ) λs + 2µs
118 Standing Waves between Singularities

and the other for a solenoidal field


ρs b2
− curl curl u2 = ü2 − . (3.4)
µs µs

By defining two new constants


s
λs + 2µs µs
r
c1 = , c2 = , (3.5)
ρs ρs

the N-C equation takes the form


b
c12 ∇ div u − c22 curl curl u + = ü. (3.6)
ρs

Operating on Equation (3.1) with the div operator and on Equa-


tion (3.2) with the curl operator yields respectively

1 ∂ 2 ( div u) div b
∇2 ( div u) − =− , (3.7)
c12 ∂t2 λs + 2µs

1 ∂ 2 ( curl u) curl b
∇2 ( curl u) − =− . (3.8)
c22 ∂t2 µs

With the outer force, b, set to zero we have two wave equations
where the dilatation, div u, satisfies a wave moving with the
speed c1 , while the rotational component curl u, satisfies a wave
moving with the speed c2 . In fact the Propagation theorem for
isotropic bodies states that if a body is isotropic, then a wave
The Navier-Cauchy Equation 119

is either longitudinal, in which case c = c1 , or transversal, in


which case c = c2 [3, page 256]. This splitting of the Navier-
Cauchy equation into one irrotational and one solenoidal part,
allows us to examine these two parts separately and thereby
simplifies the strain-stress relation immensely by reducing the
elastic constants to only one single constant (the wave speed) in
each equation (c1 6= c2 ). We see from Equation (3.5) that the
two wave speeds are relatedp to each other with a fixed constant
given by the relation c1 = 2 + λs /µs · c2 , where c1 might be
about the double of c2 (also dubbed c without the index in the
text to follow). Notice also that all information of curl u is lost
in Equation (3.7) and all information of div u in Equation (3.8).
The energy in a deformation field with no surface trajectories is
given by Kelvin’s theorem [3, page 208]:
Z h i
1 2 1 2 1 2
E= 2 ρ s u̇ + 2 (λs + 2µ)(div u) + 2 µ(curl u) dv.
B
(3.9)
The theorem states that to a curl u, a div u, and a velocity field
u̇ there always corresponds an energy equal to E, but it does not
tell exactly where in the field the energy is to be found. With
this restriction in mind, the local energy density, e, in a spatial
continuum of infinite extension can all the same be define as1
e= 1
2 ρs u̇2 + 1
2 (λs + 2µs )( div u)2 + 1
2 µs ( curl u)2 . (3.10)
1 The corresponding expression for the energy density in an electromag-

netic field has the same limitation, but nonetheless it is usually interpreted
as the local energy density.
120 Standing Waves between Singularities

3.2 Scalar longitudinal waves in the


spatial continuum
By setting φ = div u in Equation (3.7) and consider a field
without any outer forces, we get the wave equation
1 ∂ 2φ
∇2 φ = , (3.11)
c12 ∂t2
or the even simpler wave equation for plane, longitudinal waves
∂2φ 1 ∂2φ
= . (3.12)
∂x2 c12 ∂t2
It can be solved in several ways, but here I will apply
d’Alembert’s solution. Let
ξ ≡ c1 t − x
η ≡ c1 t + x.
Since x = f1 (ξ, η) and t = f2 (ξ, η) we have by the chain rule
∂ ∂ ∂ξ ∂ ∂η ∂ ∂
= · + · =− + ,
∂x ∂ξ ∂x ∂η ∂x ∂ξ ∂η
∂ ∂ ∂ξ ∂ ∂η ∂ ∂
= · + · = c1 + c1 .
∂t ∂ξ ∂t ∂η ∂t ∂ξ ∂η
By applying these operators on Equation (3.12) it reduces to
∂2φ
= 0.
∂ξ∂η
Scalar longitudinal waves in the spatial continuum 121

This partial differential equation can be integrated in two steps,


and the general solution is
φ = f (ξ) + g(η) or φ = f (c1 t − x) + g(c1 t + x).
The two functions f (c1 t−x) and g(c1 t+x) represent by a certain
time t = t0 a disturbance, or wave-formation, which when it first
is started, propagates with velocity c1 in the x-axis’s positive or
negative direction respectively.
In order to find how spherical waves propagate from a distur-
bance center, I will write Equation (3.11) in polar coordinates
φ = φ(r, v, w):
1 ∂  2 ∂φ  1 ∂2φ 1 ∂  ∂φ 
r + + sin w
r2 ∂r ∂r r2 sin2 w ∂v 2 r2 sin w ∂w ∂w
2
1 ∂ φ
= 2 2.
c1 ∂t
In this connection I will only consider the spherical symmetric
case so ∂φ/∂v = ∂φ/∂w = 0, and we acquire
1 ∂  2 ∂φ  1 ∂2φ
2
r = 2 2,
r ∂r ∂r c1 ∂t
which by the substitution φ = ψ/r goes over in the simpler
∂2ψ 1 ∂2ψ
2
= 2 2.
∂r c1 ∂t
This equation is formally equivalent to Equation (3.12) and has
a solution corresponding to what is the case for plane waves:
ψ = f (c1 t − r) + g(c1 t + r).
122 Standing Waves between Singularities

Hence the general expression for a spherical wave spreading from


a given wave-center is given by

f (c1 t − r)
div u = . (3.13)
r
The other solution is for a spherical wave converging towards a
focal point in space:

g(c1 t + r)
div u = . (3.14)
r

3.3 Irrotational standing waves


In this section I will investigate if standing waves can occur be-
tween hypothetical concentric rigid shells embedded in the spa-
tial continuum, especially in order to see if standing waves can
form between the inside of a spherical shell and the center node.
I perform the thought experiment that imbedded in the spatial
continuum is a hypothetical rigid, undeformable sphere inside
which there might be a standing compression wave bouncing
back and forth between the center and the firm shell. In or-
der to see if such a standing wave is possible, I will set up the
Navier-Cauchy equation and try to solve it with these border
conditions. I must, however emphasis that the N-C equation
only is valid for small deformations, in fact only when the de-
formations are infinitesimal. When the deformations are signifi-
cant, we have got to bring in the full power of the Navier-Stokes
equation. The main difference from the N-C equation is that
Irrotational standing waves 123

the time derivative has got to be modified to2

d ∂
(·) = (·) + (v · ∇)(·), (3.15)
dt ∂t
e.g. ü = ∂ u̇/∂t + (u̇∇)u̇. Hence the result may at best be
suggestive.
First we write the Navier-Cauchy equation for an irrotational
field with no external forces (Eq. 3.3):

1
grad div u = ü,
c21

where c1 is the speed of longitudinal waves, and then let us con-


sider a central symmetric deformation where u = f (r)r̂. From
the identities
f
div (f r̂) = 2 + f ′, (3.16)
r
and

grad f = f ′ r̂, (3.17)

we obtain the identity

2f ′ 2f 
grad div f = f ′′ + − 2 r̂. (3.18)
r r
2 The N-S equation a also incorporates viscosity, which is not present in

the spatial continuum.


124 Standing Waves between Singularities

With these properties inserted, the N-C equation takes the form
u′ u ü
u′′ + 2 − 2 2 r̂ = 2 r̂. (3.19)
r r c1
By the product method, u(r, t) = F (r) · G(t) · r̂, we have
2 2 1
F ′′ G + F ′ G − 2 F G = 2 F G̈,
r r c1
F ′′ 2 F′ 2 1 G̈
+ − 2 = 2 .
F r F r c1 G
Since the left and right side of this expression only depend on
the arguments r and t respectively, they can both be set to the
same constant −p2 , and hence they can be separated into the
two equations:
2 (−2 + p2 r2 )
F ′′ + F ′ + F =0 (3.20)
r r2
G̈ + c21 p2 G = 0 (3.21)

Equation (3.21) has the general solution

G(t) = C1 sin(c1 pt) − C1 cos(c1 pt). (3.22)

Equation (3.20) may be solved by the Frobenius method (see


e.g. [5, Sec. 4.4]):
Any differential equation of the form
b(x) ′ c(x)
y ′′ + y + 2 y=0
x x
Irrotational standing waves 125

where the functions b(x) and c(x) are analytic at x=0, has at
least one solution that can be represented in the form

y(x) = xr (a0 + a1 x + a2 x2 + · · · ) (3.23)

where the exponent r may be any (real or complex) number (and


r is chosen so that a0 6= 0).
The equation also has a second solution (such that these two
solutions are linearly independent) that may be similar to (3.23)
(with a different r and different coefficients) or may contain a
logarithmic term.
Temporary replacing r with x and F with y in Equation (3.20)
yields
2 ′ −2 + p2 x2
y ′′ + y + y = 0. (3.24)
x x2
This equation has at least one solution of the form

X ∞
X
y(x) = xr am xm = am rm+r .
m=0 m=0

The first and second derivative becomes



X

y (x) = am (m + r)xm+r−1 ,
m=0


X
y ′′ (x) = am (m + r)(m + r − 1)xm+r−2 .
m=0
126 Standing Waves between Singularities

We expand b(x) and c(x) in power series with coefficients

b0 = 2, b1,2,··· = 0, c0 = −2, c1 = 0, c2 = p2 , c3,4,··· = 0.

The indicial equation r(r − 1) + b0 r + c0 = 0 takes the form

r(r − 1) + 2r − 2 = 0
r1 = 1, r2 = −2. (3.25)

Here we have a case 3 situation (roots differing by an integer,


r1 − r2 = 3). We also note that r1 − r2 > 0 so we have the two
solutions

y1 (x) = xr1 (a0 + a1 x + a2 x2 + · · · ), (3.26)


r2 2
y2 (x) = ky1 (x) ln x + x (A0 + A1 x + A2 x + · · · ),
(3.27)

where k may or may not be equal to zero.


The solution and derivatives inserted into the original equation

X
x2 am (m + r)(m + r − 1)xm+r−2
m=0

X
+ 2x am (m + r)xm+r−1
m=0

X ∞
X
−2 am xm+r + p2 x2 am xm+r = 0.
m=0 m=0
Irrotational standing waves 127

From the indicial equation (3.25) we choose r = 1 and collect


the parts with the same power of x

as (s + 1)sxs+1 + 2as (s + 1)xs+1 − 2as xs+1


+ p2 as−2 xs+1 = 0,
as (s2 + s + 2s + 2 − 2) + as−2 p2 = 0,
as · s(s + 3) = −as−2 p2 ,
−p2
as = as−2 ,
s(s + 3)
128 Standing Waves between Singularities

and develop the even coefficients of an

a0 · 0(0 + 3) = 0,
a0 = a,
(−p2 ) (−p2 )
a2 = a0 =a ,
2·5 2·5
(−p2 ) (−p2 )2
a4 = a2 =a ,
4·7 2·5·4·7
2 2 3
(−p ) (−p )
a6 = a4 =a ,
6·9 2·5·4·7·6·9
·····················
m 3 · (m + 2)
am = a(−1) 2 · pm ,
(m + 3)!
m
h 3 3 i
= a(−1) 2 · pm − , m = 0, 2, 4, 6, · · · ,
(m + 2)! (m + 3)!
h 3 3 i
a2n = a(−1)n p2n − , n = 0, 1, 2, 3, · · · .
(2n + 2)! (2n + 3)!
(3.28)

The odd coefficients of an are all zero:

a1 · 1(1 + 3) = 0,
a1 = 0,
(−p2 )
a3 = a1 = 0,
3·6
a5 = a7 = a9 = · · · = 0.
Irrotational standing waves 129

According to (3.26) the first solution is:


X
y1 =xr1 a2n x2n
n=0

X h 3 3 i
=x a0 (−1)n p2n − x2n
n=0
(2n + 2)! (2n + 3)!
x p2 x3 p4 x5 x p2 x3 p4 x5 
=3a0 − + − ··· − + − + ···
2! 4! 6! 3! 5! 7!
3a0 h  (px)2 (px)4 (px)6 i
= 2 1− 1− + − + ···
p x 2! 4! 6!
3a0 h  (px)3 (px)5 (px)7 i
3 2
− px + px − + − + ··· ,
p x 3! 5! 7!
3a0 3a0
y1 = 3 2 sin (px) − 2 cos (px). (3.29)
p x p x

There may be another solution (3.27) of the form

y2 = ky1 ln(x) + (A0 + A1 x + A2 x2 + · · · ),

where k might or might not be zero. First I try the solution


where k is set to zero. From the indicial Equation (3.25) we
130 Standing Waves between Singularities

choose r2 = −2 and collect the parts with the same power of x.

Am (m − 2)(m − 3)xm−2 + 2Am (m − 2)xm−2


− 2Am xm−2 + p2 Am xm = 0,
As [(s − 2)(s − 3) + 2(s − 2) − 2] = −p2 As−2 ,
As (s − 3)s = −p2 As−2 , (3.30)
2
−p
As = As−2 , (s 6= 0, 3),
s(s − 3)

and develop the coefficients of An

A0 (0 − 3)0 = 0,
A0 = A0 ,
(−p2 )
A2 = −A0 ,
2·1
−p2 (−p2 )2
A4 = −A2 = A0 ,
4·1 1·2·4
−p2 (−p2 )3
A6 = −A4 = A0 ,
3·6 1·2·3·4·6
······
n
n p (n − 1)
An = −A0 (−1) 2 , n = 0, 2, 4, · · · ,
n!
n n
n
h p p i
An = A0 (−1) 2 − , n = 2, 4, 6, · · · ,
n! (n − 1)!
h p2s p2s i
A2s = A0 (−1)s − , s = 1, 2, 3, · · · .
(2s)! (2s − 1)!
Irrotational standing waves 131

We see from (3.30) that A1 = 0, but that A3 may still be 6= 0.


We get the odd coefficients of An :

A1 (1 − 3)1 = 0,
A1 = 0,
A3 (3 − 3)3 = A1 (−p2 ) = 0,
A3 = A3 ,
(−p2 )
A5 = A3 ,
5·2
(−p2 ) (−p2 )2
A7 = A5 = A3 ,
7·4 2·4·5·7
······
n−3 pn−3 · 3 · (n − 1)
An = A3 (−1) 2 · , n = 3, 5, 7, · · · ,
n!
h p2s p2s i
A2s+3 = 3A3 (−1)s − , s = 0, 1, 2, · · · .
(2s + 2)! (2s + 3)!

We first seek the partial solution of (3.27) with the odd coeffi-
132 Standing Waves between Singularities

cients of An set to zero:

∞  p2s
h X p2s  2s i
y21 = x−2 A0 + A0 (−1)s − x
s=1
(2s)! (2s − 1)!
 (px)2 (px)4 (px)6 
= A0 x−2 1 − + − + ···
2! 4! 6!
 (px)1 (px) 3
(px) 5 
+ A0 px−1 − + − ··· ,
1! 3! 5!
A0 p A0
y21 = sin(px) + 2 cos(px).
x x

Next we seek the partial solution of (3.27) with the even coeffi-
Irrotational standing waves 133

cients of An set to zero:


X
y22 =x−2 A2s+3 x2s+3
s=0

X h 1 1 i
=x−2 3A3 (−1)s p2s − x2s+3
s=0
(2s + 2)! (2s + 3)!
x p2 x3 p4 x5
=3A3 − + − ···
2! 4! 6!
x p2 x3 p4 x5 
···− + − + ···
3! 5! 7!
3A3 h  (px)2 (px)4 (px)6 i
= 2 1− 1− + − + ···
p x 2! 4! 6!
3A3 h  (px)3 (px)5 (px)7 i
− px + px − + − + · · · ,
p3 x2 3! 5! 7!
3A3 3A3
y22 = 3 2 sin (px) − 2 cos (px).
p x p x

We notice that the partial solution y22 is identical to the partial


solution y1 , hence the complete solution has got to be of the
form

y = ky1 ln x + y21 + y22 .


134 Standing Waves between Singularities

We can test this equation in order to see if k 6= 0.

y = [y1 ] · k ln x + [y21 + y22 ]


1
y ′ = [y1′ ] · k ln x + y1 + [y2′ 1 + y2′ 2 ]
x
2y ′ 1
y ′′ = [y1′′ ] · k ln x + 1 − y1 2 + [y2′′1 + y2′′2 ].
x x

We already know that the expressions in square brackets are


partial solutions to Equation (3.20) and zeroes out, hence it
only remains to see if the rest of the terms may be another
partial solution with k 6= 0 or if k has got to be zero. We have

2 1 1
x2 (y1′ − y1 2 ) + 2x(y1 )
x x x
=2xy1′ − y1 + 2y1
=2xy1′ + y1
2a 2a ap
=2x[− 3 sin(px) + 2 cos(px) + sin(px)]
px x x
a a
+ 2 sin(px) − cos(px)
px x
3a 3a
= cos(px) − 2 sin (px) + 2ap sin(px).
x px

This expression cannot be zero for any range of r, so k has got


to be zero, and the solution of the differential equation is given
Irrotational standing waves 135

by
1h 3A3 i
y= A0 p sin(px) − 2 cos(px)
x p
1 h 3A3 i
+ 2 sin(px) + A0 cos(px) ,
x p3
and by reentering F = y and r = x we finally acquire the
solution of Equation (3.20):
1h 3A3 i
F = A0 p sin(pr) − 2 cos(pr) (3.31)
r p
1 h 3A3 i
+ 2 sin(pr) + A0 cos(pr) .
r p3
By (3.22) and (3.31) the displacement vector, u(r, t) = F G · r̂,
becomes

u(r, t) =[C1 sin(c1 pt) − C2 cos(c1 pt)] (3.32)


n1h 3A3 i
· A0 p sin(pr) − 2 cos(pr)
r p
1 h 3A3 io
+ 2 sin(pr) + A 0 cos(pr) r̂.
r p3
Equation (3.32) is the complete solution for concentric waves in
the spatial continuum. By a suitable choice of constants we can
investigate how standing waves may occur between imaginary
fixed spheres in the spatial continuum, but more important, we
can see if a standing wave may occur between the inside of a
fixed sphere and the center node. There might be a possibility
136 Standing Waves between Singularities

div u

u
A>0

A=0
0

A<0

r
0 R1

Figure 3.1: Standing wave between the inside of a rigid shell


with radius R1 and the center node.
Irrotational standing waves 137

as we can see from the graph in Figure 3.1. A0 has got to be


zero (else |u| would raise beyond any limits when r approaches
zero), the radius of the sphere is given by the choice of p, and
A3 dictates the amplitude of the standing wave. Under these
conditions there might be established a standing wave between
r → 0 and r = R1 , which is oscillating with a frequency, f ,
given by c1 p = 2πf hence

c1 p 2πf
f= , and p = . (3.33)
2π c1

The displacement function u(r) approaches zero as r → 0 and


r = Rn , which is the necessary condition that there might be
a standing wave between these two borders. This situation is
covered by the expression
h sin(pr) cos(pr) i
u(r, t) = A cos(c1 pt) · − r̂ (3.34)
p3 r 2 p2 r

when C1 and A0 are set to zero, and A = −3A3 C2 .


In order to find the radii of the possible spheres, we set u(r, t) =
0 which gives

tan(pRn ) = pRn , (3.35)

where Rn are possible radii of the firm shell. There is a solution


for every tan(pRn ) = pRn , and we also notice that the first
radius to be considered is when pR1 ≈ 4.49341 radians.
By deriving Equation (3.34) one and two times with respect on
138 Standing Waves between Singularities

t we acquire the velocity and acceleration fields


h sin(pr) cos(pr) i
u̇(r, t) = −Ac1 sin(c1 pt) · − r̂, (3.36)
p2 r 2 pr

and
h sin(pr) cos(pr) i
ü(r, t) = −Ac1 2 cos(c1 pt) · − r̂. (3.37)
pr2 r

In order to see that u, u̇, and ü really approach zero when


r → 0, we can apply l’Hôpital’s rule:

sin(pr) − pr cos(pr) [sin(pr) − pr cos(pr)]′


lim = lim
r→0 (pr)2 r→0 [(pr)2 ]′
1
= lim sin(pr) = 0. (3.38)
r→0 2

By applying the identity, div (f (r)r̂) = 2f (r)/r + f ′ (r), on


Equation (3.34), we acquire

2u ∂u
div u = +
r ∂r
h 2 sin(pr) 2 cos(pr)
= A cos(c1 pt) −
p3 r 3 p2 r 2
2 sin(pr) cos(pr) cos(pr) p sin(pr) i
− + 2 2 + 2 2 +
p3 r 3 p r p r p2 r
sin(pr)
div u = A cos(c1 pt) , (3.39)
pr
Irrotational standing waves 139

and by the identity, grad (f (r)) = f ′ (r)r̂

h sin(pr) cos(pr) i
grad div u = −A cos(c1 pt) − r̂. (3.40)
pr2 r

It is noteworthy that even if the node itself is a mathematical


singularity such that r 6= 0, the limit of both the displacement u
and the divergence of u have finite limits when r is approaching
zero3 .
We easily verify from Equation (3.40) and (3.37) that the N-C
equation for a solenoidal field (3.3) is satisfied.
The amplitude of the displacement from inside a sphere of radius
r = π/p where div u is positive is given by

π/p
sin(pr)
Z
D= A 4πr2 dr
0 pr
4πA h sin(pr) r cos(pr) iπ/p
= −
p p2 p 0
4π 2 A
= . (3.41)
p3

This thought experiment shows that standing waves may be es-


tablished between a single node and a fixed concentric shell, but
it also suggests that individual points in space may be nodes in
a more complicated system of standing waves.
3 Note sin(x)
that limx→0 x
= 1.
140 Standing Waves between Singularities

3.4 Solenoidal standing waves

The Navier-Cauchy equation for solenoidal deformations is

1
curl curl u = − ü. (3.42)
c2

Solving this equation for a deformation around a singularity may


be a bit tricky, so first I will try to guess a solution of the form:

u(r, t) = g(r, t)(m̂ × r̂), (3.43)

where m̂ is a fixed direction in space.


By the identities curl (φA) = φ curl A + grad φ × A, curl (c ×
r) = 2c, curl r = 0, grad (cr) = c, and A × (B × C) =
(AC)B − (AB)C where c is a constant vector, r a radius vector,
and φ a scalar field, we develop

g
curl [g(m̂ × r̂)] = curl [ (m̂ × r)]
r
g g
= curl (m̂ × r) + grad × (m̂ × r)
r r
2g g′ g
= m̂ + ( − 2 )r̂ × (m̂ × r)
r r r
2g ′ g
= m̂ + (g − )[(r̂ · r̂)m̂ − (r̂ · m̂)r̂
r r
g  g′ g
= + g ′ m̂ − − 2 (r · m̂)r̂,
r r r
Solenoidal standing waves 141

and further
g g
curl curl [g(m̂ × r̂)] =( + g ′ ) curl m̂ + grad ( + g ′ ) × m̂
r r
g′ g g′ g
− − 2 (r · m̂) curl r̂ − grad [ − 2 (r · m̂)] × r̂
r r r r
g′ g ′′ g′ g ′
=( − 2 + g )(r̂ × m̂) − − 2 r̂(r · m̂)] × r̂
r r r r
g′ g
− − 2 grad (r · m̂) × r̂
r r

g g g′ g
=( − 2 + g ′′ )(r̂ × m̂) − − 2 (m̂ × r̂)
r r r r
2g ′ 2g
=(g ′′ + − 2 )(r̂ × m̂).
r r

With these terms inserted into Equation (3.42) we acquire

2g ′ 2g 1
(g ′′ + − 2 )(r̂ × m̂) = 2 g̈(r̂ × m̂), (3.44)
r r c

or the scalar differential equation

2g ′ 2g 1
g ′′ + − 2 = 2 g̈, (3.45)
r r c
which is exactly like Equation (3.19) except for the value of c,
and it gives the same solution, hence
h sin(qr) cos(qr) i
u = M cos(cqt) − (r̂ × m̂), (3.46)
q 3 r2 q2 r
142 Standing Waves between Singularities

By deriving Equation (3.46) one and two times with respect on


t we acquire the velocity and acceleration fields
h sin(qr) cos(qr) i
u̇ = −M c sin(cqt) · − (m̂ × r̂), , (3.47)
q 2 r2 qr

and
h sin(qr) cos(qr) i
ü = −M c2 cos(cpt) · − (m̂ × r̂).. (3.48)
qr2 r

It follows from the above equations and some straightforward


work that
h sin(qr) cos(qr) i
curl curl u =M cos(cqt) − (m̂ × r̂),
qr2 r
(3.49)
which together with Equation (3.48) inserted into the Navier-
Cauchy equation confirms that Equation (3.46) is a solution.
We notice that the irrotational and solenoidal components of u
are orthogonal to each other as expected.
It might be of interest to see how curl u behave when r ap-
proaches zero. By applying l’Hôpital’s rule we find in a straight
forward way that
lim curl u = 23 M cos(cqt)m̂. (3.50)
r→0

This interesting result confirms that curl u - and hence also


curl u̇ - has a finite value in the vicinity of an oscillating singu-
larity.
Standing waves between singularities 143

3.5 Standing waves between singular-


ities
In the preceding section I discussed the possibility that there
might be a standing wave between a node and a concentric firm
shell. The most important result was that a singularity may
form the one endpoint in a standing wave. An imaginary firm
shell is of course not realizable in a spatial continuum of infinite
extension, but singularities are indeed possible entities. The
question therefore naturally arises if standing waves can still be
formed in the spatial continuum without the participation of
any firm shells.
By the identity

sin x cos y = 12 [sin(x + y) + sin(x − y)], (3.51)

Equation (3.39) can be rewritten into:


A
div u = cos(c1 pt) sin(pr)
pr
A  
= sin(pr + c1 pt) + sin(pr − c1 pt) , (3.52)
2pr
which can be interpreted as the sum of two spherical waves
bouncing back and forth in opposite directions between the firm
shell and the center node. The two waves can be said to be
reflected successively from the firm shell and the singularity at
the center. We realize that standing waves can be treated as the
superposition of two or more progressive free waves. This makes
144 Standing Waves between Singularities

the mathematics simpler, and we can investigate more complex


systems of standing waves. In this case the waves move with
the speed, c1 . The frequency is f = c1 p/2π, and the wavelength
is λ = 2π/p.
Next let us have a look at the energy which is involved in the
standing wave between the center node and the shell. At the
time t = 0 the whole energy is in the div u-field because the
u̇-field equals zero all over the considered volume. By Equation
(3.10) the energy inside a thin shell at the distance r from the
center is
λ + 2µ
de = ( div u)2 · 4πr2 dr,
2
and inside a shell with radius Rn

2π(λ + 2µ)A2 Rn 2
Z
ERn = sin (pr)dr
p2 0
2π(λ + 2µ)A2 h Rn sin(2pRn ) i
= 2
− .
p 2 4p

We notice that the total energy inside the sphere approaches


infinity when Rn grows without any limit. Hence a single oscil-
lating node cannot exist in the spatial continuum.
Let us, however, make a new thought experiment. Inside a
sphere with a great radius and many wavelengths of standing
waves, there are two oscillating nodes oppositely placed at a
short distance, z1 = +d and z2 = −d, from the center, and let
the the two nodes oscillate in opposite phase with each other
Standing waves between singularities 145

Figure 3.2: Considering a point in spherical coordinates

(see Figure 3.2). Then by adjusting Rn a little, it should be


possible to let two progressive waves move from one node via
the firm shell and encounter the other node in tune with the
oscillation of that node, where it is reflected. We then would
have a system of two separate waves bouncing back and forth
between the two nodes via the firm shell. The superposition of
the free waves would form a standing wave, and the problem is
now to find the total energy in the new system. If it is finite
and the energy is kept in the surrounding of the singularities,
it may indicate that we could get rid of the reflecting shell and
end up with a pure oscillating dipole.

First I will only consider the energy in a volume element dV at


146 Standing Waves between Singularities

a distance r from the center that is great in comparison with


the distance 2d between the nodes. Then the angle between
the axis through the two nodes and the direction towards the
volume element can be considered to be the same, namely φ,
from both nodes and the center point. The divergence in the
volume element at P is the superposition of the divergence from
the two nodes at the time t = 0
h sin p(r + d cos φ) sin p(r − d cos φ) i
div u = A −
p(r + d cos φ) p(r − d cos φ)
A 
≈ sin(pr + pd cos φ) − sin(pr − pd cos φ)
pr
A
= cos(pr) · sin(pd cos φ).
pr

The energy in a zone between Rm ≫ d and Rn when n ≫ m in


the two half spheres is
λs + 2µs
ZZZ
E= ( div u)2 dV
2
λs + 2µs Rn π π A2 n
Z Z Z
=2 2 r2
cos2 (pr) · sin2 (pd cos φ)r2 sin φ dφ dθ
2 Rm 0 0 p
(λs + 2µs )A2 Rn
Z Z π Z π
2 2
= cos (pr)dr · sin (pd cos φ) sin φ dφ dθ
p2 Rm 0 0

(λs + 2µs )A2 Rn cos φ sin(pd cos φ) cos(pd cos φ)


Z
2
= cos (pr)dr − +

p2 Rm 2 2pd
2h Z Rn
π(λs + 2µs )A sin(2pd) i
E= 1− cos2 (pr)dr.
p2 2pd Rm
Standing waves between singularities 147

The property sin(2pd)/2pd has a maximum value of unity when


2dp approaches zero, hence the total energy has no upper bounds
except when either d or p approaches zero. Hence we can con-
clude that a single isolated dipole cannot exist because it would
take an unlimited amount of energy to keep it up. Notice by
the way that E is vanishing in the equatorial direction from the
origin and increases when φ shrinks towards zero (i.e. it has its
maximum value in the z-direction).
The next possibility I will examine is whether a chain of os-
cillating nodes may exist. So let an infinite chain of oscillating
nodes with the same strength be organized along a straight line
separated with a distance d and organized such that one node
always oscillate in opposite phase with its two adjacent nodes.
Hence the next node that oscillates in the same phase is at a
distance λ1 = 2d. The divergence in a point P at a distance r
from the chain and an offset z along the chain from node #0,
is given by the superposition of the effects from all the nodes in
the chain

∞ p
2 2
n sin p r + (nd − z)
X
div u = A cos(c1 pt) (−1) .

p
p r 2 + (nd − z)2
n=−∞

Let pd = π, and 0 < z < d, then we acquire


c1
div u =A cos(2π t)
2d
p πr 
( d )2 + (πn − πz
∞ 2
n sin d )
X
· (−1) p πr 2 .

πz 2
n=−∞
( d ) + (πn − d )
148 Standing Waves between Singularities

This formula should be valid for all of space, but let us see what
div u amounts to strictly along the chain axis where r = 0.

sin(πn − πz

d )

c1 X
(−1)n

div u = A cos(2π t) πz
2d n=−∞ πn − d
∞ πz n
n sin(− d )(−1)

c1 X
= A cos(2π t) (−1)
2d n=−∞ πn − πzd


c1 πz  X (−1)n
= A cos(2π t) sin ,
d d n=−∞ |nπ − zπ
d |

which is a alternating harmonic series that can be shown to be


convergent.
As stated above, the result is only valid strictly along the chain
axis, but a simple consideration shows that the result is valid
also when P is outside the axis. First take the sum when −m <
n < m. It is of course finite. Then let m be so great that
md ≫ r. Then we can set r = 0 for the rest of the series, and
the reminder will also be finite making the whole series finite.
Even when r grows beyond all limits, the effect from one node
will be canceled out by the effect from its neighbor nodes.
This proves that a string of oscillating nodes may have a finite
energy per unit length. Say that an extreme situation in the
spatial continuum, with release of an enormous amount of en-
ergy, initiated a boiling phase where nodes of the type described
above where formed. Then the nodes would have got to orga-
nize in a way that required the least amount of energy, i.e. along
Chains of irrotational and solenoidal oscillating nodes 149

strings. Hence if oscillating nodes at all can be present in the


spatial continuum, they will tend to organize along strings.
The important result, that standing waves may form between
nodes in long strings, opens for the possibility that space can
be filled with a plethora of faint oscillating nodes that organize
and reorganize along strings - not necessarily formed as straight
lines - in an ever changing pattern.

3.6 Chains of irrotational and


solenoidal oscillating nodes
Since div u has a finite value at every point in space, we could
express it as a function of type
c1 πz
div u = A(r, z) cos(2π t) sin( ),
2d d
where A(r, z) is the amplitude at the point P . The oscillatory
frequency is
c1
f1 = . (3.53)
2d
By the identity, cos a sin b = 1/2 sin(a + b) − 1/2 sin(a − b), we
can decompose the function above into

1 n π  π o
div u = A(r, z) sin (c1 t + z) − sin (c1 t − z) ,
2 d d
150 Standing Waves between Singularities

which can be interpreted as the superposition of two waves mov-


ing in opposite directions along the chain of nodes with velocity
c1 .

In much the same way as considered above, it should even be


possible to assume a solenoidal standing wave in another chain
of nodes that is given by a function of the type

1 n  2π   2π o
curl u = A(r, θ, z) sin (ct + z) − sin (ct − z) .
2 d d

The main difference between these two set of waves, is that the
latter moves with the velocity c, which is probably about half
of the velocity c1 , and the frequency

c
f2 = . (3.54)
2d

Both sets of deformation fields can, however, be expressed as two


progressive waves moving in opposite directions in space, and
the resultant of these waves are standing waves along separate
strings in space.

So far all the deductions are done on the basis of the Navier-
Cauchy equation, but in the N-C representation there is no cou-
pling between the two types of oscillations. Hence a chain with
both types of oscillation represented in the same set of nodes is
not feasible in the N-C representation, but that doesn’t mean
that a more thorough examination rules out this possibility.
The Navier-Stokes equation and Coupled oscillations 151

3.7 The Navier-Stokes equation and


Coupled oscillations
In this section I will discuss the possibility that the two basic
types of oscillation can occur in the same set of nodes. In the
Navier-Cauchy representation the two wave equations are inde-
pendent of each other and cannot interact, but not so in the
Navier-Stokes representation of a continuous medium. Disre-
garding viscosity and the influence of any outer forces, the N-S
equation takes the form
h ∂ u̇ i
(λs + 2µs ) grad div u − µs curl curl u = ρs + (u̇ · ∇)u̇ ,
∂t

or by the identity, grad (A · A) = 2[A × curl A + (A · ∇)A]:


h ∂ u̇ 1 i
(λs + 2µs ) grad div u − µs curl curl u = ρs + grad u̇2 − u̇ × curl u̇
∂t 2
where the additional term is the directional derivative of u̇ with
respect on t along the tangent vector to the velocity field. We
now turn back to the central case with standing waves bouncing
back and forth between a center node and an imaginary firm
concentric sphere. The displacement vector u can as seen earlier
be split into a radial component u1 = f (r, t)r̂ and a tangential
component u2 = g(r, t)(m̂ × r̂), which are orthogonal to each
other. We see that the equation no longer can be split into
two independent equations. We have at our hands two different
oscillation systems which have a mutual coupling between them.
152 Standing Waves between Singularities

I think that the Navier-Stokes equation is too complicated to


be solved by conventional methods, so instead of solving the
equation, I will try to figure out what the new terms will do to
the partial solutions of the Navier-Cauchy equation.
Let us first single out the term ρ(u̇ × curl u̇) and see what im-
pact it may have on the oscillatory system. The effect is only
significant with relatively great deformations in the near vicinity
of the node. Here the N-S equation takes the form (note that
u̇1 ⊥ u̇2 , div u2 = 0, curl u1 = 0):
(λ + 2µ) grad div u1 − µ curl curl u2
∂ u̇1 ∂ u̇2 1 1
=ρ + + grad u̇12 + grad u̇22 − u̇1 × curl u̇2 − u̇2 × curl u
∂t ∂t 2 2
The terms in this formula are representing different components
of force. If we multiply them with the velocity at a site, we get
the power, or energy transfer per units of time and volume,
that are transferred between the different fields. For example
the formula
∂ u̇1
u̇1 · (λ + 2µ) grad div u1 = u̇1 · ρ ,
∂t
represent the energy transferred between the local potential and
kinetic energies in an irrotational oscillation per time unit at a
given moment. If both sides are negative, the energy transfer
is from potential to kinetic energy, and if they are positive, the
energy transfer is from kinetic to potential energy. By the way,
this component of the oscillation is between compression and
rarefaction and will not result in any net displacement of spatial
mass.
The Navier-Stokes equation and Coupled oscillations 153

Now let us multiply the N-S equation with the velocity u̇ =


u̇1 + u̇2 , and neglect all terms that are orthogonal to each other
(note that the vector product of two vectors is orthogonal to
both of them). We obtain
∂ u̇1 ∂ u̇2
c12 u̇1 · grad div u1 − c22 u̇2 · curl curl u2 = u̇1 · + u̇2 ·
∂t ∂t
1
+ u̇ · grad u̇2 − u̇2 · (u̇1 × curl u̇2 ) − u̇1 · (u̇2 × curl u̇2 ).
2
The two last terms representing components of the power ex-
erted by the two fields can be set like

w1 = u̇1 · (u̇2 × curl u̇2 ),


w2 = u̇2 · (u̇1 × curl u̇2 ).

We see by the sequence in which the three terms occur that


when one of the properties is positive, then the other is nega-
tive and vice versa. Clearly this shows that there may be an
energy transfer back and forth between the irrotational and the
solenoidal field in the vicinity of an oscillating node. Hence the
two fields are coupled and there might be some normal modes
of oscillation where the two fields can oscillate within the same
firm shell with a common frequency. The three vectors u̇1 , u̇2 ,
and curl u̇2 make up a parallelepiped with volume w1 = w2 .
Hence the energy that is given up from one field is exactly like
the energy received by the other field. It is also implicit that
the two velocity fields has got to be a quarter out of phase with
each other such that the maximum displacement coincides with
the maximum rotation. By (3.36) and (3.47) we could end up
154 Standing Waves between Singularities

with two equations of the form

u̇1 = A1 sin 2πνf (r)r̂


(3.55)
u̇2 = A2 sin (2πν + ψ)f (r)(r̂ × m̂),

where ψ is the phase shift between the two fields. According to


the above assumption it might be like π/4.
In Figure (3.3) I have tried to visualize some possible standard
modes of oscillation which could occur depending of the initial
condition. In the left columns the energy is shifted from dis-
placement to rotation, and in the right columns the energy is
shifted from rotation to displacement. In the upper row the
oscillation remains symmetric, while in the lower row the os-
cillation becomes asymmetric, meaning that the inflation (or
rarefaction) phase is stronger than the rarefaction (or inflation)
phase, or the right/left orientation is stronger than the cor-
responding left/right orientation. I also suspect that if these
modes of oscillation (or something like them) at all can occur in
a string of nodes, the lower ones would curl up along an inter-
wound line while the two upper ones could occur along a straight
line. Well, so much for speculations.
Finally I will try to estimate what effect the term grad ( 12 ρ u̇2 )
can have on the displacement of spatial mass. The gradient of
the kinetic energy density, 12 ρ u̇2 , can be taken up by the force
(λs + 2µs ) grad div u1 , hence we can set up the equation
ρs
(λs + 2µs ) grad div u = grad u̇2 ,
2
The Navier-Stokes equation and Coupled oscillations 155

and because u = u1 + u2 , u1 ⊥ u2 , and div u2 = 0 we get

ρs
(λs + 2µs ) grad div u = grad (u̇12 + u̇22 ),
2

and further, since the deformation is zero at infinity, we obtain

ρs
div u = (u̇ 2 + u̇22 ).
2(λs + 2µs ) 1

The displacement D from a volume V is given by


Z
D= div u dV.
V

By Equation (3.55) we acquire the displacement from inside a


sphere with radius R

ρs nZ R 2
D= A1 sin(2πνt)f (r) 4πr2 dr
2(λ + 2µ) 0
Z RZ π o
 2
+ A2 sin(2πνt + ψ)f (r) sin φ 2πr sin φ r dφ dr
0 0
R
ρs
Z
 2 2
A sin (2πνt) + 23 A22 sin2 (2πνt + ψ) · f (r)2 4πr2 dr.

=
2(λ + 2µ) 1 0

If we set ψ = ±π/2, the two oscillations are shifted one quarter


of a period in relation to each other, and moreover if

A12 = 2/3A22 (3.56)


156 Standing Waves between Singularities

the displacement is independent of time and we get


Z R
ρs
Z
2
( div u′1 )dV = A f (r)2 4πr2 dr,
V 2(λ + 2µ) 1 0
1 ρs 2
Z
= u̇ dV,
(λ + 2µ) V 2 1

or

E′
D= ,
λ + 2µ

where E ′ is the kinetic energy inside the sphere.


In the Navier-Cauchy representation of a string of oscillating
node there is no net displacement of spatial mass from the sys-
tem - the compression into one node is like the displacement
from another node - but the above equation shows that the pres-
ence of kinetic energy will inevitably lead to a net displacement
of spatial mass from the entire string. The energy component
E1 is only a part of the total energy, E, in the system, so if
it should happen to be like one third of the total energy, the
displacement amounts to
E
D= ,
3(λ + 2µ)

which corresponds to the energy-displacement relation that I


found in the paper Elastodynamics in a continuum of infinite
extension.
The Navier-Stokes equation and Coupled oscillations 157

Rotation enhanced mode Displacement enhanced mode

Asymmetric rotation enhanced mode Asymmetric displacement enhanced mode

Figure 3.3: Coupled oscillations. The blue solid lines repre-


sent the div u-field and the red solid lines the curl u̇-lines. The
smaller graphs in the middle represent the transfer of movement
between the two main oscillations, and the thick solid lines are
the resulting amplitudes.
158 Standing Waves between Singularities
Chapter 4

Spatial Continuum
Mechanics

A traditional elastic medium like steel consists of material par-


ticles bound together by electromagnetic forces. The medium
gets its inertial properties from the masses of these particles,
and if it is compressed, its mass density increases because more
particles are squeezed into a smaller volume. On the other hand
mass is proportional with energy according to Einstein’s rela-
tion, E = M c2 . In this paper I will discuss a couple of ways a
true spatial continuum can acquire inertia and how mass den-
sity changes by being compressed. Finally I will discuss how the
wave speed is affected.
160 Spatial Continuum Mechanics

4.1 Strain
In this section I will introduce some necessary tensors in order
to describe the deformation of an initially homogeneous spatial
continuum.
In a true spatial continuum it will be possible to displace spa-
tial points away from their original positions. We consider the
configuration of spatial points B0 at time t0 in a 3D Euclidian
space E3. In B0 , space is undeformed and unstressed. The po-
sition vector of a point P0 of B0 relative to the origin O of an
orthogonal Cartesian coordinate system is denoted by

X = X i ii , where X i : X 1 , X 2 , X 3

are Lagrangian or material coordinates and ii = ii are unit


vectors along the X i -axes. We now suppose the spatial con-
tinuum to take at a certain time t a new configuration B in
E3. Thus, the point P0 is moved into the position P which will
be determined with respect to the same origin by the position
vector

x = xi ii , where xi : x1 , x2 , x3

are called Euler or spatial coordinates.


We now assume the mapping of B0 into B such that the cor-
respondence of the point P0 and P is one to one and may be
described by the transformation

xi = xi (X 1 , X 2 , X 3 , t) → x = x(X, t),
Strain 161

which is reversible

X i = X i (x1 , x2 , x3 , t) → X = X(x, t).

This condition is satisfied if the functions xi and X i are single


valued and at least once continuously differentiable with respect
to their arguments. In addition the Jacobian must be positive:
∂xi
J = > 0.

∂X j

We now imbed a convective coordinate system into the spa-


tial continuum, i.e. the coordinate system undergo the same
deformations as the spatial continuum itself. In such a coor-
dinate system any point will maintain the same coordinates,
Θi : (Θ1 , Θ2 , Θ3 ), within the course of the deformation. It fol-
lows that

X i = X i (Θ1 , Θ2 , Θ3 ) → X = X(Θ1 , Θ2 , Θ3 ), (4.1)

and

xi = xi (Θ1 , Θ2 , Θ3 ) → x = x(Θ1 , Θ2 , Θ3 ). (4.2)

The coordinate system will in general be curvilinear, but it will


always be possible to select it as Cartesian coordinates for ex-
ample Θi = xi in B at a given time t, which would make the
corresponding coordinate lines in B0 curvilinear. I will keep
this possibility in mind for later use. For now we suppose the
mapping of B0 into B to be described by curvilinear coordinates
Θi .
162 Spatial Continuum Mechanics

The line elements in the undeformed state can be derived from


(4.1)
∂Θi ∂X i j
dΘi = dX j , dX i = dΘ ,
∂X j ∂Θj
and, similarly in the deformed state, from (4.2) at a fixed time
t
∂Θi j ∂xi
dΘi = dx , dxi = dΘj .
∂xj ∂Θj
From (4.1) we also derive the base vectors related to P0
∂X ∂X k
Gi = X,i = = ik ,
∂Θi ∂Θi

∂Θi k
Gi = i ,
∂X k
and accordingly from (4.2) the base vectors related to P
∂x ∂xk
gi = x,i = i
= ik ,
∂Θ ∂Θi

∂Θi k
gi = i ,
∂xk
The position of point P relative to P0 is called the displacement
vector and denoted u. We introduce for u two sets of compo-
nents Ui and ui
u = u(Θi , t) = (x − X) = Ui Gi = ui gi .
Strain 163

defined with respect to the undeformed basis Gi and the de-


formed basis gi , respectively.
The vectorial line elements dX and dx related respectively to
the material and spatial coordinates X and x are given by
∂X
dX = dΘi = Gi dΘi ,
∂Θi

∂x
dx = dΘi = gi dΘi .
∂Θi

The deformation gradient is defined by:

F = gi ⊗ Gi , (4.3)

and accordingly the inverse tensors by

F−1 = Gi ⊗ gi . (4.4)

For later use we introduce the Lagrangian gradient of x refer-


ring to the undeformed state

∂x ∂x ∂Θi j
grad x = j
⊗ ij = i
⊗ i = gi ⊗ Gi = F,
∂X ∂Θ ∂X j
(4.5)

and similarly the spatial or Euler gradient

∂X j ∂X ∂Θi j
grad X = ⊗ i = ⊗ i = Gi ⊗ gi = F−1 .
∂xj ∂Θi ∂xj
164 Spatial Continuum Mechanics

Hence

F = grad (X + u) = G + grad u. (4.6)

Note that G = I, the identity tensor in the undeformed base.


By contracting (4.3) and (4.4) by Gi and gi respectively we
acquire

FGi = (gm ⊗ Gm )Gi = gm (Gm · Gi ) = gm δim = gi


(4.7)

F−1 gi = (Gm ⊗ gm )gi = Gm (gm · gi ) = Gm δim = Gi


(4.8)

from which it clearly follows that the deformation gradient


transforms the convective curvilinear coordinate system from
the undeformed to the deformed basis and vise versa. Further-
more by contracting (4.7) and (4.8) by dΘi we obtain

gi dΘi = (FGi )dΘi = F(Gi dΘi ) → dx = FdX,

Gi dΘi = (F−1 gi )dΘi = F−1 (gi dΘi ) → dX = F−1 dx,

from which it is clear that F can also be used to transform a


line element in the undeformed basis to a line element in the
deformed basis and conversely. These formulas clearly are inde-
pendent of which coordinate system we choose to use.
Strain 165

As with any second-order tensor the polar decomposition theo-


rem states that the deformation gradient F can be multiplica-
tively decomposed in the form

F = RU = vR (4.9)

into a rotation tensor R, which is orthogonal, and a right stretch


tensor U or a left stretch tensor v, which are supposed to be
positive definite and symmetric. These translations can geomet-
rically be interpreted as a stretch of a volume element by U, a
rotation by R, and a translation by u, or a translation by u, a
rotation by R, and a stretch by v.
In general any symmetric second-order tensor, in this case U and
v, possesses three eigenvalues λi and three orthogonal principal
axes which can be determined by the eigenvectors Ni . If the
tensor U refers to the orthonormal basis Ni
ij
U = U Ni ⊗ Nj

it possesses only three non-vanishing components


ii ij
U = U ii = λi , U = U ij = 0 f or i 6= j.

This permits to represent U by the spectral decomposition the-


orem as
3
X
U= λi Ni ⊗ Ni .
i=1
166 Spatial Continuum Mechanics

The eigenvalues λi are determined through the condition


det(U − λI) = 0,
and the unit vectors Ni satisfying the homogenous equation
(U − λi I)Ni = 0 (4.10)
determines the mutually orthogonal principal axes of U.
In accordance with (4.9) we now express U by
U = RT F = RT vR.
Next we insert the above expression for U into (4.10) and con-
tract it from the left side by R. This delivers
(RRT vR − λi RI)Ni = 0

(v − λi I)RNi = 0
by taking the orthogonality condition RRT = I and the identity
RI = IR into account. By inserting a new unit vector
ni = RNi , Ni = RT ni (4.11)
it becomes
(v − λi I)ni = 0.
Thus the spectral decomposition of v is given by
3
X
v= λi ni ⊗ ni
i=1

with the same eigenvalues λi as for the spectral decomposition


of U and the principal direction given by (4.11).
Stress-strain relations 167

Figure 4.1: Deformation of volume element with edges showing


in principal directions Ni or ni as a 2D illustration.

4.2 Stress-strain relations

In this section I will postulate a stress-strain relationship that


is valid for great deformations. I will also consider a general
deformation on top of a uniform compression of space.
In the preceding section it was pointed out that strain can be
168 Spatial Continuum Mechanics

represented as a translation by u, a rotation by R, and a stretch


by v. It was also pointed out that like any symmetric tensors,
the stretch tensor v can be represented by the spectral decom-
position theorem
3
X
v= λi ni ⊗ ni , (4.12)
i=1

where the eigenvalues λi are the principal values and the or-
thogonal base vectors ni are the associated principal directions
of the stretch tensor v.
The spatial continuum is per definition homogeneous, isotropic,
and elastic. Hence a deformation as shown in Figure 4.1 will
be closely related to real internal stress that generally can be
expressed by the second order symmetric Cauchy stress tensor

σ = σ(F),

which has got to be some function of the deformation gradient


F. Moreover, since it is a second order symmetric tensor, it can
be decomposed by the same decomposition theorem in much of
the same way as the stretch tensor v, and since the spatial con-
tinuum by definition is isotropic, it will have the same principal
axes as v, so
3
X
σ= γi ni ⊗ ni ,
i=1

where the eigenvalues γi are the principal values of σ.


Stress-strain relations 169

Figure 4.2: Stress acting on a volume element with edges show-


ing in principal directions Ni or ni as a 2D illustration.

Aside from the more or less obvious assumption that γi is some


function of λi , it is for several reasons not a suitable way to
go further along the classical route leading to the definition of
the forth-order material tensor C, so here I have come to a
point where I have got to delve into some calculated guesses
and assume some more or less probable connections.
Let two opposite directed forces, say outwards along the n1 -
axes, act on two complementary surfaces of a small cube in
170 Spatial Continuum Mechanics

the spatial continuum, see Figure 4.2. In order to keep the


volume unchanged from the undeformed to the deformed state,
we apply four mutually equal compressible forces on the four
faces of the cube. Hence J = λ1 ·λ3 ·λ3 = 1. My first assumption
is that the force in the n1 -direction is proportional to ln(λ1 )
keeping J constant like 1. Next let the force in the n1 -direction
be such that λ1 remains unchanged, i.e. λ1 = 1. The forces
in the remaining two directions are such that the volume of
the cube is changed to some arbitrary values, e.g. dV /dV0 =
1 · λ2 · λ3 = J. My next assumption is that the forces acting
on the faces orthogonal to those in the n1 -dirction needed to
keep λ1 unchanged is proportional to ln(J). Finally, if both the
volume and the stretch along one or more of the principal axes
are changed, I will combine these two assumptions and postulate
that the stress-strain relation for the spatial continuum is given
by

γi = 2µ ln(λi ) + 2µβ ln(J) = 2µ ln(λi · J β ), (4.13)

where 2µ and β are constants1 . Hence the Euler stress tensor


is given by the fairly simple relation

3
X
2µ ln λi · J β ni ⊗ ni .

σ= (4.14)
i=1

1 I have chosen to include the number 2 because then µ turns up to be

identical with the Lamé elastic constant.


Stress-strain relations 171

Generally
3
X
ln I = ln ni ⊗ ni ,
i=1

and for any scalar a we have

ln(aI) = ln[I + (aI − I)] = ln[I + I(a − 1)]


= I(a − 1) − 12 I2 (a − 1)2 + 31 I3 (a − 1)3 · · ·
= I (a − 1) − 12 (a − 1)2 + 13 (a − 1)3 · · ·
 

= I ln a, (4.15)

so by (4.12) Equation (4.14) can finally be write


3
X 3
X
σ = 2µβ ln (J ni ⊗ ni ) + 2µ ln (λi ni ⊗ ni )
i=1 i=1
= 2µβ ln(J I) + 2µ ln v. (4.16)

There is still another situation to be considered. Suppose that


space initially is compressed and then gets an additional defor-
mation. The goal would then be to consider what additional
stress this deformation would cause. In Figure (4.3) I have il-
lustrated the situation as a 2D representation. We first have a
uniform compression of space bringing a volume element from
B0 to B and then an arbitrary deformation bringing the volume
element from B to B. The principal axes for B and B follow by
172 Spatial Continuum Mechanics

Figure 4.3: Stress caused by two successive deformations, one


pure compression followed by an infinitesimal strain, as a 2D
illustration.
Stress-strain relations 173

the spectral decomposition theorem, and since the deformation


between B0 and B is uniform without rotation, Ni can be cho-
sen at will like Ni . The first step can be described by the the
relation
3
X
σ = 2µ ln(J β λ N1 ⊗ N1 ),
i=1

and the next step would bring the stress tensor up to


3
X
σ = 2µ ln(J β · δJ β · λ · δλ1 n1 ⊗ n1 )
i=1
X3 3
X
= 2µ ln(J β λ n1 ⊗ n1 ) + 2µ ln(δJ β · δλ1 n1 ⊗ n1 ).
i=1 i=1
(4.17)

The change of stress can be represented by the difference be-


tween these two stress components and we finally arrive at
3
X
δσ = 2µ ln(δJ β · δλ1 n1 ⊗ n1 ),
i=1

or by changing the notation δσ, δJ and δλi into σ, J and λi


respectively we get
3
X
σ = 2µ ln(J β λ1 n1 ⊗ n1 ). (4.18)
i=1
174 Spatial Continuum Mechanics

Hence the additional strain-stress relation for deformations in a


uniformly compressed or expanded space is the same as it is in
the undeformed space.

4.3 Velocity and acceleration


In order to find the acceleration of points in the spatial con-
tinuum special care has got to be taken if spatial description
is considered. The following notation for the ’material’ time
derivative in spatial notation is adopted from the mechanics of
solids [1, page 115 ff.]
Df
= ḟ ,
Dt
where ḟ is given by

∂ ḟ
ḟ = + ( grad f )v.
∂t
The velocities, v, of spatial points are given by the intrinsic
function
∂u
v = u̇ = + ( grad u)v,
∂t
and the acceleration by
∂ u̇
a = ü = + ( grad u̇)u̇.
∂t
The strain-stress relation for small deformations 175

By infinitesimal deformations the last term is negligible and the


acceleration is simply
∂2u
a = ü = .
∂t2

4.4 The strain-stress relation for small


deformations
In this section I will show that a small variation on top of a
great compression of space obeys the same law that we find in
a classical continuum.
The spatial continuum is per definition isotropic, and a uniform
compression of space will not alter this property. It can be
shown that in an isotropic continuum a stress function σ(v)
can be obtained simply by setting v = (FFT )1/2 , see [1, page
147]. By this relation and (4.6) the strain-stress relation (4.16)
can be developed into
σ = 2µβ ln(J I) + 2µ ln(FFT )1/2
= 2µβ ln(λ1 λ2 λ3 I) + 2µ 21 ln F + ln FT


= 2µβ(ln λ1 + ln λ2 + ln λ3 )I
h i
+ 2µ 12 ln(I + grad u) + ln I + ( grad u)T .
By linearization of the tensor valued function of σ with respect
on u we get
Lσ(u, ∆u) = σ(u) + ∆σ(u, ∆u), (4.19)
176 Spatial Continuum Mechanics

where the second term on the right hand side is the Gateaux-
derivative defined by
d
∆σ(u, ∆u) = σ(u + ε∆u).

Equation (4.19) can be developed into
∆λ1 ∆λ2 ∆λ3
Lσ(u, ∆u) = σ(u) + 2µβ( + + )I
λ1 λ2 λ3
1 h ( grad ∆u) ( grad ∆u)T i
+ 2µ + .
2 I + grad u I + ( grad u)T ε=0
By the identity (u = 0) ⇒ (λi = 1 ∧ grad u = 0) we have
Lσ(u = 0, ∆u) = 2µβ(∆λ1 + ∆λ2 + ∆λ3 )I
i
+ 2µ 12 ( grad ∆u) + ( grad ∆u)T .


(4.20)
We now introduce the linear tensor valued function at the point
where u = 0
1
grad ∆u + ( grad ∆u)T

ǫ= (4.21)
2

From (4.12) we have that


tr v = vii = λ1 + λ2 + λ3 ,
so
∆λ1 + ∆λ2 + ∆λ3 = tr (∆v). (4.22)
The strain-stress relation for small deformations 177

By (4.9) and (4.5) we have

v = FRT = (I + grad u)RT

so
 ∂v 
tr (∆v) = tr ∆( grad u)
∂( grad u)
 T   
= tr R grad (∆u) = tr grad (∆u)
 
= tr sym grad (∆u) = tr ǫ. (4.23)

By (4.21), (4.22), and (4.23) Equation (4.20) finally takes the


form

σ = λ tr ǫ I + 2µǫ, (4.24)

where λ = 2µβ.
This is the same relation as we encounter in literature on the
strain stress-relation for small deformations in a homogeneous
and isotropic elastic continuum, and it follows that the elastic
potential is given by
λ
W = (tr ǫ)2 + µ tr ǫ2 ,
2
see e.g. [1, page 167].
As shown in Equation (3.1), (3.7), and (3.8), the strain-stress re-
lation (4.24) and the Cauchy equation of motion ρü = div σ+b
[1, page 126], where b is some hypothetical force per unit vol-
ume and ρ is the the inertial density to be discussed in the next
178 Spatial Continuum Mechanics

section, can be developed into the Navier-Cauchy equation

(λ + 2µ) grad div u − µ curl curl u + b = ρü, (4.25)

and further into two wave equations

1 ∂ 2 (div u)
∇2 (div u) − = 0, (4.26)
c12 ∂t2

1 ∂ 2 (curl u)
∇2 (curl u) − = 0, (4.27)
c22 ∂t2
representing longitudinal waves moving with speed c1 and
transversal waves moving with speed c2 respectively. The ve-
locities are

s
λ + 2µ
c1 = ,
ρ

µ
r
c2 = . (4.28)
ρ

The relation between c1 and c2 is


s
λ + 2µ p
c1 /c = = 2(β + 1), (4.29)
µ
Inertia and the speed of waves as a function of the compression of
space 179

and we notice that longitudinal waves are travelling with the


double of the speed of transversal waves if β should happen to
be unity. Finally we notice that both µ and λ are true universal
constants according to the definitions in Equation (4.13) and
(4.24).

4.5 Inertia and the speed of waves as


a function of the compression of
space
Elastic waves can only propagate through space if it has inertial
properties. In this section I will consider two possible ways
through which inertia can enter the stage.
The classical approach. The simplest possibility is the clas-
sical assumption that inertia is an intrinsic property of the con-
tinuum such that the spatial density in uncompressed space is
like a universal constant ρ0 and increases or decreases if space
is being compressed or inflated according to the relation
dV0
ρclassical = ρ0 = ρ0 D.
dV

By the classical approach the propagating velocity of transversal


waves as a function of compression are given by
µ
r
c2 (D) = .
ρ0 D
180 Spatial Continuum Mechanics

Let us consider small variations of the waves speed c in an area


of space that initially is compressed by a factor D. This can be
achieved by the linearization of c2 with respect to D which is
defined by

L c2 (D, ∆D) = c(D) + ∆c(D, ∆D), (4.30)

where the second term on the right-hand side as usual is the


Gateaux-derivative defined by
d
∆c2 (D, ∆D) = c(D + ε∆D) , (4.31)

dε ε=0
so
 µ 1/2 d
∆c2 (D, ∆D) = (D + ε∆D)−1/2

ρ0 dε ε=0
 µ 1/2  1 
= − D + ε∆D)−3/2 · ∆D

ρ0 2 ε=0
1 µ ∆D
r
=− · . (4.32)
2 ρ0 D D

Generally we have the identity


1 ∆V
ZZ
div u = lim u n dA = lim , (4.33)
V →0 V V →0 V
A

and the inverse of the Jacobian is given by

V0
D = J −1 = lim ,
V →0 V
Inertia and the speed of waves as a function of the compression of
space 181

so by (4.33)
∂D −V0 ∆V
∆D = ∆V = lim ∆V = −D lim
∂V V →0 V 2 V →0 V
= − D div (∆u), (4.34)

where div (∆u) is the spatial divergence in Euler coordinates.


By (4.34) and (4.32) we finally acquire from (4.30) the variations
of c as a function of small deformations u in a space that initially
is compressed by a factor D corresponding to a wave speed of c
µ
r
c2 = (1 + 12 div u) = c(1 + 12 div u). (4.35)
ρ0 D

In later discussions it is the gradient of c2 that is of most interest


so we write

grad c2 = α c grad div u, (4.36)

where

α = 12 . (4.37)

We notice that by this approach the wave speed of transversal


waves varies as a function of small displacements u independent
of how much space initially is compressed or inflated.
The spatial approach. Another possibility is that space in its
undeformed state has no inertia at all and only gets its inertia by
the energy going into it by being compressed or inflated. This is
more in line with how earthly matter gets its mass according to
182 Spatial Continuum Mechanics

the famous equation e = mc2 where e is the energy dencity in a


body and c is the speed of light. The property m then becomes
the definition of mass. If this is how Nature works, the spatial
mass density might be given by

ρs = e/κ 2 ,

where e is the deformation energy per unit volume going into


space by compression or inflation, and κ is a universal constant.
This is a very interesting possibility because it would mean that
the speed of any elastic waves will approach infinity if space
is in a nearly undeformed state. In order to investigate this
possibility, it will be necessary to find the energy going into
space by compression. For simplicity I will consider a space
where the constant β is set to unity.
By uniform compression of space, a little sphere with radius r0
will shrink to a sphere with radius r. The force acting on a unit
surface element of the sphere is given by the Cauchy stress
vector

t = σn

where n is the unit vector normal to the surface. Since the


pressure is uniform in all directions the Jacobian is given by

dV
J= = λ3 where λ = λ1 = λ2 = λ3 .
dV0

The Euler stress vector is normal to any surface we choose,


Inertia and the speed of waves as a function of the compression of
space 183

e.g. on a surface whose unit normal points in the n1 -direction


3
X
t1 = σn1 = 2µ ln λ4 (ni ⊗ ni )n1 = 2µ ln λ4 n1
i=1

Hence t is normal to the surface of the sphere and is pointing


outwards by stretch. The energy going into the sphere as po-
tential energy is given by the surface of the sphere times the
Cauchy stress vector times the enlargement of the radius, so
according to (4.18) we acquire
dEr = 4πr2 (σn)ndr
= 8πµr2 ln(λ4 )dr
r
= 32πµr2 ln dr
r0
= 32πµr2 (ln r − ln r0 ). (4.38)
The whole energy going into the sphere if it is inflated or deflated
from an undeformed state (or previous uniformly compressed or
expanded state) with radius r0 to its final radius r is given by
the definite integral
Z r
Er =32πµ (r2 ln r − r2 ln r0 )dr
r0
h r3 r3 r3 ir
=32πµ ln r − − ln r0
3 9 3 r0
h r3 r 3
r 3
r 3
=32πµ ln r − 0 ln r0 − + 0
3 3 9 9
r3 r0 3 i
− ln r0 + ln r0
3 3
184 Spatial Continuum Mechanics

32πµr3 h r 3i
= 3 ln r − 3 ln r0 + 1 − 03
9 r
8µ h  r 3
0
 r 3 i
0
= Vr 1 − − ln
3 r r
Er 8µ  −1 −1
= J − 1 − ln J ], (4.39)
Vr 3
where Vr is the volume of the deformed sphere. If we define the
compression of space as D = J −1 = dV0 /dV we finally get for
the energy density of the compressed space measured by Euler
or spatial coordinates
8µ  
es = (D − 1) − ln D . (4.40)
3
If we chose to postulate that ρ is a function of the energy density
of space, then it is given by
8µ  
ρs = 2
(D − 1) − ln D . (4.41)

By (4.41) and (4.28) the speed of transversal waves in a space


with varying compression is given by
 1/2  −1/2
3
c2 (D) = κ 8 (D − 1) − ln D . (4.42)

We notice that the wave-speed only is depending on the constant


κ and the compression of the spatial continuum.
Next let us consider small variations of the waves speed c in an
area of space that initially is compressed by a factor D0 . This
Inertia and the speed of waves as a function of the compression of
space 185

can be achieved by the linearization of c with respect to D which


is defined by


L c2 (D, ∆D) = c(D) + ∆c(D, ∆D) ,

c2 =c, D=D0

where the second term on the right-hand side is the Gateaux-


derivative defined by (4.31).
Thus we obtain from (4.42) by means of (4.31)

∆c2 (D, ∆D)


 3  12 d h i− 21
=κ (D + ε∆D − 1) − ln(D + ε∆D)
8 dε
c D−1 ∆D
=−
2 (D − 1) − ln D D
c ln D −1
= 1− div (∆u). (4.43)
2 D−1
Finally we acquire for variations of c2 as a function of small
deformations u = ∆u in a space that initially is compressed by
a factor D corresponding to a wave speed of c

c ln D −1
c2 = c + 1− div u, (4.44)
2 D−1
and the gradient of c is given by

grad c2 = α c grad div u, (4.45)


186 Spatial Continuum Mechanics

where
1 ln D −1
α= 1− , (4.46)
2 D−1
or by (4.40)

4(D − 1)µ
α= . (4.47)
3es
By the spatial approach the gradient of the wave speed is pro-
portional to grad div u as it was by the classical approach only
differing from that by a somewhat smaller constant depending
on the initial compression of space.
Chapter 5

Confined Energy and


Gravity

An elastic continuum might be compressed and then released in


a controlled manner such that it undergoes a continuous expan-
sion. In this paper I will show that confined disturbance energy
in such a space of infinite or nearly infinite extension will cre-
ate around itself a gradually decreasing compression field that
implies a gradually increasing propagation speed of transversal
waves with distance from the energy packet. A small test energy
packet in the vicinity of the confined energy will be accelerated
towards the confined energy, or alternatively feel a pull from it
if it is hindered from moving freely in space. In this way pack-
188 Confined Energy and Gravity

ages of confined disturbance energy will influence each other


and move around exactly like masses do under the influence of
gravity in open space.

5.1 Confined energy in the spatial


continuum
In this section I will assume that disturbance energy may be
confined in distinct regions of space where the energy for some
reason is hindered from spreading in space. First I will show
that such confined energy will displace a certain amount of the
spatial continuum regardless of how it is distributed. How it
comes that it is confined will not be discussed in this paper, but
if it is restricted to a certain area of space, it will exert an out-
ward directed pressure that is counteracted by an equally strong
grad div u-field. Next I will show that small, almost point-like,
energy packets in an expanding continuum will create a scalar
div u-field in the environment where they are situated.
In another paper, Elastodynamics in a Continuum of Infinite
Extention, I made a comparison between the Navier-Cauchy
equation1 ,

(λs + 2µs ) grad div u − µs curl curl u + b = ρs ü, (5.1)


1 In order to distinguish the constants of the N-C equation from those

in the Maxwell equations, the index s is used to indicate that the property
refers to the spatial continuum, i.e. ρs is defined as the mass density of the
spatial continuum, and λs and µs are Lamé’s elastic constants.
Confined energy in the spatial continuum 189

and Maxwell’s electrodynamic equations. Provided that sinks


and sources can be realized in a spatial continuum of infinite ex-
tension, I found a complete match between the two sets of equa-
tions. The wave speed for transversal and longitudinal waves
were found to be
µs
r
c= , (5.2)
ρs

and
s
λs + 2µs
c1 = , (5.3)
ρs

respectively. I also found that confined disturbance energy gov-


erned by the divergence-free part of the Navier-Cauchy equation,
creates a second order tensor T, the stress energy tensor, given
by
 
e Sx /c Sy /c Sz /c
 Sx /c −σxx −σxy −σxz 
Tαβ =   Sy /c −σyx −σyy −σyz  ,
 (5.4)
Sz /c −σzx −σzy −σzz

The condition that confined disturbance energy, that is suffi-


ciently smoothly distributed in space, will exert forces on what-
ever keeps it at bay. Here I will assume that there might be
conditions in space, which cause the energy to be kept inside
”bodies” that generally keep their form unchanged over a rea-
sonable span of time. Hence the forces have got to be taken
190 Confined Energy and Gravity

up by the continuum itself, and the principle of conservation of


energy is covered by the condition ∇ · T = 0. This condition is
fulfilled if the trace of T is like zero so
e = σxx + σyy + σzz .
It follows that in an isotropic stress field with σxx = σyy = σzz ,
we acquire
σxx = σyy = σzz = 13 e. (5.5)

Assume that isotropic radiation energy in the spatial continuum


is present as a function of position alone, but not of time, i.e.
u = u(x, y, z). The energy may then vary from place to place
and the resulting radiation pressure will exert a body force b on
an infinitesimal volume element of space. The body force in the
x-direction can be found by taking the difference of the forces
that act on the two opposite surfaces (dy · dz) at (x, y, z) and
(x + dx, y, z)
bx · dx · dy · dz = 31 e(x, y, z) · dy · dz − 13 e(x + dx, y, z) · dy · dz
1 e(x + dx, y, z) − e(x, y, z)
bx = −
3 dx
1 ∂e
=− . (5.6)
3 ∂x
This expression can be expanded to imply the two other spatial
directions, and we get
1  ∂e ∂e ∂e 
bx i + by j + bz k = − i+ j+ k
3 ∂x ∂y ∂z
b = − 13 grad e. (5.7)
Confined energy in the spatial continuum 191

Thus confined isotropic radiation that is a function of position


alone, will act as a body force which can be inserted into the
Navier-Cauchy equation in order to find the deformation it im-
poses on the spatial continuum. Since the force is expressed as
the gradient of a potential, it cannot cause any rotational defor-
mation, and moreover since the energy distribution is presumed
to be stationary, we can use the irrotational part of Navier-
Cauchy equation (5.1) with ü set to zero to find the deformation
1
(λs + 2µs ) grad div u = 3 grad e. (5.8)

If we choose to set the displacement vector, u, like zero in the


undeformed space, i.e. the space outside the confined energy,
then both div u and e are falling down to zero outside the vol-
ume, V , were the radiation energy is confined, so Equation (5.8)
can be solved
e
div u = . (5.9)
3(λs + 2µs )

The physical meaning of (5.8) is that the outwards directed force


caused by the confined radiation energy, e, is counteracted by
an inwards directed force caused by the gradient of the displace-
ment. The grad div u-field represents the force that keeps the
radiation at bay. The deduction above, however, does not prove
that the radiation has got to be confined (the reason for that
must be sought elsewhere), only that if there is a certain amount
of confined energy, then there has got to be a displacement that
sets up an exact balance of forces between the expanding force
of the confined energy and the gradient of the pressure in space.
192 Confined Energy and Gravity

Moreover, from the divergence theorem


I Z
(A · n) df = div A dV (5.10)
V V

we can infer that the displacement is independent of how densely


the energy is distributed in space; a certain amount of con-
fined energy will always displace the same amount of the spatial
continuum regardless of how it is distributed. Hence the total
displacement from a volume of space where an amount of dis-
turbance energy, E, is confined is given by

E
D= , (5.11)
3(λs + 2µs )

or conversely

E = 3(λs + 2µs )D. (5.12)

5.2 The expanding spatial continuum


The spatial continuum can be considered to behave like being
confined in a huge spherical container with receding walls such
that the space it occupies expands in all directions. Huge in
this context is taken to mean that the receding boundaries are
so far away from the center of the sphere, S, that no signal can
reach out to the boundaries during the course of the observation.
Thus if the expansion is uniform, the speed with which a spatial
The expanding spatial continuum 193

element is moving is given by the relation

v(r) = Hr or u̇ = Hr (5.13)

where H is a very small but not necessarily constant coefficient,


and r is the radius vector from S to the spatial element.

Figure 5.1: Velocities in expanding space.

Now, let us see what the space is looking like when considered
from another origin, O, that is drifting along with the space
around it at a distance a from the center of the sphere, S (see
Figure 5.1). An arbitrary point P is located at a distance r from
194 Confined Energy and Gravity

S and a distance r′ from O. We have

r = a + r′ .

Seen from S the point is moving with a speed Hr, and seen from
the new origin, O, it is moving with speed Hr minus the speed
with which the origin itself is moving, namely Ha, so

v′ = Hr − Ha
= H(r − a)
= Hr′ = ṙ′ . (5.14)

If an observer at O cannot see the border of space, he will have


no possibility whatsoever to decide that he himself is moving
with a speed vo , so according to (5.14), the space around him
will be looking like it is uniformly expanding in all direction
from himself.
I shall assume that the speed, v ′ , is constant so the distance
between the two points is given by

r′ = v ′ T, (5.15)

where T is the time elapsed since the two points were close to
each other. From (5.14) and (5.15) we can eliminate the distance
r′ between the points and the receding speed v ′ , and acquire

1
H= . (5.16)
T
Confined energy in expanding space 195

The time derivative of H is


1
Ḣ = − = −H 2 , (5.17)
T2
so if the time elapsed from the points were close together is great
compared with the time of observation, the value of H can be
considered to be constant.
By (5.14) and (5.17) we see that the acceleration of any point
of the spatial continuum is nil as expected:

v̇′ = H r˙′ + Ḣr′


= H · Hr′ − H 2 r′
= 0.

There are no dynamic forces at work in empty space.

5.3 Confined energy in expanding


space
Let a small amount of energy E be evenly distributed within
a small sphere with radius, r. According to Equation (5.5) the
pressure on the inside of the sphere is given by

p = 13 e.

In an expanding space the surface will recede with a speed H ·


r, and hence the confined energy within the sphere will loose
196 Confined Energy and Gravity

energy at a pace given by the force on the spherical surface


times the velocity

∂E
= − 13 e · 4πr2 · Hr
∂t
= − 43 πr3 e · H
= −E · H. (5.18)

Now, let such energy packets be sufficiently smoothly dis-


tributed throughout space. Then the mean energy density in
space is given by
m
1 X
e = lim En ,
V →ǫ V
n=1

where En is the energy in each cell and m is the number of cells


within a volume, V , that is decreasing towards a small volume
ǫ, which still is great enough to contain many cells. The spatial
volume containing the energy packets is loosing energy at a pace
given by
m
∂e 1 X ∂En
= lim
∂t V →ǫ V
n=1
∂t
m
1 X
= lim −En · H,
V →ǫ V
n=1
∂e
= −e · H. (5.19)
∂t
Confined energy in expanding space 197

We can now differentiate e once more with respect on time to


acquire
∂2e
= −ėH − eḢ. (5.20)
∂t2
By (5.19) and (5.17) this expression reduces to

∂2e
= 2eH 2 .
∂t2
Combined with (5.9) it yields

∂ 2 ( div u) ∂2 h e i
=
∂t2 ∂t2 3(λs + 2µs )
2
∂ ( div u) 2eH 2
= . (5.21)
∂t2 3(λs + 2µs )

In space outside of the confined energy there is no body force,


but instead we have got to take into account the dynamic force
ρs ü, and since space in this area is supposed to be free of any
rotation we can without loss of generality take the divergence
of Navier-Cauchy equation (5.1)

∂ 2 ( div u)
(λs + 2µs ) div grad ( div u) − ρs = 0,
∂t2
and insert the result from (5.21) to acquire

2ρs eH 2
∇2 ( div u) = . (5.22)
3(λs + 2µs )2
198 Confined Energy and Gravity

We first introduce a new property ̺(x) (without an index) that


is related to e(x) through the relation
def
e(x) = ̺(x)c2 ,
and insert it into the above equation
2ρs H 2 c2
∇2 ( div u) = ̺. (5.23)
3(λs + 2µs )2
Next we define a new variable Φ(x) and a temporary constant
a that can be chosen at will later
def
Φ(x) = −a div u, (5.24)
and let us also draw together several properties above to a new
property G defined by
def aµs H 2
G = . (5.25)
6π(λs + 2µs )2
The propagation speed for transversal waves is given by c2 =
µs /ρs , and with this property inserted we end up with
∇2 Φ = −4πG̺. (5.26)

We notice that G can be interpreted as the gravitational con-


stant2 , Φ as the gravitational potential, and ̺ as the mass den-
sity of the confined energy. Then Equation (5.26) is an exact
2 In fact G is not a constant, but a function of H. Both are, however,

functions of T −1 where T is supposed to be very great, hence the variation


of G in the course of any time of observation is extremely small.
Confined energy in expanding space 199

match of the classical equation for Newton potential. This Pois-


son’s equation can be solved under rather general condition, so
if ̺(r′ ) is a function of the radius vector r′ , then Φ(r) at the
position r is given by the well known relation

̺(r′ )dv ′
Z
Φ(r) = G ,
|r − r′ |

where dv ′ is a small volume element around the position r′ and


the integration is taken over the whole space. If the total mass
equivalence is defined as
Z
def
M = ̺ dV,
V

and is concentrated in a small area around r′ = 0, then the


potential at a distance r from it is given by

GM
Φ= .
r
The gradient of Φ is given by

GM r
grad Φ = − , (5.27)
r2 |r|

or for later use by reinserting div u from (5.24) we get

1 GM r
grad div u = . (5.28)
a r2 |r|
200 Confined Energy and Gravity

5.4 Movements of energy packets in a


space with varying wave speeds
In this section I will show that disturbance energy that moves
along filiform paths with velocity c within closed packages, will
be accelerated towards bodies of confined energy, or else acted
upon by forces like forces in a gravitational field if they are
hindered from moving freely in space.
In my paper Elastodynamics in in a continuum of infinite ex-
tension I discussed the property of momentum in the spatial
continuum. I found that momentum is related to energy trans-
port in solenoidal fields given by

S = e c.

The time derivative of the property

S
p=
c2
produces a force

∂p
f= .
∂t
Hence p can be interpreted as the momentum of the energy flow.
In another paper Standing waves between singularities in an
elastic continuum of infinite extension I found that standing
waves may organize between singularities along strings in the
Movements of energy packets in a space with varying wave speeds
201

spatial continuum. Standing waves may be seen as the superpo-


sition of progressive waves moving in opposite directions along
the string. Here I will assume that elementary material particles
are comparable to energy packets moving with the speed of c
along such strings, and that the momentum of each particle is
like
En
pn = c,
c2
where pn is the momentum of particle number n in a collection
of m particles making up a ”material body”. A further discus-
sion of this assumption will not be performed in this paper, but
has to be to be done in another connection.
Consider a test body of the type assumed above situated in a
space where c may be a function of position, but not of time.
We take the time derivative of pn , and by applying the chain
rule and the directional derivative along the space curve of c,
we acquire

dpn d  En  En dcn ds
= cn = 2 .
dt dt c2 c ds dt
The term ds/dt is simply like c, and dcn /ds is the directional
derivative of cn in the direction of ĉn , so we obtain

dpn En En
= − 2 · (ĉn ∇)cn · c = − 2 · (cn ∇)cn .
dt c c
By the mathematical identity grad (A·B) = (A∇)B+(B∇)A+
202 Confined Energy and Gravity

A × curl B + B × curl A, we obtain


dpn En
= − 2 12 grad c2 − cn × curl cn

dt c
En 
= − 2 c grad c − cn × curl cn .
c

We can now sum all the time derivatives of pn and obtain


m m m
d X c grad c X 1 X
pn = En − En (cn × curl cn )
dt n=1 c2 n=1 c2 n=1

The last terms, concerning the interior movements of the energy


packets within the body, are in all possible directions and will
cancel out, and we finally acquire
dP
= M c grad c, (5.29)
dt
where P is the resulting momentum of all the packages and
def
M = E/c2 can be dubbed the mass of the body.
If the body moves through space with the velocity v, then the
mean energy flow density within the body is given by

S = En cn = ev.

Hence the mean momentum density, pn , in the body is given by

S En cn e
pn = = = 2 v.
c2 c2 c
Movements of energy packets in a space with varying wave speeds
203

It follows that the momentum of the whole body is

P = M v.

The time derivative of P is also given by

dP dv
=M = M g,
dt dt
where g is the acceleration of the body. If we compare this
expression with that in Equation (5.29) we see that the acceler-
ation of the body is given by

g = c grad c.

We can conclude that a body situated in a field with varying


c gets an acceleration, which we can call the acceleration of
gravity, that is independent of the mass of the body. Note also
that neither the energy nor the mass of the body is changing
when it accelerates in a grad c-field, hence there are no external
forces acting on the body. It is the confined energy that changes
from a more chaotic movement to getting an ordered component
in the direction of g. Newton’s force of gravity is non existing in
open space, it only occurs when a body is hindered from moving
freely, but then it has to be kept back by a force that equals the
force needed to give it an acceleration like g. The force is given
by

F = M g. (5.30)
204 Confined Energy and Gravity

5.5 Newton’s Gravitational law


In this final section I will show that two separate concentrations
of confined energy will attract each other by the same law as
that discovered by Newton some 300 years ago, and finally es-
timate what values µs and ρs should have in order to make the
comparison feasible.
First we have got to find out how the wave speed, c, varies
with compression of the spatial continuum. The question is not
quite trivial. In another paper, Spatial continuum mechanics,
I have addressed this question from two points of view. One
that the spatial continuum has an intrinsic mass density that
only changes when a volume element is compressed to a smaller
volume, and the other that the spatial continuum has no initial
mass density at all and only gets its inertial properties by the
energy that goes into it by compression. Here I will only consider
the first alternative.
Let a volume element, V0 , be changed to a volume V = V0 +∆V .
Then we can set up the relation:

ρs V0 = ρV,
1 1 V0 + ∆V
= ,
ρ ρs V0

and multiply both sides with the property µs

µs µs ∆V 
= 1+ .
ρ ρs V0
Newton’s Gravitational law 205

Then by the identity

∆V
lim = div u,
V0 →0 V0

we get the variation of c as a function of div u

c2 = c02 (1 + div u),

and finally by taking the divergence on both sides of the equa-


tion, we obtain

c grad c = α c02 grad div u, (5.31)

where α by this approach is like 12 . Another approach where ρs


is assumed to be proportional to the energy density caused by
compression of the spatial continuum, yields an α that is greater
depending on the grade of compression.
Now let us introduce a mass M1 into space, which generates a
potential given by Equation (5.27)

GM1 r
grad Φ = − .
r2 |r|

In this field at a distance r we introduce another test mass, M2 ,


which is kept from moving and hence be acted upon with a force
given by (5.30)

F21 = −M2 · c grad c.


206 Confined Energy and Gravity

Note that the force is acting on M2 and the radius vector is from
M1 to M2 , hence the negative sign. From (5.31) we get c grad c,
and from (5.28) we get the value of grad div u, and with these
two properties inserted the mutual attraction between the two
bodies above becomes
α c02 GM1 M2
F =
a r2

Now let the temporary constant a (5.24) be given by a = α c02 .


Then the equation above finally reduces to
M1 M2
F =G ,
r2
and according to (5.25) the value of G becomes

α c02 H 2 µs
G= · . (5.32)
6π (λs + 2µs )2

This is Newton’s equation for the attraction between two heav-


enly bodies and needs no additional comments.
It is noteworthy that confined wave energy will displace a pro-
portional amount of the spatial continuum regardless how it
is distributed. Hence the displacement is also a measure of the
amount of confined energy. It is also noteworthy that it is only in
an expanding space that there can be any density gradient that
can be associated with a gravitational field. In a non-expanding
space grad div u ≡ 0 outside any energy concentrations, and no
fields that can be resembled with gravity can occur.
A numerical comparison 207

5.6 A numerical comparison


In this final section I will try to estimate which approximate
values the fundamental constants µs and λs should have in order
to make the defined spatial continuum comparable to empty
space. To make Equation (5.32) simpler, I will estimate the
value of λs to 2µs , and α to 12 . This will certainly keep G
within its numerical order of magnitude. With λs ≈ 2µs and
α ≈ 12 inserted into Equation (5.32), µs becomes

c2 H 2
µs ≈ ,
192πG
and accordingly since c2 = µs /ρs

H2
ρs ≈ .
192πG

Newton’s gravitational constant G = 6.673 × 10−8 cm3 /g sec2 .


The Hubble Space Telescope Key Project Team has measured
Hubble’s constant with an uncertainty of 10 percent to3 H =
70 km s−1 Mpc−1 = 2.3 × 10−18 sec−1 . The speed of light c =
3.0 × 1010 cm/sec. With these values inserted the spatial mass
density amounts to
3
ρs ≈ 1.3 × 10−31 g/cm ,
3 In a press release from May 9. 1996 at http:
oposite.stsci.edu/pubinfo/press-releases/96-21.txt the most probable
value is reported to be between 68 and 78 km s−1 Mpc−1 .
208 Confined Energy and Gravity

which is fairly close to the estimated mean spatial density of


matter in the universe4 , and finally Lamé’s elastic constant

µs ≈ 1.2 · 10−10 g/sec2 cm. (5.33)


Bibliography

[1] Yavuz Başar and Dieter Weichert. Nonlinear Continuum


Mechanics of Solids. Springer, 1999.

[2] Richard P. Feynman, Robert B. Leighton, and Matthew


Sands. The Feynman Lectures on Physics.
Addison-Wesley, 1977.

[3] S. Flűgge, editor. Mechanics of Solids II, volume VIa/2 of


Encyclopedia of Physics. Springer, 1972.

[4] Egil Hylleraas. Matematisk og teoretisk fysikk. Grøndal &


Søns forlag, 1950.

[5] Erwin Kreyszig. Advanced Engineering Mathematics.


JOHN WILEY and SONS, INC, 8 edition, 1999.

[6] Lawrence E. Malvern. Introduction to the Mechanics of a


Continuous Medium. Prentice-Hall, 1969.
210 BIBLIOGRAPHY

[7] D. H. Menzel, editor. Fundamental Formulas of Physics.


Dover Publications, 1960.
[8] Charles W. Misner, Kip S. Thorne, and John Archibald
Wheeler. GRAVITATION. W. H. Freeman and Company,
1972.
[9] Seth J. Putterman. Sonoluminescence: Sound into Light.
Scientific American, 272:32–37, February 1995.
[10] Sir Edmund Whittaker. A History of the Theories of
Aether and Electricity, volume I and II. Philosophical
Library, 1951.
[11] Martin V. Zombeck. Handbook of Space, Astronomy and
Astrophysics. Cambridge University Press, 1982.
Index

acceleration, 83 curved space-time, 75


actual stress, 85
deformation gradient, 163
Big Bang, 33, 41 density, 83
body force, 84, 190 displacement field, 83
Bohm, David, 61 displacement vector, 162
bubble, 3, 34, 37, 42 Einstein field equation, 76
Einstein, Albert, 24
Cauchy-Poisson theorem, 84 elastic continuum, 79
chains of oscillating nodes, elasticity tensor, 85
147 elasticity waves, 179
classical approach, 12 electronS, 52
color charge, 67 energy flow vector, 94, 100
confined energy, 25, 192 energy transport, 93
convective coordinate system, entangle, 24, 58, 68
161 equation of motion, 84
Cosmic Background Euler coordinates, 160, 181
Radiation, 40 expanding space., 193
Coulomb gauge, 107
coupled oscillations, 151 Faraday’s law, 13
212 INDEX

fermion, 65 Linear Theory of Elasticity, 79


local energy density, 92, 119
Gauss’s law, 13 local frame, 74
General Relativity, 4, 24 longitudinal wave, 89, 119
geodesic, 75 Lorentz frame, 16
gravitation, 17, 69, 77, 204, Lorenz gauge, 107
206
gravitational constant, 198 mass of radiation, 111
gravitational potential, 3 material coordinates, 160
Matter wave, 64
hadron, 40 matter wave, 59, 64
Helmholtz’s Theorem, 87, 116 Maxwell’s equations, 13
Helmholtz, Hermann von, 18 Maxwell’s stress tensor, 114
hidden dependency, 20, 21, 98 Maxwell, James Clerk, 14
hidden variable, 61 metric tensor, 75, 77
Huygens, Christiaan, 28 Minkowski’s space, 105
hydraulic jump, 39 Minkowski, Hermann, 16
motion, 83
isotropic radiation, 190
Navier, Claude, 17
Kelvin’s theorem, 18, 92, 119 Navier-Cauchy, 88, 117
Kelvin, Lord, 14 Navier-Cauchy equation, 17,
Kelvins cord, 15 86, 89, 116, 119
Kronecker delta, 86 neutral mode, 54, 57
neutrino, 57
Lagrangian coordinates, 160 Newton’s second law, 26
Lamé’s elastic moduli, 86 Newton, Isaac, 26
left stretch tensor, 165 nonlocality, 61
lepton, 40, 65
linear momentum, 83 Occam’s razor, 23
INDEX 213

Occam, William of, 23 sinks, 95


orthogonal, 165 solenoidal standing wave, 140
solitary wave, 39
photon, 49 sonoluminecence, 3
pilot wave, 60, 62 sonoluminescence, 34
Planck’s constant, 45 source, 54
Poisson, Siméon-Denis, 69 sources, 95
polar decomposition theorem, spatial approach, 12
165 spatial continuum, 3, 11, 18,
positive definite, 165 24, 33, 55, 85
positronium, 58 spatial coordinates, 160, 184
postamble, 43, 50, 52 spatial divergence, 181
Poynting’s vector, 113 Special Relativity, 3, 23, 75
preamble, 43, 49, 52 spectral decomposition
principal axes, 166 theorem, 165, 168
propagation theorem, 89, 118 spherical waves, 121
Spin, 49
quarks, 65 spin, 55
Standing waves, 29
radiation pressure, 190 standing waves, 139
relativity, 17, 69, 72 Stokes, George Gabriel, 17
residual stress, 85 stress, 169
Ricci tensor, 77 stress energy tensor, 72, 99,
Riemannian geometry, 75 103, 189
right stretch tensor, 165 stress field, 84
rotation tensor, 165 stress vector, 84
successive deformations, 172
sink, 54
sink density, 19, 21, 97 transversal, 89, 119
214 INDEX

uniform compression, 171


universal frame, 74

vector potential, 106


vectorial line element, 163
velocity, 83

Wave equations, 28

You might also like