You are on page 1of 17

Downloaded from http://rsta.royalsocietypublishing.

org/ on June 26, 2016

Phil. Trans. R. Soc. A (2008) 366, 1717–1733


doi:10.1098/rsta.2007.2181
Published online 25 January 2008

From Maxwell’s theory of Saturn’s rings


to the negative mass instability
B Y R ENATO F EDELE *
Dipartimento di Scienze Fisiche, Universitá Federico II and INFN,
Complesso Universitario di Monte S. Angelo, via Cintia, 80126 Napoli, Italy

The impact of Maxwell’s theory of Saturn’s rings, formulated in Aberdeen ca 1856, is


discussed. One century later, Nielsen, Sessler and Symon formulated a similar theory to
describe the coherent instabilities (in particular, the negative mass instability) exhibited
by a charged particle beam in a high-energy accelerating machine. Extended to systems
of particles where the mutual gravitational attraction is replaced by the electric
repulsion, Maxwell’s approach was the conceptual basis to formulate the kinetic theory
of coherent instability (Vlasov–Maxwell system), which, in particular, predicts the
stabilizing role of the Landau damping. However, Maxwell’s idea was so fertile that, later
on, it was extended to quantum-like models (e.g. thermal wave model), providing the
quantum-like description of coherent instability (Schrödinger–Maxwell system) and its
identification with the modulational instability (MI ). The latter has recently been
formulated for any nonlinear wave propagation governed by the nonlinear Schrödinger
equation, as in the statistical approach to MI (Wigner–Maxwell system). It seems that
the above recent developments may provide a possible feedback to Maxwell’s original
idea with the extension to quantum gravity and cosmology.
Keywords: Saturn’s rings; charged particle beams; negative mass instability;
thermal wave model; coherent instabilities; modulation instability

1. A brief historical note on Saturn’s rings studies

The rings of Saturn are a series of planetary rings orbiting around the planet.
The major rings are labelled. They consist largely of ice and dust. There are
several gaps between the rings, all of which are caused by the gravitational pull
of one or more of Saturn’s moons affecting the orbits of the tiny particles that
comprise the rings. Icy particles spread out into large, flat rings and make up
Saturn’s ring system that can be seen with even low-power telescopes on the
Earth’s surface.
Improved telescopes have allowed astronomers to better ‘resolve’ the images of
the rings. In fact, one of the large rings was resolved into two rings (B and C),
and it was found that the A ring shows a gap (the so-called Encke division). The
presence of another ring (the so-called D ring) closer to the planet than the C
ring was also observed.
*renato.fedele@na.infn.it
One contribution of 20 to a Theme Issue ‘James Clerk Maxwell 150 years on’.

1717 This journal is q 2008 The Royal Society


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1718 R. Fedele

In 1610, Galileo Galilei was the first person to observe Saturn’s rings, although
with his weak telescope he could barely resolve them, and thought they were two
moons on either side of the planet.
After 45 years, in 1655, Christian Huygens was the first scientist to propose
that there was a ring surrounding Saturn (the rings were actually discovered by
him in 1659). He proposed that Saturn was surrounded by a solid ring.
In 1660, the poet Jean Chapelain, a friend of Huygens, suggested that Saturn’s
rings are made up of a large number of very small satellites. Chapelain’s
suggestion did not receive enthusiasm among the astronomers. However, 200
years later, James Clerk Maxwell arrived at similar conclusion.
A valuable discovery of Giovanni Cassini in 1675 showed the existence of a gap
between the two large (A and B) rings, now called the Cassini division in honour
of this astronomer.
During the period between 1856 and 1859, the theoretical studies of Maxwell
showed that the rings could not be solid, but rather a swarm of particles. A solid
ring would become unstable and break up. He carried out a careful theoretical
treatment and concluded that the rings could not be solid or liquid, since the
mechanical forces acting upon rings of such immense size would break them up.
He suggested that, instead, the rings were composed of a vast number of
individual solid particles rotating in separate concentric orbits at different
speeds. He reported this theory in his final article on the subject entitled ‘On the
stability of the motion of Saturn’s rings’, published in the Proceedings of the
Royal Society of Edinburgh in 1859 (Niven 1890).1

2. Some details of Maxwell’s idea about Saturn’s rings

Maxwell supposed that the dust particles move around the planet at the angular
frequency that decreases when the energy increases. Under the action of the
gravitational field of the planet, the dust particles move along closed orbits. But
this is not sufficient to have a stable motion. In fact, what confines the dust is the
mutual gravitational interaction between the particles: this way he proved that
dust forms a stable system forced by Saturn’s gravitational field to execute
revolutions. In 1895, James Keeler proved that Maxwell’s hypothesis was correct
when he measured the Doppler shifts of different parts of the rings and found that
the outer parts of the ring system orbited at a slower speed than the inner parts.
The rings obeyed Kepler’s third law and, therefore, must be made of millions of
tiny bodies each orbiting Saturn as a tiny mini-Moon.
One century later, in the period between 1958 and 1959, an analogous
mechanism was put forward by Nielsen et al. (1959). It was pointed out that
Maxwell’s mechanism would be more interesting if the particles repel each other.
An example of this kind was given by the above authors in particle accelerators
(for an historical review, see Lawson (1988)).
1
Most of the historical information given in this article are available at the following web
sites: http://en.wikipedia.org/wiki/Rings_of_Saturn; www.enchantedlearning.com/subjects/
astronomy/planets/saturn/saturnrings.shtml. Craig Howie, http://heritage.scotsman.com/pro-
files.cfm?cidZ1&idZ39592005 (article). Nick Strobel, www.astronomynotes.com/solarsys/s16.
htm#A9.5.1 (article). Calvin J. Hamilton, www.solarviews.com/eng/saturnbg.htm (article).

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1719

In fact, in a circular accelerating machine, a charged particle beam is forced to


execute millions or billions of revolutions under the action of electromagnetic
(e.m.) forces instead of the gravitational field of the planet.

3. The theory of Nielsen, Sessler and Symon

Due to their electric charges, the particles of the beam repel each other.
Consequently, one could think that in this case, due to the space charge
repulsion, the beam motion is unstable. In reality, the physical conditions in an
accelerating machine are richer than the ones occurring in the mechanism
proposed by Maxwell, because the angular frequency variation is related to the
energy variation through a threshold called the ‘transition energy’. But,
nevertheless, Maxwell’s idea was very useful for Nielsen, Sessler and Symon
and constituted the basic idea of their mechanism. Maxwell’s idea was, in fact,
extended to the collective behaviour of a charged particle system in a circular
accelerating machine (e.g. a storage ring).

(a ) Kinematics of a charged particle beam in a storage ring


Let us consider a relativistic charged particle beam travelling with velocity bc
throughout the pipe of a circular accelerating machine. In order to describe the
kinematics of the beam, it is useful to introduce the concept of a synchronous
particle as a particle that does not change its energy E and the radius R of its circular
orbit. This particle rotates at an angular velocity uZbc/R and has a linear
momentum pZm 0gbc, where m 0 is the particle rest mass and g is the relativistic
Lorentz factor.
It is useful to adopt a co-moving frame of coordinates, whose origin coincides
with the position of the synchronous particle. During motion, an arbitrary
particle of the beam has, in general, the physical quantities displaced with
respect to the ones of the synchronous particle. Note that
Du Db DR
u
Z
b
K
R ð3:1Þ
and
Dp Dg Db
Z C : ð3:2Þ
p g b
Furthermore, it is easy to see that
Dg
Z g2 bDb ð3:3Þ
g
and
Dp 1 DE
Z 2 : ð3:4Þ
p b E
To maintain an arbitrary particle with a given linear momentum on a circular
orbit of radius R, an axial magnetic field with amplitude BZB(R) is necessary.
Consequently,
 
Dp R dB DR
Z 1C : ð3:5Þ
p B dR R

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1720 R. Fedele

Then, by combining the previous relationships, we easily get


 
Du 1 1 DE
Z 2
K ac ; ð3:6Þ
u g b2 E
where we have introduced the ‘momentum compaction’ as
 
DR=R R dB K1
ac Z Z 1C : ð3:7Þ
Dp=p B dR
By introducing the ‘transition energy’, i.e.
1
gT Z ; ð3:8Þ
ac
and the ‘slip factor’, i.e.
1 1 1
hZ 2
K ac h 2 K 2 ; ð3:9Þ
g g gT
equation (3.6) can be finally cast as
Du h DE
Z 2 : ð3:10Þ
u b E
Equation (3.10) relates the energy displacement to the angular frequency
displacement of an arbitrary particle with respect to the synchronous one.
Therefore, it describes the kinematics of the longitudinal motion of an arbitrary
particle of the beam and has to be coupled with the ones describing the dynamics
of the system. Furthermore, from equation (3.10) we observe that: (i) for h!0
(above transition energy) the angular frequency decreases as the energy increases
and (ii) for hO0 (below transition energy) the angular frequency increases as the
energy decreases.
Let us now describe the longitudinal dynamics associated with an arbitrary
particle of the beam.

(b ) Dynamics of a charged particle beam in a storage ring


Owing to the interaction with the surroundings, the beam space charge
(capacitive effect) is in competition with the magnetic inductive effect. In order
to describe the interaction between the beam and the surroundings, let us
consider a beam propagating through the accelerator pipe and denote with l the
longitudinal number density (number of particles per unit length along the
propagation direction). In general, l depends on both the transverse and
longitudinal coordinates and time. For the sake of simplicity, let us consider
the one-dimensional case in which the beam is considered transversally
flat and therefore l depends only on the longitudinal coordinate z and time t,
i.e. lZl(z, t). Correspondingly, the current associated with the beam is
I(z, t)Zqbcl(z, t), where q is the particle charge.
During its motion, the beam is a source of both radial electric and azimuthal
magnetic fields. Furthermore, owing to its interaction with the surroundings
through the image charges and image currents (mainly originated on the pipe
walls), a self-interaction (collective force) is produced.

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1721

Table 1. Coherent instability scheme of a monochromatic ½ f 0 ðpÞf dðpÞ coasting beam in the case
of a purely reactive impedance Z I h XL K XC . (The instability corresponding to h!0 is usually
referred to as ‘negative mass instability’.)

Z IO0 (inductive) Z I!0 (capacitive)

hO0 (below transition energy) unstable stable


h!0 (above transition energy) stable unstable

In the case of perfectly conducting walls and coasting beam, the Lorentz–
Maxwell system of equations easily leads to the following expression for the total
longitudinal force per unit length acting on the beam (Lawson 1988)
vl
F Z k 0 ðXL K XC Þ ; ð3:11Þ
vz
where k 0 is a positive constant proportional to the square of q and XL and XC are
the inductive reactance (magnetic effect of the beam current) and capacitive
reactance (space charge effect) per unit length, respectively. Note that F does not
depend on the sign of the particle charge.

(c ) The physical mechanism of the instability


Physically, particles with a speed very close to c cannot travel much faster by
acceleration but can increase their momentum and consequently move with a
larger radius of curvature. Above the transition energy, particles that normally
repel each other seem to experience an attractive force. This is known as the
negative mass effect just for its similarity with the formation of planetary rings
with an otherwise attractive gravitational force outside the Roche limit. In
general, to have a stable or unstable beam motion more than a combination is
possible (for a detailed explanation, see table 1 and figures 1–3). This way,
Maxwell’s mechanism is the gravitational analogue of the case in which the
system is above the transition energy and the inductive effect overcomes the
space charge effect only. But, of course, when the space charge is dominating,
the beam particles are paradoxically drawn in the opposite direction towards
which the total force is acting, like the behaviour of particles with negative mass.

4. Subsequent developments: kinetic theory of coherent instabilities


(Vlasov–Maxwell system)

To provide a more general description of the coherent instability, Boltzmann’s


kinetic theory is applied. For a collisionless charged particle beam, the Boltzmann
transport equation (Lawson 1988) has no usual collision integral. However, taking
into account the two-particle correlation effects due to the e.m. interactions, a
macroscopic collective Lorentz force appears in the Boltzmann equation (mean field
approximation). Such an equation is usually known as the Vlasov equation, widely
used in plasma physics (Nicholson 1983). Thus, for a transversally flat beam, the
Vlasov equation, governing its longitudinal phase–space evolution, has to be
coupled with Maxwell’s equations for the self-consistent e.m. fields. (The set of

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1722 R. Fedele

(a) (b)

Figure 1. Longitudinal perturbed density of a coasting beam: (a) periodic density perturbation
plotted with respect to the longitudinal direction and (b) periodic density perturbation as seen
around the orbit of the synchronous particle.

(a) <0 >0 (b) >0 <0


<0 >0 >0 <0

0 0

Figure 2. Physical mechanism of the coherent instability of a coasting beam below transition
energy (hO0): (a) for a capacitive impedance, the force is acting longitudinally towards the minima
of the perturbation, but it accelerates (decelerates) the particles located at the front (back) of the
maxima and therefore stabilizes the system and (b) for an inductive impedance, the force is acting
longitudinally towards the maxima but it decelerates (accelerates) the particles located at the front
(back) of the perturbation and therefore destabilizes the system.

equations comprising the Vlasov equation and Maxwell’s equations, the so-called
Vlasov–Maxwell system, govern the motion of the particles under the action of the
e.m. fields that they actually generate.)
By introducing the first-order perturbations in the physical quantities
(e.g. distribution function and e.m. fields) the Fourier analysis of the Vlasov–
Maxwell system in the case of a coasting beam leads to the following dispersion
relation (Chao 1993):
  ðN
Z df0 =dp
1 Z ia0 h dp; ð4:1Þ
k KN pKu=k
where fZf0(p) is the equilibrium momentum distribution ( p being the longitudinal
conjugate momentum coordinate); Z is the complex longitudinal coupling
impedance of the system; and a0 is a positive real constant. To obtain equation
(4.1), the first-order perturbations of the physical quantities have been assumed to
be proportional to exp[i(kzKut)]. The real part of Z accounts for the resistive
properties of the walls of the machine’s vacuum chamber. The inclusion of the

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1723

(a) <0 >0 (b) >0 <0


>0 <0 <0 >0

0 0

Figure 3. Physical mechanism of the coherent instability of a coasting beam above transition
energy (h!0): (a) for a capacitive impedance, the force is acting longitudinally towards the minima
of the perturbation, but it decelerates (accelerates) the particles located at the front (back) of the
maxima and therefore destabilizes the system and (b) for an inductive impedance, the force is
acting longitudinally towards the maxima, but it accelerates (decelerates) the particles located at
the front (back) of the perturbation and therefore stabilizes the system.

resistive effects extends the instability analysis to an arbitrary coupling impedance


presented above for a purely reactive coupling impedance only. By denoting with
Z R and Z I the real and imaginary parts of Z, respectively, the dispersion relation
(4.1) can be cast in the following way:
ð  K1
df0 =dp
VR C iVI Z i dpGip df0 =dpðbph Þ ; ð4:2Þ
PV pK bph
Ð
where VR ZKa0 hZ R =k, VI ZKa0 hZ I =k, bph Z u=k and the symbol PV stands for
the Cauchy principal value. This equation determines a relationship between VR,
VI and bph. In principle, bph is a complex quantity. Thus, we put: bph h gR C igI .
Consequently, we can plot curves in the VRKVI plane for a given equilibrium
distribution function f0(p) and for different growth rates gI. A qualitative
behaviour of these plots is given in figure 4, where f0 ðpÞf ð1Kp2=4Þ2 is assumed.
This picture describes the coherent instability (for instance, the negative mass
instability) in circular accelerating machines in which coherent instabilities
compete with the stabilizing effect of Landau damping. We would like to stress
that figure 4 represents a sort of universal stability chart predicted by the
Vlasov–Maxwell system. Any impedance Z leading to a (VR, VI) pair belonging
to the area surrounded by the curve with gIZ0 corresponds to a stable operation.
Furthermore, in the limit of a monochromatic beam, i.e. f0 ðpÞf dðpÞ, all the
curves become parabolas and the stability region collapses into the positive part
of the imaginary axis. In this limit, the case Z RZ0 recovers the instability table
(table 1) of a coasting beam.

5. Subsequent developments: quantum-like theory of coherent


instability (Schrödinger–Maxwell system)

(a ) The thermal wave model


The description of the coherent instabilities given above can be alternatively
provided by the quantum formalism (quantum-like description, see Fedele &
Shukla (1995)). The charged particle beam dynamics has been successfully
described by a quantum formalism in a number of problems of particle

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1724 R. Fedele

(a) VI (b)
I1 I2 I3
I

I=0

–1 0 1 VR
0

Figure 4. Qualitative instability chart in the VR–VI plane, as described by the Vlasov–Maxwell
system, for a non-monochromatic coasting beam. (a) The region inside the curve with gIZ0 plus
the adjacent straight line located above on the imaginary axis represents the stability region. For
increasing values of gI (gIO0), the curves plotted cover the entire instability region. They are a
sort of ‘deformed parabolas’. However, as gI increases, their shape becomes more similar to the
parabolas. The deformation of the parabolas is due to the stabilizing effect provided by the Landau
damping that exists for a p-distribution with non-negligible spread. (b) Qualitative plot of
f0(p) f(1Kp2/4)2 corresponding to the chart displayed in (a ).

accelerators by means of the so-called thermal wave model (TWM; Fedele &
Miele 1991). In particular, within the TWM framework, the longitudinal
dynamics of particle bunches is described in terms of a complex wave function,
J(z, s), where s is the distance of propagation (i.e. sZct) and z is the longitudinal
extension of the particle beam measured in the moving frame of reference.
The particle density, l(z, s), is related to the wave function according to
l(z, s)ZjJ(z, s)j2 (Fedele et al. 1993; Anderson et al. 1999a) and the collective
longitudinal evolution of the beam in a circular high-energy accelerating machine
is governed by the following Schrödinger-like equation for J, viz.

vJ e2 h v2 J
ie C KU ðz; sÞJ Z 0; ð5:1Þ
vs 2 vz 2

where e is the longitudinal beam emittance and U(z, s) is a dimensionless effective


potential energy. Note that here 1/h plays the role of an effective mass associated
with the beam as a whole. In general, equation (5.1) has to be coupled with the
e.m. fields (via Maxwell’s equations) that allows one to determine U (Schrödinger–
Maxwell system). In particular, U is related to the self-consistent e.m. fields
accounting for the collective interaction between the beam and the surroundings
and, consequently, to the longitudinal coupling impedance Z. Let us assume that
no external sources of e.m. fields are present and the effects of charged particle
radiation damping are negligible. Then, the self-interaction between the beam
and the surroundings, due to the image charges and the image currents originated
on the walls of the vacuum chamber, makes U a functional of the beam density.
It can proven that, for a coasting beam of an unperturbed density l0 travelling in
a high-energy circular accelerating machine, the governing equation (5.1) can be

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1725

cast as the following integro-differential equation (Anderson et al. 1999a):


ðx
vJ a v2 J  2   
i C 2
2
C X jJj KjJ0 j J C RJ jJðx 0 ; sÞj2 KjJ0 j2 dx 0 Z 0; ð5:2Þ
vs 2 vx 0

where X f Z I , Rf Z R , aZeh and J0 is a complex constant such that l0ZjJ0j2.


Equation (5.2) belongs to the family of nonlinear Schrödinger (NLS) equations
governing the propagation and dynamics of wave packets in the presence of non-
local effects. The modulational instability (MI) of such an integro-differential
equation has been investigated for the first time in literature by Anderson et al.
(1999a). Some non-local effects associated with the collective particle beam
dynamics have been recently described with this equation (Johannisson et al.
2004). Note that, by comparing equation (5.1) with equation (5.2), U is now
expressed by the functional
 ðz 
 2 2
  0
 0
U ZKa X jJj KjJ0 j C R jJðz ; sÞj KjJ0 j dz h U ½jJj2 :
2 2
ð5:3Þ
0

(b ) MI of a monochromatic coasting beam and its identification


with coherent instability
In order to provide a suitable quantum-like description of the coherent
instability, we cast equation (5.2) in the pair of Madelung fluid equations
(Madelung 1926). To this end, let us assume
 
pffiffiffiffiffiffiffiffiffiffiffiffiffi i
J Z lðz; sÞ exp Qðz; sÞ ; ð5:4Þ
a
then substitute (5.4) in (5.2). After separating the real from the imaginary parts,
we get the following representation of (5.2) in terms of a pair of coupled fluid
equations (continuity and motion, respectively):
vl v
C ðlV Þ Z 0; ð5:5Þ
vs vz
  " #
v v vU a2 v 1 v2 l1=2
CV V ZK C ; ð5:6Þ
vs vz vz 2 vz l1=2 vz 2

where the current velocity V is given by


vQðz; sÞ
V ðz; sÞ Z: ð5:7Þ
vz
Under the conditions assumed above, let us consider a monochromatic coasting
beam travelling in a circular high-energy machine with an unperturbed velocity
V0 and an unperturbed density l0ZjJ0j2 (equilibrium state). In these conditions,
all the particles of the beam have the same velocity and their collective
interaction with the surroundings is absent. Within the Madelung’s fluid
representation, the beam can be thought as a fluid with both current velocity and
density which are uniform and constant (monochromatic beam). In this state,
the Madelung fluid equations (5.5) and (5.6) vanish identically. Let us now

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1726 R. Fedele

>0 ZI
ZR

I3

I1 < I2 < I3

I=0 I2
I1

Figure 5. Qualitative plots of the MI curves in the plane ZR–ZI of a coasting beam below the
transition energy, i.e. hO0 (Anderson et al. 1999a). The bold face vertical straight lines represent
the stability region (uIZ0).

introduce small perturbations in V(z, s) and l(z, s), i.e.


V Z V0 C V1 ; jV1 j/ jV0 j; ð5:8Þ

l Z l0 C l1 ; jl1 j/ l0 : ð5:9Þ
By introducing (5.8) and (5.9) in the pair of equations (5.5) and (5.6), after
linearizing, and assuming that V1, l1 f expðikz K iusÞ, we finally get the
following dispersion relation:

u 2  
Z a2 k 2
K V0 Z ial0 C ; ð5:10Þ
k k 4
where we have introduced the complex quantity Z Z RC ikX h ZR C iZI ,
proportional to the longitudinal coupling impedance per unit length of the
beam. In general, in equation (5.10) u is a complex quantity, i.e. u h uR C iuI . If
uIs0, an instability takes place in the system. It has been recently proven that
such an instability is a sort of MI predicted by the integro-differential NLS
equation (5.2) (Fedele et al. 1993; Anderson et al. 1999a). If now we substitute
the complex form of u in equation (5.10), separating the real from the imaginary
parts, the dispersion relation can be cast as
ekl 1 u2I ek 3
ZI ZKh 20 ZR2 C Ch : ð5:11Þ
4uI h ekl0 4l0
This equation fixes, for any values of the wavenumber k and any values of
the growth rate u I, a relationship between real and imaginary parts of the
longitudinal coupling impedance. For each uIs0, running the values of the
slip factor h, it describes two families of parabolas in the complex plane ZRKZI.
Each pair (ZR, ZI) in this plane represents a working point of the accelerating
machine. Consequently, each parabola is the locus of the working points
associated with a fixed growth rate of the MI. According to figures 5 and 6,
below the transition energy (g!gT), h is positive and therefore the instability
parabolas have a negative concavity, while above the transition energy (gOgT),
since h is negative, the instability parabolas have a positive concavity (negative

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1727

<0 ZI
I1 I2
I=0

I1
< I2
< I3

I3

ZR

Figure 6. Qualitative plots of the MI curves in the plane ZR–ZI of a coasting beam above the
transition energy, i.e. h!0 (Anderson et al. 1999a). The bold face vertical straight lines represent
the stability region (uIZ0).

mass instability). Let us suppose that hO0. From equation (5.11) one can easily
see that, approaching uIZ0, the parabolas reduce asymptotically to a straight
line lower unlimited located on the imaginary axis, as shown in figure 5. If h!0,
in the same limit, parabolas reduce to a straight line upper unlimited located on
the imaginary axis, as shown in figure 6. The straight line represents the only
possible region below (above) the transition energy where the system is
modulationally stable against small perturbations in both density and velocity
of the beam, with respect to their unperturbed values l0 and V0, respectively.
(Note that density and velocity are directly connected with amplitude and phase,
respectively, of the wave function J.) Any other point of the complex plane
belongs to an instability parabola (uIs0).
In the limit of small dispersion, i.e. ek/1, the second term of the r.h.s. of
equation (5.10) can be neglected and equation (5.11) reduces to
ekl0 2 1 u2I
ZI zKh ZR C : ð5:12Þ
4u2I h ekl0
Furthermore, for purely reactive impedances (ZRh0), equation (5.11) reduces to
the cubic NLS equation and the corresponding dispersion relation gives (note
that in this case uRZV0k)
 
u2I ZI a2 k 2
Z ehl 0 K ; ð5:13Þ
k2 k 4
from which it is easily seen that the system is modulationally unstable (u2I O 0)
under the following conditions:
hZI O 0 ðhX I O 0Þ; ð5:14Þ

ehk 2
l0 O : ð5:15Þ
4X I
Condition (5.15) implies that the instability threshold is given by the non-zero
minimum intensity l0m Z ehk 2 =4X I .

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1728 R. Fedele

Remarkably, we observe that condition (5.14) is the well-known Lighthill


criterion associated with the standard cubic NLS equation (Lighthill 1965, 1967).
On the other hand, according to what has been described in §3, it reproduces the
coherent instability condition (see table 1) of a monochromatic coasting beam in
the presence of a purely reactive impedance.
Furthermore, equation (5.12) predicts the MI associated with the cubic
integro-differential NLS equation (5.2). In particular, the MI is predicted in
terms of parabolas in the plane ZR–ZI, but, remarkably, it coincides with the
results of coherent instability of a monochromatic coasting beam in the presence
of an arbitrary longitudinal coupling impedance presented in §4.
To complete the above correspondence, one has to extend the MI analysis to
the case of a non-monochromatic beam. This is done in §5c.

(c ) Statistical approach to MI analysis of a non-monochromatic


coasting beam
The dispersion relation (5.10) allows one to write an expression for the
‘longitudinal coupling admittance’ of the coasting beam, Yh1/Z:
ial0
kY Z : ð5:16Þ
ðu=k K V0 Þ2 Ka2 k 2 =4
Let us now consider a non-monochromatic coasting beam. Such a system may be
thought as an ensemble of incoherent coasting beams with different unperturbed
velocities (white beam). Let us call f0(V ) the distribution function of the velocity
at the equilibrium. The subsystem corresponding to a coasting beam collecting
the particles having velocities between V and VCdV has an elementary
admittance dY. Replacing l0 with f0(V )dV in equation (5.16), the expression
for the elementary admittance is easily given as follows:
iaf0 ðV ÞdV
k dY Z : ð5:17Þ
ðV Ku=kÞ2 Ka2 k 2 =4
All the elementary coasting beams in which we have divided the system suffer the
same electric voltage per unit length along the longitudinal direction. This means
that the total admittance of the system is the sum of all the elementary
admittances, as happens for a system of electric wires all connected in parallel.
Therefore, ð
f0 ðV ÞdV
kY Z ia : ð5:18Þ
ðV Ku=kÞ2 Ka2 k 2 =4
Of course, this dispersion relation can also be cast in the following way:
 ð
Z f0 ðV ÞdV
1 Z ia ; ð5:19Þ
k ðV Ku=kÞ2 Ka2 k 2 =4
where ZZ1/Y stands for the total impedance of the system which is the inverse
of the total admittance.
An interesting equivalent form of equation (5.19) can be obtained (Fedele
et al. 2006). To this end, we first observe that the following identity holds:
 
1 1 1 1
Z K :
ðV Ku=kÞ2 Ka2 k 2 =4 ak ðV Kak=2ÞKu=k ðV C ak=2ÞKu=k

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1729

Then, using this identity, equation (5.19) can be easily cast in the following form:
  ð ð 
Z 1 f0 ðp1 C ak=2Þdp1 f0 ðp2 Kak=2Þdp2
1 Zi K ; ð5:20Þ
k k p1 Ku=k p2 Ku=k

and finally in the following form:


 ð
Z f0 ðV C ak=2ÞK f0 ðV Kak=2Þ dV
1 Z ia : ð5:21Þ
k ak V Ku=k

We soon observe that, assuming that f0(V ) is proportional to d(VKV0), from


equation (5.21) we easily recover the dispersion relation for the case of a
monochromatic coasting beam (see equation (5.10)). In general, equation (5.21)
takes into account the non-monochromatic character of the velocity distribution.
This dispersion relation extends the standard MI analysis to a non-
monochromatic beam thought of as a statistical ensemble of monochromatic
coasting beams. It predicts a Landau-type damping, a phenomenon very similar
to the well-known Landau damping, predicted by L. D. Landau in 1946 for
plasma waves (Landau 1946). The latter is recovered in the limit of small
dispersion, i.e. ek/1. In fact, the dispersion relation (5.21) reduces to a form
very similar to dispersion relation (4.1) obtained in classical kinetic theory
(Vlasov–Maxwell system).

(d ) Quantum-like kinetic description of coherent instability


(Wigner–Maxwell system)
Equation (5.21) can be also obtained within the kinetic description provided
by the quasidistribution (Wigner 1932; Ville 1948; Moyal 1949), as it has been
done for the first time in the context of the TWM (Anderson et al. 1999b; Fedele
et al. 2000) soon extended to nonlinear optics (Fedele & Anderson 2000; Hall
et al. 2002; Helczynski et al. 2002), plasma physics (Fedele et al. 2002; Marklund
2005), surface gravity waves (Onorato et al. 2003) and lattice vibration physics
(molecular crystals, Visinescu & Grecu 2003; Grecu & Visinescu 2004, 2005).
According to the former quantum kinetic approaches (Klimontovich & Silin
1960; Alber 1978), the basic idea is to transit from the configuration space
description, where the NLS equation governs the particular wave-envelope
propagation, to the phase space, where an appropriate kinetic equation is able to
show a random version of the MI. This has been accomplished by using the
mathematical tool provided by the ‘quasidistribution’ (the Fourier transform of
the density matrix) that is widely used for quantum systems. In fact, for any
nonlinear system, whose dynamics is governed by the NLS equation, one can
introduce a two-point correlation function that plays a role similar to the one
played by the density matrix of a quantum system (Landau 1927; Weyl 1931).
Consequently, in the TWM framework, the governing kinetic equation is nothing
but a sort of nonlinear von Neumann–Weyl (or Wigner–Moyal) equation
(Wigner 1932; Moyal 1949) coupled with the effective potential (via Maxwell’s
equations) accounting for the interaction with the beam and the surroundings.
A linear stability analysis of the von Neumann–Weyl equation for the
quasidistribution leads to a dispersion relation that coincides with equation (5.12).

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1730 R. Fedele

6. The impact of Maxwell’s theory during the twentieth century


and concluding remarks

In this section, we describe the impact that Maxwell’s theory of Saturn’s rings
has produced on the physics of the twentieth century. To this end, we present a
summary of the above results in the form of a historical discussion.
In this paper, we have outlined that, 100 years later, by replacing the
gravitational interaction with the e.m. one, Maxwell’s theory of Saturn’s rings
was extended to the more rich context of charge particle beam dynamics in high-
energy circular accelerators. With this extension, an important transfer of
knowledge from planetary physics to accelerator physics allows one to conceive
the physical mechanism of coherent instability (including the negative mass
instability). During the sixth and seventh decades, this mechanism soon
stimulated the development of the general theory of coherent instability based
on the kinetic theory (Vlasov–Maxwell system). Another important process
meanwhile already known in physics was the MI. Since the 6th decade, the MI
(also known as the Benjamin–Feir instability, Benjamin & Feir 1967; Zakharov
1968; Yuen 1982) was already referred to as ‘a general phenomenon encountered
when a quasi-monochromatic wave is propagating in a weak nonlinear medium,
whose dynamics is governed essentially by a NLS equation coupled with a set of
equations accounting for the interaction between the wave and the medium’. It
was discussed and even observed experimentally in many fields of physics, such
as deep waters physics (ocean gravity waves), plasma physics (electrostatic and
e.m. plasma waves) and nonlinear optics (Kerr media, optical fibres). More
recently, it has also been studied in electrical transmission lines, matter wave
physics (Bose–Einstein condensates), lattice vibrations physics (molecular
crystals) and in the physics of anti-ferromagnetism (dynamics of the spin waves).
For a review of the above developments of the MI studies, see Karpman
(1975), Sulem & Sulem (1999) and Abdulaev et al. (2002).
For a long time, the theories of coherent instability and MI were neither
compared, one to the other, nor connected to each other. Around the beginning
of the 9th decade, a quantum-like theory of charged particle beam dynamics was
proposed in the literature. On the basis of valuable physical analogies, the TWM
was formulated (Fedele & Miele 1991). Later on, by means of the TWM, a
connection with the conventional description of a particle accelerator in the
presence of collective effects was given for the first time in terms of a standard
cubic NLS equation (Fedele et al. 1993) and an integro-differential cubic NLS
equation (Anderson et al. 1999b). In this way, the coherent instability was easily
re-described as a sort of MI process. It is worth pointing out that this discovery
has shown, in particular, that Maxwell’s original idea on Saturn’s ring stability
was also imported and extended to a new context where the ‘elements’ involved
in the dynamics of the system were not the particles but the waves.
Classical statistical approaches or Vlasov-type kinetic equations to describe
systems of waves were used since the 1960s to describe wave interactions and/or
the phenomenon of the Landau damping (see Dawson 1960; Hasselmann 1962;
Vedenov & Rudakov 1965; Bingam et al. 1997; Zakharov 1999). More recently,
within the context of the quantum-like kinetic description based on the Wigner
quasidistribution, a statistical approach to MI has been formulated (Anderson
et al. 1999a; Fedele & Anderson 2000; Fedele et al. 2000, 2006). Remarkably, this

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1731

approach, which has been applied to a number of physical problems, unifies the
classical descriptions of coherent instability and MI in one general approach to the
instability of any weakly nonlinear system in which the propagation of partially
incoherent waves is involved (Fedele et al. 2002; Hall et al. 2002; Helczynski et al.
2002; Onorato et al. 2003; Santos et al. 2007). Consequently, it can be thought as
the generalization of Maxwell’s original idea to a random wave system.
The studies of the extreme physics exhibited by astrophysical objects, such as
the superfluid two-stream instability neutrons stars (Andersson et al. 2003),
Bose–Einstein condensation properties in collapsing and exploding stars (Ball
2001), density waves in spiral structures (Lin & Bertin 1995) and the quasi-linear
kinetic approach to the planetary rings (Griv et al. 2003), have already imported,
directly or indirectly, the basic idea of Maxwell’s theory.
The very rapid developments in quantum gravity have recently shown
valuable applications of the quantum formalism in astrophysics and cosmology
(Penrose 1996, 1997, 1998; Moroz et al. 1998; Tod & Moroz 1999). It is reasonable
to assume that suitable Schrödinger-like equations, coupled with a set of
equations describing the gravitational interaction, govern the collective dynamics
of the astrophysical systems in terms of a complex wave function as it has been
assumed for the entire Universe (Hartle & Hawking 1983; Hawking 1984).
As a natural consequence, the quantum-like kinetic description of the
coherent/MI and the related statistical approach may provide a random version
of Maxwell’s theory whether applied to planetary problems or to different
astrophysical environments. In particular, a kinetic theory, based on the
Schrödinger–Newton (Schrödinger–Einstein) system or Wigner–Newton (Wigner–
Einstein) system, is underway and it would provide a more general description of the
instability of such systems.
It should be noted that, when the knowledge of the quantum-like formulation
of the coherent/MI is transferred back to astrophysics and cosmology, it is
natural that one would need to determine, in the gravitational context, a
possible counterpart of the e.m. repulsion. The latter has been detected through
the accelerated expansion of the Universe and might be interpreted as an
effective ‘repulsive gravitation’. A suitable theory of MI including this effect is
also underway.

References
Abdulaev, F. Kh., Darmanyan, S. A. & Garnier, J. 2002 Modulational instability of electro-
magnetic waves in inhomogeneous and discrete media. Prog. Opt. 44, 303–365.
Alber, I. E. 1978 The effects of randomness on the stability of two-dimensional surface wavetrains.
Proc. R. Soc. A 363, 525–546. (doi:10.1098/rspa.1978.0181)
Anderson, D., Fedele, R., Vaccaro, V. G., Lisak, M., Berntson, A. & Johansson, S. 1999a
Modulational instabilities within the thermal wave model of high-energy charged-particle beam
dynamics. Phys. Lett. A 258, 244–248. (doi:10.1016/S0375-9601(99)00374-6)
Anderson, D., Fedele, R., Vaccaro, V. G., Lisak, M., Berntson, A. & Johansson, S. 1999b
Quantum-like description of modulational instability and landau damping in the longitudinal
dynamics of high-energy charged-particle beams. In Proc. 1998 ICFA workshop on Nonlinear
Collective Phenomena in Beam Physics, Arcidosso, Italy, 1–5 September 1998 (eds
S. Chattopadhyay, M. Cornacchia & C. Pellegrini), p. 197. New York, NY: AIP Press.
Andersson, N., Comer, G. L. & Prix, R. 2003 Are pulsar glitches triggered by a superfluid two-
stream instability? Phys. Rev. Lett. 90, 091101. (doi:10.1103/PhysRevLett.90.091101)

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

1732 R. Fedele

Ball, P. 2001 Tabletop astrophysics. Nature 411, 628–630. (doi:10.1038/35079770)


Benjamin, T. B. & Feir, J. E. 1967 The disintegration of wave trains on deep water. Part 1.
Theory. J. Fluid Mech. 27, 417–430. (doi:10.1017/S002211206700045X)
Bingam, R., Mendonca, J. T. & Dawson, J. M. 1997 Photon Landau damping. Phys. Rev. Lett. 78,
247–249. (doi:10.1103/PhysRevLett.78.247)
Chao, A. (ed.) 1993 Physics of collective beam instabilities in high energy accelerators. New York,
NY: Wiley.
Dawson, J. M. 1960 Plasma oscillations of a large number of electron beams. Phys. Rev. 118,
381–389. (doi:10.1103/PhysRev.118.381)
Fedele, R. & Anderson, D. 2000 A quantum-like Landau damping of an e.m. wavepacket. J. Opt.
B: Quantum Semiclassical Opt. 2, 207–213. (doi:10.1088/1464-4266/2/2/325)
Fedele, R. & Miele, G. 1991 A thermal-wave model for relativistic-charged-particle beam
propagation. Nuovo Cimento D 13, 1527–1544. (doi:10.1007/BF02457189)
Fedele, R. & Shukla, R. K. (eds) 1995 Quantum-like models and coherent effects. Singapore: World
Scientific.
Fedele, R., Miele, G., Palumbo, L. & Vaccaro, V. G. 1993 Thermal wave model for nonlinear
longitudinal dynamics in particle accelerators. Phys. Lett. A 179, 407–413. (doi:10.1016/0375-
9601(93)90099-L)
Fedele, R., Anderson, D. & Lisak, M. 2000 Role of Landau damping in the quantum-like theory of
charged-particle beam dynamics. In Proc. 7th European Particle Accelerator Conf.—EPAC2000,
Vienna, Austria, 26–30 June 2000 (eds W. A. Mitaroff, C. Petit-Jean-Genaz, J. Poole, M. Regler &
J. L. Laclare), p. 1489. Geneva, Switzerland: CERN Scientific Information Service.
Fedele, R., Shukla, P. K., Onorato, M., Anderson, D. & Lisak, M. 2002 Landau damping of partially
incoherent Langmuir waves. Phys. Lett. A 303, 61–66. (doi:10.1016/S0375-9601(02)01201-X)
Fedele, R., Anderson, D. & Lisak, M. 2006 How the coherent instabilities of an intense high-energy
charged-particle beam in the presence of nonlocal effects can be explained within the context of
the Madelung fuid description. Eur. Phys. J. B 49, 275–281. (doi:10.1140/epjb/e2006-00067-3)
Grecu, D. & Visinescu, A. 2004 Modulational instability of some nonlinear continuum and discrete
systems. AIP Conf. Proc. 729, 389–395. (doi:10.1063/1.1814755)
Grecu, D. & Visinescu, A. 2005 Modulational instability of coupled nonlinear equations:
Manakov’s system. Rom. J. Phys. 50, 137–146.
Griv, E., Gedalin, M. & Yuan, C. 2003 On the stability of Saturn’s rings: a quasi-linear kinetic
theory. Month. Not. R. Astron. Soc. 342, 1102–1116. (doi:10.1046/j.1365-8711.2003.06608.x)
Hall, B., Lisak, M., Anderson, D., Fedele, R. & Semenov, V. E. 2002 Statistical theory for
incoherent light propagation in nonlinear media. Phys. Rev. E 65, 035602. (doi:10.1103/
PhysRevE.65.035602)
Hartle, J. B. & Hawking, S. W. 1983 Wave function of the Universe. Phys. Rev. D 44, 2960–2975.
(doi:10.1103/PhysRevD.28.2960)
Hasselmann, K. 1962 On the non-linear energy transfer in a gravity-wave spectrum. Part I.
General theory. J. Fluid Mech. 12, 481–500. (doi:10.1017/S0022112062000373)
Hawking, S. W. 1984 The quantum state of the Universe. Nucl. Phys. B 239, 257–276. (doi:10.
1016/0550-3213(84)90093-2)
Helczynski, L., Anderson, D., Fedele, R., Hall, B. & Lisak, M. 2002 Propagation of partially
incoherent light in nonlinear media via the Wigner transform method. IEEE J. Sel. Top.
Quantum Electron. 8, 408–412. (doi:10.1109/JSTQE.2002.1016342)
Johannisson, P., Anderson, D., Lisak, M., Marklund, M., Fedele, R. & Kim, A. 2004 Nonlocal
effects in high-energy charged-particle beams. Phys. Rev. E 69, 066501. (doi:10.1103/PhysRevE.
69.066501)
Karpman, V. I. (ed.) 1975 Nonlinear waves in dispersive media. Oxford, UK: Pergamon Press.
Klimontovich, Y. & Silin, V. 1960 The spectra of systems of interacting particles and collective
energy losses during passage of charged particles through matter. Sov. Phys. Usp. 3, 84. (doi:10.
1070/PU1960v003n01ABEH003260)

Phil. Trans. R. Soc. A (2008)


Downloaded from http://rsta.royalsocietypublishing.org/ on June 26, 2016

From Maxwell to negative mass instability 1733

Landau, L. 1927 Das Dämpfungsproblem in der Wellenmechanik. Z. Phys. 45, 430–441. (doi:10.
1007/BF01343064)
Landau, L. D. 1946 On the vibrations of the electronic plasma. J. Phys. USSR 10, 25–34.
Lawson, J. D. (ed.) 1988 The physics of charged particle beams, 2nd edn. Oxford, UK: Clarendon.
Lighthill, M. J. 1965 Contributions to the theory of waves in non-linear dispersive systems. J. Inst.
Appl. Math. 1, 269–306. (doi:10.1093/imamat/1.3.269)
Lighthill, M. J. 1967 Some special cases treated by the whitham theory. Proc. R. Soc. A 299,
28–53. (doi:10.1098/rspa.1967.0121)
Lin, C. C. & Bertin, G. 1995 Global wave patterns in galaxies: their generation and maintenance in
waves in astrophysics. Ann. NY Acad. Sci. 773, 125–144. (doi:10.1111/j.1749-6632.1995.
tb12166.x)
Madelung, E. 1926 Quanten theorie in hydrodynamischer form. Z. Phys. 40, 332.
Marklund, M. 2005 Classical and quantum kinetics of the Zakharov system. Phys. Plasmas 12,
082110. (doi:10.1063/1.2012147)
Moroz, I. M., Penrose, R. & Tod, K. P. 1998 Spherically-symmetric solutions of the Schrödinger–
Newton equations. Class. Quantum Grav. 15, 2733–2742. (doi:10.1088/0264-9381/15/9/019)
Moyal, J. E. 1949 Quantum mechanics as a statistical theory. Proc. Camb. Philos. Soc. 45, 99–124.
Nicholson, D. R. 1983 Introduction to plasma theory. New York, NY: Wiley.
Nielsen, C. E., Sessler, A. M., & Symon, K. R. 1959 In Proc. Int. Conf. on High Energy
Accelerators and Instrumentation, CERN, Geneva, 14–19 September 1959 (ed. L. Kowarski).
Geneva, Switzerland: CERN Scientific Information Service.
Niven, W. D. (ed.) 1890 The scientific papers of James Clerk Maxwell, vol. 1, pp. 288–377.
Cambridge, UK: Cambridge University Press.
Onorato, M., Osborne, A., Fedele, R. & Serio, M. 2003 Landau damping and coherent structures in
narrow-banded 1C1 deep water gravity waves. Phys. Rev. E 67, 046305. (doi:10.1103/
PhysRevE.67.046305)
Penrose, R. 1996 On gravity’s role in quantum state reduction. Gen. Rel. Grav. 28, 581–600.
(doi:10.1007/BF02105068)
Penrose, R. (ed.) 1997 The large, the small, and the human mind. Cambridge, UK: Cambridge
University Press.
Penrose, R. 1998 Quantum computation, entanglement and state reduction. Phil. Trans. R. Soc. A
356, 1927–1939. (doi:10.1098/rsta.1998.0256)
Santos, J. E., Silva, L. O. & Bingham, R. 2007 White-light parametric instabilities in plasmas.
Phys. Rev. Lett. 98, 235001. (doi:10.1103/PhysRevLett.98.235001)
Sulem, P. & Sulem, C. (eds) 1999 Nonlinear Schrödinger equation. Berlin, Germany: Springer.
Tod, P. & Moroz, I. M. 1999 An analytical approach to the Schrödinger–Newton equations.
Nonlinearity 12, 201–216. (doi:10.1088/0951-7715/12/2/002)
Vedenov, A. A. & Rudakov, L. I. 1965 Interaction of waves in continuous media. Sov. Phys. Dokl.
9, 1073.
Ville, J. 1948 Theorie et applications de signal analytique. Cables Transm. 2, 61.
Visinescu, A. & Grecu, D. 2003 Statistical approach of the modulational instability of the discrete
self-trapping equation. Eur. Phys. J. B 34, 225–229. (doi:10.1140/epjb/e2003-00215-3)
Weyl, H. 1931 Gruppentheorie und Quantenmechanik. (Transl. The theory of groups and quantum
mechanics). New York, NY: Dover Publications.
Wigner, E. 1932 On the quantum correction for thermodynamic equilibrium. Phys. Rev. 40, 749.
(doi:10.1103/PhysRev.40.749)
Yuen, H. C. 1982 Nonlinear dynamics of deep-water gravity waves. Adv. Appl. Mech. 22, 67.
Zakharov, V. E. 1968 Stability of periodic waves of finite amplitude on the surface of a deep fluid.
J. Appl. Mech. Tech. Phys. 9, 190–194. (doi:10.1007/BF00913182)
Zakharov, V. 1999 Statistical theory of gravity and capillary waves on the surface of a finite-depth
fluid. Eur. J. Mech. B/Fluids 18, 327–344. (doi:10.1016/S0997-7546(99)80031-4)

Phil. Trans. R. Soc. A (2008)

You might also like