Part III Classical and Quantum Solitons
Based on lectures by N. S. Manton and D. Stuart
Notes taken by Dexter Chua
Easter 2017
Solitons are solutions of classical field equations with particle-like properties. They
are localised in space, have finite energy and are stable against decay into radiation.
The stability usually has a topological explanation. After quantisation, they give rise
to new particle states in the underlying quantum field theory that are not seen in
perturbation theory. We will focus mainly on kink solitons in one space dimension, on
gauge theory vortices in two dimensions, and on Skyrmions in three dimensions.
Pre-requisites
This course assumes you have taken Quantum Field Theory and Symmetries, Fields
and Particles. The small amount of topology that is needed will be developed during
the course.
Reference
N. Manton and P. Sutcliffe, Topological Solitons, CUP, 2004
Contents
0 Introduction
1 φ
4
kinks
1.1 Kink solutions
1.2 Dynamic kink
1.3 Soliton interactions
1.4 Quantization of kink motion
1.5 Sine-Gordon kinks
2 Vortices
2.1 Topological background
2.2 Global U (1) Ginzburg–Landau vortices
2.3 Abelian Higgs/Gauged Ginzburg–Landau vortices
2.4 Bogomolny/self-dual vortices and Taubes’ theorem
2.5 Physics of vortices
2.6 Jackiw–Pi vortices
3 Skyrmions
3.1 Skyrme field and its topology
3.2 Skyrmion solutions
3.3 Other Skyrmion structures
3.4 Asymptotic field and forces for B = 1 hedgehogs
3.5 Fermionic quantization of the B = 1 hedgehog
3.6 Rigid body quantization
0 Introduction
Given a classical field theory, if we want to “quantize” it, then we find the
vacuum of the theory, and then do perturbation theory around this vacuum. If
there are multiple vacua, then what we did was that we arbitrarily picked a
vacuum, and then expanded around that vacuum.
However, these field theories with multiple vacua often contain soliton so-
lutions. These are localized, smooth solutions of the classical field equations,
and they “connect multiple vacuums”. To quantize these solitons solutions, we
fix such a soliton, and use it as the “background”. We then do perturbation
theory around these solutions, but this is rather tricky to do. Thus, in a lot of
the course, we will just look at the classical part of the theory.
Recall that when quantizing our field theories in perturbation theory, we
obtain particles in the quantum theory, despite the classical theory being com-
pletely about fields. It turns out solitons also behave like particles, and they are
a new type of particles. These are non-perturbative phenomena. If we want to
do the quantum field theory properly, we have to include these solitons in the
quantum field theory. In general this is hard, and so we are not going to develop
this a lot.
What does it mean to say that solitons are like particles? In relativistic field
theories, we find these solitons have a classical energy. We define the “mass”
M
of the soliton to be the energy in the “rest frame”. Since this is relativistic, we
can do a Lorentz boost, and we obtain a moving soliton. Then we obtain an
energy-momentum relation of the form
E
2
P · P = M
2
.
This is a Lorentz-invariant property of the soliton. Together with the fact that
the soliton is localized, this is a justification for thinking of them as particles.
These particles differ from the particles of perturbative quantum fields, as
they have rather different properties. Interesting solitons have a topological
character different from the classical vacuum. Thus, at least naively, they cannot
be thought of perturbatively.
There are also non-relativistic solitons, but they usually don’t have interpre-
tations as particles. These appear, for example, as defects in solids. We will not
be interested in these much.
What kinds of theories have solitons? To obtain solitons, we definitely need
a non-linear field structure and/or non-linear equations. Thus, free field theories
with quadratic Lagrangians such as Maxwell theory do not have solitons. We
need interaction terms.
Note that in QFT, we dealt with interactions using the interaction picture.
We split the Hamiltonian into a “free field” part, which we solve exactly, and
the “interaction” part. However, to quantize solitons, we need to solve the full
interacting field equations exactly.
Having interactions is not enough for solitons to appear. To obtain solitons,
we also need some non-trivial vacuum topology. In other words, we need more
than one vacuum. This usually comes from symmetry breaking, and often gauge
symmetries are involved.
In this course, we will focus on three types of solitons.
In one (space) dimension, we have kinks. We will spend 4 lectures on this.
In two dimensions, we have vortices. We will spend 6 lectures on this.
In three dimensions, there are monopoles and Skyrmions. We will only
study Skyrmions, and will spend 6 lectures on these.
These examples are all relativistic. Non-relativistic solitons include domain walls,
which occur in ferromagnets, and two-dimensional “baby” Skyrmions, which are
seen in exotic magnets, but we will not study these.
In general, solitons appear in all sorts of different actual, physical scenarios
such as in condensed matter physics, optical fibers, superconductors and exotic
magnets. “Cosmic strings” have also been studied. Since we are mathematicians,
we probably will not put much focus on these actual applications. However, we
can talk a bit more about Skyrmions.
Skyrmions are solitons in an effective field theory of interacting pions, which
are thought to be the most important hadrons because they are the lightest.
This happens in spite of the lack of a gauge symmetry. While pions have no
baryon number, the associated solitons have a topological charge identified with
baryon number. This baryon number is conserved for topological reasons.
Note that in QCD, baryon number is conserved because the quark number
is conserved. Experiments tried extremely hard to find proton decay, which
would be a process that involves baryon number change, but we cannot find
such examples. We have very high experimental certainty that baryon number is
conserved. And if baryon number is topological, then this is a very good reason
for the conservation of baryon number.
Skyrmions give a model of low-energy interactions of baryons. This leads
to an (approximate) theory of nucleons (proton and neutron) and larger nuclei,
which are bound states of any number of protons and neutrons.
For these ideas to work out well, we need to eventually do quantization.
For example, Skyrmions by themselves do not have spin. We need to quantize
the theory before spins come out. Also, Skyrmions cannot distinguish between
protons and neutrons. These differences only come up after we quantize.
1 φ
4
kinks
1.1 Kink solutions
In this section, we are going to study
φ
4
kinks. These occur in 1 + 1 dimensions,
and involve a single scalar field
φ
(
x, t
). In higher dimensions, we often need
many fields to obtain solitons, but in the case of 1 dimension, we can get away
with a single field.
In general, the Lagrangian density of such a scalar field theory is of the form
L =
1
2
µ
φ∂
µ
φ U(φ)
for some potential
U
(
φ
) polynomial in
φ
. Note that in 1 + 1 dimensions, any
such theory is renormalizable. Here we will choose the Minkowski metric to be
η
µν
=
1 0
0 1
,
with µ, ν = 0, 1. Then the Lagrangian is given by
L =
Z
−∞
L dx =
Z
−∞
1
2
µ
φ∂
µ
φ U(φ)
dx,
and the action is
S[φ] =
Z
L dt =
Z
L dx dt.
There is a non-linearity in the field equations due to a potential
U
(
φ
) with
multiple vacua. We need multiple vacua to obtain a soliton. The kink stability
comes from the topology. It is very simple here, and just comes from counting
the discrete, distinct vacua.
As usual, we will write
˙
φ =
φ
t
, φ
0
=
φ
x
.
Often it is convenient to (non-relativistically) split the Lagrangian as
L = T V,
where
T =
Z
1
2
˙
φ
2
dx, V =
Z
1
2
φ
02
+ U(φ)
dx.
In higher dimensions, we separate out
µ
φ into
˙
φ and φ.
The classical field equation comes from the condition that
S
[
φ
] is stationary
under variations of
φ
. By a standard manipulation, the field equation turns out
to be
µ
µ
φ +
dU
dφ
= 0.
This is an example of a Klein–Gordon type of field equation, but is non-linear if
U is not quadratic. It is known as the non-linear Klein–Gordon equation.
We are interested in a soliton that is a static solution. For a static field, the
time derivatives can be dropped, and this equation becomes
d
2
φ
x
2
=
dU
dφ
.
Of course, the important part is the choice of U! In φ
4
theory, we choose
U(φ) =
1
2
(1 φ
2
)
2
.
This is mathematically the simplest version, because we set all coupling constants
to 1.
The importance of this U is that it has two minima:
φ
U(φ)
1 1
The two classical vacua are
φ(x) 1, φ(x) 1.
This is, of course, not the only possible choice. We can, for example, include
some parameters and set
U(φ) = λ(m
2
φ
2
)
2
.
If we are more adventurous, we can talk about a φ
6
theory with
U(φ) = λφ
2
(m
2
φ
2
)
2
.
In this case, we have 3 minima, instead of 2. Even braver people can choose
U(φ) = 1 cos φ.
This has infinitely many minima. The field equation involves a
sin φ
term,
and hence this theory is called the sine-Gordon theory (a pun on the name
Klein–Gordon, of course).
The sine-Gordon theory is a special case. While it seems like the most
complicated potential so far, it is actually integrable. This implies we can find
explicit exact solutions involving multiple, interacting solitons in a rather easy
way. However, integrable systems is a topic for another course, namely IID
Integrable Systems.
For now, we will focus on our simplistic
φ
4
theory. As mentioned, there are
two vacuum field configurations, both of zero energy. We will in general use
the term field configuration to refer to fields at a given time that are not
necessarily solutions to the classical field equation, but in this case, the vacua
are indeed solutions.
If we wanted to quantize this
φ
4
theory, then we have to pick one of the vacua
and do perturbation theory around it. This is known as spontaneous symmetry
breaking. Of course, by symmetry, we obtain the same quantum theory regardless
of which vacuum we expand around.
However, as we mentioned, when we want to study solitons, we have to
involve both vacua. We want to consider solutions that “connect” these two
vacua. In other words, we are looking for solutions that look like
x
φ
a
This is known as a kink solution.
To actually find such a solution, we need the full field equation, given by
d
2
φ
dx
2
= 2(1 φ
2
)φ.
Instead of solving this directly, we will find the kink solutions by considering the
energy, since this method generalizes better.
We will work with a general potential
U
with minimum value 0. From
Noether’s theorem, we obtain a conserved energy
E =
Z
1
2
˙
φ
2
+
1
2
φ
02
+ U(φ)
dx.
For a static field, we drop the
˙
φ
2
term. Then this is just the
V
appearing in the
Lagrangian. By definition, the field equation tells us the field is a stationary
point of this energy. To find the kink solution, we will in fact find a minimum
of the energy.
Of course, the global minimum is attained when we have a vacuum field, in
which case
E
= 0. However, this is the global minimum only if we don’t impose
any boundary conditions. In our case, the kinks satisfy the boundary conditions
φ
(
) = 1”, “
φ
(
−∞
) =
1” (interpreted in terms of limits, of course). The kinks
will minimize energy subject to these boundary conditions.
These boundary conditions are important, because they are “topological”.
Eventually, we will want to understand the dynamics of solitons, so we will want
to consider fields that evolve with time. From physical considerations, for any
fixed
t
, the field
φ
(
x, t
) must satisfy
φ
(
x, t
)
vacuum
as
x ±∞
, or else the
field will have infinite energy. However, the vacuum of our potential
U
is discrete.
Thus, if
φ
is to evolve continuously with time, the boundary conditions must
not evolve with time! At least, this is what we expect classically. Who knows
what weird tunnelling can happen in quantum field theory.
So from now on, we fix some boundary conditions
φ
(
) and
φ
(
−∞
), and
focus on fields that satisfy these boundary conditions. The trick is to write the
potential in the form
U(φ) =
1
2
dW (φ)
dφ
2
.
If
U
is non-negative, then we can always find
W
in principle we take the
square root and then integrate it. However, in practice, this is useful only if we
can find a simple form for
W
. Let’s assume we’ve done that. Then we can write
E =
1
2
Z
φ
02
+
dW
dφ
2
!
dx
=
1
2
Z
φ
0
dW
dφ
2
dx ±
Z
dW
dφ
dφ
dx
dx
=
1
2
Z
φ
0
dW
dφ
2
dx ±
Z
dW
=
1
2
Z
φ
0
dW
dφ
2
dx ± (W (φ()) W (φ(−∞))).
The second term depends purely on the boundary conditions, which we have
fixed. Thus, we can minimize energy if we can make the first term vanish! Note
that when completing the square, the choice of the signs is arbitrary. However, if
we want to set the first term to be 0, the second term had better be non-negative,
since the energy itself is non-negative! Hence, we will pick the sign such that
the second term is 0, and then the energy is minimized when
φ
0
= ±
dW
dφ
.
In this case, we have
E = ±(W () W (−∞)).
These are known as the Bogomolny equation and the Bogomolny energy bound.
Note that if we picked the other sign, then we cannot solve the differential
equation φ
0
= ±
dW
dφ
, because we know the energy must be non-negative.
For the φ
4
kink, we have
dW
dφ
= 1 φ
2
.
So we pick
W = φ
1
3
φ
3
.
So when
φ
=
±
1, we have
W
=
±
2
3
. We need to choose the + sign, and then we
know the energy (and hence mass) of the kink is
E M =
4
3
.
We now solve for φ. The equation we have is
φ
0
= 1 φ
2
.
Rearranging gives
1
1 φ
2
dφ = dx,
which we can integrate to give
φ(x) = tanh(x a).
This
a
is an arbitrary constant of integration, labelling the intersection of the
graph of φ with the x-axis. We think of this as the location of the kink.
Note that there is not a unique solution, which is not unexpected by trans-
lation invariance. Instead, the solutions are labeled by a parameter
a
. This is
known as a modulus of the solution. In general, there can be multiple moduli,
and the space of all possible values of the moduli of static solitons is known as
the moduli space. In the case of a kink, the moduli space is just R.
Is this solution stable? We obtained this kink solution by minimizing the
energy within this topological class of solutions (i.e. among all solutions with
the prescribed boundary conditions). Since a field cannot change the boundary
conditions during evolution, it follows that the kink must be stable.
Are there other soliton solutions to the field equations? The solutions are
determined by the boundary conditions. Thus, we can classify all soliton solutions
by counting all possible combinations of the boundary conditions. We have, of
course, two vacuum solutions
φ
1 and
φ
1. There is also an anti-kink
solution obtained by inverting the kink:
φ(x) = tanh(x b).
This also has energy
4
3
.
1.2 Dynamic kink
We now want to look at kinks that move. Given what we have done so far, this
is trivial. Our theory is Lorentz invariant, so we simply apply a Lorentz boost.
Then we obtain a field
φ(x, t) = tanh γ(x vt),
where, as usual
γ = (1 v
2
)
1/2
.
But this isn’t all. Notice that for small
v
, we can approximate the solution
simply by
φ(x, t) = tanh(x vt).
This looks like a kink solution with a modulus that varies with time slowly. This
is known as the adiabatic point of view.
More generally, let’s consider a “moving kink” field
φ(x, t) = tanh(x a(t))
for some function
a
(
t
). In general, this is not a solution to the field equation,
but if ˙a is small, then it is “approximately a solution”.
We can now explicitly compute that
˙
φ =
da
dt
φ
0
.
Let’s consider fields of this type, and look at the Lagrangian of the field theory.
The kinetic term is given by
T =
Z
1
2
˙
φ
2
dx =
1
2
da
dt
2
Z
φ
02
dx =
1
2
M
da
dt
2
.
To derive this result, we had to perform the integral
R
φ
02
d
x
, and if we do
that horrible integral, we will find a value that happens to be equal to
M
=
4
3
.
Of course, this is not a coincidence. We can derive this result from Lorentz
invariance to see that the result of integration is manifestly M.
The remaining part of the Lagrangian is less interesting. Since it does not
involve taking time derivatives, the time variation of
a
is not seen by it, and we
simply have a constant
V =
4
3
.
Then the original field Lagrangian becomes a particle Lagrangian
L =
1
2
M ˙a
2
4
3
.
Note that when we first formulated the field theory, the action principle
required us to find a field that extremizes the action among all fields. However,
what we are doing now is to restrict to the set of kink solutions only, and then
when we solve the variational problem arising from this Lagrangian, we are
extremizing the action among fields of the form
tanh
(
x a
(
t
)). We can think of
this as motion in a “valley” in the field configuration space. In general, these
solutions will not also extremize the action among all fields. However, as we
said, it will do so “approximately” if ˙a is small.
We can obtain an effective equation of motion
M¨a = 0,
which is an equation of motion for the variable a(t) in the moduli space.
Of course, the solution is just given by
a(t) = vt + const,
where
v
is an arbitrary constant, which we interpret as the velocity. In this
formulation, we do not have any restrictions on
v
, because we took the “non-
relativistic approximation”. This approximation breaks down when v is large.
There is a geometric interpretation to this. We can view the equation of
motion
M¨a
= 0 as the geodesic equation in the moduli space
R
, and we can think
of the coefficient
M
as specifying a Riemannian metric on the moduli space. In
this case, the metric is (a scalar multiple of) the usual Euclidean metric (da)
2
.
This seems like a complicated way of describing such a simple system, but
this picture generalizes to higher-dimensional systems and allows us to analyze
multi-soliton dynamics, in particular, the dynamics of vortices and monopoles.
We might ask ourselves if there are multi-kinks in our theory. There aren’t in
the
φ
4
theory, because we saw that the solutions are classified by the boundary
conditions, and we have already enumerated all the possible boundary conditions.
In more complicated theories like sine-Gordon theory, multiple kinks are possible.
However, while we cannot have two kinks in
φ
4
theory, we can have a kink
followed by an anti-kink, or more of these pairs. This actually lies in the “vacuum
sector” of the theory, but it still looks like it’s made up of kinks and anti-kinks,
and it is interesting to study these.
1.3 Soliton interactions
We now want to study interactions between kinks and anti-kinks, and see how
they cause each other to move. So far, we were able to label the position of the
particle by its “center”
a
, and thus we can sensibly talk about how this center
moves. However, this center is well-defined only in the very special case of a
pure kink or anti-kink, where we can use symmetry to identify the center. If
there is some perturbation, or if we have a kink and an anti-kink, it is less clear
what should be considered the center.
Fortunately, we can still talk about the momentum of the field, even if
we don’t have a well-defined center. Indeed, since our theory has translation
invariance, Noether’s theorem gives us a conserved charge which is interpreted
as the momentum.
Recall that for a single scalar field in 1 +1 dimensions, the Lagrangian density
can be written in the form
L =
1
2
µ
φ∂
µ
φ U(φ).
Applying Noether’s theorem, to the translation symmetry, we obtain the energy-
momentum tensor
T
µ
ν
=
L
(
µ
φ)
ν
φ δ
µ
ν
L =
µ
φ∂
ν
φ δ
µ
ν
L.
Fixing a time and integrating over all space, we obtain the conserved energy and
conserved momentum. These are
E =
Z
−∞
T
0
0
dx =
Z
−∞
1
2
˙
φ
2
+
1
2
φ
02
+ U(φ)
dx,
P =
Z
−∞
T
0
1
dx =
Z
−∞
˙
φφ
0
dx.
We now focus on our moving kink in the adiabatic approximation of the
φ
4
theory. Then the field is given by
φ = tanh(x a(t)).
Doing another horrible integral, we find that the momentum is just
P = M ˙a.
This is just as we would expect for a particle with mass M!
Now suppose what we have is instead a kink-antikink configuration
x
φ
a
a
Here we have to make the crucial assumption that our kinks are well-separated.
Matters get a lot worse when they get close to each other, and it is difficult
to learn anything about them analytically. However, by making appropriate
approximations, we can understand well-separated kink-antikink configurations.
When the kink and anti-kink are far away, we first pick a point
b
lying
in-between the kink and the anti-kink:
x
φ
a
a
b
The choice of
b
is arbitrary, but we should choose it so that it is far away
from both kinks. We will later see that, at least to first order, the result of our
computations does not depend on which
b
we choose. We will declare that the
parts to the left of
b
belongs to the kink, and the parts to the right of
b
belong
to the anti-kink. Then by integrating the energy-momentum tensor in these two
regions, we can obtain the momentum of the kink and the anti-kink separately.
We will focus on the kink only. Its momentum is given by
P =
Z
b
−∞
T
0
1
dx =
Z
b
−∞
˙
φφ
0
dx.
Since T
µ
ν
is conserved, we know
µ
T
µ
ν
= 0. So we find
t
T
0
1
+
x
T
1
1
= 0.
By Newton’s second law, the force
F
on the kink is given by the rate of change
of the momentum:
F =
d
dt
P
=
Z
b
−∞
t
T
0
1
dx
=
Z
b
−∞
x
T
1
1
dx
= T
1
1
b
=
1
2
˙
φ
2
1
2
φ
02
+ U
b
.
Note that there is no contribution at the
−∞
end because it is vacuum and
T
1
1
vanishes.
But we want to actually work out what this is. To do so, we need to be more
precise about what our initial configuration is. In this theory, we can obtain it
just by adding a kink to an anti-kink. The obvious guess is that it should be
φ(x)
?
= tanh(x + a) tanh(x a),
but this has the wrong boundary condition. It vanishes on both the left and the
right. So we actually want to subtract 1, and obtain
φ(x) = tanh(x + a) tanh(x a) 1 φ
1
+ φ
2
1.
Note that since our equation of motion is not linear, this is in general not a
genuine solution! However, it is approximately a solution, because the kink
and anti-kink are well-separated. However, there is no hope that this will be
anywhere near a solution when the kink and anti-kink are close together!
Before we move on to compute
˙
φ
and
φ
0
explicitly and plugging numbers
in, we first make some simplifications and approximations. First, we restrict
our attention to fields that are initially at rest. So we have
˙
φ
= 0 at
t
= 0. Of
course, the force will cause the kinks to move, but we shall, for now, ignore what
happens when they start moving.
That gets rid of one term. Next, we notice that we only care about the
expression when evaluated at
b
. Here we have
φ
2
1
0. So we can try to
expand the expression to first order in φ
2
1 (and hence φ
0
2
), and this gives
F =
1
2
φ
02
1
+ U(φ
1
) φ
0
1
φ
0
2
+ (φ
2
1)
dU
dφ
(φ
1
)
b
.
We have a zeroth order term
1
2
φ
02
1
+
U
(
φ
1
). We claim that this must vanish.
One way to see this is that this term corresponds to the force when there is no
anti-kink φ
2
. Since the kink does not exert a force on itself, this must vanish!
Analytically, we can deduce this from the Bogomolny equation, which says
for any kink solution φ, we have
φ
0
=
dW
dφ
.
It then follows that
1
2
φ
02
=
1
2
dW
dφ
2
= U(φ).
Alternatively, we can just compute it directly! In any case, convince yourself
that it indeed vanishes in your favorite way, and then move on.
Finally, we note that the field equations tell us
dU
dφ
(φ
1
) = φ
00
1
.
So we can write the force as
F =
φ
0
1
φ
0
2
+ (φ
2
1)φ
00
1
b
.
That’s about all the simplifications we can make without getting our hands dirty.
We might think we should plug in the
tanh
terms and compute, but that is too
dirty. Instead, we use asymptotic expressions of kinks and anti-kinks far from
their centers. Using the definition of tanh, we have
φ
1
= tanh(x + a) =
1 e
2(x+a)
1 + e
2(x+a)
1 2e
2(x+a)
.
This is valid for
x a
, i.e. to the right of the kink. The constant factor of 2 in
front of the exponential is called the amplitude of the tail. We will later see that
the 2 appearing in the exponent has the interpretation of the mass of the field
φ
.
For φ
2
, take the approximation that x a. Then
φ
2
1 = tanh(x a) 1 2e
2(xa)
.
We assume that our
b
satisfies both of these conditions. These are obviously
easy to differentiate once or twice. Doing this, we obtain
φ
0
1
φ
0
2
= (4e
2(x+a)
)(4e
2(xa)
) = 16e
4a
.
Note that this is independent of
x
. In the formula, the
x
will turn into a
b
, and
we see that this part of the force is independent of
b
. Similarly, the other term is
(φ
2
1)φ
00
1
= (2e
2(xa)
)(8e
2(x+a)
) = 16e
4a
.
Therefore we find
F = 32e
4a
,
and as promised, this is independent of the precise position of the cutoff
b
we
chose.
We can write this in a slightly more physical form. Our initial configuration
was symmetric around the
y
-axis, but in reality, only the separation matters.
We write the separation of the pair as s = 2a. Then we have
F = 32e
2s
.
What is the interpretation of the factor of 2? Recall that our potential was given
by
U(φ) =
1
2
(1 φ
2
)
2
.
We can do perturbation theory around one of the vacua, say
φ
= 1. Thus, we
set φ = 1 + η, and then expanding gives us
U(η)
1
2
(2η)
2
=
1
2
m
2
η
2
,
where m = 2. This is the same “2” that goes into the exponent in the force.
What about the constant factor of 32? Recall that when we expanded the
kink solution, we saw that the amplitude
A
of the tail was
A
= 2. It turns out if
we re-did our theory and put back the different possible parameters, we will find
that the force is given by
F = 2m
2
A
2
e
ms
.
This is an interesting and important phenomenon. The mass
m
was the per-
turbative mass of the field. It is something we obtain by perturbation theory.
However, the same mass appears in the force between the solitons, which are
non-perturbative phenomenon!
This is perhaps not too surprising. After all, when we tried to understand
the soliton interactions, we took the approximation that
φ
1
and
φ
2
are close to
1 at b. Thus, we are in some sense perturbing around the vacuum φ 1.
We can interpret the force between the kink and anti-kink diagrammatically.
From the quantum field theory point of view, we can think of this force as
being due to meson exchange, and we can try to invent a Feynman diagram
calculus that involves solitons and mesons. This is a bit controversial, but at
least heuristically, we can introduce new propagators representing solitons, using
double lines, and draw the interaction as
¯
K
K
So what happens to this soliton? The force we derived was positive. So the
kink is made to move to the right. By symmetry, we will expect the anti-kink to
move towards the left. They will collide!
What happens when they collide? All our analysis so far assumed the kinks
were well-separated, so everything breaks down. We can only understand this
phenomenon numerically. After doing some numerical simulations, we see that
there are two regimes:
If the kinks are moving slowly, then they will annihilate into meson radia-
tion.
If the kinks are moving very quickly, then they often bounce off each other.
1.4 Quantization of kink motion
We now briefly talk about how to quantize kinks. The most naive way of doing
so is pretty straightforward. We use the moduli space approximation, and then
we have a very simple kink Lagrangian.
L =
1
2
M ˙a
2
.
This is just a free particle moving in
R
with mass
M
. This
a
is known as the
collective coordinate of the kink. Quantizing a free particle is very straightforward.
It is just IB Quantum Mechanics. For completeness, we will briefly outline this
procedure.
We first put the system in Hamiltonian form. The conjugate momentum to
a is given by
P = M ˙a.
Then the Hamiltonian is given by
H = P ˙a L =
1
2M
P
2
.
Then to quantize, we replace
P
by the operator
i~
a
. In this case, the quantum
Hamiltonian is given by
H =
~
2
2M
2
a
2
.
A wavefunction is a function of
a
and
t
, and this is just ordinary QM for a single
particle.
As usual, the stationary states are given by
ψ(a) = e
iκa
,
and the momentum and energy (eigenvalues) are
P = ~κ, H = E =
~
2
κ
2
2M
=
P
2
2M
.
Is this actually “correct”? Morally speaking, we really should quantize the
complete 1 + 1 dimensional field theory. What would this look like?
In normal quantum field theory, we consider perturbations around a vacuum
solution, say
φ
1, and we obtain mesons. Here if we want to quantize the
kink solution, we should consider field oscillations around the kink. Then the
solution contains both a kink and a meson. These mesons give rise to quantum
corrections to the kink mass M .
Should we be worried about these quantum corrections? Unsurprisingly, it
turns out these quantum corrections are of the order of the meson mass. So we
should not be worried when the meson mass is small.
Meson-kink scattering can also be studied in the full quantum theory. To
first approximation, since the kink is heavy, mesons are reflected or transmitted
with some probabilities, while the momentum of the kink is unchanged. But
when we work to higher orders, then of course the kink will move as a result.
This is all rather complicated.
For more details, see Rajaraman’s Solitons and Instantons, or Weinberg’s
Classical Solutions in Quantum Field Theory.
The thing that is really hard to understand in the quantum field theory
is kink-antikink pair production. This happens in meson collisions when the
mesons are very fast, and the theory is highly relativistic. What we have done
so far is perturbative and makes the non-relativistic approximation to get the
adiabatic picture. It is very difficult to understand the highly relativistic regime.
1.5 Sine-Gordon kinks
We end the section by briefly talking about kinks in a different theory, namely
the sine-Gordon theory. In this theory, kinks are often known as solitons instead.
The sine-Gordon theory is given by the potential
U(φ) = 1 cos φ.
Again, we suppress coupling constants, but it is possible to add them back.
The potential looks like
φ
U(φ)
2π 4π2π4π
Now there are infinitely many distinct vacua. In this case, we find we need to
pick W such that
dW
dφ
= 2 sin
1
2
φ.
Static sine-Gordon kinks
To find the static kinks in the sine-Gordon theory, we again look at the Bogomolny
equation. We have to solve
dφ
dx
= 2 sin
1
2
φ.
This can be solved. This involves integrating a
csc
, and ultimately gives us a
solution
φ(x) = 4 tan
1
e
xa
.
We can check that this solution interpolates between 0 and 2π.
x
φ
0
2π
a
Unlike the
φ
4
theory, dynamical multi-kink solutions exist here and can be
derived exactly. One of the earlier ways to do so was via acklund transforms,
but that was very complicated. People later invented better methods, but they
are still not very straightforward. Nevertheless, it can be done. Ultimately, this
is due to the sine-Gordon equation being integrable. For more details, refer to
the IID Integrable Systems course.
Example. There is a two-kink solution
φ(x, t) = 4 tan
1
v sinh γx
cosh γvt
,
where, as usual, we have
γ = (1 v
2
)
1/2
.
For v = 0.01, this looks like
x
φ
2π
2π
t = 0
t = ±400
Note that since
φ
(
x, t
) =
φ
(
x, t
), we see that this solution involves two
solitons at first approaching each other, and then later bouncing off. Thus, the
two kinks repel each other. When we did kinks in
φ
4
theory, we saw that a kink
and an anti-kink attracted, but here there are two kinks, which is qualitatively
different.
We can again compute the force just like the
φ
4
theory, but alternatively,
since we have a full, exact solution, we can work it out directly from the solution!
The answers, fortunately, agree. If we do the computations, we find that the
point of closest approach is 2 log
2
v
if v is small.
There are some important comments to make. In the sine-Gordon theory, we
can have very complicated interactions between kinks and anti-kinks, and these
can connect vastly different vacua. However, static solutions must join 2
and
2(
n ±
1)
π
for some
n
, because if we want to join vacua further apart, we will
have more than one kink, and they necessarily interact.
If we have multiple kinks and anti-kinks, then each of these things can have
their own velocity, and we might expect some very complicated interaction
between them, such as annihilation and pair production. But remarkably, the
interactions are not complicated. If we try to do numerical simulations, or use
the exact solutions, we see that we do not have energy loss due to “radiation”.
Instead, the solitons remain very well-structured and retain their identities. This,
again, is due to the theory being integrable.
Topology of the sine-Gordon equation
There are also a lot of interesting things we can talk about without going into
details about what the solutions look like.
The important realization is that our potential is periodic in
φ
. For the
sine-Gordon theory, it is much better to think of this as a field modulo 2
π
, i.e.
as a function
φ : R S
1
.
In this language, the boundary condition is that
φ
(
x
) = 0
mod
2
π
as
x ±∞
.
Thus, instead of thinking of the kink as joining two vacua, we can think of it as
“winding around the circle” instead.
We can go further. Since the boundary conditions of
φ
are now the same on
two sides, we can join the ends of the domain
R
together, and we can think of
φ
as a map
φ : S
1
S
1
instead. This is a compactification of space.
Topologically, such maps are classified by their winding number , or the degree,
which we denote
Q
. This is a topological (homotopy) invariant of a map, and
is preserved under continuous deformations of the field. Thus, it is preserved
under time evolution of the field.
Intuitively, the winding number is just how many times we go around the
circle. There are multiple (equivalent) ways of making this precise.
The first way, which is the naive way, is purely topological. We simply have
to go back to the first picture, where we regard
φ
as a real value. Suppose the
boundary values are
φ(−∞) = 2n
π, φ() = 2n
+
π.
Then we set the winding number to be Q = n
+
n
.
Topologically, we are using the fact that
R
is the universal covering space of
the circle, and thus we are really looking at the induced map on the fundamental
group of the circle.
Example. As we saw, a single kink has Q = 1.
x
φ
0
2π
Thus, we can think of the Q as the net soliton number.
But this construction we presented is rather specific to maps from
S
1
to
S
1
.
We want something more general that can be used for more complicated systems.
We can do this in a more “physics” way. We note that there is a topological
current
j
µ
=
1
2π
ε
µν
ν
φ,
where
ε
µν
is the anti-symmetric tensor in 1 + 1 dimensions, chosen so that
ε
01
= 1.
In components, this is just
j
µ
=
1
2π
(
x
φ,
t
φ).
This is conserved because of the symmetry of mixed partial derivatives, so that
µ
j
µ
=
1
2π
ε
µν
µ
ν
φ = 0.
As usual, a current induces a conserved charge
Q =
Z
−∞
j
0
dx =
1
2π
Z
−∞
x
φ dx =
1
2π
(φ() φ(−∞)) = n
+
n
,
which is the formula we had earlier.
Note that all these properties do not depend on
φ
satisfying any field equa-
tions! It is completely topological.
Finally, there is also a differential geometry way of defining
Q
. We note that
the target space S
1
has a normalized volume form ω so that
Z
S
1
ω = 1.
For example, we can take
ω =
1
2π
dφ.
Now, given a mapping
φ
:
R S
1
, we can pull back the volume form to obtain
φ
ω =
1
2π
dφ
dx
dx.
We can then define the degree of the map to be
Q =
Z
φ
ω =
1
2π
Z
−∞
dφ
dx
dx.
This is exactly the same as the formula we obtained using the current!
Note that even though the volume form is normalized on
S
1
and has integral
1, the integral when pulled back is not 1. We can imagine this as saying if we
wind around the circle
n
times, then after pulling back, we would have pulled
back
n
“copies” of the volume form, and so the integral will be
n
times that of
the integral on S
1
.
We saw that these three definitions gave the same result, and different
definitions have different benefits. For example, in the last two formulations, it
is not a priori clear that the winding number has to be an integer, while this is
clear in the first formulation.
2 Vortices
We are now going to start studying vortices. These are topological solitons in
two space dimensions. While we mostly studied
φ
4
kinks last time, what we are
going to do here is more similar to the sine-Gordon theory than the
φ
4
theory,
as it is largely topological in nature.
A lot of the computations we perform in the section are much cleaner when
presented using the language of differential forms. However, we shall try our
best to provide alternative versions in coordinates for the easily terrified.
2.1 Topological background
Sine-Gordon kinks
We now review what we just did for sine-Gordon kinks, and then try to develop
some analogous ideas in higher dimension. The sine-Gordon equation is given by
2
θ
t
2
2
θ
x
2
+ sin θ = 0.
We want to choose boundary conditions so that the energy has a chance to
be finite. The first part is, of course, to figure out what the energy is. The
energy-momentum conservation equation given by Noether’s theorem is
t
θ
2
t
+ θ
2
x
2
+ (1 cos θ)
+
x
(θ
t
θ
x
) =
µ
P
µ
= 0.
The energy we will be considering is thus
E =
Z
R
P
0
dx =
Z
R
θ
2
t
+ θ
2
x
2
+ (1 cos θ)
dx.
Thus, to obtain finite energy, we will want
θ
(
x
)
2
n
±
π
for some integers
n
±
as x ±∞. What is the significance of this n
±
?
Example. Consider the basic kink
θ
K
(x) = 4 tan
1
e
x
.
Picking the standard branch of tan
1
, this kink looks like
x
φ
0
2π
This goes from θ = 0 to θ = 2π.
To better understand this, we can think of
θ
as an angular variable, i.e. we
identify
θ θ
+ 2
for any
n Z
. This is a sensible thing because the energy
density and the equation etc. are unchanged when we shift everything by 2
.
Thus, θ is not taking values in R, but in R/2πZ
=
S
1
.
Thus, for each fixed t, our field θ is a map
θ : R S
1
.
The number
Q
=
n
+
n
equals the number of times
θ
covers the circle
S
1
on going from
x
=
−∞
to
x
= +
. This is the winding number, which is
interpreted as the topological charge.
As we previously discussed, we can express this topological charge as the
integral of some current. We can write
Q =
1
2π
Z
θ(R)
dθ =
1
2π
Z
−∞
dθ
dx
dx.
Note that this formula automatically takes into account the orientation. This is
the form that will lead to generalization in higher dimensions.
This function
dθ
dx
appearing in the integral has the interpretation as a topo-
logical charge density. Note that there is a topological conservation law
µ
j
µ
=
j
0
t
+
j
1
x
= 0,
where
j
0
= θ
x
, j
1
= θ
t
.
This conservation law is not a consequence of the field equations, but merely a
mathematical identity, namely the commutation of partial derivatives.
Two dimensions
For the sine-Gordon kink, the target space was a circle
S
1
. Now, we are concerned
with the unit disk
D = {(x
1
, x
2
) : |x|
2
< 1} R
2
.
We will then consider fields
Φ : D D.
In the case of a sine-Gordon kink, we still cared about moving solitons. However,
here we will mostly work with static solutions, and study fields at a fixed time.
Thus, there is no time variable appearing.
Using the canonical isomorphism
R
2
=
C
, we can think of the target space
as the unit disk in the complex plane, and write the field as
Φ = Φ
1
+ iΦ
2
.
However, we will usually view the D in the domain as a real space instead.
We will impose some boundary conditions. We pick any function
χ
:
S
1
R
,
and consider
g = e
: S
1
S
1
= D D.
Here
g
is a genuine function, and has to be single-valued. So
χ
must be single-
valued modulo 2π. We then require
Φ
D
= g = e
.
In particular, Φ must send the boundary to the boundary.
Now the target space
D
has a canonical choice of measure
1
2
. Then
we can expect the new topological charge to be given by
Q =
1
π
Z
D
1
2
=
1
π
Z
D
det
Φ
1
x
1
Φ
1
x
2
Φ
2
x
1
Φ
2
x
2
!
dx
1
dx
2
.
Thus, the charge density is given by
j
0
=
1
2
ε
ab
ε
ij
Φ
a
x
i
Φ
b
x
j
.
Crucially, it turns out this charge density is a total derivative, i.e. we have
j
0
=
V
i
x
i
for some function
V
. It is not immediately obvious this is the case. However, we
can in fact pick
V
i
=
1
2
ε
ab
ε
ij
Φ
a
Φ
b
x
j
.
To see this actually works, we need to use the anti-symmetry of
ε
ij
and observe
that
ε
ij
2
Φ
b
x
i
x
j
= 0.
Equivalently, using the language of differential forms, we view the charge density
j
0
as the 2-form
j
0
= dΦ
1
2
= d(Φ
1
2
) =
1
2
d(Φ
1
2
Φ
2
1
).
By the divergence theorem, we find that
Q =
1
2π
I
D
Φ
1
2
Φ
2
1
.
We then use that on the boundary,
Φ
1
= cos χ, Φ
2
= sin χ,
so
Q =
1
2π
I
D
(cos
2
χ + sin
2
χ) dχ =
1
2π
I
D
dχ = N.
Thus, the charge is just the winding number of g!
Now notice that our derivation didn’t really depend on our domain being
D
. It could have been any region bounded by a simple closed curve in
R
2
. In
particular, we can take it to be a disk D
R
of arbitrary radius R.
What we are actually interested in is a field
Φ : R
2
D.
We then impose asymptotic boundary conditions
Φ g = e
as |x| . We can still define the charge or degree by
Q =
1
π
Z
R
2
j
0
dx
1
dx
2
=
1
π
lim
R→∞
Z
D
R
j
0
dx
1
dx
2
This is then again the winding number of g.
It is convenient to rewrite this in terms of an inner product.
R
2
itself has an
inner product, and under the identification
R
2
=
C
, the inner product can be
written as
(a, b) =
¯ab + a
¯
b
2
.
Use of this expression allows calculations to be done efficiently if one makes use
of the fact that for real numbers a and b, we have
(a, b) = (ai, bi) = ab, (ai, b) = (a, bi) = 0.
In particular, we can evaluate
(iΦ, dΦ) = (iΦ
1
Φ
2
,
1
+ i
2
) = Φ
1
2
Φ
2
1
.
This is just (twice) the current V we found earlier! So we can write our charge
as
Q =
1
2π
lim
R→∞
I
|x|=R
(iΦ, dΦ).
We will refer to (
i
Φ
,
dΦ) as the current, and the corresponding charge density is
j
0
=
1
2
d(iΦ, dΦ).
This current is actually a familiar object from quantum mechanics: recall
that for the Schr¨odinger’s equation
i
Φ
t
=
1
2m
∆Φ + V (x, ∆ =
2
.
the probability
R
|
Φ
|
s
d
x
is conserved. The differential form of the probability
conservation law is
1
2
t
, Φ) +
1
2m
· (iΦ, Φ) = 0.
What appears in the flux term here is just the topological current!
2.2 Global U (1) Ginzburg–Landau vortices
We now put the theory into use. We are going to study Ginzburg–Landau vortices.
Our previous discussion involved a function taking values in the unit disk
D
.
We will not impose such a restriction on our vortices. However, we will later see
that any solution must take values in D.
The potential energy of the Ginzburg–Landau theory is given by
V (Φ) =
1
2
Z
R
2
(Φ, Φ) +
λ
4
(1 , Φ))
2
dx
1
dx
2
.
where λ > 0 is some constant.
Note that the inner product is invariant under phase rotation, i.e.
(e
a, e
b) = (a, b)
for χ R. So in particular, the potential satisfies
V (e
Φ) = V (Φ).
Thus, our theory has a global U(1) symmetry.
The Euler–Lagrange equation of this theory says
∆Φ =
λ
2
(1 |Φ|
2
.
This is the ungauged Ginzburg–Landau equation.
To justify the fact that our Φ takes values in
D
, we use the following lemma:
Lemma.
Assume Φ is a smooth solution of the ungauged Ginzburg–Landau
equation in some domain. Then at any interior maximum point
x
of
|
Φ
|
, we
have |Φ(x
)| 1.
Proof. Consider the function
W (x) = 1 |Φ(x)|
2
.
Then we want to show that
W
0 when
W
is minimized. We note that if
W
is
at a minimum, then the Hessian matrix must have non-negative eigenvalues. So,
taking the trace, we must have
W
(
x
)
0. Now we can compute
W
directly.
We have
W = 2(Φ, Φ)
W =
2
W
= 2(Φ, ∆Φ) 2(Φ, Φ)
= λ|Φ|
2
W 2|∇Φ|
2
.
Thus, we can rearrange this to say
2|∇Φ|
2
+ W = λ|Φ|
2
W.
But clearly 2
|∇
Φ
|
2
0 everywhere, and we showed that
W
(
x
)
0. So we
must have W (x
) 0.
By itself, this doesn’t force
|
Φ
|
[0
,
1], since we could imagine
|
Φ
|
having no
maximum. However, if we prescribe boundary conditions such that
|
Φ
|
= 1 on
the boundary, then this would indeed imply that
|
Φ
|
1 everywhere. Often, we
can think of Φ as some “complex order parameter”, in which case the condition
|Φ| 1 is very natural.
The objects we are interested in are vortices.
Definition
(Ginzburg–Landau vortex)
.
A global Ginzburg–Landau vortex of
charge
N >
0 is a (smooth) solution of the ungauged Ginzburg–Landau equation
of the form
Φ = f
N
(r)e
iNθ
in polar coordinates (r, θ). Moreover, we require that f
N
(r) 1 as r .
Note that for Φ to be a smooth solution, we must have
f
N
(0) = 0. In fact, a
bit more analysis shows that we must have
f
N
=
O
(
r
N
) as
r
0. Solutions for
f
N
do exist, and they look roughly like this:
r
f
N
In the case of
N
= 1, we can visualize the field Φ as a vector field on
C
.
Then it looks like
This is known as a 2-dimensional hedgehog.
For general
N
, it might be more instructive to look at how the current looks
like. Recall that the current is defined by (
i
Φ
,
dΦ). We can write this more
explicitly as
(iΦ, dΦ) = (if
N
e
iNθ
, (df
N
)e
iNθ
+ if
N
N dθe
iNθ
)
= (if
N
, df
N
+ if
N
N dθ).
We note that
if
N
and d
f
N
are orthogonal, while
if
N
and
if
N
N
d
θ
are parallel.
So the final result is
(iΦ, dΦ) = f
2
N
N dθ.
So the current just looks this:
As
|x|
, we have
f
N
1. So the winding number is given as before,
and we can compute the winding number of this system to be
1
2π
lim
R→∞
I
(iΦ, dΦ) =
1
2π
lim
R→∞
I
f
2
N
N dθ = N.
The winding number of these systems is a discrete quantity, and can make the
vortex stable.
This theory looks good so far. However, it turns out this model has a problem
the energy is infinite! We can expand out
V
(
f
N
e
iNθ
), and see it is a sum of a
few non-negative terms. We will focus on the
θ
term. We obtain
V (f
N
e
iNθ
)
Z
1
r
2
Φ
θ
2
r dr dθ
= N
2
Z
1
r
2
f
2
N
r dr dθ
= 2πN
2
Z
0
f
2
N
r
dr.
Since f
N
1 as r , we see that the integral diverges logarithmically.
This is undesirable physically. To understand heuristically why this occurs,
decompose into two components a mode parallel to Φ and a mode
perpendicular to Φ. For a vortex solution these correspond to the radial and
angular modes respectively. We will argue that for fluctuations the parallel mode
is massive, while the perpendicular mode is massless. Now given that we just
saw that the energy divergence of the vortex arises from the angular part of the
energy, we see that it is the massless mode that leads to problems. We will see
below that in gauge theories, the Higgs mechanism serves to make all modes
massive, thus allowing for finite energy vortices.
We can see the difference between massless and massive modes very explicitly
in a different setting, corresponding to Yukawa mesons. Consider the equation
u + M
2
u = f.
Working in three dimensions, the solution is given by
u(x) =
1
4π
Z
e
M|xy|
|x y|
f(y) dy.
Thus, the Green’s function is
G(x) =
e
M|x|
4π|x|
.
If the system is massless, i.e.
M
= 0, then this decays as
1
|x|
. However, if the
system is massive with
M >
0, then this decays exponentially as
|x|
. In
the nonlinear setting the exponential decay which is characteristic of massive
fundamental particles can help to ensure decay of the energy density at a rate
fast enough to ensure finite energy of the solution.
So how do we figure out the massive and massless modes? We do not have
a genuine decomposition of Φ itself into “parallel” and “perpendicular” modes,
because what is parallel and what is perpendicular depends on the local value of
Φ.
Thus, to make sense of this, we have to consider small fluctuations around a
fixed configuration Φ. We suppose Φ is a solution to the field equations. Then
δV
δΦ
= 0. Thus, for small variations Φ 7→ Φ + εϕ, we have
V + εϕ) = V (Φ) + ε
2
Z
|∇ϕ|
2
+ λ(ϕ, Φ)
2
2λ
4
(1 |Φ|
2
)|ϕ|
2
dx + O(ε
3
).
Ultimately, we are interested in the asymptotic behaviour
|x|
, in which
case 1
|
Φ
|
2
0. Moreover,
|
Φ
|
1 implies (
ϕ,
Φ) becomes a projection along
the direction of Φ. Then the quadratic part of the potential energy density for
fluctuations becomes approximately
|∇ϕ|
2
+ λ|ϕ
parallel
|
2
for large
x
. Thus, for
λ >
0, the parallel mode is massive, with corresponding
“Yukawa” mass parameter
M
=
λ
, while the perpendicular mode is massless.
The presence of the massless mode is liable to produce a soliton with slow
algebraic decay at spatial infinity, and hence infinite total energy. This is all
slightly heuristic, but is a good way to think about the issues. When we study
vortices that are gauged, i.e. coupled to the electromagnetic field, we will see
that the Higgs mechanism renders all components massive, and this problem
does not arise.
2.3 Abelian Higgs/Gauged Ginzburg–Landau vortices
We now consider a theory where the complex scalar field Φ is coupled to a
magnetic field. This is a U(1) gauge theory, with gauge potential given by a
smooth real 1-form
A = A
1
dx
1
+ A
2
dx
2
.
The coupling between Φ and
A
is given by minimal coupling: this is enacted by
introduction of the covariant derivative
DΦ = D
A
Φ = dΦ iAΦ =
2
X
j=1
D
j
Φ dx
j
.
To proceed, it is convenient to have a list of identities involving the covariant
derivative.
Proposition.
If
f
is a smooth real-valued function, and Φ and Ψ are smooth
complex scalar fields, then
D(fΦ) = (df) Φ + f ,
d(Φ, Ψ) = (DΦ, Ψ) + , DΨ).
(Here (
·
) is the real inner product defined above.) In coordinates, these take
the form
D
j
(fΦ) = (
j
f) Φ + f D
j
Φ
j
, Ψ) = (D
j
Φ, Ψ) + , D
j
Ψ).
The proofs just involve writing all terms out. The first rule is a version of
the Leibniz rule, while the second, called unitarity, is analogous to the fact that
if V, W are smooth vector fields on a Riemannian manifold, then
k
g(V, W ) = g(
k
V, W ) + g(V,
k
W )
for the Levi-Civita connection associated to a Riemannian metric g.
The curvature term is given by the magnetic field.
Definition
(Magnetic field/curvature)
.
The magnetic field, or curvature is given
by
B =
1
A
2
2
A
1
.
We can alternatively think of it as the 2-form
F = dA = B dx
1
dx
2
.
The formulation in terms of differential forms is convenient for computations,
because we don’t have to constrain ourselves to working in Cartesian coordinates
for example, polar coordinates may be more convenient.
Proposition. If Φ is a smooth scalar field, then
(D
1
D
2
D
2
D
1
)Φ = iBΦ.
The proof is again a direct computation. Alternatively, we can express this
without coordinates. We can extend D to act on
p
-forms by letting
A
act as
A
.
Then this result says
Proposition.
DDΦ = iF Φ.
Proof.
DDΦ = (d iA)(dΦ iAΦ)
= d
2
Φ id(AΦ) iA A A Φ
= id(AΦ) iA
= idA Φ + iA iA
= i(dA) Φ
= iF Φ.
The point of introducing the covariant derivative is that we can turn the global
U(1) invariance into a local one. Previously, we had a global U(1) symmetry,
where our field is unchanged when we replace Φ
7→
Φ
e
for some constant
χ R
.
With the covariant derivative, we can promote this to a gauge symmetry.
Consider the simultaneous gauge transformations
Φ(x) 7→ e
(x)
Φ(x)
A(x) 7→ A(x) + dχ(x).
Then the covariant derivative of Φ transforms as
(d iA7→ (d i(A + dχ))(Φe
)
= (dΦ + iΦdχ i(A + dχ)Φ)e
= e
(d iA.
Similarly, the magnetic field is also invariant under gauge transformations.
As a consequence, we can write down energy functionals that are invariant
under these gauge transformations. In particular, we have (using the real inner
product defined above)
(DΦ, DΦ) 7→ (e
, e
DΦ) = (DΦ, DΦ).
So we can now write down the gauged Ginzburg–Landau energy
V
λ
(A, Φ) =
1
2
Z
R
2
B
2
+ ||
2
+
λ
4
(1 |Φ|
2
)
2
d
2
x.
This is then manifestly gauge invariant.
As before, the equations of motion are given by the Euler–Lagrange equations.
Varying Φ, we obtain
(D
2
1
+ D
2
2
λ
2
(1 |Φ|
2
)Φ = 0.
This is just like the previous vortex equation in the ungauged case, but since we
have the covariant derivative, this is now coupled to the gauge potential
A
. The
equations of motion satisfied by A are
2
B = (iΦ, D
1
Φ)
1
B = (iΦ, D
2
Φ).
These are similar to one of Maxwell’s equation the one relating the curl of
the magnetic field to the current.
It is again an exercise to derive these. We refer to the complete system as
the gauged Ginzburg–Landau, or Abelian Higgs equations. In deriving them, it
is helpful to use the previous identities such as
j
, Ψ) = (D
j
Φ, Ψ) + , D
j
Ψ).
So we get the integration by parts formula
Z
R
2
(D
j
Φ, Ψ) d
2
x =
Z
, D
j
Ψ) d
2
x
under suitable boundary conditions.
Lemma.
Assume Φ is a smooth solution of the gauged Ginzburg–Landau
equation in some domain. Then at any interior maximum point
x
of
|
Φ
|
, we
have |Φ(x
)| 1.
Proof. Consider the function
W (x) = 1 |Φ(x)|
2
.
Then we want to show that
W
0 when
W
is minimized. We note that if
W
is
at a minimum, then the Hessian matrix must have non-negative eigenvalues. So,
taking the trace, we must have
W
(
x
)
0. Now we can compute
W
directly.
We have
j
W = 2(Φ, D
j
Φ)
W =
j
j
W
= 2(Φ, D
j
D
j
Φ) 2(D
j
Φ, D
j
Φ)
= λ|Φ|
2
W 2|∇Φ|
2
.
Thus, we can rearrange this to say
2|∇Φ|
2
+ W = λ|Φ|
2
W.
But clearly 2
|∇
Φ
|
2
0 everywhere, and we showed that
W
(
x
)
0. So we
must have W (x
) 0.
As before, this suggests we interpret
|
Φ
|
as an order parameter. This model
was first used to describe superconductors. The matter can either be in a
“normal” phase or a superconducting phase, and
|
Φ
|
measures how much we are
in the superconducting phase.
Thus, in our model, far away from the vortices, we have
|
Φ
|
1, and so we
are mostly in the superconducting phase. The vortices represent a breakdown of
the superconductivity. At the core of the vortices, we have |Φ| = 0, and we are
left with completely normal matter. Usually, this happens when we have a strong
magnetic field. In general, a magnetic field cannot penetrate the superconductor
(the “Meissner effect”), but if it is strong enough, it will cause such breakdown
in the superconductivity along vortex “tubes”.
Radial vortices
Similar to the ungauged case, for
λ >
0, there exist vortex solutions of the form
Φ = f
N
(r)e
iNθ
A = Nα
N
(r) dθ.
The boundary conditions are f
N
, α
N
1 as r and f
N
, α
N
0 as r 0.
Let’s say a few words about why these are sensible boundary conditions from
the point of view of energy. We want
λ
Z
R
2
(1 |Φ|
2
)
2
< ,
and this is possible only for
f
N
1 as
r
. What is less obvious is that we
also need α
N
1. We note that we have
D
θ
Φ =
Φ
θ
iA
θ
Φ = (iNf
N
iNα
N
f
N
)e
iNθ
.
We want this to approach 0 as r . Since f
N
1, we also need α
N
1.
The boundary conditions at 0 can be justified as before, so that the functions
are regular at 0.
Topological charge and magnetic flux
Let’s calculate the topological charge. We have (assuming sufficiently rapid
approach to the asymptotic values as r )
Q =
1
π
Z
R
2
j
0
(Φ) d
2
x
= lim
R→∞
1
2π
I
|x|=R
(iΦ, dΦ)
= lim
R→∞
1
2π
I
(if
N
e
iNθ
, iNf
N
e
iNθ
) dθ
=
1
2π
· N lim
R→∞
Z
2π
0
f
2
N
dθ
= N.
Previously, we understood
N
as the “winding number”, and it measures how
“twisted” our field was. However, we shall see shortly that there is an alternative
interpretation of this
N
. Previously, in the sine-Gordon theory, we could think of
N
as the number of kinks present. Similarly, here we can think of this
N
-vortex
as a superposition of
N
vortices at the origin. In the case of
λ
= 1, we will
see that there are static solutions involving multiple vortices placed at different
points in space.
We can compute the magnetic field and total flux as well. It is convenient to
use the d
A
definition, as we are not working in Cartesian coordinates. We have
dA = Nα
0
N
(r) dr dθ =
1
r
Nα
0
N
dx
1
dx
2
.
Thus it follows that
B =
1
A
2
2
A
1
=
N
r
α
0
N
.
Working slightly more generally, we assume given a smooth finite energy con-
figuration
A,
Φ, and suppose in addition that
|
Φ
|
2
1 and
r|A|
is bounded as
r , and also that D
θ
Φ = o(r
1
). Then we find that
(dΦ)
θ
= iA
θ
Φ + o(r
1
).
Then if we integrate around a circular contour, since only the angular part
contributes, we obtain
I
|x|=R
(iΦ, dΦ) =
I
|x|=R
(iΦ, iA
θ
Φ) dθ + o(1) =
I
|x|=R
A + o(1).
Note that here we are explicitly viewing
A
as a differential form so that we can
integrate it. We can then note that
|x|
=
R
is the boundary of the disk
|x| R
.
So we can apply Stokes’ theorem and obtain
I
|x|=R
(iΦ, dΦ) =
Z
|x|≤R
dA =
Z
|x|≤R
B d
2
x.
Now we let R to obtain
lim
R→∞
Z
xR
B d
2
x = lim
R→∞
I
|x|=R
A = lim
R→∞
I
|x|=R
(iΦ, dΦ) = 2πQ.
Physically, what this tells us then is that there is a relation between the
topological winding number and the magnetic flux. This is a common property
of topological gauge theories. In mathematics, this is already well known
it is the fact that we can compute characteristic classes of vector bundles by
integrating the curvature, as discovered by Chern.
Behaviour of vortices as r 0
We saw earlier that for reasons of regularity, it was necessary that
f
N
, α
N
0
as r 0.
But actually, we must have
f
N
r
N
as
r
0. This has as a consequence
that Φ
(
re
)
N
=
z
N
. So the local appearance of the vortex is the zero of an
analytic function with multiplicity N.
To see this, we need to compute that the Euler–Lagrange equations are
f
00
N
(r)
1
r
f
0
N
(r) +
(N α
N
)
2
r
2
f
N
=
λ
2
f
N
(1 f
2
N
).
With the boundary condition that
f
N
and
α
N
vanish at
r
= 0, we can approxi-
mate this locally as
f
00
N
1
r
f
0
N
+
N
2
r
2
f
N
0
since we have a
1
r
2
on the left hand side. The approximating equation is
homogeneous, and the solutions are just
f
N
= r
±N
.
So for regularity, we want the one that 0, so f
N
r
N
as r 0.
2.4 Bogomolny/self-dual vortices and Taubes’ theorem
As mentioned, we can think of the radial vortex solution as a collection of
N
vortices all superposed at the origin. Is it possible to have separated vortices all
over the plane? Naively, we would expect that the vortices exert forces on each
other, and so we don’t get a static solution. However, it turns out that in the
λ
= 1 case, there do exist static solutions corresponding to vortices at arbitrary
locations on the plane.
This is not obvious, and the proof requires some serious analysis. We will
not do the analysis, which requires use of Sobolev spaces and PDE theory.
However, we will do all the non-hard-analysis part. In particular, we will obtain
Bogomolny bounds as we did in the sine-Gordon case, and reduce the problem
to finding solutions of a single scalar PDE, which can be understood with tools
from calculus of variations and elliptic PDEs.
Recall that for the sine-Gordon kinks, we needed to solve
θ
00
= sin θ,
with boundary conditions
θ
(
x
)
0 or 2
π
as
x ±∞
. The only solutions we
found were
θ
K
(x X)
for any
X R
. This
X
is interpreted as the location of the kink. So the moduli
space of solutions is M = R.
We shall get a similar but more interesting description for the
λ
= 1 vortices.
This time, the moduli space will be
C
N
, given by
N
complex parameters
describing the solutions.
Theorem
(Taubes’ theorem)
.
For
λ
= 1, the space of (gauge equivalence classes
of) solutions of the Euler–Lagrange equations
δV
1
= 0 with winding number
N
is M
=
C
N
.
To be precise, given
N N
and an unordered set of points
{Z
1
, ··· , Z
N
}
,
there exists a smooth solution
A
(
x
;
Z
1
, ··· , Z
N
) and Φ(
x
;
Z
1
, ··· , Z
n
) which
solves the Euler-Lagrange equations
δV
1
= 0, and also the so-called Bogomolny
equations
D
1
Φ + iD
2
Φ = 0, B =
1
2
(1 |Φ|
2
).
Moreover, Φ has exactly
N
zeroes
Z
1
, ··· , Z
N
counted with multiplicity, where
(using the complex coordinates z = x
1
+ ix
2
)
Φ(x; Z
1
, ··· , Z
N
) c
j
(z Z
j
)
n
j
as
z Z
j
, where
n
j
=
|{k
:
Z
k
=
Z
j
}|
is the multiplicity and
c
j
is a nonzero
complex number.
This is the unique such solution up to gauge equivalence. Furthermore,
V
1
(A( ·, Z
1
, ··· , Z
N
), Φ( ·; Z
1
, ··· , Z
N
)) = πN ()
and
1
2π
Z
R
2
B d
2
x = N = winding number.
Finally, this gives all finite energy solutions of the gauged Ginzburg–Landau
equations.
Note that it is not immediately clear from our description that the mod-
uli space is
C
N
. It looks more like
C
N
quotiented out by the action of the
permutation group
S
N
. However, the resulting quotient is still isomorphic to
C
N
. (However, it is important for various purposes to remember this quotient
structure, and to use holomorphic coordinates which are invariant under the
action of the permutation group the elementary symmetric polynomials in
{Z
1
, . . . , Z
n
}.
There is a lot to be said about this theorem. The equation (
) tells us the
energy is just a function of the number of particles, and does not depend on
where they are. This means there is no force between the vortices. In situations
like this, it is said that the Bogomolny bound is saturated. The final statement
suggests that the topology is what is driving the existence of the vortices, as we
have already seen. The reader will find it useful to work out the corresponding
result in the case of negative winding number (in which case the holomorphicity
condition becomes anti-holomorphicity, and the sign of the magnetic field is
reversed in the Bogomolny equations).
Note that the Euler–Lagrange equations themselves are second-order equa-
tions. However, the Bogomolny equations are first order. In general, this is a
signature that suggests that interesting mathematical structures are present.
We’ll discuss three crucial ingredients in this theorem, but we will not
complete the proof, which involves more analysis than is a pre-requisite for this
course. The proof can be found in Chapter 3 of Jaffe and Taubes’s Vortices and
Monopoles.
Holomorphic structure
When there are Bogomolny equations, there is often some complex analysis
lurking behind. We can explicitly write the first Bogomolny equation as
D
1
Φ + iD
2
Φ =
Φ
x
1
+ i
Φ
x
2
i(A
1
+ iA
2
)Φ = 0.
Recall that in complex analysis, holomorphic functions can be characterized as
complex-valued functions which are continuously differentiable (in the real sense)
and also satisfy the Cauchy–Riemann equations
f
¯z
=
1
2
f
x
1
+ i
f
x
2
= 0.
So we think of the first Bogomolny equation as the covariant Cauchy–Riemann
equations. It is possible to convert this into the standard Cauchy–Riemann
equations to deduce the local behaviour at Φ.
To do so, we write
Φ = e
ω
f.
Then
f
¯z
= e
ω
Φ
¯z
ω
¯z
Φ
= e
ω
i(A
i
+ iA
2
)
2
ω
¯z
Φ.
This is equal to 0 if ω satisfies
ω
¯z
= i
A
1
+ iA
2
2
.
So the question is can we solve this? It turns out we can always solve this
equation, and there is an explicit formula for the solution. In general, if
β
is
smooth, then the equation
w
¯z
= β,
has a smooth solution in the disc {z : |z| < r}, given by
ω(z, ¯z) =
1
2πi
Z
|w|<r
β(w)
w z
dw d ¯w.
A proof can be found in the book Griffiths and Harris on algebraic geometry,
on page 5. So we can write
Φ = e
ω
f
where
f
is holomorphic. Since
e
ω
is never zero, we can apply all our knowledge
of holomorphic functions to
f
, and deduce that Φ has isolated zeroes, where
Φ (z Z
j
)
n
j
for some integer power n
j
.
The Bogomolny equations
We’ll now show that (
A,
Φ) satisfies
V
1
(
A,
Φ) =
πN
0 iff it satisfies the
Bogomolny equations, i.e.
D
1
Φ + iD
2
Φ = 0, B =
1
2
(1 |Φ|
2
).
We first consider the simpler case of the sine-Gordon equation. As in the
φ
4
kinks, to find soliton solutions, we write the energy as
E =
Z
−∞
1
2
θ
2
x
+ (1 cos θ)
dx
=
1
2
Z
−∞
θ
2
x
+ 4 sin
2
θ
2
dx
=
1
2
Z
−∞
θ
x
2 sin
θ
2
2
+ 4θ
x
sin
θ
2
!
dx
=
1
2
Z
−∞
θ
x
2 sin
θ
2
2
dx +
Z
−∞
x
4 cos
θ
2
dx
=
1
2
Z
−∞
θ
x
2 sin
θ
2
2
dx +
4 cos
θ(+)
2
+ 4 cos
θ(−∞)
2
.
We then use the kink asymptotic boundary conditions to obtain, say,
θ
(+
) = 2
π
and
θ
(
−∞
) = 0. So the boundary terms gives 8. Thus, subject to these boundary
conditions, we can write the sine-Gordon energy as
E =
1
2
Z
−∞
θ
x
2 sin
θ
2
2
dx + 8.
Thus, if we try to minimize the energy, then we know the minimum is at least 8,
and if we could solve the first-order equation
θ
x
= 2
sin
θ
2
, then the minimum
would be exactly 8. The solution we found does satisfy this first-order equation.
Moreover, the solutions are all of the form
θ(x) = θ
K
(x X), θ
K
(x) = 4 arctan e
x
.
Thus, we have shown that the minimum energy is 8, and the minimizers are all
of this form, parameterized by X R.
We want to do something similar for the Ginzburg–Landau theory. In order
to make use of the discussion above of the winding number
N
, we will make
the same standing assumptions as used in that discussion, but it is possible
to generalize the conclusion of the following result to arbitrary finite energy
configurations with an appropriate formulation of the winding number.
Lemma. We have
V
1
(A, Φ) =
1
2
Z
R
2
B
1
2
(1 |Φ|
2
)
2
+ 4|
¯
A
Φ|
2
!
d
2
x + πN,
where
¯
A
Φ =
1
2
(D
1
Φ + iD
2
Φ).
It is clear that the desired result follows from this.
Proof. We complete the square and obtain
V
1
(A, Φ) =
1
2
Z
B
1
2
(1 |Φ|
2
)
2
+ B(1 |Φ|
2
) + |D
1
Φ|
2
+ |D
2
Φ|
2
!
d
2
x.
We now dissect the terms one by one. We first use the definition of
B
d
x
1
d
x
2
=
dA and integration by parts to obtain
Z
R
2
(1 |Φ|
2
) dA =
Z
R
2
d(1 |Φ|
2
) A = 2
Z
R
2
, DΦ) A.
Alternatively, we can explicitly write
Z
R
2
B(1 |Φ|
2
) d
2
x =
Z
R
2
(
1
A
2
2
A
1
)(1 |Φ|
2
) d
2
x
=
Z
R
2
(A
2
1
|Φ|
2
A
1
2
|Φ|
2
) d
2
x
= 2
Z
R
2
A
2
, D
1
Φ) A
1
, D
2
Φ).
Ultimately, we want to obtain something that looks like
|
¯
A
Φ
|
2
. We can write
this out as
(D
1
Φ + iD
2
Φ, D
1
Φ + iD
2
Φ) = |D
1
Φ|
2
+ |D
2
Φ|
2
+ 2(D
1
Φ, iD
2
Φ).
We note that
i
Φ and Φ are always orthogonal, and
A
i
is always a real coefficient.
So we can write
(D
1
Φ, iD
2
Φ) = (
1
Φ iA
1
Φ, i∂
2
Φ + A
2
Φ)
= (
1
Φ, i∂
2
Φ) + A
2
,
1
Φ) A
1
,
2
Φ).
We now use again the fact that
, i
Φ) = 0 to replace the usual derivatives with
the covariant derivatives. So we have
(D
1
Φ, iD
2
Φ) = (
1
Φ, i∂
2
Φ) + A
2
, D
1
Φ) A
1
, D
2
Φ).
This tells us we have
Z
B(1 |Φ|
2
) + |D
1
Φ|
2
+ |D
2
Φ|
2
d
2
x =
Z
4|
¯
A
Φ|
2
+ 2(
1
Φ, i∂
2
Φ)
d
2
x.
It then remains to show that (
1
Φ, i∂
2
Φ) = j
0
(Φ). But we just write
(
1
Φ
1
+ i∂
1
Φ
2
,
2
Φ
2
+ i∂
2
Φ
1
) = (
1
Φ
1
,
2
Φ
2
) + (
1
Φ
2
,
2
Φ
1
)
= j
0
(Φ)
= det
1
Φ
1
2
Φ
1
1
Φ
2
2
Φ
2
Then we are done.
Corollary.
For any (
A,
Φ) with winding number
N
, we always have
V
1
(
A,
Φ)
πN, and those (A, Φ) that achieve this bound are exactly those that satisfy
¯
A
Φ = 0, B =
1
2
(1 |Φ|
2
).
Reduction to scalar equation
The remaining part of Taubes’ theorem is to prove the existence of solutions to
these equations, and that they are classified by
N
unordered complex numbers.
This is the main analytic content of the theorem.
To do so, we reduce the two Bogomolny equations into a scalar equation.
Note that we have
D
1
Φ + iD
2
Φ = (
1
Φ + i∂
2
Φ) i(A
1
+ iA
2
)Φ = 0.
So we can write
A
1
+ iA
2
= i(
1
+ i∂
2
) log Φ.
Thus, once we’ve got Φ, we can get A
1
and A
2
.
The next step is to use gauge invariance. Under gauge invariance, we can fix
the phase of Φ to anything we want. We write
Φ = e
1
2
(u+)
.
Then |Φ|
2
= e
u
.
We might think we can get rid of
θ
completely. However, this is not quite
true, since the argument is not well-defined at a zero of Φ, and in general we
cannot get rid of θ by a smooth gauge transformation. But since
Φ c
j
(z Z
j
)
n
j
near Z
j
, we expect we can make θ look like
θ = 2
N
X
j=1
arg(z Z
j
).
We will assume we can indeed do so. Then we have
A
1
=
1
2
(
2
u +
1
θ), A
2
=
1
2
(
1
u
2
θ).
We have now solved for
A
using the first Bogomolny equation. We then use
this to work out
B
and obtain a scalar equation for
u
by the second Bogomolny
equation.
Theorem.
In the above situation, the Bogomolny equation
B
=
1
2
(1
|
Φ
|
2
) is
equivalent to the scalar equation for u
u + (e
u
1) = 4π
N
X
j=1
δ
Z
j
.
This is known as Taubes’ equation.
Proof. We have
B =
1
A
2
2
A
1
=
1
2
2
1
u
1
2
2
2
u +
1
2
(
1
2
2
1
)θ
=
1
2
u +
1
2
(
1
2
2
1
)θ.
We might think the second term vanishes identically, but that is not true. Our
θ
has some singularities, and so that expression is not going to vanish at the
singularities. The precise statement is that (
1
2
2
1
)
θ
is a distribution
supported at the points Z
j
.
To figure out what it is, we have to integrate:
Z
|zZ
j
|≤ε
(
1
2
2
1
)θ d
2
x =
Z
|zZ
j
|=ε
1
θ dx
1
+
2
θ dx
2
=
I
|zZ
j
|=ε
dθ = 4πn
j
,
where n
j
is the multiplicity of the zero. Thus, we deduce that
(
1
2
2
1
)θ = 2π
X
δ
Z
j
.
But then we are done!
We can think of this
u
as a non-linear combination of fundamental solutions
to the Laplacian. Near the
δ
functions, the
e
u
1 term doesn’t contribute much,
and the solution looks like the familiar fundamental solutions to the Laplacian
with logarithmic singularities. However, far away from the singularities,
e
u
1
forces u to tend to 0, instead of growing to infinity.
Taubes proved that this equation has a unique solution, which is smooth on
R
2
\{Z
j
}
, with logarithmic singularities at
Z
j
, and such that
u
0 as
|z|
.
Also, u < 0.
It is an exercise to check that the Bogomolny equations imply the second-order
Euler–Lagrange equations.
For example, differentiating the second Bogomolny equation and using the
first gives
1
B = , D
1
Φ) = (Φ, iD
2
Φ).
We can similarly do this for the sine-Gordon theory.
2.5 Physics of vortices
Recall we began with the ungauged Ginzburg–Landau theory, and realized the
solitons didn’t have finite energy. We then added a gauge field, and the problem
went away we argued the coupling to the gauge field “gave mass” to the trans-
verse component, thus allowing the existence of finite energy soliton solutions. In
the book of Jaffe and Taubes there are results on the exponential decay of gauge
invariant combinations of the fields which are another expression of this effect
the Higgs mechanism. However, there is a useful and complementary way of
understanding how gauge fields assist in stabilizing finite energy configurations
against collapse, and this doesn’t require any detailed information about the
theory at all only scaling. We now consider this technique, which is known
either as the Derrick or the Pohozaev argument.
Suppose we work in
d
space dimensions. Then a general scalar field Φ :
R
d
R
`
has energy functional given by
Z
R
d
1
2
|∇Φ|
2
+ U(Φ)
d
d
x.
for some
U
. In the following we consider smooth finite energy configurations
for which the energy is stationary. To be precise, we require that the energy is
stationary with respect to variations induced by rescaling of space (as is made
explicit in the proof); we just refer to these configurations as stationary points.
Theorem
(Derrick’s scaling argument)
.
Consider a field theory in
d
-dimensions
with energy functional
E[Φ] =
Z
R
d
1
2
|∇Φ|
2
+ U(Φ)
d
d
x = T + W,
with
T
the integral of the gradient term and
W
the integral of the term involving
U.
If d = 1, then any stationary point must satisfy
T = W.
If d = 2, then all stationary points satisfy W = 0.
If d 3, then all stationary points have T = W = 0, i.e. Φ is constant.
This forbids the existence of solitons in high dimensions for this type of
energy functional.
Proof.
Suppose Φ were such a stationary point. Then for any variation Φ
λ
of Φ
such that Φ = Φ
1
, we have
d
dλ
λ=1
E
λ
] = 0.
Consider the particular variation given by
Φ
λ
(x) = Φ(λx).
Then we have
W
λ
] =
Z
R
d
U
λ
(x)) d
d
x = λ
d
Z
R
d
U(Φ(λx)) d
d
(λx) = λ
d
W [Φ].
On the other hand, since
T
contains two derivatives, scaling the derivatives as
well gives us
T
λ
] = λ
2d
T [Φ].
Thus, we find
E
λ
] = λ
2d
T [Φ] + λ
d
W [Φ].
Differentiating and setting λ = 1, we see that we need
(2 d)T [Φ] dW [Φ] = 0.
Then the results in different dimensions follow.
The
d
= 2 case is rather interesting. We can still get interesting soliton
theories if we have sufficiently large space of classical vacua {Φ : W (Φ) = 0}.
Example. In d = 2, we can take = 3 and
W (φ) = (1 |φ|
2
)
2
.
Then the set
W
= 0 is given by the unit sphere
S
2
R
3
. With
φ
constrained
to this 2-sphere, this is a
σ
-model , and there is a large class of such maps
φ
(
x
) which minimize the energy (for a fixed topology) in fact they are just
rational functions when stereographic projection is used to introduce complex
coordinates.
Derrick’s scaling argument is not only a mathematical trick. We can also
interpret it physically. Increasing
λ
corresponds to “collapsing” down the field.
Then we see that in
d
3, both the gradient and potential terms favour collapsing
of the field. However, in
d
= 1, the gradient term wants the field to expand, and
the potential term wants the field to collapse. If these two energies agree, then
these forces perfectly balance, and one can hope that stationary solitons exist.
We will eventually want to work with theories in higher dimensions, and
Derrick’s scaling argument shows that for scalar theories with energy functionals
as above this isn’t going to be successful for three or more dimensions, and
places strong restrictions in two dimensions. There are different ways to get
around Derrick’s argument. In the Skyrme model, which we are going to study
in the next chapter, there are no gauge fields, but instead we will have some
higher powers of derivative terms. In particular, by introducing fourth powers of
derivatives in the energy density, we will have a term that scales as
λ
4d
, and
this acts to stabilize scalar field solitons in three dimensions.
Now let’s see how gauge theory provides a way around Derrick’s argu-
ment without having to depart from employing only energy densities which
are quadratic in the derivatives (as is highly desirable for quantization). To
understand this, we need to know how gauge fields transform under spatial
rescaling.
One way to figure this out is to insist that the covariant derivative D
j
Φ
λ
must scale as a whole in the same way that ordinary derivatives scale in scalar
field theory. Since
j
Φ
λ
= λ(
j
Φ)(λx),
we would want A
j
to scale as
(A
j
)
λ
(x) = λA
j
(λx).
We can also see this more geometrically. Consider the function
χ
λ
: R
d
R
d
x 7→ λx.
Then our previous transformations were just given by pulling back along
χ
λ
.
Since A is a 1-form, it pulls back as
χ
λ
(A
j
dx
j
) = λA
j
(λx) dx
j
.
Then since B = dA, the curvature must scale as
B
λ
(x) = λ
2
B(λx).
Thus, we can obtain a gauged version of Derrick’s scaling argument.
Since we don’t want to develop gauge theory in higher dimensions, we will
restrict our attention to the Ginzburg–Landau model. Since we already used the
letter λ, we will denote the scaling parameter by µ. We have
V
λ
(A
µ
, Φ
µ
) =
1
2
Z
µ
4
B(µx)
2
+ µ
2
|DΦ(µx)|
2
+
λ
4
(1 |Φ(µx)|
2
)
2
1
µ
2
d
2
(µx)
=
1
2
Z
µ
2
B
2
(y) + |DΦ(y)|
2
+
λ
4µ
2
(1 |Φ(y)|
2
)
2
d
2
y.
Again, the gradient term is scale invariant, but the magnetic field term counter-
acts the potential term. We can find the derivative to be
d
dµ
µ=1
V
λ
(A
µ
, Φ
µ
) =
Z
B
2
λ
4
(1 |Φ|
2
)
2
d
2
y.
Thus, for a soliton, we must have
Z
B
2
d
2
x =
λ
4
Z
R
2
(1 |Φ|
2
)
2
d
2
x.
Such solutions exist, and we see that this is because they are stabilized by the
magnetic field energy in the sense that a collapse of the configuration induced
by rescaling would be resisted by the increase of magnetic energy which such a
collapse would produce. Note that in the case
λ
= 1, this equation is just the
integral form of one of the Bogomolny equations!
Scattering of vortices
Derrick’s scaling argument suggests that vortices can exist if
λ >
0. However, as
we previously discussed, for
λ 6
= 1, there are forces between vortices in general,
and we don’t get static, separated vortices. By doing numerical simulations,
we find that when
λ <
1, the vortices attract. When
λ >
1, the vortices repel.
Thus, when
λ >
1, the symmetric vortices with
N >
1 are unstable, as they
want to break up into multiple single vortices.
We can talk a bit more about the
λ
= 1 case, where we have static multi-
vortices. For example, for
N
= 2, the solutions are parametrized by pairs of
points in C, up to equivalence
(Z
1
, Z
2
) (Z
2
, Z
1
).
We said the moduli space is
C
2
, and this is indeed true. However,
Z
1
and
Z
2
are
not good coordinates for this space. Instead, good coordinates on the moduli
space
M
=
M
2
are given by some functions symmetric in
Z
1
and
Z
2
. One valid
choice is
Q = Z
1
+ Z
2
, P = Z
1
Z
2
.
In general, for vortex number
N
, we should use the elementary symmetric
polynomials in Z
1
, ··· , Z
N
as our coordinates.
Now suppose we set our vortices in motion. For convenience, we fix the
center of mass so that Q(t) = 0. We can then write P as
P = Z
2
1
.
If we do some simulations, we find that in a head-on collision, after they collide,
the vortices scatter off at 90
. This is the 90
scattering phenomenon, and holds
for other λ as well.
In terms of our coordinates,
Z
2
1
is smoothly evolving from a negative to a
positive value, going through 0. This corresponds to
Z
1
7→ ±iZ
1
,
Z
2
=
Z
1
.
Note that at the point when they collide, we lose track of which vortex is which.
Similar to the
φ
4
kinks, we can obtain effective Lagrangians for the dynamics
of these vortices. However, this is much more complicated.
2.6 Jackiw–Pi vortices
So far, we have been thinking about electromagnetism, using the abelian Higgs
model. There is a different system that is useful in condensed matter physics. We
look at vortices in Chern–Simons–Higgs theory. This has a different Lagrangian
which is not Lorentz invariant instead of having the Maxwell curvature term,
we have the Chern–Simons Lagrangian term. We again work in two space
dimensions, with the Lagrangian density given by
L =
κ
4
ε
µνλ
A
µ
F
νλ
(iΦ, D
0
Φ)
1
2
||
2
+
1
2κ
|Φ|
4
,
where κ is a constant,
F
νλ
=
ν
A
λ
λ
A
ν
is the electromagnetic field and, as before,
D
0
Φ =
Φ
t
iA
0
Φ.
The first term is the Chern–Simons t erm, while the rest is the Schr¨odinger
Lagrangian density with a |Φ|
4
potential term.
Varying with respect to Φ, the Euler–Lagrange equation gives the Schr¨odinger
equation
iD
0
Φ +
1
2
D
2
j
Φ +
1
κ
|Φ|
2
Φ = 0.
If we take the variation with respect to A
0
instead, then we have
κB + |Φ|
2
= 0.
This is unusual it looks more like a constraint than an evolution equation,
and is a characteristic feature of Chern–Simons theories.
The other equations give conditions like
1
A
0
=
0
A
1
+
1
κ
(iΦ, D
2
Φ)
2
A
0
=
0
A
2
1
κ
(iΦ, D
1
Φ).
This is peculiar compared to Maxwell theory the equations relate the current
directly to the electromagnetic field, rather than its derivative.
For static solutions, we need
i
A
0
=
1
κ
ε
ij
(iΦ, D
j
Φ).
To solve this, we assume the field again satisfies the covariant anti-holomorphicity
condition
D
j
Φ =
jk
D
k
Φ = 0.
Then we can write
1
A
0
= +
1
κ
(iΦ, D
2
Φ) =
1
κ
1
|Φ|
2
2
,
and similarly for the derivative in the second coordinate direction. We can then
integrate these to obtain
A
0
=
|Φ|
2
2κ
.
We can then look at the other two equations, and see how we can solve those.
For static fields, the Schr¨odinger equation becomes
A
0
Φ +
1
2
D
2
j
Φ +
1
κ
|Φ|
2
Φ = 0.
Substituting in A
0
, we obtain
D
2
j
Φ =
|Φ|
2
κ
Let’s then see if this makes sense. We need to see whether this can be consistent
with the holomorphicity condition. The answer is yes, if we have the equation
κB + |Φ|
2
= 0. We calculate
D
2
j
Φ = D
2
1
Φ + D
2
2
Φ = i(D
1
D
2
D
2
D
1
)Φ = +BΦ =
1
κ
|Φ|
2
Φ,
exactly what we wanted.
So the conclusion (check as an exercise) is that we can generate vortex
solutions by solving
D
j
Φ
jk
D
k
Φ = 0
κB + |Φ|
2
= 0.
As in Taubes’ theorem there is a reduction to a scalar equation, which is in this
case solvable explicitly:
log |Φ| =
1
κ
|Φ|
2
.
Setting ρ = |Φ|
2
, this becomes Liouville’s equation
log ρ =
2
κ
ρ,
which can in fact be solved in terms of rational functions see for example
Chapter 5 of the book Solitons in Field Theory and Nonlinear Analysis by Y.
Yang. (As in Taubes’ theorem, there is a corresponding statement providing
solutions via holomorphic rather than anti-holomorphic conditions.)
3 Skyrmions
We now move on to one dimension higher, and study Skyrmions. In recent years,
there has been a lot of interest in what people call “Skyrmions”, but what they
are studying is a 2-dimensional variant of the original idea of Skyrmions. These
occur in certain exotic magnetic systems. But instead, we are going to study the
original Skyrmions as discovered by Skyrme, which have applications to nuclear
physics.
With details to be filled in soon, hadronic physics exhibits (approximate)
spontaneously broken chiral symmetry
SU(2)
L
×SU(2)
R
Z
2
=
SO
(4), where the
unbroken group is (diagonal)
SO
(3) isospin, and the elements of
SO
(3) are
(g, g) SU(2) × SU(2).
This symmetry is captured in various effective field theories of pions (which
are the approximate Goldstone bosons) and heavier mesons. It is also a feature
of QCD with very light u and d quarks.
The special feature of Skyrmion theory is that we describe nucleons as solitons
in the effective field theory. Skyrme’s original idea was that nucleons and bigger
nuclei can be modelled by classical approximations to some “condensates” of
pion fields. To explain the conservation of baryon number, the classical field
equations have soliton solutions (Skyrmions) with an integer topological charge.
This topological charge is then identified with what is known, physically, as the
baryon number.
3.1 Skyrme field and its topology
Before we begin talking about the Skyrme field, we first discuss the symmetry
group this theory enjoys. Before symmetry breaking, our theory has a symmetry
group
SU(2) × SU(2)
(1, 1)}
=
SU(2) × SU(2)
Z
2
.
This might look like a rather odd symmetry group to work with. We can
begin by understanding the
SU
(2)
× SU
(2) part of the symmetry group. This
group acts naturally on SU(2) again, by
(A, B) · U = AUB
1
.
However, we notice that the pair (
1, 1
)
SU
(2)
×SU
(2) acts trivially. So the
true symmetry group is the quotient by the subgroup generated by this element.
One can check that after this quotienting, the action is faithful.
In the Skyrme model, the field will be valued in
SU
(2). It is convenient to
introduce coordinates for our Skyrme field. As usual, we let
τ
be the Pauli
matrices, and 1 be the unit matrix. Then we can write the Skyrme field as
U(x, t) = σ(x, t)1 + iπ(x, t) · τ .
However, the values of
σ
and
π
cannot be arbitrary. For
U
to actually lie in
SU(2), we need the coefficients to satisfy
σ, π
i
R, σ
2
+ π · π = 1.
This is a non-linear constraint, and defines what known as a non-linear
σ
-model .
From this constraint, we see that geometrically, we can identify
SU
(2) with
S
3
. We can also see this directly, by writing
SU(2) =

α β
¯
β ¯α
M
2
(C) : |α|
2
+ |β|
2
= 1
,
and this gives a natural embedding of
SU
(2) into
C
2
=
R
4
as the unit sphere,
by sending the matrix to (α, β).
One can check that the action we wrote down acts by isometries of the
induced metric on S
3
. Thus, we obtain an inclusion
SU(2) × SU(2)
Z
2
SO(4),
which happens to be a surjection.
Our theory will undergo spontaneous symmetry breaking, and the canonical
choice of vacuum will be
U
=
1
. Equivalently, this is when
σ
= 1 and
π
=
0
.
We see that the stabilizer of 1 is given by the diagonal
∆ : SU(2)
SU(2) × SU(2)
Z
2
,
since A1B
1
= 1 if and only if A = B.
Geometrically, if we view
SU(2)×SU(2)
Z
2
=
SO
(4), then it is obvious that the
stabilizer of
1
is the copy of
SO
(3)
SO
(4) that fixes the
1
axis. Indeed, the
image of the diagonal is SU(2)/1}
=
SO(3).
Note that in our theory, for any choice of
π
, there are at most two possible
choices of
σ
. Thus, despite there being four variables, there are only three degrees
of freedom. Geometrically, this is saying that
SU
(2)
=
S
3
is a three-dimensional
manifold.
This has some physical significance. We are using the
π
fields to model pions.
We have seen and observed pions a lot. We know they exist. However, as far as
we can tell, there is no
σ
meson”, and this can be explained by the fact that
σ
isn’t a genuine degree of freedom in our theory.
Let’s now try to build a Lagrangian for our field
U
. We will want to introduce
derivative terms. From a mathematical point of view, the quantity
µ
U
isn’t
a very nice thing to work with. It is a quantity that lives in the tangent space
T
U
SU(2), and in particular, the space it lives in depends on the value of U.
What we want to do is to pull this back to
T
1
SU
(2) =
su
(2). To do so, we
multiply by U
1
. We write
R
µ
= (
µ
U)U
1
,
which is known as the right current. For practical, computational purposes, it is
convenient to note that
U
1
= σ1 iπ · τ .
Using the (+
, , ,
) metric signature, we can now write the Skyrme Lagrangian
density as
L =
F
2
π
16
Tr (R
µ
R
µ
) +
1
32e
2
Tr ([R
µ
, R
ν
][R
µ
, R
ν
])
1
8
F
2
π
m
2
π
Tr(1 U).
The three terms are referred to as the
σ
-model term, Skyrme term and pion
mass term respectively.
The first term is the obvious kinetic term. The second term might seem a
bit mysterious, but we must have it (or some variant of it). By Derrick’s scaling
argument, we cannot have solitons if we just have the first term. We need a
higher multiple of the derivative term to make solitons feasible.
There are really only two possible terms with four derivatives. The alternative
is to have the square of the first term. However, Skyrme rejected that object,
because that Lagrangian will have four time derivatives. From a classical point
of view, this is nasty, because to specify the initial configuration, not only do we
need the initial field condition, but also its first three derivatives. This doesn’t
happen in our theory, because the commutator vanishes when
µ
=
ν
. The pieces
of the Skyrme term are thus at most quadratic in time derivatives.
Now note that the first two terms have an exact chiral symmetry, i.e. they
are invariant under the
SO
(4) action previously described. In the absence of
the final term, this symmetry would be spontaneously broken by a choice of
vacuum. As described before, there is a conventional choice
U
=
1
. After this
spontaneous symmetry breaking, we are left with an isospin
SU
(2) symmetry.
This isospin symmetry rotates the π fields among themselves.
The role of the extra term is that now the vacuum has to be the identity
matrix. The symmetry is now explicitly broken. This might not be immediately
obvious, but this is because the pion mass term is linear in
σ
and is minimized
when
σ
= 1. Note that this theory is still invariant under the isospin
SU
(2)
symmetry. Since the isospin symmetry is not broken, all pions have the same
mass. In reality, the pion masses are very close, but not exactly equal, and we
can attribute the difference in mass as due to electromagnetic effects. In terms
of the π fields we defined, the physical pions are given by
π
±
= π
1
±
2
, π
0
= π
3
.
It is convenient to draw the target space SU(2) as
σ = 1
σ = 1
Potential
energy
σ = 0
|π| = 1
S
3
=
SU(2)
σ
=
1 is the anti-vacuum. We will see that in the core of the Skyrmion,
σ
will
take value σ = 1.
In the Skyrme Lagrangian, we have three free parameters. This is rather few
for an effective field theory, but we can reduce the number further by picking
appropriate coefficients. We introduce an energy unit
F
π
4e
and length unit
2
eF
π
.
Setting these to 1, there is one parameter left, which is the dimensionless pion
mass. In these units, we have
L =
Z
1
2
Tr(R
µ
R
µ
) +
1
16
Tr([R
µ
, R
ν
][R
µ
, R
ν
]) m
2
Tr(1 U)
d
3
x.
In this notation, we have
m =
2m
π
eF
π
.
In general, we will think of m as being “small”.
Let’s see what happens if we in fact put
m
= 0. In this case, the lack of mass
term means we no longer have the boundary condition that
U
1 at infinity.
Hence, we need to manually impose this condition.
Deriving the Euler–Lagrange equations is slightly messy, since we have to
vary
U
while staying inside the group
SU
(2). Thus, we vary
U
multiplicatively,
U 7→ U (1 + εV )
for some
V su
(2). We then have to figure out how
R
varies, do some
differentiation, and then the Euler–Lagrange equations turn out to be
µ
R
µ
+
1
4
[R
ν
, [R
ν
, R
µ
]]
= 0.
For static fields, the energy is given by
E =
Z
1
2
Tr(R
i
R
i
)
1
16
Tr([R
i
, R
j
][R
i
, R
j
])
d
3
x E
2
+ E
4
.
where we sum
i
and
j
from 1 to 3. This is a sum of two terms the first is
quadratic in derivatives while the second is quartic.
Note that the trace functional on
su
(2) is negative definite. So the energy is
actually positive.
We can again run Derrick’s theorem.
Theorem
(Derrick’s theorem)
.
We have
E
2
=
E
4
for any finite-energy static
solution for m = 0 Skyrmions.
Proof.
Suppose
U
(
x
) minimizes
E
=
E
2
+
E
4
. We rescale this solution, and
define
˜
U(x) = U(λx).
Since U is a solution, the energy is stationary with respect to λ at λ = 1.
We can take the derivative of this and obtain
i
˜
U(x) = λ
˜
U(λx).
Therefore we find
˜
R
i
(x) = λR
i
(λx),
and therefore
˜
E
2
=
1
2
Z
Tr(
˜
R
i
˜
R
i
) d
3
x
=
1
2
λ
2
Z
Tr(R
i
R
i
)(λx) d
3
x
=
1
2
1
λ
Z
Tr(R
i
R
i
)(λx) d
3
(λx)
=
1
λ
E
2
.
Similarly,
˜
E
4
= λE
4
.
So we find that
˜
E =
1
λ
E
2
+ λE
4
.
But this function has to have a minimum at
λ
= 1. So the derivative with
respect to λ must vanish at 1, requiring
0 =
d
˜
E
dλ
=
1
λ
2
E
2
+ E
4
= 0
at λ = 1. Thus we must have E
4
= E
2
.
We see that we must have a four-derivative term in order to stabilize the
soliton. If we have the mass term as well, then the argument is slightly more
complicated, and we get some more complicated relation between the energies.
Baryon number
Recall that our field is a function
U
:
R
3
SU
(2). Since we have a boundary
condition
U
=
1
at infinity, we can imagine compactifying
R
3
into
S
3
, where
the point at infinity is sent to 1.
On the other hand, we know that
SU
(2) is isomorphic to
S
3
. Geometrically, we
think of the space and
SU
(2) as “different”
S
3
. We should think of
SU
(2) as the
“unit sphere”, and write it as
S
3
1
. However, we can think of the compactification
of
R
3
as a sphere with “infinite radius”, so we denote it as
S
3
. Of course,
topologically, they are the same.
So the field is a map
U : S
3
S
3
1
.
This map has a degree. There are many ways we can think about the degree.
For example, we can think of this as the homotopy class of this map, which is
an element of π
3
(S
3
)
=
Z. Equivalently, we can think of it as the map induced
on the homology or cohomology of S
3
.
While
U
evolves with time, because the degree is a discrete quantity, it has
to be independent of time (alternatively, the degree of the map is homotopy
invariant).
There is an explicit integral formula for the degree. We will not derive this,
but it is
B =
1
24π
2
Z
ε
ijk
Tr(R
i
R
j
R
k
) d
3
x.
The factor of 2
π
2
comes from the volume of the three sphere, and there is also
a factor of 6 coming from how we anti-symmetrize. We identify
B
with the
conserved, physical baryon number.
If we were to derive this, then we would have to pull back a normalized
volume form from
S
3
1
and then integrate over all space. In this formula, we chose
to use the
SO
(4)-invariant volume form, but in general, we can pull back any
normalized volume form.
Locally, near σ = 1, this volume form is actually
1
2π
2
dπ
1
dπ
2
dπ
3
.
Faddeev–Bogomolny bound
There is a nice result analogous to the Bogomolny energy bound we saw for
kinks and vortices, known as the Faddeev–Bogomolny bound . We can write the
static energy as
E =
Z
1
2
Tr
R
i
1
4
ε
ijk
[R
j
, R
k
]
2
!
d
3
x ± 12π
2
B.
This bound is true for both sign choices. However, to get the strongest result,
we should choose the sign such that ±12π
2
B > 0. Then we find
E 12π
2
|B|.
By symmetry, it suffices to consider the case B > 0, which is what we will do.
The Bogomolny equation for B > 0 should be
R
i
1
4
ε
ijk
[R
j
, R
k
] = 0.
However, it turns out this equation has no non-vacuum solution.
Roughly, the argument goes as follows by careful inspection, for this to
vanish, whenever
R
i
is non-zero, the three vectors
R
1
, R
2
, R
3
must form an
orthonormal frame in
su
(2). So
U
must be an isometry. But this isn’t possible,
because the spheres have “different radii”.
Therefore, true Skyrmions with B > 0 satisfy
E > 12π
2
B.
We get a lower bound, but the actual energy is always greater than this lower
bound. It is quite interesting to look at the energies of true solutions numerically,
and their energy is indeed bigger.
3.2 Skyrmion solutions
The simplest Skyrmion solution has baryon number
B
= 1. We will continue to
set m = 0.
B = 1 hedgehog Skyrmion
Consider the spherically symmetric function
U(x) = cos f(r)1 + i sin f (r)
ˆ
x · τ .
This is manifestly in
SU
(2), because
cos
2
f
+
sin
2
f
= 1. This is known as a
hedgehog, because the unit pion field is
ˆ
x
, which points radially outwards. We
need some boundary conditions. We need
U 1
at
. On the other hand, we
will see that we need U 1 at the origin to get baryon number 1. So f π
as r 0, and f 0 as r . So it looks roughly like this:
r
f
π
After some hard work, we find that the energy is given by
E = 4π
Z
0
f
02
+
2 sin
2
f
r
2
(1 + f
02
) +
sin
4
f
r
4
r
2
dr.
From this, we can obtain a second-order ordinary differential equation in
f
,
which is not simple. Solutions have to be found numerically. This is a sad truth
about Skyrmions. Even in the simplest
B
= 1
, m
= 0 case, we don’t have an
analytic expression for what f looks like. Numerically, the energy is given by
E = 1.232 × 12π
2
.
To compute the baryon number of this solution, we plug our solution into the
formula, and obtain
B =
1
2π
2
Z
0
sin
2
f
r
2
df
dr
· 4πr
2
dr.
We can interpret
df
dr
as the radial contribution to
B
, while there are two factors
of
sin f
r
coming from the angular contribution due to the i sin f(r)
ˆ
x · τ term.
But this integral is just an exact differential. It simplifies to
B =
1
π
Z
π
0
2 sin
2
f df.
Note that we have lost a sign, because of the change of limits. We can integrate
this directly, and just get
B =
1
π
f
1
2
sin 2f
π
0
= 1,
as promised.
Intuitively, we see that in this integral,
f
goes from 0 to
π
, and we can think
of this as the field U wrapping around the sphere S
3
once.
More hedgehogs
We can consider a more general hedgehog with the same ansatz, but with the
boundary conditions
f(0) = , f() = 0.
In this case, the same computations gives us
B
=
n
. So in principle, this gives a
Skyrmion of any baryon number. The solutions do exist. However, it turns out
they have extremely high energy, and are nowhere near minimizing the energy.
In fact, the energy increases much faster than
n
itself, because the Skyrmion
“onion” structure highly distorts each
B
= 1 Skyrmion. Unsurprisingly, these
solutions are unstable.
This is not what we want in hadronic physics, where we expect the energy to
scale approximately linearly with
n
. In fact, since baryons attract, we expect
the solution for B = n to have less energy than n times the B = 1 energy.
We can easily get energies approximately
n
times the
B
= 1 energy simply
by having very separated baryons, and then since they attract, when they move
towards each other, we get even lower energies.
A better strategy rational map approximation
So far, we have been looking at solutions that depend very simply on angle. This
means, all the “winding” happens in the radial direction. In fact, it is a better
idea to wind more in the angular direction.
In the case of the B = 1 hedgehog, the field looks roughly like this:
σ = 1
σ = 1
In our
B >
1 spherically symmetric hedgehogs, we wrapped radially around
the sphere many times, and it turns out that was not a good idea.
Better is to use a similar radial configuration as the
B
= 1 hedgehog, but
introduce more angular twists. We can think of the above
B
= 1 solution as
follows we slice up our domain
R
3
(or rather,
S
3
since we include the point
at infinity) into 2-spheres of constant radius, say
S
3
=
[
r[0,]
S
2
r
.
On the other hand, we can also slice up the
S
3
in the codomain into constant
σ
levels, which are also 2-spheres:
S
3
=
[
σ
S
2
σ
.
Then the function
f
(
r
) we had tells us we should map the 2-sphere
S
2
r
into the
two sphere
S
2
cos f(r)
. What we did, essentially, was that we chose to map
S
2
r
to
S
2
cos f(r)
via the “identity map”. This gave a spherically symmetric hedgehog
solution.
But we don’t have to do this! Pick any function
R
:
S
2
S
2
. Then we
can construct the map Σ
cos f
R
that sends
S
2
r
to
S
2
cos f(r)
via the map
R
. For
simplicity, we will use the same
R
for all
r
. If we did this, then we obtain a
non-trivial map Σ
cos f
R : S
3
S
3
.
Since
R
itself is a map from a sphere to a sphere (but one dimension lower),
R
also has a degree. It turns out this degree is the same as the degree of the
induced map Σ
cos f
R
! So to produce higher baryon number hedgehogs, we simply
have to find maps R : S
2
S
2
of higher degree.
Fortunately, this is easier than maps between 3-spheres, because a 2-sphere is
a Riemann surface. We can then use complex coordinates to work on 2-spheres.
By complex analysis, any complex holomorphic map between 2-spheres is given
by a rational map.
Pick any rational function
R
k
(
z
) of degree
k
. This is a map
S
2
S
2
. We
use coordinates
r, z
, where
r R
+
and
z C
=
C {∞}
=
S
2
. Then we can
look at generalized hedgehogs
U(x) = cos f(r)1 + i sin f (r)
ˆ
n
R
k
(z)
· τ ,
with
f
(0) =
π
,
f
(
) = 0, and
ˆ
n
R
k
is the normalized pion field
ˆ
π
, given by the
unit vector obtained from
R
k
(
z
) if we view
S
2
as a subset of
R
3
in the usual
way. Explicitly,
ˆ
n
R
=
1
1 + |R|
2
(
¯
R + R, i(
¯
R R), 1 |R|
2
).
This construction is in some sense a separation of variables, where we separate
the radial and angular dependence of the field.
Note that even if we find a minimum among this class of fields, it is not a
true minimum energy Skyrmion. However, it gets quite close, and is much better
than our previous attempt.
There is quite a lot of freedom in this construction, since we are free to pick
f
(
r
), as well as the rational function
R
k
(
z
). The geometric degree
k
of
R
k
(
z
) we
care about is the same as the algebraic degree of R
k
(z). Precisely, if we write
R
k
(z) =
p
k
(z)
q
k
(z)
,
where
p
k
and
q
k
are coprime, then the algebraic degree of
R
k
is the maximum of
the degrees of
p
k
and
q
k
as polynomials. Since there are finitely many coefficients
for
p
k
and
q
k
, this is a finite-dimensional problem, which is much easier than
solving for arbitrary functions. We will talk more about the degree later.
Numerically, we find that minimal energy fields are obtained with
R
1
(z) = z R
2
(z) = z
2
R
3
(z) =
3iz
2
1
z
3
3iz
R
4
(z) =
z
4
+ 2
3iz
2
+ 1
z
4
2
3iz
2
+ 1
.
The true minimal energy Skyrmions have also been found numerically, and are
very similar to the optimal rational map fields. In fact, the search for the true
minima often starts from a rational map field.
Constant energy density surfaces of Skyrmions up to baryon number 8
(for m = 0), by R. A. Battye and P. M. Sutcliffe
We observe that
for n = 1, we recover the hedgehog solution.
for n = 2, our solution has an axial symmetry.
for
n
= 3, the solution might seem rather strange, but it is in fact the
unique solution in degree 3 with tetrahedral symmetry.
for n = 4, the solution has cubic symmetry.
In each case, these are the unique rational maps with such high symmetry. It
turns out, even though these are not the exact Skyrmion solutions, the exact
solutions enjoy the same symmetries.
The function f can be found numerically, and depends on B.
Geometrically, what we are doing is that we are viewing
S
3
as the suspension
Σ
S
2
, and our construction of Σ
cos f
R
from
R
is just the suspension of maps.
The fact that degree is preserved by suspension can be viewed as an example of
the fact that homology is stable.
More on rational maps
Why does the algebraic degree of
R
k
agree with the geometric degree? One way
of characterizing the geometric degree is by counting pre-images. Consider a
generic point c in the target 2-sphere, and consider the equation
R
k
(z) =
p
k
(z)
q
k
(z)
= c
for a generic c. We can rearrange this to say
p
k
cq
k
= 0.
For a generic
c
, the
z
k
terms do not cancel, so this is a polynomial equation of
degree
k
. Also, generically, this equation doesn’t have repeated roots, so has
exactly
k
solutions. So the number of points in the pre-image of a generic
c
is
k
.
Because
R
k
is a holomorphic map, it is automatically orientation preserving
(and in fact conformal). So each of these
k
points contributes +1 to the degree.
In the pictures above, we saw that the Skyrmions have some “hollow polyhe-
dral” structures. How can we understand these?
The holes turn out to be zeroes of the baryon density, and are where energy
density is small. At the center of the holes, the angular derivatives of the Skyrme
field U are zero, but the radial derivative is not.
We can find these holes precisely in the rational map approximation. This
allows us to find the symmetry of the system. They occur where the derivative
dR
k
dz
= 0. Since R =
p
q
, we can rewrite this requirement as
W (z) = p
0
(z)q(z) q
0
(z)p(z) = 0.
W (z) is known as the Wronskian.
A quick algebraic manipulation shows that
W
has degree at most 2
k
2,
and generically, it is indeed 2
k
2. This degree is the number of holes in the
Skyrmion.
We can look at our examples and look at the pictures to see this is indeed
the case.
Example. For
R
4
(z) =
z
4
+ 2
3iz
2
+ 1
z
4
2
3iz
2
+ 1
,
the Wronskian is
W (z) = (4z
3
+ 4
3iz)(z
4
2
3iz
2
+ 1) (z
4
+ 2
3iz
2
+ 1)(4z
3
4
3iz).
The highest degree
z
7
terms cancel. But there isn’t any
z
6
term anywhere either.
Thus W turns out to be a degree 5 polynomial. It is given by
W (z) = 8
3i(z
5
z).
We can easily list the roots they are z = 0, 1, i, 1, i.
Generically, we expect there to be 6 roots. It turns out the Wronksian has
a zero at
as well. To see this more rigorously, we can rotate the Riemann
sphere a bit by a obius map, and then see there are 6 finite roots. Looking
back at our previous picture, there are indeed 6 holes when B = 4.
It turns out although the rational map approximation is a good way to find
solutions for
B
up to
20, they are hollow with
U
=
1 at the center. This is
not a good model for larger nuclei, especially when we introduce non-zero pion
mass.
For m 1, there are better, less hollow Skyrmions when B 8.
3.3 Other Skyrmion structures
There are other ways of trying to get Skyrmion solutions.
Product Ansatz
Suppose
U
1
(
x
) and
U
2
(
x
) are Skyrmions of baryon numbers
B
1
and
B
2
. Since
the target space is a group SU(2), we can take the product
U(x) = U
1
(x)U
2
(x).
Then the baryon number is
B
=
B
1
+
B
2
. To see this, we can consider the
product when
U
1
and
U
2
are well-separated, i.e. consider
U
(
x a
)
U
2
(
x
) with
|a|
large. Then we can see the baryon number easily because the baryons are
well-separated. We can then vary
a
continuously to 0, and
B
doesn’t change
as we make this continuous deformation. So we are done. Alternatively, this
follows from an Eckmann–Hilton argument.
This can help us find Skyrmions with baryon number
B
starting with
B
well-separated
B
= 1 hedgehogs. Of course, this will not be energy-minimizing,
but we can numerically improve the field by letting the separation vary.
It turns out this is not a good way to find Skyrmions. In general, it doesn’t
give good approximations to the Skyrmion solutions. They tend to lack the
desired symmetry, and this boils down to the problem that the product is
not commutative, i.e.
U
1
U
2
6
=
U
2
U
1
. Thus, we cannot expect to be able to
approximate symmetric things with a product ansatz.
The product ansatz can also be used for several
B
= 4 subunits to construct
configurations with baryon number 4
n
for
n Z
. For example, the following is
a B = 31 Skyrmion:
B = 31 Skyrmion by P. H. C. Lau and N. S. Manton
This is obtained by putting eight
B
= 4 Skyrmions side by side, and then cutting
off a corner.
This strategy tends to work quite well. With this idea, we can in fact find
Skyrmion solutions with baryon number infinity! We can form an infinite cubic
crystal out of
B
= 4 subunits. For
m
= 0, the energy per baryon is
1
.
038
×
12
π
2
.
This is a very close to the lower bound!
We can also do other interesting things. In the picture below, on the left, we
have a usual
B
= 7 Skyrmion. On the right, we have deformed the Skyrmion
into what looks like a
B
= 4 Skyrmion and a
B
= 3 Skyrmion. This is a cluster
structure, and it turns out this deformation doesn’t cost a lot of energy. This
two-cluster system can be used as a model of the lithium-7 nucleus.
B = 7 Skyrmions by C. J. Halcrow
3.4 Asymptotic field and forces for B = 1 hedgehogs
We now consider what happens when we put different
B
= 1 hedgehogs next
to each other. To understand this, we look at the profile function
f
, for
m
= 0.
For large r, this has the asymptotic form
f(r)
C
r
2
.
To obtain this, we linearize the differential equation for
f
and see how it behaves
as
r
and
f
0. The linearized equation doesn’t determine the coefficient
C
, but the full equation and boundary condition at
r
= 0 does. This has to be
worked out numerically, and we find that C 2.16.
Thus, as
σ
1, we find
π C
x
r
3
. So the
B
= 1 hedgehog asymptotically
looks like three pion dipoles. Each pion field itself has an axis, but because we
have three of them, the whole solution is spherically symmetric.
We can roughly sketch the Skyrmion as
+
+
+
Note that unlike in electromagnetism, scalar dipoles attract if oppositely oriented.
This is because the fields have low gradient. So the lowest energy arrangement
of two B = 1 Skyrmions while they are separated is
+
+
+
+
+
+
The right-hand Skyrmion is rotated by 180
about a line perpendicular to the
line separating the Skyrmions.
These two Skyrmions attract! So two Skyrmions in this “attractive channel”
can merge to form the B = 2 torus, which is the true minimal energy solution.
+
+
+
+
+
The blue and the red fields have no net dipole, even before they merge. There
is only a quadrupole. However, the field has a strong net green dipole. The whole
field has toroidal symmetry, and these symmetries are important if we want to
think about quantum states and the possible spins these kinds of Skyrmions
could have.
For B = 4 fields, we can begin with the arrangement
+
+
+
+
+
+
+
+
+
+
+
+
To obtain the orientations, we begin with the bottom-left, and then obtain the
others by rotating along the axis perpendicular to the faces of the cube shown
here.
This configuration only has a tetrahedral structure. However, the Skyrmions
can merge to form a cubic B = 4 Skyrmion.
3.5 Fermionic quantization of the B = 1 hedgehog
We begin by quantizing the
B
= 1 Skyrmion. The naive way to do this is to
view the Skyrmion as a rigid body, and quantize it. However, if we do this, then
we will find that the result must have integer spin. However, this is not what
we want, since we want Skyrmions to model protons and nucleons, which have
half-integer spin. In other words, the naive quantization makes the Skyrmion
bosonic.
In general, if we want to take the Skyrme model as a low energy effective
field theory of QCD with an odd number of colours, then we must require the
B = 1 Skyrmion to be in a fermionic quantum state, with half-integer spin.
As a field theory, the configuration space for a baryon number
B
Skyrme field
is
Maps
B
(
R
3
SU
(2)), with appropriate boundary conditions. These are all
topologically the same as
Maps
0
(
R
3
SU
(2)), because if we fix a single element
U
0
Maps
B
(
R
3
SU
(2)), then multiplication by
U
0
gives us a homeomorphism
between the two spaces. Since we imposed the vacuum boundary condition, this
space is also the same as Maps
0
(S
3
S
3
).
This space is not simply connected. In fact, it has a first homotopy group of
π
1
(Maps
0
(S
3
S
3
)) = π
1
(Ω
3
S
3
) = π
4
(S
3
) = Z
2
.
Thus,
Maps
0
(
S
3
S
3
) has a universal double cover. In our theory, the wave-
functions on
Maps
(
S
3
S
3
) are not single-valued, but are well-defined functions
on the double cover. The wavefunction Ψ changes sign after going around a
non-contractible loop in the configuration space. It can be shown that this is
not just a choice, but required in a low-energy version of QCD.
This has some basic consequences:
(i) If we rotate a 1-Skyrmion by 2π, then Ψ changes sign.
(ii)
Ψ also changes sign when one exchanges two 1-Skyrmions (without rotating
them in the process). This was shown by Finkelstein and Rubinstein.
This links spin with statistics. If we quantized Skyrmions as bosons, then
both (i) and (ii) do not happen. Thus, in this theory, we obtain the
spin-statistics theorem from topology.
(iii)
In general, if
B
is odd, then rotation by 2
π
is a non-contractible loop,
while if
B
is even, then it is contractible. Thus, spin is half-integer if
B
is
odd, and integer if B is even.
(iv)
There is another feature of the Skyrme model. So far, our rotations are
spatial rotations. We can also rotate the value of the pion field, i.e. rotate
the target 3-sphere. This is isospin rotation. This behaves similarly to
above. Thus, isospin is half-integer if B is odd, and integer if B is even.
3.6 Rigid body quantization
We now make the rigid body approximation. We allow the Skyrmion to translate,
rotate and isorotate rigidly, but do not allow it to deform. This is a low-
energy approximation, and we have reduced the infinite dimensional space of
configurations to a finite-dimensional space. The group acting is
(translation) × (rotation) × (isorotation).
The translation part is trivial. The Skyrmion just gets the ability to move, and
thus gains momentum. So we are going to ignore it. The remaining group acting
is
SO
(3)
× SO
(3). This is a bit subtle. We have
π
1
(
SO
(3)) =
Z
2
, so we would
expect
π
1
(
SO
(3)
× SO
(3)) = (
Z
2
)
2
. However, in the full theory, we only have a
single
Z
2
. So we need to identify a loop in the first
SO
(3) with a loop in the
second SO(3).
Our wavefunction is thus a function on some cover of
SO
(3)
× SO
(3). While
SO
(3)
×SO
(3) is the symmetry group of the full theory, for a particular classical
Skyrmion solution, the orbit is smaller than the whole group. This is because
the Skyrmion often is invariant under some subgroup of
SO
(3)
× SO
(3). For
example, the
B
= 3 solution has a tetrahedral symmetry. Then we require our
wavefunction to be invariant under this group up to a sign.
For a single
SO
(3), we have a rigid-body wavefunction
|J L
3
J
3
i
, where
J
is the spin,
L
3
is the third component of spin relative to body axes, and
J
3
is
the third component of spin relative to space axes. Since we have two copies of
SO(3), our wavefunction can be represented by
|J L
3
J
3
i |I K
3
I
3
i.
I
3
is the isospin we are physically familiar with. The values of
J
3
and
I
3
are not
constrained, i.e. they take all the standard (2
J
+ 1)(2
I
+ 1) values. Thus, we are
going to suppress these labels.
However, the symmetry of the Skyrmion places constraints on the body
projections, and not all values of J and I are allowed.
Example.
For the
B
= 1 hedgehog, there is a lot of symmetry. Given an axis
ˆ
n
and an angle
α
, we know that classically, if we rotate and isorotate by the same
axis and angle, then the wavefunction is unchanged. So we must have
e
ˆ
n·L
e
ˆ
n·K
|Ψi = |Ψi.
It follows, by considering α small, and all
ˆ
n, that
(L + K) |Ψi = 0.
So the “grand spin”
L
+
K
must vanish. So
L·L
=
K·K
. Recall that
L·L
=
J·J
and K · K = I · I. So it follows that we must have J = I.
Since 1 is an odd number, for any axis
ˆ
n, we must also have
e
i2π
ˆ
n·L
|Ψi = |Ψi.
So I and J must be half-integer.
Thus, the allowed states are
J = I =
n
2
for some n 1 + 2Z. If we work out the formula for the energy in terms of the
spin, we see that it increases with
J
. It turns out the system is highly unstable
if n 5, and so
1
2
and
3
2
are the only physically allowed values.
The
J
=
I
=
1
2
states corresponds to
p
and
n
, with spin
1
2
. The
J
=
I
=
3
2
correspond to the
++
,
+
,
0
and
baryons, with spin
3
2
.
Example.
If
B
= 2, then we have a toroidal symmetry. This still has one
continuous SO(2) symmetry. Our first constraint becomes
e
iαL
3
e
i2αJ
3
|Ψi = |Ψi.
Note that we have 2
α
instead of
α
. If we look at our previous picture, we see
that when we rotate the space around by 2π, the pion field rotates by 4π.
There is another discrete symmetry, given by turning the torus upside down.
This gives
e
L
1
e
K
1
|Ψi = |Ψi.
The sign is not obvious from what we’ve said so far, but it is correct. Since we
have an even baryon number, the allowed states have integer spin and isospin.
States with isospin 0 are most interesting, and have lowest energy. Then the
K operators act trivially, and we have
e
iαL
3
|Ψi = |Ψi, e
L
1
|Ψi = |Ψi.
We have reduced the problem to one involving only body-fixed spin projection,
because the first equation tells us
L
3
|Ψi = 0.
Thus the allowed states are |J, L
3
= 0i.
The second constraint requires
J
to be odd. So the lowest energy states with
zero isospin are
|1, 0i
and
|3, 0i
. In particular, there are no spin 0 states. The
state
|1, 0i
represents the deuteron. This is a spin 1, isospin 0 bound state of
p
and n. This is a success.
The
|3, 0i
states have too high energy to be stable, but there is some evidence
for a spin 3 dibaryon resonance that decays into two ∆’s.
There is also a 2-nucleon resonance with
I
= 1 and
J
= 0, but this is not a
bound state. This is also allowed by the Skyrme model.
Example.
For
B
= 4, we have cubic symmetry. The symmetry group is
rather complicated, and imposes various constraints on our theory. But these
constraints always involve + signs on the right-hand side. The lowest allowed
state is |0, 0i |0, 0i, which agrees with the α-particle. The next I = 0 state is
|4, 4i +
r
14
5
|4, 0i + |4, 4i
!
|0, 0i
involving a combination of
L
3
values. This is an excited state with spin 4, and
would have rather high energy. Unfortunately, we haven’t seen this experimentally.
It is, however, a success of the Skyrme model that there are no
J
= 1
,
2
,
3 states
with isospin 0.