2Vortices

III Classical and Quantum Solitons



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.