Sie sind auf Seite 1von 13

Quantum Field Theory

Example Sheet 1
Michelmas Term 2011
Solutions by:

Johannes Hofmann jbh38@cam.ac.uk


Laurence Perreault Levasseur L.Perreault-Levasseur@damtp.cam.ac.uk
David Morris dmm49@cam.ac.uk
Marcel Schmittfull M.Schmittfull@damtp.cam.ac.uk

Note: In the conventions of this course, the Minkowski metric is g = diag (1, −1, −1, −1), or ‘mostly minus.’

Exercise 1
Lagrangian:
" 2 2 #
Z a  
σ ∂y T ∂y
L= dx − (1)
0 2 ∂t 2 ∂x

Express y(x, t) as a Fourier series:


r ∞
2X  nπx 
y(x, t) = qn (t) sin . (2)
a n=1 a

The derivative of y(x, t) with respect to x and t is


r ∞
∂y(x, t) 2X  nπx 
= q̇n (t) sin (3)
∂t a n=1 a
r ∞
∂y(x, t) 2 X nπ  nπx 
= qn (t) cos , (4)
∂x a n=1 a a

where we abbreviate q̇ ≡ ∂q/∂t. Substituting Eqs. (3) and (4) in (1) gives
1X a nmπ 2
Z   nπx   mπx   nπx   mπx 
L= dx σ q̇n (t)q̇m (t) sin sin −T q n (t)q m (t) cos cos . (5)
a n,m 0 a a a2 a a

Using the orthonormality relations


2 a
Z  nπx   mπx 
dx sin sin = δnm and (6)
a 0 a a
2 a
Z  nπx   mπx 
dx cos cos = δnm (7)
a 0 a a
we see that the terms with n 6= m vanish. We obtain:
XZ a 
σ 2 T  nπ 2 2

L= dx q̇ (t) − qn (t) . (8)
n 0 2 n 2 a

The Euler-Lagrange equations are given by


d ∂L ∂L
− = 0, n = 0, 1, 2, . . . . (9)
dt ∂ q̇n ∂qn
With
∂L
= σ q̇n and (10)
∂ q̇n
∂L  nπ 2
= −T qn (11)
∂qn a
we obtain the equations of motion:
T  nπ 2
q̈n (t) + qn (t) = 0, (12)
σ a
q
nπ T
where n = 0, 1, 2, . . .. These are the equations of motion of harmonic oscillators with frequency ωn = a σ.

1
Exercise 2
The scalar field φ(x) transforms under a Lorentz transformation xµ → x0µ = Λµ ν xν as

φ(x) → φ0 (x0 ) = φ(x) = φ(Λ−1 x0 ).

Note that this is an active transformation, i.e. the new field at the new coordinate equals the old field at the old coordinate.
Using
∂ ∂xν ∂
∂µ0 ≡ = = (Λ−1 )ν µ ∂ ν
∂x0µ ∂x0µ ∂xν
Now,

∂ µ ∂µ φ(x) + m2 φ(x) → ∂ 0µ ∂µ0 φ0 (x0 ) + m2 φ0 (x0 )


= g µν (Λ−1 )ρ µ ∂ρ (Λ−1 )κ ν ∂κ φ(x) + m2 φ(x)
= g ρκ ∂ρ ∂κ φ(x) + m2 φ(x). (13)

In the last step we used that Λ is a Lorentz transformation and so its inverse Λ−1 preserves the inverse Minkowski metric
g µν ,
(Λ−1 )ρ µ g µν (Λ−1 )κ ν = g ρκ .
Eq. (13) therefore demonstrates that if φ(x) fulfils the Klein-Gordon equation

∂ µ ∂µ φ(x) + m2 φ(x) = 0,

then φ(Λ−1 x) likewise fulfils it.

Exercise 3
Lagrangian density:
λ ∗ 2
L = ∂µ φ∗ ∂ µ φ − m2 φ∗ φ − (φ φ) . (14)
2
Euler-Lagrange equation for φ∗ :
∂L ∂L

− ∂µ = 0. (15)
∂φ ∂(∂µ φ∗ )
With
∂L
= −m2 φ − λ(φ∗ φ)φ and (16)
∂φ∗
∂L
= ∂ µ φ. (17)
∂(∂µ φ)
We obtain the equation of motion for φ∗ :

∂µ ∂ µ φ + m2 φ + λ(φ∗ φ)φ = 0. (18)

Similarly, we can calculate the field equation for φ. The result is the complex conjugate of Eq. (18).

Consider the U(1) transformation

φ(x) → eiα φ(x), and φ† (x) → eiα φ∗ (x), (19)

where α ∈ [0, 2π) is a constant. It is straightforward to check that the Lagrangian (14) is invariant under this transformation.
This is an example of a global symmetry, i.e., the symmetry transformation acts in the same way on the fields at each point
in space and time.

The infinitesimal transformation of (19) is:

φ → φ + δφ, where δφ = iαφ, and (20)


φ∗ → φ∗ + δφ∗ , δφ∗ = −iαφ. (21)

We can check explicitly that the Lagrangian density is invariant under this transformation:

δL = ∂µ (δφ∗ )∂ µ φ + ∂µ φ∗ ∂ µ (δφ) − m2 (δφ∗ φ + φ∗ δφ) − λ(φ∗ φ)(φ∗ δφ + δφ∗ φ)


= −iα∂µ φ∗ ∂ µ φ + iα∂µ φ∗ ∂ µ φ + iαm2 (φ∗ φ − φ∗ φ) − iαλ(φ∗ φ)(φ∗ φ − φ∗ φ) = 0. (22)

2
According to Noether’s theorem, a global symmetry implies the existence of a conserved current. If L → L + α∆L, where
∆L = ∂µ J µ , then the conserved Noether current is given by the formula (Schroeder & Peskin p. 18)
∂L
jµ = ∆ φa − J µ , (23)
∂(∂µ φa )
where a labels all the fields in the theory. In this case there are two independent fields φ and φ∗ , and
∆ φ = iφ , ∆ φ∗ = −iφ , ∆ L = 0.
The conserved current is thus
∂L ∂L
jµ = (iφ) + (−iφ∗ )
∂(∂µ φ) ∂(∂µ φ∗ )
= −i {φ∗ (∂ µ φ) − (∂ µ φ∗ )φ} . (24)
Notice that this quantity is real, and we can check explicitly that it is indeed conserved:
∂µ j µ = −i (φ∗ ∂µ (∂ µ φ) − (∂ µ φ∗ )(∂µ φ) + (∂µ φ∗ )(∂ µ φ) − ∂µ (∂ µ φ∗ )φ)
= −i (φ∗ ∂µ (∂ µ φ) − ∂µ (∂ µ φ∗ )φ)
= −i φ∗ −m2 φ − λ(φ∗ φ)φ − −m2 φ∗ − λ(φ∗ φ)φ∗ φ = 0.
    
(25)

In the last step, we used the equation of motion for φ and φ , Eq. (18).

Exercise 4
The Lagrangian will certainly be invariant under SO(3) rotations of the fields φa , a = 1, 2, 3, since only the ‘length’ φa φa
enters the Lagrangian. Infinitesimally, consider the transformation of φa φa under a rotation of the fields φa by an infinitesimal
angle θ
φa → φa + θabc nb φc , i.e. ∆ φa = abc nb φc (26)
where na is a constant unit vector. φa φa transforms as
φa φa → (φa + θabc nb φc )(φa + θade nd φe )
= φa φa + θabc nb φc φa + θade nd φe φa + O(θ2 )
= φa φa + O(θ2 ).
In the last step we used that abc is skew-symmetric under exchange of c and a whereas φc φa is symmetric under this
exchange, and therefore the sum vanishes. The same applies to the other term linear in θ. Dropping the θ2 term we find
that φa φa is indeed invariant under (26). Similarly one can also show that ∂µ φa ∂ µ φa is invariant under (26). Therefore L is
invariant under (26).

To obtain the associated Noether current, we refer to the formula (23) above. In our case J = 0 and ∆φa = abc nb φc , so
that
j µ = (∂ µ φa )abc nb φc .
The Noether current j µ implies a conserved charge
Z Z
Q ≡ d3 x j 0 = d3 x φ̇a abc nb φc .

This is conserved (Q̇ = 0) for any unit vector n. If we choose nb = δbd for d = 1, 2, 3 (i.e. n is any of the standard basis
vectors of R3 ), we obtain three linearly independent charges
Z Z
Qd = d3 x adc φ̇a φc = − d3 x dac φ̇a φc .

Let us finally show explicitly that these charges are conserved using the equations of motion. These are given by
∂µ ∂ µ φa + m2 φa = φ̈a − ∇2 φa + m2 φa = 0.
Then we get
Z
d d
Qa = − d3 x abc φ̇b φc
dt dt
Z
= − d3 x abc (φ̈b φc + φ̇b φ̇c )
Z
= − d3 x abc (φ̈b φc )
Z
− d3 x abc ∇2 φb φc − m2 φb φc
  
=
Z
= d3 x abc ∇φb · ∇φc = 0.

3
Here we used the skew-symmetry of the  tensor several times and integrated by parts to go to the last line. We assume that
the field falls off sufficiently fast so that we can neglect the boundary term of the partial integration.

Exercise 5
Under a Lorentz transformation, the vector xµ transforms as:

xµ → x0µ = Λµν xν ; (27)

and such transformations leave invariant the form

gµν xµ xν = gµν x0µ x0ν . (28)

Replacing the x0µ s by their definition in terms of Λµν ,

gµν x0µ x0ν = gµν (Λµα xα ) Λν β xβ = gµν Λµα Λν β xα xβ = gαβ Λαµ Λβν xµ xν .
  
(29)

In the last step, we have interchanged the indices α and β with µ and ν. Now, equating (28) with the last step in (29), it
follows straightforwardly that:
gµν = gαβ Λαµ Λβν (30)

Now, let us consider an infinitesimal Lorentz transformation of the form:

Λµν = δ µν + ω µν ; |ω|  1 (31)

Using (30), we find that, keeping only the terms up to first order in ω:

gµν = δ αµ + ω αµ gαβ δ βν + ω βν = gµν + ωµν + ωνµ + O ω 2 ,


  
(32)

which requires that ωµν is an antisymmetric tensor for the above transformation to indeed be a Lorentz transformation.

For those of you who are interested in such matters, we introduce an alternative, more mathematically sound means of
thinking about the concept of an ‘infinitesimal’ Lorentz transform. Consider a one-parameter family of Lorentz transfor-
mations, that is a differentiable function Λ : R → SO(3, 1), i.e. for each value of the parameter t, Λ (t) defines a Lorentz
transformation. Take the family so-defined to be such that Λ(0) = I, the identity.

Then for all t,


gµν = gαβ Λ(t)α µ Λ(t)β ν . (33)
Looking ‘infinitesimally’ corresponds to differentiating at the identity, i.e. consider
d  
α β
gαβ Λ(t) µ Λ(t) ν = gαβ Λ0 (0)α µ Λ(0)β ν + gαβ Λ(0)α µ Λ0 (0)β ν
dt t=0
= gαν Λ0 (0)α µ + gµβ Λ0 (0)β ν .

However, according to (33), this is precisely


d   d

gαβ Λ(t)α µ Λ(t)β ν

= gµν = 0.
dt t=0 dt t=0

Therefore, if we define a matrix ω by


ω α µ = Λ0 (0)α µ ,
we have in all that
0 = gαν ω α µ + gµβ ω β ν = ωνµ + ωµν .
To make the connection with our earlier ‘infinitesimal’ form of the Lorentz transformation Λµ ν , we can write down the Taylor
expansion of Λ(t) for t small,

Λ(t)µ ν = Λ(0)µ ν + t Λ0 (0)µ ν + O(t2 )


= δ µ ν + t ω µ ν + O(t2 ).

Besides being completely rigorous, the benefit of this approach is that Λ infinitesimal is characterized by a parameter t being
small, while ωµν is any skew-symmetric matrix and not an infinitesimal object. One-parameter families of transformations,
and in particular one-parameter subgroups (i.e. smooth homomorphisms Λ : R → G), are precisely how one defines the Lie
algebra of a group and its exponential map in a more general setting - i.e. when the group isn’t given as being embedded in
Gl(n, R).

The form of the infinitesimal rotations and boosts are given by:

4
• Rotation:
A generic rotation by an angle θ about the x3 -axis is given by:
 
1 0 0 0
 0 cos θ sin θ 0 
 . (34)
 0 − sin θ cos θ 0 
0 0 0 1

Taylor expanding the trigonometric functions up to first order in θ, we find that an infinitesimal rotation around the
x3 -axis is given by:  
0 0 0 0
 0 0 1 0 
δ µν + ω µν = δ µν + θ 
 0 −1 0 0  .
 (35)
0 0 0 0
As a check on the above, exponentiate the matrix
 
0 0 0 0
 0 0 1 0 
ω=θ 
 0 −1
.
0 0 
0 0 0 0

In order to do this, calculate that  


0 0 0 0
 0 −1 0 0 
ω2 = θ2 
 0 0 −1 0  ,

0 0 0 0
so that
   
0 0 0 0 0 0 0 0
 0 1 0 0   0 0 −1 0 
ω 2k = (−1)k θ2k 
 0
 , k≥1 ω 2k+1 = (−1)k θ2k+1   , k≥0
0 1 0   0 1 0 0 
0 0 0 0 0 0 0 0

This facilitates the calculation


∞ ∞
X 1 2k X 1
exp (ω) = I+ ω + ω 2k+1
2k! (2k + 1)!
k=1 k=1
   
1 0 0 0 (∞ ) 0 0 0 0
 0 0 0 0  X 1 0 1 0 0 
(−1)k θ2k 

 0 0 0 0 +
=  
 0 0 1

2k! 0 
k=0
0 0 0 1 0 0 0 0
 
(∞ ) 0 0 0 0
X 1  0 0 −1 0 
+ (−1)k θ2k+1   0 1 0 0 ,

(2k + 1)!
k=1
0 0 0 0

and the desired expression for this finite transformation follows from identifying the two series here as, respectively,
cos θ and sin θ.

• Boost:
A generic boost along the x1 -axis is given by:
 
γ −γv 0 0
 −γv γ 0 0 
 . (36)
 0 0 1 0 
0 0 0 1

Recall that the rapidity φ is defined via tanh φ = v, so that the above boosts are parameterized by
 
cosh φ − sinh φ 0 0
 − sinh φ cosh φ 0 0 
 . (37)
 0 0 1 0 
0 0 0 1

5
Taylor expanding to first order around φ = 0 (i.e. v = 0), the infinitesimal boost by v is found to be
 
0 −1 0 0
µ µ µ
 −1 0 0 0 
δ ν +ω ν =δ ν +φ   0
. (38)
0 0 0 
0 0 0 0
Again, we check that the above ‘infinitesimal’ transformation gives the correct finite transformation when we exponen-
tiate the matrix  
0 −1 0 0
 −1 0 0 0 
ω=φ  0
.
0 0 0 
0 0 0 0
Again, calculate the square  
1 0 0 0
2 2  0 1 0 0 

ω =φ  ,
0 0 0 0 
0 0 0 0
so that the even and odd powers are
   
1 0 0 0 0 −1 0 0
2k 2k  0 1 0 0   −1 0 0 0 
ω 2k+1 = θ2k+1

ω =θ   , k≥1   , k≥0
0 0 0 0   0 0 0 0 
0 0 0 0 0 0 0 0
This facilitates the calculation
∞ ∞
X 1 2k X 1
exp (ω) = I+ ω + ω 2k+1
2k! (2k + 1)!
k=1 k=1
   
0 0 0 0 (∞ ) 1 0 0 0
 0 0 0 0  X 1 0 1 0 0 
θ2k 

= 
 0 0 1 0 
+ 
2k!  0 0 0 0 
k=0
0 0 0 1 0 0 0 0
 
(∞ ) 0 −1 0 0
X 1 −1 0 0 0 
θ2k+1 

+  0
,
(2k + 1)! 0 0 0 
k=1
0 0 0 0
and the desired expression follows from identifying the two series here as, respectively, cosh φ and sinh φ.
Mathematical Remark: As was pointed out by some of you in class, there is some subtlety involved in
the above operations:
1. If one takes a group, is it true that exponentiating (a representation of) its algebra gives an element
of (a representation of) the original group? Seen from a different perspective, can two groups have
the same algebra?
2. If the answer to the previous question is positive, can one obtain all elements in this way?
Unfortunately, the answer to both of these questions is in the negative. Fortunately, the answer is almost
as nice as one would wish, and is certainly much more interesting than a bland affirmative.

Aspects of the relevant theory may be covered in the course ‘Symmetries and Particles’ and is covered
in the course ‘Lie Algebras and Their Representations,’ though, needless to say, in great abstraction.
Otherwise, see Warner, Foundations of Differentiable Manifolds for general background or Fulton &
Harris, Representation Theory for full details.

Exercise 6
From the previous exercise, a four-vector xµ transforms under an infinitesimal Lorentz transformation as
xµ → x0µ = Λµν xν = xµ + t ω µν xν + O(t2 ) , t small. (39)
Similarly, a scalar field φ transforms according to
φ(x) → φ0 (x0 ) = φ Λ−1 x .

(40)

6
We need to know what Λ−1 looks like finitesimally. From the definition (30), the inverse to any Lorentz transformation Λ is
simply µ
Λ−1 ν = g µα gνβ Λβ α .
Note that if the metric were Euclidean, this would simply say ΛT = Λ−1 . In our case, the inverse to the above infinitesimal
transformation is (neglecting terms at order t2 )

(Λ−1 )µ ν = δµ ν + t ω β α g µα gνβ
= δµ ν + t ωνα g µα
= δµ ν − t ωαν g µα
= δµν − t ωµ ν ,

(42)

In the second line we use the skew-symmetry of ωνα .

Now, Taylor expanding φ to first order in t, we obtain:

φ(xµ ) → φ0 (x0µ ) = φ(xµ ) − t ω µν xν ∂µ φ(xµ ). (43)

The Lagrangian density, L, is a Lorentz scalar and will therefore transform in a like manner to φ under an infinitesimal
Lorentz transformation:
L(xµ ) 7−→ L(x0µ ) = L(xµ ) − t ω µν xν ∂µ L(xµ )
Hence, δL = −t ω µν xν ∂µ L. If you are unconvinced by this reasoning, check explicitly that the Lagrangian for scalar field
theory, given the transformation for φ, does indeed change in this way.

However,

∂µ (ω µν xν L) = ∂µ (ω µν xν ) L + ω µν xν ∂µ L
= ω µν ∂µ (xν ) L + ω µν xν ∂µ L
= ω µν δµν L + ω µν xν ∂µ L
= ω µν xν ∂µ L; (44)

where in the second step we have used that ω µν is a constant tensor, and in the last step we made use of the skew-symmetry
of ωµν to infer that ω µµ = 0. Therefore, we obtain that the variation of the Lagrangian density is a total derivative:

∆L = −∂µ (ω µν xν L) . (45)

There is a conserved current associated with the above symmetry, given by the formula (23) from earlier. Inserting the
expressions derived above for ∆φ and J µ , we obtain:
 
∂L
j µ = − ω αβ xβ ∂α φ − δ µα L

∂ (∂µ φ)

According to the standard definition in field theory, the bracketed quantity is precisely the energy-momentum tensor, T µα .

The total charge Q is given by:


Z
Q ≡ d3 x j 0 (46)
Z
= − d3 x ω µν xν T 0µ . (47)

• For pure spatial rotations, take ω i j 6= 0, else ω µ ν = 0. In this case:


Z
Q = − d3 x ω j k xk T 0j
Z
= − d3 x ωjk xk T 0j
Z
1
d3 x ωjk xj T 0k − xk T 0j

=
2
Z
ωjk
d3 x xj T 0k − xk T 0j

= (48)
2

7
In the third line the skew-symmetry was used to throw away the symmetric part of xk T 0j .

Any skew 3 × 3 array (ωjk ) may be written uniquely as a linear combination of the three skew matrices ωi defined via
(ωi )jk = ijk . We therefore obtain three linearly independent charges
Z
1
d3 x xj T 0k − xk T 0j ,

Qi = ijk (49)
2
which have the interpretation of angular momenta.
• For a Lorentz boost along the xi direction, set ω 0 k = δik and ω µ ν = 0 otherwise. This implies ω k 0 = δik also, since
(using our metric conventions)
ω k 0 = −ωk0 = ω0k = ω 0 k = δik
Therefore, in this case:
Z Z
Qi = − d3 x ω 0k xi T 00 − d3 x ω k0 x0 T 0k
Z
= − d3 x xi T 00 − x0 T 00


Z
d3 x x0 T 0i − xi T 00

=

where to arrive at the final expression we raise the indices 0 and i in each term picking up factors of +1 and −1,
respectively.

Since Qi is conserved, taking its derivative with respect to time gives zero:
Z  Z 
d d 3 0 i0 d 3 i 00
Qi = 0 ⇒ 0 = d xx T − d xxT (50)
dt dt dt
Z Z Z 
3 i0 3 d i0
 d 3 i 00
= d x T +t d x T − d xxT (51)
dt dt
Z  Z Z
d d
d3 x xi T 00 d3 x T i0 + t d3 x T i0

⇒ = (52)
dt dt
(53)

To interpret this equation, recall now that T 00 is the energy density of the quantum field and that T 0k is its linear momentum
density. The first term on the right-hand side is the total (physical) linear momentum (i.e. the space integral of the linear
momentum density) and it is therefore conserved by conservation of momentum. Similarly, the last term on the RHS vanishes
since the linear momentum is conserved. Therefore, we find that the LHS is a constant.

Concerning its interpretation, we find that the center of energy of the field travels with constant speed. We can see from the
fact that the spatial integral indicates some average position of the energy of the field.

Exercise 7
In Maxwell Theory, the dynamics of a co-vector field Aµ (x) are governed by the Lagrangian
1 1
L = − Fµν Fστ η µσ η ντ = − Fµν F µν
4 4
in terms of the field-strength (or Faraday) tensor

Fµν = ∂µ Aν − ∂ν Aµ .

Under the ‘gauge’ transformation


Aµ 7−→ Aµ + ∂µ ξ, (54)
for ξ any smooth function, the field-strength changes according to
2 2
Fµν 7−→ ∂µ (Aν + ∂ν ξ) − ∂ν (Aµ + ∂µ ξ) = ∂µ Aν − ∂ν Aµ + ∂µν ξ − ∂νµ ξ
= Fµν

Where in the last line we employed the fact that mixed partial derivatives commute. We see then that under (54), the
field-strength, and therefore the Lagrangian, is unchanged - i.e. it is gauge invariant.

8
Now, for space-time translations xµ 7→ xµ − aµ , the field transforms by Taylor’s Theorem as
Aµ (x) 7−→ Aµ + aν ∂ν Aµ (x) + O(a2 ),
ie. δAµ = aν ∆ν Aµ , where ∆ν Aµ = ∂ν Aµ . For the field-strength,
Fµν 7−→ Fµν + aλ ∂µλ
2
Aν − aλ ∂νλ
2
Aµ + O(a2 ) = Fµν + ∂λ aλ Fµν + O(a2 ),


with everything evaluated at the space-time point xµ . Putting this transformation into L above, we see that the first order
change in the Lagrangian is
1 1
L 7−→ L − ∂λ aλ Fµν Fστ η µσ η ντ = L − ∂λ aλ Fµν F µν
 
2 4
or δL = ∂λ (aλ L), a divergence.

By Noether’s theorem, we obtain conserved currents T µν for each space-time direction given by the formula
∂L
T µν = ∆ν Aλ − Jνµ ,
∂ ∂µ Aλ
where, in this instance, Jνµ = δνµ L. Calculate that
∂ Fστ
= δσµ δτν − δσν δτµ
∂ ∂µ Aν
so that
∂L 1 ∂ Fστ στ
= − F
∂ ∂µ Aν 2 ∂ ∂µ Aν
1
= − (δσµ δτν − δσν δτµ ) F στ
2
1
= − (F µν − F νµ )
2
= F νµ , (55)
by skew-symmetry of the field-strength (i.e. Fµν = −Fνµ ). Altogether, the array T µν , or Energy-Momentum Tensor, is given
by
1
T µν = F λµ ∂ν Aλ + δνµ F στ Fστ .
4
Let us raise the ν index,
1
T µν = F λµ ∂ ν Aλ + η µν F στ Fστ .
4
Manifestly, this quantity is not a symmetric tensor and cannot be gauge invariant, since under (54)
F λµ ∂ ν Aλ 7−→ F λµ ∂ ν Aλ + F λµ ∂ ν ∂λ ξ,
the second term of which does not, in general, vanish; very undesirable properties for an energy-momentum distribution.

Instead, let us examine the following tensor


Θµν = T µν − F λµ ∂λ Aν
1
= F λµ ∂ ν Aλ − F λµ ∂λ Aν + η µν F στ Fστ
4
λµ ν 1 µν στ
= F (∂ Aλ − ∂λ Aν ) + η F Fστ
4
λµ σν 1 µν στ
= F Fσλ η + η F Fστ
4
λµ ν 1 µν στ
= F F λ + η F Fστ
4
which manifestly is gauge invariant. Furthermore
1
Θνµ = F λν F µλ + η νµ F στ Fστ
4
1
= −F νλ F µλ + η µν F στ Fστ
4
1
= −F λ F + η µν F στ Fστ
ν µλ
4
1
= F νλ F λµ + η µν F στ Fστ
4

9
where we interchange indices ν ↔ λ in the second line, picking up one factor of (−1), simultaneously raise/lower λ in the
third and swap indices again in the final line.

This new tensor is traceless since


1
Θµµ = F λµ Fµλ + δµµ F στ Fστ
4
= −F µλ Fµλ + F στ Fστ
= 0.

Finally, we show that Θµν is also conserved. However, a Noether current is only conserved when the fields are ‘on-shell,’ i.e.
satisfy the field equations, which we must therefore calculate:
 
∂L ∂L
0 = ∂µ −
∂ ∂µ Aν ∂Aν
= ∂µ F νµ ,

where the calculation (55) from earlier was used. Now,the divergence of Θµν can either be found by long-hand (extremely
tedious) or using the a-priori fact that T µν is conserved.

∂µ Θµν = ∂µ T µν − ∂µ F λµ ∂λ Aν − F λµ ∂µλ
2

Aν ,

the first term of which vanishes by Noether’s theorem, ∂µ T µν = 0, the second vanishes by the field equations and the
remaining term F λµ ∂µλ
2
Aν vanishes since F λµ is skew in λ µ but ∂µλ
2
Aν is symmetric in λ µ.

The new object Θµν therefore defines a symmetric, gauge-invariant, trace-free and conserved tensor and is thus a candidate
for a physical energy-momentum tensor. Notice also that when the fields are on-shell, the difference T µν − Θµν is simply a
divergence, so that the two tensors describe the same physics.

Exercise 8
The Lagrangian
1 1
L = − Fµν F µν + m2 Cµ C µ , where Fµν = ∂µ Cν − ∂ν Cµ ,
4 2
has the same dependence on the field derivatives ∂C as the Lagrangian governing electrodynamics in the previous question,
∂L
= F νµ .
∂ ∂µ Cν

The field equations are therefore


 
∂L ∂L
0 = ∂µ −
∂ ∂µ Cν ∂Cν
= ∂µ F νµ − m2 C ν . (56)

Upon taking the divergence we find


2
0 = ∂µν F νµ − m2 ∂ν C ν .
The first term vanishes due to symmetry of second partials and skew symmetry of the field-strength, leaving

m2 ∂ν C ν = 0.

If m is non-zero, we see that the fields satisfy


∂µ C µ = 0.
To obtain the field equation for C0 , set ν = 0 in (56),

0 = ∂µ F 0µ − m2 C 0
∂µ F 0µ = ∂i F 0i
= ∂i ∂ 0 C i − ∂ i C 0


= ∂i Ċ i − ∂i ∂ i C 0
⇒ 0 = ∂i Ċ i − ∂i ∂ i C 0 − m2 C 0
⇒ ∂i ∂ C0 + m2 C0 = ∂i Ċ i ,
i

10
where we variously use the fact that (in the conventions of this course) the Minkowski metric is diag(1, −1, −1, −1), so that
∂0 = ∂ 0 and C0 = C 0 . The field component C0 therefore satisfies the inhomogeneous Helmholtz equation

∇2 + m2 C0 = ∂i Ċ i ,


which, subject to sufficiently nice asympototic behaviour, possesses a unique solution for C0 . For instance, the Green’s
function for the operator ∇2 + m2 ,
eim|x|
G(x) = ,
4π|x|
provides the following solution
∂i Ċ i (t, y) im|y−x|
Z
C0 (t, x) = d3 y e .
R3 4π|y − x|
µ
Recall that the momenta Π conjugate to the Cµ are defined by
∂L
Πµ = .
∂ Ċµ
Again using (55), this gives the following expressions

µ 0µ 0 ,µ=0
Π = −F =
−∂ 0 C i + ∂ i C 0 ,µ=i

The velocities of the dynamically relevant variables Ci are thus given by (lowering the i)

Ċi = −Πi + ∂i C 0 .

Having found this inverse relation between the momenta and the dynamical velocities, the Hamiltonian density can now be
computed

H = Πµ Ċµ − L
 1 1 1
= Πi −Πi + ∂i C 0 + F 0i F0i + F ij Fij − m2 Cµ C µ
2 4 2
The term quadratic in F ij contains no time derivatives and so may be left intact, however, the second term must be
re-expressed as a function of the Πµ ,

F 0i = ∂0C i − ∂iC 0
= Ċ i − ∂ i C 0
= Πi

The complete expression for the Hamiltonian density is therefore


1 1 1
H[Πµ , Cµ ] = − Πi Πi + F ij Fij − m2 C µ Cµ + C 0 ∂i Πi − ∂i Πi C 0
 
2 4 2
Where the additional term −Πi ∂ i C0 has been written as

C 0 ∂i Πi − ∂i Πi C 0 ,
 

which involves an irrelevant 3-divergence term. Since the remainder of the Hamiltonian contains no derivatives in C 0 , C 0
may be regarded as a multiplier, that, in the m = 0 theory, imposes the constraint ∇ · Π = m2 C 0 = 0, which is precisely
Gauss’ Law.

Exercise 9
Let us begin in the greatest generality, i.e. work in (n + 1)-dimensional Minkowski space-time. We are given that the group
(R+ , ×) of positive reals under multiplication acts on the space of field configurations via the action

(λ · φ)(x) ≡ λ−D φ(λ−1 x)

for some D, the scaling dimension of φ. In other (perhaps less technical) words, we have a transformation

φ(x) 7−→ λ−D φ(λ−1 x) , λ > 0.

We show that this is a symmetry of the action


Z
S[φ] = dn+1 x L (φ(x), ∂ φ(x)) ,

11
where the Lagrangian density is given as
1 1
L (φ(x), ∂ φ(x)) = ∂µ φ(x)∂ µ φ(x) − m2 φ(x)2 − gφ(x)p ,
2 2
for real constants m, p and g. We assume p 6= 2, else the third term simply serves to modify the mass.

Recall that a group action is a symmetry iff we have

S[λ · φ] = S[φ] , ∀λ ∈ R+ ,

i.e. if the above transformation leaves S unchanged. So we must show that


Z 
1 −2D 
S[λ · φ] = dn+1 x ∂µ φ(λ−1 x) ∂ µ φ(λ−1 x)
  
λ
2

1 2 −2D
− m λ φ(λ−1 x)2 − gλ−pD φ(λ−1 x)
2

does not depend on λ. Making the subtitution x0 = λ−1 x, then as λ > 0,



dn+1 x = |λn+1 | dn+1 x0 = λn+1 dn+1 x0 , ∂µ = λ−1
∂x0µ
and, abbreviating

= ∂µ0 ,
∂x0µ
we find Z  
1 −2D−2 0 1
S[λ · φ] = dn+1 x0 λn+1 λ ∂µ φ(x0 )∂ 0µ φ(x0 ) − m2 λ−2D φ(x0 )2 − gλ−pD φ(x0 )p ,
2 2
which, since we can replace x0 with x as a dummy variable in the integral, is the original value of the action iff each term in
the integrand,
λ−2D−2+n+1 ∂µ φ ∂ µ φ , m2 λ−2D+n+1 φ2 , gλ−pD+n+1 φp , (57)
does not depend on λ. Invariance of the first term entails
1
D= (n − 1).
2
Having found D, invariance of the second term occurs iff m = 0. Finally, the third term is invariant either when g = 0, else
if g 6= 0, then
n+1
p=2 .
n−1
In the case n = 3, therefore D = 1 and p = 4.

Having established that this is a symmetry, we find the conserved current associated to the ‘infinitesimal’ transformation in
the case n = 3, i.e. we look at the first order changes in the fields and Lagrangian. The change in φ to first order is, by the
chain rule,

d
∆φ(x) = (λ · φ)(x)
dλ λ=1

d −1 −1

= λ φ(λ x)
dλ λ=1
= −φ(x) − xσ ∂σ φ(x).

The Lagrangian transforms as

L ((λ · φ)(x), [∂(λ · φ)] (x)) = λ−4 L φ(λ−1 x), ∂φ (λ−1 x)




Since L on the right depends on λ only through the space-time dependence of the fields, we find the ‘infinitesimal change’ in
L to be, again by the chain rule,

d
∆L = {L ((λ · φ)(x), [∂(λ · φ)] (x))}
dλ λ=1
= −4L (φ(x), ∂φ (x)) − xσ ∂σ [L (φ(x), ∂φ (x))] .

As was to be expected from general principles, this is precisely the divergence

∆L = ∂σ J σ , where J σ = −xσ L.

12
The conserved current provided by Noether’s theorem has the same form as (23) above,

∂L
jµ = ∆φ − J µ .
∂ ∂µ φ

Plugging in the expressions found above, we eventually obtain the conserved current associated to our scaling symmetry

j µ = −∂ µ φ (φ + xσ ∂σ φ) + xµ L.

Once the field equations have been found, it is straightforward to check that this is indeed conserved.

13

Das könnte Ihnen auch gefallen