You are on page 1of 63

Introduction to Quantum Field Theory Professor John Boccio Physics 131 - Particle Physics Seminar Fall Semester 2009

- Swarthmore College
1. Background Information Relativistic Notation When dealing with non-orthogonal coordinates, it is of crucial importance to distinguish between contravariant coordinates x and covariant coordinates x . The distinction arises as follows, consider the set of 2-dimensional non-orthogonal coordinates on the 2 plane show below.

x x 2 x2 60o x1 x 1
1

^ e

Non-orthogonal coordinates on the plane

2 ^ e 1 1

Now consider the coordinates of the point x in the (1,2) basis. In terms of the unit vectors e1 and e2 (where the (1,2) subscripts are labels, not indices; e1 and e2 are vectors, not coordinates), we can write the vector to the point x as

x = x1e1 + x 2 e2

(1.1)

which defines the contravariant coordinates x1 and x 2 (as shown in the figure). The covariant coordinates ( x1 , x2 ) are defined by

x1,2

x e1,2

(1.2)

which are also shown in the figure. Note that for orthogonal axes in flat (Euclidean) space there is no distinction between covariant and contravariant coordinates, which is how you made it this far without worrying about the distinction. However, away from Euclidean space(in particular, in Minkowski space-time) the distinction is crucial. Given the two sets of coordinates, it is simple to take the scalar product of two vectors. From the definitions above, we have 1 1 + y 2 e2 = y1 x e1 + y 2 x e2 xy = x y e (1.3) = y1 x1 + y 2 x2 = y1 x1 + y2 x 2

so scalar products are always obtained by pairing upper and lower indices. The relation between contravariant and covariant coordinates is straightforward to derive:
2

xi = x ei = x1e1 + x 2 e2 ei = x j ei e j gij x j
where we have defined the metric tensor

(1.4)

cos 60 1 gij ei e j = 1 cos 60

(1.5)

Note that we are using the Einstein summation convention - repeated indices (always paired-upper and lower) are implicitly summed over. One can also define the metric tensor with raised and mixed indices in a natural way, gij gik g k gik g jl g kl (1.6) j Note that g j = j , the Kronecker delta. The metric tensor g ij indices in a natural way.
k k

raises (1.7)

x i = g ij x j

Minkowski space is a simple solution in which we use non-orthogonal basis vectors, because time and space look different. The contravariant components of the four-vector x ( = 0,1, 2, 3) are (t, r ) = (t, x, y, z ) = ( x 0 , x1, x 2 , x 3 ) . The flat Minkowski space metric is
3

g = g

1 0 0 0 0 1 0 0 = 0 0 1 0 0 0 0 1

(1.8)

x = g x = ( t, r ) . The scalar This is used to raise and lower indices: product of two four-vectors is written as 0 0 (1.9) a b = a b = a g b = g a b = a b a b

It easily follows that this is Lorentz invariant, that is, a ' b ' = a b Note, that as before, repeated indices are summed over, and upper indices are always paired with lower indices. If we define a contraction in the following way:
B = a b g B = B = a b = a b = contraction

then this ensures that the result of the contraction is a Lorentz scalar. If you get an expression like a b (this is not a scalar because the upper and lower indices are not paired) or (worse) a b c d (which indices are paired with which?) you have probably made a mistake. If in doubt, it is sometimes helpful to include explicit summations until you get the hang of it. Remember, this notation was designed to make your life easier!
4

Under a Lorentz tranformation, a four-vector transforms according to matrix multiplication: (1.10) x ' = x where the 4 4 matrix defines the Lorentz transformation. Special cases of include space rotations and "boosts", which look as follows: 0 0 0 1
0 cos (rotation about z-axis) = 0 sin 0 0 sin cos 0 0 0 1

(1.11)

v v (boost in x direction) = 0 0 0 0

0 0 1 0

0 0 0 1

(1.12)

with = (1 v 2 )1/2 . The set of all Lorentz transformations may be defined as those transformations which leave g invariant:

g = g

(1.13)

To see how derivatives transform under Lorentz transformations, we note that the variation (1.14) = x
x

is a scalar and we should therefore like to write it as = x . Thus, we obtain


5

( )

= , , , x t x y z
= , , , x t x y z

(1.15) (1.16)

and

Thus, / x transforms as a covariant (lower indices) four-vector. Note that (1.17) A = 0 A0 + j A j and

2 = 2 2 = (1.18) t The energy and momentum of a particle together form the components P = ( E, p ) . of its 4-momentum
Finally, we make use (particularly in the later discussion of Dirac fields) of the completely antisymmetric tensor (often known as the Levi-Civita tensor). It is defined by

1 if ( , , , ) is an even permutation of (0,1,2,3) = 1 if ( , , , ) is an odd permutation of (0,1,2,3) 0 if ( , , , ) is not a permutation of (0,1,2,3)


0123 = 0123 = 1

(1.19)

Note that you must be careful with raised or lowered indices, since

You can easily verify that (like the metric tensor g ) is a relativistically invariant tensor, that is, that under a Lorentz transformation the properties (1.19) still hold. Fourier Transforms We will frequently need to go back and forth between the position( x) and momentum(or wavenumber) ( p or k ) space descriptions of a function, via the Fourier transform. As you should recall, the Fourier transform f (k) allows any function f (x) to be expanded on a continuous basis of plane waves. In quantum mechanics, plane waves correspond to eigenstates of momentum, so Fourier transforming a field will allow us to write it as a sum of modes with definite momentum, which is frequently a very useful thing to do. In n dimensions, we therefore write dnk (1.20) f (x) = f (k)eikx n (2 ) and then it is simple to show that

f (k) = d n x f (x)eikx

(1.21)

We have introduced two conventions here which we will stick to in the rest of these lectures, the sign of the exponentials (we could have easily reversed the signs) and the placement of the factors of 2 .
7

The latter convention will prove very convenient because it allows us to easily keep track of the powers of 2 - every time you see a n it comes with a factor of (2 ) n , while d n x 's have no such d k factors. Also remember that in Minkowski space, , k x = Et k x where E = k0 and t = x0 . The Dirac Delta "Function" We will frequently be making use in these lectures of the Dirac delta function (x) , which satisfies

dx (x) = 1

(1.22) (1.23)

and

(x) = 0 , x 0

Similarly, in n dimensions, we may define the n-dimensional delta function (n) (1.24) (x) (x0 ) (x1 )....... (xn ) which satisfies

The -function can be written as the Fourier transform of a constant, 1

d n x (n) (x) = 1

(1.25)

(n) (x) =

(2 )

d n peipx

(1.26)
8

We will also make use of the 1-dimensional step function

1 x > 0 (x) = 0 x < 0


which satisfies

(1.27)

d (x) = (x) dx

(1.28)

Note that the symbol x will sometimes denote an n-dimensional vector with components x , as in (1.24), and sometimes a single coordinate, as in (1.27) - it should be clear from the context. For clarity, however, we will usually distinguish 3-vectors( x ) from 4-vectors ( x or x ). 2. High Energy(Particle) Physics This field attempts to uncover the fundamental structure of both matter and forces, i.e., physics at ultra-short distances. What is short? Furthest galaxies 10 26 m 18 Shortest distances probed 10 m 15 Gyr 5 1017 s Age of universe Shortest particle lifetime 10 24 s This implies Particle Physics is about laws common to > 40 orders of magnitude or more.
9

Basic components of Particle Physics: 1.05 10 34 J s Quantum Mechanics Relativity - c 3 10 8 m / s The basic object from Quantum Mechanics is quantum = "bundle of waves" = "wave packet" deBroglie = h / p size energy = E = hc / c Thus, (Energy) (size) which means that high-energy physics is about short distances. Consider the familiar problem of a particle in a box. In the non-relativistic description, we can localize the particle in an arbitrarily small region, as long as we accept an arbitrarily large uncertainty in its momentum. But relativity tells us that this description must break down if the box gets too small. Let the particle have mass and be confined in a container with a reflecting wall of size L . The uncertainty in the particle's momentum is therefore of order / L . In the relativistic regime, this translates to an uncertainty of order c / L in the particle's energy.

10

For L small enough, L / c (where / c C is the Compton wavelength of the particle), the uncertainty in the energy of the system is large enough for particle creation to occur, i.e., particle-antiparticle pairs can pop out of the vacuum, making the number of particles in the container uncertain! The physical state of the system is a quantum-mechanical superposition of states with different particle number. Even the vacuum state - which in an interacting quantum theory is not the zero-particle state, but rather the state of lowest energy - is complicated. The smaller the distance scale you look at, the more complex its structure. There is therefore no sense in which it is possible to localize a particle in a region smaller than its Compton wavelength. In atomic physics, where NRQM works very well, this does not introduce any problems. The Compton wavelength of an electron (mass = 0.511 MeV / c 2 ), is 1 1
= 2 ( c ) = (197 MeV ) fm 4 10 11 cm c c 0.511 MeV

or about 10 Bohr radii. So there is no problem localizing an electron on atomic scales and the relativistic corrections due to multi-particle states are small. On the other hand, the up and down quarks which make up a proton have masses of order 10 MeV ( C 20 fm ) and are confined to a region the size of a proton, or about 1 fm . Clearly, the internal structure of the proton is much more complex than a simple three quark system and relativistic effects will be huge.
11

Thus, there is no such thing in relativistic quantum mechanics as the two, one or even zero body problem! In principle, one is always dealing with the infinite body problem. Thus, except in very simple toy models (typically in one spatial dimension), it is impossible to solve any relativistic quantum system exactly. Even the nature of the vacuum state in the real world, a horribly complex sea of quark-antiquark pairs, gluons, electron-positron pairs as well as more exotic beasts like Higgs condensates and gravitons, is totally intractable analytically. Nevertheless, as we shall see in this course, even incomplete (usually perturbative) solutions will give us a great deal of understanding and predictive power. As a general conclusion, you cannot have a consistent, relativistic, single-particle quantum theory. So we will have to set up a formalism to handle many-particle systems. Furthermore, the standard position operator from NRQM does not make any sense in a relativistic theoryX - the basis of NRQM does not exist, since particles cannot be localized } to arbitrarily small regions. So the {x first casualty of relativistic QM is the position operator and it will not arise in the formalism we will develop. There is a second, closely elated problem which arises in a relativistic quantum theory, which is that of causality. In both relativistic and nonrelativistic quantum mechanics observables correspond to hermitian operators. In NRQM, however, observables are not attached to space-time points - one simply talks about the position operator, the momentum operator, and so on.
12

However, in a relativistic theory we have to be more careful, because making a measurement forces the system into an eigenstate of the corresponding operator. Unless we are careful about only defining observables locally, i.e., having different observables at each space-time point, we will run into trouble with causality, because observables separated by spacelike separations will be able to interfere with one another. To see this let us construct a "naive" relativistic theory as an obvious relativistic generalization of NRQM. As we will see, the theory violates causality - a single free particle will have a nonzero amplitude to be found to have traveled faster than the speed of light. Consider a free, spinless particle of mass . The state of the particle is completely determined by its 3-momentum k , that is, the components of the momentum form a complete set of commuting observables. We may choose as a set of basis states the set of momentum eigenstates { k } :

P k =k k

(2.1)
P

where is the momentum operator. Note that is an operator on the Hilbert space, while the components k are just numbers. These states are normalized
13

(3) k k' = k k'

(2.2)

and satisfy the completeness relation

d k k k =1
3

(2.3)

An arbitrary state eigenstates

is a linear combination of momentum


3

= d k ( k ) k (k ) = k

(2.4)

The time evolution of the system is determined by the Schrodinger equation i (t) = H (t) (2.5)

where the operator H is the Hamiltonian of the system. The solution of (2.5) is (assuming the Hamiltonian is independent of t)

(t ') = eiH (t 't ) (t)


In NRQM, for a free particle of mass

(2.6) ,

2 k H k = k 2

14

Now to see how causality is violated we write the transition amplitude for particle propagation from x0 at t=0 to x at time t . We have for the unitary time translation operator

iHt it U(t) = x e x0 = x e 1 3 it = d pe 3 (2 ) 1 3 it = d pe 3 (2 )
p2 + m2

2 Pop + m 2

3 it x0 = d p x e

2 Pop + m 2

p p x0
(2.7)

x p p x0 e
ip( x x0 )

p2 + m2

where we have inserted an identity operator (completeness relation) and used ip x x p =e x0 = 0 and consider x >> t , i.e., well outside the light Now take cone, go to one dimension for simplicity and using a stationary phase technique to deal with the exponent. This gives

exponent = E = ipx it p 2 + m 2 E = 0 = ix it p t 2 p2 x2 = 2 2 2 p + m2 p +m p

(x 2 t 2 )p 2 + x 2 m 2 = 0 imx p= x2 t 2
15

Then we have

imx x2 m2 2 U(t) exp it 2 2 + m + i x 2 2 x t x t t 2 m2 = exp it 2 2 x t = exp m x 2 t 2 2 2 x t mx 2

which is exponentially small, but non-zero, implying that causality is violated. Alternatively, we can proceed as follow. Suppose we write (using a Taylor expansion) 2 2 4 k 2 k = k + m m + + O( k ) k0 2m The amplitude then becomes
1 3 it k ik ( xb xa ) = d ke e 3 (2 )

U(t)RQM

Moving to polar coordinates for gives

k and doing the angular integrals

U(t)RQM

i ik xb xa it k = e 3 kdke (2 ) r

16

Now consider this as a contour integral

We can deform the contour like this:

17

Changing variables to y=ik yields

U(t)RQM

1 y xb xa = ydye sinh(t y 2 m 2 2 2 r 0

which is not zero when


2

2 (tb t a ) xb xa < 0

This implies that x is no good(makes no physical sense) as a local operator with good causality properties and we need a multi-particle formalism with particles and antiparticles on the same footing(QFT).
As we will see, QFT solves this problems by virtue of the fact that there are both particles and antiparticles propagating across the space-like interval and their amplitudes cancel one another, thereby preserving causality. For processes that do not violate causality, i.e., involve a time-like separation, this cancellation does not occur even though both particle and antiparticle propagation is occurring. 3. Review of SHO "All particle physics is the SHO" "All perturbation theory relies on it ....."
18

Recall that for a single SHO

p2 1 H= + m 2 X 2 2m 2
First, we apply a transformation P p= , q = X m m so that [ P, X ] = i [ p, q ] = i and

2 H = ( p + q2 ) 2

Now introducing annihilation and creation operators

1 1 a= (q + ip) , a = (q ip) 2 2
we get and while

H = (a a + 1 / 2)

a, a = 1 H , a = a

and

[ H , a ] = a
a
is an
19

which means that a is the creation operator and annihilation operator.

The number operator is N = a a . It has eigenstates also eigenstates of H . The normalized states are

, which are

1 n = a n!
where
0

( )

is the vacuum(ground) state.

A few more useful relations are

a n = n n 1

, a n = n +1 n +1

and, in the Heisenberg representation,

da(t) i = [ a(t), H ] = a(t) a(t) = aei t dt


4. The Electromagnetic Field - no sources, i.e., "free" field The electromagnetic fields satisfy Maxwell's equations.
The choice of a gauge such as the Coulomb(radiation) gauge A = 0 , i ( k x t ) which for a plane wave state A(x,t) = A0 e becomes k A = 0 . In

this gauge, the equation of motion for where

A=0 1 2 = 2 2 2 c t

becomes

20

and

1 A B=A , E = c t

for which the Hamiltonian or energy is given by 1 2 2 3 H rad = E + B d x 2

Let look at Fourier space. In order to keep things most comprehensible we will go to a finite volume of space(rather than the continuum). We will use periodic boundary conditions(BC) in each direction of length L . Ultimately, the volume V must cancel out of all calculations (so we can take the infinite volume or continuum limit). For periodic BC, the momenta in the box are multiples of in each direction 2 k= ( n1, n2 , n3 ) , ni = 0, 1,....... L so that in the infinite volume limit

2 / L

L3 V 3 = = (2 )3 (2 )3 d k states n1 ,n2 ,n3 k

In the finite volume, we use periodic BC in the form

A(0, y, z,t) = A(L, y, z,t),.......

21

In this case, the functions

1 ik x r ( k )e V

, r = 1, 2

form a complete set of transverse, orthonormal vector fields. Here r ( k ) s ( k ) = rs , r ( k ) k = 0 , r, s = 1, 2 where the latter equation is required by the radiation gauge

k A= 0

This setup corresponds to a linear polarization basis with polarizations orthogonal to k . We could also employ circular polarizations, but that would be less convenient for the moment, since it is easiest to keep everything explicitly real. This now allows us to expand A( x,t) as a Fourier series:
* ik x c ik x A( x,t) = r ( k ) ar ( k,t)e + ar ( k,t) e r 2V k k where k = c k . Note that k has dimensions (sec)1 . The normalization factor is chosen for specific reasons as we shall see later. The construction is such that A is explicitly real. 2 1/2

From

A= 0,

we find

2 2 a ( k,t) = k ar ( k,t) 2 r t
22

This is the equation of motion for a harmonic oscillator for each of the infinitely large number of (r, k ) choices. The solution is

i t ar ( k,t) = ar ( k )e k

If we work out H rad we find

H rad

* = k ar ( k ) ar ( k )
k r

or it looks like an infinite number of harmonic oscillators. 2nd Quantization - We now quantize the electromagnetic field by hypothesizing that each of these infinite number of harmonic oscillator equations should be quantized in exactly the way we quantize a single harmonic oscillator system and see what happens. The way it is actually done (we will do this later) is to: 1. treat the fields themselves (at each ( x,t) location) as coordinates 2. use the Lagrangian/Hamiltonian of the field theory to determine the momentum conjugate to the fields when treated as coordinates

23

3. introduce the analogue of the usual QM commutation relation between "coordinates" and "momenta" - this is where the and other factors introduced in the normalization earlier come into play. There is an that we require for the commutator of the "coordinate" A field with its conjugate "momentum" ( A / t) . 4. Then we reformulate QM commutators in terms of operators in the Fourier transform space. The above operators are precisely the ar( k ) in the case of the E&M field and so the ar ( k ) and ar ( k )* objects are no longer simple numbers, but rather operators that work exactly like the a and a operators of the harmonic oscillator. It is just that there is one such operator for every (r, k ) choice. ar ( k ) Now, back to 2nd quantization of the E&M field. We promote * and ar ( k ) to operators ar ( k ) and ar ( k ) obeying the commutation relations ar ( k ), as ( k ') = rs kk ' ar ( k ), as ( k ') = ar ( k ), as ( k ') = 0 and write the quantized Hamiltonian operating in the space of states defined below as

H op

= k ar ( k )ar ( k ) + r k

1 2

24

Note that the commutation relations assume that each "mode" (r, k )
is completely independent of every other mode, i.e., all these pseudo-harmonic oscillator systems are independent of one another. Note: We can now crudely understand a bit better about the in A the normalization factor of the expansion. Roughly, the commutator between the "coordinate" A and its conjugate "momentum" takes the form

j ij i (3) A ( x,t), t A ( x ',t) = i ( x x ')

implying that the is needed when we have harmonic oscillator type normalization for the a and a commutators as above. Continuing on, clearly the states nr ( k ) ar ( k ) nr ( k ) = 0 nr ( k )! obey

N r ( k ) nr ( k ) = ar ( k )ar ( k ) nr ( k ) = nr ( k ) nr ( k )

The eigenfunctions of the radiation Hamiltonian take the form

ki ri

nri ( ki ) = n1 ( k1 ) n2 ( k1 ) n1 ( k2 ) n2 ( k2 ) .......

and have energy, which is computed as


25

H op
ki ri

nri ( ki ) = E nri ( ki )
ki ri

given by

1 E = k nr ( k ) + 2 r k

Raising and lowering works as usual, e.g. ar ( k ) ......, nr ( k ),...... = nr ( k ) ......, nr ( k ) 1,...... and correspondingly, the eigenvalue of the energy operator H op is reduced by k = c k . If we write for the momentum operator the analogue of what we wrote for the energy operator H op , something that we will justify later (it simply follows from the E&M flux S vector ), we have

Pop = k nr ( k ) + r k

1 = k N r ( k ) 2 r k

and then ar ( k ) also reduces the momentum eigenvalue of the state by k . We could also get an expression for the angular momentum (using the circular polarization basis) and by computation show that a single photon state was one with a single unit of spin.
General Process: (1) Write down the classical expression for energy, momentum, angular momentum
26

(2) Rewrite these expressions by substituting in the "Fourier" expansion of the A field (but treating the a 's and a 's as operators) and performing the d 3 x to get an operator form of the classical expression. (3) Compute what happens when the operator form of the quantity acts upon a state Interpretation (1) ar ( k ) is an annihilation operator that removes one photon in the mode ( k, r) of energy k and momentum k from the system a a state. Thus, we have a result in which the and operators "look, feel, taste, ..." like they annihilate and create a single photon. If further investigation reveals no conflict with this interpretation, then probably our guess of how to formulate multi-photon states is correct. (2) Our states can have an arbitrary number of photons. In fact, we can have any number of photons of the same (k, r) value, which says that Bose statistics is implicit. Clearly something in all of this will have to change when we consider fermions.

27

(3) You might worry that the ground state (with no photons) has infinite energy given by

1 k 2 k r
i.e., by the sum of all the "zero-point" energies of the infinite number of harmonic oscillators. However, this energy is unobservable in the sense that all we can detect are excitations relative to the ground state. (4) Fortunately, the ground state does have zero momentum. Non-zero momentum would be observable in a very real sense. (5) We have not yet checked causality and we delay this for a while. How should we check causality in the context of the operator field A( x,t) ? We should require

A( x,t), A( x ',t ') = 0


2

whenever

2 (t t ') ( x x ') < 0

i.e., when the two operator fields are located at a space-like separations they should commute. As we will see later this turns out to be true.
28

(6) However, if we do the same thing for a spin-1/2 Dirac field ( x,t) , we encounter a problem.
ik x k t First, we note that the analogues of the "plane r ( k )e waves" are the Dirac spinor forms us ( k )eik x E t and similarly for the complex conjugate. The algebraic form of u s ( k ) is completely fixed by the requirement that ( x,t) obey the Dirac equation. The result is that ( x,t) takes the form 1/2 ik x E t ik x + E t 1 k k bs ( k )us ( k )e ( x,t) = + ds ( k )vs ( k )e 2VE
k

Employing this form and assuming commutation relations like those for E&M a operators for the b and d operators we do not get zero when computing

[ ( x,t), ( x ',t ')]

for space-like separations. We must switch everything from commutation to anticommutation and get 0 anticommutator for space-like separation of two Dirac fields. Will this be acceptable? We will see that it is since the Dirac fields themselves are not observable. The observable operators (bilinear forms like ( x,t) ( x,t) constructed from them will commute for space-like separation.
29

(7) A bit of notation. It is often convenient to divide the operator field A into its positive and negative frequency components which correspond to the part with the annihilation operator and the part with the creation operator, respectively.

+ A( x,t) = A ( x,t) + A ( x,t)


1/2

with

ik x t + c 2 k A ( x,t) = r ( k )ar ( k )e 2V k r k c 2 ik x + k t A ( x,t) = r ( k )ar ( k )e r 2V k k 1/2

ei t when operated on by Note:

Eop = i

gives E = . t

Summary of differences between 2nd and 1st quantization: 2nd quantization refers to the quantization of fields, in which the ( x,t) remain fields themselves become operators. The coordinates simple numbers.

1st quantization refers to MRQM or RQM in which we made the x


coordinate into an operator and obtained a wave equation for a non-operator wave function.

30

The 2nd-quantized theory, of course, gives the same results as the 1st-quantized approach when the energies and such are small enough that only 1 particle needs to be considered, with creation and annihilation processes being negligible. One final note: By direct calculation state with definite number of photons E same state = 0
since the E field, like the A field from which it is computed, either annihilates or creates an extra photon and then the resulting state does not overlap with the final state. To get a non-zero expectation value for E requires that the state in question be a superposition of states with different numbers of photons, i.e., like a wave packet in QM can only have a non-zero expectation value for both position and momentum if it is a superposition of many plane waves. Thus, the usual classical E fields seen in the laboratory result from a coherent superposition of states with different, but large numbers (in order to have small fluctuations), of photons.

31

5. Classical Fields and Particles Nonrelativistic Classical Mechanics - Nonrelativistic means that time t is singled out as special. System coordinates are q a (t) and system velocities are q a (t) where a = index(label). The Lagrangian determines the dynamics, where

L(q a , q a ,t) = T V S = Ldt = action


ti tf

We have Hamilton's principle of stationary action under transformation a a a

q (t) q (t) + q (t)


tf

q a (ti ) = 0 = q a (t f )
so that

L a L a S = 0 = dt a q + a q q q ti
d q = qa dt
a

Since

( )

if we integrate by parts(using the boundary conditions) we get

L d L a S = 0 = dt a a q dt q q ti
32

tf

where

qa

is arbitrary. Therefore, we have

L d L a =0 a q dt q
which are the Euler-Lagrange equations of motion. The canonical momentum conjugate to q a is L pa = a q and the Hamiltonian , where space" variables, is given by
H (q a , pa ) (q a , pa )

are so-called "phase

a L H = pa q
a

with canonical equations of motion

H = qa pa

H = pa a q

i.e., two 1st-order equations for each index(label). Then if


H =0 t

we have

dH H a H = aq + pa = 0 dt q pa

(remember Einstein summation convention). This result corresponds to energy conservation from time-independence. We will develop a much more powerful version of this shortly.
33

Poisson Brackets These are defined by so that we can write

[ f , g ]PB

f g g f = a a q pa q pa

f + [ f , H ]PB = [ f , H ]PB t

for any f ( p, q)

where will assume from now on that f / t = 0 . We can also show that

q a , q b PB = 0 = [ pa , pb ]PB
Now for NRQM we a (1) promote the q , pa to operators (2) the Poisson brackets to commutators This very important idea is the correspondence principle. (2) implies that
a a q a , pb PB = b q a , pb = i b
a q a , pb PB = b

where you must remember we set = 1 This implies that operators motion
O

satisfy Heisenberg equations of


34

d i O = O, H dt
This is the Heisenberg Picture where operators depend on time t, but state vectors do not. In the Schrodinger Picture state vectors depend on time t, but operators do not so that

(t)

= (t 0 )

H S

The pictures are related by a unitary transformation. Thus, (t) satisfies d i (t) S = H (t) S dt (t) S = eiH (t t0 ) (t 0 ) S

so that the relevant unitary operator(time translation operator) is

U = eiH (t t0 )
Note that observables (matrix elements of operators) agree in the two pictures. Relativistic Classical Field Theory

x x . "t is no longer special; replace with (x ) " Coordinate x is no good; move to field
35

The action is then

a (x), a (x)) S = d x L(
4

where

is the Lagrangian density the Lagrangian is given by L = d 3x L

Variations are now

a a + a

and by analogy, the Euler-Lagrange equations are


L L =0 a a ( )

(remember summation convention). The momenta conjugate to


a

are

L ( a )
0 a is singled out and called the canonical momentum. Then the Hamiltonian (density!) is

0 H = a 0 a L

and the Hamiltonian is

H = d 3x H
36

0 The is the object appearing in the Poisson a ( a sometimes) a : brackets with the field

a (x), b (y) PB = (3) (x y)


The simplest possible field has no indices, (x) . It is a real function of x . 2 1 What should we write for the kinetic energy term? Clearly, 2 mx is not relativistically invariant. The simplest guess for a field analog that is relativistic is

We can always do this since by a redefinition of scale(magnitude) part of can be removed by

( )( ) , = 1 for now 2

the

b ( )( ) ( b )( b ) 2 2
and defining a new

as new = b .

Now what should a mass term look like? Let us try

1 m 2 2 2
Using these guesses we will assume that the simplest Lagrangian for is
37

= ( )( ) 1 m 2 2 L 2 2

Now we find

. We have L = = ( )( ) ( ) 2 ( )

where is the Minkowski metric and we use it to raise and lower indices in SR. We then have

= ( + ) = 2

Therefore,

H = 0 0 L

1 0 i = ( )( 0 ) ( 0 )( ) (i )( ) + m 2 2 2 2 2
0

where i = 1,2,3 only. Now

(i )( ) = (i )( j ) = ij (i )( j ) = ( )
i ij

because we use (+,-,-,-) signature which implies that

0 = 0 , i = i
Thus,
= ( )2 + ( )2 + 1 m 2 2 H 2 2 2
38

We must have H 0 for stability of the vacuum which then implies that = +1 = . Finally, we obtain 1 2 2 4 1 S free [ ] = d x ( )( ) m 2 2 We now work out the equations of motion: L L S = 0 = a = m 2 ( ) ( )

+ m2 = 0
which is the Klein-Gordon equation for a relativistic scalar field. This is the relativistic version of the Schrodinger equation. Interpretation of : This turns out to be a bad theory! Let us try to interpret as a wavefunction. The solutions to the above Klein-Gordon equation are plane waves with

2 2 k k + m = 0 E = k + m

=e

ik x

What is the meaning of the negative energy solution?

39

For fermions Dirac had a pseudo-solution to skirt around this problem. He proposed that the negative energy states were all filled with electrons (an infinite sea) and because of the Pauli exclusion principle no positive energy electron could drop below zero energy and the world would be stable! He also proposed that adding an energy = 2mc 2 would eject an electron from the sea and it would look like the creation of an electron with positive energy and the creation of a hole (empty spot in the infinite sea) with negative energy, which he said was an anti-particle, the positron! For scalar (spin 0) bosons represented by the field there is no exclusion principle and thus we have a problem - the vacuum will be unstable. The wave function approach of QM does not give us a multi-particle formalism and that is the source of the problem, so we must find a better way. That will be QFT and we will not need constructs like the Dirac sea! One more nail in the coffin: suppose we cling to the notion of as a wave function. Consider a particle incident on a step(barrier) 1 potential in one dimension ( x = x ) where

V0 x > 0 region II V (x) = 0 x < 0 region I


40

We then have plane wave solutions

I = eip1 x + Reip1 x II = eip2 x


where

E=

p12 + m 2

, E = V0 +

2 p2 + m 2

Imposing matching conditions I (0) = II (0) , ' I (0) = ' II (0) gives and thus

1+ R = T
2 p1 T= p1 + p2

p1 (1 R) = p2T
p1 p2 , R= p1 + p2
(same as NRQM)

Now let us separate off the rest energy

E = m + , = kinetic energy
We now ask the question: for a given V0 ? Let us look at whether

, what happens as we vary

p2 =
2 2

( E V0 )
2

m2
2

is a real number, which implies that there can be transmission.

p = ( E V0 ) m = (( V0 ) + m ) m 2
2

= (( V0 ) + m ) ( V0 )

41

(a) (b) (c) (d)

if if if if

2m < ( V0 ) < 0

( V0 ) > 2m 0 < ( V0 ) < 2m

( V0 ) < 2m

, , , ,

then then then then

p2 p2 p2 p2

is is is is

real - OK real - OK imaginary - OK real - TRANSMISSION !!!!

This makes no physical sense! Why? Basically, what is going wrong is the assumption that encodes only particles and not anti-particles. The trouble with as a single-particle wave function arises whenever V0 exceeds the rest energy of a particle-anti-particle pair. Q(uantum) F(ield) T(heory) Momentum p is conserved. This means that p is a proper operator (unlike x ). Thus, we will use the momentum basis. We make a 3+1 split and later show relativistic invariance. For the basis we go to the so-called "Fock space" where multi-particle states are tensor products of single-particle states. Two-particle case: we must have exchange symmetry for bosons, i.e.,

k1 , k2 = k1 k2 k1 , k2 = + k2 , k1

k2 , k1 = k2 k1

42

Properties: (1) Orthogonality

(3) (3) (3) (3) k '1 , k '2 k1 , k2 = k1 k '1 k2 k '2 + k1 k '2 k2 k '1

) (

) (

(2) Eigenstates

H k1 , k2 = k1 + k2 k1 , k2 P k1 , k2 = k1 + k2 k1 , k2

( (

and so on for n > 2 particle states. (3) The vacuum state is

0 0 =1 , H 0 = 0 , P 0 = 0
3

with

(4) Completeness

1 3 3 I = 0 0 + d k1 k1 k1 + d k1d k2 k1 , k2 k1 , k2 + .... 2!

where the 2! corrects overcounting due to exchange symmetry. The algebra here can get very messy, so we switch to the occupation # basis: { n } . If we put space in a box of side length L ( L eventually), then we have 2 k= n L
43

and

where n( k ) = occupation # in state with momentum k . We then have H = k N( k ) , P = k N( k ) k k We then define operators a and a for each k (these are discrete when we are in a box)
ak , ak ' = kk ' , ak , ak ' = 0 , ak , ak ' = 0

N( k ) n = n( k ) n

In the continuum limit ( L ) this just goes over to

ak , ak ' = ( k k ')
(3)

Then, we build Fock space states as


ak 0 = 0 k k1 = ak1 0 ,

k1 , k2 = ak1 ak2 0

etc

so that

H = k ak ak + 1 / 2 d 3 k ak ak + 1 / 2 k

in continuum

Relativistic Normalization Under Lorentz transformations

x ' = x
Since k is a vector, we have

k ' = k

44

But what about the vector ? One way to determine its Lorentz transformation rule is to use the completeness relation which says (among other things!) that
k'

3 d k k k = d k' k ' k '


3

Now we can find

by using the transformation properties of d 3 k .

The "mass shell" hyperboloid (shown below) is the surface in 2 2 k-space where the relation k = m holds("real" particles).

Light Cone

Suppose we are interested in positive-energy states. Then under the integral we would restrict the integration using

(k m ) (k )
2 2 0
45

which is Lorentz invariant (LI). Now

d k (k m ) (k )
4 2 2 0

is also LI and
4 2 2 0 4

d k (k m ) (k ) = d k ( k

( )

0 2

2 2 0 k m ) (k )

(k 0 k ) d 3k 4 0 =d k (k ) = 2 k 2 k

where

2 = k + m2
k

and we used the general formula

1 ( f (k)) = (k) f '(k)


The final result is LI. This implies that we should put a factor 2 k beside k of to get something "relativistically normalized":
2 k k = 2 k ' k '

Therefore, we define

k = ( 2 ) 2 k k
3

This has the orthogonality relation


3 (3)

k ' k = ( 2 ) 2 k ( k k ')

The factor in front makes the delta function relativistic.


46

Now we make the big jump...... = sum(integral) over creation and annihilation operators representing particles and antiparticles on equal footing d 3k ik x ik x (x) = ak e + ak e 3 ( 2 ) 2 k

We first need to check the canonical commutation relations to be L sure ....! We have
0 =
( )
0

= 0 =

d 3k

( 2 )

2 k
3

((i )a e
k
ik x k

ik x

+ (+i k )a e

ik x k

= i

(a e ( 2 ) 2
k d 3k
il x

ak e

ik x

)
ik y k

and therefore, the equal-time commutators are

[ (x),

(y)]ET

=
3

(a e ( 2 ) 2
d 3l
3 l l

+a e

il x l

),

(a e ( 2 ) 2
k d 3k
3

ak e

ik y

ET

= i = i

2 ( 2 ) 2 ( 2 ) d 3ld 3 k

d 3ld 3 k

k il x ik y il x ik y + al e , ak e ak e al e ET l

)(

ilx iky k (3) ilx +iky d 3 k ik ( x y ) (3) (3) (l k )e = i + ( k l )e = i ( x y) 3 e l ( 2 )

47

as expected. Note that

[ (x),

(y)]ET

= [ (t, x), 0 (t, y)]

In a similar manner, we have

[ (t, x), (t, y)] = 0 = [ 0 (t, x), 0 (t, y)]

Now let us construct the Hamiltonian. 1 1 2 2 2 2 3 3 1 H = d xH = d x ( 0 ) + ( ) + m 2 2 2 2 1 d 3k 2 2 2 2i k t 2i k t 2 k + m k e ak a k + e = ak a k + k + m 2 + k ak ak + ak ak 2 2


k

Now using the mass shell definition 1 3 H = d k k ak ak + ak ak 2 Now (3) ak , al = ( k l ) or ak ak = ak ak + (3) (0) Remember that
(3) (0)

is "very singular" Therefore,


3

k (3) H = d k k ak ak + (0) 2
The last term is the sum of the zero-point energies(the vacuum energy) of all k -modes; it is infinite!
48

In QFT, we subtract this out, i.e., we ignore it. Equivalently, we assume that all operators are "normal ordered :...:" which means ak terms to the left of all ak terms This that all terms have all means that

1 3 1 3 : H : = d k k : ak ak + ak ak : = d k k : ak ak : + : ak ak : 2 2 1 3 = d k k ak ak + ak ak = d 3 k k ak ak 2
and we have no infinities! In gravity, even a vacuum energy gravitates!! If we work in the box, put in an upper momentum cutoff at k = kmax , then
4 S(zero-point) kmax

The theoretically "natural" value of kmax kmax m planck = theory 1019 GeV
exp t 10 12 GeV

is

Experimentally, , so that we have a factor of 10120 problem in the calculation of S(zero-point) . It is hoped that string theory or quantum gravity research can straighten out this problem. Let us now check out that the Heisenberg equations of motion work out. We should find
49

i [ H , ]ET =

d dt

, i [ H , 0 ]ET =

d 0 dt

Let us prove the first equation. We have


d 3l 3 ilx ilx i [ H , ]ET = i d k k ak ak , al e + al e ( 2 )3 2 l d 3 kd 3l = i k eilx ak ak , al + eilx ak ak , al 3 ( 2 ) 2 l

Now

ak ak , al = ak ak al al ak ak = ak al ak al ak ak = al ak al , ak ak al ak ak (3) = (l k )ak

and similarly, We then get

a a , a = ( k l )ak
k k l (3)

i [ H , ]ET = i = 0

d 3 kd 3l

( 2 )

2 l

k ( ak eilx + ak eilx )

A similar derivation holds for the other equation. Proof complete.

50

All ordering ambiguities for products of field operators can be resolved by "normal" ordering - :all ak to the left: . This ordering is unique because

[a , a ] = a , a
k l k

=0

So free QFT is SHOs and quanta are field excitations. Technically, basis states n where n = # quanta and energy is quantized in units of k . But what really is the field operator? Let us get a handle on it by using it as a hammer and hit some states with it. (A) Hitting the vacuum:

(x) 0 = =

d k

( 2 )
d 3k

2 k

ak eikx + ak eikx

0 eikx k

( 2 )

2 k

eikx

k =

( 2 )

d 3k
3

2 k

which is a linear combination of plane waves(momentum eigenstates). This expression is nicely LI. What is the overlap with state
l

?
51

l (x) l = =

( 2 ) ( 2 )

d 3k
3

2 k 2 k

eikx l k
ikx

d 3k
3

( 2 ) 2 k ( k l )
3 (3)

= eilx
so the 1-particle answer looks like "RQM". Hitting a 2-particle state:

(x) k, l = =
Now

d q

( 2 )
d 3q
3

2 q

aq eiqx + aq eiqx

k, l 2 ( 2 )3 k l
3

( 2 )

2 q

q, k,l eiqx

2 k l ( 2 ) iqx 3 + d q e aq ak al 0 2 q

aq ak al 0 = aq , ak + ak aq al 0 (3) = (q k )al 0 + ak aq , al + alaq 0 (3) (3) = (q k )al 0 + (q l )ak 0

So

(x) k, l =

( 2 )

d 3q
3

2 q

q, k,l eiqx + eikx l + eilx k


52

which is the sum of a 3-particle state and two 1-particle states. Therefore, hitting a 2-particle state with hammer
l with wave function eikx k with wave function eilx
(x)

gives

and a relativistically invariant sum over


q, k,l with wave function eiqx

In general, hitting an n-particle state with the (n-1)- and (n+1)-particle states. Covariance and Causality/Locality We had

operator gives

[ (0, x), (0, y)] = 0


d 3 kd 3l ak eikx + ak eikx , al eilx + aleilx [ (x), (y)] = (2 )3 2 k 2 l d 3k eik( x y) eik( y x ) = (2 )3 2 k = i + (x y) i + (y x)

What about unequal times?

where

d 3k i + (x y) = eik( x y) (2 )3 2 k
53

is manifestly LI. Now, LI implies this can only depend on (x y) as

(x y) (x y) = (x y) (x y) = (x y)

For causality, we worry about spacelike intervals . But if 2 x 0 = y 0 , then (x y)2 = 0 x y < 0 which is always true. Now we know that the equal time commutator is OK. Therefore, in general it is OK at spacelike separations (that is what LI means). So
(x y)2 < 0 , [ (x), (y)] = 0

(x y)2 < 0

is the statement of "microscopic causality". In NRQM, any Hermitian operator is a physical observable. Not so in QFT! - Causality is more restrictive. Measurements are local, which implies that the algebra of observables must be local. Let us elaborate a bit more on the Lagrangian approach. Let us take a real example from solid state physics to build up our sense of what is going on. Imagine for the moment that you have a discretized glob of jelly or a solid state lattice. Each particle of the jelly or ion on the lattice will have its own coordinate and its own momentum, and, in addition, these particles or ions will interact with one another in some fashion.
54

How would we quantize this system? i = 1,..., N(N = very large) We would introduce a qi and pi , particle and some potential

for each

V (x , x )
i j i, j

and require

[ q , p ] = i
i i

ij

Here i and j denote the location of the ions within the lattice, while qi is a displacement coordinate of the ion relative to its central location(equilibrium position), and pi is the conjugate momentum for this displacement coordinate. If we carried this through, then we would find it easiest to deduce the excitations of this system by defining creation and annihilation operators for (in the lattice example) "phonon" excitations(lattice vibrations) on the lattice. You had and the would have a "vacuum" state 0 in which all the ions simply their zero-point energy (like a single harmonic oscillator) ak then there would be creation operators that would excite lattice as a whole to contain a phonon described by momentum

k : k = ak 0

55

This state would be a state describing a coherent "wave-like" motion of the lattice as a whole. The precise energy of the phonon state would be determined by such things as the restoring force keeping each ion at its lattice site location and on the potential describing interactions between different ions on the lattice. Now imagine going to the continuum limit of the ion lattice, which would be sort of like a jelly. At every ( x,t) location there would be a particle, whose coordinate location we could denote by ( x,t). This latter coordinate location can be thought of as the coordinate describing the displacement of the ion relative to its central location ( x,t) . It is this latter "displacement" coordinate that is the one that should be quantized (just like a simple harmonic coordinate is really a displacement coordinate). We would then want to quantize in a continuum sort of way by defining the momentum ( x,t) of this particle (actually it is a momentum density) and require that

(3) [ ( x,t), ( x ',t)] = i ( x x ') , = 1

Note that the delta function makes sense in that we integrate over a small volume in x ' and think of ( x) as the coordinate of another small volume, and think of each little volume as being a pseudo-particle(pp), then the above description would be equivalent to
56

If there was more than one degree of freedom at each ( x,t) (such as
spin degrees of freedom or the like), we would attach some appropriate index to and and require (3) ( x), ( x ',t) = i ( x x ') Field Theory In field theory, the coordinates are no longer the coordinates of some ion or jelly component, but rather coordinates in an abstract sense, i.e., really "fields" that appear in a Lagrangian density. But, we quantize in exactly the same way as a coordinate. One introduces a Lagrangian(density) from which the field equations follow via Hamilton's principle. One introduces momenta(density) operators conjugate to the fields and imposes canonical commutation relations between the fields and these conjugate momenta. Everything follows from the Lagrangian density. In particular, all conservation laws follow from the symmetries of the Lagrangian. This approach gives exactly the results we have just described in the E&M field case.
57

[coord of pp centered at x, conj mom of pp centered at x ']= i xx '

One must use manifestly relativistic covariant notation. We restrict to theories which can be derived by means of a variational principle from the action S() = d 4 x L( , )

where r is some index (like components of A ) often related to spin and r is the derivative of the field r .
The fields r may be real or complex. In the case that they are complex, the real and imaginary components (or equivalently and * ) are treated as independent objects. We postulate that the equations of motion (EOM) follow from a variational principle as follows. Let

r (x) r (x) + r (x)


with vanishing on the surface () bounding the region We require S() = 0 under this variation. This leads to (dropping the r index for the moment) L 4 L S() = 0 = d x + ( ) L L 4 L 4 0 = d x + d x x x
58

where the last line follows from partial integration and ( ) = x The last term in the last line is zero after using Gauss's theorem in 4D to convert the volume integration to a surface integration and using the fact that = 0 on the surface. Then using the fact that the at each point within the volume is independent of other points leads to

L L =0 x
In order to get the conjugate momenta right, we temporarily discretize space and time by dividing space up into small cells of equal volume xi labeled by index i = 1,2,3,.... We approximate the values of the fields within each cell by the value at the center of the cell. Think qi (t) (i,t) ( xi ,t) . Write the Lagrangian of the system of cells as

L(t) = xi Li ( (i,t), (i,t), (i ',t))


i

where i' denotes the label of neighboring cells. The conjugate momenta to the qi are
59

L L pi (t) = i (i,t) xi qi (i,t)


The Hamiltonian of the discrete system is then given by

H = pi qi L = xi i (i,t) (i,t) Li
i i

In the continuum limit(using the notation x = ( x,t)), we define


L (x) =
and and

L(t) d x L( , )
3

H (x) = (x) (x) L( , )


H
will be

Since L does not depend explicitly on time, constant in time (will prove later).
The Klein-Gordon Example

In this case we have a real scalar field (x) . To derive the Klein-Gordon equation, we proceed just as in NRQM. (1) Write down the relationship between energy and momentum, in the relativistic case:
60

2 E p 2 = 0
2

where is a constant with dimension mass = (length)1 that we identify as the particle's mass. (2) Replace (using

= c =1
Ei t ,

p i

and operate the resulting differential equation on the one-particle wave function which we call . The result is
2 2 2 2 2 2 t 2 + ( x,t) = 0 = t 2 + ( x,t)

or where

+ ( x,t) = 0
2

2 2 = 2 t which is the KG equation.

The Lagrangian that gives this equation of motion is

= 1 2 2 L 2

)
61

Proof: We must first compute, using dummy indices , and the metric tensor, for the first term in L ,

L 1 1 = g = g + 2 2

1 = g + g = 2
Thus, we have

The EOM is

+ 2 ) (x) = 0 (

L = 2 L = = x L = = x x x L L = 0 + 2 = 0 x

Being the KG equation implies that the quantized degrees of freedom that will be associated with it will turn out to be spinless neutral bosons. Note that the L form is quite abstract until this connection is made.
62

The conjugate momentum is (by direct computation)

L (x) = = (x)

Similarly, the Hamiltonian density is


1 H = L = 2 2 ( )2 2 2 2 1 2 = + ( )2 + 2 2 2

We then 2nd quantize by (1) Imposing the "usual" canonical commutation rules ( j,t), p( j ',t)] = ( j,t), ( j ',t) x j ' = i jj ' x j [ with other commutators equal to zero. In the continuum limit, you can think of dividing by obtain ( x,t), ( x ',t)] = i (3) ( x x ') [ with other commutators equal to zero. Note that the commutation relations are at equal time, as always is the case for standard quantization. (2) In the KG case, the important non-zero commutator reduces to ( x ',t) = i (3) ( x x ') ( x,t),
63

xj

to

You might also like