Sei sulla pagina 1di 54

Solutions to problems in

\Gauge Theories in Particle Physics", Third Edition, volume 1:


\From Relativistic Quantum Mechanics to QED"

Chapter 2
2.1 The quantum numbers of the `anything' must be B = 1 Q = 0 in both
cases, together with S = ;1 in the rst case and S = +1 in the second
consider what are the lightest single-particle or two-particle states with
these quantum numbers.
2.2 (i) 4-momentum conservation gives p + k = p + k which may be written
0 0

as p = (k;k )+p squaring both sides gives p 2 = q2+2p(k;k )+M 2c2,


0 0 0 0

and the result follows by evaluating the dot product in the frame such
that p = (Mc 0). For highly relativistic electrons E  cjkj E  0

cjk j, whence q2c2 = (E ; E )2 ; c2(k ; k )2  ;2EE + 2EE cos  =


0 0 0 0 0

;4EE sin2 =2. For elastic scattering p 2 = M 2c2, and hence from the
0 0

rst result q2 = ;2M (E ; E ). Combining this with the (approximate)


0

formula for q2 in terms of  gives the formula for E . 0

(ii) E = 4:522 GeV (negelecting the electron mass).


0

(iii) The invariant mass of the produced hadronic state is (p 2=c2)1=2


0

using the rst displayed equation and the (approximate) expression for
q2 in terms of  gives 1238 MeV=c2 (approximately equal to the mass
of the I = 3=2  resonance).
(iv) The term `quasi-elastic peak' in this context means the value of E 0

corresponding to a collision between the incident electron and a single


nucleon in the He nucleus, ignoring the binding energy of the struck
nucleon (i.e. it is treated like elastic scattering). Since the binding
energy is of order 10 MeV, very much less than the electron energy,
this is expected to be a good rst approximation to the kinematics. It
gives E = 355:6 MeV.
0

The struck nucleon is, however, moving by virtue of being bound inside
the He nucleus (it has a bound state wavefunction and an associated
momentum distribution). Rather than attempting a more realistic cor-
rection based on the nucleon wavefunction, we can get a fair idea of
a typical nucleon momentum by using the `uncertainty relation' esti-
mate p  h =R where R is the nuclear radius. Here R  1:5fm, and
p  130MeV=c, which gives a struck nucleon speed of order v=c  0:14.
(Note that He is a tightly bound nucleus, and the formula for the nu-
clear radius R = 1:1  (A)1=3 is not really applicable - of course, we
are making rough estimates anyway.) Considering con gurations with
the outgoing electron moving parallel/antiparallel to the struck nucleon
gives a typical shift in E of order  50 MeV. Note that this is quite a
0

bit bigger than the nucleon's binding energy - the relativistic transfor-
mation has ampli ed the eect.
2.3 (i) 5.1 eV.
(ii) (a) 1.29 (b) 0.75
(iii) h1  2i = +1 for S = 1 ;3 for S = 0. Hyper ne splitting =
8.45 10 4 eV.
;

(iv) 0.57.
2.4 One-gluon exchange con nement.
Ground state expected to be at the minimum of E (r) as a function of
r, i.e. at r0 such that dE (r)=drj(r=r0) = 0.
Egr(cc)  3:23GeV.
Threshold for production of `open charm' (DD states) opens at about
3.73 GeV.
2.5 E = 2a cos 2  a j+i = cos j1i + sin j2i j;i = ; sin j1i + cos j2i.
System is in state cos j+i ; sin j;i at time t = 0 evolves to


cos expi(2a cos 2 + a)t=h ]j+i ; sin expi(2a cos 2 ; a)t=h ]j;i
at time t amplitude of j2i in this state is
cos  sin fexpi(2a cos 2 + a)t=h ] ; expi(2a cos 2 ; a)t=h ]g
modulus squared of this gives result.
2.6 (iv) d = 2 R = 1mm: MP3+d  1:54 TeV.
Chapter 3
3.1 From the inverse transformation and the chain rule
@ = @t @ + @x1 @
@t @t @t @t @x1
0 0 0

we nd "  !#
@ = @ ;v ; @
@t 0
@t @x1
and similarly for ; @x@1 .
0

3.2 Six. F 21 = B 3 F 32 = B 1 F 13 = B 2 F 10 = E 1 F 20 = E 2 F 30 = E 3.
Consider  = 1 component of (3.17) LHS is
@F 1 = @2F 21 + @3F 31 + @0F 01
 !1
= @B 3 @B 2 @E 1
; ; = ( r  B ) 1 ; @E
@x @x
2 3 @t @t
which veri es the rst component of (3.8).
3.3 It is advisable in such manipulations to let the whole expression act
on an arbitrary function F (x) say, which can be removed at the end.
Thus
;
@ fe iqf (x)F (x)g
p^e iqf (x)F (x) = ;i @x ;

= ;i(;iq) @f
@x e iqf (x)F + e iqf (x) ; i @F (x)
; ;

@x
@f
= (;q) @x e i qf (x )
;
F +e i qf
; (x ) p^F (x)
multiplying this from the left by eiqf (x) gives the result after F is re-
moved.
Chapter 4
4.1(b) = i(

_ ;
_
) j = i 1(
r
; (r
)
).
  ;  

For
= N e ip x = 2jN j2E j = 2jN j2p j  = 2jN j2p .
; 
4.2(ii) Consider for example i = ; i . Multiply from the left by to ob-
tain i = ; i . Now take the Trace and use Tr( i ) = Tr( i 2) =
Tr( i).
4.4(b) (  a)(  b) = iai j bj (sum on i and j ) = i j aibj
= (ij 1 + iijk k )aibj = a  b1 + i  (a  b):
4.6(i) (  u^ )2 = 1. Hence   u^ 12 (1 +   u^ )
= 21 (  u^ + 1)
. Projection
operator for   u^ = ;1 is 12 (1 ;   u^ ):
(ii) Take
 !  i !  1 !  !
N 1 N 1 +

+ = 2 (1+u^ ) 0 = 2 sin ei 1 ; cos cos  sin  e cos 2 =2
0 = N sin =2 cos =2ei
;

 i !
and choose N = 1=(cos =2) for normalization. Possible
is ; sin =
;
2e :
;
cos =2
4.8 Consider jy for example. We have
0

jy = ! y ! = ! e i
x =2 y ei
x=2!:
0 0y 0 y ;

Now
ei
x=2 = 1 + ix =2 + 12 (ix =2)2 + 3!1 (ix =2)3 + : : :
= 1 + ix =2 ; 21 ( =2)2 ; 3!1 ix( =2)3 + : : :
= cos =2 + ix sin =2
as in (4.80). Hence
jy = ! (cos =2 ; ix sin =2) y (cos =2 + ix sin =2)]!:
0 y

But
 ! !  !  !
x y = 0 x 0 0 y 0 x y 0 i z
x y 0 = x y 0 = i z 0 = i z
similarly y x = ;i z and x y x = ; y 2x = ; y : Hence
jy = ! cos2 =2 y + 2 sin =2 cos =2 z ; sin2 =2 y ]!
0 y

= ! cos y + sin z ]!
y

= cos jy + sin jz :
4.10 Under (4.83):
  =   =  e i n^ =2 ei n^ =2:
0 0 0y 0 y ;  

 !  !
 0 1 0
Now = 0   = 0 ;1 , and = , so that

e i n^ =2 = ei n^ =2
;  

and the exponentials cancel.


Under (4.90): use x = ; x.
4.11 (a)     +     = 2g .
(b) Taking the dagger of (i @ ;m) = 0 gives ;i(@ )  ;m = 0,
or  (i  @; + m) = 0 where the notation  @; means that the derivative
y y y

y y

acts on what is to the left of it, i.e. on  . Multiply this equation from
y

the right by  0 = and use   =   (obvious for  = 0, check it


for the space components) this gives (i ;@ + m) = 0:
y

6
(c) @(  ) = ( 6 ;@ ) + (6 @ ) = ;m
 + m  = 0.
4.13 V^KG = iq(@A + A@) ; q2AA.
4.14 (ii) (  r)(  A) = i@i j Aj  = i j @i(Aj )
= (ij 1 + iijk k )@i(Aj ) = @i(Ai) + iijk k @i(Aj )
= r  (A) + i  fr  (A)g:
4.15 (i) For a massive particle the direction of its momentum reverses on
transforming to a frame moving parallel to, and faster than, it this is
not possible for a massless particle which moves at the speed of light.
(ii) We give the solution in a form which adapts to the general case.
We have (E ; xpx )uR = 0. Mulitiplying from the left by (1 ; 12 xx)
and inserting a unit operator (to rst order in x) it follows that
(1 ; 12 xx)(E ; xpx)(1 ; 21 xx)(1 + 12 xx)uR = 0:
But (1 + 12 xx)uR = uR, and
0

(1 ; 12 xx)(E ; xpx)(1 ; 21 xx) = E ; xpx ; E xx + xpx


(to rst order in x), which is just E ; xpx , whence (E ; xpx)uR = 0
0 0 0 0 0

as required.
(iv) See part (i). Since, for m 6= 0, we can reverse the helicity by a
Lorentz transformation, we expect that in a Lorentz-invariant theory
mass terms will mix states of opposite helicity.
(v) let BR = (1 + 12 xx) BL = (1 ; 21 xx), so that uR = BRuR uL =
0 0

BLuL: Equation (4.98) is (E ; xpx)


= m. So we have
BL(E ; xpx)BLBL 1
= mBL. But BL 1 = BR, and part (ii) showed
; ;

that BL(E ; xpx)BL = E ; xpx. Hence it follows that (E ; xpx )


=
0 0 0 0 0

m if
= BR
and  = BL. Similarly for (4.99).
0 0 0

Chapter 5
5.1 Consider the term
R L  @ 2 dx. We have
0 @x
@
= X A (t)  r  sin  rx  :
1

@x r=1 r L L
 2
The quantity ` @ @x ' is the square of this entire series - i.e. it involves
not only terms of the form
    x 
2  2   2x 
2
A1(t) L sin L  A2(t) L sin L 
etc., but also all the `cross' terms such as
     2   2x 
A1(t) L sin x L  A 2 (t ) L sin L :
To make sure we are including all these cross terms, we use a dierent
summation index for the two independent summations when writing
the product:
Z L  @
!2 Z L "X  r   rx #
1

dx = Ar(t) L sin L
0 @x 0 r=1
"X  s   sx #

1

As(t) L sin L dx
s=1
X  r   s  Z L  rx   sx 
sin L sin L dx: ( )
1

= Ar(t)As(t) L L
rs=1 0
We need to evaluate the integral:
R L sin  r x  sin  s x  dx
0 L
R n h i Lh io
= 0 2 cos (r ; s) L ; cos (r + s) xL dx
L 1 x
n h i h ioL
= 21 (rL s) sin (r ; s) xL ; (rL+s) sin (r + s) xL 0 :
;

Here both r and s are positive integers, so the second term vanishes.
The rst does also, except when r = s, in which case it has the value
L=2 as can be seen by evaluating the integral directly for r = s. So we
have
Z L  rx   sx 
sin L sin L = L=2 for r = s
0
= 0 for r 6= s:
This means that in the sums over r and s in ( ), only the term in which
r = s survives, after the integral over x has been done. So ( ) becomes
X1  r 2
2
(Ar(t)) L L
r=1
and Z L 1  @
!2 L X 1
1
2
c @x dx = 2 !r2A2r :
0 2 r=1 2
Similarly for the
_ 2 term.
5.2 S = R0t0  12 mx_ 2 + mgx]dt.
Z t0 1
(a) x = at : S(a) = ( 2 ma2 + mgat)dt = 12 ma2t0 + 21 mgat20:
0
Choose `a' so that end point of this trajectory coincides with end point
of the Newtonian trajectory in (b), which is at x = 12 gt20. So we set
at0 = 12 gt20 i.e. a = 12 gt0. Then S(a) = 83 mg2t30 = 0:375(mg2 t30). (b):
S(b) = 13 mg2t30 = 0:3333(mg2 t30). (c): b = 12 gt0 1 S(c) = 207 mg2t30 =
;

0:35(mg2t30). So S(b) < S(c) < S(a).


5.3 (a) We want to verify that q^_ = ;i^q H^ ] is consistent with q^_ = p^=m.
We have
^q H^ ] = ^q 21m p^2 + 21 m!2q^2] = 21m ^q p^2]
= 21m fp^ ^q p^] + ^q p^] p^g using A
^ B^ C^ ] = B^ A
^ C^ ] + A
^ B^ ] C^
= 21m f2i^pg = mi p^
whence ;i^q H^ ] = p^=m.
5.4 (a) From (5.61) and (5.62), q^ = (^a +^a )=(2m!) 12 . So 21 m!2q^2 = 14 !(^a +
y

a^ )2. Similarly for the 21m p^2 term.


y

(b) ^a ^a ] = 1: (^a a^ + 12 )! a^] = ! ^a ^a a^] = ! fa^ ^a ^a] + ^a  a^] a^g =
y y y y y

;!^a.
(c) We rst prove that a^(^a )n ; (^a )na^ = n (^a )n 1, by induction: (i) it
y y y ;

is true for n = 1 (ii) multiplying from the left by ^a and using ^a a^ = y y

a^^a ; 1 in the rst term proves the result for n + 1 assuming it's true
y

for n. Letting this result act on j0i we deduce a^(^a )n j0i = n(^a )n 1 j0i. y y ;

Hence h0j(^a)n (^a )nj0i = h0j(^a)n 1 ^a(^a )nj0i = nh0j(^a)n 1 (^a )n 1 j0i, and
y ; y ; y ;

using induction again (or iterating this step another n ; 1 times) it


follows that h0j(^a)n(^a )nj0i = n!, provided j0i is normalized.
y

(d) n^ jni = p1 ^a a^(^a )nj0i = p1 a^ fn(^a )n 1 + (^a )na^gj0i


y y y y ; y

n! n!
= n p1 (^a )n j0i = njni: y

n!
5.6 (a) From (5.116),
Z dpk ^a(k)eikx i!t + a^ (k)e ikx+i!t]

^(x t) =
1
; y ;

2 2!
;1

and so ^ (y t) is given by


Z dpk (;i! )^a(k )eik y i! t + (i! )^a (k )e ik y+i! t]:
0

^ (y t) =
1
0 0 0 0
0 0 ; 0 y 0 ;

;1 2 2! 0

Note the use of a dierent integration variable in ^ (x t), to ensure that
the operators inside the integral for
^ are treated independently of those
inside the integral for ^ . In the commutator 
^(x t) ^ (y t)] there are
therefore four types of term: ^a(k) ^a(k )] ^a (k) a^ (k )], which vanish
0 y y 0

by the second line of (5.117), and ^a(k) a^ (k )] ^a (k) ^a(k )]: The rst
y 0 y 0

of these latter two yields


Z Z
dk dk (21 )2 p 1 eikx i!t e ik y+i! t(i! )2(k ; k )
1 1
0 0
0 ; ; 0 0

;1 ;1 4!! 0

Z
= dk 21 i! 21! eik(x y) = 2i (x ; y)
1
;

;1

using (E.25). The ^a (k) a^(k )] term gives the same.
y 0

(b) Z

^(x1 t1)
^(x2 t2)] = d3p k Z 1
d3pk 1 0

(2) 2E
3 ;1 (2) 2E
3 ;1
0

(^a(k)e ik x1 + ^a (k)eik x1 ) (^a(k )e ik x2 + a^ (k )eik x2 )]:


;  y  0 ;
0
 y 0
0


The surviving terms are the `^a ^a ]' and `^a  a^]' ones which give
y y

Z 1
dkp Z 1
dkp e ik x1 eik x2 (2)33(k ; k )
0
0
;   0

;1 (2)3 2E (2)3 2E
;1
0

;eik x1 e ik x2 (2)33(k ; k )]
 ;
0
 0

Z d3k e ik (x1 x2 ) ; eik (x1 x2 )]


=
1
;  ;  ;

(2)32E
;1

with k  (x1 ; x2) = E (t1 ; t2) ; k  (x1 ; x2). For t1 = t2 the integral is
Z d3k
ik (x1 x2 ) ; e ik (x1 x2 )]
(2)32E e
 ; ;  ;

which vanishes since the integrand is an odd function of k. So D is


Lorentz invariant (see Appendix E) and vanishes when t1 = t2. It
therefore also vanishes at all points which can be connected to `t1 = t2'
by a Lorentz transformation. Noting that for t1 = t2 the invariant
interval (x1 ; x2)2 is spacelike (i.e. is less than zero), we deduce that
D vanishes for all x1 x2 such that (x1 ; x2)2 < 0.
5.7 We have
Z dpk ^a(k)eikx i!t + a^ (k)e ikx+i!t]

^(x t) =
1
; y ;

2 2!
Z
;1

^
@
= 1
dpk ^a(k)(ik)eikx i!t + a^ (k)(;ik)e ikx+i!t]
; y ;

@x ;1 2 2!
and so
 !
1 Z dx @
2 = 1 Z
1 1
k ^a(k)(ik)eikx i!t + ^a (k)(;ik)e ikx+i!t]:
dp ; y ;

2 @x 2 (2) 2!
Z
;1 ;1

1
k ^a(k )(ik )eik x i! t + a^ (k )(;ik )e ik x+i! t]:
dp 0
0 0
0
;
0
y 0 0 ;
0 0

;1 (2) 2! 0

There are four terms when the brackets are multiplied out. For exam-
ple, the rst is
1 Z dx Z dpk Z dpk (ik)(ik )^a(k)^a(k )ei(k+k )x i(!+! )t:
0
0 0
0
;
0

2 2 2! 2 2! 0

The integral over x yields (2)(k + k ) the k -integral can then be


0 0

done leading to
1 Z dp k 1 p1 (ik)(;ik)^a(k)^a(;k)e 2i!t ( ) ;

2 (2) 2! 2 2!
using ! = jk j = k = !. The three other terms can be reduced
0 0

similarly. Meanwhile,
Z dx
^ (x t) =
^_ (x t) = p ^a(k)(;i!)eikx i!t + a^ (k)(i!)e ikx+i!t]
; y ;

2 2!
and so
1 Z ^ 2dx = 1 Z dx Z dpk ^a(k)(;i!)eikx i!t + a^ (k)(i!)e ikx+i!t]: ; y ;

2 2
Z dk 2 2!
p ^a(k )(;i! )eik x i! t + ^a (k )(i! )e ik x+i! t]:
0
0 0 0 0
0 0 ; y 0 0 ;

2 2!0

For the `^aa^' term the x-integral yields


1Z dpk (;i!)2^a(k)^a(;k)e 2i!t ;

2 (2) 2!]2
which cancels against ( ). Similarly, the `^a ^a ' terms cancel, and the
y y

`^aa^ ' and `^a a^' terms give (5.119).


y y

5.8 (a) We have



^_ (x t) = ;i
^(x t) H^ ]
Z  ^ !2
= ;i
(x t) 2 f^ (y t) + @@y
gdy]:
^ 1 2


^ commutes with (@
^=@y)2, and the non-vanishing term on the RHS is
1 Z
^
;i 2 
(x t) ^ 2(y t)dy]
; i Z
= 2 f^(y t)
^(x t) ^ (y t)] + 
^(x t) ^(y t)]^(y t)gdy
; i Z
= 2 f^(y t)i(x ; y) + i(x ; y)^(y t)gdy = ^ (x t)
as required.
Chapter 6
6.1 The RHS is
1 Z dt Z t1 dt f^(t )f^(t ) + Z dt f^(t )f^(t ) : (A)
1 1

1 2 1 2 2 2 1
2 ;1 ;1 t1
The second term is
1 Z dt Z dt f^(t )f^(t ):
1 1

1 2 2 1
2 ;1 t1
With the usual convention for double integrals, in this expression the
integral over t2 is understood to be performed rst, holding t1 xed,
and then t1 is integrated over. We proceed by changing the order of
integration in this double integral. In order not to make a mistake, it
is a good idea to draw a diagram of the t1 (horizontal) - t2 (vertical)
plane, with the region of integration shaded over: it is the whole of the
region lying above the line t1 = t2. When we write the integral in the
other order (i.e. doing rst the t1 integral and then the t2 one) we must
be careful to arrange the limits so that the required region is covered.
The result is
1 Z dt Z t2 dt f^(t )f^(t ):
1

2 1 2 1
2 ;1 ;1

We now rename t1 as t2, and t2 as t1, showing that this term is in fact
equal to the rst in (A).
6.2 In the rest frame of C,
q q
mC = EA + EB = m2A + p2 + m2B + p2:
q
Taking the term m2A + p2 to the LHS, squaring, cancelling the p2,
and squaring again we arrive at
(m2A + m2C ; m2B)2 = 4m2C(m2A + p2):
This gives jpj as required.
6.3 (iii)
@ (t ; t )
^(x  t )
^(x  t ) + (t ; t )
^(x  t )
^(x  t )]
@t1 1 2 1 1 2 2 2 1 2 2 1 1

= (t1 ; t2)
^(x1 t1)
^(x2 t2) + (t1 ; t2)( @t@
^(x1 t1))
^(x2 t2)
1
;(t2 ; t1)
^(x2 t2)
^(x1 t1) + (t2 ; t1)
^(x2 t2)( @t@
^(x1 t1))
1
The  functions pick out only the point t1 = t2 in the terms multiplying
them, and at this (equal-time) point the
^ elds commute according
to (5.105), leading to the result ((
^_ (x1 t1) is short for @
^(x1 t1)=@t1).
(iv) Dierentiating the result of (iii) with respect to t1 we get
@ 2 fT (
^(x  t )
^(x  t )g = (t ; t )^(x  t )
^(x  t ) + (t ; t )
!^(x  t )
^(x  t )
1 1 2 2 1 2 1 1 2 2 1 2 1 1 2 2
@t21
;(t2 ; t1)
^(x2 t2)^(x1 t1) + (t2 ; t1)
^(x2 t2)
!^(x1 t1):
Again the  functions force t1 to equal t2 in the terms multiplying them,
which become

^(x1 t1)
^(x2 t2)] (t1 ; t2) = ;i(x1 ; x2)(t1 ; t2)
using (5.104) and (E.32). The remaining two terms are just T f
!^(x1 t1)
^(x2 t2)g.
Since the operator ;@ 2=@x21 + m2 doesn't involve t1, it `passes through'
the -functions in the T -product, acting just on
^(x1 t1):
( 2 ! )
@ 2 @
(; @x2 +m )T f
^(x1 t1)
^(x2 t2)g = T (; @x2 + m )
^(x1 t1))
^(x2 t2) :
2 2
1 1
Hence  2 !
@ ; @ 2 + m2 T f
^(x  t )
^(x  t )g
1 1 2 2
@t21 @x21
( 2 ! )
@ @ 2
= ;i(x1 ; x2)(t1 ; t2) + T ( @t2 ; @x2 + m2)
^(x1 t1)
^(x2 t2) :
1 1
The equation of motion for the free KG eld
^ implies that the second
term vanishes.
6.4
Z d3k
0 ^
h0ja^A (pA )
A(x1)j0i = h0ja^A (pA ) (2)3p2E ^aA(k)e ik x1 +^aA(k)eik x1 ]j0i
0 ;  y 

k
q 2 2
with Ek = mA + k . The ^aA(k) term gives zero on j0i. To evaluate
the `^aA^aA' term, we use (6.46):
y

a^A(pA )^aA(k) = (2)33(pA ; k) + a^A(k)^aA(pA )


0 y 0 y 0

leading to
Z
h0ja^A (pA )
A(x1)j0i = h0j (2)d3pk2E eik x1 (2)33(pA ; k)j0i
3
^
0  0

k
= q 1 eipA x1  0


2EA
0

where EA = (m2A + pA 2)1=2:


0 0

6.5
h0jT f
^C(x1)
^C(x2)gj0i = h0j((t1 ; t2)
^C(x1)
^C(x2) + (t2 ; t1)
^C (x2)
^C(x1)j0i
Z d3 k
= h0j (t1 ; t2) (2)3p2! ^aC(k)e ik x1 + a^C(k)eik x1 ]
;  y 

k
Z )
 (2)d3pk2! ^aC(k )e ik x2 + ^aC(k )eik x2 ]j0i + similar term multiplying (t2 ; t1)
3
0
0 0
0 ;  y 0 

k 0

q q
where k0 = !k = m2C + k2 and k0 = !k = m2C + k 2: The conditions
0 0 0

h0ja^C (k) = a^C(k )j0i = 0 reduce the (t1 ; t2) term to


y 0

Z Z
h0j(t1 ; t2) (2)d3pk2! (2)d3pk2! a^C (k)^aC(k )e ik x1 eik x2 j0i:
3 3 0
0
y 0 ;  

k k 0

Using
a^C(k)^aC(k ) = (2)33(k ; k ) + a^C(k )^aC(k)
y 0 0 y 0

this becomes
Z Z
(t1 ; t2) (2)d3pk2! (2)d3pk2! (2)33(k ; k )e ik x1 eik x2
3 3 0
0
0 ;  

k k 0

Z d3 k
= (2)32! (t1 ; t2)e i!k (t1 t2)+ik (x1 x2): ; ;  ;

k
Similarly for the (t2 ; t1) term.
6.6 We write (6.91) as
ZZ
(;ig)2 d4x1d4x2ei(pA 0
; pB ) x1 ei(pB pA ) x2

0
; 
f (x1 ; x2)
where f (x1 ; x2) is given by the RHS of (6.98). Introducing x =
x1 ; x2 X = (x1 + x2)=2, the integral becomes
ZZ
d4xd4X ei(pA pB) (X +x=2) ei(pB pA ) (X x=2) f (x)
0
; 
0
;  ;

since the Jacobian for the change of variables is unity (for instance,
dt1dt2 ! d(t1 ; t2)d(t1 + t2)=2, etc.) The only place X appears is in the
exponent, allowing the X -integral to be done using the 4-dimensional
version of (E.26) we then get
Z
(2)  (pA ; pB + pB ; pA ) d4x ei(pA pB) (pB pA)] x=2 f (x):
4 4 0 0
0
; ;
0
; 
The 4-momentum -function ensures that in the exponent pA ; pB is 0

equal to ;(pB ; pA ), each being equal to q. Re-instating the factor


0

(;ig)2 we obtain
Z
(;ig)2(2)44(pA + pB ; pA ; pB ) d4x eiq x f (x)
0 0 

which is (6.99). Substituting now the RHS of (6.98) for f (x) and per-
forming the x-integration gives a factor (2)24(q ; k), and nally per-
forming the k-integration gives (6.100).
6.7 The analogue of (6.91) for this contraction is
ZZ
(;ig) 2 d4x1d4x2 ei(pA+pB ) x1 e i(pA+pB) x2 h0jT (
^C(x1)
^C(x2))j0i:
0 0
 ; 

Introducing x and X as in problem 6.6, this becomes


Z
(;ig)2(2)44(pA + pB ; pA ; pB) d4x ei(pA+pB ) x f (x):
0 0 

Inserting (6.98) for f (x) and performing the x and then the k integrals
leads to the same expression as (6.100) but with q replaced by pA + pB
- that is, (6.101).
6.8 In the CM frame,
q q
pA = ( mA + k  k) pB = ( m2A + k 2 k )
2 2 0 0 0

and jkj = jk j by energy conservation. Hence


0

u = (pA ; pB )2 = (0 (k ; k ))2 = ;(k ; k )2 0:


0 0 0

6.9 Inq the frame in which B is initially at rest, pB = (mB 0) pA =


( m2A + p2 p) and
q
(pA  pB )2 ; m2Am2B = (mB m2A + p2)2 ; m2Am2B
= m2Bp2:
Hence (pA  pB )2 ; m2Am2B]1=2 = mBjpj. But in this frame mB = EB,
and jpj = EA jvj.
Chapter 7
7.2 (a) The calculation is similar to that in problem 5.7.

Z Z (Z
N^0 _ y _
= i (

^ ;
^
^ ) d x = i d x
y 3 3 k (^a (k)eik x + ^b(k)e ik x)
d3 p y  ; 

(2) 2!
3
Z d3k
 p (;i! ^a(k )e ik x + i! ^b (k )eik x)
0
0 0
0 0 ;  0 y 0 

(2)3 2! 0

Z d3k Z d3k )
; (2)3p2! (^a(k)e + b (k)e ) (2)3p2! (i! ^a (k )eik x ; i! ^b(k )e
^
0
i k x
;  i k
y x  0 y 0
0
 0 0 ik x )
;
0

:
0

After multiplying out the brackets, the terms a^ ^b and a^^b cancel using y y

^a (k) ^b (k )] = ^a(k) ^b(k )] = 0. The term a^ a^ is


y y 0 0 y

Z Z d3k Z d3k
dx p p ! a^ (k)^a(k )ei(! ! )t i(k k ) x:
0
3 0 y 0 ;
0
; ;
0


(2) 2! (2) 2!
3 3 0

The x-integral gives (2)33(k ; k ), which forces ! = ! and allows0 0

the k -integral to be done giving


0

Z d3k 1
(2)3 2 a^ (k)^a(k):
y

The term a^^a gives the same when normally ordered (throwing away a
y

divergent piece) similarly the ^b ^b terms. y

7.3 The commutator is


Z d3k Z d3 k
p ik x1 +^b (k )eik x2 ) p (^a (k )eik x2 +^b(k )e
0

 a
(^ ( k )e ;  y  y 0  0 ; ik x2 )]:


(2)3 2! (2)3 2! 0

The non-vanishing terms involve ^a(k) a^ (k )] and ^b (k) ^b(k )]. The y 0 y 0

rst of these gives


Z d3k Z d3k
p p e ik x1 eik x2 (2)33(k ; k )
0
0
;   0

(2)3 2! (2)3 2! 0
Z 3
= (2d)k32! e ; ik (x1 x2 )
 ;

and the second gives


Z
; (2d)k32! eik (x1 x2):
3
 ;

As long as (x1 ; x2) is spacelike ((x1 ; x2)2 < 0) we can transform


x1 ; x2 into ;(x1 ; x2) by a continuous Lorentz transformation, and
so the second term cancels the rst for (x1 ; x2)2 < 0.
Similar manipulations show that the two terms in the commutator can
be re-written as follows. The rst is
Z d3k
ik (x1 x2 ) = h0j
^(x )
^ (x )j0i
(2)32! e 1 2
;  ; y

which has the interpretation that a particle is created at x2 and de-


stroyed at x1. The second is
Z d3 k
; (2)32! eik (x1 x2) = ;h0j
^ (x2)
^(x1)j0i
 ; y

which corresponds to the creation of an antiparticle at x1 and its de-


struction at x2. Compare the comments following (7.32). Thus for
the commutator to vanish in the spacelike region, the contributions of
these two processes must cancel, and this requires the components of
`k' to be the same in each case - that is, the masses of the particle and
antiparticle must be identical.
7.5 The LHS of the anticommutator (7.44) is
8Z
< k X (^cs(k)u(k s)e i!t+ik x + d^ (k)v(k s)ei!t ik x)
d3p
: (2)3 2! s=12
;  y ; 
s

Z 9
d3 k X =
ik y + d^ (k )v (k  s )e i! t+ik y ) :
p
0

(^cs (k )u (k  s )ei! t
0 0
0 0


y y y
s
0 0 0 ;  0 0 0 ; 

(2)3 2! 0 0

s =12
0
0
From (7.41) the only non-vanishing terms involve fc^s (k) ^cs (k )g and y 0

fd^s (k) d^s (k )g, which give


0

y 0
0

Z d3k Z d3k X X
p p
0

(2)3 2! (2)3 2! s=12 s =12 0


0

u(k s)u (k  s )e i(! ! )t eik x e ik y (2)33(k ; k )ss + similar term in vv ]


0
y 0 y
0 0 ; ;  ;  0
0

Z 3
= (2d)3k2!  u(k s)u (k s)eik (x y ) +
X Xy
v(k s)v (k s)e ik (x y)]: (C )
 ; y ;  ;

s=12 s=12
We need to evaluate Ps u(k s)u (k s) and Ps v(k s)v (k s). Prob-
y y

lem 7.8 below deals with similar expressions and gives some hints. First,
note that `uu ' is a matrix (u u is single number), and uu is its (  )
y y y

element. We can calculate the matrix u(k s)u (k s) straightforwardly y

using (4.105):

s !  
u(k s)u (k s) = (! + m)  k
s
y

s
s !+km 
y y 

0 !+ms s s
s  k
1


= (! + m) @  k s s  k s !s+m k A :
y y 

!+m

!+m

!+m
 y  y 

We now use Ps
s
s = 1 (see problem 7.8) to obtain
y

0 !+km 1
X 1
u(k s)u (k s) = (! + m) @  k k2 A :
y


s !+m

(!+m)2
Since k2 = !2 ; m2 this becomes
 !
!+m k (u)
k !;m
which is in fact the matrix m +   k + ! using (4.31). For the `P vv ' y

term, we rst change the integration


P uu ' term.variable k to ;k so as to have
the
P v(k s)v (k s) we use (4.114) with p replaced by ;that
same exponent as the ` This means y
in evaluating
s
y
k (and E by !),
which gives  !
! ; m ;  k : (v)
;  k ! ; m
The sum of (u) and (v) is just 2! times the unit 4  4 matrix, whose
(  ) element is  . Inserting this into (C ) gives (7.44).
7.6
Z Z
x (2d)3pk 2! X^cs(k)u (k s)eik x + d^s (k)v (k s)e ik x]
3
N^ = d3 y y  y ; 

s
Z d3 k X
 p ^cs (k )u (k  s )e ik x + d^s (k )v (k  s )eik x]:
0
0 y 0
0 y 0 0 ;  0 y 0 0 

(2)3 2!
0 0

s
0
0

Consider the term `^csc^s '. The integral over x gives (2)33(k ; k ) and
y 0

the k -integral can then be done giving the term


0

Z d3 k X
(2)32! ss c^s(k)^cs (k)u (k s)u(k s):
y y
0

Taking
1
2 to be orthogonal (as is conventional), problem 4.12 im-
plies that u (k s)u(k s ) = 2!ss , so that this term is
y 0
0

Z d3 k X
(2)3 s c^s (k)^cs(k):
y

The `d^s d^s ' term is handled similarly, using v (k s)v(k s ) = 2!ss .
y y 0
0

Now consider the `^csd^s ' term: the integral over x yields (2)33(k +
0

y y

k ) so that we need to evaluate u ((! k) s)v((! ;k) s ). It is easily


0

0 y 0

veri ed from (4.105) and (4.114) that this vanishes similarly for the
`d^sc^s ' term.
0

Z Z
H^ D = d3xf^_^ ; L^D g = d3x^ (;i  r + m)^ y

Z Z d3 k X
= dx 3 p  c^ (k)u (k s)eik x + d^s (k)v (k s)e ik x] y y  y ; 

(2)3 2! s s
Z 3k X
 (;i  r + m) (2d)3p  c^s (k )u(k  s )e ik x + d^s (k )v(k  s )eik x]:
0
0 y 0
0 0 0 ;  0 0 0 

2! s
0 0
0
0

Now note that


(;i  r + m)u(k  s )e ik x = (  k + m)u(k  s )e ik x
0 0 ;
0
 0 0 0 ;
0


= ! u(k  s )e ik x: 0 0 0 ;
0

In the `^csc^s ' term, the x-integral can be done, and then the k -integral,
y
0
0

leading to
Z d3k X
(2)32! ss c^s(k)^cs (k)u (k s)u(k s )!:
y y 0
0

Use of u (k s)u(k s ) = 2!ss then gives the rst term in (7.50). For
y 0

the `d^sd^s ' term note that


0

y
0

(;i  r + m)v(k  s )eik x = (;  k + m)v(k  s )eik x


0 0
0
 0 0 0
0


= ;! v(k  s )eik x: 0 0 0
0


Performing the x-integral and then the k -integral leads to the second 0

term of (7.50). The `^csd^s ' and `d^sc^s ' terms vanish as in the calculation
y y

of N^ .
7.8 The solution to problem 7.5 showed that
X
u(k s)u (k s) = ( m +   k + !): y

s
Hence
X
u(k s)u(k s) = ( 2m +   k + !) = (m ;   k + !)
s
= (6 k + m):
7.9
h0jT (^(x1)^(x2))j0i =
Z
k X c^s(k)u(k s)e ik x1 + d^ (k)v(k s)eik x1 ]
d3p
h0j(t1 ; t2) s
;  y 

(2)3 2! s
Z d3k X
 p c^s (k)u (k  s )eik x2 + d^s (k )v (k  s )e ik x2 ]j0i
0

 y 0 0
0
 0 0 0 ;
0


(2) 2! s
3 0 0
0
0

Z
k  c^ (k)u (k s)eik x2 + d^s (k)v (k s)e ik x2 ]
X
;h0j(t2 ; t1) (2d)3p
3
y  ; 

2! s s
Z d3k X
 (2)3p2!  c^s (k)u(k  s )e ik x1 + d^s (k )v(k  s )eik x1 ]j0i
0
0 y 0
0 0 ;  0 0 0 
0 0

s
0
0
where in the exponents k0 = (k2 + m2)1=2 !, and k0 = (k 2 + 0 0

m2)1=2 ! . In the (t1 ; t2) part, the terms in d^s and d^s vanish,
0 y

using d^j0i = 0 = h0jd^ from (7.36) and in the (t2 ; t1) part, the
0

terms in c^s and c^s vanish similarly. In the (t1 ; t2) part we then use
y

fc^s(k) c^s (k )g = (2)33(k ; k )ss to reduce it to


0

y 0 0
0 0

Z 3
(t1 ; t2) (2d)k32! u(k s)u (k s)e ik (x1 x2 ) =
X ;  ;

Z s d3 k
(t1 ; t2) (2)32! (6 k + m) e ik (x1 x2 ):
;  ;

Similarly, the (t2 ; t1) part reduces to


Z d3k X
;(t2 ; t1) (2)32! v(k s)v (k s)e ik (x2 x1 ) =
;  ;

Z s d3k
;(t2 ; t1) (2)32! (6 k ; m) e ik (x2 x1):
;  ;

So altogether we get
Z d3k (t ; t )(6 k + m) e i!(t1 t2 )+ik (x1 x2)
h0jT (^(x1)^(x2))j0i = (2)32! 1 2 
; ;  ;

;(t2 ; t1)(6 k ; m) e i!(t2 t1)+ik (x2 x1)]


; ;  ;

where 6 k = ! ;   k. We need to show that this is equal to


Z d4 k
e ik (x1 x2 ) i(6 k + m) : (T )
k 2 ; m2 + i
;  ;

(2)4
Consider the integral over k0 in (T ). There are poles at k0 = (k2 + m2 ; i)1=2
which is in the lower half k0-plane, and which tends to the real value
! as  ! 0 and at k0 = ;(k2 + m2 ; i)1=2 which is in the upper
half k0-plane, and tends to ;! as  ! 0. On the other hand, the
k0-dependence of the exponential is
e ik0(t1 t2 ):
; ;

We must therefore close the k0-contour (see Appendix F) in the upper


half k0-plane for t1 ; t2 < 0, picking up the contribution from the pole
at k0 = ;! and in the lower half k0-plane for t1 ; t2 > 0, picking up
the contribution of the pole at k0 = !. This gives
Z d3k  ;2i 
(t1 ; t2) (2)4 2! i(! ;   k + m) e i!(t1 t2 )+ik (x1 x2)
; ;  ;

Z d3k  2i 
+(t2 ; t1) i( ;! ;   k + m) i!(t t )+ik (x1 x2 ):
 e 1 2
(2) ;2!
;  ;

In the second integral, change k to ;k to obtain


Z d3 k
(t1 ; t2) (2)32! (! ;   k + m) e i!(t1
; ; k x x2 )
t2 )+i ( 1
 ;

Z d3k
;(t2 ; t1) (2)32! (! ;   k ; m) e i!(t2; ; k x x1 )
t1 )+i ( 2
 ;

as required.
7.10 The Euler-Lagrange equations for A are (compare (5.134))
 !
 @ L
@ @ (@ A ) = @A @L 


where in this case


L = ; 12 F @ A ; jem  A
with F = @A ; @ A. We use a `brute force' approach. Consider
the E-L equations written out in full for one component of A , say A1:
 !  !  !  !
0 @ L 1 @ L 2 @ L 3
@ @ (@ 0A1) +@ @ (@ 1A1) +@ @ (@ 2A1) +@ @ (@ 3A1) = @A @ L @L :
1

We need to identify the terms in L which involve @ 0A1 @ 1A1 @ 2A1 @ 3A1.
First, note that @ 1A1 does not appear since F = 0 when  =  . The
relevant terms are contained in
; 12 (F01@ 0A1 + F10@ 1A0 + F12@ 1A2 + F21@ 2A1 + F13@ 1A3 + F31@ 3A1)
= ; 21 f(@0A1 ; @1A0)@ 0A1 + (@1A0 ; @0A1)@ 1A0 + : : :g
Now use @0A1 = ;@ 0A1 @1A2 = @ 1A2, etc, to pick out just the terms
involving A1, which are
1 (@ 0A1)2 ; 1 (@ 2A1)2 ; 1 (@ 3A1)2
2 2 2
;(@ A )(@ A ) + (@ A )(@ A ) + (@ A )(@ A3): (R)
0 1 1 0 2 1 1 2 3 1 1

We then nd
9
@ (@ 0A1 ) = @ A ; @ A = @1A0 ; @0A1 >
@ 0 1 1 0
=
L

@ (@ 2A1 ) = @ A ; @ A = @1A2 ; @2A1 > (D)


@ L 1 2 2 1
@ L
@ (@ 3A1 ) = @ 1A3 ; @ 3A1 = @1A3 ; @3A1: 
Substituting these expressions into the E-L equations for A1 we obtain
@ 0(@1A0 ; @0A1) + @ 2(@1A2 ; @2A1) + @ 3(@1A3 ; @3A1) = ;jem 1 (S )
or
(@ 0@0 + @ 1@1 + @ 2@2 + @ 3@3)A1 ; @1(@ 0A0 + @ 1A1 + @ 2A2 + @ 3A3) = jem 1
which is the  = 1 component of (7.62) (written with the index 
lowered).
Equations (D) can be written in covariant form as
@L = @ A ; @ A :
@ (@ A )    
Those skilled in manipulation of indices in tensor calculus can derive
this result directly from the covariant Lagrangian (7.62) then follows
trivially.
7.11 (b) The condition (7.90) is
(A(k2)g + B (k2)k k)(;k2g + kk ) = g  :
Multiplying out the brackets,
;k2A(k2)g  + A(k2)k k ; k2B (k2)k k + B (k2)k k2k = g  
giving (7.91).
7.12 Referring to (R) of problem 7.10, the addition of the term ; 12 (@A)2
to L means that we must add to (R) the additional terms
; 21 (@1A1)2 ; (@1A1)(@0A0 + @2A2 + @3A3):
This produces a non-zero value for @ L=@ (@ 1A1), namely
@ L = ;(@ A ) + (@ A ; @ A ; @ A ):
1 1 0 0 2 2 3 3
@ (@ 1A1)
Including then the contribution of @ 1(@ L=@ (@ 1A2)) in (S ) of problem
7.10 (with jem 1 set to zero) leads to the required equation of motion
for A1.
7.13 The (k )'s are given in (7.102)-(7.104), from which we note that,
for each value of , the only non-vanishing component of (k ) is
that in which  = , which equals 1. Consider then the case  = 0 in
(7.115): the  = 1, 2 and 3 terms all vanish, and the only surviving
 -component is  = 0, for which the LHS is -1, verifying the result.
Similarly for the other cases.
7.14 For (7.119): following the same steps as in (7.87)-(7.91), we require
(M 1) where
;

M  = (;k2g + (1 ; 1=)k k ):
Let
(M 1) = A(k2)g + B (k2)k k :
;

The condition g  = M  (M 1 ) gives


;

g  = (;k2g + (1 ; 1=)k k )(A(k2)g + B (k2)k k )


= ;k2A(k2)g  ; k2B (k2)k k + A(k2)(1 ; 1=)k k + B (k2)(1 ; 1=)k k k2:
To match coe"cients of g  on both sides we require A = ;1=k2, and
then the vanishing of the k k term gives B = (1 ; )=(k2 )2.
7.15 The Hamiltonian density is
H^ = H^ S + H^ S
0
where H^ S is the Hamiltonian density for the free complex scalar eld,
H^ S = ^ ^ + r
^  r
^ + m2
^
^
y y y

and H^ S is the interaction part. But also the general de nition of the
0

Hamiltonian is
H^ = ^
^_ + ^
^_ ; L^: y y

Now
^
^ = @ L_ =
^_ ; iq
^ A^0
y y

@
^
from (7.134) - (7.136), and
^ =
^_ + iq
^A^0:
y

Substituting for
^_ and
^_ in favour of ^ and ^ , we obtain
y y

H^ = ^ (^ ; iq
^A^0) + ^ (^ + iq
^ A^0)
y y y

;f(^ + iq
^ A^0)(^ ; iq
^A^0) ; r
^  r
^ ; m2
^
^g ; L^int
y y y y

= ^ ^ + r
^  r
^ + m2
^
^ ; q2
^
^(A^0)2 ; L^int
y y y y

which establishes (7.139).


7.16 Consider rst
@2 h0jT (
^(x1)
^ (x2)j0i =
y

@2 f(t1 ; t2)h0j


^(x1)
^ (x2)j0i + (t2 ; t1)h0j
^ (x2)
^(x1)j0ig
y y

where @2 = @=@x2 (similarly for @1). If  = i, a spatial index,


then the derivative passes through the -functions and we get sim-
ply h0jT (
^(x1)@2i
^ (x2))j0i. If  = 0, then the dierentiation produces
y

extra terms
;(t1 ; t2)h0j
^(x1)
^ (x2)j0i + (t2 ; t1)h0j
^ (x2)
^(x1)j0i
y y

which cancel using (7.32) (at equal times (x1 ; x2)2 is spacelike). So
@2 h0jT (
^(x1)
^ (x2))j0i = h0jT (
^(x1)@2
^ (x2))j0i:
y y
Now consider
@1@2 h0jT (
^(x1)
^ (x2))j0i = @1h0jT (
^(x1)@2
^ (x2))j0i =
y y

@1f(t1 ; t2)h0j
^(x1)@2
^ (x2)j0i + (t2 ; t1)h0j@2
^ (x2)
^(x1)j0ig
y y
:
As before, if  = j the derivative passes through the -functions and
we get h0jT (@1j
^(x1)@2
^ (x2)j0i: On the other hand, if  = 0 we get
y

(t1 ; t2)h0j
^(x1)@2
^ (x2)j0i + (t1 ; t2)h0j@10
^(x1)@2
^ (x2)j0i
y y

;(t1 ; t2)h0j@2
^ (x2)
^(x1)j0i + (t2 ; t1)h0j@2
^ (x2)@10
^(x1)j0i:
y y

= h0jT (@10
^(x1)@2
^ (x2))j0i + h0j
^(x1) @2
^ (x2)]j0i(t1 ; t2):
y y

If  is a spatial index, the equal time commutator vanishes, as can be


seen by dierentiating (7.32) with respect to a spatial component of
x2. But if  = 0 then the equal time commutator is

^(x1)
^_ (x2)](t1 ; t2) = i3(x1 ; x2)(t1 ; t2):
y

These results establish (7.140).


Chapter 8
8.2 We use (8.25), and the mode expansion (7.16) in (8.23), to obtain
 (x)js+  pi = iep4EE h0ja^(p )
^ @ 
^ ; (@ 
^ )
^]^a (p)j0i
hs+  p j^jem
0 0 y y y
s
0

p Z d3k
= ie 4EE h0ja^(p )f p ^a (k )eik x + ^b(k )e ik x]
0
0 0
0 0 y 0  0 ; 

(2) 2!
3 0

Z d3k
 (2)3p2! ;ika^(k)e ik x + ik^b (k)eik x]ga^ (p)j0i
;  y  y

;(@ 
^ )
^ term y

where E =
pM 2 + p2 E = qM 2 + p 2 ! = qM 2 + k2 and ! =
q
0 0 0

M 2 + k 2. We evaluate the
^ @ 
^ term. The a^ and ^b operators com-
0 y

mute with each other, so we can move the ^b(k) all the way to the right
where it will give zero on j0i similarly the ^b (k ) will give zero on h0j. y 0

This term therefore reduces to


p Z d3k Z d3 k
ie 4EE h0ja^(p ) p a^ (k )e p ;ik^a(k)e ik xa^ (p)j0i:
0

0 0 i k x y 0
0
 ;  y

(2)3 2! 0
(2)3 2!
We then use
a^(k)^a (p) = a^ (p)^a(k) + (2)33(k ; p)
y y

and similarly for the product a^(p )^a (k ), together with 0 y 0

a^(k)j0i = h0ja^ (k ) = 0 y 0

to get rid of all the operators. The -functions allow all the momentum
integrals to be performed, setting k = p and k = p so that ! = 0 0

E ! = E  k = p , and k  = p . This term becomes simply


0 0 0 0

ep e i(p p ) x ; ;
0


the (@ 
^ )
^ term supplies the p  contribution to (8.27).
y 0

8.3 We use (8.32) to obtain


 (x)js  pi = iep4EE h0j^b(p )fsame expression for
hs  p j^jem
; 0 ; 0
s
0

the current as in problem 8:2g^b (p)j0i: y

This time, in the


^ @ 
^ term of the current, we can move a^(k ) to the
y 0

right to annihilate on j0i, and ^a (k ) to the left to annihilate on h0j.


y 0

This leaves an expression of the form


h0j^b(p ):::::^b(k )::::^b (k)::::^b (p)j0i:
0 0 y y

But we must remember to normally order the operators in the current:


this means that the above expression should in fact be
h0j^b(p ):::::^b (k)::::^b(k )::::^b (p)j0i:
0 y 0 y

We now use the commutation relations for ^b(p ) and ^b (k) to get ^b (k) 0 y y

anihilating on j0i, and similarly for ^b(k ) and ^b (p) to get ^b(k ) anni- 0 y 0

hilating on j0i. This leaves delta-functions (p ; k ) and (k ; p ), 0 0


allowing the momentum integrals (in this rst term) to be performed,
setting k = p  k = p, and hence ! = E  ! = E k = p and k = p.
0 0 0 0 0 0

The result is
;ep  e i(p p ) x
0 ; ;
0


for this rst term. The second supplies the ;ep part, and the whole
matrix element is indeed the negative of the one in problem 8.2.
8.4 We consider a Lorentz transformation along the x1-axis. Then
jx =  x  =
0 0y 0
 ex#=2 xex#=2
y

=  xex#
y

=  x(cosh # + x sinh #)


y

= cosh # jx + sinh #
as in (4.86), where we have used (4.90), and (4.91) with #=2 replaced
by #.
8.5 (a) We are assuming E = E , as in the required application. Then
0

  0
1 1
u (k  s = 1)u(k s = 1) = (E + m)
1
1 (E+km) @  k
1 A
0
y 0 0 y y 

(E +m)
(  ! )
k  k  
= (E + m) 1 +
(E + m)2 + i (E + m)2
1
1 k  k y
0 0

which reduces immediately to the given expression.


8.5 (b) We have  !  !

= 0  and
= 01 :
1 1 2

Also  !
A A
  A = Ax + iAy ;Az :
z x ; i A y

Then, by straightforward matrix multiplication we obtain, for instance,



1   A
2 = Ax ; iAy :
y
Generally,

i   A
j = (  A)ij :
y

(c) The expression S of (8.46) is


S = 12 fju s =1us=1 j2 + ju s =1us=2j2 + ju s =2us=1j2 + ju s =2us=2j2g:
0y
0
0y
0
0y
0
0y
0

Part (a) evaluated u s =1us=1, and u s =2us=2 is the same but with
1
0y
0
0y
0

replaced by
2. Since
1
2 = 0, we nd
y

u s =1us=2 = (E + m) i
(E +k m)2k

1 2 y 0
0y
0

and similarly for u s =2us=1. The result follows after noting that
0y
0

j
1   A
1j2 + j
1   A
2j2 + j
2   A
1j2 + j
2   A
2j2 = 2A2:
y y y y

(d)
8" #2 9
< =
(E = m)2 : 1 + k cos  2 + (k ) sin 4 
2 2 2 2
S = (E + m) (E + m)
(  
2  E ; m 2 )
= E ; m
(E + m) 1 + E + m cos  + E + m sin 
2 2

= f(E + m) + (E ; m) cos ]2 + (E ; m)2 sin2 g


= (E + m)2 + 2k2 cos  + (E ; m)2
= 2(E 2 + m2 + k2 cos )
= 2E 2 + m2 + k2(1 ; 2 sin2 =2)]
4E 2(1 ; Ek 2 sin2 =2)
2
=
and the result follows using v = jkj=E .
8.6 The manipulations are similar to those in problem 8.2 except that an-
ticommutation relations are used.
8.7   = ( 0  0). So
  = ( 0    0 ) = ( 0  0):
y y y y
Hence
 0   0 = (( 0)3  0( 0)2) = ( 0  0) =   :
y

8.8
Tr(6 k + m) (6 k + m)  ] = Tr(6 k   6 k  ) + mTr(  6 k  )
0 0

+mTr(6 k    ) + m2Tr(    ):
0

The terms linear in m vanish by (8.73) (8.74) gives Tr(   ) = 4g ,


and (8.75) implies that
Tr6 k   6 k  ] = 4k k + k  k ; k  kg ]:
0 0 0 0

These results establish (8.78).


8.9
L00 = 22k 0k0 ; (k  k)] + 2m2:
0 0

Here, k0 = (m2 + k2)1=2 = k 0 E k  k = k0k 0 ; k  k = E 2 ; k2 cos .


0 0 0 0

So
L00 = 2E 2 + k2 cos ] + 2m2
= 2E 2 + m2 + k2 cos ]
and the result follows as in problem 8.5(d).
8.11 Using (8.27), (8.55), and the Fourier transform of (7.119) (compare
(6.98) and (7.58)), the term written explicitly in (8.93) becomes
(;i)2 Z Z d4x d4x fe(p + p ) e i(p p ) x1 0

1 2
0 ; ; 

2
Z 4
 (2dl)4 e il (x1 x2) : i;g + (1l2; )ll =l ] :;eu(k  s )  u(k s)e i(k k ) x2 g:
2 0
;  ; 0 0 ; ; 

The integral over x1 gives (2)44(p + l ; p ), and that over x2 gives 0

(2)44(l + k ; k). The integral over l can now be done using one of
0

these delta functions, which sets l = k ; k = q everywhere, including


0

in the other delta function. This gives one half times (8.95) - the other
half is supplied by the `(x1 $ x2)' part of (8.93).
8.12 Using
@e i(p p ) x = ;i(p ; p )e i(p p ) x
; ;
0
 0 ; ;
0


we nd from (8.27)
@hs+  p j^jem
0  (x)js+  pi = ;ie(p2 ; p 2 )e i(p p ) x
s
0 ; ;
0


which vanishes since p2 = p 2 = M 2. Similarly,


0

@he  k  s j^jem
; 0 0  (x)je  k si = ieu (6 k ; 6 k )ue i(k k ) x
e
; 0 0 ; ;
0


which vanishes since 6 ku = mu and u 6 k = mu : (The particles in the 0 0 0

external states are `on-shell' - i.e. their wavefunctions satisfy the free-
particle equations of motion, or equivalently their 4-momenta satisfy
the energy-momentum condition for a free particle.]
8.13 We have
L T = 2k  k + k  k + (q2=2)g ](p + p )(p + p ) :
0 0 0 0

It is advantageous to write p = p + q, so that (p + p ) = (2p + q),


0 0

and similarly for (p + p ) , and then drop the qq part using (8.188).
0

Then
L T = 82p  k p  k + (q2=2)p2 ] 0

= 82p  k p  k + (q2=2)M 2 ]: 0

8.14 We have (correcting the misprint in (8.134))


Z
q x Z x =a
F (q ) = e (x)d x = e iq x e8a3 d3x
;j j
2 i ; 3  ; 

Z
= 8a3 e i q x cos  e x =a jxj djxj 2d(cos ):
1 ; j jj j ;j j

The integral over cos  can be done with the result


1 ei q x ; e i q x 
ijqjjxj
j jj j ; j jj j

so that F (q2) becomes


Z
F (q2) = 4a31ijqj jxje x =a (ei q x ; e i q x ) djxj:
1
;j j j jj j ; j jj j

0
One way of doing the jxj-integral is to write it as
@ Z e x =a (ei q x ; e i q x ) djxj
1

@ (;1=a) 0
;j j j jj j ; j jj j

where now the simple exponential integrals can be done, and the dif-
ferentiation with respect to a performed, after using
@ = a2 @ :
@ (;1=a) @a
This leads to (8.135).
8.15 After replacing  by k , (8.161) becomes (after correcting the misprints
in it)
;e2 (k   )u(p  s )  ((p6 p++k6 k)2+;mm)2 6 ku(p s)
 0 0 0 0

;e2 (k   )u(p  s )6 k ((p6 p;;k6 k)2++mm)2   u(p s):


0
 0 0 0 0

In the rst term, replace 6 ku(p s) by


(6 k + 6 p ; 6 p)u(p s) = (6 k + 6 p ; m)u(p s) = (6 p + 6 k ; m)u(p s)
0 0 0 0

using the Dirac equation and 4-momentum conservation. Also note


that
(6 p + 6 k + m)(6 p + 6 k ; m) = (p + k)2 ; m2:
The rst term is then
;e2 (k   )u  u:  0 0

In the second term, replace u(p  s )6 k by 0 0

u(p  s )(6 k + 6 p ; 6 p) = u(p  s )(6 k ; 6 p + m) = ;u(p  s )(6 p ; 6 k ; m):


0 0 0 0 0 0 0 0 0 0

The second term becomes


e2 (k   )u  u  0 0

which cancels the rst. Similarly when  (k   ) is replaced by k .  0 0 0


8.16 (a) We work in the massless limit, as in section 8.6.3. We have
X (s) (u) e4 X
Me Me = su ; ;


 ss 0 ss


0 0 0

 (k   )(k   )(k ) (k )u(p  s )  (6 p+6 k)  u(p s)u(p s)  (6 p;6 k ) u(p  s )
 0 0 0 0  0 0 0 0 0

e4 g g Trf  (6 p + 6 k) 6 p  (6 p ; 6 k ) 6 p g
= su  
0 0

e4 Trf  (6 p + 6 k) 6 p (6 p ; 6 k ) 6 p g:
= su  
0 0

We now use
 6 p (6 p ; 6 k ) = ;2(6 p ; 6 k ) 6 p 0 0

from (J.5), and then


  (6 p + 6 k)(6 p ; 6 k ) = 4(p + k)  (p ; k )
0 0

from (J.4), and nally


Tr(6 p6 p ) = 4p  p0 0

from (J.30) to obtain


X (s) e (p + k)  (p ; k ) p  p: 4
Me M(ue) = ;32 su
; ;
 0 0

 ss
0 0

Now
(p + k)  (p ; k ) = p2 ; p  k + k  p ; k  k
0 0 0

= ;p  k + k  p ; k  k 0 0

since p  k = p  k in the massless limit for all external lines (as follows
0 0

from squaring p ; k = p ; k). Hence 0 0

(p + k)  (p ; k ) = k  (p ; p ; k ) = ;k2:
0 0 0

This therefore vanishes when the photon is on-shell (k2 = 0).


8.17 The interference term calculated in problem 8.16 (a) had the value
X (s) (u) e4 (p + k)  (p ; k ) p  p:
Me Me = ;32 su
; ;
 0 0

 ss
0 0

This time (with k2 = ;Q2) we have


(p + k)  (p ; k ) = p2 ; p  k + k  p ; k  k
0 0 0

= ;p  k + k  p + Q2=2 ; k  k 0 0 0 0

using k  p = k  p + Q2=2 (from squaring k + p = k + p ). So


0 0 0 0

(p + k)  (p ; k ) = Q2=2 + k  (p ; p ; k) = Q2=2 ; (k )2 = Q2=2:


0 0 0 0

Hence using p  p = ;t=2, we obtain


0

1 X M(s) M(u) = e4 2tQ2 : 

4  ss e e 0 su
0
; ;

There are two such terms in the product


 (s)  
Me + M(ue) M(se) + M(ue) 
; ; ; ;


which are equal.


8.18 (a) We calculate L Me using (8.185) and (8.190).
L Me = 4k k + k k + (q2=2)g ]2pp + (q2=2)g ]
0 0

= 44k  p k  p + q2k  k + q2p2 + 4(q2=2)2]


0 0

= 44k  p k  p + q2(;q2=2) + q2M 2 + (q2)2]


0

= 82k  p k  p + (q2)2=4 + q2M 2 =2]:


0

In `laboratory' frame with p = (M 0), neglecting the electron mass


so that ! = jkj k ! = jk j k , and using (8.219), the preceding
0 0 0

expression becomes
L Me = 82kk M 2 + 4k2 k 2 sin4 =2 ; 2kk M 2 sin2 =2]
0 0 0

= 82kk M 2(1 ; sin2 =2) + 4k2 k 2 sin4 =2]


0 0

= 16kk M 2 cos2 =2]1 + 2 M kk sin2 =2 tan2 =2]


0
0

2
= 16kk M cos =2]1 ; (q =2M 2 ) tan2 =2]
2 2 2
0
The rst factor is exactly that in (K.24), leading to the `no-structure'
cross section.
(b) Returning to line 3 of part (a):
L Me = 44k  p k  p + q2k  k + q2p2 + 4(q2=2)2]:
0 0

With all particles massless, s = 2k  p t = ;2k  k , and u = ;2p  k .


0 0

Hence our expression becomes


4;us ; t2=2 + t2] = 4;us + t2=2]:
d =dt is then given by
d = 1 (4 )2 4; u + t2 ]:
dt 16 t2 s 2s2
In the massless case we have s + t + u = 0 so that t2 = (u + s)2, leading
to the required expression for d =dt.
From (K.41), (K.46) and (K.50) we nd
k 2 2 dt :
0

dy = 2kM 2k 2 0

But in this frame s = (k + p)2 = 2kM . Hence dt = sdy and the result
is established.
8.19 (a) We shall label the 4-momentum and spin of the ingoing e by k s, ;

of the ingoing e+ by k1 s1, of the outgoing  by p  r , and of the


; 0 0

outgoing + by p1 r1. The Mandelstam variables are


Q2 = (k+k1)2 = (p +p1)2 t = (k1;p1)2 = (k;p )2 u = (k;p1)2 = (k1;p )2:
0 0 0

The amplitude is
iev(k1 s1) u(k s) ;Qig2 ieu(p  r ) v(p1 r1):
0 0

Note that the v  u factor depends only on the e and e+ momenta,
;

while the u v factor depends only on the momenta of the  and + . ;
(b) The spin-averaged squared cross matrix element is then
 2 !X
jMj2 = 4 Qe 2 v(k1 s1) u(k s)u(k s) v(k1 s1)
1
ss1
X
 u(p  r ) v(p1 r1)v(p1 r1)  u(p  r )
0 0 0 0

r r1
0

= (4 ) L(e) L()


2

Q 4
where
L(e) = 12 Tr(6 k1 ; m) (6 k + m) ]
and
L() = 21 Tr(6 p + M )  (6 p1 ; M )  ]
0

using (7.61).
(c) Using the trace theorems as in (8.78), the lepton tensors are
L(e) = 2k1 k + k1 k ; (k1  k)g ]
and
L() = 2p p1 + p  p1 ; (p  p1)g ]
0 0 0

where now (in the massless limit) Q2=2 = p  p1 = k1  k. Hence


0

 2!
jMj = Qe 2 42p  k1 p1  k + 2p  k p1  k1 ; Q2p  p1 ; Q2k1  k + (Q2)2]
2 0 0 0

 2! 2
= Qe 2 42 u4 + 2 t4 ; Q2Q2=2 ; Q2Q2=2 + (Q2)2]
2

 2!
= Qe 2 2t2 + u2]:

In the massless limit, all 3-momenta have equal modulus which (slightly
confusingly) we denote by k. Then
t = ;2k2(1 ; cos ) u = ;2k2(1 + cos )
so that
t2 + u2 = 8k4 (1 + cos2 ):
Since Q2 = 4k2 the result follows.
Crossing symmetry implies that the amplitude for
e (k s) + e+(k1 s1) !  (p  r ) + + (p1 r1)
; ; 0 0

is equal to (minus) the amplitude for


e (k s) +  (;p1 ;r1) !  (p  r ) + e (;k1 ;s1):
; ; ; 0 0 ;

We can therefore obtain this amplitude from the one calculated in sec-
tion 8.7 by making the replacements
p r ! ;p1 ;r1 and k  s ! ;k1 ;s1:
0 0

The spin labels disappear at the Trace stage, of course, and


q2(section 8:7) = (k ; k )2 ! (k + k1)2 = Q2:
0

So (8.185) and (8.186) become


(8:185) ! 2;k1k ; k1 k + (Q2=2)g ]
and
(8:186) ! 2;p p ; p  p + (Q2=2)g ]
0 0

leading to exactly the same result for jMj2.


(d) Using the formula (6.129) for the dierential cross section for elastic
scattering in the centre of mass, and replacing jMj2 by jMj2, we
obtain
Z d Z
= d# = 6412Q2 162 2(1 + cos2 ) 2 d cos 
 3  !1
2
= 4Q2 2 cos  + 3 cos
1
;

= 4 =3Q2 2

as required.
8.20 We have
 1     ;     
u(p )i 2M q u(p) = iu(p ) 2 i
0 0

2M (p ; p )u(p) 0

= ; 4M1 u(p )( 6 p ;  6 p ; 6 p   + 6 p )u(p)


0 0 0

= ; 4M1 u(p ) 6 p + 6 p  ]u(p) + 1 u(p ) u(p):


0 0 0

2
Now
u(p )  6 p u(p) = u(p )   pu(p)
0 0 0 0

= u(p )(2g ;   )p u(p)


0 0

= u(p )2p  u(p) ; u(p )6 p  u(p)


0 0 0 0

= u(p )2p  u(p) ; M u(p )  u(p)


0 0 0

and similarly
u(p )6 p u(p) = u(p )2p u(p) ; M u(p ) u(p):
0 0 0

These formulae establish the required result.


Chapter 9
9.1 Using (8.206) with F1 = 1 and  = 0 for the current matrix elements,
and cancelling a factor of e2, we obtain
Wel = 8M 1 X u(p s)  u(p  s )u(p  s )  u(p s)(2)44(p + q ; p ) 1 d3p
0 0 0 0 0
0

ss 0 (2)3 2E 0

= 8M 1 Trf (6 p + M )  (6 p + M )g(2)44(p + q ; p ) 1 d3p


0 0
0

(2)3 2E 0

where the Trace is 2M  (see (8.186)). Equation (9.104) then becomes


 !2
d = 4q 1  (2 )4  4(p+q ;p ) d p d3k :
3 0 0

2 4(k  p)2 ; m2M 2]1=2 L  M 2E (2)3 2! (2)3


0
0 0

This is precisely the formula which which yields the cross section for
elastic e scattering (see Appendix K for the $ux and phase space
factors).
9.2 (a) Omitting the terms involving q and q from the expression (9.10)
for W  , we need to evaluate
L W  = 2k k + k  k + (q2=2)g ];g W1 + (p p =M 2 )W2]
0 0

= 2(;2k  k ; 2q2)W1 + (2p  k p  k + q2p2=2)W2=M 2 ]:


0 0

In the `laboratory' system, and neglecting the electron mass (compare


(8.217) and (8.218)),
p  k = ! M p  k = !M q2 = ;2k  k 
0 0 0

and so
(2p  k p  k + q2p2=2)=M 2 = 2!! + q2=2 = 2!! ; k  k :
0 0 0 0

Writing as usual k = ! = jkj k = ! = jk j, we have 0 0 0

k  k = kk (1 ; cos ) = 2kk sin2 =2


0 0 0

and
2!! ; k  k = kk (1 + cos ) = 2kk cos2 =2:
0 0 0 0

Hence
L W  = 4kk 2W1 sin2 =2 + W2 cos2 =2]
0

and, from (9.104),


 !2
d = 4q 1 2 =2+W cos2 =2] d k :
3 0

2 4(k  p)2 ; m2M 2]1=2 4 M 4 kk 2 W 1 sin 2


2! (2)3
0
0

Now (q2)2 = 16k2 k 2 sin4 =2, and (neglecting the electron mass)
0

(k  p)2 ; m2M 2]1=2 = kM . Also,


d3k = k 2dk d#:
0 0 0

Hence
d = 2 2W sin2 =2 + W cos2 =2]dk d#
4k2 sin4 =2 1 2
0

as required.
(b) We have
Q2 = 2kk (1 ; cos ) and  = k ; k :
0 0

Hence
d cos  dk =  @Q2 1 @Q2  dQ2 d
0

 @ cos  @k 
 @ cos  @k 
0
@ @
0

=  1  dQ2 d = 1 dQ2 d:


 2kk 2k(1 ; cos ) 
0
2kk0

 0 1 
Using d# = 2d cos  and the result of part (a), this leads to (9.16)
from (9.12).
Similarly, since
x = Q2=2M and y = =k
we have
dQ2d =  @x 1 @x  dx dy
 @Q@y2 @ 
@y 
 @Q2 @
=  1 1 Q2  = 2Mk dx dy
 2M ; 2M2 
 0 k
1 
= 2Mk2y dx dy
as in (9.19), leading straightforwardly to the formula for d2 =dxdy.
9.3 (a) The transverse polarization vectors are given in (9.38). These satisfy
( = 1)  p = 0 in the laboratory frame, and also (9.40). Hence in
the product  W  , with W  given by (9.10), only the contraction


with ;g W1 survives, leading to


1 X  () ()W  = 1 ;( = 1) ( = 1) ; ( = ;1) ( = ;1)]W
2 = 1   1
  

2  


= 12 1 + 1]W1 = W1
and (9.46) follows from (9.45).
(b) From (9.47) the longitudinal/scalar virtual photon cross section is
S = (42 =K ) ( = 0) ( = 0)W 


where W  given by (9.10), and where ( = 0) is real and given by


(9.41), and satis es q   = 0 (see (9.49)). Thus in the contractions
with W  , terms involving q and q can be dropped. The W1 term in
`   W ' is then simply (;  )W1 = ;W1 (note (9.42)), while the W2
term is
1 ( = 0)  p ( = 0)  p W = 1 (q3M )(q3M )W
2 2
M2 M 2Q2 
( q 3)2 Q 2 + (q 0)2  2!
= Q2 W2 = Q2 W2 = 1 + Q2 W2
and (9.48) follows from these results.
From (9.46) we obtain
W1 = 4K2 T
and substituting this into (9.48) gives the required result for W2.
9.4 In the limit m ! 0 the spinors
and  satisfy   p
= E
and
  p = ;E respectively, where in both cases E = jpj: Hence

satis es   p
=

jpj
which shows it has positive helicity (compare (4.67)) similarly  has
negative helicity.
(b) For example,
 1 + 5   1 ; 5  1
PR PL = 2 2 = 4 1 ;  2 ] = 0:
5

When m 6= 0, the operators PR and PL still project out the


and 
components of the 4-component spinor, but these 2-component objects
are no longer (with m 6= 0) helicity eigenstates (since, for example,
(  p=jpj)
is no longer equal to
or ;
).
(c) We may write
u u = u (PR + PL) 0 (PR + PL)u:
y

We exploit the fundamental relation  5 = ;5  (see (J.11)). Con-


sider one `cross' term:
u PR  0  PLu = u  0PL  PL u
y y

= u  0 PR PLu
y

= 0:
Similarly for the term u PL 0 PR u. The only surviving terms are
y

u PR 0 PR u + u PL 0 PL u


y y

which is just uR u + uL u. Hence `R' states connect only to `R'
states, and similarly for `L' states, and so helicity (in the massless
limit) is conserved.
Note that `uR' could perhaps more clearly be written as
uR
since we form it by taking the dagger of uR and then multiplying by
 0 - i.e. we take the Dirac `bar' of uR. uR is however the conventional
notation.
(d) In this case a typical cross term is
u PR  0  5PLu =
y
u  0PL  5PL u
y

= u  0 PR 5PLu
y

= u  0 5PR PLu
y

= 0
and again helicity is conserved.
(e) The Dirac mass term is
^^ = ^ (PR + PL ) 0(PR + PL ):
y ^
Consider a `diagonal' term:
y
^ PR 0PR^ = ^  0PL PR^ = 0
y

and similarly for the other diagonal term. Only the `L-R' and `R-L'
terms survive (the daggers in the printed answer are a misprint).
9.5 Neglecting the positron and proton masses, their 4-momenta are pe+ =
(k 0 0 ;k), say, and pp = (p 0 0 p). Then
WCM2 = (p + + p )2 = (k + p)2 ; (k ; p)2 = 4kp:
e p
p
So WCM = 2 kp = 300:3 GeV.
A leptoquark of mass Mlq formed as a resonance state of the e+ and the
struck quark would appear as a peak in the eective mass of the e+ and
the quark, at an eective mass q equal to M2lq. In a simple parton model
picture, this efective mass is (pe+ + xpp) . So we expect a peak when
p2e+ + 2x pe+  pp + x2p2p = Mlq2 
or, neglecting the positron and proton masses, at x = Mlq2 =WCM
2 .

9.6 (a) Using (9.92) for d2 =dx1 dx2, we have


d = Z dx dx 4 2 X e2q (x )q (x ) + q (x )q (x )](q2 ; sx x )
dq2 1 2 2
9q a a a 1 a 2 a 1 a 2 1 2
 2Z X
= 9q2 dx1 dx2 e2aqa(x1)qa(x2) + qa(x1)qa(x2)] 1s (q2=s ; x1x2)
4
a
with the help of (E.29), and then writing 1=s = x1x2=q2 and  = q2=s
we obtain the desired formula.
(b)
 @q2 @q2 
dq2 dxF =  @x @x1 @x2  dx1 dx2
F @xF 
 @x1 @x2 
=  sx1 2 sx 1  dx dx
;1  1 2
= ;s(x1 + x2) dx1 dx2:
The minus sign can be absorbed by appropriate choice of limits in the
q2 ; xF integration. In the variables (x1 x2), the integration is over the
square 0 x1 1 0 x2 1. Consider performing the integration
holding x1 xed and integrating over x2, and then integrating over x1.
Take x1 = 1=2 as an example, with x2 running from x2 = 0 to x2 = 1.
In the q2 ; xF plane, this line maps into the line 2q2=s + xF = 1=2, and
it is traversed in the sense of q2=s increasing (from 0 to 1/2) but xF
decreasing (from 1/2 to -1/2). We can reverse the sense in which xF is
covered by invoking the minus sign from the determinant.
The variables x1 and x2 are given in terms of xF and  by x1 ; x2 = xF
and x2 = =x1. So we have
x1 ; =x1 = xF:
Solving for x1 (which is greater than 0) we nd
x1 = 12 xF + (x2F + 4 )1=2]
and hence
x2 = 12 ;xF + (x2F + 4 )1=2]
so that x1 + x2 = (x2F +  )1=2. Hence
2X
d2 = 4 2 e2aqa(x1)qa(x2) + qa(x1)qa(x2)] dx1 dx2
9q a
= (x +1 x )s f: : :g dq2 dxF
1 2
2 X
= (x2 +14 )1=2 4 9 q 4 e2aqa(x1)qa(x2) + qa(x1)qa(x2)] dq2 dxF
F a
which leads to the desired expression.
9.7 Let the 4-momenta of the incoming q and q be k and k1 respectively,
and let those of the outgoing  and + be p and p1. Then the q ;
; 0

q ;  vertex, for scalar quarks, is proportional to (k ; k1), while the


 ;  ; + vertex is same as in problem 8.19(b). Thus in evaluating
;

the unpolarized cross section we need the contraction


T = (k ; k1)(k ; k1) (p p1 + p p1 ; (Q2=2)g )
0 0

= 2p  (k ; k1) p1  (k ; k1) ; (Q2=2)(k ; k1)2:


0
Introduce the Mandelstam variables
s = (k+k1 )2 = Q2 t = (k;p )2 = (k1;p1)2 and u = (k;p1)2 = (k1;p )2:
0 0

Then, negelecting lepton masses,


T = 2; 2t + u2 ]; u2 + 2t ] ; (Q2=2)(;Q2)
= ; 12 (t ; u)2 + 21 (Q2)2
= ; 12 (4k2 cos )2 + 21 (4k2)2
/ (1 ; cos2 )
where k is the CM momentum.
Chapter 10
10.1 We need to calculate the Jacobian for the change of variables
fx1 x2 x3 x4g ! fx y z X g: (A)
Each of these variables is in fact a 4-vector, with (therefore) four com-
ponents, as is explicit in the notation d4x1 etc. Thus there are a total
of 16 variables altogether, and we certainly don't want to deal with a
16  16 determinant. We can however write the 16-dimensional inte-
gration element as
dx01 dx02 dx03 dx04] dx11 dx12 dx13 dx14] dx21 etc]
and imagine doing the transformation rst for the ` 0 ' components of
the two sets in (A):
fx01 x02 x03 x04g ! fx0 y0 z0 X 0g
then for the ` 1 ' components, etc. The total Jacobian will then be the
product of four 4  4 Jacobians, one for each of the components. But
since they all have exactly the same form, we shall suppress the explicit
`component' label, and write simply
 @x @x @x @x 
 @x@y1 @x@y2 @x@y3 @x@y4 
 
J =  @x@z1 @x@z2 @x@z3 @x@z4 
 @x1 @x2 @x3 @x4 
 @x
@X @X @X @X 
1 @x2 @x3 @x4
for one of the determinants. Evaluating the partial derivatives, we
obtain
 1 0 ;1 0 
 
0 1
J =  0 0 1 ;1 
 0 ; 1
 
1=4 1=4 1=4 1=4 
8   9
>
<  1 0 ;1   0 1 ; 1 >
=
= 14 > 0 1 ;1  ;  0 0 ;1 >
: 1 1 1   1 1 1 
= 1:
Hence the full Jacobian is unity.
Inverting the relations giving fx y z X g in terms of fx1 x2 x3 x4g,
we nd
x1 = 14 (4X + 3x ; y + 2z)
x2 = 14 (4X ; x + 3y ; 2z)
x3 = 14 (4X ; x ; y + 2z)
x4 = 14 (4X ; x ; y ; 2z):
The exponentials in (10.3) then become
p pB p pA
ei( A 4 ) (4X +3x y+2z) ei( B 4 ) (4X x+3y 2z)
0 0
; ;
 ;  ; ;

which reduces to the exponentials given in (10.4). The propagator


factors only depend on the dierence in the arguments of the two elds
(see (6.98)).
10.2 The integral is
Z1 " #1
1 1 1
= ; (B ; A) A + x(B ; A)
0 A + x(B ; A)]2 0
1 1 1
1
= A ; B B ; A = AB :

10.3 The : : :] bracket is


k2 ; xk2 ; (1 ; x)m2A + xq2 ; 2x q  k + xk2 ; xm2B + i]
= k2 ; 2x q  k + xq2 ; xm2B ; (1 ; x)m2A + i]
= (k ; xq)2 ; x2q2 + xq2 ; xm2B ; (1 ; x)m2A + i]
= k 2 ; f;x(1 ; x)q2 + xm2B + (1 ; x)m2Ag + i]
0

which is just k 2 ;  + i]. Note that we have replaced `(1 ; x)i + i'
0

by `i' since all that matters is the sign of this in nitesimal imaginary
part, and (1 ; x) can never be negative.
10.4 We have
%2] 2 ;g2 Z 1 Z  u2du :
C (q ) = 82 0 0 (u2 + )3=2
One way of evaluating the integral over u is to substitute u = 1=2 tan .
Then
du = 1=2 sec2  d
and
(u2 + )3=2 = 3=2 sec3 :
The integral over u becomes
Z tan2  Z sin2 
sec2  d = d
sec3  cos
Z 1  
= cos  ; cos  d
 2 + tan1=2) ; sin  !
= ln(sec
= ln (u +1) u ; &
 = 2 +  1 = 2 (u + )1=2 :
2

This expression vanishes at u = 0 and has the value


(  2 + )1=2 ! )
ln & + (& &
; (& + )1=2
1=2
at u = & (correcting the formula given in (10.51)).
10.5 %2]
C (q ) is given by (10.42) and its dependence on q is contained in the
2 2
quantity  of (10.43). Hence
d%2] Z Z d4k ;2
C (q ) = ig 2 1 dx d
2 0

dq 2 0 (2) (k ;  + i) dq2


4 2 0 3
Z1 Z 4
= ig2 dx (2d k)4 (k 2 ; 2 + i)3 x(1 ; x):
0

0 0

The denominator of the k integral now behaves like (k )6 at large values


0 0

of k , which will ensure convergence. In more detail, we now have, in


0

place of (10.44), a k 0 integral of the form


0

Z dk 00
1 @2 Z dk 0 0

0
=
(k 0)2 ; A]3 2 @A2 (k 0)2 ; A]
0
which will result in a denominator (u2 + )5=2 in (10.50), so that the
u-integral is manifestly convergent, by counting powers of u at large
values of u.
10.6 Consider the counter term
Lct = 12 ZC @
^phC@ 
^phC
where normal-ordering is understood. Remembering that `L = T ; V ',
the corresponding interaction potential is ;Lct, and in the Dyson-Wick
expansion it is `;i' times the interaction that appears. Hence the one-C
matrix element required is
i Z hC kj@
^ @ 
^ jC ki
 phC phC
2 C
Z d3k1
= i 2EZC h0ja^C (k) p ^a (k )(;ik1)e ik1 x+^aC(k1)(ik1)eik1 x]
y

(2)3 2E1 C 1
;  

2
Z
 (2d)3pk22E ^aC(k2)(;ik2)e ik2 x + ^aC(k2)(ik2)eik2 x] a^C(k)j0i
3
;  y  y

2
where normal-ordering is understood but not shown explicitly. The
only terms which give a non-zero vev are those in which the number of
a^ operators is equal to the number of ^a operators. One of these is
y

h0ja^C(k)(ik1)^aC(k1)(;ik2)^aC(k2)^aC(k)j0i:
y y

As usual, we write
^aC(k)^aC(k1) = a^C(k1)^aC(k) + (2)33(k ; k1)
y y

the ^aC(k1) then annihilating on h0j, and similarly with a^C(k2)^aCj0i


y y

which yields a factor (2)33(k ; k2). The k1 and k2 integrations can


then be done, setting k1 = k and k2 = k, and hence E1 = E , and
E2 = E , and k1 = k k2 = k. This term therefore nally yields
i Z k2:
2 C
The other surviving term is (after normal-ordering)
h0ja^C(k)(ik2)^aC(k2)(;ik1)^aC(k1)^aC(k)j0i:
y y

This contributes exactly the same result, removing the factor 12 as re-
quired. The 1-C matrix element of the counter term involving
^2phC is
very similar, lacking the k2 factor coming from the gradients.
Chapter 11
11.1 The Feynman rule for the counter term in the fermion propagator is
given by (a) of (11.7). The fermion analogue of (10.63) is then
S = 6 p ; m + (Z ; 1)i6 p ; m ; 2](p) :
2
For 6 p  m we expand 2](p) as
d 2] 
 (p)   (6 p = m) + (6 p ; m) d6 p  
2] 2]
p=m 6

so that
S  i 
6 p ; m + (Z2 ; 1)6 p ; m ; 2](6 p = m) ; (6 p ; m) d
dp2] p=m
6 6

= 
d
2]  d
2] 
i 
6 pZ2 ; dp p=m ] + m dp p=m ;  (6 p = m) ; m ; m
6 6 6 6
2]

where Z2 and m must be chosen so that this expression is equal to


i=(6 p ; m). The coe"cient of 6 p therefore yields
d 2] 
Z2 = 1 + d6 p  
p=m 6

while the mass terms may be written as


d 2] 
m d6 p  ; 2](6 p = m) ; m ; m = m(Z2 ; 1) ; 2](6 p = m) ; (m0Z2 ; m) ; m
6 p= m
= ;m + (m ; m0)Z2 ; 2](6 p = m)
whence we require
m0 ; m = ;Z2 12](6 p = m):
;

When these values are substituted back into the rst expression for S ,
we obtain (11.10).
11.2 Consider QED for de niteness, with interaction Lagrangian
L^int(x) = ;q^(x) ^(x):
The amplitude for a typical closed fermion loop will involve (compare
(10.3)) a sequence of fermion propagators starting from a certain space-
time point and ending at the same point:
h0jT (^(x1)^(x2))j0ih0jT (^(x2)^(x3))j0i : : : h0jT (^(xN )^(x1))j0i: (A)
Note that each propagator involves elds from dierent interaction La-
grangians in a term of given order in the Dyson expansion analogous to
(6.42), since each interaction involves two elds at the same point. In
a given `Dyson' term, however, the interactions may be written in any
order within the overall time-ordering symbol T , for example (referring
to (A)) as
h0j : : : T f: : : ^(x1)A6 (x1)^(x1)]^(x2)A6 (x2)^(x2)] : : : ^(xN )A6 (x1)^(xN )]g : : : j0i
without introducing any sign changes. The product of contractions
shown in (A) is now obtained after permuting ^(x1) through an odd
number of fermion elds (namely the single eld ^(x1) and the appro-
priate number of pairs ^(x2)^(x2) : : :  ^(xN )^(xN )), which produces
an overall minus sign.
11.3 The amplitude arises from the second-order term in the Dyson expan-
sion, which involves the T -product
T (^(x1)  ^ (x )  
1 ^ (x2) ^ (x2 ))


where we have indicated the (summed over) Dirac matrix indices   


explicitly. Figure 11.2(b) corresponds to the contraction
 h0jT (^ (x )
: : :   
1 ^ (x2))j0i h0jT (^ (x2 )^ (x1))j0i : : :

which has the structure
 S (x ; x )  S (x ; x )
  1 2   2 1
X
=  S (x1 ; x2)  S (x2 ; x1)]

which exhibits the required Trace.
11.4 The result is (up to a factor of 2) the same as (8.78) with k = k + q.
0

11.5 We have, from (11.32),


% 2]
 (q ) = % (q ) ; % (0)
2 2] 2 2]

where from (11.23)


Z
%2] ( q 2) = 8ie2 1 dx x(1 ; x)I (x q 2)
 0
with Z d4k
I (x q2) =
0
1 
(2) (k ;  (x q2) + i)2
4 2 0

where
 (x q2) = ;x(1 ; x)q2 + m2:
As noted in the text, the k -integral in I (x q2) has been evaluated in
0

section 10.3.1 and 10.3.2, with the result (correcting (10.51))


(  2 +  (x q 2))1=2 ! )
i
I (x q ) = 82 ln
2 & + (& &
; (&2 +  (x q2))1=2 
1=2(x q2) 

where & is the ultraviolet cut o. The quantity % 2]  (q ) involves the
2
x-integral of the subtracted quantity
I (x q2) ; I (x q2 = 0)
and we shall see that this has a well-de ned nite limit as & ! 1. We
have
(  2 +  (x q 2))1=2 1=2(x q 2 = 0) !
2 2 i & + (&
I (x q );I (x q = 0) = 82 ln & + (&2 +  (x q2 = 0))1=2  1=2 2
  (x q )
)
& &
; (&2 +  (x q2))1=2 + (&2 +  (x q2 = 0))1=2 :
 
Now
lim & + (&2 +  (x q2))1=2 = 1
 & + (&2 +  (x q2 = 0))1=2
!1

and
; & &

lim
!1 (& +  (x q ))
2 2 1 = 2 + (& +  (x q2 = 0))1=2 = 0:
2

Noting also that  (x q2 = 0) = m2, we obtain


2 Z1 " 2 #
%  (q ) = ; 22
2] 2 e m
dx x(1 ; x) ln  (x q2)
0 
which is equivalent to (11.36).
11.7 For q2 = ;q2  m2 we write
Z
% 2] ( q 2 ) = 2 1 dx x(1 ; x) ln1 + q 2x(1 ; x)=m2 ]
  Z0
2 1 q 2
  0 dx x(1 ; x) m2 x(1 ; x)
q2 Z 1 dx x2(1 ; x)2
= 2m 2 0
which leads to (11.38).
11.8 In this case we are interested in q2 = ;jq2j  m2. We write
Z1  2 !
2] 2 2 jq j
%  (jq j) =  dx x(1 ; x) ln m2 x(1 ; x) + 1
0
Z ( 2 )
2 1 j q j m 2
=  0 dx x(1 ; x) ln m2 x(1 ; x)1 + jq2jx(1 ; x) ]
Z1 (  2!  2 !)
2 j q j
=  0 dx x(1 ; x) ln m2 + lnx(1 ; x)] + O jmq2j :
Then using  2
Z1 x x 2 !1
x ln x dx = 2 ln x ; 4 = ; 14
0 0
and Z1  !1
ln x dx = x3 ln x ; x3 = ; 91 
3 3
0
x2
0
together with
Z1 Z1
x(1 ; x) ln(1 ; x) dx = x(1 ; x) ln xdx
0 0
we arrive at (11.54).
11.9 The e contributes 0.01639, the  0.00825 and the  0.00388 for a total
of 0.0285.
11.10 In the system h = c = 1, with one independent dimension which we
take to be mass (see Appendix B), the electromagnetic eld strength
tensor F^  has dimension M 2, so that the Maxwell action
Z
; 14 d4xF^  F^
is dimensionless. Hence (F^  F^ )2 has dimension M 8, and therefore
the coupling constant `E ' must have dimension M 4 so that ;

Z
E d4x (F^  F^ )2
is dimensionless.
In terms of the elds E^ and B^ ,
F^  F^ = 2(B^ 2 ; E^ 2):
Hence
E (F^  F^ )2 = 4E (B^ 2 ; E^ 2)2:
This fourth-order (in the elds) interaction will contribute to  ; 
scattering, which does not occur in the classical (Maxwell) theory. See
for example The Quantum Theory of Fields, volume 1, by S. Weinberg
(CUP), pages 533-4.

Potrebbero piacerti anche