Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative...

61
4 Perturbative Evaluation of the Path Integral: λϕ 4 Theory In the following, we proceed with the conventional (perturbative) evaluation of the Green’s functions in Euclidean space. We start from W E [J ]= e -Z E [J ] = N Z Dϕe - R d 4 ¯ x[ 1 2 ¯ μϕ ¯ μ ϕ+ 1 2 m 2 ϕ 2 +V (ϕ)-] , (4.1) where N is an arbitrary (infinite) normalization constant. The connected Green’s functions are given by G (N ) E x 1 , ··· , ¯ x N )= - δ N Z E [J ] δJ 1 ··· δJ N J =0 . (4.2) They will be calculated by perturbing in the potential V . For simplicity in the following, we neglect the subscript E and the bar over x, which indicate Euclidean space. Later when confusion with Minkowski space might occur, they will be reinstated. Using the trick of Section 3.3, we obtain W [J ]= Ne -hV ( δ δJ )i e -Z 0 [J ] , (4.3) where Z 0 [J ]= - 1 2 hJ (xF (x - y)J (y)i xy , (4.4) and Δ F (x - y)= Z d 4 p (2π) 4 e ip(x-y) p 2 + m 2 . (4.5) A little algebraic rearrangement yields 1

Transcript of Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative...

Page 1: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4

Perturbative Evaluation of the Path Integral: λϕ4

Theory

In the following, we proceed with the conventional (perturbative) evaluation

of the Green’s functions in Euclidean space. We start from

WE [J ] = e−ZE [J ] = N

∫Dϕe−

∫d4x[ 1

2∂µϕ∂µϕ+ 1

2m2ϕ2+V (ϕ)−Jϕ] , (4.1)

where N is an arbitrary (infinite) normalization constant. The connected

Green’s functions are given by

G(N)E (x1, · · · , xN ) = − δNZE [J ]

δJ1 · · · δJN

∣∣∣∣J=0

. (4.2)

They will be calculated by perturbing in the potential V . For simplicity in

the following, we neglect the subscript E and the bar over x, which indicate

Euclidean space. Later when confusion with Minkowski space might occur,

they will be reinstated. Using the trick of Section 3.3, we obtain

W [J ] = N e−〈V ( δδJ )〉 e−Z0[J ] , (4.3)

where

Z0[J ] = −1

2〈J(x)∆F (x− y)J(y)〉xy , (4.4)

and

∆F (x− y) =

∫d4p

(2π)4

eip(x−y)

p2 +m2. (4.5)

A little algebraic rearrangement yields

1

Page 2: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

2 Perturbative Evaluation of the Path Integral: λϕ4 Theory

Z[J ] = − lnN + Z0[J ]− ln(

1 + eZ0

(e−〈V ( δ

δJ )〉 − 1)e−Z0

), (4.6)

which is ready for a perturbative expansion in the potential V . If we let

δ ≡ eZ0

(e−〈V ( δ

δJ )〉 − 1)e−Z0 , (4.7)

we arrive at

Z[J ] = − lnN + Z0[J ]− δ[J ] +1

2δ2[J ]− 1

3δ3[J ] + · · · . (4.8)

In particular, for V = λ4!ϕ

4, we can expand in powers of the dimensionless

(in four dimensions) coupling constant λ. Setting

δ = λδ1 + λ2δ2 + · · · (4.9)

we find

Z[J ] = − lnN + Z0[J ]− λδ1[J ]− λ2

(δ2[J ]− 1

2δ2

1 [J ]

)(4.10)

−λ3

(δ3[J ]− δ1[J ]δ2[J ] +

1

3δ3

1 [J ]

)+ · · · . (4.11)

From expanding the exponential in (4.1.7) we find

δ1[J ] = − 1

4!eZ0[J ]〈 δ

4

δJ4〉 e−Z0[J ] (4.12)

δ2[J ] =1

2(4!)2eZ0[J ]〈 δ

4

δJ41

〉1〈δ4

δJ42

〉2 e−Z0[J ] , etc. · · · . (4.13)

Using the explicit form (4.1.4) for Z0, we arrive at

δ1[J ] = − 1

4!

[〈∆xa∆xb∆xc∆xdJaJbJcJd〉+ 6 〈∆xx∆xa∆xbJaJb〉+ 3

⟨∆2xx

⟩](4.14)

where all variables x, a, b, c, d, are integrated over in the relevant 〈· · ·〉.Similarly, we evaluate δ2 in a slightly trickier fashion: we note that

δ2[J ] = − 1

2(4!)2eZ0[J ]〈 δ

4

δJ41

〉 e−Z0[J ]δ1[J ] , (4.15)

Page 3: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

Perturbative Evaluation of the Path Integral: λϕ4 Theory 3

by inserting e−Z0[J ]eZ0[J ] in the middle of (4.1.12). Next the expansion

δ4

δJ4e−Z0[J ] =

δ4e−Z0[J ]

δJ4+ 4

δ3e−Z0[J ]

δJ3

δ

δJ+ 6

δ2e−Z0[J ]

δJ2

δ2

δJ2(4.16)

+4δe−Z0[J ]

δJ

δ3

δJ3+ e−Z0[J ] δ

4

δJ4(4.17)

allows us to write

δ2 =1

2δ2

1 −1

2(4!)2eZ0[J ]

⟨(4δ3e−Z0[J ]

δJ31

δ

δJ1+ 6

δ2e−Z0[J ]

δJ21

δ2

δJ21

(4.18)

+4δe−Z0[J ]

δJ1

δ3

δJ31

+ e−Z0[J ] δ4

δJ41

)⟩1

δ1[J ] . (4.19)

Comparison with the expansion (4.1.10) for Z[J ] shows that the “discon-

nected” part 12δ

21 drops out. By disconnected we mean a contribution which

can be written as the product of two or more functions of J . This con-

cept will become obvious in the diagrammatic representation. The fact that

Z generates only connected pieces is true to all orders (see problem). For

example, the order λ3 contribution in (4.1.10) is connected: write

δ3 = − 1

3!

⟨eZ0VxVyVze

−Z0⟩xyz

= − 1

3!

⟨(eZ0Vxe

−Z0) (eZ0Vye

−Z0) (eZ0Vze

−Z0)⟩xyz

−1

2

⟨(eZ0Vxe

−Z0) (eZ0VyVze

−Z0)⟩xyz

+ δc3

=1

3!δ3

1 [J ] + δ1[J ]δc2[J ] + δc3[J ] .

In the above δc2, δc3 stand for the connected pieces. To arrive at this form, we

have used the fact that there are only two types of “disconnectedness”: all

three x, y, z disconnected, and only one disconnected from the other two;

and there are three ways to obtain the latter possibility. The parentheses

in (4.1.17) serve to shield other terms from the action of the derivative

operators within them. It follows that the term appearing in the expansion

of Z can be rewritten, using (4.1.18):

δ3 − δ1δ2 +1

3δ3

1 = δc3 +1

3!δ3

1 + δ1δc2 − δ1

(δc2 +

1

2δ2

1

)+

1

3δ3

1 = δc3 . (4.20)

Now, the explicit evaluation of the connected part of δ2 yields, save for the

J-independent part,

Page 4: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4 Perturbative Evaluation of the Path Integral: λϕ4 Theory

δc2[J ] = +1

2

⟨Ja∆ax

(1

6∆3xy +

1

4∆xx∆yy∆xy

)∆ybJb

⟩xyab

(4.21)

+1

8

⟨Ja∆ax∆yy∆

2xy∆xbJb

⟩xyab

(4.22)

+2

4!〈Ja∆ax∆xx∆xy∆yb∆yc∆ydJbJcJd〉xyabcd (4.23)

+3

2(4!)

⟨JaJb∆ax∆bx∆2

xy∆yc∆ydJcJd⟩xyabcd

(4.24)

+1

2(3!)2〈JaJbJc∆ax∆bx∆cx∆xy∆yd∆ye∆yfJdJeJf 〉xyabcdef .(4.25)

The resulting connected Green’s functions follow from (4.1.2):

G(2)(x1, x2) = ∆(x1 − x2)− λ

2

∫d4y∆ (x1 − y) ∆(y − y)∆(y − x2)

+λ2

6

∫d4xd4y∆(x1 − x)∆3(x− y)∆(y − x2)

+λ2

4

∫d4xd4y∆(x1 − x)∆2(x− y)∆(y − y)∆(x− x2)

+λ2

4

∫d4xd4y∆(x1 − x)∆(x− x)∆(x− y)∆(y − y)∆(y − x2)

+O(λ3) ,

G(4)(x1, x2, x3, x4) = −λ∫d4x∆(x1 − x)∆(x2 − x)∆(x3 − x)∆(x4 − x)

+λ2

2

∫d4xd4y∆2(x− y)[∆(x1 − x)∆(x2 − x)∆(x3 − y)∆(x4 − y)

+∆(x1 − x)∆(x3 − x)∆(x2 − y)∆(x4 − y)

+∆(x1 − x)∆(x4 − x)∆(x2 − y)∆(x3 − y)]

+λ2

2

∫d4xd4y∆(y − y)∆(x− y)[∆(x1 − x)∆(x2 − x)∆(x3 − x)

∆(x4 − y) + cyclic permutations] +O(λ)3) ,

and finally

G(6)(x1, · · · , x6) = λ2

∫d4xd4y∆(x− y)

∑(ijk)

∆(x1 − x)∆(xj − x)∆(xk − x)

∆(x` − y)∆(xm − y)∆(xn − y) +O(λ3) .

where the sum in the last expression runs over the triples (ijk) = (123),

Page 5: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

Perturbative Evaluation of the Path Integral: λϕ4 Theory 5

(124), (125), (126), (134), (135), (136), (145), (146), (156), with (`mn)

assuming the complementary value, i.e., (`mn) = (456) when (ijk) = (123),

etc. The remaining Green’s functions get no contribution to this order in

λ. Note that the λ0 contribution to G(2), the λ contribution to G(4) and

the λ2 contribution to G(6) were all previously obtained in the classical

approximation of the last chapter.

It is straightforward to derive the p-space Green’s functions, using (3.4.26).

We find

G(2)(p,−p) =1

p2 +m2− λ

2

1

(p2 +m2)2

∫d4q

(2π)4

1

q2 +m2

+λ2

6

1

(p2 +m2)2

∫d4q1

(2π)4

d4q2

(2π)4

d4q3

(2π)4

δ(p− q1 − q2 − q3)(2π)4

(q21 +m2)(q2

2 +m2)(q23 +m2)

+λ2

4

1

(p2 +m2)2

∫d4q

(2π)4

1

q2 +m2

∫d4`1(2π)4

d4`2(2π)4

δ(`1 − `2)(2π)4

(`21 +m2)(`22 +m2)

+λ2

4

1

(p2 +m2)2

∫d4q

(2π)4

1

q2 +m2

1

p2 +m2

∫d4`

(2π)4

1

`2 +m2

+O(λ3)

G(4)(p1, p2, p3, p4) =4∏i=1

1(p2i +m2

) − λ+1

2λ2

∫d4q

(2π)4

1

q2 +m2

4∑i=1

1

p2i +m2

+λ2

2

∫d4q1

(2π)4

d4q2

(2π)4

1

(q21 +m2)(q2

2 +m2)

∑(ij)

δ(q1 + q2 − pi − pj)(2π)4

+O(λ3) .

In the last expression, the sum ij runs over (ij) = (12), (13), (14) only.

Finally G(6) is given by (3.4.29). These expressions are clearly unwieldy.

One needs to devise a clever way of remembering how to generate them.

This is exactly what the Feynman rules achieve. We now proceed to state

them:

1. For each factor 1p2+m2 draw a line with momentum p flowing through it:

.5.2 :1

p2 +m2. (4.26)

2. For each factor of −λ/4! draw a four-point vertex with the understanding

Page 6: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

6 Perturbative Evaluation of the Path Integral: λϕ4 Theory

that the net momentum flowing into the vertex is zero:

− λ4!

11 (p1 + p2 + p3 + p4 = 0) . (4.27)

3. In order to get the contributions to G(N)(p1, · · · , pN ), draw all the possible

arrangements which are topologically inequivalent after having identified the

external legs. The number of ways a given diagram can be drawn is the

topological weight of the diagram.

4. After having conserved momentum at every vertex, integrate over the

internal loop momenta with∫ d4q

(2π)4 .

The result gives the desired Green’s function. Perhaps a more systematic

way to describe these rules is to attach to each vertex the factor− λ4!δ(Σp)(2π)4,

where Σp is the net incoming momentum at that vertex. The one integrates

over all internal momenta. In this manner one obtains an overall (2π)4δ(Σp)

where Σp is the net momentum flow into the Green’s function. For instance,

the expression (4.1.24) is diagrammatically rewritten as

.6.7.6.7.6.7.6.7.6.7.6.7 (4.28)

From the rules we can easily obtain the analytical expressions corresponding

to these diagrams:

a) We need one vertex and three propagators. There are

four ways to attach the first leg of the vertex to 1, three

ways to attach the second leg to 2. Hence the weight 14!4 ·

3 = 12 . The vertex counts for −λ. Let q be the momentum

circulating around the loop. The rules then give

−λ2

∫d4q

(2π)4(2π)4δ(p1 + p2 + q − q) 1

q2 +m2

1

(p2 +m2)2 . (4.29)

b) We need two vertices. There are four ways to attach the

first leg of the first vertex to 1, four ways to attach the first

leg of the second vertex to 2, three ways to sew the second

leg of the first vertex to the second, and two ways to sew the

third leg of the first vertex to the second. Hence the weight14! ·

14!4 · 4 · 3 · 2 = 1

6 . Note that we did not count that we

Page 7: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

Perturbative Evaluation of the Path Integral: λϕ4 Theory 7

initially had two vertices to play with. This is because the

diagrams are the same irrespective of which vertex was used.

The strength of this diagram is (−λ)2 = λ2. If we call the

momenta flowing in the internal legs q1, q2, q3, the Feynman

rules give

λ2

6

1

(p2 +m2)2

∫d4q1

(2π)4

d4q2

(2π)4

d4q3

(2π)4(2π)8 (4.30)

×δ(p1 − q1 − q2 − q3)δ(p2 + q1 + q2 + q3)

(q21 +m2)(q2

2 +m2)(q23 +m2)

.(4.31)

c) We need two vertices: four ways to attach the first leg

to 1, four ways to attach the third leg of the tied vertex to

the other, three ways to tie the fourth leg of the first vertex

to the second one. Hence 14!

14!4 ·3 ·4 ·3 = 1

4 . Strength (−λ)2.

d) We need two vertices; four ways to attach one vertex to

1, four ways to attach the other vertex to 2. This leaves three

legs from each vertex free to be tied together in one way. For

each vertex there are three ways to close the buckle. Hence14!

14!4 · 4 · 3 · 3 = 1

4 . Strength (−λ)2.

Thus we see that the Feynman rules have reduced the problem to that faced

by a child assembling a “Leggo” set. The basic tools are the propagator

(line and the vertex. With a bit of skill one can read those directly from the

Lagrangian. The same applies for the four-point function:

.7.7 .6.7.6.7.6.7.6.7.6.7 (4.32)

.7.7.7.7.7.7 (4.33)

corresponding to the analytical expression (4.1.25). We leave it to the reader

to verify the correctness of the numerical factors in (4.1.25). Let us remark

that in the expression for G(N)(p1, · · · , pN ), we will have the multiplicative

factor∏Ni=1

(p2

1 +m2)−1

corresponding to the propagation of the external

legs.

It is much simpler to deal with the Green’s functions generated by the

effective Action. Their relation to G(n) is very simple: while G(n) are con-

nected, the Γ(n) are one particle irreducible. In particular Γ(2)(p) is minus

Page 8: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

8 Perturbative Evaluation of the Path Integral: λϕ4 Theory

the inverse propagator. We have already come in contact with this result in

conjunction with the tree diagram, but the result holds true in all orders of

perturbation theory. As a result Γ(4) contains only the diagrams

.7.9.7.7.7.7.7.7 (4.34)

with no propagators for the external legs. In order to show that the Γ(n)

are one particle irreducible, it is convenient to use as a starting point the

defining equations:

δΓ[ϕ]

δϕ= −J ;

δZ[J ]

δJ= ϕ , (4.35)

and keep on differentiating them.

After all this work, let us go back to the Lagrangian where all the ingre-

dients of the Feynman rules are easily identified: the four-particle vertex

coming from the λϕ4 term, and the propagator coming from the kinetic and

mass terms. Hence, after a little bit of practice, one can just read off the

Feynman rules from L. The difficult part is to get the right signs and weight

factors in front of the diagram.

4.0.1 PROBLEMS

A. Draw the contributions in the two point function G(2) that are of order

λ3.

B. Draw the contributions to G(4) of order λ3.

C. Write the analytical expressions for the diagrams of problem A, including

the weights.

∗D. Derive the Feynman rules for V = λ4!ϕ

4 + µ3!ϕ

3.

∗∗E. Show that Z[J ] generates only connected Feynman diagrams.

∗∗∗F. Derive the form of the effective Action Seff [ϕcl] to order λ2, and show

that the one particle reducible diagrams do not appear in the Green’s func-

tions it generates.

Page 9: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.1 Divergences of Feynman Diagrams 9

4.1 Divergences of Feynman Diagrams

No sooner has the beauty of the Feynman rules sunk in that one realizes that

most of the loop integrations diverge! For instance the O(λ) contribution to

G(2) involves the integral

∫d4q

(2π)4

1

q2 +m2; (4.36)

it clearly diverges when q → ∞ since the integrand does not have enough

juice to make up for the measure. Such a divergence is called an ultraviolet

divergence. It occurs for large momenta, or equivalently, small distances [in

x-space it comes from ∆(0)], and clearly has to do with the fact that one is

taking too many derivatives with respect to J at the same point. Another

example occurs in the “fish” diagram

.6.6 ∼ λ2

2

∫d4q1

(2π)4

d4q2

(2π)4

δ(q1 + q2 − p1 − p2)

(q21 +m2)(q2

2 +m2)(4.37)

=1

(2π)4

λ2

2

∫d4q

(2π)4

1

(q2 +m2) ((q − p1 − p2)2 +m2). (4.38)

Here as q → ∞, the integral behaves as d4qq4 , which is a logarithm: log q.

It also diverges! In passing, we note that when m2 = 0, (p1 + p2)2 = 0 it

also diverges for low q. Such a divergence is called an infrared divergence.

Typically such divergences occur only in the massless case and for special

value of the external momenta [in this case p1 + p2 = 0 because we are in

Euclidean space]. For the moment we take m2 6= 0 and concentrate on the

ultraviolet divergences.

On the face of it, it is catastrophic to our program to find that our carefully

constructed Green’s functions diverge. Still, with the hindsight of History

we do not get discouraged, but try to learn more about these divergences.

We will find that they appear in a very traceable way, and that by a suitable

redefinition of the fields and coupling constants, they will disappear! This is

the miracle of renormalization which, as we shall see, occurs only for certain

theories.

By a mixture of topology and power counting we now show how to tell

where the divergences are. Consider a Feynman diagram with V vertices,

E external lines and I internal lines. For a start we assume only scalar

particles are involved.

The number of independent internal momenta is the number of loops, L,

in the diagram. The I internal momenta satisfy V −1 relations among them-

Page 10: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

10 Perturbative Evaluation of the Path Integral: λϕ4 Theory

selves (the −1 appears here because of overall momentum conservation), so

that

L = I − V + 1 . (4.39)

This relation enables us to compute the naive count of powers of momenta

for the diagram. This yields the superficial (apparent) degree of divergence

of the diagram D. To compute it, we note that there are — L independent

loop integrations, each providing, in d dimensions, d powers of momenta. —

I internal momenta, such providing a propagator with two inverse powers

of momenta. Hence

D = dL− 2I . (4.40)

We need one more relation among V , E and I. Suppose VN stands for

the number of vertices with N legs. In a diagram with VN such vertices,

we have NVN lines which are either external or internal. An internal line

counts twice because it originates and terminates at a vertex, so that

NVN = E + 2I. (4.41)

These relations allow us to express in D terms of the number of external

lines and vertices

D = d− 1

2(d− 2)E + VN

(N − 2

2d−N

). (4.42)

In four dimensions, this reduces to

D = 4− E + (N − 4)VN [four dimensions] . (4.43)

Furthermore, in the theory of interest to us, N = 4. Hence

D = 4− E [for λϕ4 theory in four dimensions] . (4.44)

The important result here is that the superficial degree of divergence does

not depend on the number of vertices, but only on the number of external

legs! Thus we have only two candidates with D ≥ 0 — G(2) with D = 2

superficial quadratic divergence. — G(4) with D = 0 superficial logarithmic

divergence. Note that these two- and four-point interactions are already

Page 11: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.1 Divergences of Feynman Diagrams 11

present in the Lagrangian, a fact which will prove crucial for renormaliza-

tion. Also, D = 0 does not necessarily mean a logarithmic divergence: the

fundamental vertex has D = 0.

This analysis does not prove that G(6), G(8), · · · which have a negative

D, converge. This is why D is called superficial. Consider a “n-particle

reducible” diagram with E external lines; it is a diagram which can be

disconnected by cutting at least n internal lines. In general, if D1 and D2

are the superficial degrees of divergence of the two blobs shown below, then

the whole n-particle reducible diagram has

D = D1 +D2 + 4(n− 1)− 2n , (4.45)

since the two blobs are connected by n propagators and n− 1 loops:

1.51 (4.46)

Note that by definition, I and II are at least n-particle reducible themselves.

In our case, when n = 1, we could have D1 = D2 = 0 and yet we would

obtain D = −2, making the diagram apparently convergent. An example of

this situation is the “dinosaur” diagram

1.51 (4.47)

which clearly diverges because of the two divergent loop integrations. An-

other example of a one particle reducible graph is

1.51 (4.48)

In this case D1 = 2, D2 = −2; it is apparently convergent, but is not because

of the “wart” on one of the legs. This should also serve to show that a four

point function may be more divergent than it would naively seem to be:

witness this quadratically divergent four-point function

.8.8 (4.49)

Let us remark that by considering Γ(n) we avoid such one particle reducible

diagrams. When n = 2, D = D1+D2 and we can haveD1 (orD2) sufficiently

negative to offset D2 (or D1) and yield a negative D. An example is pictured

by the “lobster” diagram

1.51 (4.50)

Page 12: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

12 Perturbative Evaluation of the Path Integral: λϕ4 Theory

Similarly, when n = 3, we can have diagrams like

1.51 (4.51)

[in this complicated diagram, each vertex has at most one external line em-

anating from it]. In λϕ4 theory, diagrams can be at most three-particle

reducible because any vertex attached to an external line can be discon-

nected from the diagram by cutting its remaining three legs.

The procedure of hunting for diagrams which have a negative D and yet

are divergent is now clear. Take any Feynman diagram and catalog it in

terms of its reducibility: in our case it is either one-, two-, or three-particle

reducible. It if is three-particle reducible, decompose it into units, and look

for breakups of the form

1.51 N ≥ 5 (4.52)

Then in this type of diagram, we have a primitively divergent four-point

function, where the second blob can be decomposed in the same way.

Similarly, hidden divergences in two-particle reducible diagrams will arise

in the diagrams of the form

1.51 N > 4 1.51 N > 2 (4.53)

with the same breakup to be repeated in the second blob. Finally, one-

particle reducible diagrams which can be decomposed in the form

1.51 N > 5 1.51 N > 3 (4.54)

will have hidden divergences. The same decomposition can be carried out

for the second blobs until all such structures have been uncovered. This

exhaustive catalog shows that truly convergent diagrams do not contain

hidden two- and four-point functions.

One can understand the origin of hidden ultraviolet divergences in any

diagram in a more pedestrian fashion. Consider any loop residing inside a

diagram. Integration over the loop momentum in four dimensions will lead

to a UV divergence if the loop is bounded by one or two propagators (internal

lines) at most. Any more will give UV convergence. A loop bounded by one

propagator involves only one vertex,

.8.8 (4.55)

Page 13: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.1 Divergences of Feynman Diagrams 13

leaving two free legs, which in turn may be attached to the rest of the

diagram (or one external, one attached). In this case one can isolate this

two-point function from the insides of the diagram. A loop bounded by two

propagators involves two vertices and therefore four free legs

.8.8 (4.56)

which can be attached to the rest of the diagram, or else up to three can

serve as external lines. In all these cases, one is led to isolate from the

diagram a four-point function, or a two-point function if two of the four legs

are attached together (in this case the divergence becomes quadratic). In

this way, one sees that UV divergence inside a diagram originate from such

loops and that such loops appear in two- and four-point functions nested

inside the diagram.

Thus, for the λϕ4 theory in four dimensions, a Feynman diagram is truly

convergent if its superficial degree of convergence D is positive and if it

cannot be split up into three-, two- or one-particle reducible parts of the kind

just described which can contain isolated two- and four-point function blobs.

Stated more elegantly, a Feynman diagram is convergent if its superficial

degree of convergence and that of all its subgraphs are positive. This is

known as Weinberg’s theorem, and it holds irrespective of the field theory.

This means that the generic sources of the divergences are the two- and

four-point functions and nothing else. They are the culprits! So, if we control

them in G(2) and G(4), we have the possibility of controlling the divergences

of all the other G(N)’s!

The graphs which contain the generic divergences are said to be primitively

divergent. The fact in λϕ4 theory that the primitively divergent interactions

are finite in number (two- and four-point interactions) and are of the type

that appears in the Lagrangian, is a necessary ingredient for the successful

removal of the ultraviolet divergence by clever redefinitions. A theory for

which this is possible is said to be renormalizable. We can see from (4.2.4)

that very few theories of interacting scalars satisfy these requirements (see

problem). — When d = 4, we see that D grows with the number of vertices

for which N > 4. Hence ϕ5, ϕ6, · · · theories in four-dimensions, although

perfectly reasonable classically, lead to an infinite number of primitively

divergent diagrams (the more vertices the more divergent!). In this case the

situation quickly gets out of hand and the hope of tagging the divergences

disappears, and hence the renormalizability. — When d = 2 (one space and

one time dimension) the situation is reversed. There

Page 14: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

14 Perturbative Evaluation of the Path Integral: λϕ4 Theory

D = 2− 2VN (two-dimensions) , (4.57)

and D does not depend on N , which labels the type of interaction! It

depends only on the number of vertices, and the more vertices, the more

convergent the Feynman diagram! Also D does not depend on the number

of external legs! So the only primitively divergent diagrams have one or no

vertex. Since divergences occur because of loop integrations, this means that

divergences occur only when a leg from one vertex is connected to the same

vertex, and not from the interaction between two or more vertices. Such

self-inflicted divergences are called “normal ordering” divergences. In two

dimensions, the only ultraviolet divergences come from “normal ordering,”

and not from the type of interaction. Finally, we note that when d ≥ 7,

there are no theories with a finite number of primitively divergent graphs.

The last theory in higher dimensions is λϕ3 in six dimensions, where λ

is dimensionless, since ϕ now has dimension −2. There the primitively

divergent diagrams are few since V does not appear in the expression for D

D = 6− 2E (ϕ3 in six dimensions) , (4.58)

so that the one-, two- and three-point functions are primitively divergent

(quartic, quadratic and logarithmic, respectively). This theory, although

having an unsatisfactory potential unbounded from below, is interesting in

that it shares with the more complicated gauge theories the property of

asymptotic freedom.

To summarize this section, we have noticed the appearance of ultraviolet

divergences in Feynman diagrams with loops (the bad news), but we have

seen that we can, at least in our theory, narrow them down as coming only

from two primitively divergent Green’s functions (the good news). Hence,

if we can arrange to stop divergences from appearing in G(2) and G(4), we

have a hope of stemming the flood and obtaining convergent answers!

4.1.1 PROBLEMS

A. In four dimensions, find all primitively divergent diagrams for the ϕ3

theory. For each, give examples in lowest order of perturbation theory.

vskip .5cmB. In three dimensions (d = 3), list all theories of interacting

scalars with a finite number of primitively divergent graphs. Give graphical

examples for each.

Page 15: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.2 Dimensional Regularization of Feynman Integrals 15

C. Repeat B, when d = 5, and show that when d ≥ 7 there are no theories

with a finite number of primitively divergent graphs.

D. For d = 2 3, 5, 6, find the dimensions of the various coupling constants in

the theories where there is a finite number of primitively divergent graphs.

4.2 Dimensional Regularization of Feynman Integrals

In the following, we proceed to evaluate the Feynman diagrams. The proce-

dure is straightforward for the UV convergent ones, while special measures

have to be taken to evaluate the divergent ones. In those we are confronted

with integrals of the form

I4(k) =

∫ +∞

−∞d4`F (`, k) , (4.59)

where for large `, F behaves either as `−2 or `−4. The basic idea behind the

technique of dimensional regularization is that by lowering the number of di-

mensions over which one integrates, the divergences trivially disappear. For

instance, if F → `−4, then in two dimensions the integral (4.3.1) converges

at the UV end.

Mathematically, we can introduce the function

I(ω, k) =

∫d2ω`F (`, k) , (4.60)

as a function of the (complex) variable ω. Evaluate it in a domain here I

has no singularities in the ω plane. Then invent a function I ′(ω, k) which

has well-defined singularities outside of the domain of convergence. We say

by analytic continuation that I and I ′ are the same function.

A nice example, which is the basis for the method of analytic continuation,

is the difference between the Euler and Weierstrass representations of the

Γ-function. For <z > 0, the Euler representation is

Γ(z) =

∫ ∞0

dt e−ttz−1 . (4.61)

As such, it diverges when <z < 0, because as t approaches zero, the integral

behaves as dt/t1+|<z|, which leads to an infinity. Starting from (4.3.3) we

can split up the troublesome integration limit

Page 16: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

16 Perturbative Evaluation of the Path Integral: λϕ4 Theory

Γ(z) =∞∑n=0

(−1)n

n!

∫ α

0dt tn+z−1 +

∫ ∞α

dt e−ttz−1 , (4.62)

where α is totally arbitrary. The second integral is well-defined even when

<z < 0 as long as α > 0. The first integral has simple poles whenever z is

a negative integer or zero. We find

Γ(z) =

∞∑n=0

(−1)n

n!

αn+z

(z + n)+

∫ ∞α

dt e−ttz−1 . (4.63)

This form is valid everywhere in the z-plane. Furthermore, it should not

depend on the arbitrary coefficient α (you can check that dΓdα = 0). When

α = 1, it is the Weierstrass representation of the Γ-function. Still, to isolate

the singularities we did introduce an arbitrary scale in the process, although

the end result is independent of it.

Our problem is that integral expressions like (4.3.2) are like Euler’s. We

want to find the equivalent of Weierstrass’ representation. Our procedure

will be as follows: 1) establish a finite domain of convergence for the loop

integral in the ω-plane. For divergent integrals, it will typically lie to the

left of the ω = 2 line; 2) construct a new function which overlaps with the

loop integral in its domain of convergence, but is defined in a larger domain

which encloses the point ω = 2; 3) take the limit ω → 2.

We now show how this is done in the case of one-loop diagrams, following

the procedures of ‘t Hooft and Veltman, Nucl. Phys. 44B, 189 (1972). Let

us split up the domain of integration as

d2ω`→ d4` d2ω−4` . (4.64)

Next in the 2ω− 4 space, introduce polar coordinates, and call L the length

of the 2ω − 4 dimensional `-vector. The integral now reads

I =

∫d4`

∫dΩ2ω−4

∫ ∞0

dLL2ω−5 1

(L2 + `2 +m2). (4.65)

Integration over the angles can be performed (see Appendix B), with the

result

I =2πω−2

Γ(ω − 2)

∫d4`

∫ ∞0

dLL2ω−5 1

(L2 + `2 +m2). (4.66)

This expression is not well-defined because it is UV divergent for ω ≥ 1,

Page 17: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.2 Dimensional Regularization of Feynman Integrals 17

and the integration over L diverges at the lower end (“infrared”) whenever

ω ≤ 2. Thus, there is no overlapping region in the ω plane where I is well-

defined. The IR divergence is, however, an artifact of the break up of the

measure. Observe that by writing

L2ω−6 =1

ω − 2

d

dL2

(L2)ω−2

, (4.67)

and integrating by parts over L2, and throwing away the surface term, we

obtain

I =πω−2

Γ(ω − 1)

∫d4`

∫ ∞0

dL2(L2)ω−2

(− d

dL2

)1

L2 + `2 +m2, (4.68)

where we have used Γ(ω − 1) = (ω − 2)Γ(ω − 2). Now the representation

(4.3.9) has an infrared divergence for ω ≤ 1 and the same UV divergence

for ω ≥ 1, that is, still no overlapping region of convergence. So we do it

again, and obtain

I =πω−2

Γ(ω)

∫d4`

∫ ∞0

dL2(L2)ω−1

(− d

dL2

)2 1

L2 + `2 +m2, (4.69)

an expression which is well-defined for 0 < ω < 1. Note that we had to

move the IR convergence region two units to obtain a nonzero region of

convergence. Had the loop integral been logarithmically divergent, one such

step would have sufficed.

Having obtain an expression for I convergent in a finite domain (in this

case 0 < ω < 1), we want to continue I of (4.3.10) to the physical point

ω = 2. It goes as follows: insert in the integrand the clever expression

1 =1

5

(∂L

∂L+∂`µ∂`µ

), (4.70)

and integrate by parts in the region of convergence. We obtain

I = −1

5

2πω−2

Γ(ω)

∫d4`

∫ ∞0

dL2

[`µ

∂`µ+ L2 ∂

∂L2

] (L2)ω−1

(L2 + `2 +m2)3 . (4.71)

A little bit of algebra gives, after reexpressing the right-hand side in terms

of I,

Page 18: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

18 Perturbative Evaluation of the Path Integral: λϕ4 Theory

I = − 3m2

ω − 1

2πω−2

Γ(ω)

∫d4`

∫ ∞0

dL2

(L2)ω−1

(L2 + `2 +m2)4 . (4.72)

This expression displays explicitly a pole at ω = 1. The integral now diverges

at the upper end when ω ≥ 2. So we repeat the process and reinsert (4.3.11)

in (4.3.13). The result is predictable:

I =2 · 3 · 4m4

(ω − 1)(ω − 2)

πω−2

Γ(ω)

∫d4`

∫ ∞0

dL2

(L2)ω−1

(L2 + `2 +m2)5 . (4.73)

This is the desired result. The only hint of a divergence comes from the

simple pole at ω = 2, since the integral itself now converges.

Let us summarize. We first define a finite integral in the ω-plane to be

what we mean by∫d2ω`F (`, k): in this case it is the expression given by

(4.3.10), and constitutes our starting point. Then if the region of conver-

gence does not include ω = 2, we continue analytically by means of the trick

(4.3.11).

It would be nice to show that for a convergent integral, the procedure

that leads to (4.3.10) indeed gives the right answer. Take as an example the

convergent integral

I =

∫d2ω`

1

(`2 +m2)6 . (4.74)

It is easy to see that the same procedure leads to the expression

I(ω) =πω−2

Γ(ω − 1)

∫d4`

∫ ∞0

dL2(L2)ω−2

(− d

dL2

)1

(L2 + `2 +m2)6 ,

(4.75)

which is perfectly finite at ω = 2. We find

I(2) =

∫d4`

(−1)

(L2 + `2 +m2)6

∣∣∣∣∞0

=

∫d4`

1

(`2 +m2)6 , (4.76)

as desired. Therefore, the procedure is entirely consistent.

After all this ponderous work, let us turn to a more cavalier interpretation

of integrals in 2ω-dimensions, as derived in Appendix B.

If we were to blindly plug in the formulae of Appendix B, we would obtain

Page 19: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.2 Dimensional Regularization of Feynman Integrals 19

∫d2ω`

(`2 +m2)= πω

Γ(1− ω)

Γ(1)

1

(m2)1−ω . (4.77)

We expand around ω = 2, using for n = 0, 1, 2, · · · and ε→ 0 the formula

Γ(−n+ε) =(−1)n

n!

[1

ε+ ψ(n+ 1) +

1

[π2

3+ ψ2(n+ 1)− ψ′(n+ 1)

]+O(ε2)

],

(4.78)

where

ψ(n+ 1) = 1 +1

2+ · · ·+ 1

n− γ ,

[ψ(s) =

d ln Γ(s)

ds

], (4.79)

ψ′(n+ 1) =π2

6−

n∑k=1

1

k2, ψ′(1) =

π2

6(4.80)

γ being the Euler-Mascheroni constant

ψ(1) = −γ = −0.5772 · · · . (4.81)

The result is

limω→2

∫d2ω`

(`2 +m2)= −π2m2

[1

2− ω+ ψ(2)

]. (4.82)

It can be shown that this is the same result as that obtained by integrating

(4.3.14).

From now on, we are going to use the naive formulae of Appendix B,

and not concern ourselves with their justification since we know they are

legitimate, thanks to ‘t Hooft and Veltman.

4.2.1 PROBLEMS

A. Show that the naive procedure of Appendix B and the more careful one

outlines in this section lead to the same results for the integrals

I =

∫d2ω`

1

(`2 +m2), I =

∫d2ω`

1

(`2 +m2)2. (4.83)

B. Prove the last four formulae of Appendix B.

Page 20: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

20 Perturbative Evaluation of the Path Integral: λϕ4 Theory

4.3 Evaluation of Feynman Integrals

Using the techniques of the previous section and the formulae of Appendix

B, we proceed to evaluate the low order Feynman diagrams for λϕ4 theory.

We have seen that the technique of dimensional regularization is predi-

cated on the evaluation of Feynman integrals in 2ω dimensions, where the

coupling constant λ is no longer dimensionless. We find it convenient to

redefine it in terms of a dimensionless coupling constant by the artifact

λold = λnew

(µ2)2−ω

, (4.84)

where λnew is dimensionless and µ2 is an arbitrary constant with the di-

mension of mass. Alternatively, we can say that we evaluate the Green’s

functions for the theory defined by the action

Sω[ϕ] =

∫d2ωx

[1

2∂µϕ∂µϕ+

1

2m2ϕ2 +

λ

4!

(µ2)2−ω

ϕ4

]. (4.85)

The Feynman rules for this theory are the same as the one for the theory

in four-dimensions with three exceptions: 1) the scalar product between

vectors is summed over their 2ω components, 2) the loop integrals appear

with∫

d2ω`(2π)2ω and 3) the vertex strength −λ is replaced by (−λ)

(µ2)2−ω

.

Let us evaluate the lowest order diagrams for this theory. We start with

the “tadpole” diagram (in conventional terms such a diagram comes from

not normal ordering the interaction term). It is given by

.7.7 ≡ T =1

2(−λ)

(µ2)2−ω ∫ d2ω`

(2π)2ω

1

(`2 +m2)(4.86)

= − λm2

2(4π)2

(4πµ2

m2

)2−ωΓ(1− ω) , (4.87)

where we have used (B–16), and have kept m2 in front because the diagram

has dimension of mass squared. By expanding around ω = 2, we obtain

T = − λm2

32π2

[1 + (2− ω) ln

4πµ2

m2+ · · ·

] [− 1

2− ω− ψ(2) + · · ·

](4.88)

=λm2

32π2

1

2− ω+ ψ(2)− ln

m2

4πµ2+O(2− ω)

. (4.89)

Observe that in this formula, the judicious introduction of the arbitrary scale

Page 21: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.3 Evaluation of Feynman Integrals 21

µ2 allows us to keep track of dimensions, and that the pole from the expan-

sion of Γ cancels against the first order term in the expansion of(

4πµ2

m2

)2−ω,

leaving us with a finite contribution as ω → 2. This feature will be present

in the evaluation of all divergent graphs. We conclude that the divergence

in T appears as a simple pole, and that the finite part of T , which in this

case does not depend on the external momenta, is totally arbitrary because

a change in µ2 affects it.

The next diagram is the “fish”

11 p1 + p2 + p3 + p4 = 0 . (4.90)

The Feynman rules give [p = p1 + p2]

11 =1

2(−λ)2

(µ2)4−2ω

∫d2ω`

(2π)2ω

1

(`2 +m2)

1

[(`− p)2 +m2]. (4.91)

When there are more than one propagator taking part in a loop integration,

it is convenient to introduce the Feynman parametrization based on the

formula

1

Da11 D

a22 · · ·D

akk

=Γ(a1 + a2 + · · · ak)Γ(a1)Γ(a2) · · ·Γ(ak)

∫ 1

0(4.92)

· · ·∫ 1

0dx1 · · · dxk

δ(1− x1 − · · · − xk)xa1−11 · · ·xak−1

k

(D1x1 + · · ·Dkxk)a1+···+ak .(4.93)

It allows for a convenient rearrangement of the loop momenta. In this case

we use it in the form

1

[`2 +m2] [(`− p)2 +m2]=

∫ 1

0

dx

[`2 +m2 − 2` · p(1− x) + p2(1− x)]2.

(4.94)

The denominator can be rewritten in the form

`′2 +m2 + p2x(1− x) , (4.95)

where

`′ = `− p(1− x) . (4.96)

Since we are dealing with convergent integrals, we use d2ω`′ = d2ω` and

Page 22: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

22 Perturbative Evaluation of the Path Integral: λϕ4 Theory

relabel the loop integral from ` to `′. The result of these manipulations

yields

.7.7 =λ2

2

(µ2)4−2ω

∫ 1

0dx

∫d2ω`

(2π)2ω

1

[`2 +m2 + p2x(1− x)]2. (4.97)

Thanks to this trick, we can now use (B–16) and integrate, obtaining(4.98)

.7.7 =λ2

2

(µ2)4−2ω

∫ 1

0dx

Γ(2− ω)

(4π)ω1

[m2 + p2x(1− x)]2−ω. (4.99)

Before expanding, remember that this diagram has dimensions(µ2)2−ω

,

which we show explicitly. After expansion we obtain to O(2− ω)

.6.6 (4.100)

=(µ2)2−ω λ2

32π2

∫ 1

0dx

1

2− ω+ ψ(1)− ln

(m2 + p2x(1− x)

4πµ2

)(4.101)

=(µ2)2−ω λ2

32π2

[1

2− ω+ ψ(1)−

∫ 1

0dx ln

(m2 + p2x(1− x)

4πµ2

)].(4.102)

Again, observe that this time the finite part depends not only on µ2, which

is arbitrary, but also on the external momenta. Let us emphasize that

this arbitrariness in the finite part is generic to the method because of the

separation of a divergent expression into a divergence plus a finite part [after

all ∞+ 5 =∞+ 6!].

There remains to integrate over the Feynman parameter x. Since x(1−x)

is always positive [0 < x(1 − x) < 1/4] over the range of integration, the

argument of the logarithm is always positive, making the integral easy to

evaluate. We use the formula

∫ 1

0dx ln

[1 +

4

ax(1− x)

]= −2 +

√1 + a ln

(√1 + a+ 1√1 + a− 1

), a > 0 .

(4.103)

The result is then

.7.7 =(µ2)2−ω λ2

32π2

[1

2− ω+ ψ(1) + 2 + ln

4πµ2

m2(4.104)

√1− 4m2

p2ln

1 + 4m2

p2 + 1√1 + 4m2

p2 − 1

+O(2− ω)

]. (4.105)

Page 23: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.3 Evaluation of Feynman Integrals 23

In the evaluation of the four-point function, there will be three such con-

tributions with p = p1 + p2, p = p1 + p3, and p = p1 + p4, corresponding

to the s-, t- and u-channel contributions [caution: here all momenta are in-

coming]. This diagram is computed in the Euclidean domain; continuation

to Minkowski space will entail changing the sign of p2 and carefully inter-

preting the result. As it stands, however, the finite part has no interesting

analytical structure as long as p2 > 0. We will return to this point when

the interpretation of the result is discussed in Minkowski space. Using the

same techniques, we compute the “double scoop” diagram:

.7.7 =1

4λ2(µ2)4−2ω

∫d2ω`

(2π)2ω

1

`2 +m2

∫d2ωq

(2π)2ω

1

(q2 +m2)2 (4.106)

(4.107)

=λ2m2

1024π4

[1

(2− ω)2+

1

(2− ω)

[2 ln

4πµ2

m2+ ψ(2) + ψ(1)

](4.108)

+2 ln2 4πµ2

m2+ 2 ln

4πµ2

m2[ψ(2) + ψ(1)] +

1

2

([ψ(2) + ψ(1)]2(4.109)

2π2

3− ψ′(2)− ψ′(1)

)+O(ω − 2)

]. (4.110)

Note the appearance of a double pole and the arbitrariness of the residue of

the simple pole and of the finite part.

Finally, we calculate in detail the “setting sun” diagram

1.51.5 (4.111)

It is worth doing in some detail as it is a two-loop diagram. Call the diagram

Σ(p). The Feynman rules give

Σ(p) =λ2(µ2)4−2ω

6

∫d2ω`

(2π)2ω

∫d2ωq

(2π)2ω

1

(`2 +m2)(q2 +m2) ([q + p− `]2 +m2)(4.112)

In diagrams involving several loops, the ultraviolet divergences will find

their way into parametric integrals. For convenience one wants to have

as few divergences as possible in these integrals. This means that special

techniques have to be applied on (4.4.19) before introducing the Feynman

parameters, unless the integral is at worse logarithmically divergent. Using

the trick

Page 24: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

24 Perturbative Evaluation of the Path Integral: λϕ4 Theory

1 =1

[∂`µ∂`µ

+∂qµ∂qµ

](4.113)

in the integrand of (4.4.19) and integrating by parts, we find

Σ(p) = − 1

λ2

6

(µ2)4−2ω

∫d2ω`

(2π)2ω

∫d2ωq

(2π)2ω

(`µ

∂`µ+ qµ

∂qµ

)(4.114)

· 1

(`2 +m2)(q2 +m2) ([q + p− `]2 +m2), (4.115)

where we have discarded the surface terms, in keeping with the philosophy

of analytic continuation we discussed in conjunction with one-loop diagrams

[see ‘t Hooft and Veltman, Nucl. Phys. B44, 189 (1972), and also Curtright

and Ghandour, Annals of Physics, 106, 209 (1977)].

Explicit differentiation gives

Σ(p) = − 1

(2ω − 3)

λ2

6

(µ2)4−2ω

∫d2ω`

(2π)2ω

∫d2ωq

(2π)2ω(4.116)

× 3m2 + p · (p+ q − `)(q2 +m2)(`2 +m2) ([q − `+ p]2 +m2)2 (4.117)

= − 1

2ω − 3

λ2

6

(µ2)4−2ω [

3m2K(p) + pµKµ(p)], (4.118)

where

K(p) =

∫d2ω`

(2π)2ω

∫d2ωq

(2π)2ω

1

(q2 +m2)2(`2 +m2) [(q − `+ p)2 +m2](4.119)

and

Kµ(p) =

∫d2ω`

(2π)2ω

∫d2ωq

(2π)2ω

(p+ q − `)µ(q2 +m2)(`2 +m2) [(q − `+ p)2 +m2]2

,

(4.120)

where we have freely made several linear changes of variables over loop

momenta. Now K(p) is log divergent and Kµ(p) is linearly divergent. We

first evaluate K(p).

It is prudent to introduce Feynman parameters one loop at a time, starting

from the most divergent one. So doing we obtain

Page 25: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.3 Evaluation of Feynman Integrals 25

K(p) =

∫d2ω`

(2π)2ω

∫d2ωq

(2π)2ω

1

(q2 +m2)2

∫ 1

0dx

1

[`2 +m2 + (p+ q)2x(1− x)]2.

(4.121)

Integrate over `, using (B–16), to obtain

K(p) =Γ(2− ω)

(4π)ω

∫ 1

0dx

∫d2ωq

(2π)2ω

1

(q2 +m2)2

1

[m2 + (p+ q)2x(1− x)]2−ω.

(4.122)

Proceed to use (4.4.8) to rewrite

K(p) =Γ(4− ω)

(4π)ω

∫ 1

0dx [x(1− x)]ω−2

∫ 1

0dy y1−ω(1− y)

∫d2ωq

(2π)2ω(4.123)

×[q2 + p2y(1− y) +m2

(1− y +

y

x(1− x)

)]ω−4

. (4.124)

Integration over q finally gives

K(p) =Γ(4− 2ω)

(4π)2ω

∫ 1

0dx [x(1− x)]ω−2

∫ 1

0dy y1−ω(1− y) (4.125)

×[p2y (1− y) +m2

(1− y +

y

x(1− x)

)]2ω−4

.(4.126)

It is convenient to introduce

2− ω ≡ ε(> 0 because of the analytic continuation) (4.127)

and expand (4.4.29) around ε = 0. The parametric integral has a pole at

ε = 0 coming from y = 0. Then set

K(p) =Γ(2ε)

(4π)4−2ε

∫ 1

0dx [x(1− x)]−ε

∫ 1

0dy y−1+ε(1− y) (4.128)

quad×[p2y(1− y) +m2

(1− y +

y

x(1− x)

)]−2ε

.(4.129)

Using

y−1+ε =1

ε

d

dyyε , (4.130)

Page 26: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

26 Perturbative Evaluation of the Path Integral: λϕ4 Theory

and integrating by parts, we find

K(p) =Γ(2ε)

(4π)4−2ε

1

ε

∫ 1

0dx [x(1− x)]−ε

∫ 1

0dy yε

×

1 + 2ε(1− y)d

dyln

[p2y(1− y) +m2

(1− y +

y

x(1− x)

)]×[p2y(1− y) +m2

(1− y +

y

x(1− x)

)]−2ε

. (4.131)

Now the parametric integral can be done by expanding around ε = 0.

The evaluation of Kµ proceeds in a similar way, giving

pµKµ = p2 Γ(2ε)

(4π)4−2ε

∫ 1

0dx [x(1− x)]−ε

∫ 1

0dy yε(1− y) (4.132)

×[p2y(1− y) +m2

(1− y +

y

x(1− x)

)]−2ε

. (4.133)

Here the parametric integral converges so that the p2 singularity is only at

the simple pole.

Expanding around ε = 0 gives

K(p) =Γ(2ε)

(4π)4−2ε

1

ε

1 + ε− 2ε lnm2 +O(ε2)

(4.134)

pµKµ(p) = p2 Γ(2ε)

(4π)4−2ε

[1

2+O(ε)

]. (4.135)

The O(ε2)[O(ε)] terms in the parametric integral for K(p)[pµKµ(p)] are very

complicated, and give contributions to the finite part of Σ(p).

Putting it all together we find

Σ(p) =λ2

6 (16π2)2

[3m2

2ε2+

3m2

ε

[3

2+ ψ(1) + ln

4πµ2

m2

]+

1

4εp2 + finite

].

(4.136)

Observe that we now have arbitrariness at the level of the simple pole (as

well as at the level of the finite part).

The finite part of Σ(p) is very difficult to obtain. It cannot be done

in closed form and necessitates the introduction of the “dilogarithm” (or

Spence) function defined by

Page 27: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.4 Renormalization 27

Li2(x) ≡ −∫ 1

0

dt

tlog(1− xt) . (4.137)

Experience shows that when masses are present, the finite parts of two-loop

diagrams are quite lengthy to evaluate.

4.3.1 PROBLEMS

A. Show that for a > 0

I(a) =

∫ 1

0dx ln

[1 +

4

ax(1− x)

]= −2 +

√1 + a ln

√1 + a+ 1√1 + a− 1

. (4.138)

∗B. Now let z = 4a . Find the singularity structure of I(z) in the complex

z-plane and specify I(z) when z is real.

C. Derive the form of the finite part of Σ(p) when m2 = 0.

D. Derive the expression for pµKµ [Eq. (4.4.34)].

∗∗E. Express K(p) to order ε2 and write the resulting integrals in terms of

dilogarithms.

4.4 Renormalization

In the previous sections we have shown how to evaluate Feynman diagrams.

In λϕ4 theory we found that some diagrams are ultraviolet divergent, with

the divergences showing up only in two- and four-point Green functions (the

primitively divergent diagrams). When dimensional regularization was used

to regularize these diagrams, the infinities resulted in poles in the dimension

plane in ε = 2− n/2 > 0, n being the number of space-time dimensions; in

addition, the finite part of these diagrams was arbitrary, depending in our

scheme on a mysterious mass parameter µ.

We are now going to show how to eliminate these poles order by order in

λ. The technique is very simple: alter the Feynman rules at each order so as

to obtain a finite result as ε → 0. As a first example, consider the tadpole

diagram

Page 28: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

28 Perturbative Evaluation of the Path Integral: λϕ4 Theory

1.5 = m2 λ

2

[1

ε+ ψ(2)− ln m2 +O(ε)

], (4.139)

where

λ ≡ λ

16π2, m2 ≡ m2

4πµ2. (4.140)

This infinity can be hidden away by adding to L the additional term

m2

[1

ε+ F1(ε, m2)

]ϕ2 , (4.141)

which we consider to be an extra interaction term. Here F1 is an arbitrary

dimensionless function, analytic as ε → 0; its presence reflects the arbi-

trariness of the procedure. This extra term results in a new Feynman rule

indicated by

1.5 = −1

2m2λ

[1

ε+ F1

]. (4.142)

Thus if we compute the inverse propagator to (λ), we find (remember that

the correction to the inverse propagator picks up a minus sign for inverting)

.7.7 = .7.7 + .7.7 + .7.7 +O(λ2) (4.143)

Γ(2)new(p) = p2 +m2

[1− 1

2λ(ψ(2)− ln m2 − F1

)]+O(λ2) .(4.144)

This very naive procedure makes the theory finite to order λ. The extra term

(4.5.3) is called a counterterm. It is crucial to note that its dependence on

the field (and its derivatives) is the same as that of a term already appearing

in L (in this case the mass term).

Now we proceed to order λ2. The one particle irreducible (1PI) four-point

function is given by

.7.7 .7.7.7.7 (4.145)

.7.9.7.9 O(λ2) (4.146)

Γ(4)(p1, p2, p2, p4) = −µ2ελ

[1− 3

(1

ε+ ψ(1) + 2− ln m2 − 1

3A(s, t, u)

+O(ε)

)]+O(λ3) , (4.147)

Page 29: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.4 Renormalization 29

where

A(s, t, u) =∑

z=s,t,u

(1 +

4m2

z

)1/2

ln

√1 + 4m2

z + 1√1 + 4m2

z − 1, (4.148)

and where s, t and u are Mandelstam variables

s = (p1 + p2)2 t = (p1 + p3)2 u = (p1 + p4)2 . (4.149)

As it stands Γ(4) is divergent as ε → 0. To remedy this situation, we add

yet another term to L

1

4!µ2ελ · 3λ

2

(1

ε+G1(ε, m2)

)ϕ4 , (4.150)

where G1 is an arbitrary dimensionless function of ε, analytic as ε→ 0. This

new counterterm results in an additional Feynman rule denoted by

.7.7 = −3

2µ2ελ λ

[1

ε+G1

]. (4.151)

It is added in the calculation of the new Γ(4), yielding a finite result as ε→ 0:

Γ(4)new = 2.61.5 (4.152)

= −µ2ελ

[1− 3

(−G1 + ψ(1) + 2− ln m2 − 1

3A(s, t, u)

)]+O(λ3) .(4.153)

The contribution to Γ(2) in O(λ2) can be handled in a similar way, but there

the additional Feynman rules (4.5.4) and (4.5.10) must be used. Hence we

have diagrammatically for the inverse propagator

.7.7 .7.7.7.7.7.7

.7.7.7.7.7.7.7.7 (4.154)

where the extra Feynman rules have induced to this order two new diagrams.

These are easily calculated to be (see problem)

.7.7m2

4λ2

[1

ε2+

1

ε

[ψ(1) + F1 − ln m2

]+ · · ·

], (4.155)

were we show only the poles in ε. We also have

Page 30: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

30 Perturbative Evaluation of the Path Integral: λϕ4 Theory

.7.73m2

4λ2

[1

ε2+

1

ε

[ψ(2) +G1 − ln m2

]+ · · ·

]. (4.156)

Comparing with the “double scoop” diagram of the previous section

.7.7− m2

4λ2

[1

ε2+

1

ε

[ψ(2) + ψ(1)− 2 ln m2

]+ · · ·

], (4.157)

we observe that its double pole is exactly canceled by the counterterm di-

agram (4.5.13) although the simple pole remains. This is obvious from the

point of view of diagrams since the combination .7.1 + .7.1 is by definition

finite. Adding all the diagrams, we find

1.1 = − λ2

24εp2 +

m2

2λ2

[1

ε2+

1

2ε(F1 + 3G1 − 1) + · · ·

]+O(λ3) , (4.158)

where, again, we have not shown the finite part. We are faced again with

a divergent expression as ε → 0. Lo and behold! The ln m2 term present

in the simple poles of individual diagrams have disappeared as well as the

Euler constant present in each ψ(n) but absent in the difference

ψ(n+ 1)− ψ(n) =1

n. (4.159)

In order to cancel the poles in (4.5.16) we introduce a new mass counterterm

so that the mass counterterm Feynman rule becomes

1.25 = −m2

2

[λ2

ε2+

1

ε

(λ+

λ2

4(F1 + 3G1 − 1)

)+ λ2F2 + λF1

](4.160)

where F2 is an arbitrary function of ε and m2, but finite as ε → 0. This

term is generated by the additional counterterm

m2

4ϕ2

[λ2

ε2+

1

ε

(λ+

λ2

4(F1 + 3G1 − 1)

)+ λ2F2 + λF1

]. (4.161)

This takes care of only one type of infinity in Γ(2). The other one is canceled

by adding yet another term to our ever expanding Lagrangian of the form

1

2∂µϕ∂

µϕ

[− λ2

24ε− λ2H2

(ε,m2

)], (4.162)

Page 31: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.4 Renormalization 31

H2 being arbitrary and analytic as ε → 0. Thus with all this patching up,

we have been able to eliminate the ultraviolet divergences to order λ2. It is

clear that we can continue this little game ad nauseam: calculate diagrams

to order λ3, with the original L and the counterterms (4.5.19) and (4.5.20);

then invent new counterterms which to O(λ3) are chosen to cancel the new

divergences, etc. So far in this process, the remarkable thing has been

that the counterterms needed to remove the divergences all generate new

interactions of the same type as those that were present in the original

Lagrangian; we have not been forced (to O(λ2)) to introduce counterterms

that correspond to terms of a type absent in L. If it can be shown that

this noteworthy matching feature continues to all orders in λ, we will say

that the theory is renormalizable. Here we do not attempt a proof for λϕ4

theory, but rather point out where the procedure might fail.

Consider a generic two loop diagram

1.1. (4.163)

where all external legs have been removed. The diagram is divergent due to

the various loop integrations. In accordance with our new rules for the coun-

terterms, it must be added to the three counterterm diagrams, (presented

here schematically)

.7.7 .7.7 .7.7 (4.164)

Here each • represents the lowest order counterterm vertex needed to cancel

the infinity coming from one loop diagrams. These counterterm diagrams

will therefore contain a 1ε coming from • which multiplies a ln p2 where p is

some momentum coming from the loop integration. It would therefore seem

that this would generate counterterms of the form

ln p2

ε, (4.165)

which do not correspond to any term in L because of the ln p2 residue,

which is highly nonlocal in position space. Such terms would clearly throw

a monkey-wrench in the works. This is the famous problem of overlapping

divergences. A close study of these diagrams shows that the sum of all the

diagrams shown above does not contain any poles with logarithmic residue:

they cancel against similar poles contained in the two loop diagram. We

have in fact seen as example of this miracle when the ln m2 terms canceled

from the residue of the simple pole in (4.5.16). [For more details, the reader

Page 32: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

32 Perturbative Evaluation of the Path Integral: λϕ4 Theory

is referred to the paper of ‘t Hooft and Veltman, Nucl. Phys. 448, 189

(1972).] In the proof of renormalization, it is crucial to be able to prove

that these overlapping divergences do, in fact, cancel.

Let us assume that this is so that to an arbitrarily high order in λ only

counterterms which match the original L are needed to render the theory

finite. This means that the Lagrangian which gives finite answers has the

form

Lren = L+ Lc.t. , (4.166)

where L is our original Lagrangian

L =1

2∂µϕ∂

µϕ+1

2m2ϕ2 +

λ

4!µ2εϕ4 (4.167)

and Lc.t. is the counterterm Lagrangian

Lc.t. =1

2A∂µϕ∂

µϕ+1

2m2Bϕ2 +

λ

4!µ2εCϕ4 . (4.168)

It is (by assumption, but verified to O(λ2)) exactly of the same form as

L, but with specially designed A, B and C so that the Green’s functions

generated by Lren are finite as ε→ 0. By defining new fields and parameters,

we can rewrite Lren in the form

Lren =1

2∂µϕ0∂

µϕ0 +1

2m2

0ϕ20 +

λ0

4!ϕ4

0 , (4.169)

where

ϕ0 ≡ (1 +A)1/2 ϕ ≡ Z1/2ϕ ϕ , (4.170)

m20 = m2 1 +B

1 +A= m2 (1 +B)Z−1

ϕ , (4.171)

λ0 = λµ2ε 1 + C

(1 +A)2= λµ2ε(1 + C)Z−2

ϕ , (4.172)

are called the bare field, mass and coupling constant, respectively. Note

that Lren looks exactly the same as L except for the parameters and the

field. Yet Lren leads to a finite theory while L does not. This shows that

by cleverly putting all the infinities in ϕ0, m0 and λ0, we can make the

theory finite. The infinities are then absorbed by renormalization. The

bare quantities all diverge as ε→ 0 while the (renormalized) quantities m,λ

all give finite (but so far arbitrary) values as ε → 0. The latter are to be

identified with the physical parameters of the theory. In the path integral

Page 33: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.4 Renormalization 33

formalism, one integrates over the fields; thus their rescaling by Zϕ can be

absorbed provided one rescales the source accordingly, defining a bare source

J0 = Z−1/2ϕ J , (4.173)

or a bare classical field

ϕclo = Z12ϕϕcl . (4.174)

Starting from the new Lagrangian (4.5.21) we obtain the Green’s functions

of the previous section with m and λ replaced by m0 and λ0. However, by

expressing the bare parameters in terms of the physical parameters m and

λ and by suitable renormalizing J , we obtain finite Green’s functions. For

the 1PI Green’s functions, this equality reads

Γ(n)0 (p1, · · · , pn;λ0,m0, ε) = Z−n/2ϕ Γ(n) (p1, · · · , pn;λ,m, µ, ε) , (4.175)

where the Γ(n) are finite as ε → 0. In this equation we can either regard

the bare parameters as functions of the renormalized ones or take the bare

parameters to be the independent variables; in the latter case the dressed

parameters are functions of the bare parameters. In this form we note

that the left hand side of (4.5.30) does not depend on µ while the right

hand side is explicitly as well as implicitly (through λ and m) dependent

on µ. Therefore, by differentiating (4.5.30) with respect to µ, we obtain a

differential equation that summarizes the magic of renormalization[µ∂

∂µ+ µ

∂λ

∂µ

∂λ+ µ

∂m

∂µ

∂m− n

2µ∂ lnZϕ∂µ

]Γ(n) = 0 . (4.176)

The beauty of this equation is that it only involves the renormalized Green’s

function Γ(n) which is finite as ε → 0. The various derivatives come from

the implicit dependence of Γ(n) on µ via λ and m. Define the coefficients

β

(λ,m

µ, ε

)≡ µ

∂λ

∂µ(4.177)

γd

(λ,m

µ, ε

)≡ 1

2µ∂ lnZϕ∂µ

(4.178)

γm

(λ,m

µ, ε

)=

1

2µ∂ lnm2

∂µ. (4.179)

They are analytic as ε→ 0 and dimensionless — they depend only on λ andmµ .

Page 34: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

34 Perturbative Evaluation of the Path Integral: λϕ4 Theory

On the other hand, Γ(n) has an engineering dimension equal to 4 − n +

ε(n − 2), which can be read off as the sum of its degree of homogeneity in

its dimensionful parameter, i.e.,

(µ∂

∂µ+ s

∂s+m

∂m− [4− n+ ε(n− 2)]

)Γ(n)(sp;m,λ, µ, ε) = 0 (4.180)

where we have introduced a scale s for the momenta. This equation, in

conjunction with (4.5.31) can be turned into a scaling equation for Γ(n), by

eliminating µ ∂∂µ . Taking the limit ε→ 0, we obtain

[−s ∂

∂s+ β

(λ,m

µ

)∂

∂λ+

[γm

(λ,m

µ

)− 1

]m

∂m(4.181)

−nγd(λ,m

µ

)+ 4− n

]Γ(n)(sp;m,λ, µ) = 0 . (4.182)

This equation summarizes the behavior of Γ(n) as one scales it momenta.

(An equation of this type for QED was first obtained by Gell-Mann and

Low, Phys. Rev. 95, 1300 (1954); see also Peterman and Stueckelberg

,Helv. Phys. Acta 26,499 (1953).) If we could solve it, we would know

how the Green’s functions at momenta sp are related to the same functions

at a reference p. The difficulty in solving (4.5.36) lies in the fact that the

coefficients β, γ and γm depend on two variables λ and mµ . They can be

computed explicitly order by order in perturbation theory, but they are

at the moment quite arbitrary because we have not stated what the finite

part of the counterterms is. This will be done in the next section where

various “renormalization prescriptions” will be investigated. Suffice it to

say that these coefficients depend on the way we choose the finite part of

the counterterms.

We can express the bare parameters as a Laurent series in the renormalized

parameters

= µ2ε

[a0

(λ, mµ , ε

)+∑∞

k=1

ak

(λ,mµ

)εk

](4.183)

m20 = m2

b0(λ, mµ, ε

)+∞∑k=1

bk

(λ, mµ

)εk

(4.184)

Zϕ = c0

(λ,m

µ, ε

)+∞∑k=1

ck

(λ, mµ

)εk

,(4.185)

Page 35: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.4 Renormalization 35

where a0, b0 and c0 are analytic as ε→ 0. Comparing with the counterterms

already obtained up to O(λ2), we find

a0

(λ,m

µ, ε

)= λ

(1 +

3

2λ1G1

)+O(λ3) (4.186)

b0

(λ,m

µ, ε

)= 1 +

1

2

(λF1 + λ2F2

)+ λ2H2 +O(λ3) (4.187)

c0

(λ,m

µ, ε

)= 1− λ2H2

(ε,m

µ

)+O(λ3) (4.188)

a1

(λ,m

µ

)=

3

2

λ2

16π2+O(λ3) (4.189)

b1

(λ,m

µ

)=

1

2

(λ+

λ2

4(F1 + 3G1 − 1)

)+λ2

24+O(λ3)(4.190)

b2

(λ,m

µ

)=

1

2λ2 +O(λ3) (4.191)

c1

(λ,m

µ

)= − λ

2

24+O(λ3) (4.192)

In these formulas, we remark that the a, b and c coefficients depend on mµ

only through the unknown functions F1, F2, G1, H2. This can be understood

heuristically by noting that the counterterms are used to eliminate the diver-

gences that occur at very large momentum (∼ mass) scales. There any fixed

mass parameter is not expected to play any role. Thus, as long as the coun-

terterms have order by order no finite part, we do not expect the residues

of their poles to depend on m. This is exactly what the formula (4.5.40) –

(4.5.46) reflect. This remark is at the heart of the mass-independent renor-

malization prescription we discuss in the next section. The dependence of

these coefficients on the arbitrary finite parts of the counterterms reflects

the (a priori) prescription dependence of the β- and γ-functions. It follows

that the solution of the renormalization group equation (4.5.31) is not be to

attempted before a prescription has been chosen. The technical difficulty in

finding solutions lies in the dependence of the coefficients on both λ and mµ .

However, as we shall see, there is a prescription where the coefficients become

mass independent, greatly simplifying the solutions of (4.5.31). Otherwise,

one has to solve (4.5.31) in a region where the masses can be neglected, i.e.,

where momenta are large compared with input mass parameters.

Finally, we mention that one can derive another type of renormalization

group equation, first obtained by C. Callan [Phys. Rev. D5, 3202 (1972)]

Page 36: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

36 Perturbative Evaluation of the Path Integral: λϕ4 Theory

and K. Symanzik [Comm. Math. Phys. 23, 49 (1971)]. This kind of

equation studies the variation of the Γ’s with respect to the physical mass.

The β and γ coefficients depend only on λ; the γm and µ ∂∂µ terms do not

appear but are replaced by an inhomogeneous term, which can be neglected

in the limit of small masses, or equivalently large momenta.

4.4.1 PROBLEMS

A. Compute the value of the extra counterterm diagrams (4.5.12) and (4.5.13)

including the finite part.

B. Verify that the propagator itself is finite to O(λ2) (hint — the propagator

contains one particle reducible graph).

∗∗C. In λϕ3 theory, verify that the overlapping divergences from the diagram

.7.7 (4.193)

are, in fact, canceled by the counterterm diagram that regulates the one

loop diagram

.7.7 (4.194)

4.5 Renormalization Prescriptions

In the previous section, we carried out in detail the renormalization proce-

dure in λϕ4 theory. Besides the arbitrary scale µ, the elimination of diver-

gences brought with it an extra arbitrariness reflected by the functions F1,

F2, G1, H2, · · · which constitute the finite part of the counterterms. From

the structure of the Lagrangian that gives finite answers

Lren = L+ Lc.t. , (4.195)

it is clear that the finite part of Lc.t. can be absorbed in a redefinition (or

finite renormalization) of the initial parameters appearing in L since both

L and Lc.t. have the same structure. It follows that the finite part of the

counterterms can be fixed only by defining the parameters that enter in

L. There is, however, much arbitrariness in the method used to define m,

λ and ϕ, and the choice of method is dictated by convenience, or by the

convergence properties of the perturbation theory.

In some cases it is possible to directly relate the renormalized parameters

Page 37: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.5 Renormalization Prescriptions 37

to physically measured quantities. QED is a case in point where the physical

electric charge is equated to the vertex function in the Thomson limit.

Specification of the scale at which the renormalized parameters are equated

to the relevant Green’s functions is largely arbitrary (in Euclidean space)

with one important restriction for theories that involve massless particles.

These theories give infrared divergent Green’s function for zero value of in-

put momenta. It would not be wise to choose the subtraction point at the

scale where the Green’s function diverges. Such points are to be avoided.

Later when the amplitude is continued to Minkowski space, the subtraction

scale will appear at a space-like value of input momenta and will not in-

terfere with the singularities that Green’s functions must and do have to

quality as transition amplitudes. These appear in the physical region where

at least some of the momenta are always timelike.

Let us give several examples of subtractions, also called prescriptions:

A. This first way of fixing parameters is the most common. We define

the input parameters by

Γ(2)(p , mA) ≡ p2 +m2A at p2 = 0 (4.196)

Γ(4)(p1, p2, p3, p4) ≡ −µ2ελA at pi = 0 . (4.197)

In the absence of infrared divergences which occur when m2 = 0, this pre-

scription is well-defined. We have substituted the input parameters to in-

dicate how they were defined. Note that (4.6.2) embodies two conditions

since it fixes the mass as well as the field normalization. These fix the finite

parts of the counterterms. In particular, we find that

FA1 = ψ(2)−ln m2A ; GA1 = ψ(1)−ln m2

A ; HA2 = 0 , etc. . (4.198)

Ideally, one would prefer to identify the coupling constant with Γ(4) at a

physical point where the particles are in Minkowski space and on their mass-

shell (p2M = m2).

B. One can change the subtraction point at will provided it does not

interfere with the continuation to Minkowski space or with infrared sin-

gularities. We should add that as long as the subtraction procedure is

performed on Euclidean Green functions, it will result in a spacelike sub-

traction in Minkowski space. Thus our second prescription [H. Georgi and

H. D. Politzer, Phys. Rev. D14, 1829 (1976)] is the same as A but carried

out in an arbitrary value of p:

Page 38: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

38 Perturbative Evaluation of the Path Integral: λϕ4 Theory

Γ(2)(p,mB) = p2 +m2B at p2 = M2 , (4.199)

Γ(4)(p1, p2, p3, p4) = −µ2ελB at pipj = M2

(δij −

1

4

), (4.200)

the later point chosen to that s = t = u = M2. One can, of course, choose

any value for s, t and u, and the p2 at which Γ(2) is normalized. In this

prescription, the unknown functions are now fixed to be

FB1 = ψ(2)− ln m2B ; HB

2 = 0 (4.201)

GB1 = ψ(1)− ln m2B −

∫ 1

0dx ln

[1 +

M2

m2B

x(1− x)

], etc.(4.202)

In this case the scale µ has been totally absorbed and replaced by the scale

M2, which is equally arbitrary. The numerical value of M2/m2B now clearly

becomes relevant in choosing M .

The trouble with the type of prescription outlined above is that the renor-

malization group equation of the last section is not easily solved except in

the deep Euclidean region where all masses can presumably be neglected.

Yet equating a coupling constant with the value of an amplitude at some

scale has some physical appeal even though the identification may take place

at an unphysical point. The reason is that it brings in the masses explicitly

in the calculation and allows for a ready identification of various physical

thresholds.

C. We now present a very beautiful prescription invented by ‘t Hooft and

Weinberg [Nucl. Phys. B61, 455 (1973) and Phys. Rev. D8, 3497 (1973),

respectively]. It is very simple to state and allows for a simple solution of

the “renormalization group equation” (4.5.36) of the last section. In it one

simply sets all the finite parts of the counterterms equal to zero, order by

order in the coupling, i.e.

FC1 = HC2 = GC1 = FC2 = 0 , etc. . (4.203)

Then, by comparing with Eqs. (4.5.40) – (4.5.46) of the last section, we

notice that all the a, b and c coefficients are independent of m. This pre-

scription is aptly called “mass-independent” renormalization. This mass

independence survives to arbitrarily high order; that this is, in fact, true is

not hard to understand by means of the following heuristic argument: when

the counterterms have no finite part, they just have the “bare bones” struc-

ture needed to cancel the infinite behavior at very short distances and no

more, but in this regime, i.e. at infinite momenta, all masses can presumably

Page 39: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.5 Renormalization Prescriptions 39

be neglected, provided the amplitudes are well-behaved as p→∞ — hence

the mass independence. It is true that in prescriptions A and B the mass

dependence appeared only through the finite part of the counterterms. This

enormous simplification enables us to compute the β, γ and γm coefficients

appearing in (4.5.36) in a straightforward way. For instance, we now write

λ0 = µ2ε

[λ+

∞∑k=1

ak(λ)

εk

]. (4.204)

Differentiate with respect to µ at fixed λ0 to obtain

0 = 2ε

(λ+

∞∑k=1

ak(λ)

εk

)+ µ

∂λ

∂µ

(1 +

∞∑k=1

a′k(λ)

εk

), (4.205)

with the prime denoting differentiation with respect to λ. In this formula λ

and ∂λ∂µ are analytic at ε = 0. It follows that

µ∂λ

∂µ= −2ελ− 2a1(λ) + 2λa′1(λ) , (4.206)

or taking the limit ε→ 0

β(λ) = limε→0

µ∂λ

∂µ= −2

(1− λ d

)a1(λ) , (4.207)

showing that the β-function appearing in the “renormalization group equa-

tion” depends only on λ and is determined by the residue of the simple pole

on ε. Using (4.6.12) and the fact that the residues of the various poles in ε

must vanish in (4.6.11), we find that(1− λ d

)ak+1(λ) = a′k(λ)

(1− λ d

)a1(λ) . (4.208)

The meaning of (4.6.12) is clear; a successful renormalization means that

the bare coupling does not depend on what µ is — a change in µ is accompa-

nied by a change in λ to as to leave (4.6.10) invariant. Let us now evaluate

β in perturbation theory. Using (4.5.43) we find

µ∂λ

∂µ=

3λ2

16π2+O(λ3) . (4.209)

Neglecting the terms of O(λ3), we can easily integrate (4.6.15), obtaining

Page 40: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

40 Perturbative Evaluation of the Path Integral: λϕ4 Theory

λ = λs1

1− 316π2λs ln µ

µs

, (4.210)

where λs is the value of λ at some scale µs.

It is clear from (4.6.15) that λ increases with µ. Thus, is we start with

a small λs (<< 1) at a given scale µs, the effective coupling constant will

increase with µ. In so doing, we will have to deal with larger and larger

λ and will therefore leave the domain of validity of perturbation theory:

λ << 1, or more exactly 316π2λs ln µ

µs<< 1. Thus at shorter distances, we

have to add more and more contributions to the right hand side of (4.6.15).

Thus, for λϕ4 theory, perturbation theory becomes more reliable at larger

distances, that is, in the long range properties of the interaction: the pertur-

bative approach to defining asymptotic states is to be trusted. Should the

sign of the right hand side of (4.6.15) ever be found to be negative in a field

theory, then perturbation arguments would be misleading for the definition

of the asymptotic states, but great for the short distance behavior. (We will

see that this situation occurs in Quantum Chromodynamics (QCD) which

describes the interaction among quarks — quarks are useful to describe the

short range interaction of two protons but they are not asymptotic states,

just the constituents of asymptotic states such as the proton.)

Note that the expression for β(λ) obtained here is the same as that ob-

tained via the ζ-function evaluation of determinants. This should not be too

surprising since the former method was O(~) and here we have only shown

the one loop result which is also O(~).

Let us now speculate on the possible behaviors of λ as a function of µ,

outside of perturbation theory. First of all we note that if β(λ) is given by

(4.6.15) even for large λ, then it will blow up at a scale

µ = µs exp

[16π2

3λs

], (4.211)

a very large scale if λs was small to start with. This is called the Landau

point after Landau who recognized the same behavior in QED. However,

there is no reason to believe that the one-loop contribution to β is valid for

large λ. We do not know how to calculate β for large λ, but let us present

some possible scenarios, starting from β = 0 at λ = 0, the no interaction

point:

1. β(λ) stays positive for large λ; then λ keeps increasing with the scale

in a concave or convex curve depending on the sign of β′(λ). If β(λ) blows

up for some value of λ, λ itself becomes infinite there (Landau point).

Page 41: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.5 Renormalization Prescriptions 41

2. β(λ) starts out positive for low λ then turns over and becomes nega-

tive, crossing the axis at λF :

beta(λF ) = 0, (4.212)

like

1.22 (4.213)

λF is called a fixed point because if for some reason the coupling was origi-

nally at λF , it would remain there. We can analyze the behavior of λ near

λF by expanding β about λF , yielding the equation

µ∂λ

∂µ= (λ− λF )β′(λF ) + · · · . (4.214)

We see that the sign of β′(λF ) is crucial. If β′(λF ) < 0, as in the figure,

µ∂λ∂µ is positive for λ just below λF , which drives λ to a larger value, i.e.,

towards the fixed point λF , while µ∂λ∂µ is negative when λ is above the fixed

point, therefore driving λ towards λF . We then see that λ will be driven

towards λF as µ increases: such a fixed point is called ultraviolet stable,

because λ will approach the value λF asymptotically as µ→∞, from above

or from below depending on the starting point λs which can be either above

or below λF . If there were a field theory for which β behaved as in the

curve, at very short distances λ would be more and more like λF . If λFwere small, it would mean that if we start from a small λ < λF we never

leave perturbation theory! Alternatively, if we started with λ > λF then as

the distance decreases, λ would be driven to a perturbative region. These

two situations are depicted below.

32.5 (4.215)

None of the field theories in four dimensions exhibit this behavior perturba-

tively (λF << 1).

The point λ = 0 is a fixed point for which β′(0) > 0, which means that

above it µ∂λ∂µ is positive driving λ away from it as the distance decreases.

Such a fixed point is called infrared stable. Finally, we note that for small

λ most field theories behave this way, with β(λ) starting out positive.

3. β(λ) starts out negative for low λ, decreasing in value monotonically.

This means that λ decreases monotonically with lnµ. In this case the pertur-

bation approximation becomes better at shorter distances, and λ is driven to

Page 42: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

42 Perturbative Evaluation of the Path Integral: λϕ4 Theory

zero which in this instance becomes an ultraviolet stable fixed point. Such

coupling constant behavior for low λ is exhibited by gauge theories in four

dimensions — a phenomenon known as asymptotic freedom. This will be

analyzed in detail when we study Yang-Mills gauge theories.

4. β(λ) starts out negative, turns over and becomes positive, crossing

the axis at λF .

32 (4.216)

In this case, β′(λF ) > 0 and λF is an infrared point. This means that if at

µs, λs < λF , λ will be driven towards 0, but if λs > λF it will be driven

away from λF to larger values of λ:

32 (4.217)

We can also derive the variation of the mass with µ, starting from (4.5.41)

with b0 = 1 and bk independent of mµ ; it is not hard to find that

µ∂m2

∂µ= 2λm2db1

dλ(4.218)

so that

γm(λ) = λdb1(λ)

dλ(4.219)

leading to the value in λϕ4 theory

γm(λ) =λ

16π2+

7

12

16π2

)2

+O(λ3) . (4.220)

We also get a recursion formula for the residues of the higher poles:

λdbk+1

dλ= bkλ

db1dλ− dbkdλ

(1− λ d

)a1(λ) k = 1, 2, · · · . (4.221)

Finally, we can also derive from the definition of Zϕ with c0 = 1 and ckindependent of m

µ the equation

µ∂

∂µlnZϕ = −2λ

dc1

dλ, (4.222)

leading to

Page 43: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.5 Renormalization Prescriptions 43

γd(λ) = −λdc1

dλ, (4.223)

so that in λϕ4 theory

γd(λ) =1

12

16π2

)2

+O(λ3) . (4.224)

The ck coefficients in turn obey

λdck+1

dλ= ckλ

dc1

dλ−(a1 − λ

da1

)dckdλ

, k = 1, 2, · · · . (4.225)

These recursion relations are useful in computing the residues of the higher

order poles in terms of those of the simple pole. This is where the power of

the procedure resides: if the theory is renormalizable, then many coefficients

can be calculated indirectly without the aid of Feynman diagrams.

In this ‘t Hooft–Weinberg mass independent renormalization prescription,

the renormalization group equation is easy to integrate because the m de-

pendence has dropped out of the coefficients. It now reads

[−s ∂

∂s+ β(λ)

∂λ+ (γm(λ)− 1)m

∂m+ dn − nγd(λ)

]Γ(n)(sp;m,λ, µ) = 0 .

(4.226)

Introduce scale dependent variables λ(s) and m(s) such that

s∂λ(s)

∂s= β

(λ(s)

)λ(s = 1) = λ (4.227)

s∂m(s)

∂s= m(s)

(γm(λ(s)

)− 1)

m(s = 1) = m . (4.228)

Thus (4.6.28) is easy to integrate for it becomes a first order differential

equation in s; the result is

Γ(n)(sp;m,λ, µ) = sdnΓ(n)(p; m(s), λ(s), µ) exp

−n∫ s

1

ds′

s′γd(λ(s′))

.

(4.229)

The interpretation of this equation is that under a change of scale of the

external momenta the Green’s functions scale in an unexpected way: the

coupling constant and mass scale in a nontrivial fashion and the Green’s

Page 44: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

44 Perturbative Evaluation of the Path Integral: λϕ4 Theory

functions develop beside the engineering dimension dn and anomalous di-

mension γd for each outside leg.

Suppose m had been set equal to zero initially. Then the classical theory

would have been invariant under dilatations (and conformal transformations

— see Chapter 1). This is no longer true in quantum theory where the reg-

ularization technique introduces a scale, either via a high momentum cutoff

or via the µ-parameter of dimensional regularization, which breaks dilata-

tion invariance [see S. Coleman’s lecture in Aspects of Symmetry, Cambridge

University Press, Cambridge, 1985].

Also this shows that the Green’s functions at scaled momenta give a be-

havior ruled by λ(s) and m(s) which therefore govern the physics at large

scales.

It is interesting to investigate the behavior of the Green’s functions at

large s. Suppose the theory has an ultraviolet stable fixed point at λ = λF .

At large scales λ will be driven to λF . Hence λm and γd will be driven to

γm(λF ) and γd(λF ), respectively, so that the solution of (4.6.21)

sm(s) = me∫ s1 γm(λ(s))dlns (4.230)

can be integrated to

m(s)→ ms[−1+γm(λF )] (4.231)

if we assume that the integral is dominated by the large s range. A similar

assumption about the integration over λd leads to

Γ(n)(sp;m,λ, µ)→ s(dn−nγd(λF ))Γ(n)(p;msγm(λF )−1, λF , µ) . (4.232)

Thus, if 1−γm(λF ) is positive the mass can be neglected altogether at large

scales and also γd(λF ) appears indeed as an anomalous dimension.

We see from (4.6.23) that the naive expectation that the masses decouple

from the theory at large momenta depends essentially on the integral over

the anomalous dimensions which we write as

∫ λ(s)

λdλ′

γm(λ′)

β(λ′). (4.233)

This form suggests that unless γm(λ) and β(λ) have simultaneous zeros, the

greater contributions to the integral will come from the fixed points, which

“kind of” justifies our earlier assumption.

If the ultraviolet stable fixed point is at λF = 0 (as in Yang-Mills theories),

Page 45: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.5 Renormalization Prescriptions 45

then all is well because the large momentum behavior of the theory is given

by perturbation theory, that is γm(0) = 0.

In λϕ4 theory we can integrate the various equations by using the lowest

perturbative results found for γm and γd. The results are

m(s) =m

s

(λ(s)

λ

)1/3

e7(λ(s)−λ)/(576π2) (4.234)

for the scale dependent mass and

Γ(n)(sp) ∼ sdn−n36

λ(s)−λ16π2 Γ(n)(p) (4.235)

for the anomalous dimension. These results are only valid for small mass

scales because as we have seen, the perturbative β function is positive and,

after a certain scale is reached perturbation theory loses its validity.

4.5.1 PROBLEMS

A. Given β(λ) = µ∂λ∂µ , describe the behavior of the hypothetical field theories

for which

1. β(λF ) = β′(λF ) = 0 (4.236)

2. β(λk) = 0 λk = λF +a

kk = 1, 2, · · ·∞ (4.237)

B. Given

m20 = m2

(1 +

∞∑n=1

bn(λ)

εn

)≡ m2Zm , (4.238)

show that

γm = −1

2µ∂ lnZm∂µ

= λdb1(λ)

dλ(4.239)

and that

λdbn+1

dλ= bnλ

db1dλ

+1

2

dbndλ

β(λ) n = 1, 2, · · · . (4.240)

If bn(λ) = bn, n+1λn+1 +O(λ2), derive a formula relating bn, n+1 to b1,2.

Page 46: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

46 Perturbative Evaluation of the Path Integral: λϕ4 Theory

C. Given

ϕ0 = ϕ

(1 +

∞∑n=1

cn(λ)

εn

)1/2

≡ ϕZ1/2ϕ , (4.241)

show that

γd(λ) =1

2µ∂ lnZϕ∂µ

= −λdc1

dλ(4.242)

and that

λdcn+1

dλ= cnλ

dc1

dλ+

1

2

dcndλ

β(λ) . (4.243)

D. Verify the renormalization group equation in λϕ4 theory using the results

from perturbation theory for β, γm and γd.

4.6 Prescription Dependence of the Renormalization

Group Coefficients

In the previous section we computed the value of the β and γ coefficients

using the ‘t Hooft–Weinberg mass independent renormalization prescriptions

to lowest nontrivial order in λ. We have drawn conclusions of physical

import from their behavior — yet the alert reader might wonder to what

extent these coefficients depend on the prescription for renormalization: for

a general prescription of type A or B these coefficients will, in general,

depend on masses through the finite parts of the counterterms.

Starting from Eqs. (4.5.37) – (4.5.39) we find by differentiating with re-

spect to µ at fixed bare parameters that

0 = 2ε[a0 +

a1

ε

]+µ

∂λ

∂µ

[a′0 +

a′1ε

]+

(∂m

∂µ− m

µ

)[a0 +

a1

ε

]+ higher poles ,

(4.244)

from which we deduce that

µ∂λ

∂µ= ε

(−2λ+

3λ2

16π2G1

)+

3λ2

16π2+

3

2

m

µ

λ2

16π2G1 +O(λ3) . (4.245)

In these formulae, the prime and dot denote differentiation with respect

to λ and mµ , respectively. Comparing with (4.6.15), we see the explicit

Page 47: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.6 Prescription Dependence of the Renormalization Group Coefficients 47

prescription dependence coming through G1. In particular, we note that the

prescription dependence enters in the lowest order because of the mass. In

deriving Eq. (4.7.2) it becomes evident that the order λ3 in (4.7.2) depends

directly on G1 without multiplicative mass factors. In a similar way we can

show that

γm = µ∂ lnm

∂µ=

(1

2

λ

16π2+

1

4

m

µ

λ

16π2F1

)+ε

2

λ

16π2F1 +O(λ3) , (4.246)

as well as

γd = µ∂

∂µlnZ1/2

ϕ =1

48

16π2

)2

− m

16π2

)2

H2 (4.247)

2

16π2

)2

H2 +O(λ3) . (4.248)

We let ε→ 0 and finally obtain

β

(λ,m

µ

)=

3λ2

16π2+

3

2

m

µ

λ2

16π2G1 +O(λ3) , (4.249)

γm

(λ,m

µ

)=

1

2

λ

16π2+

1

4

m

µ

λ

16π2F1 +O(λ2) , (4.250)

γd

(λ,m

µ

)=

1

48

16π2

)2

− m

16π2

)2

H2 +O(λ3) . (4.251)

These formulae indicate that when masses are present, the prescription de-

pendence enters in the lowest order. Then it is only when the mass can be

neglected that we can say that the lowest order renormalization equation co-

efficients are independent of the prescription. Typically, when prescriptions

of type A or B are used, one solves the renormalization group equation in a

regime where the mass can be neglected. If µ is the renormalization point,

we can take mµ to be exceedingly small by choosing a very large µ.

Suppose we have two renormalization prescriptions. They must be re-

lated to one another by a finite renormalization since they differ only in

the definition of the renormalized parameters. Hence the parameters in one

prescription will be related to those in the other as follows

λ′ = T

(λ,m

µ

)= λ+O(λ2) , (4.252)

Page 48: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

48 Perturbative Evaluation of the Path Integral: λϕ4 Theory

Z ′m

(λ′,

(m

µ

)′)= Zm

(λ,m

µ

)U

(λ,m

µ

); U = 1 +O(λ) ,(4.253)

Z ′ϕ

(λ′,

(m

µ

)′)= Zϕ

(λ,m

µ

)V

(λ,m

µ

); V = 1 +O(λ3) .(4.254)

In particular, it follows that, where A is some function of λ and mµ ,

β′(λ′,

(m

µ

)′)= β

(λ,m

µ

)A′ +

(m

µ

)(γm − 1) A . (4.255)

In the deep Euclidean region this reduces to an equation of the form

β′(λ′) = β(λ)A′(λ) , (4.256)

which shows that a fixed point at λ = λF results in a fixed point at

λ′F = A(λF ) 6= λF . Note that this result is not perturbative. Hence, as

can be expected, the presence (or absence) of a fixed point is prescription

independent. It can also be shown that the sign of the first derivative of β

at a fixed point is prescription independent (see problem).

The other renormalization group parameters γm and γ also have some

features which are prescription independent, notably their numerical value

at a fixed point in the deep Euclidean region.

4.6.1 PROBLEMS

A. Verify Eqs. (4.7.2), (4.7.3) and (4.7.4).

B. Verify to lowest order the renormalization group equation for Γ(2) and

Γ(4), keeping the finite parts F1, F2, H2 and G1.

C. Show that the sign of dβdλ at a fixed point λF is prescription independent,

neglecting masses.

∗D. Relate γ′(λ′) and γ′m(λ′) to γ(λ), γm(λ) and β(λ) where the prime and

unprimed system refers to two mass independent renormalization schemes

[F , G and H functions are first taken independent of mµ , but can have

numerical values]. Show that the value of γ and γm at λF is prescription

independent.

Page 49: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.7 Continuation to Minkowski Space: Analyticity 49

4.7 Continuation to Minkowski Space: Analyticity

At this stage we have obtained the finite Green’s functions at the price of

introducing an arbitrary scale. However, we know from the renormalization

group that if we alter this scale, nothing happens to the Green’s functions

because the change is compensated by a concomitant change in the renor-

malized parameters and field. There now remains to express the Green’s

functions in Minkowski space in order to make contact with reality.

This is achieved by means of analytic continuation. Consider an Euclidean

Green’s function which depends on momenta p1, · · · , pN . We first change all

the time components of the p by making them imaginary

p = (p0, pi)→ p = (p0 = ip0, pi = pi) . (4.257)

As an example let us examine what happens to the propagator. According

to the above

1

p2 +m2→ 1

−p2 +m2, (4.258)

since we are using the Minkowski metric g00 = −gii = +1. This replacement

is not entirely satisfactory because the Minkowski space expression develops

a pole at p2 = m2, i.e., when

p0 = ±√~p · ~p+m2 . (4.259)

The continuation process can be regarded as proceeding from the imaginary

to the real axis in the p0 plane in a clockwise fashion: indeed the ability

to perform this continuation rests on the avoidance of any pole. It follows

that the poles in p0 must be taken to be just below (above) the positive

(negative) real axis, that is the continuation is from

1

p2 +m2to

−1

p2 −m2 + iε, (4.260)

where ε > 0, and the limit ε→ 0+ is to be taken at the end of all calculations.

In more complicated cases the poles will be chosen so that their locations do

not interfere with the clockwise rotation of the imaginary time axis into the

real time axis. In this −iε prescription we recognize the familiar device used

to make the path integral convergent which we discussed earlier. Another

way to look at this is to compare the Feynman rules in Minkowski and

Euclidean spaces, say for ϕ4 theory:

BIG MESS

Page 50: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

50 Perturbative Evaluation of the Path Integral: λϕ4 Theory

1.02

1

p2+m2 (Euclidean)

ip2−m2+iε

(Minkowski)

(4.261)

11

−λ (Euclidean)

−iλ (Minkowski)

(4.262)

loop

integration:

d4k

(2π)4 (Euclidean)

d4k(2π)4 (Minkowski) .

(4.263)

Consider a Feynman diagram with L loops, V vertices and I internal lines.

The difference between its computation in Minkowski and Euclidean space

will be: a factor of (−i) for each vertex and propagator, and a factor of i

for each loop because d4k = id4k. In addition, m2 in the Euclidean Green’s

function is replaced by m2 − iε in the corresponding Minkowski space ex-

pression. Thus, we arrive at

G(n)M (p1, · · · , pn;m2) = (i)L+V+I(−1)IG

(n)E (p1 = p1, · · · pn = pn;m2 − iε) .

(4.264)

where we replace in the Euclidean function the momenta by their Minkowski

space continuation, i.e., p = (p0, pi) is replaced by p = (ip0, pi). Using the

topological relation L = I − V + 1, (V 6= 0), we obtain

G(n)M (p.m2) = (i)G

(n)E (p = p;m2 − iε) , (4.265)

valid for any Feynman diagram with V 6= 0. The exception to this rule is

the original propagator expression (V = 0) for which i is replaced by −i, as

can be seen from (4.8.4) by putting L = V = 0, I = 1.

As an example of the procedure, let us consider the four-point function in

Minkowski space. We have

Γ(4)M (p1, p2, p3, p4) = i

[−λ− λ2

2 · 16π2

∫ 1

0dx ln

[m2 − iε− sx(1− x)

m2 − iε+M2x(1− x)

](4.266)

+t- and u-channels

]. (4.267)

Page 51: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.7 Continuation to Minkowski Space: Analyticity 51

In the above we have multiplied the Euclidean expression by i, replaced p2

by −p2, m2 by m2 − iε. The subtraction was performed on the Euclidean

Green’s function at pipj = M2(δij − 1/4) so that Γ(4)E → −λ at this sym-

metric point (prescription B). Note that the subtraction point appears as

a parameter and it does not get changed in the continuation process. The

−iε in the denominator is not necessary since it can never vanish. However,

the numerator does suffer a change of sign for some value of s = (p1 + p2)2.

When this occurs the logarithm develops a cut in the complex s-plane. Con-

sider the argument of the log

F (s, x) ≡ m2 − iε− sx(1− x) . (4.268)

Now x(1 − x) is positive definite and varies between 0 and 1/4. Thus the

least value of s for which F vanishes is

s0 = 4m2 (4.269)

where x(1 − x) assumes its largest value. At this point Γ(4) develops a

branch point. Traditionally, one attaches to it a cut in the complex s-plane

extending from 4m2 to +∞ along the positive real s-axis. In a similar way

the u- and t-channel contributions give cuts starting at u0 and t0 = 4m2. In

view of the relation

s+ t+ u = 4m2 , (4.270)

there cuts are not all independent. These branch points can be physically

understood if we interpret Γ(4) as the scattering amplitude for two particles

with momenta p1 and p2 to scatter into two particles of momenta p3 and p4:

p1 + p2 → p3 + p4 . (4.271)

Assuming that incoming and outgoing particles are on their “mass-shells”

p2a = m2 a = 1, 2, 3, 4 , (4.272)

it is easy to see that s0 = 4m2 corresponds to the two particles having their

minimum energies Ea = m. It is only beyond s0 that the two particles have

enough energy to scatter nontrivially into two others. Consequently, s0 is

called a physical two-particle threshold. Below it Γ(4) is a real function of

its argument, but it develops an imaginary part for s > s0.

To recapitulate: by continuing Γ(4) in Minkowski space, we see a nontrivial

analytical structure emerging; this is, or course, the structure demanded by

unitarity and causality which will enable us to regard the Minkowski space

Green’s functions as transition amplitudes.

Page 52: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

52 Perturbative Evaluation of the Path Integral: λϕ4 Theory

Another example of a nontrivial analytic structure emerging as a result

of continuation in Minkowski space is given by the “setting sun” diagram.

There it is easy to see that the best way to find the branch points is to look

at the argument of the logarithm in the parametric integral. In this case the

argument of interest is

A = −y(1− y)p2 +m2

(1− y +

y

x(1− x)

). (4.273)

It will vanish when

p2 = m2

(1

y+

1

x(1− x)(1− y)

), (4.274)

and the location of the branch point will be given by the least such value of

p2. In order to find its value we have to extremize the parametric expression

that multiplies m2. In general for the two-point function, branch points will

appear at the minimum value of p2 for which

p2 = m2f(x1, x2, · · · , xn) (4.275)

where x1 · · ·xN are the Feynman parameters needed for an N -loop diagram.

The branch point is then located at

p2 = m2f(x0

1, · · · , x0N

), (4.276)

where the points x0i are determined by

∂f

∂xi= 0 at xi = x0

i , (4.277)

[it must be checked that xi = x0i is, in fact, a minimum]. Such equations

are called the Landau equations after Landau who introduced a systematic

procedure to hunt for the branch points of Feynman diagrams. Applying

this procedure to our case, we find from (4.8.14) that the minimum occurs

at

x =1

2y =

1

3, (4.278)

so that the branch point is located at

p2 = 9m2 . (4.279)

Page 53: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.8 Cross-Sections and Unitarity 53

If you recall that the form of the “setting sun”, 1.4 we see that it corresponds

to the minimum energy needed to excite three particles so it is called the

three-particle threshold.

Hence the propagator has in Minkowski space the following singularity

structure: — a pole at p2 = m2 (appropriately displaced by the iε prescrip-

tion, — a branch point at p2 = 9m2 with a cut taken traditionally along

the real positive p2 axis extending to p2 = +∞. When higher order of λ

are included, it is expected that branch points at higher values of p2 will

be encountered. This singularity structure is (of course) consistent with the

interpretation of G(2) as a propagator.

4.7.1 PROBLEMS

∗A. Using diagram and physical arguments, find the location of branch points

in Γ(4) including O(λ4).

∗B. Repeat problem A for the propagator.

∗C. Show that Γ(4) satisfies a dispersion relation that expresses its real part

in terms of its imaginary part. [Use only its perturbative value up to O(λ2).]

4.8 Cross-Sections and Unitarity

We are now almost at the end of the road. We are about to identify the

Minkowski space Green’s functions with transition amplitudes. However,

not all functions can be transition amplitudes for they must satisfy certain

requirements, notably those of unitarity and causality. As you might expect,

the Green’s functions of the previous sections are acceptable candidates.

In order to state precisely the requirements, let us review the S-matrix

formalism and apply it to λϕ4 theory.

Suppose it were possible to define states far away from the region of

interaction; in particular in the very distant past or future. Such a concept

clearly makes sense in the case of short range forces, as for instance in weak

and strong interactions. When long range forces are involved, the concept

is trickier and special care must be exercised in the definition of such states.

Schematically let the states be described by kets |α;±T 〉 where T is a very

large time and α represents a complete set of observables. These states obey

completeness and orthogonality

Page 54: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

54 Perturbative Evaluation of the Path Integral: λϕ4 Theory

∑α

|a,±T 〉〈±T, α| = 1 (4.280)

〈α,±T |β,±T 〉 = δαβ . (4.281)

If the system were a harmonic oscillator, α would denote the occupation

number, etc. It is crucial to note that these relations hold only for a given

time and therefore involve no dynamics, only kinematics. If, when T is large,

the interaction can be turned off (short range forces) the states should be

easy to recognize because they diagonalize the unperturbed Hamiltonian.

In λϕ4 theory, when m2 6= 0, there is no difficulty in recognizing these

states. They are made out of the one-particle Wigner states labeled by

m and ~p with the energy identified with +√~p2 +m2. If we adopt a more

relativistic-looking notation and denote those states by |p〉, where p stands

for the momentum four vector, they are required to satisfy

∫d4p

(2π)3|p〉θ(p0)δ(p2 −m2)〈p| = 1 , (4.282)

〈p|p′〉 = 2(2π)3√~p2 +m2 δ(~p− ~p′) . (4.283)

Then any multiparticle state will be a superposition of noninteracting one-

particle states:

|α,±∞〉 ∼ |p1, p2, · · · , pn〉 = |p1〉 ⊗ |p2〉 · · · ⊗ |pn〉 . (4.284)

In λϕ4 theory there is some justification for believing these states describe

the asymptotic states because the large scale behavior of the coupling is

such that the free Feynman propagator accurately describes signal prop-

agation, and we know that ∆F does propagate one particle states of the

type described above. [Here we are a bit cavalier since strictly speaking ∆F

propagates both positive and negative energy states.] We note in passing

that if the large distance behavior of the coupling constant were such that it

grew with distance, the identification of asymptotic states would have had

to be made only for constructs that can escape this formidable force. This

is supposedly the case for QCD where quarks are subject to such a force.

Hence, they cannot serve as asymptotic states. However, this force only

attacks objects with color and allows for the definition of asymptotic states

that have no color (hadrons).

Page 55: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.8 Cross-Sections and Unitarity 55

A question of physical interest is the computation of the transition am-

plitude

Tαβ = 〈α,+∞|β,−∞〉 . (4.285)

With Heisenberg we define an S-matrix with the property

|β,+∞〉 = S|β,−∞〉 . (4.286)

Its job is to contain all the dynamical information about the evolution of

the physical states in time. Completeness of the states at +∞ and −∞

1 =∑β

|β,+∞〉〈+∞, β| =∑β

S|β,−∞〉〈−∞, β|S† (4.287)

= SS† , (4.288)

implies that S is unitary. [You can show that S†S = 1 as well.] In physical

terms the unitarity of S means that the system cannot disappear into noth-

ing [black holes?]. Most of the time nothing will happen when states are

scattered — they are much more likely to miss each other than to interact.

For this reason we set

S = 1 + iR , (4.289)

with R containing the interesting information. Hence, it follows that

Tαβ = 〈α,−∞|S†|β,−∞〉 (4.290)

= δαβ − i〈α,−∞|R†|β,−∞〉 . (4.291)

Since the interaction is Lorentz invariant, it is convenient to take this into

account and write

Tαβ = δαβ − i(2π)4δ(4)(pα − pβ)〈α,−∞|T †|β,−∞〉 , (4.292)

where pα(pβ) is the sum of the momenta in the final (initial) state. The

transition probability over all of space-time is then given by

ωαβ =[(2π)4δ(4)(pα − pβ)

]2〈α,−∞|T †|β,−∞〉〈β,−∞|T |α,−∞〉 .

(4.293)

Page 56: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

56 Perturbative Evaluation of the Path Integral: λϕ4 Theory

The square of the δ-function is quickly understood since (2π)4δ(4)(0) is the

volume of space-time [as can be seen by putting the system in a box]. It

follows that the transition probability per element of space-time is

Ωαβ = (2π)4δ(4)(pα − pβ)|〈α|T |β〉|2 . (4.294)

This form is valid for states that satisfy (4.9.2). In our case the momentum

states are not normalized to 1 but according to (4.9.4). Hence we find, after

dividing by the normalization, that

Ω(pα|pβ) ≡ Ωpαpβ =(2π)4δ(4)(pα − pβ)

(2Eα)(2Eβ)|〈α|T |β〉|2 , (4.295)

where Eα(Eβ) stands for the product of the energies in the α- (β-) state,

each energy being given by

Ei =√~p2i +m2 . (4.296)

In scattering experiments one is usually interested in the scattering cross-

section of two particles (target and projectile) into many. It is given by

dσ(a+ b→ 1 + 2 + · · ·+N) =1

vabΩ(pa, pb|p1, p2, · · · , pN )

d3p1d3p2 · · · d3pN(2π)3N

,

(4.297)

where vab is the relative velocity of particles a and b. For equal mass particles

it is conveniently expressed as

vab =

√(pa · pb)2 −m4

EaEb. (4.298)

Putting it all together we obtain

dσ(a+ b→ 1 + 2 + · · ·N) =(2π)4δ(4) (pa + pb − p1 − · · · − pn)

4√

(pa · pb)2 −m4

(4.299)

×∣∣∣⟨papb|T |p1 · · · pN

⟩∣∣∣2 N∏i=1

d3pi2(2π)3Ei

.(4.300)

Note that the measure d3p2E is relativistic since

Page 57: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.8 Cross-Sections and Unitarity 57

d3p

2E= d4pθ(p0)δ

(p2 −m2

). (4.301)

A special case of interest is elastic scattering where N = 2. Define the center

of mass frame where

~pa + ~pb = ~p1 + ~p2 = 0 . (4.302)

Then simple kinematics yields

dσ(a+ b→ 1 + 2) =|T |2

64π2sdΩ , (4.303)

where dΩ = dϕd(cos θ), θ being the angle between the ingoing and outgoing

directions:

1.51.5 (4.304)

and s is Mandelstam’s variable

s = (pa + pb)2 . (4.305)

As you might have expected, the renormalized Green’s functions will be

identified with the matrix elements of T . Hence it is important to translate

the requirements of unitarity into conditions on T and verify they they are

met by our identification.

From the unitarity condition on S, it follows that

R− R† = iR†R = iRR† . (4.306)

This operator equation summarizes the restrictions on R due to unitarity.

For instance, let us take its matrix element between two particle states,

labeled |1, 2〉 and 〈3, 4|:

〈3, 4|R|1, 2〉 − 〈3, 4|R†|1, 2〉 = i〈3, 4|RR†|1, 2〉 . (4.307)

It is not hard to see that when the external particles are spinless

langle3, 4|R†|1, 2〉 = 〈1, 2|R†|3, 4〉 . (4.308)

Using

Page 58: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

58 Perturbative Evaluation of the Path Integral: λϕ4 Theory

〈1, 2|R†|3, 4〉 = 〈3, 4|R|1, 2〉∗ , (4.309)

we arrive at

2=〈3, 4|R|1, 2〉 = 〈3, 4|RR†|1, 2〉 . (4.310)

Now the right hand side of this equation can be rewritten by introducing a

set of intermediate states. Since we want to limit ourselves to interactions

that involve an even number of states (invariance under ϕ → −ϕ), the

lowest energy intermediate state is the two particle state |a, b〉 = |a〉|b〉.Hence, using (4.9.3), we arrive at

2=〈3, 4|R|1, 2〉 =

∫d4a d4b

(2π)6θ(a0)θ(b0)δ(a2 −m2)δ(b2 −m2)(4.311)

×〈3, 4|R|a, b〉〈a, b|R†|1, 2〉 · · · (4.312)

where · · · denotes the sum over 4-, 6-, · · · particle intermediate states. In

terms of the T -matrix of (4.9.12), this equation becomes

2=〈3, 4|T |1, 2〉 =

∫d4a d4b

(2π)2θ(a0)θ(b0)δ(a2 −m2)δ(b2 −m2)δ(a+ b− 1− 2)(4.313)

×〈3, 4|T †|a, b〉 a, b|T |1, 2〉+ · · · . (4.314)

Since both a and b are on their mass-shells, the right hand side will be

non-zero only when the |1, 2〉 initial state has enough energy to produce the

|a, b〉 intermediate state, that is, when s = (p1 + p2)2 ≥ 4m2. Thus, we see

that as a consequence of unitarity and completeness the T matrix elements

are real for s < 4m2 and acquire an imaginary part after the two-particle

threshold is crossed, and for all other higher thresholds there is a further

contribution to the imaginary part of the matrix element of T .

Let us now compare this with the four-point function obtained from per-

turbation theory.

Γ(4) = −i

[λ +

λ2

2 · 16π2

∫dx ln

[m2 − iε− sx(1− x)

m2 +M2x(1− x)

](4.315)

+(s→ t) + (s→ u) +O(λ3)

]. (4.316)

We have seen that the parametric integral does develop an imaginary part

Page 59: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.8 Cross-Sections and Unitarity 59

and has a branch point at s = 4m2, all of it consistent with the unitarity

equation (4.9.29). So we are led to identify Green’s functions with (−i)times the T matrix. In this case

Γ(4)(1, 2, 3, 4) = −i〈3, 4|T |1, 2〉 . (4.317)

We can easily check that this is true: we have already calculated the imag-

inary part of Γ(4); it is or order λ2 and appears only for s ≥ 4m2. On

the other hand, we can calculate it from the unitarity equation (4.9.29), by

putting in its right hand side the lowest perturbation vertex; this gives to

O(λ4)

2=(iΓ(4)

)=

λ2

(2π)2

∫d4a d4bθ(a0)θ(b0)δ(a2−m2)δ(b2−m2)δ(1+2−a− b).

(4.318)

We leave its verification to the reader. This yields only the lowest order

contribution to the imaginary part — the 4-, 6-, · · · particle thresholds will

add to the imaginary part but in higher orders of λ.

With this identification between Green’s functions and scattering ampli-

tudes there emerges a new way of computing the imaginary part of diagrams

using the unitarity equation. It is useful in perturbation theory because of

the quadratic nature of the right hand side of (4.9.29); for if one has com-

puted a, b|T |1, 2〉 to order λk, it will give the imaginary part to order

λk+1. This remark is important because of the optical theorem which re-

lates the imaginary part of the forward scattering amplitude to the total

cross section. This theorem is simply obtained by putting |3, 4〉 = |1, 2〉in (4.9.29) and comparing the right hand side with the integrated form of

(4.9.22).

We can arrive at a diagrammatic representation of the unitarity con-

straint, if we recall the meaning of the Feynman propagator. We have

∆F (x) =i

(2π)4

∫d4k

eikx

k2 −m2 + iε(4.319)

= θ(x0)

∫d4k

(2π)3θ(−k0)δ(k2 −m2) (4.320)

+θ(−x0)

∫d4k

(2π)3θ(+k0)δ(k2 −m2) , (4.321)

which expresses the fact that ∆F includes propagation of both positive and

negative energy states depending on the sign of x0. Now if we invent a new

Page 60: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

60 Perturbative Evaluation of the Path Integral: λϕ4 Theory

set of rules where the full propagator ∆F is replaced by θ(k0)δ(k2−m2), we

can arrive at a pictographic way to compute imaginary parts and therefore

total cross-sections. This new rule would apply only in Minkowski space, of

course. While we have the old rule

⇒ i

p2 −m2 + iε, (4.322)

invent a new one for the cut propagator

.6.6→ (2π)θ(k0)δ(p2 −m2) . (4.323)

Note that the cut propagator is not symmetric, as the shading shows. The

reason is that since we have to compute TT † on the right hand side of

(4.9.29), portions of the diagram to the left of the cut must correspond to

the conjugate of diagrams to the right of it although they may be different.

The interested reader is referred to “Diagrammar” by ‘t Hooft and Veltman,

in Particle Interactions at Very High Energies, part B, D. Speiser et al., eds.

(Plenum Press, New York, 1974), for more details.

Thus our equation (4.9.32) would read diagrammatically

11 = 11 (4.324)

The end result is that one can derive general cutting equations that express

the imaginary part of diagrams in terms of the sum of all the possible cuts.

[This is not as bad as it sounds since many cut diagrams vanish by energy

conservation for cutting a Feynman propagator restricts the energy flow to

one direction.]

This concludes our study of perturbative λϕ4 theory.

4.8.1 PROBLEMS

A. Show that dσdΩ = 1

s|T |264π2 starting from Eq. (4.9.19).

vskip .5cmB. Show that for elastic scattering of spinless particles

〈3, 4|S|1, 2〉 = 〈1, 2|S|3, 4〉 . (4.325)

C. Compute =iΓ(4) by means of the unitarity equation and compare with

the result previously obtained from perturbation theory.

Page 61: Perturbative Evaluation of the Path Integral: 4 Theoryramond/Chapter4_CUP.pdf4 Perturbative Evaluation of the Path Integral: ’4 Theory In the following, we proceed with the conventional

4.8 Cross-Sections and Unitarity 61

D. Show that in general for ε > 0

1

x+ iε= −iπδ(x) + P

(1

x

), (4.326)

where P(

1x

)is the Cauchy principal part of 1

x defined by

P(

1

x

)= −i

∫ +∞

−∞dy eixy [θ(y)− θ(−y)] . (4.327)

∗E. Compute the imaginary part of the setting sun by using the unitarity

equation

= [i.4.4 ] = 11 (4.328)

∗∗F. Given LE = 12∂µϕ∂µϕ+ 1

2m2ϕ2 + h

3!ϕ3 + λ

4!ϕ4, a) derive the Feynman

rules b) find the change of m, h, and λ with scale of O(~) c) solve the

equations of b) and interpret the result physically. You may use any

renormalization prescription, but the mass independent prescription is

strongly advised.