<<

BASICS OF QCD



Davison E. Sop er

Institute of Theoretical Science

University of Oregon, Eugene, OR 97403

ABSTRACT

This is an intro duction to the use of QCD p erturbation theory, em-

phasizing generic features of the theory that enable one to separate

short-time and long-time e ects. I also cover some imp ortant classes

of applications: -p ositron to , deeply in-

elastic , and hard pro cesses in -hadron collisions.



Supp orted by DOE Contract DE-FG03-96ER40969.

1 Intro duction

A prediction for exp eriment based on p erturbative QCD combines a particular

calculation of Feynman diagrams with the use of general features of the theory.

The particular calculation is easy at leading order, not so easy at next-to-leading

order, and extremely dicult b eyond the next-to-leading order. This calculation

of Feynman diagrams would b e a purely academic exercise if we did not use certain

general features of the theory that allow the Feynman diagrams to b e related to

exp eriment:

 the group and the running coupling;

 the existence of infrared safe observables;

 the factorization prop erty that allows us to isolate hadron structure in parton

distribution functions.

In these lectures, I discuss these structural features of the theory that allowa

comparison of theory and exp eriment. Along the way,we will discover something

ab out certain imp ortant pro cesses:

+

 e e annihilation;

 deeply inelastic scattering;

 hard pro cesses in hadron-hadron collisions.

By discussing the particular along with the general, I hop e to arm the reader with

information that sp eakers at research conferences take to b e collective knowledge|

knowledge that they assume the audience already knows.

Now here is the disclaimer. We will not learn how to do signi cant calculations

in QCD p erturbation theory. Three lectures is not enough for that.

I hop e that the reader may b e inspired to pursue the sub jects discussed here

1

in more detail. A go o d source is the Handbook of Perturbative QCD by the

CTEQ collab oration. More recently, Ellis, Stirling, and Webb er have written an

2

excellent book that covers most of the sub jects sketched in these lectures. For

the reader wishing to gain a mastery of the theory, I can recommend the recent

3 4 5

b o oks on quantum eld theory by Brown, Sterman, Peskin and Schro eder, and

6

Weinb erg. Another go o d source, including b oth theory and phenomenology,is

7

the lectures in the 1995 TASI pro ceedings, QCD and Beyond.

2 Electron- Annihilation and Jets

In this section, I explore the structure of the nal state in QCD. I b egin with the

+

kinematics of e e ! 3 par tons, then examine the b ehavior of the

+

for e e ! 3 par tons when two of the parton momenta b ecome collinear or one

parton b ecomes soft. In order to illustrate b etter what is going on,

Iintro duce a theoretical to ol, null-plane co ordinates. Using this to ol, I sketch

a space-time picture of the singularities that we nd in momentum space. The

singularities of p erturbation theory corresp ond to long-time . We see that

the structure of the nal state suggested by this picture conforms well with what

is actually observed.

I draw the distinction b etween short-time physics, for which p erturbation the-

ory is useful, and long-time physics, for which the p erturbative expansion is out

of control. Finally, I discuss how certain exp erimental measurements can prob e

the short-time physics while avoiding sensitivity to the long-time physics.

+

2.1 Kinematics of e e ! 3 Partons

+

Figure 1: for e e ! q qg .

p

+

Consider the pro cess e e ! q qg , as illustrated in Fig. 1. Let s b e the total



in the c.m. frame and let q b e the virtual (or Z b oson) momentum,





so q q = s. Let p b e the momenta of the outgoing partons (q; q; g) and let



i

0

E = p b e the of the outgoing partons. It is useful to de ne energy

i

i

fractions x by

i

2p  q E

i i

p

= : (1) x =

i

s=2 s

Then

0

Energy conservation gives

P

X

2( p )  q

i

x = =2: (3)

i

s

i

Thus only two of the x are indep endent.

i

Let  b e the angle b etween the momenta of partons i and j . We can relate

ij

these angles to the momentum fractions as follows:

2 2

2p  p =(p +p ) =(qp ) =s2qp ; (4)

1 2 1 2 3 3

2E E (1 cos  )=s(1 x ): (5)

1 2 12 3

Dividing this equation by s=2 and rep eating the argument for the two other pairs

of partons, we obtain three relations for the angles  :

ij

x x (1 cos  ) = 2(1 x ) ;

1 2 12 3

x x (1 cos  ) = 2(1 x ) ;

2 3 23 1

x x (1 cos  ) = 2(1 x ) : (6)

3 1 31 2

We learn two things immediately. First,

x < 1: (7)

i

Second, the three p ossible collinear con gurations of the partons are mapp ed into

x space very simply:

i

 ! 0 , x ! 1;

12 3

 ! 0 , x ! 1;

23 1

 ! 0 , x ! 1: (8)

31 2

The relations 0  x  1, together with x =2x x, imply that the

i 3 1 2

allowed region for (x ;x ) is a triangle, as shown in Fig. 2. The edges x =1 of

1 2 i

the allowed region corresp ond to two partons b eing collinear, as shown in Fig. 3.



The corners x = 0 corresp ond to one parton momentum b eing soft (p ! 0).

i

i

2.2 Structure of the Cross Section

One can easily calculate the cross section corresp onding to Fig. 1 and the similar

amplitude in which the attaches to the antiquark line. The result is

2 2

d x + x 1

s

1 2

= C ; (9)

F

 dx dx 2 (1 x )(1 x )

0 1 2 1 2

Figure 2: Allowed region for (x ;x ). Then x is 2 x x .

1 2 3 1 2

Figure 3: Allowed region for (x ;x ). The lab els and small pictures show the

1 2

physical con guration of the three partons corresp onding to subregions in the

allowed triangle.

P

2 2 +

where C =4=3 and  =(4 =s) Q is the total cross section for e e !

F 0

f

0

hadr ons at order . The cross section has collinear singularities:

s

(1 x ) ! 0 ; (2&3 col l inear );

1

(1 x ) ! 0 ; (1&3 col l inear ) : (10)

2

There is also a singularity when the gluon is soft: x ! 0. In terms of x and x ,

3 1 2

this singularity o ccurs when

(1 x )

1

 const: (11) (1 x ) ! 0; (1 x ) ! 0;

1 2

(1 x )

2

Let us write the cross section in a way that displays the collinear singularity

at  ! 0 and the soft singularityat E ! 0:

31 3

1 d f (E ; )

s 3 31

= C : (12)

F

 dE d cos  2 E (1 cos  )

0 3 31 3 31

Here, f (E ; ) , a rather complicated function. The only thing that we need to

3 31

know ab out it is that it is nite for E ! 0 and for  ! 0.

3 31

Now lo ok at the collinear singularity,  ! 0.Ifweintegrate over the singular

31

region holding E xed, we nd that the integral is divergent:

3

Z

1

d

d cos  = log(1): (13)

31

dE d cos 

a

3 31

Similarly,ifweintegrate over the region of the soft singularity, holding  xed,

31

we nd that the integral is divergent:

Z

a

d

dE = log(1): (14)

3

dE d cos 

0

3 31

Evidently, p erturbation theory is telling us that we should not take the p ertur-

+

bative cross section to o literally. The total cross section for e e ! hadr ons is

certainly nite, so this partial cross section cannot b e in nite. What we are seeing

is a breakdown of p erturbation theory in the soft and collinear regions, and we

should understand why.

+

Figure 4: Cross section for e e ! q qg , illustrating the singularity when the

gluon is soft or collinear with the .

Where do the singularities come from? Lo ok at Fig. 4 (in a physical gauge).

2

The scattering matrix element M contains a factor 1=(p + p ) where

1 3

2

(p + p ) =2p p =2E E (1 cos  ): (15)

1 3 1 3 1 3 31

2

Evidently,1=(p +p ) is singular when  ! 0 and when E ! 0. The collinear

1 3 31 3

singularity is somewhat softened b ecause the numerator of the Feynman diagram

contains a factor prop ortional to  in the collinear limit. (This is not exactly

31

obvious, but is easily seen by calculating. If you like arguments, you can

derive this factor from quark helicity conservation and overall angular momentum

conservation.) Wethus nd that

" #

2



31

2

jMj / (16)

2

E 

3

31

for E ! 0 and  ! 0. Note the universal nature of these factors.

3 31

Integration over the double singular region of the momentum space for the

gluon has the form

Z Z

2

E dE d cos  d

3 31

3

2

 E dE d d: (17)

3 3

31

E

3

Combining the integration with the matrix element squared gives

" #

2

Z Z

2

 dE d

31 3

31

2

d  E dE d d  d: (18)

3 3

31

2 2

E  E 

3 3

31 31

Thus wehave a double logarithmic divergence in p erturbation theory for the soft

and collinear region. With just a little enhancement of the argument, we see

that there is a collinear divergence from integration over  at nite E and a

31 3

separate soft divergence from integration over E at nite  . Essentially the

3 31

same argument applies to more complicated graphs. There are divergences when

two nal state partons b ecome collinear and when a nal state gluon b ecomes soft.

8

Generalizing further, there are also divergences when several nal state partons

b ecome collinear to one another or when several (with no net avor quantum

numb ers) b ecome soft.

Wehave seen that if weintegrate over the singular region in momentum space

with no cuto , we get in nity. The integrals are logarithmically divergent, so if we

2

integrate with an infrared cuto M ,we will get big logarithms of M =s.Thus

IR

IR

the collinear and soft singularities represent p erturbation theory out of control.

Carrying on to higher orders of p erturbation theory, one gets

2 2

1+  (big )+ (big ) +  : (19)

s

s

If this expansion is in p owers of (M ), wehave  1. Nevertheless, the

s Z s

big logarithms seem to sp oil anychance of the low order terms of p erturbation

theory b eing a go o d approximation to any cross section of interest. Is the situation

hop eless? We shall havetoinvestigate further to see.

2.3 Interlude: Null Plane Co ordinates

Figure 5: Null plane axes in momentum space.

In order to understand b etter the issue of singularities, it is helpful to intro duce

a concept that is generally quite useful in high-energy quantum eld theory,null

plane co ordinates. The idea is to describ e the momentum of a particle using

 + 1 2

momentum comp onents p =(p ;p ;p ;p ) where

p

 0 3

p =(p p )= 2: (20)

For a particle with large momentum in the +z direction and limited transverse

+

momentum, p is large and p is small. Often one chooses the plus axis so that

+

a particle or group of particles of interest have large p and small p and p .

T



Using null plane comp onents, the covariant square of p is

2 + 2

p =2p p p : (21)

T

Thus, for a particle on its mass shell, p is

2 2

p + m

T

: (22) p =

+

2p

Note also that, for a particle on its mass shell,

+

p > 0 ; p > 0 : (23)

Integration over the mass shell is

Z Z Z

3 +

1

p~ d dp

3 3 2

p

(2 )  =(2) d p  : (24)

T

+

2 2

2p

2 ~p + m 0



We also use the plus/minus comp onents to describ e a space-time p oint x :

p

 0 3

x =(x x )= 2. In describing a system of particles moving with large mo-

+

mentum in the plus direction, we are invited to think of x as \time." Classically,

+

the particles in our system follow paths nearly parallel to the x axis, evolving

+

slowly as it moves from one x = const: plane to another.

We relate momentum space to p osition space for a quantum system byFourier

transforming. In doing so, wehave a factor exp(ip  x), which has the form

+ +

p  x = p x + p x p  x : (25)

T T

+ +

Thus x is conjugate to p , and x is conjugate to p . That is a little confusing,

but it is simple enough.

2.4 Space-Time Picture of the Singularities

Figure 6: Corresp ondence b etween singularities in momentum space and the de-

velopment of the system in space-time.

 

+ 

Wenow return to the singularity structure of e e ! q qg . De ne p + p = k .

1 3

+ 2 +

Cho ose null plane co ordinates with k large and k =0. Then k =2k k

T

b ecomes small when

2 2

p p

3;T 3;T

k = + (26)

+ +

2p 2p

1 3

+ +

b ecomes small. This happ ens when p b ecomes small with xed p and p ,so

3;T

1 3

that the gluon momentum is nearly collinear with the quark momentum. It also

+ +

happ ens when p and p b oth b ecome small with p /jp j, so that the gluon

3;T 3;T

3 3

momentum is soft. (It also happ ens when the quark b ecomes soft, but there is a

numerator factor that cancels the soft quark singularity.) Thus the singularities

for a soft or collinear gluon corresp ond to small k .

Now consider the to co ordinate space. The quark

in Fig. 6 is

Z

+ + +

S (k )= dx dx dx exp (i[k x + k x k  x]) S (x): (27)

F F

+

When k is large and k is small, the contributing values of x have small x and

+

large x . Thus the propagation of the virtual quark can b e pictured in space-

+

time as in Fig. 6. The quark propagates a long distance in the x direction

b efore decaying into a quark-gluon pair. That is, the singularities that can lead

to divergent p erturbative cross sections arise from interactions that happ en a long

time after the creation of the initial quark-antiquark pair.

2.5 Nature of the Long-Time Physics

+

Figure 7: Typical paths of partons in space contributing to e e ! hadr ons,

as suggested by the singularities of p erturbative diagrams. Short wavelength

elds are represented by classical paths of particles. Long wavelength elds are

represented bywavy lines.

Imagine dividing the contributions to a scattering cross section into long-time

contributions and short-time contributions. In the long-time contributions, p er-

turbation theory is out of control, as indicated in Eq. (19). Nevertheless, the

generic structure of the long-time contribution is of great interest. This structure

is illustrated in Fig. 7. Perturbative diagrams have big contributions from space-

time histories in which partons move in collinear groups and additional partons

are soft and communicate over large distances, while carrying small momentum.

The picture of Fig. 7 is suggested by the singularity structure of diagrams at

any xed order of p erturbation theory. Of course, there could b e nonp erturbative

e ects that would invalidate the picture. Since nonp erturbative e ects can b e

invisible in p erturbation theory, one cannot claim that the structure of the nal

state indicated in Fig. 7 is known to b e a consequence of QCD. One can p oint,

however, to some cases in which one can go b eyond xed order p erturbation theory

and sum the most imp ortant e ects of diagrams of all orders (for example, Ref. 9).

In such cases, the general picture suggested by Fig. 7 remains intact.

Wethus nd that p erturbative QCD suggests a certain structure of the nal

+

state pro duced in e e ! hadr ons: the nal state should consist of jets of nearly

collinear particles plus soft particles moving in random directions. In fact, this

qualitative prediction is a qualitative success.

Given some degree of qualitative success, wemay b e b older and ask whether

p erturbative QCD p ermits quantitative predictions. If wewant quantitative pre-

dictions, we will somehowhave to nd things to measure that are not sensitiveto

interactions that happ en long after the basic hard interaction. This is the sub ject

of the next section.

2.6 The Long-Time Problem

Wehave seen that p erturbation theory is not e ective for long-time physics. But

the detector is a long distance away from the interaction, so it would seem that

long-time physics has to b e present.

Fortunately, there are some measurements that are not sensitive to long-time

+

physics. An example is the total cross section to pro duce hadrons in e e anni-

p

hilation. Here, e ects from times t  1= s cancel b ecause of .To see

why, note that the quark state is created from the vacuum by a current op era-

tor J at some time t; it then develops from time t to time 1 according to the

evolution op erator U (1;t), when it b ecomes the nal state

jN i. The cross section is prop ortional to the sum over N of this amplitude times

0

a similar complex conjugate amplitude with t replaced by a di erent time t .We

p

0 0

Fourier transform this with exp (i s (t t )), so that we can takettt to

p

P

b e of order 1= s. Now, replacing jN ihN j by the unit op erator and using the

unitarity of the evolution op erators U ,we obtain

X

0 0

h0jJ (t )U (t ; 1)jN ihN jU (1;t)J(t)j0i (28)

N

0 0 0 0

= h0jJ (t )U (t ; 1)U (1;t)J(t)j0i = h0jJ(t )U(t ;t)J(t)j0i:

Because of unitarity, the long- has canceled out of the cross section,

0

and wehave only evolution from t to t .

There are three ways to view this result. First, wehave the formal argument

given ab ove. Second, wehave the intuitive understanding that after the initial

p

and are created in a time t of order 1= s, something will happ en

with probability 1. Exactly what happ ens is long-time physics, but we don't care

ab out it since we sum over all the p ossibilities jN i. Third, we can calculate at

some nite order of p erturbation theory. Then we see infrared in nities at various

stages of the calculations, but we nd that the in nities cancel b etween real gluon

emission graphs and virtual gluon graphs. An example is shown in Fig. 8.

Figure 8: Cancellation b etween real and virtual gluon graphs. If weintegrate the

real gluon graph on the left times the complex conjugate of the similar graph with

the gluon attached to the antiquark, we will get an infrared in nity.However, the

virtual gluon graph on the right times the complex conjugate of the Born graph

is also divergent, as is the Born graph times the complex conjugate of the virtual

gluon graph. Adding everything together, the infrared in nities cancel.

We see that the total cross section if free of sensitivity to long-time physics. If

the total cross section were all you could lo ok at, QCD physics would b e a little

b oring. Fortunately, there are other quantities that are not sensitive to infrared

e ects. They are called infrared safe quantities.

To formulate the concept of infrared safety, consider a measured quantity that

is constructed from the cross sections,

d [n]

; (29)

d dE d  dE d

2 3 3 n n

+

to make n hadrons in e e annihilation. Here, E is the energy of the j th hadron

j

and = ( ; ) describ es its direction. We treat the hadrons as e ectively

j j j

massless and do not distinguish the hadron avors. Following the notation of

Ref. 10, let us sp ecify functions S that describ e the measurementwewant, so

n

that the measured quantityis

Z

1 d [2]

 

) I = ;p d S (p

2 2

2 1

2! d

2

Z

1 d [3]

  

+ d dE d S (p ;p ;p )

2 3 3 3

1 2 3

3! d dE d

2 3 3

Z

1

+ d dE d dE d

2 3 3 4 4

4!

d [4]

   

 S (p ;p ;p ;p )

4

1 2 3 4

d dE d dE d

2 3 3 4 4

+ : (30)

The functions S are symmetric functions of their arguments. In order for our

measurement to b e infrared safe, we need

 

  

S (p ;:::;(1 ) p ;p )=S (p ;:::;p ) (31)

n+1 n

1 1

n n n

for 0    1.

Figure 9: Infrared safety. In an infrared safe measurement, the three-jet

shown on the left should b e (approximately) equivalent to an ideal three-jet event

shown on the right.

What do es this mean? The physical meaning is that the functions S and

n

S are related in suchaway that the cross section is not sensitive to whether

n1

or not a mother particle divides into two collinear daughter particles that share

its momentum. The cross section is also not sensitive to whether or not a mother

particle decays to a daughter particle carrying all of its momentum and a soft

daughter particle carrying no momentum. The cross section is also not sensitive

to whether or not two collinear particles combine, or a soft particle is absorb ed

by a fast particle. All of these decay and recombination pro cesses can happ en

with large probability in the nal state long after the hard interaction. But,

by construction, they don't as long as the sum of the probabilities for

something to happ en or not to happ en is one.

Another version of the physical meaning is that for an IR-safe quantity,a

physical event with hadron jets should give approximately the same measurement

as a parton event with each jet replaced by a parton, as illustrated in Fig. 9. To

see this, we simply have to delete soft particles and combine collinear particles

until three jets have b ecome three particles.

In a calculation of the measured quantity I ,we simply calculate with partons

instead of hadrons in the nal state. The calculational meaning of the infrared

safety condition is that the infrared in nities cancel. The argument is that the

in nities arise from soft and collinear con gurations of the partons, that these

con gurations involve long times, and that the time evolution op erator is unitary.

Ihave started with an abstract formulation of infrared safety. It would b e

go o d to have a few examples. The easiest is the total cross section, for which





S (p ;:::;p )=1: (32)

n

1

n

A less trivial example is the thrust distribution. One de nes the thrust T of an

n

n particle eventas

P

n

j~p  ~uj

i



i=1



T (p ;:::;p ) = max : (33)

P

n

1

n

n

~u

j~p j

i

i=1

Here ~u is a unit vector, whichwevary to maximize the sum of the absolute values

of the pro jections of ~p on ~u. Then the thrust distribution (1= ) d =dT is de ned

i tot

by taking

 

 

S (p ;::: ;p )=(1= )  ( T T (p ;:::;p )) : (34)

n tot n

1 1

n n

It is a simple exercise to show that the thrust of an event is not a ected by

collinear parton splitting or by zero momentum partons. Therefore, the thrust

distribution is infrared safe.

Another infrared safe quantity is the cross sections to make n jets. Here

one has to de ne what one means by a jet. The de nitions used in electron-

p ositron annihilationtypically involve successively combining particles that are

nearly collinear to make the jets. A description can b e found in Ref. 11. I discuss

jet cross sections for hadron collisions in Sec. 5.4.

12

A nal example is the energy-energy correlation function, which measures

the average of the pro duct of the energy in one calorimeter cell times the energy

in another calorimeter cell. One lo oks at this average as a function of the angular

separation of the calorimeter cells.

Before leaving this sub ject, I should mention another way to eliminate sensi-

tivity to long-time physics. Consider the cross section

+

d (e e !  + X )

: (35)

dE



This cross section can b e written as a convolution of two factors, as illustrated

in Fig. 10. The rst factor is a calculated \hard scattering cross section" for

+ +

e e ! q uar k + X or e e ! g l uon + X . The second factor is a \parton decay

function" for q uar k !  + X or g l uon !  + X . These functions contain the

long-time sensitivity and are to b e measured, since they cannot b e calculated

p erturbatively.However, once they are measured in one pro cess, they can b e used

for another pro cess. This nal state factorization is similar to the initial state

factorization involving parton distribution functions, whichwe will discuss later.

(See Refs. 1, 2, and 13 for more information.)

+

Figure 10: The cross section for e e !  + X can b e written as a convolution of

a short-distance cross section (inside the dotted line) and a parton decay function.

3 The Smallest Time Scales

p

In this section, I explore the physics of time scales smaller than 1= s. One

way of lo oking at this physics is to say that it is plagued by in nities and we

can manage to hide the in nities. A b etter view is that the short-time physics

contains wonderful truths that wewould like to discover|truths ab out grand

uni ed theories, , and the like. However, quantum eld theory

is arranged so as to e ectively hide the truth from our exp erimental apparatus,

which can prob e with a time resolution of only an inverse half TeV.

I rst outline what renormalization do es to hide the ugly in nities or the b eau-

tiful truth. Then I describ e how renormalization leads to the running coupling.

Because of renormalization, calculated quantities dep end on a renormalization

scale. Ilookathow this dep endence works and how the scale can b e chosen.

Finally, I discuss how one can use exp eriment to lo ok for the hidden physics b e-

yond the Standard Mo del, taking high E jet pro duction in hadron collisions as

T

an example.

3.1 What Renormalization Do es

In anyFeynman graph, one can insert p erturbative corrections to the vertices and

the propagation of particles, as illustrated in Fig. 11. The lo op integrals in these

p

graphs will get big contributions from momenta much larger than s. That is,

there are big contributions from interactions that happ en on time scales much

p

smaller than 1= s.Ihave tried to illustrate this in the gure. The virtual vector

p

s, while the virtual uctuations that correct the b oson propagates for a time 1=

electroweak vertex and the quark propagator o ccur over a time t that can b e

p

much smaller than 1= s.

p

Let us pick an ultraviolet cuto M that is much larger than s, so that we

calculate the e ect of uctuations with 1=M<texactly, up to some order of

p erturbation theory. What, then, is the e ect of virtual uctuations on smaller

time scales, t with t<1=M but, say,tstill larger than t , where gravity

P l ank

takes over? Let us supp ose that we are willing to neglect contributions to the

p

cross section that are of order s=M or smaller compared to the cross section

14

itself. Then there is a remarkable theorem: the e ects of the uctuations are

not particularly small, but they can b e absorb ed into changes in the couplings of

the theory. (There are also changes in the masses of the theory and adjustments

to the normalizations of the eld op erators, but we can concentrate on the e ect

on the couplings.)

The program of absorbing very short-time physics into a few parameters go es

under the name of renormalization. There are several schemes available for renor-

Figure 11: Renormalization. The e ect of the very small time interactions pictured

are absorb ed into the running coupling.

malizing. Each of them involves the intro duction of some scale parameter that is

not intrinsic to the theory but tells howwe did the renormalization. Let us agree

to use MS renormalization (see Ref. 14 for details). Then weintro duce an MS

renormalization scale . A go o d (but approximate) way of thinking of  is that

the physics of time scales t  1= is removed from the p erturbative calculation.

The e ect of the small time physics is accounted for by adjusting the value of the

strong coupling, so that its value dep ends on the scale that we used: = ().

s s

(The value of the electromagnetic coupling also dep ends on .)

3.2 The Running Coupling

Figure 12: Short-time uctuations in the propagation of the gluon eld absorb ed

into the running strong coupling.

We account for time scales much smaller than 1= by using the running coupling

(). That is, a uctuation such as that illustrated in Fig. 12 can b e dropp ed

s

from a calculation and absorb ed into the running coupling that describ es the

probability for the quark in the gure to emit the gluon. The  dep endence of

() is given by a certain di erential equation, called the renormalization group s

equation (see Ref. 14):

! !

2 3

d () () ()

s s s

= ( ()) = +  : (36)

s 0 1

2

d ln( )   

One calculates the b eta function ( ) p erturbatively in QCD. The rst co e-

s

cient, with the conventions used here, is

= (33 2 N )=12 ; (37)

0 f

where N is the numb er of quark avors.

f

Of course, at time scales smaller than a very small cuto 1=M (at the \GUT

scale," say), there is completely di erentphysics op erating. Therefore, if we use

just QCD to adjust the strong coupling, we can say that we are accounting for

the physics b etween times 1=M and 1=. The value of at   M is then the

s 0

b oundary condition for the di erential equation.

Figure 13: Distance scales accounted for by explicit xed order p erturbative cal-

culation and by use of the renormalization group.

The renormalization group equation sums the e ects of short-time uctuations

of the elds. To see what one means by \sums" here, consider the result of solving

the renormalization group equation with all of the beyond set to zero:

i 0

2 2 2

(M ) ()  (M ) ( = ) ln( =M )

s s 0

s

2

2 2 2 3

+( = ) (M )+ ln ( =M )

0

s

(M)

s

= : (38)

2 2

1+( = ) (M ) ln( =M )

0 s

A series in p owers of (M )|that is the strong coupling at the GUT scale|is

s

summed into a simple function of . Here (M ) app ears as a parameter in the

s

solution.

Note a crucial and wonderful fact. The value of () decreases as  increases.

s

This is called \." Asymptotic freedom implies that QCD acts

likeaweakly interacting theory on short time scales. It is true that quarks and

gluons are strongly b ound inside , but this strong binding is the result of

weak acting collectively over a long time.

In Eq. (38), we are invited to think of the graph of ()versus . The

s

di erential equation that determines this graph is characteristic of QCD. There

could, however, b e di erentversions of QCD with the same di erential equation

but di erent curves, corresp onding to di erent b oundary values (M ). Thus

s

the parameter (M ) tells us whichversion of QCD wehave. To determine this

s

parameter, we consult exp eriment. Actually, Eq. (38) is not the most convenient

way to write the solution for the running coupling. A b etter expression is



()  : (39)

s

2 2

ln( = )

0

Here wehave replaced (M )by a di erent (but completely equivalent) parameter

s

. A third form of the running coupling is

(M )

s Z

()  : (40)

s

2

2

1+( = ) (M ) ln( =M )

0 s Z

Z

Here the value of ()at=M lab els the version of QCD that obtains in our

s Z

world.

In any of the three forms of the running coupling, one should revise the equa-

tions to account for the second term in the b eta function in order to b e numerically

precise.

3.3 The Choice of Scale

In this section, we consider the choice of the renormalization scale  in a calculated

+

cross section. Consider, as an example, the cross section for e e ! hadr ons via

virtual photon decay. Let us write this cross section in the form

1 0

2

X

4

2

A @

[1 + ] : (41)  = Q

tot

f

s

f

2

Here s is the square of the c.m. energy, is e =(4 ), and Q is the electric charge

f

in units of e carried by the quark of avor f , with f = u; d; s; c; b. The nontrivial

part of the calculated cross section is the quantity , which contains the e ects of

the strong interactions. Using MS renormalization with scale , one nds (after

15

a lot of work) that  is given by

!

2

h  i

() ()

s s

2

= + 1:4092 + 1:9167 ln  =s

 

!

3

h    i

()

s

2

2 2

+ 12:805 + 7:8186 ln  =s +3:674 ln  =s



+  : (42)

Here, of course, one should use for () the solution of the renormalization group

s

Eq. (36) with at least two terms included.

As discussed in the preceding subsection, when we renormalize with scale ,we

are de ning what we mean by the strong coupling. Thus, in Eq. (42) dep ends

s

on . The p erturbative co ecients in Eq. (42) also dep end on . On the other

hand, the physical cross section do es not dep end on :

d

=0: (43)

2

d ln 

That is b ecause  is just an artifact of howwe organize p erturbation theory, not

a parameter of the underlying theory.

Let us consider Eq. (43) in more detail. Write  in the form

1

X

n

  c () () : (44)

n s

n=1

If we di erentiate not the complete in nite sum but just the rst N terms, we get

minus the derivative of the sum from N + 1 to in nity. This remainder is of order

N +1

as ! 0. Thus

s

s

N

X

d

n N+1

c () () O( () ): (45)

n s s

2

d ln 

n=1

That is, the harder wework calculating more terms, the less the calculated cross

section dep ends on .

Since wehave not worked in nitely hard, the calculated cross section dep ends

2

on . What choice shall we make for ? Clearly,ln(=s) should not b e big.

Otherwise the co ecients c () are large and the \convergence" of p erturbation

n

theory will b e sp oiled. There are some who will argue that one scheme or the other

for cho osing  is the \b est." You are welcome to follow whichever advisor you

want. I will showyou b elow that for a well-b ehaved quantity like , the precise

choice makes little di erence, as long as you ob ey the common sense prescription

2

that ln ( =s) not b e big. 0.06

0.05

0.04

() 0.03

0.02

0.01

0

-3 -2 -1 0 1 2

p

ln (= s)

2

Figure 14: Dep endence of () on the MS renormalization scale . The falling

curveis . The atter curveis . The horizontal lines indicate the amountof

1 2

variation of  when  varies by a factor 2.

2

3.4 An Example

Let us consider a quantitative example of how() dep ends on . This will also

giveusachance to think ab out the theoretical error caused by replacing  by the

sum  of the rst n terms in its p erturbative expansion. Of course, we do not

n

know what this error is. All we can do is provide an estimate. (Our discussion

will b e rather primitive. For a more detailed error estimate for the case of the

hadronic width of the Z b oson, see Ref. 16.)

Let us think of the error estimate in the spirit of a \1  " theoretical error: we

would b e surprised if j j were much less than the error estimate and we

n

would also b e surprised if this quantitywere much more than the error estimate.

Here, one should exercise a little caution. Wehave no reason to exp ect that theory

errors are Gaussian distributed. Thus a 4  di erence b etween  and  is not

n

out of the question, while a 4  uctuation in a measured quantity with purely

statistical, Gaussian errors is out of the question.

p

Take (M )=0:117, s =34GeV , ve avors. In Fig. 14, I plot ()

s Z

versus p de ned by

p

p

 =2 s: (46)

1

approximation to (),  ()= ()= . The steeply falling curve is the order

1 s

s

Notice that if wechange  by a factor 2,  ()changes by ab out 0.006. If we

1

p

had no other information than this, we might pick ( s)0:044 as the \b est" 1

value and assign a 0:006 error to this value. (There is no sp ecial magic to the

use of a factor of 2 here. The reader can pickany factor that seems reasonable.)

Another error estimate can b e based on the simple exp ectation that the co e-

n

cients of are of order 1 for the rst few terms. (Eventually, they will grow like

s

n! Ref. 16 takes this into account, but we ignore it here.) Then the rst omitted

2

term should b e of order 1  0:020 using (34 GeV )  0:14. Since this

s

s

is bigger than the previous 0:006 error estimate, wekeep this larger estimate:

  0:044  0:020.

2

Returning now to Fig. 14, the second curve is the order approximation,

s

 (). Note that  () is less dep endentonthan  ().

2 2 1

What value would wenow take as our b est estimate of ? One idea is to

cho ose the value of  at which () is least sensitiveto . This idea is called

2

17

the principle of minimal sensitivity:

" #

d()

 =( ) ; =0: (47)

PMS PMS

d ln 

=

PMS

This prescription gives   0:0470. Note that this is ab out 0.003 away from our

previous estimate,   0:0440. Thus our previous error estimate of 0.020 was to o

big, and we should b e surprised that the result changed so little. We can makea

new error estimate by noting that  ()varies by ab out 0.0012 when  changes

2

by a factor 2 from  .Thus we might estimate that   0:0470 with an error

PMS

of 0:0012. This estimate is represented by the two horizontal lines in Fig. 14.

An alternative error estimate can b e based on the next term b eing of order

3

1  (34 GeV )  0:003. Since this is bigger than the previous 0:0012 error

s

estimate, wekeep this larger estimate:   0:0470  0:003.

I should emphasize that there are other ways to pick the \b est" value for .

18

For instance, one can use the BLM metho d, which is based on cho osing the 

that sets to zero the co ecient of the numb er of quark avors in  (). Since the

2

graph of  () is quite at, it makes very little di erence which metho d one uses.

2

3

Now let us lo ok at ()evaluated at order , (). Here we make use of

3

s

the full formula in Eq. (42). In Fig. 15, I plot  () along with  () and  ().

3 2 1

The variation of  () with  is smaller than that of  (). The improvementis

3 2

not overwhelming, but is apparent particularly at small .

It is a little dicult to see what is happ ening in Fig. 15, so I show the same

thing with an expanded scale in Fig. 16. (Here the error band based on the 

dep endence of  is also indicated. Recall that we decided that this error band 2 0.06

0.05

0.04

() 0.03

0.02

0.01

0

-3 -2 -1 0 1 2

p

log (= s)

2

Figure 15: Dep endence of () on the MS renormalization scale . The falling

curveis . The atter curveis . The still atter curveis .

1 2 3

was an underestimate.) The curve for  () has zero derivativeattwo places.

3

The corresp onding values are   0:0436 and   0:0456. If I take the b est value

of  to b e the average of these twovalues and the error to b e half the di erence,

I get   0:0446  0:0010.

4

The alternative error estimate is 1  (34 GeV )  0:0004. Wekeep the

s

larger error estimate of 0:0010.

Was the previous error estimate valid? We guessed   0:0470  0:003. Our

new b est estimate is 0.0446. The di erence is 0.0024, which is in line with our

previous error estimate. Had we used the error estimate 0:0012 based on the

 dep endence, wewould have underestimated the di erence, although wewould

not have b een to o far o .

3.5 Beyond the Standard Mo del

Wehave seen how the renormalization group enables us to account for QCD

p

s, as indicated in Fig. 17. However, at physics at time scales much smaller than

some scale t  1=M ,we run into the unknown!

How can we see the unknown in current exp eriments? First, the unknown

2

physics a ects , , sin ( ). Second, the unknown physics a ects masses

s em W

of u;d;:::;e;;:::. That is, the unknown physics (presumably) determines the

parameters of the Standard Mo del. These parameters have b een well-measured.

Thus, a Nob el prize awaits the physicist who gures out how to use a mo del for 0.05

0.048

0.046

) ( 0.044

0.042

-3 -2 -1 0 1 2

p

log (= s)

2

Figure 16: Dep endence of () on the MS renormalization scale  with an ex-

panded scale. The falling curveis . The atter curveis . The still atter

1 2

curveis . The horizontal lines represent the variation of  when  varies by

3 2

a factor 2.

the unknown physics to predict these parameters.

Figure 17: Time scales accounted for by xed order p erturbative calculations and

by use of the renormalization group.

Figure 18: New physics at a TeV scale. In the rst diagram, quarks scatter by

gluon exchange. In the second diagram, the quarks exchange a new ob ject with a

TeV mass, or p erhaps exchange some of the constituents out of which quarks are

made.

There is another way that as yet unknown physics can a ect current exp er-

iments. Supp ose that quarks can scatter by the exchange of some new particle

with a heavy mass M , as illustrated in Fig. 18, and supp ose that this mass is not

to o enormous, only a few TeV. Perhaps the new particle isn't a particle at all,

but is a pair of constituents that live inside of quarks. As mentioned ab ove, this

physics a ects the parameters of the Standard Mo del. However, unless we can

predict the parameters of the Standard Mo del, this e ect do es not help us. There

is, however, another p ossible clue. The physics at the TeV scale can intro duce

new terms into the Lagrangian that we can investigate in current exp eriments.

In the second diagram in Fig. 18, the twovertices are never at a separation in

time greater than 1=M , so that our low-energy prob es cannot resolve the details

p

of the structure. As long as we sticktolow-energy prob es, s  M , the e ect

of the new physics can b e summarized by adding new terms to the Lagrangian of

QCD. A typical term mightbe

2

g~



 

L= : (48)



2

M

2

There is a factorg ~ that represents howwell the new physics couples to quarks.

2

The most imp ortant factor is the factor 1=M . This factor must b e there: the

pro duct of eld op erators has dimension 6 and the Lagrangian has dimension 4,

so there must b e a factor with dimension 2. Taking this argument one step

further, the pro duct of eld op erators in L must have a dimension greater than

4 b ecause any pro duct of eld op erators having dimension equal to or less than

4 that resp ects the symmetries of the Standard Mo del is already included in the

Lagrangian of the Standard Mo del.

3.6 Lo oking for New Terms in the E ective Lagrangian

How can one detect the presence in the Lagrangian of a term like that in Eq. (48)?

These terms are small. Therefore we need either a high-precision exp eriment, or

an exp eriment that lo oks for some e ect that is forbidden in the Standard Mo del,

or an exp eriment that has mo derate precision and op erates at energies that are

as high as p ossible.

Let us consider an example of the last of these p ossibilities, p +p!jet + X

as a function of the transverse energy ( P ) of the jet. The new term in the

T

Lagrangian should add a little bit to the observed cross section that is not included

in the standard QCD theory. When the transverse energy E of the jet is small T

compared to M ,we exp ect

2

D ata T heor y E

T

2

/ g~ : (49)

2

T heor y M

2 2 2

Here the factorg ~ =M follows b ecause L contains this factor. The factor E

T

follows b ecause the left-hand side is dimensionless and E is the only factor with

T

dimension of mass that is available.

1

CTEQ3M CDF (Preliminary) * 1.03 D0 (Preliminary) * 1.01

0.5

0 (Data - Theory)/ Theory

-0.5 50 100 200 300 400

Et (GeV)

Figure 19: Jet cross sections from CDF and D0 compared to QCD theory. (Data

Theory)/Theory is plotted versus the transverse energy E of the jet. The

T

theory here is next-to-leading order QCD using the CTEQ3M parton distribution.

Source: Ref. 19.

20

In Fig. 19, I show a plot comparing exp erimental jet cross sections from CDF

21

and D0 (Ref. ) compared to next-to-leading order QCD theory. The theory works

ne for E < 200 GeV , but for 200 GeV

T T

deviation of just the form anticipated in Eq. (49).

This example illustrates the idea of how small distance physics b eyond the

Standard Mo del can leave a trace in the form of small additional terms in the

e ective Lagrangian that controls physics at currently available energies. However,

in this case, there is some indication that the observed e ect might b e explained

by some combination of the exp erimental systematic error and the uncertainties

22

inherent in the theoretical prediction. In particular, the prediction is sensitive

to the distributions of quarks and gluons contained in the colliding , and

the gluon distribution in the kinematic range of interest here is rather p o orly

known. In the next section, we turn to the de nition, use, and measurementof the distributions of quarks and gluons in hadrons.

4 Deeply Inelastic Scattering

Until now, I have concentrated on hard scattering pro cesses with in the

initial state. For such pro cesses, wehave seen that the hard part of the pro cess can

b e describ ed using p erturbation theory b ecause () gets small as  gets large.

s

Furthermore, wehave seen how to isolate the hard part of the interaction bycho os-

ing an infrared safe observable. But what ab out hard pro cesses in which there

are hadrons in the initial state? Since the fundamental hard interactions involve

quarks and gluons, the theoretical description necessarily involves a description

of how the quarks and gluons are distributed in a hadron. Unfortunately, the

distribution of quarks and gluons in a hadron is controlled by long-time physics.

We cannot calculate the relevant distribution functions p erturbatively (although

a calculation in lattice QCD might give them, in principle). Thus, wemust nd

how to separate the short-time physics from the parton distribution functions,

and wemust learn how the parton distribution functions can b e determined from

the exp erimental measurements.

In this section, I discuss parton distribution functions and their role in deeply

inelastic scattering (DIS). This includes e + p ! e + X and  + p !

e + X where the momentum transfer from the lepton is large. I rst outline the

kinematics of deeply inelastic scattering and de ne the structure functions F , F ,

1 2

and F used to describ e the pro cess. By examining the space-time structure of DIS,

3

we will see how the cross section can b e written as a convolution of two factors,

one of which is the parton distribution functions and the other of which is a cross

section for the lepton to scatter from a quark or gluon. This factorization involves

a scale  that, roughly sp eaking, divides the soft from the hard regime; I discuss

F

the dep endence of the calculated cross section on  . With this groundwork laid,

F

I give the MS de nition of parton distribution functions in terms of eld op erators

and discuss the evolution equation for the parton distributions. I close the section

with some comments on how the parton distributions are, in practice, determined

from exp eriment.

4.1 Kinematics of Deeply Inelastic Lepton Scattering



scatters on a hadron In deeply inelastic scattering, a lepton with momentum k



with momentum p . In the nal state, one observes the scattered lepton with

Figure 20: Kinematics of deeply inelastic scattering.

0

momentum k as illustrated in Fig. 20. The momentum transfer

  0

q = k k (50)

is carried on a photon, or a W or Z b oson.

The interaction b etween the vector b oson and the hadron dep ends on the

 

variables q and p .From these twovectors, we can build two scalars (not counting

2 2

m = p ). The rst variable is

2 2

Q = q ; (51)

2

where the minus sign is included so that Q is p ositive. The second scalar is the

dimensionless Bjorken variable,

2

Q

: (52) x =

bj

2p  q



(In the case of scattering from a nucleus containing A nucleons, one replaces p

 2

by p =A and de nes x = AQ =(2p  q ).)

bj

2

2

. One calls the scattering deeply inelastic if Q is large compared to 1 GeV

2

Traditionally, one sp eaks of the scaling limit, Q !1with x xed. Actually,

bj

2

the asymptotic theory to b e describ ed b elowworks prettywell if Q is bigger than,

2

say,4 GeV and x is anywhere in the exp erimentally accessible range, roughly

bj

4

10

bj

2 2

The invariant mass squared of the hadronic nal state is W =(p+q) . In

2

the scaling regime of large Q , one has

1 x

bj

2 2 2 2

W = m + Q  m : (53)

x

bj

This justi es saying that the scattering is not only inelastic but deeply inelastic.

 

Wehavespoken of the scalar variables that one can form from p and q .



Using the lepton momentum k , one can also form the dimensionless variable

p  q

y = : (54)

p  k

4.2 Structure Functions for DIS

One can make quite a lot of progress in understanding the theory of deeply in-

elastic scattering without knowing anything ab out QCD except its symmetries.

One expresses the cross section in terms of three structure functions, which are

2

functions of x and Q only.

bj

Supp ose that the initial lepton is a neutrino,  , and the nal lepton is a



muon. Then in Fig. 20, the exchanged vector b oson, call it V ,isaW b oson,

with mass M = M . Alternatively, supp ose that b oth the initial and nal

V W

leptons are and let the exchanged vector b oson b e a photon, with mass

M = 0. This was the situation in the original DIS exp eriments at SLACinthe

V

2

late 1960s. In exp eriments with suciently large Q , Z b oson exchange should b e

considered along with photon exchange, and the formalism describ ed b elowmust

b e augmented.

Given only the electroweak theory to tell us how the vector b oson couples to

the lepton, one can write the cross section in the form

2 3 0

4 d k C

V



d = L (k; q) W (p; q ); (55)



2

0 2 2

s 2jk j (q M )

V

4

where C is 1 in the case that V is a photon and 1=(64 sin  ) in the case that

V W



V is a W b oson. The tensor L describ es the lepton coupling to the vector b oson

and has the form

1

  0 

L = Tr (k  k  ) (56)

2

in the case that V is a photon. For a W b oson, one has

  0 

L = Tr (k  k  ) ; (57)

  + + 

where is (1 ) for a W b oson ( ! W `)or (1 + ) for a W b oson

5 5



( ! W `). See Ref. 1.



The tensor W describ es the coupling of the vector b oson to the hadronic

 

system. It dep ends on p and q .We know that it is Lorentz invariant and that

  

W = W . We also know that the current to which the vector b oson couples

is conserved (or in the case of the axial current, conserved in the absence of quark



masses, whichwe neglect here) so that q W =0. Using these prop erties, one





nds three p ossible tensor structures for W . Each of the three tensors multiplies

a structure function, F , F ,orF , which, since it is a Lorentz scalar, can dep end

1 2 3

2

only on the invariants x and Q .Thus

bj

!

q q

 

2

W = g F (x ;Q )

  1 bj

2

q

! !

p  q p  q 1

2

+ p q p q F (x ;Q )

    2 bj

2 2

q q p  q

1

  2

i p q F (x ;Q ): (58)

  3 bj

p  q

If we combine Eqs. (55,56,57,58), we can write the cross section for deeply in-

elastic scattering in terms of the three structure functions. Neglecting the hadron

2

mass compared to Q , the result is

" #

1 y y d

2

~

= N (Q ) yF + F +  (1 )F : (59)

1 2 V 3

dx dy x y 2

bj bj

~

Here the normalization factor N and the factor  multiplying F are

V 3

2

4

~

N = ;  =0; e +h ! e +X;

V

2

Q

2 2

 Q

~

;  =1;  +h ! +X; N =

V

4

2 2

4 sin ( )(Q +M )

W W

2 2

 Q

+

~

N = ;  =1; +h! +X: (60)

V

4

2 2

4 sin ( )(Q +M )

W W

In principle, one can use the y dep endence to determine all three of F ;F ;F in

1 2 3

a deeply inelastic scattering exp eriment.

4.3 Space-Time Structure of DIS

So far, wehave used the symmetries of QCD in order to write the cross section for

deeply inelastic scattering in terms of three structure functions, but wehave not

used any other dynamical prop erties of the theory.Nowwe turn to the question

of how the scattering develops in space and time.

For this purp ose, we de ne a convenient reference frame, which is illustrated

 +

in Fig. 21. Denoting comp onents of vectors v by(v ;v ;v ), wechose the frame T

Figure 21: Reference frame for the analysis of deeply inelastic scattering.

in which

1

+

p

(q ;q ;q)= (Q; Q; 0): (61)

2

We also demand that the transverse comp onents of the hadron momentum b e

zero in our frame. Then

2

Q 1 x m

bj

h

+

p

( (p ;p ;p)  ; ; 0): (62)

x Q

2

bj

Notice that in the chosen reference frame, the hadron momentum is big and the

momentum transfer is big.

Figure 22: Interactions within a fast moving hadron. The lines representworld

+

lines of quarks and gluons. The interaction p oints are spread out in x and pushed

together in x .

Consider the interactions among the quarks and gluons inside a hadron, using

+

x in the role of \time" as in Sec. 2.3. For a hadron at rest, these interactions

+

happ en in a typical time scale x  1=m, where m  300 MeV . A hadron

that will participate in a deeply inelastic scattering event has a large momentum,

+

p  Q, in the reference frame that we are using. The Lorentz transformation

from the rest frame spreads out interactions by a factor Q=m, so that

1 Q Q

+

x   = : (63)

2

m m m

This is illustrated in Fig. 22.

I o er twocaveats here. First, I am treating x as b eing of order 1. To treat

bj

small x physics, one needs to put back the factors of x , and the picture changes

bj bj

rather dramatically. Second, the interactions among the quarks and gluons in a

+

hadron at rest can take place on time scales x that are much smaller than 1=m,

as we discussed in Sec. 3. We will discuss this later on, but for nowwe start with

the simplest picture.

Figure 23: The virtual photon meets the fast moving hadron. One of the par-

tons is annihilated and recreated as a parton with a large minus comp onentof

momentum. This parton develops into a jet of particles.

What happ ens when the fast moving hadron meets the virtual photon? The

interaction with the photon carrying momentum q  Q is lo calized to within

+

x  1=Q: (64)

During this short time interval, the quarks and gluons in the are e ectively

free, since their typical interaction times are comparatively much longer.

+

Wethus have the following picture. At the moment x of the interaction, the

hadron e ectively consists of a collection of quarks and gluons (partons) that have

+ +

momenta (p ;p ). We can treat the partons as b eing free. The p are large, and

i

i i

it is convenient to describ e them using momentum fractions  :

i

+ +

 = p =p ; 0 < <1: (65)

i i

i

(This is convenient b ecause the  are invariant under b o osts along the z axis.)

i

The transverse momenta of the partons, p , are small compared to Q and can

i

b e neglected in the kinematics of the -parton interaction. The \on-shell" or

+

2

), are also very small = p =(2p \kinetic" minus momenta of the partons, p

i i

i

compared to Q and can b e neglected in the kinematics of the -parton interaction.

We can think of the partonic state as b eing describ ed byawave function

+ +

(p ;p ;p ;p ;); (66)

1 2

1 2

where indices sp ecifying and avor quantum numb ers have b een suppressed.

Figure 24: Feynman diagram for deeply inelastic scattering.

This approximate picture is represented in Feynman diagram language in

Fig. 24. The larger lled circle represents the hadron wave function . The

smaller lled circle represents a sum of sub diagrams in which the particles have

2

virtualities of order Q . All of these interactions are e ectively instantaneous on

the time scale of the intra-hadron interactions that form the wave function. The

approximate picture also leads to an intuitive formula that relates the observed

cross section to the cross section for -parton scattering:

Z

1

X

d^ () d

a

f (; )  d + O (m=Q): (67)

a=h

0 0 0 0

dE d! dE d!

0 a

In Eq. (67), the function f is a parton distribution function: f (; ) d gives

a=h

probability to nd a parton with avor a = g; u; u;d;::: , in hadron h, carrying

+

+

momentum fraction within d of  = p =p .Ifwe knew the wave functions ,we

i

would form f by summing over the number n of unobserved partons, integrating

2

j j over the momenta of the unobserved partons, and also integrating over the

n

transverse momentum of the observed parton.

0 0

The second factor in Eq. (67), d^ =dE d! , is the cross section for scattering

a

the lepton from the parton of avor a and momentum fraction  .

Ihave indicated a dep endence on a factorization scale  in b oth factors of

Eq. (67). This dep endence arises from the existence of virtual pro cesses among

+

the partons that take place on a time scale much shorter than the nominal x 

2

Q=m . I will discuss this dep endence in some detail shortly.

4.4 The Hard Scattering Cross Section

The parton distribution functions in Eq. (67) are derived from exp eriment. The

0 0

hard scattering cross sections d^ ()=dE d! are calculated in p erturbation theory,

a

using diagrams like those shown in Fig. 25. The diagram on the left is the lowest

order diagram. The diagram on the right is one of several that contributes to d^

at order ; in this diagram, the parton a is a gluon.

s

Lowest order. Higher order.

Figure 25: Some Feynman diagrams for the hard scattering part of deeply inelastic

scattering.

One can understand a lot ab out deeply inelastic scattering from Fig. 26, which

illustrates the kinematics of the lowest order diagram. Recall that in the reference

frame that we are using, the virtual vector b oson has zero transverse momentum.

The incoming parton has momentum along the plus axis. After the scattering,



the parton momentum must b e on the cone k k = 0, so the only p ossibilityis 

Figure 26: Kinematics of lowest order diagram.

that its minus momentum is nonzero and its plus momentum vanishes. That is

+ +

p + q =0: (68)

p p

+ +

2) while q = Q= 2, this implies Since p = Q=(x

bj

 = x : (69)

bj

The consequence of this is that the lowest order contribution to d^ in Eq. (67)

contains a delta function that sets  to x .Thus deeply inelastic scattering at a

bj

given value of x provides a determination of the parton distribution functions

bj

at momentum fraction  equal to x , as long as one works only to leading order.

bj

In fact, b ecause of this close relationship, there is some tendency to confuse the

2

structure functions F (x ;Q ) with the parton distribution functions f (; ). I

n bj a;h

will try to keep these concepts separate: the structure functions F are something

n

that one measures directly in deeply inelastic scattering; the parton distribution

functions are determined rather indirectly from exp eriments like deeply inelastic

scattering, using formulas that are correct only up to some nite order in .

s

4.5 Factorization for the Structure Functions

We will lo ok at DIS in a little detail since it is so imp ortant. Our ob ject is to

derive a formula relating the measured structure functions to structure functions

calculated at the parton level. Then we will lo ok at the parton level calculation

at lowest order.

Start with Eq. (67), representing Fig. 24. Wechange variables in this equation

0 0

from (E ;! )to(x ;y). We relate x to the momentum fraction  and a new

bj bj



variablex ^ that is just x with the proton momentum p replaced by the parton

bj



momentum p :

2 2

Q Q

x = =  = x:^ (70)

bj

2p  q 2p  q

That is,x ^ is the parton level version of x . The variable y is identical to the parton

bj



level version of y b ecause p app ears in b oth the numerator and denominator:

p  q p  q

y = = : (71)

p  k p  k

Thus Eq. (67) b ecomes

" #

Z

1

X

d d^ 1

a

 + O (m=Q): (72) d f ( )

a=h

dx dy  dxdy^

0

bj

a

x^=x =

bj

Now recall that, for exchange, d =(dx dy ) is related to the structure func-

bj

tions by Eq. (59):

" #

d 1y

2 2 2

~

= N (Q ) yF (x ;Q )+ F (x ;Q ) + O(m=Q): (73)

1 bj 2 bj

dx dy x y

bj bj

^

We de ne structure functions F for partons in the same way:

n

# "

d^ 1y

a

a 2 2 a 2

^ ~ ^

F (x = ; Q ) : (74) = N(Q ) y F (x = ; Q )+

bj bj

2 1

dxdy^ (x = )y

bj

We insert Eq. (74) into Eq. (72) and compare to Eq. (73). We deduce that the

structure functions can b e factored as

Z

1

X

1

2 a 2

^

F (x ;Q )  d f ( ) F (x = ; Q )+O(m=Q); (75)

1 bj a=h bj

1



0

a

Z

1

X

2 2

a

^

F (x ;Q )  d f ( ) F (x = ; Q )+O(m=Q): (76)

2 bj a=h bj

2

0

a

^ ^

A simple calculation gives F and F at lowest order:

1 2

1

2 a 2

^

Q (x = 1) + O ( ); (77) (x = ; Q F )=

bj s bj

a 1

2

a 2 2

^

(x = ; Q (x = 1) + O ( ): (78) F )=Q

bj bj s

2 a

Inserting these results into Eqs. (75) and (76), we obtain the lowest order relation

between the structure functions and the parton distribution functions:

X

1

2 2

F (x ;Q )  Q f (x )+O( )+O(m=Q); (79)

1 bj a=h bj s

a

2

a

X

2 2

)  Q x f (x )+O( )+O(m=Q): (80) F (x ;Q

bj a=h bj s 2 bj

a

a

The factor 1/2 b etween x F and F follows from the Feynman diagrams for spin

bj 1 2 1/2 quarks.

4.6  Dep endence

F

Figure 27: Deeply inelastic scattering with a gluon emission.

Ihave so far presented a rather simpli ed picture of deeply inelastic scattering

+

in which the hard scattering takes place on a time scale x  1=Q, while the

+

internal dynamics of the proton take place on a much longer time scale x 

2

Q=m . What happ ens when one actually computes Feynman diagrams and lo oks

at what time scales contribute? Consider the graph shown in Fig. 27. One nds

that the transverse momenta k range from order m to order Q, corresp onding to

2 + 2 2

energy scales k = k =2k between k  m =Q and k = Q =Q  Q, or time

2 +

< <

scales Q=m x 1=Q.

 

The prop erty of factorization for the cross section of deeply inelastic scattering,

emb o died in Eq. (67), is established by showing that the p erturbative expansion

can b e rearranged so that the contributions from long time scales app ear in the

parton distribution functions, while the contributions from short time scales ap-

p ear in the hard scattering functions. (See Ref. 23 for more information.) Thus,

2 2

in Fig. 27, a gluon emission with k  m is part of f ( ) , while a gluon emission

2 2

with k  Q is part of d^ .

Breaking up the cross section into factors asso ciated with short and long time

scales requires the intro duction of a factorization scale,  . When calculating

F

the diagram in Fig. 27, one integrates over k . Roughly sp eaking, one counts

2 2

the contribution from k < as part of the higher order contribution to ( ),

F

convoluted with the lowest order hard scattering function d^ for deeply inelastic

2 2

scattering from a quark. The contribution from 

F

of the higher order contribution to d^ convoluted with an uncorrected parton

distribution. This is illustrated in Fig. 28. (In real calculations, the split is

accomplished with the aid of dimensional , and is a little more subtle

than a simple division of the integral into two parts.)

Figure 28: Distance scales in factorization.

0 0

A consequence of this is that b oth d^ ( )=dE d! and f (;  ) dep end on

a F a=h F

 . Thus wehavetwo scales, the factorization scale  in f (;  ) and the

F F F

f=h

renormalization scale  in (). As with , the cross section do es not dep end on

s

 . Thus there is an equation d(cr oss section )=d = 0 that is satis ed to the

F F

accuracy of the p erturbative calculation used. If you work harder and calculate

to higher order, then the dep endence on  is less.

F

Often one sets  =  in applied calculations. In fact, it is rather common in

F

applications to deeply inelastic scattering to set  =  = Q.

F

4.7 Contour Graphs of Scale Dep endence

As an example, lo ok at the one jet inclusive cross section in proton-antiproton

collisions. Sp eci cally, consider the cross section d =dE d to make a collimated

T

spray of particles, a jet, with transverse energy E and rapidity  . (Here E is

T T

essentially the transverse momentum carried by the particles in the jet and  is

related to the angle b etween the jet and the b eam direction by   ln(tan(=2).)

We will investigate this pro cess and discuss the de nitions in the next section.

For now, all we need to know is that the theoretical formula for the cross sec-

tion at next-to-leading order involves the strong coupling () and two factors

s

f (x;  ) representing the distribution of partons in the two incoming hadrons.

F

a=h

There is a parton level hard scattering cross section that also dep ends on  and

 .

F

How do es the cross section dep end on  in () and  in f (x;  )? In

s F F

a=h

Fig. 29, I show contour plots of the jet cross section versus  and  at two

F

di erentvalues of E . The center of the plots corresp onds to a standard choice of

T

scales,  =  = E =2. The axes are logarithmic, representing log (2=E ) and

F T T

2

log (2 =E ). Thus  and  vary from E =8to2E in the plots.

F T F T T

2

Notice that the dep endence on the two scales is rather mild for the next-to-

leading order cross section. The cross section calculated at leading order is quite

2

sensitive to these scales, but most of the scale dep endence found at order s 2 2

1.05 1.05 1 0.90 1 0.90 0.95 0.95 0 1.00 0 1.00 Nco Nco 1.00 1.00

-1 0.95 -1 0.95

-2 -2 -2 -1 0 1 2 -2 -1 0 1 2

Nuv Nuv

E = 100 GeV E = 500 GeV

T T

Figure 29: Contour plots of the one jet inclusive cross section versus the renormal-

ization scale  and the factorization scale  . The cross section is d =dE d at

F T

 = 0 with E = 100 GeV in the rst graph and E = 500 GeV in the second.

T T

The horizontal axis in each graph represents N  log (2=E ) and the vertical

UV T

2

axis represents N  log (2 =E ). The contour lines show5%changes in the

CO F T

2

cross section relative to the cross section at the center of the gures. The c.m.

p

energy is s = 1800 GeV .

3

has b een canceled by the contributions to the cross section. One reads from

s

the gure that the cross section varies by roughly 15% in the central region of

the graphs, b oth for medium and large E . Following the argument of Sec. 3.4,

T

this leads to a rough estimate of 15% for the theoretical error asso ciated with

truncating p erturbation theory at next-to-leading order.

MS De nition of Parton Distribution Functions 4.8

The factorization prop erty, Eq. (67), of the deeply inelastic scattering cross sec-

tion states that the cross section can b e approximated as a convolution of a hard

scattering cross section that can b e calculated p erturbatively and parton distri-

bution functions f (x; ). But what are the parton distribution functions? This

a=A

question has some practical imp ortance. The hard scattering cross section is es-

sentially the physical cross section divided by the parton distribution function, so

the precise de nition of the parton distribution functions leads to the rules for

calculating the hard scattering functions.

The de nition of the parton distribution functions is to some extent a matter

of convention. The most commonly used convention is the MS de nition, which

arose from the theory of deeply inelastic scattering in the language of the \op erator

24

pro duct expansion." Here I will follow the (equivalent) formulation of Ref. 13.

For a more detailed p edagogical review, the reader may consult Ref. 25.

Using the MS de nition, the distribution of quarks in a hadron is given as the

hadron matrix element of certain quark eld op erators:

Z

1 dy

+

i p y +



f (;  )= e hpj (0;y ;0) F (0)jpi: (81)

F i i

i=h

2 2



Here jpi represents the state of a hadron with momentum p aligned so that

p =0. For simplicity, I take the hadron to have spin zero. The op erator (0),

T





evaluated at x = 0, annihilates a quark in the hadron. The op erator (0;y ;0)

i

+

recreates the quark at x = x = 0 and x = y , where we take the appropriate

T

Fourier transform in y so that the quark that was annihilated and recreated has

+ +

momentum k = p . The motivation for the de nition is that this is the hadron

matrix element of the appropriate numb er op erator for nding a quark.

There is one subtle p oint. The numb er op erator idea corresp onds to a par-

+

ticular gauge choice, A =0. If we are using any other gauge, we insert the

op erator

!

Z

y

+

F = P exp ig dz A (0;z ;0) t : (82)

a

a

0

The P indicates a path ordering of the op erators and color matrices along the

+

path from (0; 0; 0) to (0;y ;0). This op erator is the identity op erator in A =0

gauge and it makes the de nition gauge invariant.

The physics of this de nition is illustrated in Fig. 30. The rst picture (from

Fig. 23) illustrates the amplitude for deeply inelastic scattering. The fast proton

moves in the plus direction. A virtual photon kno cks out a quark, which emerges

moving in the minus direction and develops into a jet of particles. The second

picture illustrates the amplitude asso ciated with the quark distribution function.

We express F as F F where

2 1

 

Z

1

+



F = P exp +ig dz A (0;z ;0) t ;

2 a

a

y

 

Z

1

+

F = P exp ig dz A (0;z ;0) t ; (83)

1 a

a 0

DIS Parton distribution

Figure 30: Deeply inelastic scattering and the parton distribution functions.

and write the quark distribution function including a sum over intermediate states

jN i:

Z

X

1 dy

+

+ i p y



f (;  )= hpj (0;y ;0) F jN ihN jF (0)jpi: (84) e

i=h F i 2 1 i

2 2

N

Then the amplitude depicted in the second picture in Fig. 30 is hN jF (0)jpi.

1 i

The op erator annihilates a quark in the proton. The op erator F stands in for

1

the quark moving in the minus direction. The gluon eld A evaluated along a

lightlike line in the minus direction absorbs longitudinally p olarized gluons from

the color eld of the proton, just as the real quark in deeply inelastic scattering

can do. Thus the physics of deeply inelastic scattering is built into the de nition

of the quark distribution function, alb eit in an idealized way. The idealization is

not a problem b ecause the hard scattering function d^ systematically corrects for

the di erence b etween real deeply inelastic scattering and the idealization.

There is one small hitch. If you calculate anyFeynman diagrams for f (;  ),

F

i=h

you are likely to wind up with an ultraviolet-divergentintegral. The op erator

pro duct that is part of the de nition needs renormalization. This hitch is only a

small one. We simply agree to do all of the renormalization using the MS scheme

for renormalization. It is this renormalization that intro duces the scale  into

F

f (;  ). This role of  is in accord with Fig. 28: roughly sp eaking,  is

i=h F F F

the upp er cuto for what momenta b elong with the parton distribution function;

at the same time, it is the lower cuto for what momenta b elong with the hard scattering function.

What ab out gluons? The de nition of the gluon distribution function is similar

to the de nition for quarks. We simply replace the quark eld by suitable



combinations of the gluon eld A , as describ ed in Refs. 13 and 25.

4.9 Evolution of the Parton Distributions

Since weintro duced a scale  in the de nition of the parton distributions in

F

order to de ne their renormalization, there is a renormalization group equation

that gives the  dep endence

F

Z

1

X

d d

f (x;  )= P (x= ; ( )) f (;  ): (85)

F ab s F F

a=h b=h

d ln  

x

F

b

This is variously known as the evolution equation, the Altarelli-Parisi equation,

and the DGLAP (Dokshitzer-Grib ov-Lipatov-Altarelli-Parisi) equation. Note the

sum over parton avor indices. The evolution of, say, an (a = u) can

involve a gluon (b = g ) through the element P of the kernel that describ es gluon

ug

splitting intouu  .

The equation is illustrated in Fig. 31. When wechange the renormalization

scale  , the change in the probability to nd a parton with momentum fraction x

F

and avor a is prop ortional to the probability to nd such a parton with large

transverse momentum. The way to get this parton with large transverse momen-

tum is for a parton carrying momentum fraction  and much smaller transverse

momentum to split into partons carrying large transverse momenta, including the

parton that we are lo oking for. This splitting probability,integrated over the

appropriate transverse momentum ranges, is the kernel P .

ab Figure 31: The renormalization for the parton distribution functions.

The kernel P in Eq. (85) has a p erturbative expansion

!

2

( ) ( )

s F s F

(1) (2)

P (x= ; ( )) = P (x= ) + P (x= ) +  : (86)

ab s F

ab ab

 

The rst two terms are known and are typically used in numerical solutions of

the equation. To learn more ab out the DGLAP equation, the reader may consult

Refs. 1 and 25.

4.10 Determination and Use of the Parton Distributions

The MS de nition giving the parton distribution in terms of op erators is pro cess

indep endent|it do es not refer to any particular physical pro cess. These parton

distributions then app ear in the QCD formula for any pro cess with one or two

hadrons in the initial state. In principle, the parton distribution functions could

b e calculated by using the metho d of lattice QCD (see Ref. 25). Currently, they

are determined from exp eriment.

19

Currently, the most comprehensive analyses are b eing done by the CTEQ

26

and MRS groups. These groups p erform a \global t" to data from exp eriments

of several di erenttyp es. To p erform such a t, one cho oses a parameterization

for the parton distributions at some standard factorization scale  . Certain sum

0

rules that follow from the de nition of the parton distribution functions are built

into the parameterization. An example is the momentum sum rule:

Z

1

X

df (; )=1: (87)

a=h

0

a

Given some set of values for the parameters describing the f (x;  ), one can

a=h 0

determine f (x; ) for all higher values of  by using the evolution equation.

a=h

Then the QCD cross section formulas give predictions for all of the exp eriments

that are b eing used. One systematicallyvaries the parameters in f (x;  )to

a=h 0

obtain the best t to all of the exp eriments. One source of information ab out these

ts is the World Wide Web pages of Ref. 27.

If the freedom available for the parton distributions is used to t all of the

world's data, is there anyphysical content to QCD? The answer is yes: there are

lots of exp eriments, so this program won't work unless QCD is right. In fact, there

are roughly 1400 data in the CTEQ t and only ab out 25 parameters available to t these data.

5 QCD in Hadron-Hadron Collisions

When there is a hadron in the initial state of a scattering pro cess, there are

inevitably long time scales asso ciated with the binding of the hadron, even if part

of the pro cess is a short-time scattering. Wehave seen, in the case of deeply

inelastic scattering of a lepton from a single hadron, that the dep endence on these

long time scales can b e factored into a parton distribution function. But what

happ ens when two high-energy hadrons collide? The reader will not b e surprised

to learn that we then need two parton distribution functions.

I explore hadron-hadron collisions in this section. I b egin with the de nition

of a convenient kinematical variable, rapidity. Then I discuss, in turn, pro duction



of vector b osons ( , W , and Z ), heavy quark pro duction, and jet pro duction.

5.1 Kinematics: Rapidity

In describing hadron-hadron collisions, it is useful to employ a kinematic variable y

that is called rapidity. Consider, for example, the pro duction of a Z b oson plus

anything, p +p ! Z+X. Cho ose the hadron-hadron c.m. frame with the z

axis along the b eam direction. In Fig. 32, I show a drawing of the collision.

The arrows represent the momenta of the two hadrons; in the c.m. frame, these

momenta have equal magnitudes. We will want to describ e the pro cess at the

parton level, a + b ! Z + X . The two partons a and b each carry some share

of the parent hadron's momentum, but generally these will not b e equal shares.

Thus the magnitudes of the momenta of the colliding partons will not b e equal.

We will have to b o ost along the z axis in order to get to the parton-parton c.m.

frame. For this reason, it is useful to use a variable that transforms simply under

b o osts. This is the motivation for using rapidity.

 +

Let q =(q ;q ;q) b e the momentum of the Z b oson. Then the rapidityof

the Z is de ned as

!

+

1 q

y = ln : (88)

2 q

+

The four comp onents (q ;q ;q) of the Z b oson momentum can b e written in terms

of four variables, the two comp onents of the Z b oson's transverse momentum q ,

its mass M , and its rapidity:

q q

y  y

2 2 2 2

(q +M )=2; e (q +M )=2; q): (89) q =(e

Figure 32: Collisionoftwo hadrons containing partons pro ducing a Z b oson. The

c.m. frame of the two hadrons is normally not the c.m. frame of the two partons

that create the Z b oson.

The utility of using rapidity as one of the variables stems from the transfor-

mation prop erty of rapidity under a b o ost along the z axis:

+ ! + !

q ! e q ; q ! e q ; q ! q: (90)

Under this transformation,

y ! y + !: (91)

This is as simple a transformation lawaswe could hop e for. In fact, it is just

the same as the transformation law for velo cities in nonrelativistic physics in one

dimension.

Figure 33: De nition of the p olar angle  used in calculating the rapidityofa

massless particle.

Consider now the rapidity of a massless particle. Let the massless particle

emerge from the collision with p olar angle  , as indicated in Fig. 33. A simple

calculation relates the particle's rapidity y to  :

y = ln (tan(=2)) ; (m =0): (92)

Another way of writing this is

tan  =1=sinh y; (m =0): (93)

One also de nes the pseudorapidity  of a particle, massless or not, by

 = ln (tan(=2)) or tan  =1=sinh : (94)

The relation b etween rapidity and pseudorapidityis

q

2

2

sinh  = 1+m =q sinh y: (95)

T

Thus, if the particle isn't quite massless,  may still b e a go o d approximation to

y .



5.2 , W , Z Pro duction in Hadron-Hadron Collisions

Consider the pro cess

A + B ! Z + X; (96)

where A and B are high-energy hadrons. Two features of this reaction are im-

p ortant for our discussion. First, the mass of the Z b oson is large compared to

1 GeV, so that a pro cess with a small time scale t  1=M must b e involved in

Z

the pro duction of the Z . Atlowest order in the strong interactions, the pro cess

is q +q ! Z. Here the quark and antiquark are constituents of the high-energy

hadrons. The second signi cant feature is that the Z b oson do es not participate

in the strong interactions, so that our description of the observed nal state can

be very simple.



We could equally well talk ab out A + B ! W + X or A + B ! + X where

the virtual photon decays into a muon pair or an electron pair that is observed,

 28

and where the mass of the is large compared to 1 GeV. This last pro cess,

 +

A + B ! + X ! ` + ` + X , is historically imp ortant b ecause it help ed establish

the parton picture as b eing correct. The W and Z pro cesses were observed later.

In fact, these are the pro cesses by which the W and Z b osons were rst directly

29

observed.

In pro cess (96), we allow the Z b oson to haveany transverse momentum q .

(Typically, then, q will b e much smaller than M .) Since weintegrate over q and

Z

the mass of the Z b oson is xed, there is only one variable needed to describ e

the momentum of the Z b oson. Wecho ose to use its rapidity y , so that we are

interested in the cross section d =dy .

The cross section takes a factored form similar to that found for deeply inelastic

scattering. Here, however, there are two parton distribution functions:

Z Z

1 1

X

d^ (;  ) d

ab F

 : (97) d d f ( ; ) f ( ; )

A B a=A A F b=B B F

dy dy

x x

A B a;b

Figure 34: A Feynman diagram for Z b oson pro duction in a hadron-hadron col-

lision. Two partons, carrying momentum fractions  and  , participate in the

A B

hard interaction. This particular Feynman diagram illustrates an order contri-

s

bution to the hard scattering cross section: a gluon is emitted in the pro cess of

making the Z b oson. The diagram also shows the decay of the Z b oson into an

electron and a neutrino.

The meaning of this formula is intuitive: f ( ; )d gives the probabilityto

a=A A F A

nd a parton in hadron A; f ( ; )d gives the probability to nd a parton

B f B

b=B

in hadron B ; d^ =dy gives the cross section for these partons to pro duce the

ab

observed Z b oson. The formula is illustrated in Fig. 34. The hard scattering cross

section can b e calculated p erturbatively. Fig. 34 illustrates one particular order

contribution to d^ =dy . The integrations over parton momentum fractions

s ab

have limits x and x , which are given by

A B

q q

y y

2 2

x = e M =s; x = e M =s: (98)

A B

Equation (97) has corrections of order m=M , where m is a mass characteristic

Z

of hadronic systems, say 1 GeV. In addition, when d^ =dy is calculated to order

ab

N N +1

, then there are corrections of order .

s s

There can b e soft interactions b etween the partons in hadron A and the partons

in hadron B , and these soft interactions can o ccur b efore the hard interaction that

creates the Z b oson. It would seem that these soft interactions do not t into

the intuitive picture that comes along with Eq. (97). It is a signi cant part of

the factorization prop erty that these soft interactions do not mo dify the formula.

These intro ductory lectures are not the place to go into how this can b e. For more

information, the reader is invited to consult Ref. 23.

5.3 Heavy Quark Pro duction

Wenow turn to the pro duction of a heavy quark and its corresp onding antiquark

in a high-energy hadron-hadron collision:



A + B ! Q + Q + X: (99)

The most notable example of this is top quark pro duction. A Feynman diagram

for this pro cess is illustrated in Fig. 35.

Figure 35: Feynman graph for heavy quark pro duction. The lowest order hard

2



pro cess is g + g ! Q + Q, which o ccurs at order . This particular Feynman

s

3

diagram illustrates an order pro cess in which a gluon is emitted.

s

The total heavy quark pro duction cross section takes a factored form similar

to that for Z b oson pro duction,

Z Z

1 1

X

ab

  d d f ( ; ) f ( ; )^ ( ;): (100)

T A B a=A A F b=B B F F

T

x x

A B

a;b



As in the case of Z pro duction, QQ pro duction is a hard pro cess, with a time scale

determined by the mass of the quark: t  1=M . It is this hard pro cess that is

Q

ab

represented by the calculated cross section ^ . Of course, the heavy quark and

T

antiquark have strong interactions, and can radiate soft gluons or exchange them

with their environment. These e ects do not, however, a ect the cross section:



once the QQ pair is made, it is made. The probabilities for it to interact in various

ways must add to one. For an argument that Eq. (100) is correct, see Ref. 30.

5.4 Jet Pro duction

In our study of high-energy electron-p ositron annihilation, we discovered three

things. First, QCD makes the qualitative prediction that particles in the nal

state should tend to b e group ed in collimated sprays of hadrons called jets. The

jets carry the momenta of the rst quarks and gluons pro duced in the hard pro cess.

Second, certain kinds of exp erimental measurements prob e the short-time physics

of the hard interaction, while b eing insensitive to the long-time physics of parton

splitting, soft gluon exchange, and the binding of partons into hadrons. Such

measurements are called infrared safe. Third, among the infrared safe observables

are cross sections to make jets.

Figure 36: Sketch ofatwo-jet event at a hadron collider. The cylinder represents

the detector, with the b eam pip e along its axis. Typical hadron-hadron collisions

pro duce b eam remnants, the debris from soft interactions among the partons. The

particles in the b eam remnants have small transverse momenta, as shown in the

sketch. In rare events, there is a hard parton-parton collision, which pro duces jets

with high transverse momenta. In the event shown, there are two high P jets.

T

These ideas work for hadron-hadron collisions to o. In such collisions, there

is sometimes a hard parton-parton collision, which pro duces two or more jets, as

depicted in Fig. 36. Consider the cross section to make one jet plus anything else,

A + B ! jet + X: (101)

Let E b e the transverse energy of the jet, de ned as the sum of the absolute values

T

of the transverse momenta of the particles in the jet. Let y b e the rapidityofthe

jet. Given a de nition of exactly what it means to have a jet with transverse

energy E and rapidity y , the jet pro duction cross section takes the familiar

T

factored form

Z Z

ab

1 1

X

d d^ (;  )

F

 : (102) d d f ( ; ) f ( ; )

A B a=A A F b=B B F

dE d dE d

x x

A B

T T

a;b

What shall wecho ose for the de nition of a jet? At a crude level, high E jets

T

are quite obvious and the precise de nition hardly . However, if wewant

Figure 37: AFeynman diagram for jet pro duction in hadron-hadron collisions.

2

The leading order diagrams for A + B ! jet+X o ccur at order . This particular

s

3

diagram is for an interaction of order . When the emitted gluon is not soft or

s

nearly collinear to one of the outgoing quarks, this diagram corresp onds to a nal

state like that shown in the small sketch, with three jets emerging in addition to

the b eam remnants. Any of these jets can b e the jet that is measured in the one

jet inclusive cross section.

to make a quantitative measurement of a jet cross section to compare to next-to-

leading order theory, then the de nition do es matter. There are several p ossibil-

ities for a de nition that is infrared safe. The one most used in hadron-hadron

collisions is based on cones.

31

In the standard Snowmass Accord de nition, one imagines that the exp er-

imental calorimeter is divided into small angular cells lab eled i in  - space, as

depicted in Fig. 38. We can say that a jet consists of all the particles that fall into

certain of the calorimeter cells, or we can measure the E in each cell and build

T

the jet parameters from the cell variables (E ; ; ). We then say that a jet

Ti i i

consists of the cells inside a certain circle in  - space. The circle has a radius R ,

usually chosen as 0.7 radians, and is centered on a direction ( ; ). Thus the

J J

calorimeter cells i included in the jet ob ey

2 2 2

(  ) +(  )

i J i J

The transverse energy of the jet is de ned to b e

X

E = E : (104)

T;J T;i

i2cone

The direction of the jet is de ned to b e the direction ( ; ) of the jet axis, which

J J

is chosen to ob ey

X

1

E  ; (105)  =

T;i i J

E

T;J

i2cone

X

1

 = E  : (106)

J T;i i

E

T;J

i2cone

Of course, if one picks a trial jet direction ( ; ) to de ne the meaning of

J J

\i 2 cone" and then computes ( ; ) from these equations, the output jet di-

J J

rection will not necessarily match the input cone axis. Thus one has to treat the

equations iteratively until a consistent solution is found.

Figure 38: Jet de nition according to the Snowmass algorithm. The shading of

the squares represents the density of transverse energy as a function of azimuthal

angle  and pseudorapidity  . The cells inside the circle constitute the jet.

Note that the Snowmass algorithm for computing E ; ; is infrared safe.

T;J J J

In nitely soft particles do not a ect the jet parameters b ecause they enter the

equations with zero weight. If two particles have the same angles ; , then it

do es not matter if we join them together into one particle b efore applying the

algorithm. For example,

E  + E  =(E + E ) : (107)

T;1 T;2 T;1 T;2

Note, however, that the Snowmass de nition given ab ove is not complete. It

is p erfectly p ossible for two or more cones that are solutions to Eqs. (104, 105,

106) to overlap. One must then have an algorithm to assign calorimeter cells to

one of the comp eting jets, thus splitting the jets, or else to merge the jets. When

supplemented by an appropriate split/merge algorithm, the Snowmass de nition

is not as simple as it seemed at rst.

3

In an order p erturbative calculation, one simply applies this algorithm at

s

the parton level. At this order of p erturbation theory, there are two or three

partons in the nal state. In the case of three partons in the nal state, twoof

them are joined into a jet if they are within R of the jet axis computed from

the partonic momenta. The split/merge question do es not apply at this order of

p erturbation theory.

I showed a comparison of the theory and exp eriment for the one jet inclusive

cross section in Fig. 19.

I should record here that the actual jet de nitions used in current exp eriment

are close to the Snowmass de nition given ab ove but are not exactly the same.

Furthermore, there are other de nitions available that may come into use in the

future. There is not time here to explore the issues of jet de nitions in detail.

What I hop e to have done is give the outline of one de nition and to explore what

the issues are.

6 Epilogue

QCD is a rich sub ject. The theory and the exp erimental evidence indicate that

quarks and gluons interact weakly on short time and distance scales. But the net

e ect of these interactions extending over long time and distance scales is that

the chromo dynamic is strong. Quarks are b ound into hadrons. Outgoing

partons emerge as jets of hadrons, with each jet comp osed of sub jets. Thus QCD

theory can b e viewed as starting with simple p erturbation theory, but it do es not

end there. The challenge for b oth theorists and exp erimentalists is to extend the

range of phenomena that we can relate to the fundamental theory.

I thank F. Hautmann for reading the manuscript and helping to eliminate

some of the mistakes.

References

[1] G. Sterman et al., "Handb o ok of p erturbative QCD," Rev. Mo d. Phys. 67,

157 (1995).

[2] R. K. Ellis, W. J. Stirling, and B. R. Webb er, QCD and Col lider Physics

(Cambridge University Press, Cambridge, 1996).

[3] L. Brown, Quantum Theory (Cambridge University Press, Cambridge, 1992).

[4] G. Sterman, An Introduction to (Cambridge Univer-

sity Press, Cambridge, 1993).

[5] M. Peskin and D. V. Schro eder, An Introduction to Quantum Field Theory

(Addison-Wesley, Reading, 1995).

[6] S. Weinb erg, The Quantum Theory of Fields (Cambridge University Press,

Cambridge, 1995).

[7] QCD and Beyond, Proceedings of the 1995 Theoretical Advanced Studies In-

stitute, edited by D. E. Sop er, Boulder, 1995, (World Scienti c, Singap ore,

1996).

[8] G. Sterman, Phys. Rev. D 17, 2773 (1978); 17, 2789 (1978).

[9] J. C. Collins and D. E. Sop er, Nucl. Phys. B 193, 381 (1981).

[10] Z. Kunszt and D. E. Sop er, Phys. Rev. D 46, 192 (1992).

[11] S. Bethke, Z. Kunszt, D. E. Sop er, and W. J. Stirling, Nucl. Phys. B 370,

310 (1992).

[12] C. L. Basham, L. S. Brown, S. D. Ellis, and S. T. Love, Phys. Rev. D 19,

2018 (1979).

[13] J. C. Collins and D. E. Sop er, Nucl. Phys. B 194, 445 (1982); G. Curci,

~

W.Furmanski, and R. Petronzio, Nucl. Phys. B 175, 27 (1980).

[14] J. C. Collins, Renormalization (Cambridge University Press, Cambridge,

1984).

[15] L. Surguladze and M. Samuel, Rev. Mo d. Phys. 68, 259 (1996).

[16] D. E. Sop er and L. Surguladze, Phys. Rev. D 54, 4566 (1996).

[17] P. M. Stevenson, Phys. Rev. D 23, 2916 (1981).

[18] S. J. Bro dsky,G.P. Lepage, and P. MacKenzie, Phys. Rev. D 28, 228 (1983).

[19] H. Lai et al., e-Print Archive hep-ph/9606399, Phys. Rev. D 55 (to b e pub-

lished).

[20] CDF Collab oration (F. Ab e et al.), Phys. Rev. Lett. 77, 438 (1996); B.

Flaugher, CDF Collab oration, Proceedings of the XI Topical Workshop on

ppbar Col lider Physics,Padova, Italy,May 1996.

[21] D0 Collab oration: N. Varelas, Proceedings of the International Conference

on High Energy Physics,Warsaw, July 1996.

[22] J. Huston et al.,Phys. Rev. Lett. 77, 444 (1996).

[23] J. C. Collins, D. E. Sop er, and G. Sterman, in Perturbative QCD, edited by

A. Mueller (World Scienti c, Singap ore, 1989).

[24] W. A. Bardeen, A. J. Buras, D. W. Duke, and T. Muta, Phys. Rev. D 18,

3998 (1978).

[25] D. E. Sop er, in Lattice '96 International Symposium on Lattice Field The-

ory, edited by M. Golterman et al., St. Louis, June 1996 (Elsevier Science,

Amsterdam, to b e published).

[26] A. D. Martin, R. G. Rob erts, and W. J. Stirling, Phys. Lett. B 387, 419

(1996).

[27] P. Anandam and D. E. Sop er, \A Potp ourri of Partons",

http://zebu.uoregon.edu/~parton/.

[28] S. D. Drell and T.-M. Yan, Ann. Phys. 66, 578 (1971).

[29] UA1 Collab oration (G. Arnison et al.), Phys. Lett. B 122, 103 (1983);

UA2 Collab oration (G. Banner et al.), Phys. Lett. B 122, 476 (1983).

[30] J. C. Collins, G. Sterman, and D. E. Sop er, Nucl. Phys. B 263, 37 (1986).

[31] J. Huth et al.,inProceedings of the Summer Study on High Energy Physics

in the 1990s, Research Directions for the Decade, edited by E. Berger, Snow-

mass, Colorado, 1990 (World Scienti c, Singap ore, 1992), p. 134.