Sunteți pe pagina 1din 28

Chapter 11 Elementary Scattering Theory

As seen in the previous chapter, the bound state problem is characterized through the stationary, normalizable states in the Hilbert space. Since quantum mechanical states in principle always have to be normalized (because of the statistical interpretation of the wave functions), scattering states should be non-stationary, normalizable solutions of the Schrodinger equation. An elementary, conceptually not quite satisfactory, however, in practical applications extremely successful approach, consists of dropping the normalization condition. This allows to introduce scattering states as suitably chosen stationary, non-normalizable solutions of the Schrodinger equation. Stationary states are in principle eigenstates of the Hamiltonian. If they are supposed to be non-normalizable, they have to be special solutions from the continuous spectrum of H .

11.1 Free Motion


The force free, single particle motion (free motion) is given by a Hamiltonian, often called free Hamiltonian, ~2 H0 = Pm : (11.1) 2 Possible eigenvalues of H0 are the momentum eigenstates de ned by ~ ~ p ~ (11.2) P j 'p i = ~ j ' p i : 251

With (11.2) and (11.1) follows

~2 H0 j 'pi = Pm j 'pi = Ep j 'pi ; ~ ~ ~ 2 the "norm" of those states can be chosen as ~ p h'p j 'pi = (p 0 ? ~) : ~ ~
0

(11.3)

(11.4) ~ The speci c form of j 'pi is obtained when the explicit representation of the operator P ~ in coordinate space is employed: ~ ~ ~x h~ j P j 'pi = h r h~ j 'pi = h r 'p(~ ) = ~ 'p(~ ) ; x ~ ~ p ~x (11.5) i x ~ i which has as solution of the di erential equation 1 ~x x ~ (11.6) 'p(~ ) = (2 h)3=2 e hi p ~ := h~ j pi ; ~ x which corresponds to a plane wave. The norm is explicitly given as i ~ x i ~x 1 Z d3x e? h p ~ e h p ~ = (p 0 ? ~) ; h'p j 'pi = (2 h)3 ~ p (11.7) ~ ~ thus, the norm in a regular sense does not exist.
0

In general, the time evolution of solutions of the free Schrodinger equation is given by d (11.8) ? h dt j '(t)i = H0 j '(t)i i with the solution j '(t)i = e? hi H0 t j 'i : (11.9) With j 'pi as starting vector follows with (11.3) ~ (11.10) This is a stationary solution with the typical time dependence given in a phase factor. It is obviously not normalizable ~ ~ (11.11) h'p (t) j 'p(t)i = h'p j 'pi = (p 0 ? p) : ~ ~ ~ ~ Inserting (11.6) into (11.10) gives the explicit representation i ~x 1 'p(~ ; t) = e? hi Ept 'p(~ ) = (2 h)3=2 e h (p ~ ?Ept) ; (11.12) ~x ~ x
0 0

j 'p(t)i = e? hi Ept j 'pi : ~ ~

252

which is just the De-Broglie wave. In coordinate space, the physical interpretation of (11.12) can be readily studied. We have a plane wave. Positions with the same phase

p ~ ? Ept = constant ~ x (11.13) spread with the phase velocity 2 x ~ph = d~ = Ep p = p 2pm p = 2p p = v p ; v ^ 1 ^ (11.14) dt p m^ 2^ where p = p= j ~ j indicates the direction of the spread. This leads intuitively to the ^ ^ p fact that plane waves are used to describe the motion of quanta with a de nite starting momentum p. ~

11.2 Free Wave Packets


Especially at the beginning, it is useful to understand that in principle we should have described the free motion with a normalizable state j 'ai. We should work with a packet

j 'a i =

d3p j 'pih'p j 'ai = ~ ~

d3p j 'pi 'a (p) ; ~ ~ ~

(11.15)

where 'a (p) is the experimentally given distribution of momenta, e.g., around a speci c ~ ~ value ~a. Of course, we can assume that this distribution is nite, so that the integral p over j 'a (p) j2 exists and can be set to one. This means, we can require ~ ~

k 'a = h'a j 'a i = k2

d3p

h'a j 'pih'p j 'ai = ~ ~

d3p j 'a(p) j2 = 1 : (11.16) ~ ~

In contrast to j 'pi this normalizable state j 'ai is not an eigenvector of H0 to a de nite ~ energy E , since Z Z 2 3 p H j ' i ' (p) = 3 p p j ' i ' (p) = E j ' i : (11.17) H0 j 'a i = d 0 p ~a ~ d 2m p ~a ~ 6 ~ a ~ The general solution (11.9) of the time-dependent Schrodinger equation (11.8) reads with the initial state j 'a i:

j 'a(t)i =

d3p e? hi H0 t

j 'pi 'a(p) = ~ ~ ~

d3p? hi

p2 2m

j 'pi 'a (p) ; ~ ~ ~

(11.18)

and similarly to (11.17) the time dependence cannot be pulled in front of the integral. One has instead

j 'a(t)i 6= e? hi Et j 'a i :
253

(11.19)

Using (11.12) we obtain for the wave packet in coordinate space representation Z 1 ~x 'a (~ ; t) = (2 h)3=2 d3p e hi (p ~ ?Ept) 'a (p) : x ~ ~ (11.20) It can be shown that for a general initial state j '(0)i the explicit time evolution is given by 3=2 Z x x 2 x (11.21) d3x0 e hi mt (~ ?~ )2 '(~ ; 0) : '(~ ; t) = 2 mht x i Estimating the absolute value gives 1 m 3=2 Z d3x0 j e 2imt (~ ?~ )2 j j '(~ ) j x x j '(~ ; t) j x x j t j3=2 2 ih = constant : j t j3=2 (11.22)
0 0

Thus for the probability density, we obtain

constant : (11.23) j t j3=2 However, the total probability, which corresponds to the conservation of the number of particles is given by (~ ; t) = j '(~ ; t) j2 x x

k '(t) k2 = h'(t) j '(t)i = he? hi H t '(0) j e? hi H t '(0)i = h'(0) j '(0)i = k '(0) k2 :


0 0

(11.24)

This means that the total probability is independent of t; however, in (11.21) one has to integrate over increasingly larger areas in order to compensate the jtj1=2 behavior. 3

11.3 Scattering States


As we saw with respect to the free motion, the energy eigenvalues in the continuous spectrum are in nitely degenerate. In this simple case, it was possible to explicitly give the physically realized solutions, the plane waves. If a potential V is present, this is in general no longer the case. Here we need boundary conditions in order to pick the physically relevant solutions among the in nitely many possible solutions. If we assume that the region of the interaction V is specially restricted (short-ranged 254

Asymptotic Conditions:

(+) potential), then it seems plausible to assume that the scattering solution j p i consists ~ sc of a superposition of the free solution and a scattering piece j p i, which is given by the potential. According to the previous discussion, one chooses as the free solution the momentum state j 'pi. We, therefore, assume that the solution of the full eigenvalue ~ equation, which is associated with an incoming particle with sharp momentum ~, p 2 (+) (+) (+) H j p i = E j p i = 2pm j p i (11.25) ~ ~ ~ can be written as (+) sc (11.26) j p i = j 'p i + j p i : ~ ~ ~ In principle, this is not yet a real statement, since one can always separate a piece j 'pi ~ o any vector (which is normalized on a function). The ansatz (11.26) becomes less sc trivial if we assume that the piece j p i is caused by the potential, and thus vanishes if ~ the potential vanishes. This transition can be achieved by multiplying a given potential with a factor and one studies the limit ! 0.

sc In order to obtain a detailed requirement for j p i, we use the coordinate space repre~ sentation. Here (11.26) reads i p~ 1 (+) ~ x + sc (~ ) : (11.27) p x ~ p (~ ) = (2 h)3=2 e h ~ x sc As already mentioned, p (~ ) is in general in a complicated way given by the potential ~ x V . If one considers how the in uence of a perturbation (here a short-ranged potential) in uences a plane wave, it is reasonable to require as physically motivated boundary (+) condition that p (~ ) behaves asymptotically as an outgoing spherical wave, i.e., ~ x i 1 f (^; p) e h pr ; sc(~ ) r!1 (11.28) p x ?! (2 h)3=2 E x ^ r ~ 2 where E = 2pm . With this we x the scattering solution as the speci c solution of (11.25) for which in the coordinate space representation the condition (11.28) holds. This means they are de ned by the following requirements # " h2 r2 + V (~ ) (+) (~ ) = E (+) (~ ) ~ x p x ? 2m (11.29) ~ p x ~
i ~x 1 e hi pr 5 : (+) (~ ) r!1 4e h p ~ + f (^; p) (11.30) E x ^ p x ?! (2 h)3=2 ~ r sc The factor fE (^; p) is a measure for the magnitude of the scattering part p (~ ) of the x^ ~ x scattering solution and is called scattering amplitude. According to our assumptions fE (^; p) has to vanish if the potential is zero. The condition (11.30) is also known as x^ Sommerfeld Radiation Condition.
2 3

255

11.4 Cross Section


The scattering of a particle under the in uence of a potential can be described in the following fashion. From the incoming current ~0 of particles the scattering process creates j ~ sc of scattered particles. a current j

n . dF

j0

z-axis

Fig. 11.1

Explanation of the di erential cross section.

scattering center

Experimentally one measures at a distance r from the scattering center the probability current which penetrates through an area (11.31) dF = r2 d : Here d is a solid angle which is given by the resolution of the detector (ideally d is in nitesimally small). This current is then related to the incoming current, i.e., one considers the quantity ~ sc n ~ sc n 2 d = j ~~ dF = j ~~ r d : (11.32) j j0 j j j0 j Here ~ characterizes the location of the areal element dF through which the scattered n particles penetrate. This de nition (11.32) shows that d has the dimension of an area. The current is given by the probability density multiplied by the velocity. For the incoming plane wave, this means i ~x 1 eh p ~ 2 ~ = ~ v ~0 = (11.33) j v (2 h)3=2 ; (2 h)3=2 whereas for the scattered wave i 1 f (^; p) 1 e h pr 2 v0~ ~ sc = j sc j2 v0 ~ = n j n 3=2 E x ^ r (2 h) 256

x^ 2 0 n = j fE (^2; p) j v 3 ~ : r (2 h)

(11.34)

Here the current points in direction of ~ , i.e., it is perpendicular to the corresponding n spheres whose center is given by the scattering center. Inserting (11.33) and (11.34) into (11.32) leads to
0 0 d = j fE (^; p) j2 v d = j fE (^; p) j2 E d : x^ v x^ (11.35) E The factor r12 in (11.34) is canceled through the multiplication with the factor r2 from the area dF in (11.31). This is important since then the result (11.35) does not depend on the distance r at which the detector is positioned. Experimentally one only has to make sure that this detector is positioned outside the interaction region, where the asymptotic conditions are realized. When considering short-ranged potentials, this is trivially ful lled, however, leads to complications when considering the quantum mechanical treatment of Coulomb scattering.
s

In (11.33) and (11.34) v and v0 denote the velocity of the incoming and outgoing particles. Correspondingly the factor

v 0 = p0 = E 0 (11.36) v p E occurs in (11.35). However, in the here considered case of potential scattering v = v0, a condition which as been assumed in the asymptotic condition (11.30). Scattering with v = v0 is also called elastic scattering. From (11.35) we derive that the di erential d cross section d is given through the scattering amplitude fE (^; p) according to x^ d = j f (^; p) j2 : (11.37) E x ^ d The amplitude fE has obviously the dimension of a length, which can be immediately seen from (11.30). In (11.35) the factor (11.36) has been considered explicitly to point out that in, e.g., inelastic processes energy can be lost in the scattering process, e.g., through the excitation of composite targets, and then the factors (11.36) have to occur in the cross section.

Spatial Symmetries
if the potential is rotationally symmetric or has an axial symmetry with respect to the direction p of the incoming current, one has to expect that the scattering amplitude will ^ 257

only depend on the azimuthal angle with respect to the direction of the incoming particle. Choosing the incoming direction as z-axis, we obtain for the asymptotic condition (11.30)
i 1 r!1 (+) p (~ ) ?! (2 h)3=2 4e h ~ x
2

pz

e hi pr + fE ( ) r

3 5

(11.38)

and the di erential cross section becomes d = j f ( ) j2 : E d

(11.39)

Total Cross Section


If one sums up the scattering into all di erent directions, i.e., integrates the di erential cross section over all angles, one obtains the total cross section Z d = Z d j f (^; p) j2 : d d (11.40) tot = E x ^ If the potential has the symmetries leading to (11.39), then this de nition simpli es to
tot

d j fE ( ) j 2 =

= 2

0 Z 1

sin d d' j fE ( ) j2

?1

d cos j fE ( ) j2 :

(11.41)

In the preceding discussion we introduce the probability current densities by multiplying the probabilities with the corresponding velocities, as is customary in electro dynamics, where charge density times velocity is introduced as current density. This is a quite intuitive procedure and leads quickly to the desired results. However, we could also have used the more formal de nition of a current density ~ (~ ; t) = h =m( (~ ; t)r (~ ; t)) : jx x ~ x (11.42) m The current of the incoming particles is then obtained by inserting the plane wave:
~x ~ i ~x ~0 (~ ; t) = h =m 1 3 e? hi p ~ r e h p ~ j x m (2 h) p ~ h ip = m (2 1h)3 =m h ~ = (2 =m 3 ; h)
" #

(11.43)

258

(11.44) x ~ @ n ~ Here we used that ~ = ~ , from which follows ~ r = ~ r = @r . Now we need the n x r r scattered current at the place of the detector, i.e., large r, and thus can use (11.30). This gives # " x^ 1 =m fE (^; p) i p fE (^; p) x^ sc = ~ ~ sc(~ ; t) r!1 h jr n j x ?! m (2 h)3 r h r + higher orders in 1 : r (11.45) Aside from higher orders, which are negligible for su ciently large r, we obtain x^ 2 p ~ ~ sc r?! (2 1h)3 j fE (^2; p) j m ; n j !1 (11.46) r which corresponds to (11.34). Inserting (11.43) and (11.45) into (11.32), where now automatically p0 = p, we obtain the di erential cross section (11.37).

a result which coincides with (11.33). Similarly, we can determine the current of scattering particles in direction ~ and obtain n i h h x n ~ jrsc = ~ ~ sc(~ ; t) = m =m sc (~ ; t)~ r sc(~ ; t) n j x x " # h =m sc (~ ; t) @ sc(~ ; t) : = m x @r x

11.5 Phase Shifts


The task is to determine fE ( ; '), i.e., the solution of the Schrodinger equation 2 ? 2hm r2 p(~) + V (j ~ j) p(~) = Ep p(~) : r ~r (11.47) ~ r ~ ~ r If V is rotationally symmetric, i.e., H; L2 ] = 0, then we can write X a`(p) u`;p(r) Y`m( ; ') : ^ (11.48) p (~) = ~ r r `;m For p parallel to the z-axis, is the angle between p and the axis, and the problem is ~ ~ symmetric around the z-axis, i.e., independent of '. Thus we use 1 X (11.49) a`(p) u`;p(r) P`(cos ) : ~ r p (~) = ~ r `=0 259

Inserting in (11.47) gives as radial Schrodinger equation

d2 u (r) + p2 ? 2m V (r) ? `(` + 1) u (r) = 0 `;p dr2 `;p r2 h2


" #

(11.50)

2 where we used that Ep = 2pm . The requirement that only the solution which is regular at the origin is physically allowed gives u`;p(0) = 0 and leads to a unique solution of (11.50). Next, we need to consider the behavior of u`;p(r) for r ! 1. If the potential V (r) vanishes su ciently strong for r ! 1, then (11.50) goes asymptotically for large r to # " d2 u (r) + p2 ? `(` + 1) u (r) = 0 ; (11.51) `;p dr2 `;p r2

which has the general solution

u`;p(r) = B`;pr j`(pr) + C`;pr n`(pr) ;

(11.52)

where j`(pr) and n` (pr) are the Bessel and Neumann functions. They have the following properties:

pr << 1 : j` (pr) (pr)` = regular solution ^ n`(pr) ?(pr)?`?1 = irregular solution ^ 1 pr >> 1 : j` (pr) ; pr sin pr ? `2
!

n`(pr)

1 ? pr cos pr ? `2

Thus, we have the following asymptotic behavior for the radial component of the scattering solution:
`;p(r)

u`;p(r) r

pr>>1

sin pr ? B`;p pr

` 2

cos pr ? ? C`;p pr

` 2

(11.53)

or equivalently

u`;p(r) pr>>1 A sin pr ? `2 + `(p) : (11.54) ; `;p r pr Thus `(p) in (11.54) is a phase shift originating from the potential, which describes a shift with respect to the free wave (solution). The phase shifts ` do not only depend
260

on the potential but also on the scattering energy Ep, thus ` ` (p). (11.54) can be rewritten as " ! ! 1 sin pr ? ` + (p) = A 1 sin pr ? ` cos (p) A`;p pr ` `;p 2 pr 2 # ` ! + cos pr ? `2 sin `(p) ; (11.55) and thus we have u`;p(r) ; A cos sin pr ? `2 + A sin cos pr ? `2 : (11.56) `;p ` `;p ` r pr pr A comparison of the coe cients of (11.56) and (11.53) gives B`;p = A`;p cos ` C`;p = ?A`;p sin ` (11.57) with A`;p being a free parameter, which is determined by the overall normalization of the wave function. Thus, the radial solution can be written as u`;p(r) = A cos j (pr) ? sin n (pr)] : (11.58) `;p ` ` ` ` r Through this equation (11.58) the phase shifts ` are de ned. The radial solution R`;p(r) can be further rewritten as (11.59) R`;p(r) = u`;p(r) = A`;p j`(pr) ? tan ` n`(pr)] r with A`;p = A`;p cos ` = B`;p (11.60) and sin C tan ` = cos ` = ? B`;p : (11.61) ` `;p In practical applications, one obtains the phase shift tan ` by solving the radial Schrodinger equation (11.50) up to a value of r where the potential has su ciently fallen o (in case of V being the nucleon-nucleon (nn) interaction typical values of r are 15fm). Eq. (11.59), together with its derivative, gives two equations to determine tan ` as 0 (11.62) tan ` = p j0`(pr) ? j`(pr) p n`(pr) ? n`(pr) 261

where is the logarithmic derivative of the solution R`;p(r) = 1 d R`;p(r) : (11.63) R`;p(r) dr Once the phase shift ` is obtained, the asymptotic behavior of the scattering solution is given as (cp. 11.49) 3 2 1 cos pr ? `2 5 sin pr ? `2 X + sin ` P`(cos ) :(11.64) a`(p) 4cos ` p (~) ; ~ r pr pr `=0 Since we are looking for a solution with a de nite asymptotic behavior (11.30), we have to decompose (11.30) with respect to spherical waves. With
1

eipz = fE ( ) =
we can write (11.30) as

`=0
X

i` (2` + 1) j`(pr) P`(cos )


(2` + 1) f`(p) P`(cos ) (11.65)
3

`=0

ipr sin pr ? `2 4i` (2` + 1) + f`(p) er 5 P`(cos ) p (~) ; ~ r pr `=0 # " 1 X eipr ? (?1)` e?ipr + f (p) eipr P (cos ) = (2` + 1) 2ipr ` ` 2ipr r `=0 " # ! 1 X (2` + 1) e?ipr + (2` + 1) 1 + f (p) eipr P (cos ) : (?1)`+1 = ` ` 2ip r 2ip r `=0 (11.66) ` ` 1 1 Here we used that i` = ei` 2 and pr sin pr ? `2 = 2ipr ei(pr? 2 ) ? e?i(pr? 2 ) . In (11.64) the asymptotic form of the scattering solution was given. If one does not use this form, the original wave function reads
X

p (~) ~ r

Comparing the coe cients in (11.67) and (11.66) leads to the determination of a` (p) and f`(p) via a` (p) = (2` + 1) i` ei L (p) e2i ` (p) = 1 + 2ip f`(p) (11.68) 262

`=0

a`(p) cos `(p) j`(pr) ? sin `(p) n`(pr)] P`(cos ) :

(11.67)

from which follows or

f`(p) = 21 e2i ` (p) ? 1] ip

(11.69) (11.70)

1 1 f`(p) = p ei ` (p) sin `(p) =: p t`(p) ; where t`(p) is called the partial wave amplitude.

Thus, the solution of the Schrodinger equation for positive energies is given by 1 (+) (~) = X (2` + 1) i` ei ` (p) u`;p(r) P (cos ) (11.71) r ` p ~ r `=0 and 1 1 X fE ( ) = p (2` + 1) ei `(p) sin `(p) P`(cos )
1 1 X (2` + 1) t (p) p (cos ) : = p ` ` `=0
`=0

The di erential cross section is then given by d = j f ( ) j2 = 1 X (2` + 1)(2`0 + 1) E d p2 `` ei( ` ? ` ) sin ` sin ` P`(cos ) P` (cos )
0 0 0 0

(11.72)

and the total cross section by


tot

(11.73)

1 X = 22 p `=0 1 4 X (2` + 1) sin2 (p) = p2 ` `=0


2 2`+1)
0

d'

d dcos d ?1 2 (2` + 1)2 sin2 (p) ` 2` + 1


Z

where us used

d P` P`0 =

p;` tot

(11.74) ` ` . If we de ne a "partial wave" cross section as := 4 2 (2` + 1) sin2 ` p 1 X = p;`


`=0

(11.75)

263

then we see that for each `;

max p;`

4 p2

(2` + 1) is the upper bound for

p;`.

The expressions (11.73) and (11.74) for the di erential and total cross sections are only practically useful if only a few partial waves contribute, i.e., are di erent from zero. This is the case for short-ranged potentials and small energies, as we will see later. If too many `'s have to be taken into account, the use of a partial wave expansion becomes questionable.

Insert: Di erent view on boundary conditions:


Consider the free solution of the radial Schrodinger equation

ul;p(r) = Al;p cos l rjl (pr) ? sin l rnl(pr)] ei l sin l Gl (pr) + cos l Fl (pr)]
with

(11.76)

Fl (pr) = prjl (pr) Gl (pr) = ?prnl (pr):


Then and

(11.77) (11.78)

?F 0 (pr) ? F tan l = ? ?Gl0 (pr) ? Gl (pr) (R) l (pr) (R) l

(11.79) (R) = 1 dul; p ul;p dr r=R Here one can establish a connection to bound states. A bound state is characterized by a square integrable solution,
Z

dru2(r) = 1 l 0 rlim ul (r) ! 0: !1


i(pr?l 2 ) ;

(11.80) (11.81) (11.82)

For scattering states one has the asymptotic behavior

ul (r) e e
i(pr?l 2 )

with Ep = p2 =2 . With p i this becomes Ep = ? 2 =2 and = e r il ; where only p = i gives a normalizable wave function. Then the logarithmic derivative is given by =? : (11.83) 264

This gives an extra condition to the wave function which makes our arti cial bound state problem into an eigenvalue problem.

End insert
The most simplest case occurs if only the lowest partial wave has to be considered (` = 0), i.e., if we have s-wave scattering. Then d = sin2 0 ; (11.84) d p2 which is independent of the scattering angle . Thus, for s-wave scattering, the di erential cross section is isotropic. A di erent expression for the phase shift can be obtained in the following way. We consider the di erential equations for the free solution " # 00 + p2 ? `(` + 1) (r j (pr)) = 0 (r j` (pr)) (11.85) ` r2 and for the full solution " # 00 (r) + p2 ? `(` + 1) ? 2m V (r) u (r) = 0 : u`;p (11.86) `;p r h2 Multiplying (11.85) with u`;p(r) and (11.86) with r j` (pr) and subtracting both equations leads to d (r j (pr))0 u (r) ? (r j (pr)) u0 (r)] = ?j (pr) 2m V (r)r u (r) : (11.87) ` `;p ` ` `;p `;p dr h2 Integrating over r and taking into account the boundary condition us;p(0) = j`(0) = 0 as well as asymptotic forms ! ` r!1 1 r j`(pr) ?! p sin pr ? 2 ! ` + (p) r!1 1 u`;p(r) ?! p sin pr ? 2 ` ! !1 (r j`(pr))0 r?! cos pr ? `2 (11.88) leads to sin `(p) = ? 2m h2
Z

dr r j`(pr) V (r) u`;p(r)


265

(11.89)

where

cos pr ? ` 2

1 sin pr ? ` + (p) p 2 `

1 ? p sin pr ? `2 ! ` + (p) = 1 sin (p) cos pr ? 2 ` p `


!

has been used. The expression (11.89) is no simpli cation with respect to the form (11.62). Though no asymptotic conditions enter, the integral of (11.89) has to be solved for all values of r to give the exact solution for sin `(p). In addition, (11.89) requires the calculation of the scattering solution u`;p(r). However, if one uses the approximation (11.90) u`;p(r) pr j`(pr) ; one obtained the so-called Born approximation for the phase shift 2m 1 Z 1 dr V (r) pr j (pr)]2 : (11.91) sin `(p) ? 2 p ` 0 h The question is under which conditions can (11.91) provide a good approximation for the phase shift. This should be expected if the right-hand side of (11.94) is small compared to 1, or more precisely, if the function j` (pr) is small in the domain of V (r). For small incident energies, one can argue as follows: The function j`( ) increases close to the q origin ( = 0) from zero as `+1 . It has a turning point at = `(` + 1). Thus, we q can assume that pr j` (pr) stays small up to r = 1 `(` + 1) and that, therefore, the p approximation (11.91) is good if the range of the potential ful lls 1q (11.92) R p `(` + 1) : Formulated in a di erent way: If the range R and the incident particle energy (given p) are xed, the phase shifts ` are small for all ` values which ful ll (11.92) and then (11.91) is a good approximation. This result can also be understood semi-classically. The length

`(` + 1) h `(` + 1) b` = = = angular momentum p hp momentum is the impact parameter of the particles incident with a speci c angular momentum. From classical mechanics, we know that there is no scattering if the impact parameter becomes larger than the range. For large energies, we use that the function j`( ) is bounded for all values of the argument. Then the right-hand side of (11.91) will be small if p is so large that 2m 1 j V (r) j dr 1: h2 p 0
Z

266

(This estimate is, of course, only valid if the integral over j V (r) j exists.) Thus, for high energies the Born approximation (11.91) should be valid for all values of `. However, at high energies one should calculate relativistically. A nal remark: If we had started with a radial wave function normalized such that

u`;p(r) ?! j`(pr) + tan ` n`(pr) ;


then (11.89) becomes tan
`

(11.93) (11.94)

= ? 2m 1 h2 p

dr j`(pr) V (r) u`;p(r) :

From the Born approximation for the phase-shift (11.91) we can draw some conclusions about the relation of the sign of the phase-shift ` and the behavior of V (r), since all other terms under the integral are positive. If ` < 0, then V (r) > 0, i.e. repulsive. If ` > 0, then V (r) < 0, i.e. attractive.

Example: Hard Sphere Scattering


Imagine a hard sphere of radius R, i.e. a potential that is in nitely repulsive for r < R and zero outside the sphere. Since the wave function does not penetrate inside the sphere, the wave function must vanish for r < R. That is we obtain ( r<R (11.95) u0(kr) = 0; kr ? kR); r > R sin( A comparison with the free solution gives for the phase shift obtained for scattering from a hard sphere
0 (E ) = ?kR:

ul(kr) ei l sin(kr ? l 2 + l )

(11.96) (11.97)

Here we see that a negative phase shift usually arises from repulsive potentials. Note also, that even a very simple potentials produce energy dependent phase shifts. 267

11.6 The Low Energy Limit


Here the behavior of the phase shift at low energies will be considered. For potentials that can be assumed to vanish beyond r = R, the earlier expressions for tan `, (11.62) and (11.94) can be used to explore the low energy behavior of the phase shift. When the external kinetic energy E is small compared to the depth of the potential, the wave function inside the potential will not depend sensitively on E . The total kinetic energy at any radius is E + j V (r) j, which is for small E nearly equal to j V (r) j. Thus, we can to rst approximation consider the logarithmic derivative (11.63) to be independent of energy. If we introduce the low energy behavior of j`(pr) and n`(pr) as
`+1 ?! pr j`(pr) pr<<` 1 3 (pr) (2` + 1) 5 ?pr n`(pr) pr<<` 1 3 5(pr)(2` ? 1) ?! `

(11.98)

then we obtain from (11.62) after multiplying numerator and denominator with pR tan

pR 2`+1 1) ? p!0 ?! (` + + R R(R0)(R) 1 3 5 ((2` )? 1)]2 (2` + 1) ` 0

(11.99)

where 0(R) is the zero energy logarithmic derivative. Thus, as the energy approaches zero, the tangent of the phase shift also approaches zero as tan For ` = 0 (s-waves) (11.99) yields tan
p!0 ?! 1 ? R(R0)(R) p = ?pa0 :
0

a` p2`+1 :

(11.100)

(11.101)

The quantity a0 is usually called the zero energy scattering length, or simply the scattering length and is de ned by 268

R) a0 = R 0((R)? 1 : 0

(11.102)

If the wave function is small at r = R, the quantity R 0(R) will be large at a0 ' 1; R if, instead, u0`;p(R) is nearly zero, R 0(R) will be small and a ' ? 0 (1R) . The scattering length a0 has a simple geometric interpretation. In the low energy limit, the wave function
p!0 u0(r) r>R ei 0 sin(pr + 0 ) ?! p(r ? a0) : =

(11.103)

Thus a0 is the point nearest to the origin at which the external wave function, or its extrapolation toward the origin, vanishes. Considering the radial Schrodinger equation for p ! 0; ` = 0, i.e.,
"

d2 ? U (r) u (r) = 0 0 dr2


#

we can deduce 1. For a repulsive potential (U > 0), the curvature of u0(r) is always away from the r axis so that a0 > 0.

U(r)

Fig. 11.2

Illustration of the geometrical meaning of the scattering length a0 .

269

2. For an attractive potential 1. incapable of producing an s-wave bound state: a0 < 0 (Fig. 11.3.a) 2. capable of producing an s-wave virtual state or "zero energy resonance": a0 = 1 (Fig. 11.3.b) 3. capable of producing one s-wave bound state: a0 > 0 (Fig. 11.3.c)

u0(r) a

u0(r)

u0(r)

U(r)

U(r)

U(r)

Fig. 11.3

(a)

Illustration of the scattering length a0 for various attractive potentials U (r).

(b)

(c)

The scattering length also has an important physical signi cance. In the low energy limit, only the s-wave makes a non-zero contribution to the cross section (11.100) so that the angular distribution of the scattering is spherically symmetric and the total cross section is 4 2 p!0 2 (11.104) tot = 2 sin 0 ?! 4 a0 : p This is similar to the result one obtains for the low energy scattering of a hard sphere of radius a. Thus, the scattering length is the "e ective radius" of the target at zero energy. However, there is a factor of 4 between the quantum mechanical cross section at low 270

energies and the classical cross section for scattering from a hard sphere ( classical = a2). This can be explained by the fact that in quantum mechanics one considers probabilities, i.e. the square of wave functions. At low energies the wave length is considerably larger than the size of the target, thus the scattering can be viewed as scattering by on the entire surface of the sphere, not only the cross section area of the sphere.

11.6.1 E ective Range Expansion


Here we only consider the very low energy regime, and thus we only need to take l = 0 into account. For the scattering amplitude we have 1 f0 (k) = k ei 0 sin 0 1 1 = k sini 00 = k cos sin i0sin ? e 0? 0 1 tan 0 = 1 1 = k 1 ? i tan k cot 0 ? i 0 1 (11.105) = k cot ? ik :
0

Using the de nition of the scattering length, k cot 0 = 1=a0, we can rewrite the scattering amplitude as 1 0 f0 (k) = 1 ? ik = 1 +aa2 k2 0 a0 2k = a0 + ia02 : (11.106) 1 + a2 k 0 For the di erent pieces of the scattering amplitude we thus get
k!0 <ef0(k) ?! k!0 =mf0(k) ?!

a0 a2 k: 0

(11.107)

From Eq. (11.104) we saw that the elastic cross section in the low energy limit is given by = 4 a2. Thus in the low energy limit the optical theorem 0 (11.108) = 4k =mf0 (k) is ful lled.

Important consequence: A single, real number, a0 , completely parameterizes all low


energy scattering. This is an advantage for e.g. experiments that use low energy neutron 271

scattering to study solids. It is a disadvantage for deduction of speci c properties of the projectile-target interaction. Consider the s-wave wave function

ei 0 sin 0 G0(kr) + cos 0 F0 (kr)] k!0 ?! kr + tan 0 (r + a0 )k (11.109) This shows that outside of the range of the potential the wave function for in the low energy limit is proportional to a0 + r, and thus intercepts the axis at r = ?a0 . u0(r)
Let's go back to Eq. (11.105) and insert the e ective range expansion to the next order, i.e. 1 1 2 (11.110) k cot 0 = ? a + 2 kr0 : This leads to the scattering amplitude
0

1 f0 (k) = ? 1 ? ik ? 1 kr2 0 a0 a0 2 = ? 1 + ia k ? 1 a kr2


0 2 0 0

(11.111)

Again, this expression for the scattering amplitude ful lls the optical theorem: 4 =mf (k) = 4 ?a0 (?a0 k) = 4 a2 0 (11.112) 0 1 a kr2 + a2 k2 1 a kr2 + a2 k2 k k 1? 2 0 0 0 1? 2 0 0 0 and Z 2 d = d = 4 jf0(k)j2 = 1 ? 1 a 4kra20+ a2 k2 (11.113)
2 0 0 0

11.6.2 Relation to Bound States


Consider the scattering amplitude 1 tan 0 a0 : 1 (11.114) fl=0 ' k ei 0 sin 0 = k 1 ? i tan 1 ? ia0 k 0 f is an analytic function and has a pole at tan 0 = 1=i. Here this means that k = ?1=a0. Considering the energy we have 2 ?1 E = 2km = 2ma2 ?EB : (11.115) 0 272

Conversely, knowledge of the scattering length determines the bound state energy. E.g. from the knowledge of the deuteron binding energy one determines the triplet nucleonnucleon scattering length to be at0 p 1 = +4:32fm (11.117) EB 2m Experimentally a value at0 = 5:42fm is extracted. The di erence comes from the fact that the bound state is not truly at zero energy and some higher order terms are needed.
Note:

Thus, the knowledge of the scattering lengths seems equivalent to the knowledge of the scattering length. If there is a bound state (with a small binding energy), it determines the low energy scattering: jf0j2 2m(E1+ E ) : (11.116)
B

While the scattering length approximation does not hold if the energy gets far from zero, the relation of the poles of the scattering amplitude to bound states is general.

11.7 Levinson's Theorem


The scattering length a0 is not simply a measure of the potential size or strength. If U (r) becomes stronger, then more than one bound state can exist, and the cycle going from negative to positive scattering length can repeat. This is consistent with tan 0 = pa0, where tan is a multi-valued function. tan 0 varies periodically and discontinuously from ?1 to +1. If the potential is strong enough to have a bound state at k = 0, we must be on the second branch of the tan function, and the zero energy phase shift is rather than zero. The general result of this consideration is Levinson's Theorem (k = 0) ? (k 1) = nB (11.118) where nB is the number of bound states supported by the speci c potential U (r). Levinson's theorem relates the phase shift as zero and in nite energy to the number of bound states nB .

273

11.8 Breit-Wigner Resonances


Negative energy state (bound states) are stationary states and obey the stationary Schrodinger equation. States with positive energy, con ned in a positive potential well are con ned, but will eventually through the potential barrier. Those state are called quasi-bound states or resonances.

V(r)

Resonances

E>0
r

Bound States

E<0

Sketch of a potential supporting bound states and resonance states.

Fig. 11.6

Consider the expansion of the phase shift tan ` for p ! 0 as given in Eq. (11.99), (pR)2`+1 p!0 (11.119) tan ` ?! (` + 1) ? R 0(R) ` + R 0(R) 1 3 5 (2` ? 1)]2 (2` + 1) The denominator vanishes for R 0(R) = ?`. Thus tan ` ! 1, i.e. tan ` = 2 + n , This condition occurs for a speci c momentum pR at a speci c energy ER = pR=2 . Expanding (pR) around Er gives

Substitution of Eq. (11.120) into Eq. (11.119) leads to 1 1 2(pR)2`+1 tan ` R E ? ER 2 (2` ? q)!!]2 d(dE ) 1 = E ? E ?l 2
R

R (pR)R ?l + (E ? ER ) d(dE ) jE=Er

(11.120)

(11.121)

274

with

?l =

2(pR)2`+1 : R (2` ? q)!!]2 d(dE )

(11.122)

This leads to the Breit-Wigner resonance form of the amplitude 2 kf` = ei ` sin ` = E ? ?l =? i? =2 (11.123) R E l and the Breit-Wigner cross section in a speci c partial wave partial wave `. 4 (2` + 1) ?2=4 l (11.124) `= p2 (E ? ER )2 + ?2 =4 l where ?l is called the width of the resonance. From the derivation it is apparent that ?l 1= (pR). Thus, if ?l is small, (pR) (and thus `) varies rapidly near ER . If this is the case, the resonance will be sharp, and the total cross section tot will have a sharp peak. However, unless the resonance is very narrow, then the 1=p2 factor in tot will distort the shape of the cross section and may shift the peak from ER to a lower energy. Physically, a sharp peak in the energy dependence of the cross section indicates a dynamical origin, such as a strong attraction at that energy. If the phase shift passes rapidly through =2 (modulo ), this probably means a resonance, i.e. beam and target particle binding temporarily and then breaking up again. Resonances are poles in f`(E ) at E = ER ? i?l =2. This means q q q q p = 2 E = 2 Er ? i ?l = 2 E 1 ? i?l =2Er pr ? i ?l =2 (for small ?l ) (11.125) pr If ?l is small, the pole is right below the real positive p-axis. When taking E = p2=2 , bound state poles map to the rst Riemann sheet, resonance poles move into the lower half of the second (unphysical) Riemann energy sheet. The smaller ?l , the sharper the pole is to the real axis.
tot

peaks, the longer lived the resonance is, and the closer

11.9 The Classical Limit


Let us now examine the scattering amplitude (11.69, 11.72) in the classical limit, that is, at a su ciently high energy that the particle may be localized. At these energies, 275

many partial waves will contribute so that the partial wave expansion for f`( ) may be approximated by an integral. In addition, most of the scattering will result from high partial waves so that we can assume that the phase shift for "normal" potentials will behave as shown in Fig. 11.4.

l
E.

Fig. 11.4

The phase shift


`

as function of ` for high energies

Fig. 11.4 The phase shift

as function of ` for high energies E .

The region where ` is relatively constant will make little contribution, as we will see below. As a result, we have from (11.72) 2i ` ? X fE ( ) = (2` + 1) e 2ip 1 P`(cos ) `=0 Z 1 2i ` ' 0 d` ` e ip? 1 P`(cos ) : (11.126) For large ` and small angles, the Legendre polynomial may be replaced by a Bessel function using the expansion 1 1 : (11.127) P`(cos ) ' j0 2 ` + 2 sin 1 + 1 sin2 2 + 2 4 This greatly simpli es the integration since it avoids integrating over the order of the Leg` endre polynomial. It is also convenient to use b = p and to recognize that the momentum transferred to the scattered particle, which we shall denote by q, is q = j p ? ~ 0 j = (p2 + p02 ? 2pp0 cos )1=2 ~ p 1 = p(2(1 ? cos )) 2 = 2p sin 2 (11.128) 276

for elastic scattering, i.e., j p j = j ~ 0 j. Then (11.126) may be rewritten to obtain the ~ p semi classical approximation p Z 1 db b (e2i (b) ? 1) j (qb) : fE ( ) = i (11.129) 0 0 This formula is reminiscent of very similar formulas in the classical theory of di raction. Its physical implications can be seen most strikingly by assuming that (b) has a limiting form as suggested by Fig. 11.4, namely that it is constant for b < R and zero for b > R. In this simple case, the integral in (11.129) can be performed analytically giving

j0 (qR) (11.130) qR where j1(qR) is the rst-order Bessel function. This result, familiar from the theory of Fraunhofer di raction, gives a cross section as shown in Fig. 11.5. fE ( )

|f( )|

3.8 qR

Fig. 11.5

Di erential cross section is the classical limit.

As expected, the cross section is sharply peaked in the forward direction and is concen1 trated within the region having < ( pR ). At high energies, we can extract the physical content of (11.129) by using the method of "steepest descent." For large q the integrand will oscillate rapidly as b varies. Asymptotically, the Bessel function has the form

?! j0 (qb) qb!1 (2 qb)1=2 exp i qb ? 4 + exp ?i qb ? 4 (11.131) so that the dominant contribution to the integral will come from those values of b for which 2 (b) qb is nearly constant. The term in (11.129) that is independent of the phase
277

shift does not contribute to the scatteringP from the forward direction as can be seen away by returning to (11.126) and noting that 1 ` + 1 P`(cos ) = (1 ? cos ). Thus, `=0 2 the important region of b is determined, for xed , by the relation d (b) 1 q = p sin : (11.132) db 2 2 It may be seen from this, as pointed out earlier, that values of b, for which (b) is a constant, will not contribute to scattering out of the forward direction (q > 0).

278

S-ar putea să vă placă și