0% found this document useful (0 votes)
8 views

12 Scattering

Uploaded by

RAJESH KUMAR
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
8 views

12 Scattering

Uploaded by

RAJESH KUMAR
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 85

12 - Scattering Theory

◮ Aim of Section:
◮ Outline quantum theory of scattering.
Introduction
◮ Historically, data regarding quantum phenomena was obtained
from two main sources.
◮ Firstly, study of spectroscopic lines, and, secondly, analysis of
data from scattering experiments.
◮ Let us now examine quantum theory of scattering.
◮ We shall treat scattering as an essentially two-particle effect.
◮ As is well known, when viewed in center of mass frame, two
particles of masses m1 and m2 , and position vector x1 and x2 ,
respectively, interacting via potential V (x1 − x2 ), can be
treated as a single body of reduced mass
µ12 = m1 m2 /(m1 + m2 ), and position vector x = x1 − x2 ,
moving in fixed potential V (x).
◮ For this reason, we can, without loss of generality, focus our
study on quantum theory of particles scattered by fixed
potentials.
Fundamental Equations - I

◮ Consider time-independent scattering theory, for which


Hamiltonian of system is written

H = H0 + H1 ,

where
p2
H0 =
2m
is Hamiltonian of a free particle of mass m, and H1 represents
non-time-varying source of scattering.
◮ Let |φi be an energy eigenket of H0 ,

H0 |φi = E |φi, (1)

whose wavefunction is φ(x). This wavefunction is assumed to


be a plane wave.
Fundamental Equations - II

◮ Schrödinger’s equation for scattering problem is

(H0 + H1 ) |ψi = E |ψi, (2)

where |ψi is an energy eigenstate of total Hamiltonian whose


wavefunction is ψ(x).
◮ In general, both H0 and H0 + H1 have continuous energy
spectra: that is, their energy eigenstates are unbound.
◮ We require a solution of (2) that satisfies boundary condition
|ψi → |φi as H1 → 0.
◮ Here, |φi is a solution of free-particle Schrödinger equation,
(1), that corresponds to same energy eigenvalue as |ψi.
Fundamental Equations - III
◮ Adopting Schrödinger representation, we can write scattering
equation, (2), in form

2m
(∇2 + k 2 ) ψ(x) = hx|H1 |ψi, (3)
~2
where
~2 k 2
E = .
2m
◮ Here, |x′ i is a state whose wavefunction is δ3 (x − x′ ). It
follows that
x |x′ i = x′ |x′ i.
In other words, |x′ i is an eigenstate of position operator, x,
corresponding to eigenvalue x′ .
◮ Follows that
hx|ψi = ψ(x).
Fundamental Equations - IV
◮ (3) is known as Helmholtz equation, and can be inverted
using standard Green’s function techniques. Thus,
2m
Z
ψ(x) = φ(x) + 2 G (x, x′ ) hx′ |H1 |ψi d 3 x′ , (4)
~
where
(∇2 + k 2 ) G (x, x′ ) = δ3 (x − x′ ).
Here, δ3 (x) is a three-dimensional Dirac delta function.
◮ Note that solution (4) satisfies previously mentioned
constraint |ψi → |φi as H1 → 0.
◮ As is well known, Green’s function for Helmholtz equation is
given by
exp(±i k |x − x′ | )
G (x, x′ ) = − .
4π |x − x′ |
◮ Thus, (4) becomes
2m exp(±i k |x − x′ | ) ′
Z
±
ψ (x) = φ(x) − 2 hx |H1 |ψ ± i d 3 x′ .
~ 4π |x − x′ |
(5)
Fundamental Equations - V
◮ Let us suppose that scattering Hamiltonian, H1 , is a function
only of position operators. This implies that

hx′ |H1 |xi = V (x) δ3 (x − x′ ). (6)


◮ We can write
Z
′ ±
hx |H1 |ψ i = hx′ |H1 |x′′ ihx′′ |ψ ± i d 3 x′′
Z
= V (x′ ) δ3 (x′ − x′′ ) ψ(x′′ ) d 3 x′′ = V (x′ ) ψ ± (x′ ),

where
R ′′ use has been made of standard completeness relation
|x ihx | d 3 x′′ = 1.
′′

◮ Thus, integral equation (5) simplifies to give


2m exp(±i k |x − x′ |)
Z
±
ψ (x) = φ(x) − 2 V (x′ ) ψ ± (x′ ) d 3 x′ .
~ 4π |x − x′ |
(7)
Fundamental Equations - VI

◮ Suppose that initial state, |φi, possesses a plane-wave


wavefunction with wavevector k (i.e., it corresponds to a
stream of particles of definite momentum p = ~ k).
◮ Ket corresponding to this state is denoted |ki.
◮ Thus,
exp( i k · x)
φ(x) ≡ hx|ki = . (8)
(2π)3/2
◮ Preceding wavefunction is conveniently normalized such that

exp[−i x · (k − k′ )] 3
Z Z
′ ′ 3
hk|k i = hk|xihx|k i d x = d x
(2π)3
= δ3 (k − k′ ).
Fundamental Equations - VII
◮ Suppose that scattering potential, V (x), is non-zero only in
some relatively localized region centered on origin (x = 0).
◮ Let us calculate total wavefunction, ψ(x), far from scattering
region. In other words, let us adopt ordering r ≫ r ′ , where
r = |x| and r ′ = |x′ |.
◮ It is easily demonstrated that
|x − x′ | ≃ r − er · x′
to first order in r ′ /r , where er = x/r is a unit vector that is
directed from scattering region to observation point.
◮ Let us define
k′ = k er .
Clearly, k′ is wavevector for particles that possess same energy
as incoming particles (i.e., k ′ = k), but propagate from
scattering region to observation point.
◮ Note that
exp(±i k |x − x′ | ) ≃ exp(±i k r ) exp(∓i k′ · x′ ).
Fundamental Equations - VIII

◮ In large-r limit, (7) and (8) reduce to

exp( i k · x)
ψ ± (x) ≃
(2π)3/2
m exp(±i k r )
Z
− exp(∓i k′ · x′ ) V (x′ ) ψ ± (x′ ) d 3 x′ .
2π ~2 r
◮ First term on right-hand side of previous equation is incident
wave.
◮ Second term represents a spherical wave centered on
scattering region.
◮ Plus sign (on ψ ± ) corresponds to a wave propagating away
from scattering region, whereas minus sign corresponds to a
wave propagating toward scattering region.
◮ It is obvious that former sign represents physical solution.
Fundamental Equations - IX

◮ Thus, wavefunction far from scattering region can be written


 
1 exp( i k r ) ′
ψ(x) = exp( i k · x) + f (k , k) ,
(2π)3/2 r

where
(2π)2 m exp(−i k′ · x′ )
Z

f (k , k) = − V (x′ ) ψ(x′ ) d 3 x′
~2 (2π)3/2
(2π)2 m ′
=− hk |H1 |ψi.
~2
Fundamental Equations - X
◮ Let us define differential scattering cross-section, dσ/dΩ, as
number of particles per unit time scattered into an element of
solid angle dΩ, divided by incident particle flux.
◮ Probability current (which is proportional to particle flux)
associated with a wavefunction ψ is
~
j= Im(ψ ∗ ∇ψ).
m
◮ Thus, particle flux associated with incident wavefunction,

exp( i k · x)
,
(2π)3/2

is proportional to
~k
jincident = . (9)
(2π)3 m
Fundamental Equations - XI
◮ Likewise, particle flux associated with scattered wavefunction,

exp( i k r ) f (k′ , k)
,
(2π)3/2 r

is proportional to

~ k′ |f (k′ , k)| 2
jscattered = .
(2π)3 m r2
◮ Now, by definition,

dσ r 2 dΩ |jscattered |
dΩ = ,
dΩ |jincident |

giving

= |f (k′ , k)|2 . (10)
dΩ
Fundamental Equations - XII

◮ Thus, |f (k′ , k)|2 is differential cross-section for particles with


incident momentum ~ k to be scattered into states whose
momentum vectors are directed in a range of solid angles dΩ
about ~ k′ .
◮ Note that scattered particles possess same energy as incoming
particles (i.e., k ′ = k). This is always case for scattering
Hamiltonians of form specified in (6).
Born Approximation - I

◮ (10) is not particularly useful, as it stands, because quantity


f (k′ , k) depends on unknown ket |ψi.
◮ Recall that ψ(x) ≡ hx|ψi is solution of integral equation

exp( i k |x − x′ |)
Z
m
ψ(x) = φ(x) − V (x′ ) ψ(x′ ) d 3 x′ ,
2π ~2 |x − x′ |
(11)
where φ(x) is wavefunction of incident state.
◮ According to previous equation, total wavefunction is a
superposition of incident wavefunction and a great many
spherical waves emitted from scattering region.
◮ Strength of spherical wave emitted at a given point in
scattering region is proportional to local value of scattering
potential, V (x), as well as local value of wavefunction, ψ(x).
Born Approximation - II

◮ Suppose, however, that scattering is not particularly intense.


In this case, it is reasonable to suppose that total
wavefunction, ψ(x), does not differ substantially from incident
wavefunction, φ(x).
◮ Thus, we can obtain an expression for ψ(x) by making
substitution
exp( i k · x)
ψ(x) → φ(x) =
(2π)3/2
on right-hand side of (11).
◮ This simplification is known as Born approximation.
Born Approximation - III

◮ Born approximation yields


Z
m
′ ′ ′
V (x′ ) d 3 x′ .
 
f (k , k) ≃ − exp i (k − k ) · x
2π ~2
◮ Thus, f (k′ , k) is proportional to Fourier transform of
scattering potential, V (x), with respect to relative
wavevector, q ≡ k − k′ .
Born Approximation - IV
◮ For a spherically symmetric scattering potential,
Z ∞Z πZ 2π
m ′ ′
f (k′ , k) ≃ − 2
e i q r cos θ V (r ′ )′ r ′ 2 sin θ ′ dr ′ dθ ′ dϕ,
2π ~ 0 0 0
giving

2m
Z

f (k , k) ≃ − 2 r V (r ) sin(q r ) dr . (12)
~ q 0
◮ Hence, it is clear that, for special case of a spherically
symmetric potential, f (k′ , k) depends only on magnitude of
relative wavevector, q = k − k′ , and is independent of its
direction.
◮ Now, it is easily demonstrated that

q ≡ |k − k′ | = 2 k sin(θ/2), (13)

where θ is angle subtended between vectors k and k′ .


◮ In other words, θ is angle of scattering.
Born Approximation - V

◮ Recall that vectors k and k′ have same length, as a


consequence of energy conservation.
◮ It follows that, according to Born approximation,
f (k′ , k) = f (θ) for a spherically symmetric scattering
potential, V (r ). Moreover, f (θ) is real.
◮ Finally, differential scattering cross-section, dσ/dΩ = |f (θ)|2 ,
is invariant under transformation V → −V .
◮ In other words, pattern of scattering is identical for attractive
and repulsive scattering potentials of same strength.
Born Approximation - VI
◮ Consider scattering by a Yukawa potential,
V0 exp(−µ r )
V (r ) = , (14)
µr
where V0 is a constant, and 1/µ measures “range” of
potential.
◮ It follows from (12) that
2 m V0 1
f (θ) = − ,
~ µ q + µ2
2 2

because
Z ∞
q
exp(−µ r ) sin(q r ) dr = .
0 q 2 + µ2
◮ Thus, Born approximation yields a differential cross-section for
scattering by a Yukawa potential of form
2 m V0 2
 
dσ 1
≃ 2 .
dΩ ~2 µ 2

4 k sin (θ/2) + µ2
2
Born Approximation - VII
◮ Yukawa potential reduces to familiar Coulomb potential in
limit µ → 0, provided that V0 /µ → Z Z ′ e 2 /4π ǫ0 . Here, Z e
and Z ′ e are electric charges of two interacting particles.
◮ In Coulomb limit, previous Born differential cross-section
transforms into
2
2 m Z Z ′ e2

dσ 1
≃ .
dΩ 4π ǫ0 ~2 16 k 4 sin4 (θ/2)
◮ Recalling that ~ k is equivalent to |p|, where p is momentum
of incident particles, preceding equation can be rewritten
2
Z Z′ e 2

dσ 1
≃ 4 , (15)
dΩ 16π ǫ0 E sin (θ/2)

where E = p 2 /(2 m) is kinetic energy of incident particles.


◮ (15) is identical to well-known Rutherford scattering
cross-section formula of classical physics.
Born Approximation - VIII
◮ Born approximation is valid provided ψ(x) is not significantly
different from φ(x) in scattering region.
◮ It follows, from (11), that condition that must be satisfied in
order that ψ(x) ≃ φ(x) in vicinity of x = 0 is
exp( i k r ′ )
Z
m
V (x′ ) d 3 x′ ≪ 1. (16)
2π ~2 r′
◮ Consider special case of Yukawa potential, (14). At low
energies (i.e., k ≪ µ), we can replace exp( i k r ′ ) by unity,
giving
2 m |V0 |
≪1
~2 µ2
as condition for validity of Born approximation.
◮ Now, criterion for Yukawa potential to develop a bound state
turns out to be
2 m |V0 |
≥ 2.7, (17)
~2 µ2
provided V0 is negative.
Born Approximation - IX

◮ Thus, if potential is strong enough to form a bound state then


Born approximation is likely to break down.
◮ In high-k limit (i.e., k ≫ µ), (16) yields

2 m |V0 |
≪ 1.
~2 µ k
◮ This inequality becomes progressively easier to satisfy as k
increases, implying that Born approximation becomes more
accurate at high incident particle energies
Born Expansion - I
◮ As we have seen, quantum scattering theory requires solution
of integral equation,
exp( i k |x − x′ |)
Z
m
ψ(x) = φ(x) − 2
V (x′ ) ψ(x′ ) d 3 x′ ,
2π ~ |x − x′ |
where φ(x) = exp( i k · x)/(2π)3/2 is incident wavefunction,
and V (x) scattering potential.
◮ An obvious approach, in weak-scattering limit, is to solve
preceding equation via a series of successive approximations.
That is,
exp( i k |x − x′ |)
Z
(1) m
ψ (x) = φ(x) − V (x′ ) φ(x′ ) d 3 x′ ,
2π ~2 |x − x′ |
exp( i k |x − x′ |)
Z
m
ψ (2) (x) = φ(x) − V (x′ ) ψ (1) (x′ ) d 3 x′ ,
2π ~2 |x − x′ |
exp( i k |x − x′ |)
Z
(3) m
ψ (x) = φ(x) − V (x′ ) ψ (2) (x′ ) d 3 x′ ,
2π ~2 |x − x′ |
and so on.
Born Expansion - II
◮ Assuming that V (x) is only non-negligible relatively close to
origin, and taking limit |x| → ∞, we find that
 
1 exp( i k r ) ′
ψ(x) = exp( i k · x) + f (k , k) ,
(2π)3/2 r

where

f (k′ , k) = f (1) (k′ , k) + f (2) (k′ , k) + f (3) (k′ , k) + · · ·.


◮ First two terms in previous series, which is generally known as
Born expansion, are
Z
(1) ′ m ′ ′
f (k , k) = − e i (k−k )·x V (x′ ) d 3 x′ , (18)
2π ~2
 m 2 Z Z i k |x′ −x′′ |
i (k·x′′ −k′ ·x′ ) e
f (2) (k′ , k) = e V (x′ ) V (x′′ ) d 3 x′ d 3 x′′ .
2π ~2 |x′ − x′′ |
Born Expansion - III

◮ Of course, we recognize (18) as Born approximation discussed


previously.
◮ In other words, Born approximation essentially involves
truncating Born expansion after its first term.
◮ Incidentally, it can be proved that Born expansion converges
for all k (for Ra spherically symmetric Rscattering potential)
∞ ∞
provided; a) 0 r |V (r )| dr < ∞; b) 0 r 2 |V (r )| dr < ∞;
and; c) −|V (r )| is too weak to form a bound state.
◮ Furthermore, criterion for convergence becomes less stringent
at high k.
Partial Waves - I
◮ We can assume, without loss of generality, that incident
wavefunction is characterized by a wavevector, k, that is
aligned parallel to z-axis.
◮ Scattered wavefunction is characterized by a wavevector, k′ ,
that has same magnitude as k, but, in general, points in a
different direction.
◮ Direction of k′ is specified by polar angle θ (i.e., angle
subtended between two wavevectors), and an azimuthal angle
ϕ measured about z-axis.
◮ (12) and (13) strongly suggest that for a spherically
symmetric scattering potential [i.e., V (x) = V (r )], scattering
amplitude is a function of θ only: that is,

f (θ, ϕ) = f (θ).

◮ Let us assume that this is case.


Partial Waves - II

◮ It follows that neither incident wavefunction,


exp( i k z) exp( i k r cos θ)
φ(x) = 3/2
= , (19)
(2π) (2π)3/2

nor total wavefunction far from scattering region,


 
1 exp( i k r ) f (θ)
ψ(x) = exp( i k r cos θ) + , (20)
(2π)3/2 r

depend on azimuthal angle, ϕ.


Partial Waves - III
◮ Outside range of scattering potential, φ(x) and ψ(x) both
satisfy free-space Schrödinger equation,
(∇2 + k 2 ) ψ = 0. (21)
◮ Consider most general solution to this equation that is
independent of azimuthal angle, ϕ.
◮ Separation of variables (in spherical coordinates) yields
X
ψ(r , θ) = Rl (r ) Pl (cos θ) (22)
l=0,∞

◮ Legendre polynomials, Pl (cos θ), are related to associated


Legendre functions, Plm (cos θ), as well as spherical harmonics,
Ylm (θ, ϕ), via Pl (cos θ) = Pl0 (cos θ), and
r

Pl (cos θ) = Y 0 (θ, ϕ),
2l + 1 l
respectively.
Partial Waves - IV

◮ (21) and (22) can be combined to give

d 2 Rl dRl
r2 + 2r + [k 2 r 2 − l (l + 1)] Rl = 0.
dr 2 dr
◮ Two independent solutions to this equation are spherical
Bessel function, jl (k r ), and Neumann function, ηl (k r ), where

1 d l sin y
 
l
jl (y ) = y − , (23)
y dy y
1 d l cos y
 
ηl (y ) = −y l − . (24)
y dy y
◮ Note that spherical Bessel functions are well behaved in limit
y → 0, whereas Neumann functions become singular.
Partial Waves - V

◮ Asymptotic behavior of these functions in limit y → ∞ is

sin(y − l π/2)
jl (y ) → , (25)
y
cos(y − l π/2)
ηl (y ) → − . (26)
y
Partial Waves - VI
◮ We can write
X
exp( i k r cos θ) = al jl (k r ) Pl (cos θ),
l=0,∞

where the al are constants.


◮ Of course, there are no Neumann functions in this expansion
because they are not well behaved as r → 0 (whereas function
on left-hand side is clearly finite at r = 0).
◮ As is well known, Legendre polynomials are orthogonal
functions, Z 1
δn m
Pn (µ) Pm (µ) dµ = , (27)
−1 n + 1/2
so we can invert preceding expansion to give
Z 1
al jl (k r ) = (l + 1/2) exp( i k r µ) Pl (µ) dµ.
−1
Partial Waves - VII
◮ Now,
1
(−i)l
Z
jl (y ) = exp( i y µ) Pl (µ) dµ,
2 −1
for l = 0, ∞.
◮ Thus, a comparison of previous equations yields

al = il (2 l + 1),

giving
X
exp( i k r cos θ) = il (2 l + 1) jl (k r ) Pl (cos θ). (28)
l=0,∞

◮ Preceding expression specifies how a plane wave can be


decomposed into a series of spherical waves.
◮ Latter waves are usually referred to as partial waves.
Partial Waves - VIII
◮ Most general expression for total wavefunction outside
scattering region is
1 X
ψ(x) = [Al jl (k r ) + Bl ηl (k r )] Pl (cos θ), (29)
(2π)3/2 l=0,∞

where the Al and Bl are constants.


◮ Note that Neumann functions are allowed to appear in this
expansion, because its region of validity does not include
origin.
◮ In large-r limit, total wavefunction reduces to
1 X  sin(k r − l π/2) cos(k r − l π/2)

ψ(x) ≃ Al − Bl Pl (cos θ),
(2π)3/2 kr kr
l=0,∞

where use has been made of (25) and (26).


Partial Waves - IX

◮ Previous expression can also be written

1 X sin(k r − l π/2 + δl )
ψ(x) ≃ 3/2
Cl Pl (cos θ), (30)
(2π) l=0,∞ kr

where

Al = Cl cos δl , (31)
Bl = −Cl sin δl . (32)
Partial Waves - X
◮ (30) yields
 i (k r −l π/2+δl )
− e−i (k r −l π/2+δl )

1 X e
ψ(x) ≃ Cl Pl (cos θ),
(2π)3/2 l=0,∞ 2ik r
(33)
which contains both incoming and outgoing spherical waves.
◮ What is source of incoming waves?
◮ Obviously, they must form part of large-r asymptotic
expansion of incident wavefunction.
◮ In fact, it is easily seen from (19), (25), and (28) that
 i (k r −l π/2)
− e−i (k r −l π/2)

1 X
l e
φ(x) ≃ i (2 l + 1) Pl (cos θ),
(2π)3/2 l=0,∞ 2ik r
(34)
in large-r limit.
◮ Now, (19) and (20) give
exp( i k r )
(2π)3/2 [ψ(x) − φ(x)] = f (θ). (35)
r
Partial Waves - XI
◮ Note that right-hand side consists only of an outgoing
spherical wave.
◮ This implies that coefficients of incoming spherical waves in
large-r expansions of ψ(x) and φ(x) must be equal.
◮ It follows from (33) and (34) that

Cl = (2 l + 1) exp[ i (δl + l π/2)], (36)

which leads to
1 X sin(k r − l π/2)
φ(x) = 3/2
il (2 l + 1) Pl (cos θ),
(2π) l=0,∞ kr
(37)
1 X sin(k r − l π/2 + δl )
ψ(x) = il (2 l + 1) e i δl Pl (cos θ).
(2π)3/2 l=0,∞ kr
(38)
Partial Waves - XII

◮ Thus, it is apparent that effect of scattering is to introduce a


phase-shift, δl , into l th partial wave.
◮ Finally, (35) yields
X exp( i δl )
f (θ) = (2 l + 1) sin δl Pl (cos θ). (39)
k
l=0,∞

◮ Clearly, determining scattering amplitude, f (θ), via a


decomposition into partial waves (i.e., spherical waves), is
equivalent to determining phase-shifts, δl .
Partial Waves - XIII
◮ It is helpful to write
X 
φ+ −

φ(r) = l (r , θ) + φl (r , θ) , (40)
l=0,∞
X 
Sl φ+ −

ψ(r) = l (r , θ) + φl (r , θ) , (41)
l=0,∞

where
(2 l + 1) e−i (k r −l π)
φ−
l (r , θ) = − Pl (cos θ) (42)
(2π)3/2 2ik r
is an ingoing spherical wave, whereas
(2 l + 1) e i k r
φ+
l (r , θ) = Pl (cos θ) (43)
(2π)3/2 2 i k r
is an outgoing spherical wave.
◮ Moreover,
Sl = e i 2 δl . (44)
[See (37) and (38).]
Partial Waves - XIII

+
◮ Note that φ−l (r , θ) and φl (r , θ) are both eigenstates of
magnitude of total orbitalpangular momentum about origin
belonging to eigenvalues l (l + 1) ~.
◮ Thus, in preforming a partial wave expansion, we have
effectively separated incoming and outgoing particles into
streams possessing definite angular momenta about origin.
◮ Moreover, effect of scattering is to introduce an
angular-momentum-dependent phase-shift into outgoing
particle streams.
Partial Waves - XIV
◮ Net outward particle flux through a sphere of radius r ,
centered on origin, is proportional to
I
r 2 jr dΩ,

where j = (~/m) Im(ψ ∗ ∇ψ) is probability current.


◮ It follows that
I
~ X
r 2 jr dΩ = (2 l + 1) (|Sl | 2 − 1), (45)
8π 2 k m
l=0,∞

where use has been made of (27).


◮ Of course, net particle flux must be zero, otherwise number of
particles would not be conserved.
◮ Particle conservation is ensured by fact that |Sl | = 1 for all l .
[See (44).]
Optical Theorem - I
◮ Differential scattering cross-section, dσ/dΩ, is simply
modulus squared of scattering amplitude, f (θ). [See (10).]
◮ Total scattering cross-section is defined as
I I

σtotal = dΩ = |f (θ)| 2 dΩ
dΩ
Z 1 X X
1
I
= 2 dϕ (2 l + 1) (2 l ′ + 1) exp[ i (δl − δl ′ )]
k −1 l=0,∞ l =0,∞

× sin δl sin δl ′ Pl (µ) Pl ′ (µ) dµ, (46)

where µ = cos θ.
◮ It follows that
4π X
σtotal = (2 l + 1) sin2 δl , (47)
k2
l=0,∞

where use has been made of (27).


Optical Theorem - II

◮ A comparison of preceding expression with (39) reveals that

4π 4π
σtotal = Im [f (0)] = Im [f (k, k)], (48)
k k
because Pl (1) = 1.
◮ This result is known as optical theorem, and is a consequence
of fact that very existence of scattering requires scattering in
forward (θ = 0) direction, in order to interfere with incident
wave, and thereby reduce probability current in that direction.
Optical Theorem - III

◮ It is conventional to write
X
σtotal = σl ,
l=0,∞

where

σl = (2 l + 1) sin2 δl (49)
k2
is termed l th partial scattering cross-section: that is,
contribution to total scattering cross-section from l th partial
wave.
◮ Note that (at fixed k) maximum value for l th partial
scattering cross-section occurs when associated phase-shift,
δl , takes value π/2.
Determination of Phase-Shifts - I
◮ Let us now consider how partial wave phase-shifts, δl , can be
evaluated.
◮ Consider a spherically symmetric potential, V (r ), that
vanishes for r > a, where a is termed range of potential.
◮ In region r > a, wavefunction ψ(x) satisfies free-space
Schrödinger equation, (21).
◮ According to (29), (31), (32), and (36), most general solution
of this equation that is consistent with no incoming spherical
waves, other than those contained in incident wave, is
1 X
ψ(x) = il (2 l + 1) Al (r ) Pl (cos θ), (50)
(2π)3/2 l=0,∞

where

Al (r ) = exp( i δl ) [cos δl jl (k r ) − sin δl ηl (k r )]. (51)


Determination of Phase-Shifts - II
◮ Note that Neumann functions are allowed to appear in
previous expression, because its region of validity does not
include torigin (where V 6= 0).
◮ Logarithmic derivative of l th radial wavefunction, Al (r ), just
outside range of potential is given by

cos δl jl′ (k a) − sin δl ηl′ (k a)


 
βl+ = k a ,
cos δl jl (k a) − sin δl ηl (k a)

where jl′ (x) denotes djl (x)/dx, et cetera.


◮ Previous equation can be inverted to give

k a jl′ (k a) − βl+ jl (k a)
tan δl = . (52)
k a ηl′ (k a) − βl+ ηl (k a)

◮ Thus, problem of determining phase-shift, δl , is equivalent to


that of determining βl+ .
Determination of Phase-Shifts - III
◮ Most general solution to Schrödinger’s equation inside range
of potential (r < a) that does not depend on azimuthal angle,
ϕ, is
1 X
ψ(x) = 3/2
il (2 l + 1) Al (r ) Pl (cos θ), (53)
(2π) l=0,∞

where
ul (r )
Al (r ) = , (54)
r
and
d 2 ul
 
2 2m l (l + 1)
+ k − 2 V− ul = 0. (55)
dr 2 ~ r2
◮ Boundary condition
ul (0) = 0 (56)
ensures that radial wavefunction is well behaved at origin.
Determination of Phase-Shifts - IV

◮ We can launch a well-behaved solution of previous equation


from r = 0, integrate out to r = a, and form logarithmic
derivative [of Al (r )]

1 d(ul /r )
βl− = .
(ul /r ) dr r =a

◮ Because ψ(x) and its first derivatives are necessarily


continuous for physically acceptable wavefunctions, it follows
that
βl+ = βl− .
◮ Phase-shift, δl , is then obtained from (52).
Hard-Sphere Scattering - I
◮ Let us try out scheme outlined previously using a particularly
simple example.
◮ Consider scattering by a hard sphere, for which potential is
infinite for r < a, and zero for r > a.
◮ It follows that ψ(x) is zero in region r < a, which implies that
ul = 0 for all l .
◮ Thus,
βl− = βl+ = ∞
for all l .
◮ (52) yields
jl (k a)
tan δl = . (57)
ηl (k a)
◮ In fact, this result is most easily obtained from obvious
requirement that Al (a) = 0. [See (51).]
Hard-Sphere Scattering - II
◮ Consider l = 0 partial wave, which is usually referred to as
S-wave.
◮ (57) gives
sin(k a)/k a
tan δ0 = = − tan(k a),
− cos(k a)/ka
where use has been made of (23) and (24).
◮ It follows that
δ0 = −k a. (58)
◮ S-wave radial wavefunction is
 
cos(k a) sin(k r ) − sin(k a) cos(k r )
A0 (r ) = exp(−i k a)
kr
sin[k (r − a)]
= exp(−i k a) . (59)
kr
[See (51).]
Hard-Sphere Scattering - III

◮ Corresponding radial wavefunction for incident wave takes


form
sin(k r )
Ã0 (r ) = .
kr
[See (37), (38), (50), and (58).]
◮ It is clear that actual l = 0 radial wavefunction is similar to
incident l = 0 wavefunction, except that it is phase-shifted by
k a.
Hard-Sphere Scattering - IV
◮ Let us consider low- and high-energy asymptotic limits of
tan δl .
◮ Low energy corresponds to k a ≪ 1.
◮ In this limit, spherical Bessel functions and Neumann
functions reduce to
(k r ) l
jl (k r ) ≃ ,
(2 l + 1)!!
(2 l − 1)!!
ηl (k r ) ≃ − ,
(k r ) l+1
where n!! = n (n − 2) (n − 4) · · · 1.
◮ It follows that
−(k a) 2 l+1
tan δl = .
(2 l + 1) [(2 l − 1)!!] 2
◮ It is clear that we can neglect δl , with l > 0, with respect to
δ0 .
Hard-Sphere Scattering - V
◮ In other words, at low energy, only S-wave scattering (i.e.,
spherically symmetric scattering) is important.
◮ It follows from (10), (39), and (58) that

dσ sin2 (k a)
= ≃ a2 (60)
dΩ k2
for k a ≪ 1.
◮ Note that total cross-section,
I

σtotal = dΩ = 4π a2 ,
dΩ

is four times geometric cross-section, π a2 (i.e., cross-section


for classical particles bouncing off a hard sphere of radius a).
◮ However, low-energy scattering implies relatively long de
Broglie wavelengths, so we would not expect to obtain
classical result in this limit.
Hard-Sphere Scattering - VI

◮ Consider high-energy limit, k a ≫ 1.


◮ At high energies, by analogy with classical scattering,
scattered particles with largest angular momenta about origin
have angular momenta ~ k a (i.e., product of their incident
momenta, ~ k, and their maximum possible impact
parameters, a).
◮ Given
p that particles in l th partial wave have angular momenta
l (l + 1) ~, we deduce that all partial waves up to lmax ≃ k a
contribute significantly to scattering cross-section.
◮ It follows from (47) that

4π X
σtotal = (2 l + 1) sin2 δl . (61)
k2
l=0,lmax
Hard-Sphere Scattering - VII
◮ Making use of (57), as well as asymptotic expansions (25) and
(26), we find that

tan2 δl jl 2 (k a)
sin2 δl = = = sin 2 (k a − l π/2).
1 + tan2 δl jl 2 (k a) + ηl2 (k a)

◮ In particular,

sin2 δl + sin2 δl+1 = sin2 (k a − l π/2) + cos2 (k a − l π/2) = 1.

◮ Hence, it is a good approximation to write


2π X 2π
σtotal ≃ (2 l + 1) = (lmax + 1) 2 ≃ 2π a2 .
k2 k2
l=0,lmax

◮ This is twice classical result, which is somewhat surprising,


because we might expect to obtain classical result in
short-wavelength limit.
Hard-Sphere Scattering - VIII

◮ In fact, for hard-sphere scattering, all incident particles with


impact parameters less than a are deflected.
◮ However, in order to produce a shadow behind sphere, there
must be scattering in forward direction (recall optical theorem)
to produce destructive interference with incident plane wave.
◮ Effective cross-section associated with this forward scattering
is π a2 , which, when combined with cross-section for classical
reflection, π a2 , gives actual cross-section of 2π a2 .
Low-Energy Scattering - I

◮ In general, at low energies (i.e., when 1/k is much larger than


range of potential), partial waves with l > 0 make a negligible
contribution to scattering cross-section.
◮ It follows that, with a finite-range potential, only S-wave (i.e.,
spherically symmetric) scattering is important at such
energies.
Low-Energy Scattering - II
◮ As a specific example, let us consider scattering by a finite
potential well, characterized by V = V0 for r < a, and V = 0
for r ≥ a
◮ Here, V0 is a constant.
◮ Potential is repulsive for V0 > 0, and attractive for V0 < 0.
◮ External wavefunction is given by [see (51)]

A0 (r ) = exp( i δ0 ) [j0 (k r ) cos δ0 − η0 (k r ) sin δ0 ]


exp( i δ0 ) sin(k r + δ0 )
= ,
kr
where use has been made of (23) and (24).
◮ Internal wavefunction follows from (55). We obtain
sin(k ′ r )
A0 (r ) = B , (62)
r
where use has been made of boundary condition (56).
Low-Energy Scattering - III
◮ Here, B is a constant, and
~ 2 k′ 2
E − V0 = . (63)
2m
◮ Note that (62) only applies when E > V0 .
◮ For E < V0 , we have
sinh(κ r )
A0 (r ) = B ,
r
where
~ 2κ 2
V0 − E = . (64)
2m
◮ Matching A0 (r ), and its radial derivative, at r = a yields
k
tan(k a + δ0 ) = ′ tan(k ′ a) (65)
k
for E > V0 , and
k
tan(k a + δ0 ) = tanh(κ a)
κ
for E < V0 .
Low-Energy Scattering - IV
◮ Consider an attractive potential, for which E > V0 .
◮ Suppose that |V0 | ≫ E (i.e., depth of potential well is much
larger than energy of incident particles), so that k ′ ≫ k.
◮ As can be seen from (65), unless tan(k ′ a) becomes extremely
large, right-hand side of equation is much less than unity, so
replacing tangent of a small quantity with quantity itself, we
obtain
k
k a + δ0 ≃ ′ tan(k ′ a).
k
◮ This yields
tan(k ′ a)
 
δ0 ≃ k a − 1 .
k′ a
◮ According to (61), total scattering cross-section is given by
 ′ 2
4π 2 2 tan(k a)
σtotal ≃ 2 sin δ0 = 4π a −1 . (66)
k k′ a
Low-Energy Scattering - V

◮ Now, r
2 m |V0 | a2
k′ a = k 2 a2 + , (67)
~2
so for sufficiently small values of k a,
r
2 m |V0 | a2
k′ a ≃ .
~2
◮ It follows that total (S-wave) scattering cross-section is
independent of energy of incident particles (provided that this
energy is sufficiently small).
Low-Energy Scattering - VI

◮ Note that there are values of k ′ a (e.g., k ′ a ≃ 4.493) at which


scattering cross-section (66) vanishes, despite very strong
attraction of potential.
◮ In reality, cross-section is not exactly zero, because of
contributions from l > 0 partial waves. But, at low incident
energies, these contributions are small.
◮ It follows that there are certain values of |V0 |, a, and k that
give rise to almost perfect transmission of incident wave.
◮ This is called Ramsauer-Townsend effect, and has been
observed experimentally.
Resonant Scattering - I
◮ There is a significant exception to energy independence of
scattering cross-section at low incident energies described
previously.
◮ Suppose that quantity (2 m |V0 | a 2 /~ 2 )1/2 is slightly less than
π/2.
◮ As incident energy increases, k ′ a, which is given by (67), can
reach value π/2.
◮ In this case, tan(k ′ a) becomes infinite, so we can no longer
assume that right-hand side of (65) is small.
◮ In fact, at value of incident energy at which k ′ a = π/2, it
follows from (65) that k a + δ0 = π/2, or δ0 ≃ π/2 (because
we are assuming that k a ≪ 1).
◮ This implies that
 
4π 2 2 1
σtotal = 2 sin δ0 = 4π a .
k k 2 a2
Resonant Scattering - II

◮ Note that total scattering cross-section now depends on


energy. Furthermore, magnitude of cross-section is much
larger than that given in (66) for k ′ a 6= π/2 (because
k a ≪ 1, whereas k ′ a ∼ 1).
◮ Origin of rather strange behavior just described is easily
explained.
◮ Condition r
2 m |V0 | a2 π
2
=
~ 2
is equivalent to condition that a spherical well of depth |V0 |
possesses a bound state at zero energy.
◮ Thus, for a potential well that satisfies preceding equation,
energy of scattering system is essentially same as energy of
bound state.
Resonant Scattering - III

◮ In this situation, an incident particle would like to form a


bound state in potential well.
◮ However, bound state is not stable, because system has a
small positive energy.
◮ Nevertheless, this sort of resonant scattering is best
understood as capture of an incident particle to form a
metastable bound state, followed by decay of bound state and
release of particle.
◮ Cross-section for resonant scattering is generally far higher
than that for non-resonant scattering.
Resonant Scattering - IV
◮ We have seen that there is a resonant effect when phase-shift
of S-wave takes value π/2.
◮ There is nothing special about l = 0 partial wave, so it is
reasonable to assume that there is a similar resonance when
phase-shift of l th partial wave is π/2.
◮ Suppose that δl attains value π/2 at incident energy E0 , so
that
π
δl (E0 ) = .
2
◮ Let us expand cot δl in vicinity of resonant energy:
 
d cot δl
cot δl (E ) = cot δl (E0 ) + (E − E0 ) + · · ·
dE E =E0
 
1 dδl
=− (E − E0 ) + · · ·.
sin2 δl dE E =E0
Resonant Scattering - V
◮ Defining  
dδl (E ) 2
= ,
dE E =E0 Γ
we obtain
2
cot δl (E ) = − (E − E0 ) + · · ·.
Γ
◮ Recall, from (49), that contribution of l th partial wave to
total scattering cross-section is
4π 4π 1
σl = (2 l + 1) sin2 δl = 2 (2 l + 1) .
k2 k 1 + cot2 δl
◮ Thus,
4π Γ 2 /4
σl ≃ (2 l + 1)
k2 (E − E0 ) 2 + Γ 2 /4
which is known as Breit-Wigner formula.
◮ Variation of partial cross-section, σl , with incident energy has
form of a classical resonance curve.
Resonant Scattering - VI

◮ Quantity Γ is width of resonance (in energy).


◮ We can interpret Breit-Wigner formula as describing
absorption of an incident particle to form a metastable state,
of energy E0 , and lifetime τ = ~/Γ .
Elastic and Inelastic Scattering - I
◮ From before, for case of a spherically symmetric scattering
potential, scattered wave is characterized by
X
f (θ) = (2 l + 1) fl Pl (cos θ), (68)
l=0,∞

where
exp( i δl ) Sl − 1
fl = sin δl = (69)
k 2ik
is amplitude of l th partial wave, whereas δl is associated
phase-shift.
◮ Here,
Sl = e i 2 δl .
◮ Moreover, fact that |Sl | = 1 ensures that scattering is elastic
(i.e., that number of particles is conserved).
◮ Finally, net elastic scattering cross-section can be written
4π X X
σelastic = 2 (2 l + 1) sin2 δl = 4π (2 l + 1) |fl | 2 .
k
l=0,∞ l=0,∞
(70)
Elastic and Inelastic Scattering - II
◮ Turns out that many scattering experiments are characterized
by absorption of some of incident particles.
◮ Such absorption may induce a change in quantum state of
target, or, perhaps, emergence of another particle.
◮ Note that scattering that does not conserve particle number is
known as inelastic scattering.
◮ We can take inelastic scattering into account in our analysis
by writing
Sl = ηl e i 2 δl , (71)
where real parameter ηl is such that

0 ≤ ηl ≤ 1.
◮ It follows from (69) that
 
ηl sin(2 δl ) 1 − ηl cos(2 δl )
fl = +i .
2k 2k
Elastic and Inelastic Scattering - III
◮ Hence, according to (70), net elastic scattering cross-section
becomes
X
σelastic = 4π (2 l + 1) |fl | 2
l=0,∞
π X
(2 l + 1) 1 + ηl2 − 2 ηl cos(2 δl ) .
 
= 2 (72)
k
l=0,∞

◮ Net inelastic scattering (i.e., absorption) cross-section follows


from (9) and (45):
H 2
r (−jr ) dΩ π X
(2 l + 1) 1 − |Sl | 2

σinelastic = = 2
|jincident | k
l=0,∞
π X
(2 l + 1) 1 − ηl2 .

= 2 (73)
k
l=0,∞
Elastic and Inelastic Scattering - IV
◮ Thus, total cross-section is

σtotal = σelastic + σinelastic


2π X
= 2 (2 l + 1) [1 − ηl cos(2 δl )].
k
l=0,∞

◮ Note, from (68), (69), and (71) that

1 X
Im[f (0)] = (2 l + 1) [1 − ηl cos(2 δl )].
2k
l=0,∞

◮ In other words,

σtotal = Im[f (0)].
k
◮ Hence, we deduce that optical theorem still applies in
presence of inelastic scattering.
Elastic and Inelastic Scattering - V
◮ If ηl = 1 then there is no absorption, and l th partial wave is
scattered in a completely elastic manner.
◮ On other hand, if ηl = 0 then there is total absorption of l th
partial wave.
◮ However, such absorption is necessarily accompanied by some
degree of elastic scattering.
◮ In order to illustrate this important point, let us investigate
special case of scattering by a black sphere.
◮ Such a sphere has a well-defined edge of radius a, and is
completely absorbing.
◮ Consider short-wavelength scattering characterized by
k a ≫ 1.
◮ In this case, we expect all partial waves with l ≤ lmax , where
lmax ≃ k a, to be completely absorbed (because, by analogy
with classical physics, impact parameters of associated
particles are less than a), and all other partial waves to suffer
neither absorption nor scattering.
Elastic and Inelastic Scattering - VI
◮ In other words, ηl = 0 for 0 ≤ lmax , and ηl = 1, δl = 0 for
l > lmax .
◮ It follows from (72) and (73) that
π X π
σelastic = 2 (2 l + 1) = 2 (1 + lmax ) 2 ≃ π a2 ,
k k
l=0,lmax

and
π X π
σinelastic = (2 l + 1) = (1 + lmax ) 2 ≃ π a2 .
k2 k2
l=0,lmax

◮ Thus, total scattering cross-section is


σtotal = σelastic + σinelastic = 2π a2 .
◮ This result seems a little strange, at first, because, by analogy
with classical physics, we would not expect total cross-section
to exceed cross-section presented by sphere.
◮ Nor would we expect a totally absorbing sphere to give rise to
any elastic scattering.
Elastic and Inelastic Scattering - V

◮ In fact, this reasoning is incorrect. Absorbing sphere removes


flux proportional to π a2 from incident wave, which leads to
formation of a shadow behind sphere.
◮ However, a long way from sphere, shadow gets filled in.
◮ In other words, shadow is not visible infinitely far downstream
of sphere.
◮ Only way in which this can occur is via diffraction of some of
incident wave around edges of sphere.
◮ Actually, amount of incident wave that must be diffracted is
same amount as was removed from wave by absorption. Thus,
scattered flux is also proportional to π a2 .
Elastic and Inelastic Scattering - VI
◮ Consider low-energy scattering by a hard-sphere potential.
◮ This process is dominated by S-wave (i.e., l = 0) scattering.
◮ Moreover, phase-shift of S-wave takes form

δ0 = −k a,

where k is wavenumber of incident particles, and a is radius of


sphere.
◮ Note that low-energy limit corresponds to k a ≪ 1.
◮ It follows that
S0 = e i 2 δ0 ≃ 1 − 2 i k a.
◮ We can generalize previous analysis to take absorption into
account by writing

S0 ≃ 1 − 2 i k α,

where α is complex, k |α| ≪ 1, and Im(α) < 0.


Elastic and Inelastic Scattering - VII

◮ According to (72) and (73),


π
σelastic ≃ |S0 − 1| 2 ≃ 4π |α|2 , (74)
k2
π  4π Im(−α)
σinelastic ≃ 2 1 − |S0 | 2 ≃ . (75)
k k
◮ We conclude that low-energy elastic scattering cross-section is
again independent of incident particle velocity (which is
proportional to k), whereas inelastic cross-section is inversely
proportional to particle velocity.
◮ Consequently, as incident particle velocity decreases, inelastic
scattering becomes more and more important in comparison
with elastic scattering.
Scattering of Identical Particles - I

◮ Consider two identical particles that scatter off one another.


◮ In center of mass frame, there is no way of distinguishing a
deflection of a particle through an angle θ, and a deflection
through an angle π − θ, because momentum conservation
demands that if one of particles is scattered in direction
characterized by angle θ then other is scattered in direction
characterized by π − θ.
◮ Here, for sake of simplicity, we are assuming that scattering
potential is spherically symmetric, which implies that motion
of two particles is confined to a fixed plane passing through
origin.
Scattering of Identical Particles - II

◮ In classical mechanics, differential cross-section for scattering


is affected by identity of particles because number of particles
counted by a detector located at angular position θ is sum of
counts due to two particles, which implies that

dσclassical dσ(θ) dσ(π − θ)


= + = |f (θ)|2 + |f (π − θ)|2 .
dΩ dΩ dΩ
[See (10).]
Scattering of Identical Particles - III

◮ In quantum mechanics, overall wavefunction must be either


symmetric or antisymmetric under interchange of identical
particles, depending on whether particles in question are
bosons or fermions, respectively.
◮ If spatial wavefunction is symmetric then (20) is replaced by

eik r
 
1 i k r cos θ −i k r cos θ
ψ(x) = e + e + [f (θ) + f (π − θ)] ,
(2π)3/2 r

and associated differential scattering cross-section becomes



= |f (θ) + f (π − θ)|2 .
dΩ
[See (10).]
Scattering of Identical Particles - IV

◮ On other hand, if spatial wavefunction is antisymmetric then


(20) is replaced by

eik r
 
1 i k r cos θ −i k r cos θ
ψ(x) = e −e + [f (θ) − f (π − θ)] ,
(2π)3/2 r

and associated differential scattering cross-section is written



= |f (θ) − f (π − θ)|2 .
dΩ
Scattering of Identical Particles - V
◮ For case of two identical spin-zero (i.e., boson) particles (e.g.,
α-particles), spatial wavefunction is symmetric with respect to
particle interchange, which implies that

= |f (θ) + f (π − θ)|2
dΩ
= |f (θ)|2 + |f (π − θ)|2 + [f ∗ (θ) f (π − θ) + f (θ) f ∗ (π − θ)].
◮ Previous result differs from classical one because of
interference term (i.e., final term on right-hand side), which
leads to an enhancement of differential scattering
cross-section at θ = π/2.
◮ In fact,  

= 4 [f (π/2)] 2 ,
dΩ θ=π/2
whereas  
dσclassical
= 2 [f (π/2)] 2 .
dΩ θ=π/2
Scattering of Identical Particles - VI
◮ For case of two identical spin-1/2 (i.e., fermion) particles
(e.g., electrons or protons), overall wavefunction is
antisymmetric under particle interchange.
◮ If two particles are in spin singlet state then spatial
wavefunction is symmetric (because spin wavefunction is
antisymmetric), and
dσsinglet
= |f (θ) + f (π − θ)| 2
dΩ
2 2
= |f (θ)| + |f (π − θ)| + [f ∗ (θ) f (π − θ) + f (θ) f ∗ (π − θ)].

◮ On other hand, if two particles are in spin triplet state then


spatial wavefunction is antisymmetric (because spin
wavefunction is symmetric), which leads to
dσtriplet 2
= |f (θ) − f (π − θ)|
dΩ
= |f (θ)|2 + |f (π − θ)|2 − [f ∗ (θ) f (π − θ) + f (θ) f ∗ (π − θ)].
Scattering of Identical Particles - VII

◮ In former case, interference term leads to an enhancement


(with respect to the classical case) of differential scattering
cross-section at θ = π/2: that is,
 
dσsinglet
= 4 [f (π/2)] 2 .
dΩ θ=π/2

◮ In latter case, interference term leads to complete suppression


of scattering in direction θ = π/2: that is,
 
dσtriplet
= 0.
dΩ θ=π/2
Scattering of Identical Particles - VIII
◮ Consider mutual scattering of two unpolarized beams of
spin-1/2 particles.
◮ All spin states are equally likely, so probability of finding a
given pair of particles (one from each beam) in triplet state is
three times that of finding it in singlet state, which implies
that
     
dσunpolarized 1 dσsinglet 3 dσtriplet
= +
dΩ 4 dΩ 4 dΩ
= |f (θ)|2 + |f (π − θ)|2
1 ∗
− [f (θ) f (π − θ) + f (θ) f ∗ (π − θ)].
2
◮ In this case, interference term leads to incomplete suppression (with
respect to classical case) of differential scattering cross-section at
θ = π/2: that is,
 
dσunpolarized 2
= [f (π/2)] .
dΩ θ=π/2

You might also like