1 areas are missing area_end: [Area(name='partialwaves', attrs={'area': 'partialwaves', 'collapse': 'true'}, visible=True)]
# ii.-scattering-theory

Scattering theory

Learning goals for this week (check afterwards that you understand these concepts)

  • Understanding the concept of the scattering experiments
  • QM formulation of the scattering experiment
  • Cross section, differential cross section
  • Integral equation for scattering and the Green's function of the scattering problem
  • Complex integration; residue theorem; transforming an integral over a real axis to a closed complex line integral
  • Solving for the Green's function via complex integration
  • Born series solution of the integral equation; Born approximation
# ii.-1.-basic-concept-cross-section

Basic concept: cross section

The structure of matter is probed by doing

  1. Spectroscopy: observe transitions between the different bound states at the system (atoms, molecules, ) - analyzed with the methods of the time-dependent perturbation theory, often with harmonic perturbations.

  2. Scattering experiments: analyze the final states and deduce the properties of target and/or beam particles (for example the structure and properties of nuclei and elementary particles, or nature of electron transport inside solids)

A typical scattering experiment (with a fixed target) could look like this:

It thus consists of two parts:

  • beam:

    • consists of particles A (electrons, positrons, protons, neutrons, photons, phonons, )

    • is nearly monoenergetic and well collimated: (allows considering a given input energy/momentum)

  • target: (fixed target macroscopic sample)

    • consists often of a large number of particles B (e.g., H-gas in a container, stack of lead plates, etc.)

    • or an impurity potential in a solid (dislocations, vacancies, and other irregularities of the ion lattice)

One can generally separate the scattering to two categories:

  • Elastic scattering: Same particles in the initial and final states, i.e., total kinetic energy is conserved

  • Inelastic scattering: final state particles initial state particles, or the structure of A,B can change.

For example: could be for example

  1. : elastic scattering

  2. : excitation of

  3. : ionization of

  4. : annihilation

  5. : positronium formation

(ii)-(v) are inelastic

The "final state" particles are taken to a detector, which can then measure for example the magnitude of their momentum (or their kinetic energy) as a function of the angles , . In addition, the detector can be sensitive to the polarization, spin, or some other internal property of the output beam.

A key quantity in scattering experiments is the cross section for scattering = number of these scattering events per unit time, per one target particle B, divided by the flux of incoming beam particles A

Let's illustrate it here:

Let us assume that the detector observes the number of scattering events in a unit time. Obviously, the more flux and the larger , the larger must be:

# crosssection

We can therefore define the scattering cross section as

It depends on the scattering process considered, i.e., on the probability of scattering, i.e., on the strength and range of the interaction and the beam energy

Here is an effective "cross section" which the beam particles see. It is also related to the probability for the scattering to happen. These have the physical units:

Units used in nuclear and particle physics: 1 b = 1 barn = 10 m = 100 fm.

Often, it is not easy to detect the particles scattered in all possible directions, or there is extra information in the precise angle to which the scattering takes place. Therefore, one often also considers the differential cross section where is the number of particles per unit time, scattered into the infinitesimal solid angle : Image

For example experimental figures on scattering cross sections in particle physics, see for example K. Nakamura, et al., Review of Particle Physics, J. Phys. G: Nucl. Part. Phys. 37, 075021 (2010).

You can get more information on particle physics measurements in the course FYSS4300 Particle Physics.

In this chapter,

  1. We formulate the general 3d scattering problem, where the wave function far away from the scattering region satisfies The rest of the chapter deals with methods aiming to solve .

  2. We use the Green's function technique to transform the Schrödinger equation to an integral equation. This is then solved with the Born approximation. It is especially useful for weak potentials, where the deviation of the particle's initial trajectory is small.

  3. Another approach is the partial wave analysis, where the dependence is expanded in terms of the spherical harmonics. This method is useful especially for low-energy particles (compared to the potential strength), since for that case is a weak function of

Alternatively to the lecture notes, you may also follow Ch. 11 in Griffiths' book.

# ii.2-formulation-of-the-3d-scattering-problem-boundary-conditions-of-the-wave-function-cross-section-optical-theorem

Formulation of the 3d scattering problem: boundary conditions of the wave function, cross section, optical theorem

We consider here the elastic scattering of two spinless, nonrelativistic particles. In laboratory frame, the incoming particles have a certain momentum and the target is at rest (sometimes referred to as the target rest frame). In the center-of-momentum (CMS) frame, the total momentum of an incoming particle and a target particle is zero:

In the CMS frame the total momentum of the system vanishes, and hence , the relevant energy is the relative kinetic energy between the particles, and the relevant position is the relative position, . This is equivalent to a scattering of a particle of a reduced mass of a potential .

In order to be able to calculate the cross section, we must assume a finite-range, localized (and often spherically symmetric) potential ; :

We proceed analogously to what was done in QM I in the case of a 1D problem (Griffiths, Ch. II). Instead of using calculationally much more difficult wave packets and their time evolution to describe particles and their motion, we assume a stationary situation, i.e., that the beam has been "turned on" already for a while and that we thus have a steady flux of incoming and scattered particles. This "turning on" is described in more detail in the chapter on time-dependent phenomena. We thus consider the scattering problem within the following framework: Image That is, at , the particles behave as free, .

Let us first specify the incoming plane wave in the 3d case, far away from the scattering center . The stationary Schrödinger equation reads in the position representation This is solved with the plane wave solution (prefactor comes from the properties of Fourier transformation) This is hence the form of the wave function for the beam of particles far away from the scattering center.

For a radially symmetric potential, we can make the separating Ansatz (see Griffiths, Ch. IV). There, satisfies the radial Schrödinger equation i.e., we assume to decay faster than the centrifugal term at large . In other words,

The general solution to this can be specified in terms of spherical Bessel () and Neumann functions: For large arguments, they have the asymptotic behavior

# radialequationsolution

For a large , the solution to the radial equation can hence be written in terms of plane waves:

Identification in terms of incoming and outgoing waves can be done by considering the probability flux, or probability current density This can be derived from the general continuity equation defining a relation between the probability density and probability current density.

QM expression for the current density obtained from the continuity equation

For the plane wave solution describing the incoming wave, , we get In defining the cross section, we need the amount (per unit time) of scattered particles moving through a sphere surface at : Recall the spherical-coordinates representation of gradient: so that

1=jin

Comparing to the generic solution to the radial equation we hence want to see what happens with . Direct calculation gives The total particle flux is thus independent of the radial coordinate , which is desired because there are no sources or sinks of particles. Now we can identify what the signs mean:
"+" outgoing spherical wave
"-" incoming spherical wave
Moreover, we find that the proper asymptotic behavior of an outgoing spherical wave is .

Now we are in position of defining the quantum scattering problem:

# quantumscatteringproblem

Quantum scattering problem

With a potential decaying faster than , find the solutions of the spherical Schrödinger equation or fulfilling the boundary condition

# psifaraway

Here

For the wave function modulus squared to integrate (over all space) to probabilities (dimensionless number), the units of the wave function have to be . This unit is thus ''hidden'' in the normalization constant. The square bracket is dimensionless, so the scattering amplitude must have units .

We have two remaining problems: What is if is known? What is the relation between and the scattering cross section ?

Note that is a generalization of the transmission and reflection probability amplitudes from the one-dimensional scattering problem.

To answer the second question, let us check the particle fluxes for our Ansatz asymptotic wave function. The total flux is

Therefore, we find that the total current consists of three parts. The first of these is nothing but the flux of incoming particles, denoted above. To understand the second, let us consider the number of particles scattered into a solid angle in a unit time per target particle Here In other words, and therefore

Using from Eq. (1) we can hence define the cross section, yielding and where As the scattering amplitude has units , it makes sense that has units of the (differential) cross section, i.e., .

We have thus answered the second question above: finding the (absolute value of the) scattering amplitude amounts to calculating the differential cross section. In retrospect it is quite clear that the two quantities must contain the same information, because they are the only ingredients left open after specifying the scattering setup!

1=jin

Interference current and the optical theorem

# ii.-3.-integral-equation-for-potential-scattering

Integral equation for (potential) scattering

After having specified the scattering problem, we should try to solve it at least in some particular cases. In other words, we wish to solve the 3d differential equation

# 3ddiffeq

with the asymptotic boundary condition Note that here the vector , is some quantity specified for the problem. In particular, the direction of the vector specifies a reference direction.

To solve the problem, let us first define

  1. The solution of the homogeneous () equation:

  2. Green's function satisfying

# integralequation

The differential equation can now be cast in the form of an integral equation: We wish to use this integral equation to help find a general scheme for solving the scattering problem. This is done below, but before we manage to do so, we need to discuss the Green's function in more detail.

Green's functions do not only depend on the differential equation, but also on the boundary conditions. Now, in order to fulfill the boundary conditions of the scattering problem, has to be taken to be the incoming plane wave, and we have to define the Green's function suitably, so that i.e., it should correspond to the outgoing spherical wave.

Let us make a Fourier expansion of , by assuming that it only depends on the difference of its arguments: It satisfies or

In order to compute , we need to consider in the complex plane, , and set at the end of the calculation. But how should it tend to 0, from positive or negative values? As a matter of fact, the choice of the sign of determines different Green's functions.

We hence have , and using the inverse Fourier transform

# GFequation

we get , where This integral can be computed as an integral in the complex -plane, using the residue theorem. This means that we need to introduce some mathematical machinery that comes handy in various physics problems, but especially those dealing with Green's functions.

If you already know how to compute the integral via the residue theorem, you may skip directly to the discussion of that calculation.

# ii.3.1-functions-of-complex-variable-and-their-integration

Functions of complex variable and their integration

Let us start by the definition of a complex derivative This should be "isotropic", i.e., independent of the direction of : Image In other words,

Let us separate the real and imaginary parts of , where , . Then and

# cauchyriemann

Equating the real and imaginary parts separately yields These are known as the Cauchy-Riemann equations.

# contour-integrals-viivaintegraalit

Contour integrals (viivaintegraalit)

Consider a closed path in the complex plane: Let us moreover define the winding direction to be counterclockwise (arrows in the figure). The contour integral along can be defined in parts, Let to be analytic on and within above. Consider the contour integral along :

The last result was obtained by using the Cauchy-Riemann equations: both integrands vanish. In other words, This results allows us to state the

Cauchy's integral theorem

A contour integral of a complex function over a closed and simply connected (no holes) region vanishes if is analytic on and within its interior. In other words, has no singularities inside or on .

Now let us start transforming the contour. For example, take : Image Assume that there are no holes or singularities within , i.e., the shaded region in the figure. Then, Cauchy's integral theorem states that Changing the direction in this then yields In this way, we can "shrink" the contours. Note that the area spanned by is smaller than the one spanned by . However, the integrals are the same --- and not necessarily zero!

We can generalize this result to many loops, which may contain singularities at :
Note that all contours run in the counterclockwise direction.


A contour integral of a complex function over a closed and simply connected (no holes) region vanishes if is analytic on and within its interior. In other words, has no singularities inside or on .


Suppose now that has an isolated singularity at . This means that is analytic in the area surrounding but not exactly at . By the definition of analyticity, cannot be written in terms of its Taylor series. Rather, we have to use the Laurent series of Let us assume there is an , so that . Then Above, we call
a pole of order
the residue of at Let now have a pole of order at ,

Take as a circle around : i.e., , . The contour integral becomes

This result can be generalized to the case of many singularities at We have derived the Residue theorem, a very important result.

The residues can be obtained from

  1. Expanding as a Laurent series and taking the coefficient or, more easily,

  2. using a pocket formula for a pole of order (can be verified trivially)

In the quite often encountered special case of a simple pole at we have and Let us now use these results for the evaluation of the Green's function.

# ii.3.2-evaluation-of-the-greens-function

Evaluation of the Green's function

We can finally come back to our initial problem of calculating the Green's function. We would like to calculate

Since , the poles of are at and , or and since the poles at are simple poles of order 1. Let us calculate the residues: These are the residues of at .

Wait! We have a simple integral over real values of . Where is the contour? The trick is to consider the integral over the real axis as part of a closed contour that closes along the complex plane as shown in the following picture (where we have chosen ):

We should hence "close" the contour along the upper or lower complex half-plane. But which one should we use? This we see below. Either way, we can see that one of the poles is inside the contour , and the other one is outside. Anyway, the idea is the following: the residue theorem says

The first term inside the square brackets is the one we wish to compute. Now our aim is to choose the contour or so that the second integral vanishes. Let us choose first , i.e., close the contour in the upper half-plane: Image

In this case, the pole is inside , and is outside it. In other words,

For , we can set , . Let us show that the contour integral along vanishes as . We have Now take the limit . The absolute value of tends to

The integral along contour running along the upper half-plane therefore vanishes: note that the length of the contour is , and hence increases linearly with an increasing . However, the integrand vanishes exponentially with an increasing . Exponential function wins a linear one, and thus the result vanishes. We can see why we have to close the contour through the upper half plane, as for we would have (again ) and The integral along the lower half-plane would hence not vanish for .

In general which half-plane to close the contour depends on the case. For exponential functions with poles in the upper half-plane, we can prove the

Jordan's lemma: If , when , (upper half-plane), and , then

Proof, using and hence for a fixed :

Now let us apply Jordan's lemma to our problem: and we have that , so it fulfills the assumptions. This means that as , the integral of of the contour over the upper half-plane vanishes. Moreover, from the equation for the Green's function we have for

In other words, the Green's function is Let us quickly check what would happen with the choise . In this case the pole at would be outside and the pole inside it:

Now you should show that we get and the corresponding Green's function would be In other words, Next, we show that it is that we want for the scattering problem, to describe an outgoing spherical wave.

# ii.3.3-integral-equation-for-potential-scattering-ieps

Integral equation for potential scattering (IEPS)

Now, finally, substitute the above Green's function back to the integral equation for scattering, still allowing for either sign of . We get

# IEPS

Here in some vicinity of . We should check the boundary condition at (not ) and choose the Green's function (+ or -) accordingly. Therefore, we should assume , and expand: Note that we identified (defined) here the wave vector of the scattered particle, . As we are describing elastic scattering (energy of the incoming particle equals that of the outgoing particle), it has the same magnitude as the incoming particle, but a different direction.

Inserting this expansion into the integral equation for potential scattering yields in the lowest order Comparing this to the desired form, we find that in order to get the proper behavior, we have to choose in the Green's function .

Now recall that Moreover, comparing to the boundary condition, we can find an expression for the scattering amplitude or This we need for the calculation of the differential scattering cross section.

# ii.4.-solving-the-ieps-born-approximation

Solving the IEPS: Born approximation

Let us then solve the IEPS via an iteration method, using the fact that the integral equation contains on both sides of the equation. Let us hence substitute the left hand side to the right hand side. We get

Substituting repeatedly to this equation hence generates the Born series of Note that each new term contains a higher-order power of . It is hence an expansion in .

Let us write the scattering amplitude in a similar series. For illustration, let us define the wave vector of the incoming particle or "initial state wave vector" by and that of the outgoing particle or "final state wave vector" by . The Born series of the scattering amplitude is We can interpret this graphically and physically as follows:

The Green's function is also called the propagator, which describes the propagation of a free particle from to . The interpretation of the Born series is thus that consists of a series of processes containing scattering events, where the incoming particle scatters off at , after which it moves as a free particle to the point , where it scatters off again, and so on. Obviously, if is very weak relative to the incident energy of the scattering particle(s), the contribution from multiple scatterings to the total scattering amplitude is small. In this case only the first term in the Born series is of relevance.

It can be shown that the Born series converges for all , if does not have bound states, i.e., is weak enough, and for (sufficiently) large , if , where and , where . We can then define the

Born approximation = the first term in the Born series for

Let us examine this further by defining the momentum exchange vector By straightforward trigonometry (see the picture) one can show that Note that the momentum exchange vector depends on the angle . Using we can hence write the Born approximated form of the scattering amplitude in terms of the Fourier transform of the potential and hence

For a spherically symmetric potential, we have

So, for a spherically symmetric potential we have Remember that the -dependence is in .

Example: Yukawa potential

Phew! There are a lot of technical details in the above sections. What should you take along? Well, at least

  • Definition of the scattering problem, total and differential cross section

  • Quantum scattering problem and the scattering amplitude

  • Green's function and integral equation

  • Solving the scattering Green's function: small imaginary term into wave number; residue theorem (complex integration)

  • Using residue theorem to evaluate certain types of integrals

  • Once we have the Green's function, for "weak enough" potentials we can solve the scattering problem via Born series

  • Note how the Born approximated version couples to the scattering amplitude.

Partial wave expansion

These are the current permissions for this document; please modify if needed. You can always modify these permissions from the manage page.