Rutherford scattering

Last updated
Figure 1. In a cloud chamber, a 5.3 MeV alpha particle track from a lead-210 pin source near point 1 undergoes Rutherford scattering near point 2, deflecting by an angle of about 30deg. It scatters once again near point 3, and finally comes to rest in the gas. The target nucleus in the chamber gas could have been a nitrogen, oxygen, carbon, or hydrogen nucleus. It received enough kinetic energy in the elastic collision to cause a short visible recoiling track near point 2. (The scale is in centimeters.) AlphaTrackRutherfordScattering3.jpg
Figure 1. In a cloud chamber, a 5.3 MeV alpha particle track from a lead-210 pin source near point 1 undergoes Rutherford scattering near point 2, deflecting by an angle of about 30°. It scatters once again near point 3, and finally comes to rest in the gas. The target nucleus in the chamber gas could have been a nitrogen, oxygen, carbon, or hydrogen nucleus. It received enough kinetic energy in the elastic collision to cause a short visible recoiling track near point 2. (The scale is in centimeters.)

In particle physics, Rutherford scattering is the elastic scattering of charged particles by the Coulomb interaction. It is a physical phenomenon explained by Ernest Rutherford in 1911 [1] that led to the development of the planetary Rutherford model of the atom and eventually the Bohr model. Rutherford scattering was first referred to as Coulomb scattering because it relies only upon the static electric (Coulomb) potential, and the minimum distance between particles is set entirely by this potential. The classical Rutherford scattering process of alpha particles against gold nuclei is an example of "elastic scattering" because neither the alpha particles nor the gold nuclei are internally excited. The Rutherford formula (see below) further neglects the recoil kinetic energy of the massive target nucleus.

Contents

Rutherford scattering is now exploited by the materials science community in an analytical technique called Rutherford backscattering.

Key experiments

Hans Geiger, working in Rutherford's lab, did a series experiments in 1908 showing that alpha particles are "scattered" as they pass through thin layers of mica, and foils of gold and aluminum. In following year, joined by undergraduate Ernest Marsden, they did a series of experiments to untangle confusing results they observed. [2] :263

A critical discovery was made by Geiger and Marsden in 1909 when they performed the gold foil experiment in collaboration with Rutherford, in which they fired a beam of alpha particles (helium nuclei) at foils of gold leaf. [3] At the time of the experiment, the atom was thought to be analogous to a plum pudding (as proposed by J. J. Thomson), with the negatively-charged electrons (the plums) studded throughout a positive spherical matrix (the pudding). The scattering in this model was proposed to occur by many repeated collisions with the electrons and the alpha particles should only be deflected by small angles as they pass through.

However, the intriguing results showed that around 1 in 8,000 [2] :264 alpha particles were deflected by very large angles (over 90°), while the rest passed through with little deflection. From this, Rutherford concluded that the majority of the mass was concentrated in a minute, positively-charged region (the nucleus) surrounded by electrons. When a (positive) alpha particle approached sufficiently close to the nucleus, it was repelled strongly enough to rebound at high angles. The small size of the nucleus explained the small number of alpha particles that were repelled in this way. Rutherford showed, using the method outlined below, that the size of the nucleus was less than about 10−14 m (how much less than this size, Rutherford could not tell from this experiment alone; see more below on this problem of lowest possible size). As a visual example, Figure 1 shows the deflection of an alpha particle by a nucleus in the gas of a cloud chamber.

Maximum nuclear size estimate

Assuming there are no external forces and initially the alpha particles are far from the nucleus, the inverse-square law between the charges on the alpha particle and nucleus gives the potential energy gained by the particle as it approaches the nucleus. For head-on collisions between alpha particles and the nucleus, all the kinetic energy of the alpha particle is turned into potential energy and the particle stops and turns back. Where the particle stops, the potential energy matches the original kinetic energy: [4] :620 [5] :320

Rearranging:

For an alpha particle:

The distance from the alpha particle to the center of the nucleus (rmin) at this point is an upper limit for the nuclear radius. Substituting these in gives the value of about 2.7×10−14 m, or 27  fm. (The true radius is about 7.3 fm.) The true radius of the nucleus is not recovered in these experiments because the alphas do not have enough energy to penetrate to more than 27 fm of the nuclear center, as noted, when the actual radius of gold is 7.3 fm.

Rutherford's 1911 paper [1] started with a slightly different formula suitable for head-on collision with a sphere of positive charge:

Here is the turning point distance and is the radius of the positive charge. The first term is the Coulomb repulsion used above. This form assumes the alpha particle could penetrate the positive charge. At the time of Rutherford's paper Thomson's plum pudding model proposed a positive charge with the radius of an atom, thousands of times larger than the rmin found above.

Derivation

[ citation needed ]

The differential cross section can be derived from the equations of motion for two charged point particles interacting through a central potential. In general, the equations of motion describing two particles interacting under a central force can be decoupled into the center of mass and the motion of the particles relative to one another. Consider the situation where one particle (labelled 1), with mass and charge with the elementary charge, is incident from very far away with some initial speed on a second particle (labelled 2) with mass and charge initially at rest. For the case of light alpha particles scattering off heavy nuclei, as in the experiment performed by Rutherford, the reduced mass, essentially the mass of the alpha particle and the nucleus off of which it scatters, is essentially stationary in the lab frame.

Substituting into the Binet equation, with the origin of coordinate system for particle 1 on the target (scatterer, particle 2), yields the equation of trajectory as

where u = 1/r and b is the impact parameter.

The general solution of the above differential equation is

and the boundary condition is

Solving the equations u → 0 using those boundary conditions:

At the initial position of the alpha particle, where

and its derivative du/1/b using those boundary conditions. By similar triangles, if y is the coordinate going up from the bottom of the page,

We can obtain


is the angle of closest approach (the periapsis).

At the deflection angle Θ after collision, :

Then deflection angle Θ can be expressed as:

b can be solved to give

To find the scattering cross section from this result consider its definition

Given the Coulomb potential and the initial kinetic energy of the incoming particles, the scattering angle Θ is uniquely determined by the impact parameter b. Therefore, the number of particles scattered into an angle between Θ and Θ + dΘ must be the same as the number of particles with associated impact parameters between b and b + db. For an incident intensity I, this implies the following equality

For a radially symmetric scattering potential, as in the case of the Coulomb potential, dΩ = 2π sin Θ dΘ, yielding the expression for the scattering cross section

Plugging in the previously derived expression for the impact parameter b(Θ) we find the Rutherford differential scattering cross section

This same result can be expressed alternatively as

where α1/137 is the dimensionless fine structure constant, EK0 is the initial non-relativistic kinetic energy of particle 1 in MeV, and ħc197 MeV·fm.

Extension to situations with relativistic particles and target recoil

The extension of low-energy Rutherford-type scattering to relativistic energies and particles that have intrinsic spin is beyond the scope of this article. For example, electron scattering from the proton is described as Mott scattering, [6] with a cross section that reduces to the Rutherford formula for non-relativistic electrons. If no internal energy excitation of the beam or target particle occurs, the process is called "elastic scattering", since energy and momentum have to be conserved in any case. If the collision causes one or the other of the constituents to become excited, or if new particles are created in the interaction, then the process is said to be "inelastic scattering".

Target recoil can be handled by converting from the center of mass frame to a lab frame. [7] :85 In the lab frame, denoted by a subscript L, the scattering angle for a general central potential is

For , . For a heavy particle 1, and , that is, the incident particle is deflected through a very small angle.

For any central potential, the differential cross-section in the lab frame is related to that in the center-of-mass frame by

where

See also

Related Research Articles

In physics, the cross section is a measure of the probability that a specific process will take place in a collision of two particles. For example, the Rutherford cross-section is a measure of probability that an alpha particle will be deflected by a given angle during an interaction with an atomic nucleus. Cross section is typically denoted σ (sigma) and is expressed in units of area, more specifically in barns. In a way, it can be thought of as the size of the object that the excitation must hit in order for the process to occur, but more exactly, it is a parameter of a stochastic process.

<span class="mw-page-title-main">Centripetal force</span> Force directed to the center of rotation

A centripetal force is a force that makes a body follow a curved path. The direction of the centripetal force is always orthogonal to the motion of the body and towards the fixed point of the instantaneous center of curvature of the path. Isaac Newton described it as "a force by which bodies are drawn or impelled, or in any way tend, towards a point as to a centre". In Newtonian mechanics, gravity provides the centripetal force causing astronomical orbits.

In optics, Lambert's cosine law says that the radiant intensity or luminous intensity observed from an ideal diffusely reflecting surface or ideal diffuse radiator is directly proportional to the cosine of the angle θ between the observer's line of sight and the surface normal; I = I0 cos θ. The law is also known as the cosine emission law or Lambert's emission law. It is named after Johann Heinrich Lambert, from his Photometria, published in 1760.

<span class="mw-page-title-main">Bremsstrahlung</span> Electromagnetic radiation due to deceleration of charged particles

In particle physics, bremsstrahlung is electromagnetic radiation produced by the deceleration of a charged particle when deflected by another charged particle, typically an electron by an atomic nucleus. The moving particle loses kinetic energy, which is converted into radiation, thus satisfying the law of conservation of energy. The term is also used to refer to the process of producing the radiation. Bremsstrahlung has a continuous spectrum, which becomes more intense and whose peak intensity shifts toward higher frequencies as the change of the energy of the decelerated particles increases.

<span class="mw-page-title-main">Solid angle</span> Measure of how large an object appears to an observer at a given point in three-dimensional space

In geometry, a solid angle is a measure of the amount of the field of view from some particular point that a given object covers. That is, it is a measure of how large the object appears to an observer looking from that point. The point from which the object is viewed is called the apex of the solid angle, and the object is said to subtend its solid angle at that point.

<span class="mw-page-title-main">Tautochrone curve</span> Curve for which the time to roll to the end is equal for all starting points

A tautochrone curve or isochrone curve is the curve for which the time taken by an object sliding without friction in uniform gravity to its lowest point is independent of its starting point on the curve. The curve is a cycloid, and the time is equal to π times the square root of the radius over the acceleration of gravity. The tautochrone curve is related to the brachistochrone curve, which is also a cycloid.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

<span class="mw-page-title-main">Klein–Nishina formula</span> Electron-photon scattering cross section

In particle physics, the Klein–Nishina formula gives the differential cross section of photons scattered from a single free electron, calculated in the lowest order of quantum electrodynamics. It was first derived in 1928 by Oskar Klein and Yoshio Nishina, constituting one of the first successful applications of the Dirac equation. The formula describes both the Thomson scattering of low energy photons and the Compton scattering of high energy photons, showing that the total cross section and expected deflection angle decrease with increasing photon energy.

In quantum physics, the scattering amplitude is the probability amplitude of the outgoing spherical wave relative to the incoming plane wave in a stationary-state scattering process. At large distances from the centrally symmetric scattering center, the plane wave is described by the wavefunction

<span class="mw-page-title-main">Etendue</span> Measure of the "spread" of light in an optical system

Etendue or étendue is a property of light in an optical system, which characterizes how "spread out" the light is in area and angle. It corresponds to the beam parameter product (BPP) in Gaussian beam optics. Other names for etendue include acceptance, throughput, light grasp, light-gathering power, optical extent, and the AΩ product. Throughput and AΩ product are especially used in radiometry and radiative transfer where it is related to the view factor. It is a central concept in nonimaging optics.

Radiative transfer is the physical phenomenon of energy transfer in the form of electromagnetic radiation. The propagation of radiation through a medium is affected by absorption, emission, and scattering processes. The equation of radiative transfer describes these interactions mathematically. Equations of radiative transfer have application in a wide variety of subjects including optics, astrophysics, atmospheric science, and remote sensing. Analytic solutions to the radiative transfer equation (RTE) exist for simple cases but for more realistic media, with complex multiple scattering effects, numerical methods are required. The present article is largely focused on the condition of radiative equilibrium.

A theoretical motivation for general relativity, including the motivation for the geodesic equation and the Einstein field equation, can be obtained from special relativity by examining the dynamics of particles in circular orbits about the Earth. A key advantage in examining circular orbits is that it is possible to know the solution of the Einstein Field Equation a priori. This provides a means to inform and verify the formalism.

<span class="mw-page-title-main">Pendulum (mechanics)</span> Free swinging suspended body

A pendulum is a body suspended from a fixed support so that it swings freely back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back towards the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

<span class="mw-page-title-main">Axis–angle representation</span> Parameterization of a rotation into a unit vector and angle

In mathematics, the axis–angle representation parameterizes a rotation in a three-dimensional Euclidean space by two quantities: a unit vector e indicating the direction (geometry) of an axis of rotation, and an angle of rotation θ describing the magnitude and sense of the rotation about the axis. Only two numbers, not three, are needed to define the direction of a unit vector e rooted at the origin because the magnitude of e is constrained. For example, the elevation and azimuth angles of e suffice to locate it in any particular Cartesian coordinate frame.

The scattering length in quantum mechanics describes low-energy scattering. For potentials that decay faster than as , it is defined as the following low-energy limit:

In mathematics, vector spherical harmonics (VSH) are an extension of the scalar spherical harmonics for use with vector fields. The components of the VSH are complex-valued functions expressed in the spherical coordinate basis vectors.

The direct-quadrature-zerotransformation or zero-direct-quadraturetransformation is a tensor that rotates the reference frame of a three-element vector or a three-by-three element matrix in an effort to simplify analysis. The DQZ transform is the product of the Clarke transform and the Park transform, first proposed in 1929 by Robert H. Park.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

In physics, and especially scattering theory, the momentum-transfer cross section is an effective scattering cross section useful for describing the average momentum transferred from a particle when it collides with a target. Essentially, it contains all the information about a scattering process necessary for calculating average momentum transfers but ignores other details about the scattering angle.

References

  1. 1 2 Rutherford, E. (1911). "LXXIX. The scattering of α and β particles by matter and the structure of the atom" (PDF). The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 21 (125): 669–688. doi:10.1080/14786440508637080. ISSN   1941-5982.
  2. 1 2 Heilbron, John L. (1968). "The Scattering of α and β Particles and Rutherford's Atom". Archive for History of Exact Sciences. 4 (4): 247–307. ISSN   0003-9519.
  3. Geiger, H.; Marsden, E. (1909). "On a Diffuse Reflection of the α-Particles". Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences. 82 (557): 495–500. Bibcode:1909RSPSA..82..495G. doi: 10.1098/rspa.1909.0054 . Archived from the original on January 2, 2008.
  4. "Electrons (+ and -), Protons, Photons, Neutrons, Mesotrons and Cosmic Rays" By Robert Andrews Millikan. Revised edition. Pp. x+642. (Chicago: University of Chicago Press; London: Cambridge University Press, 1947.)
  5. Cooper, L. N. (1970). "An Introduction to the Meaning and Structure of Physics". Japan: Harper & Row.
  6. "Hyperphysics link".
  7. Goldstein, Herbert. Classical Mechanics. United States, Addison-Wesley, 1950.

Textbooks