Beam emittance

Last updated
Samples of a bivariate normal distribution, representing particles in phase space, with position horizontal and momentum vertical. GaussianScatterPCA.svg
Samples of a bivariate normal distribution, representing particles in phase space, with position horizontal and momentum vertical.

In accelerator physics, emittance is a property of a charged particle beam. It refers to the area occupied by the beam in a position-and-momentum phase space. [1]

Contents

Each particle in a beam can be described by its position and momentum along each of three orthogonal axes, for a total of six position and momentum coordinates. When the position and momentum for a single axis are plotted on a two dimensional graph, the average spread of the coordinates on this plot are the emittance. As such, a beam will have three emittances, one along each axis, which can be described independently. As particle momentum along an axis is usually described as an angle relative to that axis, an area on a position-momentum plot will have dimensions of length × angle (for example, millimeters × milliradian). [1] :78–83

Emittance is important for analysis of particle beams. As long as the beam is only subjected to conservative forces, Liouville's Theorem shows that emittance is a conserved quantity. If the distribution over phase space is represented as a cloud in a plot (see figure), emittance is the area of the cloud. A variety of more exact definitions handle the fuzzy borders of the cloud and the case of a cloud that does not have an elliptical shape. In addition, the emittance along each axis is independent unless the beam passes through beamline elements (such as solenoid magnets) which correlate them. [2]

A low-emittance particle beam is a beam where the particles are confined to a small distance and have nearly the same momentum, which is a desirable property for ensuring that the entire beam is transported to its destination. In a colliding beam accelerator, keeping the emittance small means that the likelihood of particle interactions will be greater resulting in higher luminosity. [3] In a synchrotron light source, low emittance means that the resulting x-ray beam will be small, and result in higher brightness. [4]

Definitions

The coordinate system used to describe the motion of particles in an accelerator has three orthogonal axes, but rather than being centered on a fixed point in space, they are oriented with respect to the trajectory of an "ideal" particle moving through the accelerator with no deviation from the intended speed, position, or direction. Motion along this design trajectory is referred to as the longitudinal axis, and the two axes perpendicular to this trajectory (usually oriented horizontally and vertically) are referred to as transverse axes. The most common convention is for the longitudinal axis to be labelled and the transverse axes to be labelled and . [1] :66–70

Emittance has units of length, but is usually referred to as "length × angle", for example, "millimeter × milliradians". It can be measured in all three spatial dimensions.

Geometric transverse emittance

When a particle moves through a circular accelerator or storage ring, the position and angle of the particle in the x direction will trace an ellipse in phase space. (All of this section applies equivalently to and ) This ellipse can be described by the following equation: [1] :81

where x and x are the position and angle of the particle, and are the Courant–Snyder (Twiss) parameters, calculated from the shape of the ellipse.

The emittance is given by , and has units of length × angle. However, many sources will move the factor of into the units of emittance rather than including the specific value, giving units of "length × angle ×." [2] :335–336

This formula is the single particle emittance, which describes the area enclosed by the trajectory of a single particle in phase space. However, emittance is more useful as a description of the collective properties of the particles in a beam, rather than of a single particle. Since beam particles are not necessarily distributed uniformly in phase space, definitions of emittance for an entire beam will be based on the area of the ellipse required to enclose a specific fraction of the beam particles.

If the beam is distributed in phase space with a Gaussian distribution, the emittance of the beam may be specified in terms of the root mean square value of and the fraction of the beam to be included in the emittance.

The equation for the emittance of a Gaussian beam is: [1] :83

where is the root mean square width of the beam, is the Courant-Snyder , and is the fraction of the beam to be enclosed in the ellipse, given as a number between 0 and 1. Here the factor of is shown on the right of the equation, and would often be included in the units of emittance, rather than being multiplied in to the computed value. [2] :335–336

The value chosen for will depend on the application and the author, and a number of different choices exist in the literature. Some common choices and their equivalent definition of emittance are: [1] :83

0.15
0.39
0.87
0.95

While the x and y axes are generally equivalent mathematically, in horizontal rings where the x coordinate represents the plane of the ring, consideration of dispersion can be added to the equation of the emittance. Because the magnetic force of a bending magnet is dependent on the energy of the particle being bent, particles of different energies will be bent along different trajectories through the magnet, even if their initial position and angle are the same. The effect of this dispersion on the beam emittance is given by:

where is the dispersion at location s, is the ideal particle momentum, and is the root mean square of the momentum difference of the particles in the beam from the ideal momentum. (This definition assumes F=0.15) [1] :91

Longitudinal emittance

The geometrical definition of longitudinal emittance is more complex than that of transverse emittance. While the and coordinates represent deviation from a reference trajectory which remains static, the coordinate represents deviation from a reference particle, which is itself moving with a specified energy. This deviation can be expressed in terms of distance along the reference trajectory, time of flight along the reference trajectory (how "early" or "late" the particle is compared to the reference), or phase (for a specified reference frequency).

In turn, the coordinate is generally not expressed as an angle. Since represents the change in z over time, it corresponds to the forward motion of the particle. This can be given in absolute terms, as a velocity, momentum, or energy, or in relative terms, as a fraction of the position, momentum, or energy of the reference particle. [1] :32

However, the fundamental concept of emittance is the same—the position of the particles are plotted along one axis of a phase space plot, the rate of change in that position over time is plotted on the other axis, and the emittance is a measure of the area occupied on that plot.

One possible definition of longitudinal emittance is given by:

where the integral is taken along a path which tightly encloses the beam particles in phase space. Here is the reference frequency and the longitudinal coordinate is the phase of the particles relative to a reference particle. Longitudinal equations such as this one often must be solved numerically, rather than analytically. [3] :218

RMS emittance

The geometric definition of emittance assumes that the distribution of particles in phase space can be reasonably well characterized by an ellipse. In addition, the definitions using the root mean square of the particle distribution assume a Gaussian particle distribution.

In cases where these assumptions do not hold, it is still possible to define a beam emittance using the moments of the distribution. Here, the RMS emittance () is defined to be, [5]

where is the variance of the particle's position, is the variance of the angle a particle makes with the direction of travel in the accelerator ( with along the direction of travel), and represents an angle-position correlation of particles in the beam. This definition is equivalent to the geometric emittance in the case of an elliptical particle distribution in phase space.

The emittance may also be expressed as the determinant of the variance-covariance matrix of the beam's phase space coordinates where it becomes clear that quantity describes an effective area occupied by the beam in terms of its second order statistics.

Depending on context, some definitions of RMS emittance will add a scaling factor to correspond to a fraction of the total distribution, to facilitate comparison with geometric emittances using the same fraction.

RMS emittance in higher dimensions

It is sometimes useful to talk about phase space area for either four dimensional transverse phase space (IE , , , ) or the full six dimensional phase space of particles (IE , , , , , ). The RMS emittance generalizes to full three dimensional space as shown:

In the absences of correlations between different axes in the particle accelerator, most of these matrix elements become zero and we are left with a product of the emittance along each axis.

Normalized emittance

Although the previous definitions of emittance remain constant for linear beam transport, they do change when the particles undergo acceleration (an effect called adiabatic damping). In some applications, such as for linear accelerators, photoinjectors, and the accelerating sections of larger systems, it becomes important to compare beam quality across different energies. Normalized emittance, which is invariant under acceleration, is used for this purpose.

Normalized emittance in one dimension is given by:

The angle in the prior definition has been replaced with the normalized transverse momentum , where is the Lorentz factor and is the normalized transverse velocity.

Normalized emittance is related to the previous definitions of emittance through and the normalized velocity in the direction of the beam's travel (): [6]

The normalized emittance does not change as a function of energy and so can be used to indicate beam degradation if the particles are accelerated. For speeds close to the speed of light, where is close to one, the emittance is approximately inversely proportional to the energy. In this case, the physical width of the beam will vary inversely with the square root of the energy.

Higher dimensional versions of the normalized emittance can be defined in analogy to the RMS version by replacing all angles with their corresponding momenta.

Measurement of Emittance

Quadrupole Scan Technique

One of the most fundamental methods of measuring beam emittance is the quadrupole scan method. The emittance of the beam for a particular plane of interest (i.e., horizontal or vertical) can be obtained by varying the field strength of a quadrupole (or quadrupoles) upstream of a monitor (i.e., a wire or a screen). [4]

A schematic of the accelerator optics used in the quadrupole scan technique. Quadrupole Scan Beamline.png
A schematic of the accelerator optics used in the quadrupole scan technique.

The properties of a beam can be described as the following beam matrix.

where is the derivative of x with respect to the longitudinal coordinate. The forces experienced by the beam as it travels down the beam line and passes through the quadrupole(s) are described using the transfer matrix (referenced to transfer maps page) of the beam line, including the quadrupole(s) and other beam line components such as drifts:

Here is the transfer matrix between the original beam position and the quadrupole(s), is the transfer matrix of the quadrupole(s), and is the transfer matrix between the quadrupole(s) and the monitor screen. During the quadrupole scan process, and stay constant, and changes with the field strength of the quadrupole(s).

The final beam when it reaches the monitor screen at distance s from its original position can be described as another beam matrix :

The final beam matrix can be calculated from the original beam matrix by doing matrix multiplications with the beam line transfer matrix :

Where is the transpose of .

Now, focusing on the (1,1) element of the final beam matrix throughout the matrix multiplications, we get the equation:

Here the middle term has a factor of 2 because .

Now divide both sides of the above equation by , the equation becomes:

Which is a quadratic equation of the variable . Since the RMS emittance RMS is defined to be the following.

The RMS emittance of the original beam can be calculated using its beam matrix elements:

To obtain the emittance measurement, the following procedure is employed:

  1. For each value (or value combination) of the quadrupole(s), the beam line transfer transfer matrix is calculated to determine values of and .
  2. The beam propagates through the varied beam line, and is observed at the monitor screen, where the beam size is measured.
  3. Repeat step 1 and 2 to obtain a series of values for and , fit the results with a parabola .
  4. Equate parabola fit parameters with original beam matrix elements: , , .
  5. Calculate RMS emittance of the original beam:

If the length of the quadrupole is short compared to its focal length , where is the field strength of the quadrupole, its transfer matrix can be approximated by the thin lens approximation:

Then the RMS emittance can be calculated by fitting a parabola to values of measured beam size versus quadrupole strength .

By adding additional quadrupoles, this technique can be extended to a full 4-D reconstruction. [7]

Mask-Based Reconstruction

A schematic of mask based reconstruction. A charged particle beam is blocked by a grid and the profile is analyzed at a screen to the right. Pepper Pot Mask-Based Emittance Measurement.png
A schematic of mask based reconstruction. A charged particle beam is blocked by a grid and the profile is analyzed at a screen to the right.

Another fundamental method for measuring emittance is by using a predefined mask to imprint a pattern on the beam and sample the remaining beam at a screen downstream.  Two such masks are pepper pots [8] and TEM grids [9] .  A schematic of the TEM grid measurement is shown below.

By using the knowledge of the spacing of the features in the mask one can extract information about the beam size at the mask plane.  By measuring the spacing between the same features on the sampled beam downstream, one can extract information about the angles in the beam.  The quantities of merit can be extracted as described in Marx et al. [10]

The choice of mask is generally dependent on the charge of the beam; low-charge beams are better suited to the TEM grid mask over the pepper pot, as more of the beam is transmitted.

Emittance of electrons versus heavy particles

To understand why the RMS emittance takes on a particular value in a storage ring, one needs to distinguish between electron storage rings and storage rings with heavier particles (such as protons). In an electron storage ring, radiation is an important effect, whereas when other particles are stored, it is typically a small effect. When radiation is important, the particles undergo radiation damping (which slowly decreases emittance turn after turn) and quantum excitation causing diffusion which leads to an equilibrium emittance. [11] When no radiation is present, the emittances remain constant (apart from impedance effects and intrabeam scattering). In this case, the emittance is determined by the initial particle distribution. In particular if one injects a "small" emittance, it remains small, whereas if one injects a "large" emittance, it remains large.

Acceptance

The acceptance, also called admittance, [12] is the maximum emittance that a beam transport system or analyzing system is able to transmit. This is the size of the chamber transformed into phase space and does not suffer from the ambiguities of the definition of beam emittance.

Conservation of emittance

Lenses can focus a beam, reducing its size in one transverse dimension while increasing its angular spread, but cannot change the total emittance. This is a result of Liouville's theorem. Ways of reducing the beam emittance include radiation damping, stochastic cooling, and electron cooling.

Emittance and brightness

Emittance is also related to the brightness of the beam. In microscopy brightness is very often used, because it includes the current in the beam and most systems are circularly symmetric[ clarification needed ]. Consider the brightness of the incident beam at the sample,

where indicates the beam current and represents the total emittance of the incident beam and the wavelength of the incident electron.

The intrinsic emittance , describing a normal distribution in the initial phase space, is diffused by the emittance introduced by aberrations . The total emittance is approximately the sum in quadrature. Under the assumption of uniform illumination of the aperture with current per unit angle , we have the following emittance-brightness relation,


See also

Related Research Articles

In quantum mechanics, identical particles are particles that cannot be distinguished from one another, even in principle. Species of identical particles include, but are not limited to, elementary particles, composite subatomic particles, as well as atoms and molecules. Quasiparticles also behave in this way. Although all known indistinguishable particles only exist at the quantum scale, there is no exhaustive list of all possible sorts of particles nor a clear-cut limit of applicability, as explored in quantum statistics.

<span class="mw-page-title-main">Maxwell–Boltzmann distribution</span> Specific probability distribution function, important in physics

In physics, the Maxwell–Boltzmann distribution, or Maxwell(ian) distribution, is a particular probability distribution named after James Clerk Maxwell and Ludwig Boltzmann.

<span class="mw-page-title-main">Pauli matrices</span> Matrices important in quantum mechanics and the study of spin

In mathematical physics and mathematics, the Pauli matrices are a set of three 2 × 2 complex matrices which are Hermitian, involutory and unitary. Usually indicated by the Greek letter sigma, they are occasionally denoted by tau when used in connection with isospin symmetries.

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

In quantum mechanics, the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the product of the accuracy of certain related pairs of measurements on a quantum system, such as position, x, and momentum, p. Such paired-variables are known as complementary variables or canonically conjugate variables.

<span class="mw-page-title-main">Fermi–Dirac statistics</span> Statistical description for the behavior of fermions

Fermi–Dirac statistics is a type of quantum statistics that applies to the physics of a system consisting of many non-interacting, identical particles that obey the Pauli exclusion principle. A result is the Fermi–Dirac distribution of particles over energy states. It is named after Enrico Fermi and Paul Dirac, each of whom derived the distribution independently in 1926. Fermi–Dirac statistics is a part of the field of statistical mechanics and uses the principles of quantum mechanics.

<span class="mw-page-title-main">Fokker–Planck equation</span> Partial differential equation

In statistical mechanics and information theory, the Fokker–Planck equation is a partial differential equation that describes the time evolution of the probability density function of the velocity of a particle under the influence of drag forces and random forces, as in Brownian motion. The equation can be generalized to other observables as well. The Fokker-Planck equation has multiple applications in information theory, graph theory, data science, finance, economics etc.

<span class="mw-page-title-main">Hooke's law</span> Physical law: force needed to deform a spring scales linearly with distance

In physics, Hooke's law is an empirical law which states that the force needed to extend or compress a spring by some distance scales linearly with respect to that distance—that is, Fs = kx, where k is a constant factor characteristic of the spring, and x is small compared to the total possible deformation of the spring. The law is named after 17th-century British physicist Robert Hooke. He first stated the law in 1676 as a Latin anagram. He published the solution of his anagram in 1678 as: ut tensio, sic vis. Hooke states in the 1678 work that he was aware of the law since 1660.

<span class="mw-page-title-main">Helmholtz free energy</span> Thermodynamic potential

In thermodynamics, the Helmholtz free energy is a thermodynamic potential that measures the useful work obtainable from a closed thermodynamic system at a constant temperature (isothermal). The change in the Helmholtz energy during a process is equal to the maximum amount of work that the system can perform in a thermodynamic process in which temperature is held constant. At constant temperature, the Helmholtz free energy is minimized at equilibrium.

The Ising model, named after the physicists Ernst Ising and Wilhelm Lenz, is a mathematical model of ferromagnetism in statistical mechanics. The model consists of discrete variables that represent magnetic dipole moments of atomic "spins" that can be in one of two states. The spins are arranged in a graph, usually a lattice, allowing each spin to interact with its neighbors. Neighboring spins that agree have a lower energy than those that disagree; the system tends to the lowest energy but heat disturbs this tendency, thus creating the possibility of different structural phases. The model allows the identification of phase transitions as a simplified model of reality. The two-dimensional square-lattice Ising model is one of the simplest statistical models to show a phase transition.

In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a finite-dimensional oriented vector space endowed with a nondegenerate symmetric bilinear form. Applying the operator to an element of the algebra produces the Hodge dual of the element. This map was introduced by W. V. D. Hodge.

In physics, the S-matrix or scattering matrix relates the initial state and the final state of a physical system undergoing a scattering process. It is used in quantum mechanics, scattering theory and quantum field theory (QFT).

<span class="mw-page-title-main">Granular material</span> Conglomeration of discrete solid, macroscopic particles

A granular material is a conglomeration of discrete solid, macroscopic particles characterized by a loss of energy whenever the particles interact. The constituents that compose granular material are large enough such that they are not subject to thermal motion fluctuations. Thus, the lower size limit for grains in granular material is about 1 μm. On the upper size limit, the physics of granular materials may be applied to ice floes where the individual grains are icebergs and to asteroid belts of the Solar System with individual grains being asteroids.

<span class="mw-page-title-main">LSZ reduction formula</span> Connection between correlation functions and the S-matrix

In quantum field theory, the Lehmann–Symanzik–Zimmerman (LSZ) reduction formula is a method to calculate S-matrix elements from the time-ordered correlation functions of a quantum field theory. It is a step of the path that starts from the Lagrangian of some quantum field theory and leads to prediction of measurable quantities. It is named after the three German physicists Harry Lehmann, Kurt Symanzik and Wolfhart Zimmermann.

A multipole expansion is a mathematical series representing a function that depends on angles—usually the two angles used in the spherical coordinate system for three-dimensional Euclidean space, . Similarly to Taylor series, multipole expansions are useful because oftentimes only the first few terms are needed to provide a good approximation of the original function. The function being expanded may be real- or complex-valued and is defined either on , or less often on for some other .

<span class="mw-page-title-main">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

In solid-state physics, the tight-binding model is an approach to the calculation of electronic band structure using an approximate set of wave functions based upon superposition of wave functions for isolated atoms located at each atomic site. The method is closely related to the LCAO method used in chemistry. Tight-binding models are applied to a wide variety of solids. The model gives good qualitative results in many cases and can be combined with other models that give better results where the tight-binding model fails. Though the tight-binding model is a one-electron model, the model also provides a basis for more advanced calculations like the calculation of surface states and application to various kinds of many-body problem and quasiparticle calculations.

Intrabeam scattering (IBS) is an effect in accelerator physics where collisions between particles couple the beam emittance in all three dimensions. This generally causes the beam size to grow. In proton accelerators, intrabeam scattering causes the beam to grow slowly over a period of several hours. This limits the luminosity lifetime. In circular lepton accelerators, intrabeam scattering is counteracted by radiation damping, resulting in a new equilibrium beam emittance with a relaxation time on the order of milliseconds. Intrabeam scattering creates an inverse relationship between the smallness of the beam and the number of particles it contains, therefore limiting luminosity.

The purpose of this page is to provide supplementary materials for the ordinary least squares article, reducing the load of the main article with mathematics and improving its accessibility, while at the same time retaining the completeness of exposition.

<span class="mw-page-title-main">Matrix representation of Maxwell's equations</span>

In electromagnetism, a branch of fundamental physics, the matrix representations of the Maxwell's equations are a formulation of Maxwell's equations using matrices, complex numbers, and vector calculus. These representations are for a homogeneous medium, an approximation in an inhomogeneous medium. A matrix representation for an inhomogeneous medium was presented using a pair of matrix equations. A single equation using 4 × 4 matrices is necessary and sufficient for any homogeneous medium. For an inhomogeneous medium it necessarily requires 8 × 8 matrices.

<span class="mw-page-title-main">Courant–Snyder parameters</span> Set of quantities in accelerator physics

In accelerator physics, the Courant–Snyder parameters are a set of quantities used to describe the distribution of positions and velocities of the particles in a beam. When the positions along a single dimension and velocities along that dimension of every particle in a beam are plotted on a phase space diagram, an ellipse enclosing the particles can be given by the equation:

References

  1. 1 2 3 4 5 6 7 8 Edwards, D. A.; Syphers, M. J. (1993). An introduction to the physics of high energy accelerators. New York: Wiley. ISBN   978-0-471-55163--8.
  2. 1 2 3 Conte, Mario; MacKa, W (2008). An introduction to the physics of particle accelerators (2nd ed.). Hackensack, N.J.: World Scientific. pp. 35–39. ISBN   9789812779601.
  3. 1 2 Wiedemann, Helmut (2007). Particle accelerator physics (3rd ed.). Berlin: Springer. p. 272. ISBN   978-3-540-49043-2.
  4. 1 2 Minty, Michiko G.; Zimm, Frank (2003). Measurement and Control of Charged Particle Beams. Berlin, Heidelberg: Springer Berlin Heidelberg. p. 5. ISBN   3-540-44187-5.
  5. Peggs, Stephen; Satogata, Todd. Introduction to accelerator dynamics. Cambridge, United Kingdom. ISBN   978-1-316-45930-0. OCLC   1000434866.
  6. Wilson, Edmund (2001). An Introduction To Particle Accelerators. Oxford University Press. ISBN   9780198520542.
  7. Prat, Eduard; Aiba, Masamitsu (2014-03-13). "General and efficient dispersion-based measurement of beam slice parameters". Physical Review Special Topics - Accelerators and Beams. 17 (3). doi: 10.1103/physrevstab.17.032801 . hdl: 20.500.11850/81803 . ISSN   1098-4402.
  8. Jackson, G. (1996-07-01). "Fermilab Recycler Ring: Technical design report. Revision 1.1". doi:10.2172/426912.{{cite journal}}: Cite journal requires |journal= (help)
  9. Marx, D.; Giner Navarro, J.; Cesar, D.; Maxson, J.; Marchetti, B.; Assmann, R.; Musumeci, P. (2018-10-15). "Single-shot reconstruction of core 4D phase space of high-brightness electron beams using metal grids". Physical Review Accelerators and Beams. 21 (10). doi: 10.1103/physrevaccelbeams.21.102802 . ISSN   2469-9888. S2CID   126088358.
  10. Marx, D.; Giner Navarro, J.; Cesar, D.; Maxson, J.; Marchetti, B.; Assmann, R.; Musumeci, P. (2018-10-15). "Single-shot reconstruction of core 4D phase space of high-brightness electron beams using metal grids". Physical Review Accelerators and Beams. 21 (10). doi: 10.1103/physrevaccelbeams.21.102802 . ISSN   2469-9888. S2CID   126088358.
  11. http://www.slac.stanford.edu/pubs/slacreports/slac-r-121.html Archived 2015-05-11 at the Wayback Machine The Physics of Electron Storage Rings: An Introduction by Matt Sands
  12. Lee, Shyh-Yuan (1999). Accelerator physics. World Scientific. ISBN   978-9810237097.