Microcanonical ensemble

Last updated

In statistical mechanics, the microcanonical ensemble is a statistical ensemble that represents the possible states of a mechanical system whose total energy is exactly specified. [1] The system is assumed to be isolated in the sense that it cannot exchange energy or particles with its environment, so that (by conservation of energy) the energy of the system does not change with time.

Contents

The primary macroscopic variables of the microcanonical ensemble are the total number of particles in the system (symbol: N), the system's volume (symbol: V), as well as the total energy in the system (symbol: E). Each of these is assumed to be constant in the ensemble. For this reason, the microcanonical ensemble is sometimes called the NVE ensemble.

In simple terms, the microcanonical ensemble is defined by assigning an equal probability to every microstate whose energy falls within a range centered at E. All other microstates are given a probability of zero. Since the probabilities must add up to 1, the probability P is the inverse of the number of microstates W within the range of energy,

The range of energy is then reduced in width until it is infinitesimally narrow, still centered at E. In the limit of this process, the microcanonical ensemble is obtained. [1]

Applicability

Because of its connection with the elementary assumptions of equilibrium statistical mechanics (particularly the postulate of a priori equal probabilities), the microcanonical ensemble is an important conceptual building block in the theory. [2] It is sometimes considered to be the fundamental distribution of equilibrium statistical mechanics. It is also useful in some numerical applications, such as molecular dynamics. [3] [4] On the other hand, most nontrivial systems are mathematically cumbersome to describe in the microcanonical ensemble, and there are also ambiguities regarding the definitions of entropy and temperature. For these reasons, other ensembles are often preferred for theoretical calculations. [2] [5] [6]

The applicability of the microcanonical ensemble to real-world systems depends on the importance of energy fluctuations, which may result from interactions between the system and its environment as well as uncontrolled factors in preparing the system. Generally, fluctuations are negligible if a system is macroscopically large, or if it is manufactured with precisely known energy and thereafter maintained in near isolation from its environment. [7] In such cases the microcanonical ensemble is applicable. Otherwise, different ensembles are more appropriate—such as the canonical ensemble (fluctuating energy) or the grand canonical ensemble (fluctuating energy and particle number).

Properties

Thermodynamic quantities

The fundamental thermodynamic potential of the microcanonical ensemble is entropy. There are at least three possible definitions, each given in terms of the phase volume function v(E), which counts the total number of states with energy less than E (see the Precise expressions section for the mathematical definition of v):

  • the Boltzmann entropy [note 1]
  • the 'volume entropy'
  • the 'surface entropy'

In the microcanonical ensemble, the temperature is a derived quantity rather than an external control parameter. It is defined as the derivative of the chosen entropy with respect to energy. [8] For example, one can define the "temperatures" Tv and Ts as follows:

Like entropy, there are multiple ways to understand temperature in the microcanonical ensemble. More generally, the correspondence between these ensemble-based definitions and their thermodynamic counterparts is not perfect, particularly for finite systems.

The microcanonical pressure and chemical potential are given by: [9]

Phase transitions

Under their strict definition, phase transitions correspond to nonanalytic behavior in the thermodynamic potential or its derivatives. [10] Using this definition, phase transitions in the microcanonical ensemble can occur in systems of any size. This contrasts with the canonical and grand canonical ensembles, for which phase transitions can occur only in the thermodynamic limit—i.e., in systems with infinitely many degrees of freedom. [10] [11] Roughly speaking, the reservoirs defining the canonical or grand canonical ensembles introduce fluctuations that "smooth out" any nonanalytic behavior in the free energy of finite systems. This smoothing effect is usually negligible in macroscopic systems, which are sufficiently large that the free energy can approximate nonanalytic behavior exceedingly well. However, the technical difference in ensembles may be important in the theoretical analysis of small systems. [11]

Information entropy

For a given mechanical system (fixed N, V) and a given range of energy, the uniform distribution of probability P over microstates (as in the microcanonical ensemble) maximizes the ensemble average −⟨log P. [1]

Thermodynamic analogies

Early work in statistical mechanics by Ludwig Boltzmann led to his eponymous entropy equation for a system of a given total energy, S = k log W, where W is the number of distinct states accessible by the system at that energy. Boltzmann did not elaborate too deeply on what exactly constitutes the set of distinct states of a system, besides the special case of an ideal gas. This topic was investigated to completion by Josiah Willard Gibbs who developed the generalized statistical mechanics for arbitrary mechanical systems, and defined the microcanonical ensemble described in this article. [1] Gibbs investigated carefully the analogies between the microcanonical ensemble and thermodynamics, especially how they break down in the case of systems of few degrees of freedom. He introduced two further definitions of microcanonical entropy that do not depend on ω - the volume and surface entropy described above. (Note that the surface entropy differs from the Boltzmann entropy only by an ω-dependent offset.)

The volume entropy Sv and associated Tv form a close analogy to thermodynamic entropy and temperature. It is possible to show exactly that

(P is the ensemble average pressure) as expected for the first law of thermodynamics. A similar equation can be found for the surface (Boltzmann) entropy and its associated Ts, however the "pressure" in this equation is a complicated quantity unrelated to the average pressure. [1]

The microcanonical Tv and Ts are not entirely satisfactory in their analogy to temperature. Outside of the thermodynamic limit, a number of artefacts occur.

The preferred solution to these problems is avoid use of the microcanonical ensemble. In many realistic cases a system is thermostatted to a heat bath so that the energy is not precisely known. Then, a more accurate description is the canonical ensemble or grand canonical ensemble, both of which have complete correspondence to thermodynamics. [14]

Precise expressions for the ensemble

The precise mathematical expression for a statistical ensemble depends on the kind of mechanics under consideration—quantum or classical—since the notion of a "microstate" is considerably different in these two cases. In quantum mechanics, diagonalization provides a discrete set of microstates with specific energies. The classical mechanical case involves instead an integral over canonical phase space, and the size of microstates in phase space can be chosen somewhat arbitrarily.

To construct the microcanonical ensemble, it is necessary in both types of mechanics to first specify a range of energy. In the expressions below the function (a function of H, peaking at E with width ω) will be used to represent the range of energy in which to include states. An example of this function would be [1]

or, more smoothly,

Quantum mechanical

Example of microcanonical ensemble for a quantum system consisting of one particle in a potential well.
Ensemble quantum 1DOF all states.png
Plot of all possible states of this system. The available stationary states displayed as horizontal bars of varying darkness according to |ψi(x)|2.
Ensemble quantum 1DOF microcanonical.png
An ensemble containing only those states within a narrow interval of energy. As the energy width is taken to zero, a microcanonical ensemble is obtained (provided the interval contains at least one state).
The particle's Hamiltonian is Schrödinger-type, Ĥ = U(x) + p2/2m (the potential U(x) is plotted as a red curve). Each panel shows an energy-position plot with the various stationary states, along with a side plot showing the distribution of states in energy.

A statistical ensemble in quantum mechanics is represented by a density matrix, denoted by . The microcanonical ensemble can be written using bra–ket notation, in terms of the system's energy eigenstates and energy eigenvalues. Given a complete basis of energy eigenstates |ψi, indexed by i, the microcanonical ensemble is[ citation needed ]

where the Hi are the energy eigenvalues determined by (here Ĥ is the system's total energy operator, i. e., Hamiltonian operator). The value of W is determined by demanding that is a normalized density matrix, and so

The state volume function (used to calculate entropy) is given by

The microcanonical ensemble is defined by taking the limit of the density matrix as the energy width goes to zero, however a problematic situation occurs once the energy width becomes smaller than the spacing between energy levels. For very small energy width, the ensemble does not exist at all for most values of E, since no states fall within the range. When the ensemble does exist, it typically only contains one (or two) states, since in a complex system the energy levels are only ever equal by accident (see random matrix theory for more discussion on this point). Moreover, the state-volume function also increases only in discrete increments, and so its derivative is only ever infinite or zero, making it difficult to define the density of states. This problem can be solved by not taking the energy range completely to zero and smoothing the state-volume function, however this makes the definition of the ensemble more complicated, since it becomes then necessary to specify the energy range in addition to other variables (together, an NVEω ensemble).

Classical mechanical

Example of microcanonical ensemble for a classical system consisting of one particle in a potential well.
Ensemble classical 1DOF all states.png
Plot of all possible states of this system. The available physical states are evenly distributed in phase space, but with an uneven distribution in energy; the side-plot displays dv/dE.
Ensemble classical 1DOF microcanonical.png
An ensemble restricted to only those states within a narrow interval of energy. This ensemble appears as a thin shell in phase space. As the energy width is taken to zero, a microcanonical ensemble is obtained.
Each panel shows phase space (upper graph) and energy-position space (lower graph). The particle's Hamiltonian is H = U(x) + p2/2m, with the potential U(x) shown as a red curve. The side plot shows the distribution of states in energy.

In classical mechanics, an ensemble is represented by a joint probability density function ρ(p1, … pn, q1, … qn) defined over the system's phase space. [1] The phase space has n generalized coordinates called q1, … qn, and n associated canonical momenta called p1, … pn.

The probability density function for the microcanonical ensemble is:

where

Again, the value of W is determined by demanding that ρ is a normalized probability density function:

This integral is taken over the entire phase space. The state volume function (used to calculate entropy) is defined by

As the energy width ω is taken to zero, the value of W decreases in proportion to ω as W = ω (dv/dE).

Based on the above definition, the microcanonical ensemble can be visualized as an infinitesimally thin shell in phase space, centered on a constant-energy surface. Although the microcanonical ensemble is confined to this surface, it is not necessarily uniformly distributed over that surface: if the gradient of energy in phase space varies, then the microcanonical ensemble is "thicker" (more concentrated) in some parts of the surface than others. This feature is an unavoidable consequence of requiring that the microcanonical ensemble is a steady-state ensemble.

Examples

Ideal gas

The fundamental quantity in the microcanonical ensemble is , which is equal to the phase space volume compatible with given . From , all thermodynamic quantities can be calculated. For an ideal gas, the energy is independent of the particle positions, which therefore contribute a factor of to . The momenta, by contrast, are constrained to a -dimensional (hyper-)spherical shell of radius ; their contribution is equal to the surface volume of this shell. The resulting expression for is: [15]

where is the gamma function, and the factor has been included to account for the indistinguishability of particles (see Gibbs paradox). In the large limit, the Boltzmann entropy is

This is also known as the Sackur–Tetrode equation.

The temperature is given by

which agrees with analogous result from the kinetic theory of gases. Calculating the pressure gives the ideal gas law:

Finally, the chemical potential is

Ideal gas in a uniform gravitational field

The microcanonical phase volume can also be calculated explicitly for an ideal gas in a uniform gravitational field. [16]

The results are stated below for a 3-dimensional ideal gas of particles, each with mass , confined in a thermally isolated container that is infinitely long in the z-direction and has constant cross-sectional area . The gravitational field is assumed to act in the minus z direction with strength . The phase volume is

where is the total energy, kinetic plus gravitational.

The gas density as a function of height can be obtained by integrating over the phase volume coordinates. The result is:

Similarly, the distribution of the velocity magnitude (averaged over all heights) is

The analogues of these equations in the canonical ensemble are the barometric formula and the Maxwell–Boltzmann distribution, respectively. In the limit , the microcanonical and canonical expressions coincide; however, they differ for finite . In particular, in the microcanonical ensemble, the positions and velocities are not statistically independent. As a result, the kinetic temperature, defined as the average kinetic energy in a given volume , is nonuniform throughout the container:

By contrast, the temperature is uniform in the canonical ensemble, for any . [17]

See also

Notes

  1. SB is the information entropy, or Gibbs entropy, for the specific case of the microcanonical ensemble. Note that it depends on the energy width ω.
  2. (Historical note) Gibbs' original ensemble effectively set h = 1 [energy unit]×[time unit], leading to unit-dependence in the values of some thermodynamic quantities like entropy and chemical potential. Since the advent of quantum mechanics, h is often taken to be equal to Planck's constant in order to obtain a semiclassical correspondence with quantum mechanics.
  3. In a system of N identical particles, C = N! (factorial of N). This factor corrects the overcounting in phase space due to identical physical states being found in multiple locations. See the statistical ensemble article for more information on this overcounting.

Related Research Articles

<span class="mw-page-title-main">Boltzmann distribution</span> Probability distribution of energy states of a system

In statistical mechanics and mathematics, a Boltzmann distribution is a probability distribution or probability measure that gives the probability that a system will be in a certain state as a function of that state's energy and the temperature of the system. The distribution is expressed in the form:

<span class="mw-page-title-main">Entropy</span> Property of a thermodynamic system

Entropy is a scientific concept, as well as a measurable physical property, that is most commonly associated with a state of disorder, randomness, or uncertainty. The term and the concept are used in diverse fields, from classical thermodynamics, where it was first recognized, to the microscopic description of nature in statistical physics, and to the principles of information theory. It has found far-ranging applications in chemistry and physics, in biological systems and their relation to life, in cosmology, economics, sociology, weather science, climate change, and information systems including the transmission of information in telecommunication.

In physics, statistical mechanics is a mathematical framework that applies statistical methods and probability theory to large assemblies of microscopic entities. It does not assume or postulate any natural laws, but explains the macroscopic behavior of nature from the behavior of such ensembles.

In physics, specifically statistical mechanics, an ensemble is an idealization consisting of a large number of virtual copies of a system, considered all at once, each of which represents a possible state that the real system might be in. In other words, a statistical ensemble is a set of systems of particles used in statistical mechanics to describe a single system. The concept of an ensemble was introduced by J. Willard Gibbs in 1902.

<span class="mw-page-title-main">Maxwell–Boltzmann statistics</span> Statistical distribution used in many-particle mechanics

In statistical mechanics, Maxwell–Boltzmann statistics describes the distribution of classical material particles over various energy states in thermal equilibrium. It is applicable when the temperature is high enough or the particle density is low enough to render quantum effects negligible.

<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.

In physics, a partition function describes the statistical properties of a system in thermodynamic equilibrium. Partition functions are functions of the thermodynamic state variables, such as the temperature and volume. Most of the aggregate thermodynamic variables of the system, such as the total energy, free energy, entropy, and pressure, can be expressed in terms of the partition function or its derivatives. The partition function is dimensionless.

<span class="mw-page-title-main">Negative temperature</span> Physical systems hotter than any other

Certain systems can achieve negative thermodynamic temperature; that is, their temperature can be expressed as a negative quantity on the Kelvin or Rankine scales. This should be distinguished from temperatures expressed as negative numbers on non-thermodynamic Celsius or Fahrenheit scales, which are nevertheless higher than absolute zero.

In statistical mechanics, the grand canonical ensemble is the statistical ensemble that is used to represent the possible states of a mechanical system of particles that are in thermodynamic equilibrium with a reservoir. The system is said to be open in the sense that the system can exchange energy and particles with a reservoir, so that various possible states of the system can differ in both their total energy and total number of particles. The system's volume, shape, and other external coordinates are kept the same in all possible states of the system.

In statistical mechanics, a canonical ensemble is the statistical ensemble that represents the possible states of a mechanical system in thermal equilibrium with a heat bath at a fixed temperature. The system can exchange energy with the heat bath, so that the states of the system will differ in total energy.

A property of a physical system, such as the entropy of a gas, that stays approximately constant when changes occur slowly is called an adiabatic invariant. By this it is meant that if a system is varied between two end points, as the time for the variation between the end points is increased to infinity, the variation of an adiabatic invariant between the two end points goes to zero.

<span class="mw-page-title-main">Thermodynamic beta</span> Measure of the coldness of a system

In statistical thermodynamics, thermodynamic beta, also known as coldness, is the reciprocal of the thermodynamic temperature of a system:

<span class="mw-page-title-main">Microstate (statistical mechanics)</span> Specific microscopic configuration of a thermodynamic system

In statistical mechanics, a microstate is a specific microscopic configuration of a thermodynamic system that the system may occupy with a certain probability in the course of its thermal fluctuations. In contrast, the macrostate of a system refers to its macroscopic properties, such as its temperature, pressure, volume and density. Treatments on statistical mechanics define a macrostate as follows: a particular set of values of energy, the number of particles, and the volume of an isolated thermodynamic system is said to specify a particular macrostate of it. In this description, microstates appear as different possible ways the system can achieve a particular macrostate.

The mathematical expressions for thermodynamic entropy in the statistical thermodynamics formulation established by Ludwig Boltzmann and J. Willard Gibbs in the 1870s are similar to the information entropy by Claude Shannon and Ralph Hartley, developed in the 1940s.

The concept entropy was first developed by German physicist Rudolf Clausius in the mid-nineteenth century as a thermodynamic property that predicts that certain spontaneous processes are irreversible or impossible. In statistical mechanics, entropy is formulated as a statistical property using probability theory. The statistical entropy perspective was introduced in 1870 by Austrian physicist Ludwig Boltzmann, who established a new field of physics that provided the descriptive linkage between the macroscopic observation of nature and the microscopic view based on the rigorous treatment of a large ensembles of microstates that constitute thermodynamic systems.

The grand potential is a quantity used in statistical mechanics, especially for irreversible processes in open systems. The grand potential is the characteristic state function for the grand canonical ensemble.

<span class="mw-page-title-main">Fundamental thermodynamic relation</span>

In thermodynamics, the fundamental thermodynamic relation are four fundamental equations which demonstrate how four important thermodynamic quantities depend on variables that can be controlled and measured experimentally. Thus, they are essentially equations of state, and using the fundamental equations, experimental data can be used to determine sought-after quantities like G or H (enthalpy). The relation is generally expressed as a microscopic change in internal energy in terms of microscopic changes in entropy, and volume for a closed system in thermal equilibrium in the following way.

<span class="mw-page-title-main">Introduction to entropy</span> Non-technical introduction to entropy

In thermodynamics, entropy is a numerical quantity that shows that many physical processes can go in only one direction in time. For example, you can pour cream into coffee and mix it, but you cannot "unmix" it; you can burn a piece of wood, but you cannot "unburn" it. The word 'entropy' has entered popular usage to refer a lack of order or predictability, or of a gradual decline into disorder. A more physical interpretation of thermodynamic entropy refers to spread of energy or matter, or to extent and diversity of microscopic motion.

<span class="mw-page-title-main">Boltzmann's entropy formula</span> Equation in statistical mechanics

In statistical mechanics, Boltzmann's equation is a probability equation relating the entropy , also written as , of an ideal gas to the multiplicity, the number of real microstates corresponding to the gas's macrostate:

<span class="mw-page-title-main">Thermal fluctuations</span> Random temperature-influenced deviations of particles from their average state

In statistical mechanics, thermal fluctuations are random deviations of a system from its average state, that occur in a system at equilibrium. All thermal fluctuations become larger and more frequent as the temperature increases, and likewise they decrease as temperature approaches absolute zero.

References

  1. 1 2 3 4 5 6 7 8 9 Gibbs, Josiah Willard (1902). Elementary Principles in Statistical Mechanics. New York: Charles Scribner's Sons.
  2. 1 2 Balescu, Radu (1975), Equilibrium and Nonequilibrium Statistical Mechanics, John Wiley & Sons, ISBN   978-0-471-04600-4
  3. Pearson, Eric M.; Halicioglu, Timur; Tiller, William A. (1985). "Laplace-transform technique for deriving thermodynamic equations from the classical microcanonical ensemble". Physical Review A. 32 (5): 3030–3039. Bibcode:1985PhRvA..32.3030P. doi:10.1103/PhysRevA.32.3030. ISSN   0556-2791. PMID   9896445.
  4. Lustig, Rolf (1994). "Statistical thermodynamics in the classical molecular dynamics ensemble. I. Fundamentals". The Journal of Chemical Physics. 100 (4): 3048–3059. Bibcode:1994JChPh.100.3048L. doi:10.1063/1.466446. ISSN   0021-9606.
  5. Hill, Terrell L. (1986). An Introduction to Statistical Thermodynamics. Dover Publications. ISBN   978-0-486-65242-9.
  6. Huang, Kerson (1967). Statistical Mechanics. John Wiley & Sons.
  7. Hilbert, Stefan; Hänggi, Peter; Dunkel, Jörn (2014). "Thermodynamic laws in isolated systems". Physical Review E. 90 (6): 062116. arXiv: 1408.5382 . Bibcode:2014PhRvE..90f2116H. doi:10.1103/PhysRevE.90.062116. hdl: 1721.1/92269 . ISSN   1539-3755. PMID   25615053. S2CID   5365820.
  8. "The Microcanonical Ensemble". chem.libretexts. Retrieved May 3, 2020.
  9. Hill, Terrell L. (1986). An Introduction to Statistical Thermodynamics. Dover Publications. ISBN   978-0-486-65242-9.
  10. 1 2 Nigel Goldenfeld; Lectures on Phase Transitions and the Renormalization Group, Frontiers in Physics 85, Westview Press (June, 1992) ISBN   0-201-55409-7
  11. 1 2 Dunkel, Jörn; Hilbert, Stefan (2006). "Phase transitions in small systems: Microcanonical vs. canonical ensembles". Physica A: Statistical Mechanics and Its Applications. 370 (2): 390–406. arXiv: cond-mat/0511501 . Bibcode:2006PhyA..370..390D. doi:10.1016/j.physa.2006.05.018. ISSN   0378-4371. S2CID   13900006.
  12. Jörn Dunkel; Stefan Hilbert (2013). "Inconsistent thermostatistics and negative absolute temperatures". Nature Physics. 10 (1): 67–72. arXiv: 1304.2066 . Bibcode:2014NatPh..10...67D. doi:10.1038/nphys2815. S2CID   16757018.
  13. See further references at https://sites.google.com/site/entropysurfaceorvolume/
  14. Tolman, R. C. (1938). The Principles of Statistical Mechanics. Oxford University Press.
  15. Kardar, Mehran (2007). Statistical Physics of Particles. Cambridge University Press. pp. 105–109. ISBN   978-0-521-87342-0.
  16. Roman, F L; White, J A; Velasco, S (1995). "Microcanonical single-particle distributions for an ideal gas in a gravitational field". European Journal of Physics. 16 (2): 83–90. Bibcode:1995EJPh...16...83R. doi:10.1088/0143-0807/16/2/008. ISSN   0143-0807. S2CID   250840083.
  17. Velasco, S; Román, F L; White, J A (1996). "On a paradox concerning the temperature distribution of an ideal gas in a gravitational field". European Journal of Physics. 17 (1): 43–44. doi:10.1088/0143-0807/17/1/008. ISSN   0143-0807. S2CID   250885860.