fbpx
Wikipedia

Fermi–Dirac statistics

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.[1][2] Fermi–Dirac statistics is a part of the field of statistical mechanics and uses the principles of quantum mechanics.

Fermi–Dirac statistics applies to identical and indistinguishable particles with half-integer spin (1/2, 3/2, etc.), called fermions, in thermodynamic equilibrium. For the case of negligible interaction between particles, the system can be described in terms of single-particle energy states. A result is the Fermi–Dirac distribution of particles over these states where no two particles can occupy the same state, which has a considerable effect on the properties of the system. Fermi–Dirac statistics is most commonly applied to electrons, a type of fermion with spin 1/2.

A counterpart to Fermi–Dirac statistics is Bose–Einstein statistics, which applies to identical and indistinguishable particles with integer spin (0, 1, 2, etc.) called bosons. In classical physics, Maxwell–Boltzmann statistics is used to describe particles that are identical and treated as distinguishable. For both Bose–Einstein and Maxwell–Boltzmann statistics, more than one particle can occupy the same state, unlike Fermi–Dirac statistics.

Comparison of average occupancy of the ground state for three statistics

History edit

Before the introduction of Fermi–Dirac statistics in 1926, understanding some aspects of electron behavior was difficult due to seemingly contradictory phenomena. For example, the electronic heat capacity of a metal at room temperature seemed to come from 100 times fewer electrons than were in the electric current.[3] It was also difficult to understand why the emission currents generated by applying high electric fields to metals at room temperature were almost independent of temperature.

The difficulty encountered by the Drude model, the electronic theory of metals at that time, was due to considering that electrons were (according to classical statistics theory) all equivalent. In other words, it was believed that each electron contributed to the specific heat an amount on the order of the Boltzmann constant kB. This problem remained unsolved until the development of Fermi–Dirac statistics.

Fermi–Dirac statistics was first published in 1926 by Enrico Fermi[1] and Paul Dirac.[2] According to Max Born, Pascual Jordan developed in 1925 the same statistics, which he called Pauli statistics, but it was not published in a timely manner.[4][5][6] According to Dirac, it was first studied by Fermi, and Dirac called it "Fermi statistics" and the corresponding particles "fermions".[7]

Fermi–Dirac statistics was applied in 1926 by Ralph Fowler to describe the collapse of a star to a white dwarf.[8] In 1927 Arnold Sommerfeld applied it to electrons in metals and developed the free electron model,[9] and in 1928 Fowler and Lothar Nordheim applied it to field electron emission from metals.[10] Fermi–Dirac statistics continues to be an important part of physics.

Fermi–Dirac distribution edit

For a system of identical fermions in thermodynamic equilibrium, the average number of fermions in a single-particle state i is given by the Fermi–Dirac (F–D) distribution,[11][nb 1]

 

where kB is the Boltzmann constant, T is the absolute temperature, εi is the energy of the single-particle state i, and μ is the total chemical potential. The distribution is normalized by the condition

 

that can be used to express   in that   can assume either a positive or negative value.[12]

At zero absolute temperature, μ is equal to the Fermi energy plus the potential energy per fermion, provided it is in a neighbourhood of positive spectral density. In the case of a spectral gap, such as for electrons in a semiconductor, the point of symmetry μ is typically called the Fermi level or—for electrons—the electrochemical potential, and will be located in the middle of the gap.[13][14]

The Fermi–Dirac distribution is only valid if the number of fermions in the system is large enough so that adding one more fermion to the system has negligible effect on μ.[15] Since the Fermi–Dirac distribution was derived using the Pauli exclusion principle, which allows at most one fermion to occupy each possible state, a result is that   .[nb 2]

The variance of the number of particles in state i can be calculated from the above expression for  ,[17][18]

 

Distribution of particles over energy edit

 
Fermi function   with μ = 0.55 eV for various temperatures in the range 50 K ≤ T ≤ 375 K

From the Fermi–Dirac distribution of particles over states, one can find the distribution of particles over energy.[nb 3] The average number of fermions with energy   can be found by multiplying the Fermi–Dirac distribution   by the degeneracy   (i.e. the number of states with energy  ),[19]

 

When  , it is possible that  , since there is more than one state that can be occupied by fermions with the same energy  .

When a quasi-continuum of energies   has an associated density of states   (i.e. the number of states per unit energy range per unit volume[20]), the average number of fermions per unit energy range per unit volume is

 

where   is called the Fermi function and is the same function that is used for the Fermi–Dirac distribution  ,[21]

 

so that

 

Quantum and classical regimes edit

The Fermi–Dirac distribution approaches the Maxwell–Boltzmann distribution in the limit of high temperature and low particle density, without the need for any ad hoc assumptions:

  • In the limit of low particle density,  , therefore   or equivalently  . In that case,  , which is the result from Maxwell-Boltzmann statistics.
  • In the limit of high temperature, the particles are distributed over a large range of energy values, therefore the occupancy on each state (especially the high energy ones with  ) is again very small,  . This again reduces to Maxwell-Boltzmann statistics.

The classical regime, where Maxwell–Boltzmann statistics can be used as an approximation to Fermi–Dirac statistics, is found by considering the situation that is far from the limit imposed by the Heisenberg uncertainty principle for a particle's position and momentum. For example, in physics of semiconductor, when the density of states of conduction band is much higher than the doping concentration, the energy gap between conduction band and fermi level could be calculated using Maxwell-Boltzmann statistics. Otherwise, if the doping concentration is not negligible compared to density of states of conduction band, the Fermi–Dirac distribution should be used instead for accurate calculation. It can then be shown that the classical situation prevails when the concentration of particles corresponds to an average interparticle separation   that is much greater than the average de Broglie wavelength   of the particles:[22]

 

where h is the Planck constant, and m is the mass of a particle.

For the case of conduction electrons in a typical metal at T = 300 K (i.e. approximately room temperature), the system is far from the classical regime because   . This is due to the small mass of the electron and the high concentration (i.e. small  ) of conduction electrons in the metal. Thus Fermi–Dirac statistics is needed for conduction electrons in a typical metal.[22]

Another example of a system that is not in the classical regime is the system that consists of the electrons of a star that has collapsed to a white dwarf. Although the temperature of white dwarf is high (typically T = 10000 K on its surface[23]), its high electron concentration and the small mass of each electron precludes using a classical approximation, and again Fermi–Dirac statistics is required.[8]

Derivations edit

Grand canonical ensemble edit

The Fermi–Dirac distribution, which applies only to a quantum system of non-interacting fermions, is easily derived from the grand canonical ensemble.[24] In this ensemble, the system is able to exchange energy and exchange particles with a reservoir (temperature T and chemical potential μ fixed by the reservoir).

Due to the non-interacting quality, each available single-particle level (with energy level ϵ) forms a separate thermodynamic system in contact with the reservoir. In other words, each single-particle level is a separate, tiny grand canonical ensemble. By the Pauli exclusion principle, there are only two possible microstates for the single-particle level: no particle (energy E = 0), or one particle (energy E = ε). The resulting partition function for that single-particle level therefore has just two terms:

 

and the average particle number for that single-particle level substate is given by

 

This result applies for each single-particle level, and thus gives the Fermi–Dirac distribution for the entire state of the system.[24]

The variance in particle number (due to thermal fluctuations) may also be derived (the particle number has a simple Bernoulli distribution):

 

This quantity is important in transport phenomena such as the Mott relations for electrical conductivity and thermoelectric coefficient for an electron gas,[25] where the ability of an energy level to contribute to transport phenomena is proportional to  .

Canonical ensemble edit

It is also possible to derive Fermi–Dirac statistics in the canonical ensemble. Consider a many-particle system composed of N identical fermions that have negligible mutual interaction and are in thermal equilibrium.[15] Since there is negligible interaction between the fermions, the energy   of a state   of the many-particle system can be expressed as a sum of single-particle energies,

 

where   is called the occupancy number and is the number of particles in the single-particle state   with energy  . The summation is over all possible single-particle states  .

The probability that the many-particle system is in the state  , is given by the normalized canonical distribution,[26]

 

where  , e  is called the Boltzmann factor, and the summation is over all possible states   of the many-particle system.   The average value for an occupancy number   is[26]

 

Note that the state   of the many-particle system can be specified by the particle occupancy of the single-particle states, i.e. by specifying   so that

 

and the equation for   becomes

 

where the summation is over all combinations of values of   which obey the Pauli exclusion principle, and   = 0 or 1 for each  . Furthermore, each combination of values of   satisfies the constraint that the total number of particles is  ,

 

Rearranging the summations,

 

where the   on the summation sign indicates that the sum is not over   and is subject to the constraint that the total number of particles associated with the summation is  . Note that   still depends on   through the   constraint, since in one case   and   is evaluated with   while in the other case   and   is evaluated with    To simplify the notation and to clearly indicate that   still depends on   through   , define

 

so that the previous expression for   can be rewritten and evaluated in terms of the  ,

 

The following approximation[27] will be used to find an expression to substitute for   .

 

where  

If the number of particles   is large enough so that the change in the chemical potential   is very small when a particle is added to the system, then  [28]  Taking the base e antilog[29] of both sides, substituting for  , and rearranging,

 

Substituting the above into the equation for  , and using a previous definition of   to substitute   for  , results in the Fermi–Dirac distribution.

 

Like the Maxwell–Boltzmann distribution and the Bose–Einstein distribution the Fermi–Dirac distribution can also be derived by the Darwin–Fowler method of mean values.[30]

Microcanonical ensemble edit

A result can be achieved by directly analyzing the multiplicities of the system and using Lagrange multipliers.[31]

Suppose we have a number of energy levels, labeled by index i, each level having energy εi  and containing a total of ni  particles. Suppose each level contains gi  distinct sublevels, all of which have the same energy, and which are distinguishable. For example, two particles may have different momenta (i.e. their momenta may be along different directions), in which case they are distinguishable from each other, yet they can still have the same energy. The value of gi  associated with level i is called the "degeneracy" of that energy level. The Pauli exclusion principle states that only one fermion can occupy any such sublevel.

The number of ways of distributing ni indistinguishable particles among the gi sublevels of an energy level, with a maximum of one particle per sublevel, is given by the binomial coefficient, using its combinatorial interpretation

 

For example, distributing two particles in three sublevels will give population numbers of 110, 101, or 011 for a total of three ways which equals 3!/(2!1!).

The number of ways that a set of occupation numbers ni can be realized is the product of the ways that each individual energy level can be populated:

 

Following the same procedure used in deriving the Maxwell–Boltzmann statistics, we wish to find the set of ni for which W is maximized, subject to the constraint that there be a fixed number of particles, and a fixed energy. We constrain our solution using Lagrange multipliers forming the function:

 

Using Stirling's approximation for the factorials, taking the derivative with respect to ni, setting the result to zero, and solving for ni yields the Fermi–Dirac population numbers:

 

By a process similar to that outlined in the Maxwell–Boltzmann statistics article, it can be shown thermodynamically that   and  , so that finally, the probability that a state will be occupied is:

 

See also edit

Notes edit

  1. ^ The F-D Distribution is a type of mathematical function called a logistic function or sigmoid function.
  2. ^ Note that   is also the probability that the state   is occupied, since no more than one fermion can occupy the same state at the same time and  .
  3. ^ These distributions over energies, rather than states, are sometimes called the Fermi–Dirac distribution too, but that terminology will not be used in this article.

References edit

  1. ^ a b Fermi, Enrico (1926). "Sulla quantizzazione del gas perfetto monoatomico". Rendiconti Lincei (in Italian). 3: 145–9., translated as Zannoni, Alberto (1999-12-14). "On the Quantization of the Monoatomic Ideal Gas". arXiv:cond-mat/9912229.
  2. ^ a b Dirac, Paul A. M. (1926). "On the Theory of Quantum Mechanics". Proceedings of the Royal Society A. 112 (762): 661–77. Bibcode:1926RSPSA.112..661D. doi:10.1098/rspa.1926.0133. JSTOR 94692.
  3. ^ (Kittel 1971, pp. 249–50)
  4. ^ . Science-Week. 4 (20). 2000-05-19. OCLC 43626035. Archived from the original on 2009-04-11. Retrieved 2009-01-20.
  5. ^ Schücking (1999). "Jordan, Pauli, Politics, Brecht and a variable gravitational constant". Physics Today. 52 (10): 26. Bibcode:1999PhT....52j..26S. doi:10.1063/1.882858.
  6. ^ Ehlers; Schücking (2002). "Aber Jordan war der Erste". Physik Journal (in German). 1 (11): 71–72. hdl:11858/00-001M-0000-0013-5513-D.
  7. ^ Dirac, Paul A. M. (1967). Principles of Quantum Mechanics (revised 4th ed.). London: Oxford University Press. pp. 210–1. ISBN 978-0-19-852011-5.
  8. ^ a b Fowler, Ralph H. (December 1926). "On dense matter". Monthly Notices of the Royal Astronomical Society. 87 (2): 114–22. Bibcode:1926MNRAS..87..114F. doi:10.1093/mnras/87.2.114.
  9. ^ Sommerfeld, Arnold (1927-10-14). "Zur Elektronentheorie der Metalle" [On Electron Theory of Metals]. Naturwissenschaften (in German). 15 (41): 824–32. Bibcode:1927NW.....15..825S. doi:10.1007/BF01505083. S2CID 39403393.
  10. ^ Fowler, Ralph H.; Nordheim, Lothar W. (1928-05-01). "Electron Emission in Intense Electric Fields". Proceedings of the Royal Society A. 119 (781): 173–81. Bibcode:1928RSPSA.119..173F. doi:10.1098/rspa.1928.0091. JSTOR 95023.
  11. ^ (Reif 1965, p. 341)
  12. ^ Landau, L. D., & Lifshitz, E. M. (2013). Statistical Physics: Volume 5 (Vol. 5). Elsevier.
  13. ^ (Blakemore 2002, p. 11)
  14. ^ Kittel, Charles; Kroemer, Herbert (1980). Thermal Physics (2nd ed.). San Francisco: W. H. Freeman. p. 357. ISBN 978-0-7167-1088-2.
  15. ^ a b (Reif 1965, pp. 340–342)
  16. ^ (Kittel 1971, p. 245, Figs. 4 and 5)
  17. ^ Pearsall, Thomas (2020). Quantum Photonics, 2nd edition. Graduate Texts in Physics. Springer. doi:10.1007/978-3-030-47325-9. ISBN 978-3-030-47324-2.
  18. ^ (Reif 1965, p. 351) Eq. 9.7.7 where  .
  19. ^ Leighton, Robert B. (1959). Principles of Modern Physics. McGraw-Hill. p. 340. ISBN 978-0-07-037130-9. Note that in Eq. (1),   and   correspond respectively to   and   in this article. See also Eq. (32) on p. 339.
  20. ^ (Blakemore 2002, p. 8)
  21. ^ (Reif 1965, p. 389)
  22. ^ a b (Reif 1965, pp. 246–8)
  23. ^ Mukai, Koji; Jim Lochner (1997). . NASA's Imagine the Universe. NASA Goddard Space Flight Center. Archived from the original on 2009-01-18.
  24. ^ a b Srivastava, R. K.; Ashok, J. (2005). "Chapter 6". Statistical Mechanics. New Delhi: PHI Learning Pvt. Ltd. ISBN 9788120327825.
  25. ^ Cutler, M.; Mott, N. (1969). "Observation of Anderson Localization in an Electron Gas". Physical Review. 181 (3): 1336. Bibcode:1969PhRv..181.1336C. doi:10.1103/PhysRev.181.1336.
  26. ^ a b (Reif 1965, pp. 203–6)
  27. ^ See for example, Derivative - Definition via difference quotients, which gives the approximation f(a+h) ≈ f(a) + f '(a) h .
  28. ^ (Reif 1965, pp. 341–2) See Eq. 9.3.17 and Remark concerning the validity of the approximation.
  29. ^ By definition, the base e antilog of A is eA.
  30. ^ Müller-Kirsten, H. J. W. (2013). Basics of Statistical Physics (2nd ed.). World Scientific. ISBN 978-981-4449-53-3.
  31. ^ (Blakemore 2002, pp. 343–5)

Further reading edit

fermi, dirac, statistics, type, quantum, statistics, that, applies, physics, system, consisting, many, interacting, identical, particles, that, obey, pauli, exclusion, principle, result, fermi, dirac, distribution, particles, over, energy, states, named, after. 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 1 2 Fermi Dirac statistics is a part of the field of statistical mechanics and uses the principles of quantum mechanics Fermi Dirac statistics applies to identical and indistinguishable particles with half integer spin 1 2 3 2 etc called fermions in thermodynamic equilibrium For the case of negligible interaction between particles the system can be described in terms of single particle energy states A result is the Fermi Dirac distribution of particles over these states where no two particles can occupy the same state which has a considerable effect on the properties of the system Fermi Dirac statistics is most commonly applied to electrons a type of fermion with spin 1 2 A counterpart to Fermi Dirac statistics is Bose Einstein statistics which applies to identical and indistinguishable particles with integer spin 0 1 2 etc called bosons In classical physics Maxwell Boltzmann statistics is used to describe particles that are identical and treated as distinguishable For both Bose Einstein and Maxwell Boltzmann statistics more than one particle can occupy the same state unlike Fermi Dirac statistics Comparison of average occupancy of the ground state for three statisticsContents 1 History 2 Fermi Dirac distribution 2 1 Distribution of particles over energy 3 Quantum and classical regimes 4 Derivations 4 1 Grand canonical ensemble 4 2 Canonical ensemble 4 3 Microcanonical ensemble 5 See also 6 Notes 7 References 8 Further readingHistory editBefore the introduction of Fermi Dirac statistics in 1926 understanding some aspects of electron behavior was difficult due to seemingly contradictory phenomena For example the electronic heat capacity of a metal at room temperature seemed to come from 100 times fewer electrons than were in the electric current 3 It was also difficult to understand why the emission currents generated by applying high electric fields to metals at room temperature were almost independent of temperature The difficulty encountered by the Drude model the electronic theory of metals at that time was due to considering that electrons were according to classical statistics theory all equivalent In other words it was believed that each electron contributed to the specific heat an amount on the order of the Boltzmann constant kB This problem remained unsolved until the development of Fermi Dirac statistics Fermi Dirac statistics was first published in 1926 by Enrico Fermi 1 and Paul Dirac 2 According to Max Born Pascual Jordan developed in 1925 the same statistics which he called Pauli statistics but it was not published in a timely manner 4 5 6 According to Dirac it was first studied by Fermi and Dirac called it Fermi statistics and the corresponding particles fermions 7 Fermi Dirac statistics was applied in 1926 by Ralph Fowler to describe the collapse of a star to a white dwarf 8 In 1927 Arnold Sommerfeld applied it to electrons in metals and developed the free electron model 9 and in 1928 Fowler and Lothar Nordheim applied it to field electron emission from metals 10 Fermi Dirac statistics continues to be an important part of physics Fermi Dirac distribution editFor a system of identical fermions in thermodynamic equilibrium the average number of fermions in a single particle state i is given by the Fermi Dirac F D distribution 11 nb 1 n i 1e ei m kBT 1 displaystyle bar n i frac 1 e varepsilon i mu k text B T 1 nbsp where kB is the Boltzmann constant T is the absolute temperature ei is the energy of the single particle state i and m is the total chemical potential The distribution is normalized by the condition in i N displaystyle sum i bar n i N nbsp that can be used to express m m T N displaystyle mu mu T N nbsp in that m displaystyle mu nbsp can assume either a positive or negative value 12 At zero absolute temperature m is equal to the Fermi energy plus the potential energy per fermion provided it is in a neighbourhood of positive spectral density In the case of a spectral gap such as for electrons in a semiconductor the point of symmetry m is typically called the Fermi level or for electrons the electrochemical potential and will be located in the middle of the gap 13 14 The Fermi Dirac distribution is only valid if the number of fermions in the system is large enough so that adding one more fermion to the system has negligible effect on m 15 Since the Fermi Dirac distribution was derived using the Pauli exclusion principle which allows at most one fermion to occupy each possible state a result is that 0 lt n i lt 1 displaystyle 0 lt bar n i lt 1 nbsp nb 2 Fermi Dirac distribution nbsp Energy dependence More gradual at higher T n 0 5 displaystyle bar n 0 5 nbsp when e m displaystyle varepsilon mu nbsp Not shown is that m displaystyle mu nbsp decreases for higher T 16 nbsp Temperature dependence for e gt m displaystyle varepsilon gt mu nbsp The variance of the number of particles in state i can be calculated from the above expression for n i displaystyle bar n i nbsp 17 18 V ni kBT mn i n i 1 n i displaystyle V n i k rm B T frac partial partial mu bar n i bar n i 1 bar n i nbsp Distribution of particles over energy edit nbsp Fermi function F e displaystyle F varepsilon nbsp with m 0 55 eV for various temperatures in the range 50 K T 375 KFrom the Fermi Dirac distribution of particles over states one can find the distribution of particles over energy nb 3 The average number of fermions with energy ei displaystyle varepsilon i nbsp can be found by multiplying the Fermi Dirac distribution n i displaystyle bar n i nbsp by the degeneracy gi displaystyle g i nbsp i e the number of states with energy ei displaystyle varepsilon i nbsp 19 n ei gin i gie ei m kBT 1 displaystyle begin aligned bar n varepsilon i amp g i bar n i amp frac g i e varepsilon i mu k rm B T 1 end aligned nbsp When gi 2 displaystyle g i geq 2 nbsp it is possible that n ei gt 1 displaystyle bar n varepsilon i gt 1 nbsp since there is more than one state that can be occupied by fermions with the same energy ei displaystyle varepsilon i nbsp When a quasi continuum of energies e displaystyle varepsilon nbsp has an associated density of states g e displaystyle g varepsilon nbsp i e the number of states per unit energy range per unit volume 20 the average number of fermions per unit energy range per unit volume is N e g e F e displaystyle bar mathcal N varepsilon g varepsilon F varepsilon nbsp where F e displaystyle F varepsilon nbsp is called the Fermi function and is the same function that is used for the Fermi Dirac distribution n i displaystyle bar n i nbsp 21 F e 1e e m kBT 1 displaystyle F varepsilon frac 1 e varepsilon mu k rm B T 1 nbsp so that N e g e e e m kBT 1 displaystyle bar mathcal N varepsilon frac g varepsilon e varepsilon mu k rm B T 1 nbsp Quantum and classical regimes editThe Fermi Dirac distribution approaches the Maxwell Boltzmann distribution in the limit of high temperature and low particle density without the need for any ad hoc assumptions In the limit of low particle density n i 1e ei m kBT 1 1 displaystyle bar n i frac 1 e varepsilon i mu k rm B T 1 ll 1 nbsp therefore e ei m kBT 1 1 displaystyle e varepsilon i mu k rm B T 1 gg 1 nbsp or equivalently e ei m kBT 1 displaystyle e varepsilon i mu k rm B T gg 1 nbsp In that case n i 1e ei m kBT NZe ei kBT displaystyle bar n i approx frac 1 e varepsilon i mu k rm B T frac N Z e varepsilon i k rm B T nbsp which is the result from Maxwell Boltzmann statistics In the limit of high temperature the particles are distributed over a large range of energy values therefore the occupancy on each state especially the high energy ones with ei m kBT displaystyle varepsilon i mu gg k rm B T nbsp is again very small n i 1e ei m kBT 1 1 displaystyle bar n i frac 1 e varepsilon i mu k rm B T 1 ll 1 nbsp This again reduces to Maxwell Boltzmann statistics The classical regime where Maxwell Boltzmann statistics can be used as an approximation to Fermi Dirac statistics is found by considering the situation that is far from the limit imposed by the Heisenberg uncertainty principle for a particle s position and momentum For example in physics of semiconductor when the density of states of conduction band is much higher than the doping concentration the energy gap between conduction band and fermi level could be calculated using Maxwell Boltzmann statistics Otherwise if the doping concentration is not negligible compared to density of states of conduction band the Fermi Dirac distribution should be used instead for accurate calculation It can then be shown that the classical situation prevails when the concentration of particles corresponds to an average interparticle separation R displaystyle bar R nbsp that is much greater than the average de Broglie wavelength l displaystyle bar lambda nbsp of the particles 22 R l h3mkBT displaystyle bar R gg bar lambda approx frac h sqrt 3mk rm B T nbsp where h is the Planck constant and m is the mass of a particle For the case of conduction electrons in a typical metal at T 300 K i e approximately room temperature the system is far from the classical regime because R l 25 displaystyle bar R approx bar lambda 25 nbsp This is due to the small mass of the electron and the high concentration i e small R displaystyle bar R nbsp of conduction electrons in the metal Thus Fermi Dirac statistics is needed for conduction electrons in a typical metal 22 Another example of a system that is not in the classical regime is the system that consists of the electrons of a star that has collapsed to a white dwarf Although the temperature of white dwarf is high typically T 10000 K on its surface 23 its high electron concentration and the small mass of each electron precludes using a classical approximation and again Fermi Dirac statistics is required 8 Derivations editGrand canonical ensemble edit The Fermi Dirac distribution which applies only to a quantum system of non interacting fermions is easily derived from the grand canonical ensemble 24 In this ensemble the system is able to exchange energy and exchange particles with a reservoir temperature T and chemical potential m fixed by the reservoir Due to the non interacting quality each available single particle level with energy level ϵ forms a separate thermodynamic system in contact with the reservoir In other words each single particle level is a separate tiny grand canonical ensemble By the Pauli exclusion principle there are only two possible microstates for the single particle level no particle energy E 0 or one particle energy E e The resulting partition function for that single particle level therefore has just two terms Z exp 0 m e kBT exp 1 m e kBT 1 exp m e kBT displaystyle begin aligned mathcal Z amp exp big 0 mu varepsilon k rm B T big exp big 1 mu varepsilon k rm B T big amp 1 exp big mu varepsilon k rm B T big end aligned nbsp and the average particle number for that single particle level substate is given by N kBT1Z Z m V T 1exp e m kBT 1 displaystyle langle N rangle k rm B T frac 1 mathcal Z left frac partial mathcal Z partial mu right V T frac 1 exp big varepsilon mu k rm B T big 1 nbsp This result applies for each single particle level and thus gives the Fermi Dirac distribution for the entire state of the system 24 The variance in particle number due to thermal fluctuations may also be derived the particle number has a simple Bernoulli distribution DN 2 kBT d N dm V T N 1 N displaystyle big langle Delta N 2 big rangle k rm B T left frac d langle N rangle d mu right V T langle N rangle big 1 langle N rangle big nbsp This quantity is important in transport phenomena such as the Mott relations for electrical conductivity and thermoelectric coefficient for an electron gas 25 where the ability of an energy level to contribute to transport phenomena is proportional to DN 2 displaystyle big langle Delta N 2 big rangle nbsp Canonical ensemble edit It is also possible to derive Fermi Dirac statistics in the canonical ensemble Consider a many particle system composed of N identical fermions that have negligible mutual interaction and are in thermal equilibrium 15 Since there is negligible interaction between the fermions the energy ER displaystyle E R nbsp of a state R displaystyle R nbsp of the many particle system can be expressed as a sum of single particle energies ER rnrer displaystyle E R sum r n r varepsilon r nbsp where nr displaystyle n r nbsp is called the occupancy number and is the number of particles in the single particle state r displaystyle r nbsp with energy er displaystyle varepsilon r nbsp The summation is over all possible single particle states r displaystyle r nbsp The probability that the many particle system is in the state R displaystyle R nbsp is given by the normalized canonical distribution 26 PR e bER R e bER displaystyle P R frac e beta E R displaystyle sum R e beta E R nbsp where b 1 kBT displaystyle beta 1 k rm B T nbsp e bER displaystyle scriptstyle beta E R nbsp is called the Boltzmann factor and the summation is over all possible states R displaystyle R nbsp of the many particle system The average value for an occupancy number ni displaystyle n i nbsp is 26 n i Rni PR displaystyle bar n i sum R n i P R nbsp Note that the state R displaystyle R nbsp of the many particle system can be specified by the particle occupancy of the single particle states i e by specifying n1 n2 displaystyle n 1 n 2 ldots nbsp so that PR Pn1 n2 e b n1e1 n2e2 n1 n2 e b n1 e1 n2 e2 displaystyle P R P n 1 n 2 ldots frac e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots displaystyle sum n 1 n 2 ldots e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots nbsp and the equation for n i displaystyle bar n i nbsp becomes n i n1 n2 ni Pn1 n2 n1 n2 ni e b n1e1 n2e2 niei n1 n2 e b n1e1 n2e2 niei displaystyle begin alignedat 2 bar n i amp sum n 1 n 2 dots n i P n 1 n 2 dots amp frac displaystyle sum n 1 n 2 dots n i e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots n i varepsilon i cdots displaystyle sum n 1 n 2 dots e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots n i varepsilon i cdots end alignedat nbsp where the summation is over all combinations of values of n1 n2 displaystyle n 1 n 2 ldots nbsp which obey the Pauli exclusion principle and nr displaystyle n r nbsp 0 or 1 for each r displaystyle r nbsp Furthermore each combination of values of n1 n2 displaystyle n 1 n 2 ldots nbsp satisfies the constraint that the total number of particles is N displaystyle N nbsp rnr N displaystyle sum r n r N nbsp Rearranging the summations n i ni 01ni e b niei i n1 n2 e b n1e1 n2e2 ni 01e b niei i n1 n2 e b n1e1 n2e2 displaystyle bar n i frac displaystyle sum n i 0 1 n i e beta n i varepsilon i quad sideset i sum n 1 n 2 dots e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots displaystyle sum n i 0 1 e beta n i varepsilon i qquad sideset i sum n 1 n 2 dots e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots nbsp where the i displaystyle i nbsp on the summation sign indicates that the sum is not over ni displaystyle n i nbsp and is subject to the constraint that the total number of particles associated with the summation is Ni N ni displaystyle N i N n i nbsp Note that S i displaystyle Sigma i nbsp still depends on ni displaystyle n i nbsp through the Ni displaystyle N i nbsp constraint since in one case ni 0 displaystyle n i 0 nbsp and S i displaystyle Sigma i nbsp is evaluated with Ni N displaystyle N i N nbsp while in the other case ni 1 displaystyle n i 1 nbsp and S i displaystyle Sigma i nbsp is evaluated with Ni N 1 displaystyle N i N 1 nbsp To simplify the notation and to clearly indicate that S i displaystyle Sigma i nbsp still depends on ni displaystyle n i nbsp through N ni displaystyle N n i nbsp define Zi N ni i n1 n2 e b n1e1 n2e2 displaystyle Z i N n i equiv sideset i sum n 1 n 2 ldots e beta n 1 varepsilon 1 n 2 varepsilon 2 cdots nbsp so that the previous expression for n i displaystyle bar n i nbsp can be rewritten and evaluated in terms of the Zi displaystyle Z i nbsp n i ni 01ni e b niei Zi N ni ni 01e b niei Zi N ni 0 e beiZi N 1 Zi N e beiZi N 1 1 Zi N Zi N 1 ebei 1 displaystyle begin alignedat 3 bar n i amp frac displaystyle sum n i 0 1 n i e beta n i varepsilon i Z i N n i displaystyle sum n i 0 1 e beta n i varepsilon i qquad Z i N n i 8pt amp frac quad 0 quad e beta varepsilon i Z i N 1 Z i N e beta varepsilon i Z i N 1 6pt amp frac 1 Z i N Z i N 1 e beta varepsilon i 1 quad end alignedat nbsp The following approximation 27 will be used to find an expression to substitute for Zi N Zi N 1 displaystyle Z i N Z i N 1 nbsp ln Zi N 1 ln Zi N ln Zi N N ln Zi N ai displaystyle begin alignedat 2 ln Z i N 1 amp simeq ln Z i N frac partial ln Z i N partial N amp ln Z i N alpha i end alignedat nbsp where ai ln Zi N N displaystyle alpha i equiv frac partial ln Z i N partial N nbsp If the number of particles N displaystyle N nbsp is large enough so that the change in the chemical potential m displaystyle mu nbsp is very small when a particle is added to the system then ai m kBT displaystyle alpha i simeq mu k rm B T nbsp 28 Taking the base e antilog 29 of both sides substituting for ai displaystyle alpha i nbsp and rearranging Zi N Zi N 1 e m kBT displaystyle Z i N Z i N 1 e mu k rm B T nbsp Substituting the above into the equation for n i displaystyle bar n i nbsp and using a previous definition of b displaystyle beta nbsp to substitute 1 kBT displaystyle 1 k rm B T nbsp for b displaystyle beta nbsp results in the Fermi Dirac distribution n i 1e ei m kBT 1 displaystyle bar n i frac 1 e varepsilon i mu k rm B T 1 nbsp Like the Maxwell Boltzmann distribution and the Bose Einstein distribution the Fermi Dirac distribution can also be derived by the Darwin Fowler method of mean values 30 Microcanonical ensemble edit A result can be achieved by directly analyzing the multiplicities of the system and using Lagrange multipliers 31 Suppose we have a number of energy levels labeled by index i each level having energy ei and containing a total of ni particles Suppose each level contains gi distinct sublevels all of which have the same energy and which are distinguishable For example two particles may have different momenta i e their momenta may be along different directions in which case they are distinguishable from each other yet they can still have the same energy The value of gi associated with level i is called the degeneracy of that energy level The Pauli exclusion principle states that only one fermion can occupy any such sublevel The number of ways of distributing ni indistinguishable particles among the gisublevels of an energy level with a maximum of one particle per sublevel is given by the binomial coefficient using its combinatorial interpretation w ni gi gi ni gi ni displaystyle w n i g i frac g i n i g i n i nbsp For example distributing two particles in three sublevels will give population numbers of 110 101 or 011 for a total of three ways which equals 3 2 1 The number of ways that a set of occupation numbers ni can be realized is the product of the ways that each individual energy level can be populated W iw ni gi igi ni gi ni displaystyle W prod i w n i g i prod i frac g i n i g i n i nbsp Following the same procedure used in deriving the Maxwell Boltzmann statistics we wish to find the set of ni for which W is maximized subject to the constraint that there be a fixed number of particles and a fixed energy We constrain our solution using Lagrange multipliers forming the function f ni ln W a N ni b E niei displaystyle f n i ln W alpha left N sum n i right beta left E sum n i varepsilon i right nbsp Using Stirling s approximation for the factorials taking the derivative with respect to ni setting the result to zero and solving for ni yields the Fermi Dirac population numbers ni giea bei 1 displaystyle n i frac g i e alpha beta varepsilon i 1 nbsp By a process similar to that outlined in the Maxwell Boltzmann statistics article it can be shown thermodynamically that b 1kBT textstyle beta frac 1 k rm B T nbsp and a mkBT textstyle alpha frac mu k rm B T nbsp so that finally the probability that a state will be occupied is n i nigi 1e ei m kBT 1 displaystyle bar n i frac n i g i frac 1 e varepsilon i mu k rm B T 1 nbsp See also edit nbsp Wikimedia Commons has media related to Fermi Dirac distribution Grand canonical ensemble Pauli exclusion principle Complete Fermi Dirac integral Fermi level Fermi gas Maxwell Boltzmann statistics Bose Einstein statistics Parastatistics Logistic functionNotes edit The F D Distribution is a type of mathematical function called a logistic function or sigmoid function Note that n i displaystyle bar n i nbsp is also the probability that the state i displaystyle i nbsp is occupied since no more than one fermion can occupy the same state at the same time and 0 lt n i lt 1 displaystyle 0 lt bar n i lt 1 nbsp These distributions over energies rather than states are sometimes called the Fermi Dirac distribution too but that terminology will not be used in this article References edit a b Fermi Enrico 1926 Sulla quantizzazione del gas perfetto monoatomico Rendiconti Lincei in Italian 3 145 9 translated as Zannoni Alberto 1999 12 14 On the Quantization of the Monoatomic Ideal Gas arXiv cond mat 9912229 a b Dirac Paul A M 1926 On the Theory of Quantum Mechanics Proceedings of the Royal Society A 112 762 661 77 Bibcode 1926RSPSA 112 661D doi 10 1098 rspa 1926 0133 JSTOR 94692 Kittel 1971 pp 249 50 History of Science The Puzzle of the Bohr Heisenberg Copenhagen Meeting Science Week 4 20 2000 05 19 OCLC 43626035 Archived from the original on 2009 04 11 Retrieved 2009 01 20 Schucking 1999 Jordan Pauli Politics Brecht and a variable gravitational constant Physics Today 52 10 26 Bibcode 1999PhT 52j 26S doi 10 1063 1 882858 Ehlers Schucking 2002 Aber Jordan war der Erste Physik Journal in German 1 11 71 72 hdl 11858 00 001M 0000 0013 5513 D Dirac Paul A M 1967 Principles of Quantum Mechanics revised 4th ed London Oxford University Press pp 210 1 ISBN 978 0 19 852011 5 a b Fowler Ralph H December 1926 On dense matter Monthly Notices of the Royal Astronomical Society 87 2 114 22 Bibcode 1926MNRAS 87 114F doi 10 1093 mnras 87 2 114 Sommerfeld Arnold 1927 10 14 Zur Elektronentheorie der Metalle On Electron Theory of Metals Naturwissenschaften in German 15 41 824 32 Bibcode 1927NW 15 825S doi 10 1007 BF01505083 S2CID 39403393 Fowler Ralph H Nordheim Lothar W 1928 05 01 Electron Emission in Intense Electric Fields Proceedings of the Royal Society A 119 781 173 81 Bibcode 1928RSPSA 119 173F doi 10 1098 rspa 1928 0091 JSTOR 95023 Reif 1965 p 341 Landau L D amp Lifshitz E M 2013 Statistical Physics Volume 5 Vol 5 Elsevier Blakemore 2002 p 11 Kittel Charles Kroemer Herbert 1980 Thermal Physics 2nd ed San Francisco W H Freeman p 357 ISBN 978 0 7167 1088 2 a b Reif 1965 pp 340 342 Kittel 1971 p 245 Figs 4 and 5 Pearsall Thomas 2020 Quantum Photonics 2nd edition Graduate Texts in Physics Springer doi 10 1007 978 3 030 47325 9 ISBN 978 3 030 47324 2 Reif 1965 p 351 Eq 9 7 7 where b 1 kBT a m kBT n i ϵi n i m displaystyle beta 1 k rm B T quad alpha mu k rm B T quad frac partial bar n i partial epsilon i frac partial bar n i partial mu nbsp Leighton Robert B 1959 Principles of Modern Physics McGraw Hill p 340 ISBN 978 0 07 037130 9 Note that in Eq 1 n e displaystyle n varepsilon nbsp and ns displaystyle n s nbsp correspond respectively to n i displaystyle bar n i nbsp and n ei displaystyle bar n varepsilon i nbsp in this article See also Eq 32 on p 339 Blakemore 2002 p 8 Reif 1965 p 389 a b Reif 1965 pp 246 8 Mukai Koji Jim Lochner 1997 Ask an Astrophysicist NASA s Imagine the Universe NASA Goddard Space Flight Center Archived from the original on 2009 01 18 a b Srivastava R K Ashok J 2005 Chapter 6 Statistical Mechanics New Delhi PHI Learning Pvt Ltd ISBN 9788120327825 Cutler M Mott N 1969 Observation of Anderson Localization in an Electron Gas Physical Review 181 3 1336 Bibcode 1969PhRv 181 1336C doi 10 1103 PhysRev 181 1336 a b Reif 1965 pp 203 6 See for example Derivative Definition via difference quotients which gives the approximation f a h f a f a h Reif 1965 pp 341 2 See Eq 9 3 17 and Remark concerning the validity of the approximation By definition the base e antilog of A is eA Muller Kirsten H J W 2013 Basics of Statistical Physics 2nd ed World Scientific ISBN 978 981 4449 53 3 Blakemore 2002 pp 343 5 Further reading editReif F 1965 Fundamentals of Statistical and Thermal Physics McGraw Hill ISBN 978 0 07 051800 1 Blakemore J S 2002 Semiconductor Statistics Dover ISBN 978 0 486 49502 6 Kittel Charles 1971 Introduction to Solid State Physics 4th ed New York John Wiley amp Sons ISBN 978 0 471 14286 7 OCLC 300039591 Retrieved from https en wikipedia org w index php title Fermi Dirac statistics amp oldid 1215552319, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.