Lipid bilayer: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Monkbot
Line 1: Line 1:
{{Statistical mechanics|cTopic=[[Particle statistics|Particle Statistics]]}}
Claude is her title and she totally digs that name. Arizona has always been my residing location but my spouse wants us to transfer. The favorite hobby for my children and me is playing crochet and now I'm attempting to earn money with it. She is currently a cashier but soon she'll be on her own.<br><br>Check out my page: [http://Www.Light4America.com/profile.php?u=DaO84 Www.Light4America.com]
 
In [[quantum statistics]], a branch of [[physics]], '''Fermi–Dirac statistics''' describes a distribution of particles in certain [[physical system|systems]] comprising many [[identical particles]] that obey the [[Pauli exclusion principle]]. It is named after [[Enrico Fermi]] and [[Paul Dirac]], who each discovered it independently, although Enrico Fermi defined the statistics earlier than Paul Dirac.<ref name='Fermi1926' /><ref name='Dirac1926'/>
 
Fermi–Dirac (F–D) statistics applies to identical particles with [[half-integer]] [[spin (physics)|spin]] in a system in [[thermodynamic equilibrium]]. Additionally, the particles in this system are assumed to have negligible mutual [[interaction]]. This allows the many-particle system to be described in terms of single-particle [[Energy eigenstates|energy states]]. The result is the F–D distribution of particles over these states and includes the condition that no two particles can occupy the same state, which has a considerable effect on the properties of the system. Since F–D statistics applies to particles with half-integer spin, these particles have come to be called [[fermion]]s. It is most commonly applied to [[electron]]s, which are fermions with [[spin&nbsp;1/2]]. Fermi–Dirac statistics is a part of the more general field of [[statistical mechanics]] and uses the principles of [[quantum mechanics]].
 
==History==
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 [[electron]]s than were in the [[electric current]].<ref name='Kittel1971Cel249'>{{harv|Kittel|1971|pp=249–50}}</ref>  It was also difficult to understand why the [[field electron emission|emission currents]], generated by applying high electric fields to metals at room temperature, were almost independent of temperature.
 
The difficulty encountered by 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]]&nbsp;''k''.
This statistical problem remained unsolved until the discovery of F–D statistics.
 
F–D statistics was first published in 1926 by [[Enrico Fermi]]<ref name='Fermi1926'>{{cite journal| title=Sulla quantizzazione del gas perfetto monoatomico| journal=Rendiconti Lincei| language=Italian| year=1926| first=Enrico| last=Fermi| volume=3| issue=| pages=145–9}}, translated as {{cite arxiv| title=On the Quantization of the Monoatomic Ideal Gas| eprint=cond-mat/9912229 | date=1999-12-14| last1=Zannoni |first1=Alberto (transl.) | class=cond-mat.stat-mech }}</ref> and [[Paul Dirac]].<ref name='Dirac1926'>{{cite journal| title=On the Theory of Quantum Mechanics| journal=Proceedings of the Royal Society, Series A| year=1926| first=Paul A. M.| last=Dirac| authorlink=Paul Dirac| coauthors=| volume=112| issue=762| pages=661–77| jstor=94692|accessdate=| doi=10.1098/rspa.1926.0133 |bibcode = 1926RSPSA.112..661D }}</ref>  According to an account, [[Pascual Jordan]] developed in 1925 the same statistics which he called ''[[Wolfgang Pauli|Pauli]] statistics'', but it was not published in a timely manner.<ref name='Science-Week2000'>{{cite journal| title=History of Science: The Puzzle of the Bohr–Heisenberg Copenhagen Meeting| journal=[[ScienceWeek|Science-Week]]| date=2000-05-19| first=|last=| coauthors=| volume=4| issue=20| pages=| url=http://scienceweek.com/2000/sw000519.htm | oclc=43626035| location=Chicago| accessdate=2009-01-20 }}</ref>  According to Dirac, it was first studied by Fermi, and Dirac called it Fermi statistics and the corresponding particles fermions.<ref name='Dirac1958'>{{cite book | last=Dirac| first=Paul A. M.| authorlink=Paul Dirac| title=Principles of Quantum Mechanics| edition=revised 4th| publisher=Oxford University Press| year=1967| location=London| pages=210–1| url=http://books.google.com/?id=XehUpGiM6FIC&pg=PA210| isbn=978-0-19-852011-5 }}</ref>
 
F–D statistics was applied in 1926 by [[Ralph Fowler|Fowler]] to describe the collapse of a [[star]] to a [[white dwarf]].<ref name='Fowler1926'>{{cite journal| title=On dense matter| journal=Monthly Notices of the Royal Astronomical Society|date=December 1926| first=Ralph H.| last=Fowler| authorlink=Ralph Fowler| volume=87| pages=114–22| bibcode=1926MNRAS..87..114F }}</ref> In 1927 [[Arnold Sommerfeld|Sommerfeld]] applied it to electrons in metals<ref name='sommerfeld1927'>{{cite journal| title=Zur Elektronentheorie der Metalle|language=German|trans_title=On Electron Theory of Metals| journal=[[Naturwissenschaften]]| date=1927-10-14| first=Arnold| last=Sommerfeld| authorlink=Arnold Sommerfeld| volume=15| issue=41| pages=824–32| doi=10.1007/BF01505083|bibcode = 1927NW.....15..825S }}</ref> and in 1928 [[Ralph Fowler|Fowler]] and [[Lothar Wolfgang Nordheim|Nordheim]] applied it to [[field electron emission]] from metals.<ref name='Fowler1928'>{{cite journal| title=Electron Emission in Intense Electric Fields| journal=Proceedings of the Royal Society A| date=1928-05-01| first=Ralph H.| last=Fowler| first2=Lothar W.| last2=Nordheim| volume=119| issue=781| pages=173–81| doi=10.1098/rspa.1928.0091| jstor=95023| url=http://rspa.royalsocietypublishing.org/content/119/781/173.full.pdf| format=PDF| author-link=Lothar Wolfgang Nordheim |bibcode = 1928RSPSA.119..173F }}</ref> Fermi–Dirac statistics continues to be an important part of physics.
 
==Fermi–Dirac distribution==
For a system of identical fermions, the average number of fermions in a single-particle state&nbsp;<math>i</math>, is given by the Fermi–Dirac (F–D) distribution,<ref name='Reif1965dist341'>{{harv|Reif|1965|p=341}}</ref>
<br>
:<math> \bar{n}_i = \frac{1}{e^{(\epsilon_i-\mu) / k T} + 1} </math>
where  ''k'' is [[Boltzmann's constant]], ''T'' is the absolute [[temperature]], <math>\epsilon_i \ </math>  is the energy of the single-particle state <math>i</math>,  and ''μ'' is the [[total chemical potential]]. At zero temperature, ''μ'' is equal to the [[Fermi energy]] plus the potential energy per electron.  For the case of electrons in a semiconductor, <math>\mu\ </math> is typically called the [[Fermi level]] or [[electrochemical potential]].<ref name=Blakemore2002p11>{{harv|Blakemore|2002|p=11}}</ref><ref name=KittelKroemer1980>{{cite book | last=Kittel| first=Charles| authorlink1=Charles Kittel| first2=Herbert| last2=Kroemer| title=Thermal Physics| edition=2nd| publisher=W. H. Freeman| year=1980| location=San Francisco| pages=357| url=http://books.google.com/?id=c0R79nyOoNMC&pg=PA357| isbn=978-0-7167-1088-2| author-link2=Herbert Kroemer }}</ref>
 
The F–D 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  <math>\mu\ </math>.<ref name='Reif1965dist'>{{harv|Reif|1965|pp=340–2}}</ref> Since the F–D distribution was derived using the [[Pauli exclusion principle]], which allows at most one electron to occupy each possible state, a result is that <math>0 < \bar{n}_i  < 1</math>  .<ref>Note that <math> \bar{n}_i </math> is also the probability that the state <math>i</math> is occupied, since no more than one fermion can occupy the same state at the same time and <math>0 < \bar{n}_i  < 1</math>.</ref>
 
<center> <gallery Caption="Fermi–Dirac distribution" widths="400px" heights="300px">
Image:FD e mu.svg|'''Energy dependence.''' More gradual at higher ''T''.  {{nowrap|1=<math> \bar{n}</math> = 0.5}}  when {{nowrap|1=<math> \epsilon \;</math> = <math>\mu \; </math>.}} Not shown is that <math>\mu \ </math> decreases for higher ''T''.<ref name='Kittel1971dist245'>{{harv|Kittel|1971|p=245, Figs. 4 and 5}}</ref> <br> <center>
Image:FD kT e.svg|<center>'''Temperature dependence''' for <math> \epsilon > \mu \ </math> . </center>
</gallery><small>(Click on a figure to enlarge.)</small></center>
 
===Distribution of particles over energy===
[[File:fermi.gif|thumb|300px|right|Fermi function F(<math>\epsilon \ </math>) vs. energy <math>\epsilon \ </math>, with μ&nbsp;=&nbsp;0.55&nbsp;eV and for various temperatures in the range 50K&nbsp;≤&nbsp;T&nbsp;≤&nbsp;375K.]]
 
The above Fermi–Dirac distribution gives the distribution of identical fermions over single-particle energy states, where no more than one fermion can occupy a state. Using the F–D distribution, one can find the distribution of identical fermions over energy, where more than one fermion can have the same energy.<ref>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.</ref>
 
The average number of fermions with energy <math>\epsilon_i \ </math>  can be found by multiplying the F–D distribution <math> \bar{n}_i  \  </math> by the [[degenerate energy level|degeneracy]]  <math> g_i \ </math> (i.e. the number of states with energy <math>\epsilon_i \ </math> ),<ref name='Leighton1959'>{{cite book | last=Leighton| first=Robert B.| authorlink=Robert B. Leighton| coauthors = | title=Principles of Modern Physics| publisher=McGraw-Hill| year=1959 | pages=340| isbn=978-0-07-037130-9 }}<br>
Note that in Eq. (1), <math>n(\epsilon) \,</math> and <math> n_s \,</math> correspond respectively to <math>\bar{n}_i </math> and <math> \bar{n}(\epsilon_i)  </math> in this article.  See also Eq. (32) on p. 339.</ref>
 
:<math> \begin{alignat}{2}
\bar{n}(\epsilon_i) & = g_i \  \bar{n}_i \\
      & = \frac{g_i}{e^{(\epsilon_i-\mu) / k T} + 1} \\
\end{alignat} </math>
<br>
When <math> g_i  \ge 2 \ </math>, it is possible that <math>\ \bar{n}(\epsilon_i) > 1 </math>  since there is more than one state that can be occupied by fermions with the same energy  <math> \epsilon_i \ </math>.
 
When a quasi-continuum of energies <math> \epsilon \ </math> has an associated [[density of states]] <math> g( \epsilon ) \ </math> (i.e. the number of states per unit energy range per unit volume<ref name=Blakemore2002p8>{{harv|Blakemore|2002|p=8}}</ref>) the average number of fermions per unit energy range per unit volume is,
 
:<math>  \bar { \mathcal{N} }(\epsilon)  = g(\epsilon) \  F(\epsilon) </math>
<br>
where <math>F(\epsilon) \ </math> is called the Fermi function and is the same [[function (mathematics)|function]] that is used for the F–D distribution <math>  \bar{n}_i  </math>,<ref name='Reif1965FermiFnc'>{{harv|Reif|1965|p=389}}</ref>
 
:<math> F(\epsilon) = \frac{1}{e^{(\epsilon-\mu) / k T} + 1} </math>
so that,
:<math>  \bar { \mathcal{N} }(\epsilon) = \frac{g(\epsilon)}{e^{(\epsilon-\mu) / k T} + 1} </math> .
 
==Quantum and classical regimes==
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]]. Using this approach, it can be shown that the classical situation occurs if the concentration of particles corresponds to an average interparticle separation <math> \bar{R} </math> that is much greater than the average [[de Broglie wavelength]] <math> \bar{\lambda} </math> of the particles,<ref name='Reif1965classical'>{{harv|Reif|1965|pp=246–8}}</ref>
:<math>\bar{R} \ \gg \ \bar{\lambda} \ \approx \ \frac{h}{\sqrt{3mkT}} </math>
 
where <big><math>h</math></big> is [[Planck's constant]], and  <big><math>m</math></big> is the [[mass]] of a particle.
 
For the case of conduction electrons in a typical metal at ''T'' = 300[[Kelvin|K]] (i.e. approximately room temperature), the system is far from the classical regime because <math> \bar{R} \approx \bar{\lambda}/25 </math> . This is due to the small mass of the electron and the high concentration (i.e. small <math>\bar{R}</math>) of conduction electrons in the metal. Thus Fermi–Dirac statistics is needed for conduction electrons in a typical metal.<ref name='Reif1965classical' />
 
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 white dwarf's temperature is high (typically ''T'' = 10,000K on its surface<ref name='Mukai1997'>{{cite web|url=http://imagine.gsfc.nasa.gov/docs/ask_astro/answers/970512e.html |title=Ask an Astrophysicist |accessdate= |last=Mukai |first=Koji |coauthors=Jim Lochner |year=1997 |work=NASA's Imagine the Universe |publisher=NASA Goddard Space Flight Center |archiveurl=http://www.webcitation.org/5dxyeNTKh |archivedate=2009-01-20 }}</ref>), its high electron concentration and the small mass of each electron precludes using a classical approximation, and again Fermi–Dirac statistics is required.<ref name='Fowler1926' />
 
==Three derivations of the Fermi–Dirac distribution==
 
=== Derivation starting with grand canonical ensemble ===
 
The Fermi-Dirac distribution, which applies only to a quantum system of non-interacting fermions, is easily derived from the [[grand canonical ensemble]].<ref name="sriva">Chapter 6 of {{cite isbn|9788120327825}}</ref> 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 [[microstate (statistical mechanics)|microstate]]s 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:
:<math> \begin{align}\mathcal Z & = \exp(0(\mu - 0)/k_B T) + \exp(1(\mu - \epsilon)/k_B T) \\ & = 1 + \exp((\mu - \epsilon)/k_B T)\end{align}</math>
and the average particle number for that single-particle substate is given by
:<math> \langle N\rangle = k_B T \frac{1}{\mathcal Z} \left(\frac{\partial \mathcal Z}{\partial \mu}\right)_{V,T} = \frac{1}{\exp((\epsilon-\mu)/k_B T)+1} </math>
This result applies for each single-particle level, and thus gives the Fermi-Dirac distribution for the entire state of the system.<ref name="sriva"/>
 
The variance in particle number (due to [[thermal fluctuations]]) may also be derived (the particle number has a simple [[Bernoulli distribution]]):
:<math> \langle (\Delta N)^2 \rangle = k_B T \left(\frac{d\langle N\rangle}{d\mu}\right)_{V,T} = \langle N\rangle (1 - \langle N\rangle) </math>
This quantity is important in transport phenomena such as the Mott relations for electrical conductivity and [[Thermoelectric_effect#Charge carrier_diffusion|thermoelectric coefficient]] for an electron gas,<ref>{{cite doi|10.1103/PhysRev.181.1336}}</ref> where the ability of an energy level to contribute to transport phenomena is proportional to <math>\langle (\Delta N)^2 \rangle</math>.
 
===Derivations starting with canonical distribution===
 
It is also possible to derive Fermi–Dirac statistics in the [[canonical ensemble]].
 
====Standard derivation====
Consider a many-particle system composed of ''N'' identical fermions that have negligible mutual interaction and are in thermal equilibrium.<ref name='Reif1965dist'>{{harv|Reif|1965|pp=340–2}}</ref> Since there is negligible interaction between the fermions, the energy <math>E_R</math> of a state <math>R</math> of the many-particle system can be expressed as a sum of single-particle energies,
 
:<math> E_R = \sum_{r} n_r \epsilon_r \; </math>
 
where <math> n_r </math> is called the occupancy number and is the number of particles in the single-particle state <math>r</math> with energy <math>\epsilon_r \;</math>. The summation is over all possible single-particle states <math>r</math>.
 
The probability that the many-particle system is in the state <math>R</math>, is given by the normalized [[canonical distribution]],<ref name='Reif1965canonical'>{{harv|Reif|1965|pp=203–6}}</ref>
 
:<math>P_R = \frac { e^{-\beta E_R} }
{ \displaystyle \sum_{R'} e^{-\beta E_{R'}} } </math>
 
where <math> \beta\;</math><math> = 1/kT</math>, &nbsp; <math>k</math>&nbsp;is [[Boltzmann's constant]], <math>T</math> is the absolute [[temperature]], ''e''<sup><math>\scriptstyle -\beta E_R</math></sup> is called the [[Boltzmann factor]], and the summation is over all possible states <math>R'</math> of the many-particle system. &nbsp; The average value for an occupancy number <math>n_i \;</math> is<ref name='Reif1965canonical'/>
 
:<math>\bar{n}_i \ = \  \sum_{R} n_i \ P_R  </math>
 
Note that the state <math>R</math> of the many-particle system can be specified by the particle occupancy of the single-particle states, i.e. by specifying <math>n_1,\, n_2,\, ... \;,</math> so that
 
:<math>P_R = P_{n_1,n_2,...} = \frac{ e^{-\beta (n_1 \epsilon_1+n_2 \epsilon_2+...)} }
                                                                                    {\displaystyle \sum_{{n_1}',{n_2}',...} e^{-\beta ({n_1}' \epsilon_1+{n_2}' \epsilon_2+...)} } </math>
 
and the equation for <math>\bar{n}_i</math> becomes
 
:<math>\begin{alignat} {2}
 
\bar{n}_i & = \sum_{n_1,n_2,\dots} n_i \ P_{n_1,n_2,\dots} \\
                \\
                & = \frac{\displaystyle \sum_{n_1,n_2,\dots} n_i \ e^{-\beta (n_1\epsilon_1 + n_2\epsilon_2 + \cdots + n_i\epsilon_i + \cdots)} }
                                {\displaystyle \sum_{n_1,n_2,\dots}  e^{-\beta (n_1\epsilon_1 + n_2\epsilon_2 + \cdots + n_i\epsilon_i + \cdots)} } \\
 
                        \end{alignat} </math>
 
where the summation is over all combinations of values of <math>n_1,n_2,...\;</math>&nbsp;  which obey the Pauli exclusion principle, and {{nowrap|1=<math> n_r </math> = 0 or 1}} for each <math>r</math>. Furthermore, each combination of values of <math>n_1,n_2,...\;</math> satisfies the constraint that the total number of particles is <math>N</math>,
 
:<math> \sum_{r} n_r = N \;</math> .
 
Rearranging the summations,
 
:<math> \bar{n}_i = \frac
 
{\displaystyle \sum_{n_i=0} ^1 n_i \ e^{-\beta (n_i\epsilon_i)} \quad  \sideset{ }{^{(i)}}\sum_{n_1,n_2,\dots} e^{-\beta (n_1\epsilon_1+n_2\epsilon_2+\cdots)} }
 
{\displaystyle \sum_{n_i=0} ^1 e^{-\beta (n_i\epsilon_i)} \qquad \sideset{ }{^{(i)}}\sum_{n_1,n_2,\dots} e^{-\beta (n_1\epsilon_1+n_2\epsilon_2+\cdots)} } </math>
 
where the &nbsp;<math>^{(i)}</math>  on the summation sign indicates that the sum is not over <math>n_i</math> and is subject to the constraint that the total number of particles associated with the summation is <math>N_i =  N-n_i</math>&nbsp;. Note that  <math>\Sigma^{(i)}</math> still depends on <math>n_i</math> through the <math>N_i  </math> constraint, since in one case <math>n_i=0 </math> and  <math>\Sigma^{(i)}</math> is evaluated with <math>N_i=N ,</math> while in the other case <math>n_i=1 </math> and  <math>\Sigma^{(i)}</math> is evaluated with <math>N_i=N-1  .</math> &nbsp;To simplify the notation and to clearly indicate that <math>\Sigma^{(i)}</math> still depends on <math>n_i</math> through <math>N-n_i</math>&nbsp;, define
 
:<math> Z_i(N-n_i) \equiv \ \sideset{ }{^{(i)}}\sum_{n_1,n_2,...} e^{-\beta (n_1\epsilon_1+n_2\epsilon_2+\cdots)} \;</math>
 
so that the previous expression for <math>\bar{n}_i</math>  can be rewritten and evaluated in terms of the <math>Z_i</math>,
 
:<math> \begin{alignat} {3}
\bar{n}_i \ & =  \frac{ \displaystyle \sum_{n_i=0} ^1  n_i \ e^{-\beta (n_i\epsilon_i)}  \ \  Z_i(N-n_i)}
                                                                                        { \displaystyle  \sum_{n_i=0} ^1 e^{-\beta (n_i\epsilon_i)} \qquad    Z_i(N-n_i)} \\
\\
& = \ \frac { \quad 0 \quad \; +  e^{-\beta\epsilon_i}\; Z_i(N-1)} {Z_i(N) + e^{-\beta\epsilon_i}\; Z_i(N-1)}  \\
& = \ \frac {1} {[Z_i(N)/Z_i(N-1)] \; e^{\beta\epsilon_i}+1} \quad .
\end{alignat} </math>
 
The following approximation<ref>See for example,  [[Derivative#Definition via difference quotients|Derivative - Definition via difference quotients]], which gives the approximation <big>''f(a+h) ≈ f(a) + f '(a) h''</big> .</ref> will be used to find an expression to substitute for <math> Z_i(N)/Z_i(N-1)</math> .
 
:<math>\begin{alignat} {2}
\ln Z_i(N- 1) & \simeq \ln Z_i(N) - \frac {\partial \ln Z_i(N)} {\partial N } \\
& = \ln Z_i(N) - \alpha_i \;
\end{alignat}</math>
where &nbsp; &nbsp; &nbsp;<math>\alpha_i \equiv \frac {\partial \ln  Z_i(N)} {\partial N} \ . </math>
 
If the number of particles <math>N</math> is large enough so that the change in the chemical potential <math>\mu\;</math> is very small when a particle is added to the system,  then <math>\alpha_i \simeq - \mu / kT \ .</math><ref name='Reif1965ChemPot'>{{harv|Reif|1965|pp=341–2}} See Eq. 9.3.17 and ''Remark concerning the validity of the approximation''.</ref>&nbsp;  Taking the base ''e'' antilog<ref>By definition, the base ''e'' antilog of ''A'' is ''e<sup>A</sup>''.</ref> of both sides, substituting for <math>\alpha_i \,</math>, and rearranging,
 
:<math>Z_i(N) / Z_i(N- 1) =  e^{-\mu / kT } \,</math> .
 
Substituting the above into the equation for <math>\bar {n}_i</math>, and using a previous definition of <math>\beta\;</math> to substitute  <math>1/kT </math>  for  <math>\beta\;</math>, results in the Fermi–Dirac distribution.
 
:<div style="{{divstylewhite}}; width:15em; margin:.3em"><center><math>\bar{n}_i = \ \frac {1} {e^{(\epsilon_i - \mu)/kT }+1}</math></center></div>
 
<br> <!-- extra linespace to not crowd equation with next section -->
 
====Derivation using Lagrange multipliers====
 
A result can be achieved by directly analyzing the multiplicities of the system and using [[Lagrange multipliers]].<ref name="Blakemore2002p343–5">{{harv|Blakemore|2002|pp=343–5}}</ref>
 
Suppose we have a number of energy levels, labeled by index ''i'', each level
having energy ε''<sub>i</sub>''&nbsp; and containing a total of ''n<sub>i</sub>''&nbsp; particles.  Suppose each level contains ''g<sub>i</sub>''&nbsp; 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 ''g<sub>i</sub>''&nbsp; 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 ''n<sub>i</sub>'' indistinguishable particles among the ''g<sub>i</sub>  '' sublevels of an energy level, with a maximum of one particle per sublevel, is given by the [[binomial coefficient]], using its  [[Binomial coefficient#Combinatorial interpretation|combinatorial interpretation]]
:<math>
w(n_i,g_i)=\frac{g_i!}{n_i!(g_i-n_i)!} \  .
</math>
 
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 ''n''<sub>''i''</sub> can be realized is the product of the ways that each individual energy level can be populated:
 
:<math>
W = \prod_i w(n_i,g_i) =  \prod_i \frac{g_i!}{n_i!(g_i-n_i)!}.
</math>
 
Following the same procedure used in deriving the [[Maxwell–Boltzmann statistics]],
we wish to find the set of ''n<sub>i</sub>'' 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:
 
:<math>
f(n_i)=\ln(W)+\alpha(N-\sum n_i)+\beta(E-\sum n_i \epsilon_i).
</math>
 
Using [[Stirling's approximation]] for the factorials, taking the derivative with respect to ''n<sub>i</sub>'', setting the result to zero, and solving for ''n<sub>i</sub>'' yields the Fermi–Dirac population numbers:
 
:<math>
n_i = \frac{g_i}{e^{\alpha+\beta \epsilon_i}+1}.
</math>
 
By a process similar to that outlined in the [[Maxwell-Boltzmann statistics]] article, it can be shown thermodynamically that <math>\beta = \frac{1}{kT}</math> and <math>\alpha = - \frac{\mu}{kT}</math> where <math>\mu</math> is the [[chemical potential]], ''k'' is [[Boltzmann's constant]] and ''T'' is the [[temperature]], so that finally, the probability that a state will be occupied is:
 
:<math>
\bar{n}_i = \frac{n_i}{g_i} = \frac{1}{e^{(\epsilon_i-\mu)/kT}+1}.
</math>
 
==See also==
{{commons category|Fermi-Dirac distribution}}
*[[Grand canonical ensemble]]
*[[Fermi level]]
*[[Maxwell–Boltzmann statistics]]
*[[Bose–Einstein statistics]]
*[[Parastatistics]]
 
==References==
#{{cite book | last = Reif | first = F. | coauthors = | title = Fundamentals of Statistical and Thermal Physics| publisher = McGraw–Hill | year = 1965 | ref=harv| pages =  | url = | doi = | isbn =978-0-07-051800-1}}
#{{cite book | last = Blakemore | first = J. S. | coauthors = | title = Semiconductor Statistics | publisher = Dover | year = 2002 | ref=harv| page =  | url = http://books.google.com/?id=cc4HE2YM1FIC&pg=front| doi = | isbn = 978-0-486-49502-6 }}
#{{cite book | last = Kittel | first=Charles | authorlink = Charles Kittel | title = Introduction to Solid State Physics| edition=4th| publisher=John Wiley & Sons| year=1971| ref=harv|location=New York| oclc=300039591 | isbn = 0-471-14286-7 }}
 
==Footnotes==
{{reflist|2}}
 
{{Statistical mechanics topics}}
<!-- Editors: Please do not add the probability distributions template here. The Fermi Dirac distribution is not a probability distribution. -->
 
{{DEFAULTSORT:Fermi-Dirac statistics}}
[[Category:Fermi–Dirac statistics| ]]
[[Category:Concepts in physics]]
[[Category:Quantum field theory]]
[[Category:Statistical mechanics]]

Revision as of 10:59, 27 February 2014

Claude is her title and she totally digs that name. Arizona has always been my residing location but my spouse wants us to transfer. The favorite hobby for my children and me is playing crochet and now I'm attempting to earn money with it. She is currently a cashier but soon she'll be on her own.

Check out my page: Www.Light4America.com