fbpx
Wikipedia

Bose–Einstein statistics

In quantum statistics, Bose–Einstein statistics (B–E statistics) describes one of two possible ways in which a collection of non-interacting, indistinguishable particles may occupy a set of available discrete energy states at thermodynamic equilibrium. The aggregation of particles in the same state, which is a characteristic of particles obeying Bose–Einstein statistics, accounts for the cohesive streaming of laser light and the frictionless creeping of superfluid helium. The theory of this behaviour was developed (1924–25) by Satyendra Nath Bose, who recognized that a collection of identical and indistinguishable particles can be distributed in this way. The idea was later adopted and extended by Albert Einstein in collaboration with Bose.

The Bose–Einstein statistics applies only to the particles not limited to single occupancy of the same state – that is, particles that do not obey the Pauli exclusion principle restrictions. Such particles have integer values of spin and are named bosons. Particles with half-integer spins are called fermions and obey Fermi-Dirac statistics.

Comparison of average occupancy of the ground state for three statistics

Bose–Einstein distribution

At low temperatures, bosons behave differently from fermions (which obey the Fermi–Dirac statistics) in a way that an unlimited number of them can "condense" into the same energy state. This apparently unusual property also gives rise to the special state of matter – the Bose–Einstein condensate. Fermi–Dirac and Bose–Einstein statistics apply when quantum effects are important and the particles are "indistinguishable". Quantum effects appear if the concentration of particles satisfies

 

where N is the number of particles, V is the volume, and nq is the quantum concentration, for which the interparticle distance is equal to the thermal de Broglie wavelength, so that the wavefunctions of the particles are barely overlapping.

Fermi–Dirac statistics applies to fermions (particles that obey the Pauli exclusion principle), and Bose–Einstein statistics applies to bosons. As the quantum concentration depends on temperature, most systems at high temperatures obey the classical (Maxwell–Boltzmann) limit, unless they also have a very high density, as for a white dwarf. Both Fermi–Dirac and Bose–Einstein become Maxwell–Boltzmann statistics at high temperature or at low concentration.

B–E statistics was introduced for photons in 1924 by Bose and generalized to atoms by Einstein in 1924–25.

The expected number of particles in an energy state i for B–E statistics is:

 

with εi > μ and where ni is the occupation number (the number of particles) in state i,   is the degeneracy of energy level i, εi is the energy of the i-th state, μ is the chemical potential, kB is the Boltzmann constant, and T is absolute temperature.

The variance of this distribution   is calculated directly from the expression above for the average number.[1]

 

For comparison, the average number of fermions with energy   given by Fermi–Dirac particle-energy distribution has a similar form:

 

As mentioned above, both the Bose–Einstein distribution and 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.

In addition to reducing to the Maxwell–Boltzmann distribution in the limit of high   and low density, B–E statistics also reduces to Rayleigh–Jeans law distribution for low energy states with
 , namely

 

History

Władysław Natanson in 1911 concluded that Planck's law requires indistinguishability of "units of energy", although he did not frame this in terms of Einstein's light quanta.[2][3]

While presenting a lecture at the University of Dhaka (in what was then British India and is now Bangladesh) on the theory of radiation and the ultraviolet catastrophe, Satyendra Nath Bose intended to show his students that the contemporary theory was inadequate, because it predicted results not in accordance with experimental results. During this lecture, Bose committed an error in applying the theory, which unexpectedly gave a prediction that agreed with the experiment. The error was a simple mistake—similar to arguing that flipping two fair coins will produce two heads one-third of the time—that would appear obviously wrong to anyone with a basic understanding of statistics (remarkably, this error resembled the famous blunder by d'Alembert known from his Croix ou Pile article[4][5]). However, the results it predicted agreed with experiment, and Bose realized it might not be a mistake after all. For the first time, he took the position that the Maxwell–Boltzmann distribution would not be true for all microscopic particles at all scales. Thus, he studied the probability of finding particles in various states in phase space, where each state is a little patch having phase volume of h3, and the position and momentum of the particles are not kept particularly separate but are considered as one variable.

Bose adapted this lecture into a short article called "Planck's law and the hypothesis of light quanta"[6][7] and submitted it to the Philosophical Magazine. However, the referee's report was negative, and the paper was rejected. Undaunted, he sent the manuscript to Albert Einstein requesting publication in the Zeitschrift für Physik. Einstein immediately agreed, personally translated the article from English into German (Bose had earlier translated Einstein's article on the general theory of relativity from German to English), and saw to it that it was published. Bose's theory achieved respect when Einstein sent his own paper in support of Bose's to Zeitschrift für Physik, asking that they be published together. The paper came out in 1924.[8]

The reason Bose produced accurate results was that since photons are indistinguishable from each other, one cannot treat any two photons having equal quantum numbers (e.g., polarization and momentum vector) as being two distinct identifiable photons. By analogy, if in an alternate universe coins were to behave like photons and other bosons, the probability of producing two heads would indeed be one-third, and so is the probability of getting a head and a tail which equals one-half for the conventional (classical, distinguishable) coins. Bose's "error" leads to what is now called Bose–Einstein statistics.

Bose and Einstein extended the idea to atoms and this led to the prediction of the existence of phenomena which became known as Bose–Einstein condensate, a dense collection of bosons (which are particles with integer spin, named after Bose), which was demonstrated to exist by experiment in 1995.

Derivation

Derivation from the microcanonical ensemble

In the microcanonical ensemble, one considers a system with fixed energy, volume, and number of particles. We take a system composed of   identical bosons,   of which have energy   and are distributed over   levels or states with the same energy  , i.e.   is the degeneracy associated with energy   of total energy  . Calculation of the number of arrangements of   particles distributed among   states is a problem of combinatorics. Since particles are indistinguishable in the quantum mechanical context here, the number of ways for arranging   particles in   boxes (for the  th energy level) would be (see image).

 
The image represents one possible distribution of bosonic particles in different boxes. The box partitions (green) can be moved around to change the size of the boxes and as a result of the number of bosons each box can contain.
 

where   is the k-combination of a set with m elements. The total number of arrangements in an ensemble of bosons is simply the product of the binomial coefficients   above over all the energy levels, i.e.

 

The maximum number of arrangements determining the corresponding occupation number   is obtained by maximizing the entropy, or equivalently, setting   and taking the subsidiary conditions   into account (as Lagrange multipliers).[9] The result for  ,  ,   is the Bose–Einstein distribution.

Derivation from the grand canonical ensemble

The Bose–Einstein distribution, which applies only to a quantum system of non-interacting bosons, is naturally derived from the grand canonical ensemble without any approximations.[10] 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. That is, the number of particles within the overall system that occupy a given single particle state form a sub-ensemble that is also grand canonical ensemble; hence, it may be analysed through the construction of a grand partition function.

Every single-particle state is of a fixed energy,  . As the sub-ensemble associated with a single-particle state varies by the number of particles only, it is clear that the total energy of the sub-ensemble is also directly proportional to the number of particles in the single-particle state; where   is the number of particles, the total energy of the sub-ensemble will then be  . Beginning with the standard expression for a grand partition function and replacing   with  , the grand partition function takes the form

 

This formula applies to fermionic systems as well as bosonic systems. Fermi–Dirac statistics arises when considering the effect of the Pauli exclusion principle: whilst the number of fermions occupying the same single-particle state can only be either 1 or 0, the number of bosons occupying a single particle state may be any integer. Thus, the grand partition function for bosons can be considered a geometric series and may be evaluated as such:

 

Note that the geometric series is convergent only if  , including the case where  . This implies that the chemical potential for the Bose gas must be negative, i.e.,  , whereas the Fermi gas is allowed to take both positive and negative values for the chemical potential.[11]

The average particle number for that single-particle substate is given by

 
This result applies for each single-particle level and thus forms the Bose–Einstein distribution for the entire state of the system.[12][13]

The variance in particle number,  , is:

 

As a result, for highly occupied states the standard deviation of the particle number of an energy level is very large, slightly larger than the particle number itself:  . This large uncertainty is due to the fact that the probability distribution for the number of bosons in a given energy level is a geometric distribution; somewhat counterintuitively, the most probable value for N is always 0. (In contrast, classical particles have instead a Poisson distribution in particle number for a given state, with a much smaller uncertainty of  , and with the most-probable N value being near  .)

Derivation in the canonical approach

It is also possible to derive approximate Bose–Einstein statistics in the canonical ensemble. These derivations are lengthy and only yield the above results in the asymptotic limit of a large number of particles. The reason is that the total number of bosons is fixed in the canonical ensemble. The Bose–Einstein distribution in this case can be derived as in most texts by maximization, but the mathematically best derivation is by the Darwin–Fowler method of mean values as emphasized by Dingle.[14] See also Müller-Kirsten.[9] The fluctuations of the ground state in the condensed region are however markedly different in the canonical and grand-canonical ensembles.[15]

Derivation

Suppose we have a number of energy levels, labeled by index  , each level having energy   and containing a total of   particles. Suppose each level contains   distinct sublevels, all of which have the same energy, and which are distinguishable. For example, two particles may have different momenta, in which case they are distinguishable from each other, yet they can still have the same energy. The value of   associated with level   is called the "degeneracy" of that energy level. Any number of bosons can occupy the same sublevel.

Let   be the number of ways of distributing   particles among the   sublevels of an energy level. There is only one way of distributing   particles with one sublevel, therefore  . It is easy to see that there are   ways of distributing   particles in two sublevels which we will write as:

 

With a little thought (see Notes below) it can be seen that the number of ways of distributing   particles in three sublevels is

 
so that
 

where we have used the following theorem involving binomial coefficients:

 

Continuing this process, we can see that   is just a binomial coefficient (See Notes below)

 

For example, the population numbers for two particles in three sublevels are 200, 110, 101, 020, 011, or 002 for a total of six which equals 4!/(2!2!). The number of ways that a set of occupation numbers   can be realized is the product of the ways that each individual energy level can be populated:

 

where the approximation assumes that  .

Following the same procedure used in deriving the Maxwell–Boltzmann statistics, we wish to find the set of   for which W is maximised, subject to the constraint that there be a fixed total number of particles, and a fixed total energy. The maxima of   and   occur at the same value of   and, since it is easier to accomplish mathematically, we will maximise the latter function instead. We constrain our solution using Lagrange multipliers forming the function:

 

Using the   approximation and using Stirling's approximation for the factorials   gives

 

Where K is the sum of a number of terms which are not functions of the  . Taking the derivative with respect to  , and setting the result to zero and solving for  , yields the Bose–Einstein population numbers:

 

By a process similar to that outlined in the Maxwell–Boltzmann statistics article, it can be seen that:

 

which, using Boltzmann's famous relationship   becomes a statement of the second law of thermodynamics at constant volume, and it follows that   and   where S is the entropy,   is the chemical potential, kB is Boltzmann's constant and T is the temperature, so that finally:

 

Note that the above formula is sometimes written:

 

where   is the absolute activity, as noted by McQuarrie.[16]

Also note that when the particle numbers are not conserved, removing the conservation of particle numbers constraint is equivalent to setting   and therefore the chemical potential   to zero. This will be the case for photons and massive particles in mutual equilibrium and the resulting distribution will be the Planck distribution.

Notes

A much simpler way to think of Bose–Einstein distribution function is to consider that n particles are denoted by identical balls and g shells are marked by g-1 line partitions. It is clear that the permutations of these n balls and g − 1 partitions will give different ways of arranging bosons in different energy levels. Say, for 3 (= n) particles and 3 (= g) shells, therefore (g − 1) = 2, the arrangement might be |●●|●, or ||●●●, or |●|●● , etc. Hence the number of distinct permutations of n + (g − 1) objects which have n identical items and (g − 1) identical items will be:

 
See the image for a visual representation of one such distribution of n particles in g boxes that can be represented as g − 1 partitions.
 
The image represents one possible distribution of bosonic particles in different boxes. The box partitions (green) can be moved around to change the size of the boxes and as a result of the number of bosons each box can contain.

OR

The purpose of these notes is to clarify some aspects of the derivation of the Bose–Einstein (B–E) distribution for beginners. The enumeration of cases (or ways) in the B–E distribution can be recast as follows. Consider a game of dice throwing in which there are   dice, with each die taking values in the set , for  . The constraints of the game are that the value of a die  , denoted by  , has to be greater than or equal to the value of die  , denoted by  , in the previous throw, i.e.,  . Thus a valid sequence of die throws can be described by an n-tuple  , such that  . Let   denote the set of these valid n-tuples:

 

 

 

 

 

(1)

Then the quantity   (defined above as the number of ways to distribute   particles among the   sublevels of an energy level) is the cardinality of  , i.e., the number of elements (or valid n-tuples) in  . Thus the problem of finding an expression for   becomes the problem of counting the elements in  .

Example n = 4, g = 3:

 
 
(there are   elements in  )

Subset   is obtained by fixing all indices   to  , except for the last index,  , which is incremented from   to  . Subset   is obtained by fixing  , and incrementing   from   to  . Due to the constraint   on the indices in  , the index   must automatically take values in  . The construction of subsets   and   follows in the same manner.

Each element of   can be thought of as a multiset of cardinality  ; the elements of such multiset are taken from the set   of cardinality  , and the number of such multisets is the multiset coefficient

 

More generally, each element of   is a multiset of cardinality   (number of dice) with elements taken from the set   of cardinality   (number of possible values of each die), and the number of such multisets, i.e.,  is the multiset coefficient

 

 

 

 

 

(2)

which is exactly the same as the formula for  , as derived above with the aid of a theorem involving binomial coefficients, namely

 

 

 

 

 

(3)

To understand the decomposition

 

 

 

 

 

(4)

or for example,   and  

 

let us rearrange the elements of   as follows

 

Clearly, the subset   of   is the same as the set

 

By deleting the index   (shown in red with double underline) in the subset   of  , one obtains the set

 

In other words, there is a one-to-one correspondence between the subset   of   and the set  . We write

 

Similarly, it is easy to see that

 
 
 

Thus we can write

 

or more generally,

 

 

 

 

 

(5)

and since the sets

 

are non-intersecting, we thus have

 

 

 

 

 

(6)

with the convention that

 

 

 

 

 

(7)

Continuing the process, we arrive at the following formula

 
Using the convention (7)2 above, we obtain the formula
 

 

 

 

 

(8)

keeping in mind that for   and   being constants, we have

 

 

 

 

 

(9)

It can then be verified that (8) and (2) give the same result for  ,  ,  , etc.

Interdisciplinary applications

Viewed as a pure probability distribution, the Bose–Einstein distribution has found application in other fields:

  • In recent years, Bose–Einstein statistics has also been used as a method for term weighting in information retrieval. The method is one of a collection of DFR ("Divergence From Randomness") models,[17] the basic notion being that Bose–Einstein statistics may be a useful indicator in cases where a particular term and a particular document have a significant relationship that would not have occurred purely by chance. Source code for implementing this model is available from the Terrier project at the University of Glasgow.
  • The evolution of many complex systems, including the World Wide Web, business, and citation networks, is encoded in the dynamic web describing the interactions between the system's constituents. Despite their irreversible and nonequilibrium nature these networks follow Bose statistics and can undergo Bose–Einstein condensation. Addressing the dynamical properties of these nonequilibrium systems within the framework of equilibrium quantum gases predicts that the "first-mover-advantage", "fit-get-rich" (FGR) and "winner-takes-all" phenomena observed in competitive systems are thermodynamically distinct phases of the underlying evolving networks.[18]

See also

Notes

  1. ^ 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.
  2. ^ Jammer, Max (1966). The conceptual development of quantum mechanics. McGraw-Hill. p. 51. ISBN 0-88318-617-9.
  3. ^ Passon, Oliver; Grebe-Ellis, Johannes (2017-05-01). "Planck's radiation law, the light quantum, and the prehistory of indistinguishability in the teaching of quantum mechanics". European Journal of Physics. 38 (3): 035404. arXiv:1703.05635. Bibcode:2017EJPh...38c5404P. doi:10.1088/1361-6404/aa6134. ISSN 0143-0807. S2CID 119091804.
  4. ^ d'Alembert, Jean (1754). "Croix ou pile". L'Encyclopédie (in French). 4.
  5. ^ d'Alembert, Jean (1754). "Croix ou pile" (PDF). Xavier University. Translated by Richard J. Pulskamp. Retrieved 2019-01-14.
  6. ^ See p. 14, note 3, of the thesis: Michelangeli, Alessandro (October 2007). Bose–Einstein condensation: Analysis of problems and rigorous results (PDF) (Ph.D.). International School for Advanced Studies. (PDF) from the original on 3 November 2018. Retrieved 14 February 2019.
  7. ^ Bose (2 July 1924). "Planck's law and the hypothesis of light quanta" (PostScript). University of Oldenburg. Retrieved 30 November 2016.
  8. ^ Bose (1924), "Plancks Gesetz und Lichtquantenhypothese", Zeitschrift für Physik (in German), 26 (1): 178–181, Bibcode:1924ZPhy...26..178B, doi:10.1007/BF01327326, S2CID 186235974
  9. ^ a b H. J. W. Müller-Kirsten, Basics of Statistical Physics, 2nd ed., World Scientific (2013), ISBN 978-981-4449-53-3.
  10. ^ Srivastava, R. K.; Ashok, J. (2005). "Chapter 7". Statistical Mechanics. New Delhi: PHI Learning Pvt. Ltd. ISBN 9788120327825.
  11. ^ Landau, L. D., Lifšic, E. M., Lifshitz, E. M., & Pitaevskii, L. P. (1980). Statistical physics (Vol. 5). Pergamon Press.
  12. ^ "Chapter 6". Statistical Mechanics. January 2005. ISBN 9788120327825.
  13. ^ The BE distribution can be derived also from thermal field theory.
  14. ^ R. B. Dingle, Asymptotic Expansions: Their Derivation and Interpretation, Academic Press (1973), pp. 267–271.
  15. ^ Ziff R. M.; Kac, M.; Uhlenbeck, G. E. (1977). "The ideal Bose–Einstein gas, revisited". Physics Reports 32: 169–248.
  16. ^ See McQuarrie in citations
  17. ^ Amati, G.; C. J. Van Rijsbergen (2002). "Probabilistic models of information retrieval based on measuring the divergence from randomness " ACM TOIS 20(4):357–389.
  18. ^ Bianconi, G.; Barabási, A.-L. (2001). "Bose–Einstein Condensation in Complex Networks". Physical Review Letters 86: 5632–5635.

References

  • Annett, James F. (2004). Superconductivity, Superfluids and Condensates. New York: Oxford University Press. ISBN 0-19-850755-0.
  • Carter, Ashley H. (2001). Classical and Statistical Thermodynamics. Upper Saddle River, New Jersey: Prentice Hall. ISBN 0-13-779208-5.
  • Griffiths, David J. (2005). Introduction to Quantum Mechanics (2nd ed.). Upper Saddle River, New Jersey: Pearson, Prentice Hall. ISBN 0-13-191175-9.
  • McQuarrie, Donald A. (2000). Statistical Mechanics (1st ed.). Sausalito, California 94965: University Science Books. p. 55. ISBN 1-891389-15-7.{{cite book}}: CS1 maint: location (link)

bose, einstein, statistics, quantum, statistics, statistics, describes, possible, ways, which, collection, interacting, indistinguishable, particles, occupy, available, discrete, energy, states, thermodynamic, equilibrium, aggregation, particles, same, state, . In quantum statistics Bose Einstein statistics B E statistics describes one of two possible ways in which a collection of non interacting indistinguishable particles may occupy a set of available discrete energy states at thermodynamic equilibrium The aggregation of particles in the same state which is a characteristic of particles obeying Bose Einstein statistics accounts for the cohesive streaming of laser light and the frictionless creeping of superfluid helium The theory of this behaviour was developed 1924 25 by Satyendra Nath Bose who recognized that a collection of identical and indistinguishable particles can be distributed in this way The idea was later adopted and extended by Albert Einstein in collaboration with Bose The Bose Einstein statistics applies only to the particles not limited to single occupancy of the same state that is particles that do not obey the Pauli exclusion principle restrictions Such particles have integer values of spin and are named bosons Particles with half integer spins are called fermions and obey Fermi Dirac statistics Comparison of average occupancy of the ground state for three statistics Contents 1 Bose Einstein distribution 2 History 3 Derivation 3 1 Derivation from the microcanonical ensemble 3 2 Derivation from the grand canonical ensemble 3 3 Derivation in the canonical approach 4 Interdisciplinary applications 5 See also 6 Notes 7 ReferencesBose Einstein distribution EditAt low temperatures bosons behave differently from fermions which obey the Fermi Dirac statistics in a way that an unlimited number of them can condense into the same energy state This apparently unusual property also gives rise to the special state of matter the Bose Einstein condensate Fermi Dirac and Bose Einstein statistics apply when quantum effects are important and the particles are indistinguishable Quantum effects appear if the concentration of particles satisfiesN V n q displaystyle frac N V geq n q where N is the number of particles V is the volume and nq is the quantum concentration for which the interparticle distance is equal to the thermal de Broglie wavelength so that the wavefunctions of the particles are barely overlapping Fermi Dirac statistics applies to fermions particles that obey the Pauli exclusion principle and Bose Einstein statistics applies to bosons As the quantum concentration depends on temperature most systems at high temperatures obey the classical Maxwell Boltzmann limit unless they also have a very high density as for a white dwarf Both Fermi Dirac and Bose Einstein become Maxwell Boltzmann statistics at high temperature or at low concentration B E statistics was introduced for photons in 1924 by Bose and generalized to atoms by Einstein in 1924 25 The expected number of particles in an energy state i for B E statistics is n i g i e e i m k B T 1 displaystyle bar n i frac g i e varepsilon i mu k text B T 1 with ei gt m and where ni is the occupation number the number of particles in state i g i displaystyle g i is the degeneracy of energy level i ei is the energy of the i th state m is the chemical potential kB is the Boltzmann constant and T is absolute temperature The variance of this distribution V n displaystyle V n is calculated directly from the expression above for the average number 1 V n k T m n i n 1 n n n 2 displaystyle V n kT frac partial partial mu bar n i langle n rangle 1 langle n rangle bar n bar n 2 For comparison the average number of fermions with energy e i displaystyle varepsilon i given by Fermi Dirac particle energy distribution has a similar form n i e i g i e e i m k B T 1 displaystyle bar n i varepsilon i frac g i e varepsilon i mu k text B T 1 As mentioned above both the Bose Einstein distribution and 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 n i g i e e i m k B T 1 1 displaystyle bar n i frac g i e varepsilon i mu k text B T pm 1 ll 1 therefore e e i m k B T 1 1 displaystyle e varepsilon i mu k text B T pm 1 gg 1 or equivalently e e i m k B T 1 displaystyle e varepsilon i mu k text B T gg 1 In that case n i g i e e i m k B T 1 Z e e i k B T displaystyle bar n i approx frac g i e varepsilon i mu k text B T frac 1 Z e varepsilon i k text B T 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 e i m k B T displaystyle varepsilon i mu gg k text B T is again very small n i g i e e i m k B T 1 1 displaystyle bar n i frac g i e varepsilon i mu k text B T pm 1 ll 1 This again reduces to Maxwell Boltzmann statistics In addition to reducing to the Maxwell Boltzmann distribution in the limit of high T displaystyle T and low density B E statistics also reduces to Rayleigh Jeans law distribution for low energy states with e i m k B T displaystyle varepsilon i mu ll k text B T namelyn i g i e e i m k B T 1 g i e i m k B T g i k B T e i m displaystyle begin aligned bar n i amp frac g i e varepsilon i mu k text B T 1 amp approx frac g i varepsilon i mu k text B T frac g i k text B T varepsilon i mu end aligned History EditWladyslaw Natanson in 1911 concluded that Planck s law requires indistinguishability of units of energy although he did not frame this in terms of Einstein s light quanta 2 3 While presenting a lecture at the University of Dhaka in what was then British India and is now Bangladesh on the theory of radiation and the ultraviolet catastrophe Satyendra Nath Bose intended to show his students that the contemporary theory was inadequate because it predicted results not in accordance with experimental results During this lecture Bose committed an error in applying the theory which unexpectedly gave a prediction that agreed with the experiment The error was a simple mistake similar to arguing that flipping two fair coins will produce two heads one third of the time that would appear obviously wrong to anyone with a basic understanding of statistics remarkably this error resembled the famous blunder by d Alembert known from his Croix ou Pile article 4 5 However the results it predicted agreed with experiment and Bose realized it might not be a mistake after all For the first time he took the position that the Maxwell Boltzmann distribution would not be true for all microscopic particles at all scales Thus he studied the probability of finding particles in various states in phase space where each state is a little patch having phase volume of h3 and the position and momentum of the particles are not kept particularly separate but are considered as one variable Bose adapted this lecture into a short article called Planck s law and the hypothesis of light quanta 6 7 and submitted it to the Philosophical Magazine However the referee s report was negative and the paper was rejected Undaunted he sent the manuscript to Albert Einstein requesting publication in the Zeitschrift fur Physik Einstein immediately agreed personally translated the article from English into German Bose had earlier translated Einstein s article on the general theory of relativity from German to English and saw to it that it was published Bose s theory achieved respect when Einstein sent his own paper in support of Bose s to Zeitschrift fur Physik asking that they be published together The paper came out in 1924 8 The reason Bose produced accurate results was that since photons are indistinguishable from each other one cannot treat any two photons having equal quantum numbers e g polarization and momentum vector as being two distinct identifiable photons By analogy if in an alternate universe coins were to behave like photons and other bosons the probability of producing two heads would indeed be one third and so is the probability of getting a head and a tail which equals one half for the conventional classical distinguishable coins Bose s error leads to what is now called Bose Einstein statistics Bose and Einstein extended the idea to atoms and this led to the prediction of the existence of phenomena which became known as Bose Einstein condensate a dense collection of bosons which are particles with integer spin named after Bose which was demonstrated to exist by experiment in 1995 Derivation EditDerivation from the microcanonical ensemble Edit In the microcanonical ensemble one considers a system with fixed energy volume and number of particles We take a system composed of N i n i textstyle N sum i n i identical bosons n i displaystyle n i of which have energy e i displaystyle varepsilon i and are distributed over g i displaystyle g i levels or states with the same energy e i displaystyle varepsilon i i e g i displaystyle g i is the degeneracy associated with energy e i displaystyle varepsilon i of total energy E i n i e i textstyle E sum i n i varepsilon i Calculation of the number of arrangements of n i displaystyle n i particles distributed among g i displaystyle g i states is a problem of combinatorics Since particles are indistinguishable in the quantum mechanical context here the number of ways for arranging n i displaystyle n i particles in g i displaystyle g i boxes for the i displaystyle i th energy level would be see image The image represents one possible distribution of bosonic particles in different boxes The box partitions green can be moved around to change the size of the boxes and as a result of the number of bosons each box can contain w i BE n i g i 1 n i g i 1 C n i n i g i 1 displaystyle w i text BE frac n i g i 1 n i g i 1 C n i n i g i 1 where C k m displaystyle C k m is the k combination of a set with m elements The total number of arrangements in an ensemble of bosons is simply the product of the binomial coefficients C n i n i g i 1 displaystyle C n i n i g i 1 above over all the energy levels i e W BE i w i BE i n i g i 1 g i 1 n i displaystyle W text BE prod i w i text BE prod i frac n i g i 1 g i 1 n i The maximum number of arrangements determining the corresponding occupation number n i displaystyle n i is obtained by maximizing the entropy or equivalently setting d ln W BE 0 displaystyle mathrm d ln W text BE 0 and taking the subsidiary conditions N n i E i n i e i textstyle N sum n i E sum i n i varepsilon i into account as Lagrange multipliers 9 The result for n i 1 displaystyle n i gg 1 g i 1 displaystyle g i gg 1 n i g i O 1 displaystyle n i g i O 1 is the Bose Einstein distribution Derivation from the grand canonical ensemble Edit The Bose Einstein distribution which applies only to a quantum system of non interacting bosons is naturally derived from the grand canonical ensemble without any approximations 10 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 That is the number of particles within the overall system that occupy a given single particle state form a sub ensemble that is also grand canonical ensemble hence it may be analysed through the construction of a grand partition function Every single particle state is of a fixed energy e displaystyle varepsilon As the sub ensemble associated with a single particle state varies by the number of particles only it is clear that the total energy of the sub ensemble is also directly proportional to the number of particles in the single particle state where N displaystyle N is the number of particles the total energy of the sub ensemble will then be N e displaystyle N varepsilon Beginning with the standard expression for a grand partition function and replacing E displaystyle E with N e displaystyle N varepsilon the grand partition function takes the formZ N exp N m N e k B T N exp N m e k B T displaystyle mathcal Z sum N exp N mu N varepsilon k text B T sum N exp N mu varepsilon k text B T This formula applies to fermionic systems as well as bosonic systems Fermi Dirac statistics arises when considering the effect of the Pauli exclusion principle whilst the number of fermions occupying the same single particle state can only be either 1 or 0 the number of bosons occupying a single particle state may be any integer Thus the grand partition function for bosons can be considered a geometric series and may be evaluated as such Z N 0 exp N m e k B T N 0 exp m e k B T N 1 1 exp m e k B T displaystyle begin aligned mathcal Z amp sum N 0 infty exp N mu varepsilon k text B T sum N 0 infty exp mu varepsilon k text B T N amp frac 1 1 exp mu varepsilon k text B T end aligned Note that the geometric series is convergent only if e m e k B T lt 1 displaystyle e mu varepsilon k text B T lt 1 including the case where ϵ 0 displaystyle epsilon 0 This implies that the chemical potential for the Bose gas must be negative i e m lt 0 displaystyle mu lt 0 whereas the Fermi gas is allowed to take both positive and negative values for the chemical potential 11 The average particle number for that single particle substate is given by N k B T 1 Z Z m V T 1 exp e m k B T 1 displaystyle langle N rangle k text B T frac 1 mathcal Z left frac partial mathcal Z partial mu right V T frac 1 exp varepsilon mu k text B T 1 This result applies for each single particle level and thus forms the Bose Einstein distribution for the entire state of the system 12 13 The variance in particle number s N 2 N 2 N 2 textstyle sigma N 2 langle N 2 rangle langle N rangle 2 is s N 2 k B T d N d m V T exp e m k B T exp e m k B T 1 2 N 1 N displaystyle sigma N 2 k text B T left frac d langle N rangle d mu right V T frac exp varepsilon mu k text B T exp varepsilon mu k text B T 1 2 langle N rangle 1 langle N rangle As a result for highly occupied states the standard deviation of the particle number of an energy level is very large slightly larger than the particle number itself s N N displaystyle sigma N approx langle N rangle This large uncertainty is due to the fact that the probability distribution for the number of bosons in a given energy level is a geometric distribution somewhat counterintuitively the most probable value for N is always 0 In contrast classical particles have instead a Poisson distribution in particle number for a given state with a much smaller uncertainty of s N c l a s s i c a l N textstyle sigma N rm classical sqrt langle N rangle and with the most probable N value being near N displaystyle langle N rangle Derivation in the canonical approach Edit It is also possible to derive approximate Bose Einstein statistics in the canonical ensemble These derivations are lengthy and only yield the above results in the asymptotic limit of a large number of particles The reason is that the total number of bosons is fixed in the canonical ensemble The Bose Einstein distribution in this case can be derived as in most texts by maximization but the mathematically best derivation is by the Darwin Fowler method of mean values as emphasized by Dingle 14 See also Muller Kirsten 9 The fluctuations of the ground state in the condensed region are however markedly different in the canonical and grand canonical ensembles 15 DerivationSuppose we have a number of energy levels labeled by index i displaystyle i each level having energy e i displaystyle varepsilon i and containing a total of n i displaystyle n i particles Suppose each level contains g i displaystyle g i distinct sublevels all of which have the same energy and which are distinguishable For example two particles may have different momenta in which case they are distinguishable from each other yet they can still have the same energy The value of g i displaystyle g i associated with level i displaystyle i is called the degeneracy of that energy level Any number of bosons can occupy the same sublevel Let w n g displaystyle w n g be the number of ways of distributing n displaystyle n particles among the g displaystyle g sublevels of an energy level There is only one way of distributing n displaystyle n particles with one sublevel therefore w n 1 1 displaystyle w n 1 1 It is easy to see that there are n 1 displaystyle n 1 ways of distributing n displaystyle n particles in two sublevels which we will write as w n 2 n 1 n 1 displaystyle w n 2 frac n 1 n 1 With a little thought see Notes below it can be seen that the number of ways of distributing n displaystyle n particles in three sublevels isw n 3 w n 2 w n 1 2 w 1 2 w 0 2 displaystyle w n 3 w n 2 w n 1 2 cdots w 1 2 w 0 2 so that w n 3 k 0 n w n k 2 k 0 n n k 1 n k 1 n 2 n 2 displaystyle w n 3 sum k 0 n w n k 2 sum k 0 n frac n k 1 n k 1 frac n 2 n 2 where we have used the following theorem involving binomial coefficients k 0 n k a k a n a 1 n a 1 displaystyle sum k 0 n frac k a k a frac n a 1 n a 1 Continuing this process we can see that w n g displaystyle w n g is just a binomial coefficient See Notes below w n g n g 1 n g 1 displaystyle w n g frac n g 1 n g 1 For example the population numbers for two particles in three sublevels are 200 110 101 020 011 or 002 for a total of six which equals 4 2 2 The number of ways that a set of occupation numbers n i displaystyle n i can be realized is the product of the ways that each individual energy level can be populated W i w n i g i i n i g i 1 n i g i 1 i n i g i n i g i displaystyle W prod i w n i g i prod i frac n i g i 1 n i g i 1 approx prod i frac n i g i n i g i where the approximation assumes that n i 1 displaystyle n i gg 1 Following the same procedure used in deriving the Maxwell Boltzmann statistics we wish to find the set of n i displaystyle n i for which W is maximised subject to the constraint that there be a fixed total number of particles and a fixed total energy The maxima of W displaystyle W and ln W displaystyle ln W occur at the same value of n i displaystyle n i and since it is easier to accomplish mathematically we will maximise the latter function instead We constrain our solution using Lagrange multipliers forming the function f n i ln W a N n i b E n i e i displaystyle f n i ln W alpha N sum n i beta E sum n i varepsilon i Using the n i 1 displaystyle n i gg 1 approximation and using Stirling s approximation for the factorials x x x e x 2 p x displaystyle left x approx x x e x sqrt 2 pi x right givesf n i i n i g i ln n i g i n i ln n i a N n i b E n i e i K displaystyle f n i sum i n i g i ln n i g i n i ln n i alpha left N sum n i right beta left E sum n i varepsilon i right K Where K is the sum of a number of terms which are not functions of the n i displaystyle n i Taking the derivative with respect to n i displaystyle n i and setting the result to zero and solving for n i displaystyle n i yields the Bose Einstein population numbers n i g i e a b e i 1 displaystyle n i frac g i e alpha beta varepsilon i 1 By a process similar to that outlined in the Maxwell Boltzmann statistics article it can be seen that d ln W a d N b d E displaystyle d ln W alpha dN beta dE which using Boltzmann s famous relationship S k B ln W displaystyle S k text B ln W becomes a statement of the second law of thermodynamics at constant volume and it follows that b 1 k B T displaystyle beta frac 1 k text B T and a m k B T displaystyle alpha frac mu k text B T where S is the entropy m displaystyle mu is the chemical potential kB is Boltzmann s constant and T is the temperature so that finally n i g i e e i m k B T 1 displaystyle n i frac g i e varepsilon i mu k text B T 1 Note that the above formula is sometimes written n i g i e e i k B T z 1 displaystyle n i frac g i e varepsilon i k text B T z 1 where z exp m k B T displaystyle z exp mu k text B T is the absolute activity as noted by McQuarrie 16 Also note that when the particle numbers are not conserved removing the conservation of particle numbers constraint is equivalent to setting a displaystyle alpha and therefore the chemical potential m displaystyle mu to zero This will be the case for photons and massive particles in mutual equilibrium and the resulting distribution will be the Planck distribution NotesA much simpler way to think of Bose Einstein distribution function is to consider that n particles are denoted by identical balls and g shells are marked by g 1 line partitions It is clear that the permutations of these n balls and g 1 partitions will give different ways of arranging bosons in different energy levels Say for 3 n particles and 3 g shells therefore g 1 2 the arrangement might be or or etc Hence the number of distinct permutations of n g 1 objects which have n identical items and g 1 identical items will be g 1 n g 1 n displaystyle frac g 1 n g 1 n See the image for a visual representation of one such distribution of n particles in g boxes that can be represented as g 1 partitions The image represents one possible distribution of bosonic particles in different boxes The box partitions green can be moved around to change the size of the boxes and as a result of the number of bosons each box can contain ORThe purpose of these notes is to clarify some aspects of the derivation of the Bose Einstein B E distribution for beginners The enumeration of cases or ways in the B E distribution can be recast as follows Consider a game of dice throwing in which there are n displaystyle n dice with each die taking values in the set 1 g displaystyle 1 dots g for g 1 displaystyle g geq 1 The constraints of the game are that the value of a die i displaystyle i denoted by m i displaystyle m i has to be greater than or equal to the value of die i 1 displaystyle i 1 denoted by m i 1 displaystyle m i 1 in the previous throw i e m i m i 1 displaystyle m i geq m i 1 Thus a valid sequence of die throws can be described by an n tuple m 1 m 2 m n displaystyle m 1 m 2 dots m n such that m i m i 1 displaystyle m i geq m i 1 Let S n g displaystyle S n g denote the set of these valid n tuples S n g m 1 m 2 m n m i m i 1 m i 1 g i 1 n displaystyle S n g left m 1 m 2 dots m n Big m i geq m i 1 m i in left 1 ldots g right forall i 1 dots n right 1 Then the quantity w n g displaystyle w n g defined above as the number of ways to distribute n displaystyle n particles among the g displaystyle g sublevels of an energy level is the cardinality of S n g displaystyle S n g i e the number of elements or valid n tuples in S n g displaystyle S n g Thus the problem of finding an expression for w n g displaystyle w n g becomes the problem of counting the elements in S n g displaystyle S n g Example n 4 g 3 S 4 3 1111 1112 1113 a 1122 1123 1133 b 1222 1223 1233 1333 c 2222 2223 2233 2333 3333 d displaystyle S 4 3 left underbrace 1111 1112 1113 a underbrace 1122 1123 1133 b underbrace 1222 1223 1233 1333 c underbrace 2222 2223 2233 2333 3333 d right w 4 3 15 displaystyle w 4 3 15 there are 15 displaystyle 15 elements in S 4 3 displaystyle S 4 3 Subset a displaystyle a is obtained by fixing all indices m i displaystyle m i to 1 displaystyle 1 except for the last index m n displaystyle m n which is incremented from 1 displaystyle 1 to g 3 displaystyle g 3 Subset b displaystyle b is obtained by fixing m 1 m 2 1 displaystyle m 1 m 2 1 and incrementing m 3 displaystyle m 3 from 2 displaystyle 2 to g 3 displaystyle g 3 Due to the constraint m i m i 1 displaystyle m i geq m i 1 on the indices in S n g displaystyle S n g the index m 4 displaystyle m 4 must automatically take values in 2 3 displaystyle left 2 3 right The construction of subsets c displaystyle c and d displaystyle d follows in the same manner Each element of S 4 3 displaystyle S 4 3 can be thought of as a multiset of cardinality n 4 displaystyle n 4 the elements of such multiset are taken from the set 1 2 3 displaystyle left 1 2 3 right of cardinality g 3 displaystyle g 3 and the number of such multisets is the multiset coefficient 3 4 3 4 1 3 1 3 4 1 4 6 4 2 15 displaystyle left langle begin matrix 3 4 end matrix right rangle 3 4 1 choose 3 1 3 4 1 choose 4 frac 6 4 2 15 More generally each element of S n g displaystyle S n g is a multiset of cardinality n displaystyle n number of dice with elements taken from the set 1 g displaystyle left 1 dots g right of cardinality g displaystyle g number of possible values of each die and the number of such multisets i e w n g displaystyle w n g is the multiset coefficient w n g g n g n 1 g 1 g n 1 n g n 1 n g 1 displaystyle w n g left langle begin matrix g n end matrix right rangle g n 1 choose g 1 g n 1 choose n frac g n 1 n g 1 2 which is exactly the same as the formula for w n g displaystyle w n g as derived above with the aid of a theorem involving binomial coefficients namely k 0 n k a k a n a 1 n a 1 displaystyle sum k 0 n frac k a k a frac n a 1 n a 1 3 To understand the decomposition w n g k 0 n w n k g 1 w n g 1 w n 1 g 1 w 1 g 1 w 0 g 1 displaystyle w n g sum k 0 n w n k g 1 w n g 1 w n 1 g 1 dots w 1 g 1 w 0 g 1 4 or for example n 4 displaystyle n 4 and g 3 displaystyle g 3 w 4 3 w 4 2 w 3 2 w 2 2 w 1 2 w 0 2 displaystyle w 4 3 w 4 2 w 3 2 w 2 2 w 1 2 w 0 2 let us rearrange the elements of S 4 3 displaystyle S 4 3 as followsS 4 3 1111 1112 1122 1222 2222 a 111 3 112 3 122 3 222 3 b 11 33 12 33 22 33 g 1 333 2 333 d 3333 w displaystyle S 4 3 left underbrace 1111 1112 1122 1222 2222 alpha underbrace 111 color Red underset 3 112 color Red underset 3 122 color Red underset 3 222 color Red underset 3 beta underbrace 11 color Red underset 33 12 color Red underset 33 22 color Red underset 33 gamma underbrace 1 color Red underset 333 2 color Red underset 333 delta underbrace color Red underset 3333 omega right Clearly the subset a displaystyle alpha of S 4 3 displaystyle S 4 3 is the same as the setS 4 2 1111 1112 1122 1222 2222 displaystyle S 4 2 left 1111 1112 1122 1222 2222 right By deleting the index m 4 3 displaystyle m 4 3 shown in red with double underline in the subset b displaystyle beta of S 4 3 displaystyle S 4 3 one obtains the setS 3 2 111 112 122 222 displaystyle S 3 2 left 111 112 122 222 right In other words there is a one to one correspondence between the subset b displaystyle beta of S 4 3 displaystyle S 4 3 and the set S 3 2 displaystyle S 3 2 We write b S 3 2 displaystyle beta longleftrightarrow S 3 2 Similarly it is easy to see that g S 2 2 11 12 22 displaystyle gamma longleftrightarrow S 2 2 left 11 12 22 right d S 1 2 1 2 displaystyle delta longleftrightarrow S 1 2 left 1 2 right w S 0 2 displaystyle omega longleftrightarrow S 0 2 varnothing Thus we can writeS 4 3 k 0 4 S 4 k 2 displaystyle S 4 3 bigcup k 0 4 S 4 k 2 or more generally S n g k 0 n S n k g 1 displaystyle S n g bigcup k 0 n S n k g 1 5 and since the setsS i g 1 for i 0 n displaystyle S i g 1 text for i 0 dots n are non intersecting we thus have w n g k 0 n w n k g 1 displaystyle w n g sum k 0 n w n k g 1 6 with the convention that w 0 g 1 g and w n 0 1 n displaystyle w 0 g 1 forall g text and w n 0 1 forall n 7 Continuing the process we arrive at the following formulaw n g k 1 0 n k 2 0 n k 1 w n k 1 k 2 g 2 k 1 0 n k 2 0 n k 1 k g 0 n j 1 g 1 k j w n i 1 g k i 0 displaystyle w n g sum k 1 0 n sum k 2 0 n k 1 w n k 1 k 2 g 2 sum k 1 0 n sum k 2 0 n k 1 cdots sum k g 0 n sum j 1 g 1 k j w n sum i 1 g k i 0 Using the convention 7 2 above we obtain the formula w n g k 1 0 n k 2 0 n k 1 k g 0 n j 1 g 1 k j 1 displaystyle w n g sum k 1 0 n sum k 2 0 n k 1 cdots sum k g 0 n sum j 1 g 1 k j 1 8 keeping in mind that for q displaystyle q and p displaystyle p being constants we have k 0 q p q p displaystyle sum k 0 q p qp 9 It can then be verified that 8 and 2 give the same result for w 4 3 displaystyle w 4 3 w 3 3 displaystyle w 3 3 w 3 2 displaystyle w 3 2 etc Interdisciplinary applications EditMain article Bose Einstein condensation network theory Viewed as a pure probability distribution the Bose Einstein distribution has found application in other fields In recent years Bose Einstein statistics has also been used as a method for term weighting in information retrieval The method is one of a collection of DFR Divergence From Randomness models 17 the basic notion being that Bose Einstein statistics may be a useful indicator in cases where a particular term and a particular document have a significant relationship that would not have occurred purely by chance Source code for implementing this model is available from the Terrier project at the University of Glasgow The evolution of many complex systems including the World Wide Web business and citation networks is encoded in the dynamic web describing the interactions between the system s constituents Despite their irreversible and nonequilibrium nature these networks follow Bose statistics and can undergo Bose Einstein condensation Addressing the dynamical properties of these nonequilibrium systems within the framework of equilibrium quantum gases predicts that the first mover advantage fit get rich FGR and winner takes all phenomena observed in competitive systems are thermodynamically distinct phases of the underlying evolving networks 18 See also EditBose Einstein correlations Bose Einstein condensate Bose gas Einstein solid Higgs boson Parastatistics Planck s law of black body radiation Superconductivity Fermi Dirac statistics Maxwell Boltzmann statisticsNotes Edit 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 Jammer Max 1966 The conceptual development of quantum mechanics McGraw Hill p 51 ISBN 0 88318 617 9 Passon Oliver Grebe Ellis Johannes 2017 05 01 Planck s radiation law the light quantum and the prehistory of indistinguishability in the teaching of quantum mechanics European Journal of Physics 38 3 035404 arXiv 1703 05635 Bibcode 2017EJPh 38c5404P doi 10 1088 1361 6404 aa6134 ISSN 0143 0807 S2CID 119091804 d Alembert Jean 1754 Croix ou pile L Encyclopedie in French 4 d Alembert Jean 1754 Croix ou pile PDF Xavier University Translated by Richard J Pulskamp Retrieved 2019 01 14 See p 14 note 3 of the thesis Michelangeli Alessandro October 2007 Bose Einstein condensation Analysis of problems and rigorous results PDF Ph D International School for Advanced Studies Archived PDF from the original on 3 November 2018 Retrieved 14 February 2019 Bose 2 July 1924 Planck s law and the hypothesis of light quanta PostScript University of Oldenburg Retrieved 30 November 2016 Bose 1924 Plancks Gesetz und Lichtquantenhypothese Zeitschrift fur Physik in German 26 1 178 181 Bibcode 1924ZPhy 26 178B doi 10 1007 BF01327326 S2CID 186235974 a b H J W Muller Kirsten Basics of Statistical Physics 2nd ed World Scientific 2013 ISBN 978 981 4449 53 3 Srivastava R K Ashok J 2005 Chapter 7 Statistical Mechanics New Delhi PHI Learning Pvt Ltd ISBN 9788120327825 Landau L D Lifsic E M Lifshitz E M amp Pitaevskii L P 1980 Statistical physics Vol 5 Pergamon Press Chapter 6 Statistical Mechanics January 2005 ISBN 9788120327825 The BE distribution can be derived also from thermal field theory R B Dingle Asymptotic Expansions Their Derivation and Interpretation Academic Press 1973 pp 267 271 Ziff R M Kac M Uhlenbeck G E 1977 The ideal Bose Einstein gas revisited Physics Reports 32 169 248 See McQuarrie in citations Amati G C J Van Rijsbergen 2002 Probabilistic models of information retrieval based on measuring the divergence from randomness ACM TOIS 20 4 357 389 Bianconi G Barabasi A L 2001 Bose Einstein Condensation in Complex Networks Physical Review Letters 86 5632 5635 References EditAnnett James F 2004 Superconductivity Superfluids and Condensates New York Oxford University Press ISBN 0 19 850755 0 Carter Ashley H 2001 Classical and Statistical Thermodynamics Upper Saddle River New Jersey Prentice Hall ISBN 0 13 779208 5 Griffiths David J 2005 Introduction to Quantum Mechanics 2nd ed Upper Saddle River New Jersey Pearson Prentice Hall ISBN 0 13 191175 9 McQuarrie Donald A 2000 Statistical Mechanics 1st ed Sausalito California 94965 University Science Books p 55 ISBN 1 891389 15 7 a href Template Cite book html title Template Cite book cite book a CS1 maint location link Retrieved from https en wikipedia org w index php title Bose Einstein statistics amp oldid 1127414905, 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.