Boson: Difference between revisions

From Citizendium
Jump to navigation Jump to search
imported>Paul Wormer
(New page: {{subpages}} In physics, a '''boson''' is an elementary particle with integral spin. According to the Pauli spin statistics postulate, systems of identical bosons are desc...)
 
imported>Meg Taylor
No edit summary
 
(7 intermediate revisions by one other user not shown)
Line 7: Line 7:


The boson is called after the Indian physicist [[Satyendra Nath Bose]] (1894–1974), who was the first to note that  [[photon]]s satisfying [[Planck's law]] for [[blackbody radiation]] obey a special kind of statistics, now called the Bose-Einstein statistics. (Photons have integral spin 1.)
The boson is called after the Indian physicist [[Satyendra Nath Bose]] (1894–1974), who was the first to note that  [[photon]]s satisfying [[Planck's law]] for [[blackbody radiation]] obey a special kind of statistics, now called the Bose-Einstein statistics. (Photons have integral spin 1.)
:*''Note: this article lacks the point of view of particle physics. Its status can only be upgraded to 1 after an elementary-particle physicist has worked on it.''
==Mathematical description==
==Mathematical description==
Let an elementary boson have  coordinates
Let an elementary boson have  coordinates
Line 15: Line 18:
The ''space coordinates'' ''x'', ''y'', and ''z''  are continuous and take on infinitely many values. The ''spin coordinate'' σ is discrete and can have 2''I''+1 different values. For bosons ''I'' is integral.  
The ''space coordinates'' ''x'', ''y'', and ''z''  are continuous and take on infinitely many values. The ''spin coordinate'' σ is discrete and can have 2''I''+1 different values. For bosons ''I'' is integral.  


A one-particle wave function and an ''N''-particle wave function are written as
A one-boson wave function and an ''N''-boson wave function are written as
:<math>
:<math>
\phi_a(\mathbf{r}_1,\sigma_1)\equiv \phi_a(1)\quad\hbox{and}\quad \Phi(1,2,\ldots,N)  
\phi_a(\mathbf{r}_1,\sigma_1)\equiv \phi_a(1)\quad\hbox{and}\quad \Phi(1,2,\ldots,N)  
=\phi_1(1)\phi_2(2)\cdots\phi_N(N),
=\phi_1(1)\phi_2(2)\cdots\phi_N(N),
</math>
</math>
where (k) stands for ('''r'''<sub>k</sub>, &sigma;<sub>k</sub>), k=1,...,''N''.  Here &Phi;(1,2,...,N) is the simplest possible ''N''-particle function (a product, which can be exact only if the bosons do not interact). The ''N''-boson function &Phi;, written here,  has as a major defect:  ''it does not satisfy Pauli's spin statistics postulate''. This postulate states that the function must be symmetric under interchange  of any two boson coordinates (a "transposition").  
where (k) stands for ('''r'''<sub>k</sub>, &sigma;<sub>k</sub>), k=1,...,''N''.  Here &Phi;(1,2,...,N) is the simplest possible ''N''-boson function (a product, which can be exact only if the bosons do not interact). The ''N''-boson function &Phi;, written here,  has as a major defect:  ''it does not satisfy Pauli's spin statistics postulate''. This postulate states that the function must be symmetric under interchange  of any two boson coordinates (under a "transposition").  
===Example===
===Examples===
As a first example we take ''N'' = 2. The following function is not symmetric under interchange of 1 and 2 (given by the permutation operator ''P''<sub>12</sub>),  unless &phi;<sub>''a''</sub> = &phi;<sub>''b''</sub>,  
As a first example we take ''N'' = 2. The following function is not symmetric under interchange of 1 and 2 (given by the transposition operator ''P''<sub>12</sub>),  unless &phi;<sub>''a''</sub> = &phi;<sub>''b''</sub>,  
:<math>
:<math>
\phi_a(1)\phi_b(2) \ne P_{12} \left[\phi_a(1)\phi_b(2)\right] = \phi_a(2)\phi_b(1)\quad\hbox{if}\quad \phi_a\ne\phi_b.
\phi_a(1)\phi_b(2) \ne P_{12} \left[\phi_a(1)\phi_b(2)\right] = \phi_a(2)\phi_b(1)\quad\hbox{if}\quad \phi_a\ne\phi_b.
</math>
</math>
The symmetrized form is symmetric, even when ''a'' &ne; ''b'', and thus is in accordance with the Pauli postulate:
The following (symmetrized) form is invariant under ''P''<sub>12</sub>, even when ''a'' &ne; ''b'', and thus obeys the Pauli postulate:
:<math>
:<math>
\phi_a(1)\phi_b(2) + \phi_a(2)\phi_b(1) = P_{12} \left[\phi_a(1)\phi_b(2) + \phi_a(2)\phi_b(1)\right]= \phi_a(2)\phi_b(1) + \phi_a(1)\phi_b(2) .
\phi_a(1)\phi_b(2) + \phi_a(2)\phi_b(1) = P_{12} \left[\phi_a(1)\phi_b(2) + \phi_a(2)\phi_b(1)\right]= \phi_a(2)\phi_b(1) + \phi_a(1)\phi_b(2) .
Line 38: Line 41:
\phi_a(1)\phi_b(2)\phi_b(3)+\phi_a(2)\phi_b(1)\phi_b(3) = P_{12}\left[ \phi_a(1)\phi_b(2) \phi_b(3) +\phi_a(2)\phi_b(1)\phi_b(3)\right] = \phi_a(2)\phi_b(1)\phi_b(3)+\phi_a(1)\phi_b(2)\phi_b(3).
\phi_a(1)\phi_b(2)\phi_b(3)+\phi_a(2)\phi_b(1)\phi_b(3) = P_{12}\left[ \phi_a(1)\phi_b(2) \phi_b(3) +\phi_a(2)\phi_b(1)\phi_b(3)\right] = \phi_a(2)\phi_b(1)\phi_b(3)+\phi_a(1)\phi_b(2)\phi_b(3).
</math>
</math>
It is easily verified that the latter function is symmetric under all 3! = 6 permutations of the three space-spin coordinates.  
It is easily verified that the latter function is not only symmetric under ''P''<sub>12</sub>, but also under all 3! = 6 permutations of the three space-spin coordinates.  


The last example shows that two bosons may occupy the same one-particle function (two bosons occupy &phi;<sub>''b''</sub>).  This is in contrast to fermions: as soon as two fermions occupy the same one-particle function, the total ''N'' fermion function vanishes.
The last example shows that two bosons may occupy the same one-particle function (two bosons occupy &phi;<sub>''b''</sub>).  This is in contrast to fermions: as soon as two fermions occupy the same one-particle function, the total ''N'' fermion function vanishes.
==Composite systems of fermions==
==Composite systems of fermions==
One may define  bosons by their collective behavior, circumventing spin.  That is, a system of ''N'' identical (not necessarily elementary) particles consists of bosons if the ''N'' particle wave function of the system is symmetric under transpositions of the particle (space plus spin) coordinates.
Let us consider a system consisting of ''N'' identical composite (non-elementary) subsystems, for instance a system consisting of ''N'' identical nuclei, where we recall that a nucleus is not elementary, but composed of neutrons and protons. One may determine whether the subsystems are bosons or fermions by considering the parity of the total wave function under the interchange of two subsystems. The  subsystems are bosons if the wave function is symmetric (has even parity) under transpositions of the subsystems. The system consists of fermions if the parity of the wave function is odd (wave function obtains a minus sign) under transposition of the identical subsystems.  


In order to show that systems consisting of fermions may behave as bosons, we must first recall that the Pauli statistics postulate requires fermionic wave functions to be antisymmetric (to change sign) under interchange of  space-spin coordinates of any two fermions. Consider two systems ''A'' and ''B'' each consisting of two fermions. The space-spin coordinates of the four identical elementary fermions are labeled 1,..., 4. The total wave function is
In order to show that composite subsystems consisting solely of fermions may behave as bosons, we recall that the Pauli statistics postulate requires fermionic wave functions to be antisymmetric (to have odd parity) under interchange of  space-spin coordinates of any two identical fermions. Consider as an example a system consisting of ''N'' = 2 subsystems ''A'' and ''B'', each itself consisting of two fermions. The fermions are labeled 1,..., 4. We assume that fermion 1 of ''A'' is identical to 3 of ''B''  and likewise that 2 and 4 are identical.  A system wave function is
:<math>
:<math>
\Phi_A(1,2)\Phi_B(3,4) \; .
\Phi_A(1,2)\Phi_B(3,4) \; .
</math>
</math>
If the fermions within ''A'' and ''B'' are strongly coupled (for instance by nuclear forces), permutations that effectively interchange 1 and 2 and/or 3 and 4 do not have to be considered,
If both fermions of ''A'' (1 and  2) and of ''B'' (3 and 4) are identical one may expect the intra-subsystem transpositions 1 &harr; 2  and 3 &harr; 4 to enter the discussion as well.  However, when there is strong coupling within the subsystems (for instance by nuclear forces), permutations that effectively interchange 1 and 2 and/or 3 and 4 are not [[feasible permutation|feasible]] and do not have to be considered. The only relevant permutation is then the transposition of ''A'' and ''B'',
so that the only permutation is
:<math>
P_{AB} = P_{13}P_{24} \quad\hbox{with}\quad P_{AB} \Phi_A(1,2)\Phi_B(3,4) = \Phi_B(1,2)\Phi_A(3,4) =\Phi_A(3,4)\Phi_B(1,2).
</math>
Since the interchange of ''A'' and ''B'' consists of two transpositions of fermions, it gives the sign (&minus;1)&times;(&minus;1) = 1 and hence the identical systems ''A'' and ''B'' are bosons (i.e., their wave function is symmetric under transposition).  The wave function must be symmetrized,
:<math>
:<math>
P_{AB} = P_{12}P_{34} \quad\hbox{with}\quad P_{AB} \Phi_A(1,2)\Phi_B(3,4) = \Phi_B(1,2)\Phi_A(3,4) =\Phi_A(3,4)\Phi_B(1,2).
\Phi_A(1,2)\Phi_B(3,4) + \Phi_B(1,2)\Phi_A(3,4)\; \quad\hbox{for}
\quad\Phi_A \ne \Phi_B.
</math>
</math>
Since this permutation consists of two transpositions of fermions, it gives the sign (-1)&times;(-1) = 1 and hence the identical systems ''A'' and ''B'' are bosons (i.e., their wave function is symmetric under transposition).  
in order to obey the Pauli postulate.


In this way one can explain why an [[ideal gas law|ideal gas]], which by definition consists of non-interacting particles, is sometimes bosonic and sometimes fermionic. For instance, an ideal gas  
In this way one can explain why an [[ideal gas law|ideal gas]], which by definition consists of non-interacting particles, is sometimes bosonic and sometimes fermionic. For instance, an ideal gas  
of H-atoms is bosonic, while that consisting of D (= <sup>2</sup>H) atoms is fermionic. Deuterium consists of three fermions: a proton, a neutron, and an electron. A simultaneous permutation of three fermions that is equivalent to the permutation of two D-atoms gives a minus sign.
of H-atoms is bosonic (an H-atom consists of two fermions: one electron and one proton), while an ideal gas consisting of D (= <sup>2</sup>H) atoms is fermionic. Deuterium consists of three fermions: a proton, a neutron, and an electron. Simultaneous transpositions within the three pairs of fermions, which is equivalent to the interchange of two D-atoms, gives (&minus;1)<sup>3</sup> = &minus;1.
 
Finally, it must be noted that for non-ideal gases the conclusions can be different. If, for example, we turn on the chemical interaction between the D-atoms in the fermionic ideal gas  of D-atoms,  D<sub>2</sub> molecules will be formed. A D<sub>2</sub> molecule consists of an even number of fermions and hence a gas of D<sub>2</sub> molecules is a boson system.
 
==Complete sets of orbital products==
Consider a [[linear space]] ''V''<sub>''M''</sub> spanned by the ''M'' boson spin-orbitals &phi;<sub>1</sub>, &phi;<sub>2</sub>, ..., &phi;<sub>''M''</sub>. We assume these orbitals to be orthonormal.  Consider the [[tensor]] power
:<math>
V_M \otimes^N \equiv \underbrace{V_M\otimes V_M\otimes \cdots\otimes V_M}_{N \mathrm{factors}}
</math>
The tensor power is of dimension ''M''<sup>''N''</sup> and the symmetrized states are in the subspace <font style="vertical-align: baseline"><math>W\equiv\mathcal{S}\big(V_M \otimes^N\big)</math></font>,  which is of dimension <math>\binom{N+M-1}{N}</math>.
An unnormalized  [[basis]] (maximum linearly independent set) for ''W'' is the following set
:<math>
\big\{ \mathcal{S}\big(\phi_{i_1}\otimes \phi_{i_2}\otimes \cdots\otimes  \phi_{i_N}\big) \;;\;
i_1 \le i_2 \le \cdots \le i_N \big\} .
</math>
'''Example'''. ''M'' = 3 and ''N'' = 3, the number of symmetrized states is <math>\binom{5}{3} = 10</math>
:<math>
\begin{align}
\mathcal{S}\big(\phi_{1}\otimes \phi_{1}\otimes \phi_{1}\big) &\equiv \mathcal{S}| 3, 0, 0\rangle \\
\mathcal{S}\big(\phi_{1}\otimes \phi_{1}\otimes \phi_{2}\big) &\equiv \mathcal{S}| 2, 1, 0\rangle \\
\mathcal{S}\big(\phi_{1}\otimes \phi_{1}\otimes \phi_{3}\big) &\equiv \mathcal{S}| 2, 0, 1\rangle \\
\mathcal{S}\big(\phi_{1}\otimes \phi_{2}\otimes \phi_{2}\big) &\equiv \mathcal{S}| 1, 2, 0\rangle \\
\mathcal{S}\big(\phi_{1}\otimes \phi_{2}\otimes \phi_{3}\big) &\equiv \mathcal{S}| 1, 1, 1\rangle \\
\mathcal{S}\big(\phi_{1}\otimes \phi_{3}\otimes \phi_{3}\big) &\equiv \mathcal{S}| 1, 0, 2\rangle \\
\mathcal{S}\big(\phi_{2}\otimes \phi_{2}\otimes \phi_{2}\big) &\equiv \mathcal{S}| 0, 3, 0\rangle \\
\mathcal{S}\big(\phi_{2}\otimes \phi_{2}\otimes \phi_{3}\big) &\equiv \mathcal{S}| 0, 2, 1\rangle \\
\mathcal{S}\big(\phi_{2}\otimes \phi_{3}\otimes \phi_{3}\big) &\equiv \mathcal{S}| 0, 1, 2\rangle \\
\mathcal{S}\big(\phi_{3}\otimes \phi_{3}\otimes \phi_{3}\big) &\equiv \mathcal{S}| 0, 0, 3\rangle \\
\end{align}
</math>
Here we exemplified the "occupation number notation" in which an arbitrary element is written as
:<math>
\mathcal{S}| n_1, n_2, \ldots, n_M\rangle \equiv \mathcal{S}\Big( (\phi_1)^{n_1}\otimes(\phi_2)^{n_2}\otimes\cdots\otimes(\phi_M)^{n_M}\Big)
</math>
where occupation numbers ''n''<sub>''k''</sub>, ''k''=1, ..., ''M'', may be zero.
A complete basis for the symmetrized space ''W'' is labeled by all possible decompositions  of ''N''
:<math>
\big\{ \mathcal{S}| n_1, n_2, \ldots, n_M\rangle\; ; \; n_1+n_2+ \cdots + n_M = N  \big\}
</math>
 
The symmetrized products  are orthogonal but not normalized. In  [[Symmetrizer|this article]] it is shown that their normalization coefficient  is the square root of a [[multinomial coefficient]]:
:<math>
K_{n_1,n_2,\ldots,n_M} = \left({ N \atop n_1\;n_2\;\ldots\;n_M}\right)^{1/2}
</math>
 
In conclusion, the following set of symmetrized states forms an orthonormal basis for ''W'':
:<math>
| n_1, n_2, \ldots, n_M\rangle_{\mathrm{sym}} \equiv K_{n_1,n_2,\ldots,n_M}\;\mathcal{S}| n_1, n_2, \ldots, n_M\rangle
</math>
where the ''M'' occupation numbers are natural numbers (positive or zero)  with the constraint that they  sum up to ''N''.
===Application===
The [[trace (mathematics)|trace]] of a linear operator ''A'' can be expressed in terms of an orthonormal basis. The canonical [[partition function (statistical physics)|partition function]]
''Q'' is the trace of  exp(&minus;&beta;''H''), where ''H'' is an energy operator and &beta; &equiv; 1/(''kT''). (Here ''k'' is [[Boltzmann's constant]] and ''T'' is the absolute [[temperature]].)
 
As an example of a statistical physics trace calculation we consider a [[Hamiltonian]] of a system of ''N'' non-interacting bosons:
:<math>
H = h(1) + h(2) + \cdots + h(N)
</math>
and assume that the one-particle space ''V''<sub>''M''</sub> is spanned by eigenstates:
:<math>
h \phi_{i} = \epsilon_i \phi_i \Longrightarrow e^{-\beta h} \phi_i = \lambda_i \phi_i \quad\hbox{with}\quad \lambda_i \equiv e^{-\beta \epsilon_i}.
</math>
 
The operator ''H'' commutes with <font style="vertical-align: text-top"><math>\mathcal{S}</math></font> so that
 
:<math>
e^{-\beta H}\; | n_1, n_2, \ldots, n_M\rangle_{\mathrm{sym}} = (\lambda_1)^{n_1}(\lambda_2)^{n_2} \cdots (\lambda_M)^{n_M} \; | n_1, n_2, \ldots, n_M\rangle_{\mathrm{sym}}
</math>
From this follows
:<math>
Q\equiv\mathrm{Tr} e^{-\beta H} = \sum_{n_1, n_2, \dots, n_M \atop n_1+n_2+\cdots+n_M = N}
(\lambda_1)^{n_1}(\lambda_2)^{n_2} \cdots (\lambda_M)^{n_M}
</math>
because the  symmetrized  states are normalized.
Compare this with the [[multinomial coefficient|multinomial]] expansion
:<math>
q^N \equiv (\lambda_1 + \lambda_2 + \cdots + \lambda_M)^{N} =
\sum_{n_1, n_2, \dots, n_M \atop n_1+n_2+\cdots+n_M = N}
\left({ N \atop n_1\;n_2\;\ldots\;n_M}\right)
(\lambda_1)^{n_1}(\lambda_2)^{n_2} \cdots (\lambda_M)^{n_M},
</math>
where ''q'' is the one-boson partition function.
In statistical physics ''M'' is usually infinite, the space ''V''<sub>''M''</sub> is a one-boson [[Hilbert space]]. When the temperature ''T'' is fairly high, not many states are occupied more than once. For most terms in the sum we have occupation numbers 1 or 0. If we make the approximation that this is true for all terms, the multinomial coefficient becomes equal to ''N''! and we obtain
:<math>
N!\;Q = N! \sum_{n_1, n_2, \dots, n_M \atop n_1+n_2+\cdots+n_M = N}
(\lambda_1)^{n_1}(\lambda_2)^{n_2} \cdots (\lambda_M)^{n_M} \approx \; q^N
\quad\Longrightarrow\quad Q \approx \frac{q^N}{N!}.
</math>
This approximation is known as [[Boltzmann statistics]].

Latest revision as of 04:59, 1 November 2013

This article is developing and not approved.
Main Article
Discussion
Related Articles  [?]
Bibliography  [?]
External Links  [?]
Citable Version  [?]
 
This editable Main Article is under development and subject to a disclaimer.

In physics, a boson is an elementary particle with integral spin. According to the Pauli spin statistics postulate, systems of identical bosons are described by totally symmetric (under permutations of the bosons) wave functions.

A composite system of an even number of fermions may behave as a boson when the coupling between the constituting fermions is strong. For instance, atomic nuclei are composed of protons and neutrons; both types of nucleons are fermions. Atomic nuclei whose mass number A (the total number of nucleons) is even are bosons. Nuclei with odd A are fermions.

A thermodynamical system of N bosons satisfies Bose-Einstein statistics.

The boson is called after the Indian physicist Satyendra Nath Bose (1894–1974), who was the first to note that photons satisfying Planck's law for blackbody radiation obey a special kind of statistics, now called the Bose-Einstein statistics. (Photons have integral spin 1.)

  • Note: this article lacks the point of view of particle physics. Its status can only be upgraded to 1 after an elementary-particle physicist has worked on it.

Mathematical description

Let an elementary boson have coordinates

The space coordinates x, y, and z are continuous and take on infinitely many values. The spin coordinate σ is discrete and can have 2I+1 different values. For bosons I is integral.

A one-boson wave function and an N-boson wave function are written as

where (k) stands for (rk, σk), k=1,...,N. Here Φ(1,2,...,N) is the simplest possible N-boson function (a product, which can be exact only if the bosons do not interact). The N-boson function Φ, written here, has as a major defect: it does not satisfy Pauli's spin statistics postulate. This postulate states that the function must be symmetric under interchange of any two boson coordinates (under a "transposition").

Examples

As a first example we take N = 2. The following function is not symmetric under interchange of 1 and 2 (given by the transposition operator P12), unless φa = φb,

The following (symmetrized) form is invariant under P12, even when ab, and thus obeys the Pauli postulate:

For the second example we take N = 3 and write the following non-symmetric function with ab:

but

It is easily verified that the latter function is not only symmetric under P12, but also under all 3! = 6 permutations of the three space-spin coordinates.

The last example shows that two bosons may occupy the same one-particle function (two bosons occupy φb). This is in contrast to fermions: as soon as two fermions occupy the same one-particle function, the total N fermion function vanishes.

Composite systems of fermions

Let us consider a system consisting of N identical composite (non-elementary) subsystems, for instance a system consisting of N identical nuclei, where we recall that a nucleus is not elementary, but composed of neutrons and protons. One may determine whether the subsystems are bosons or fermions by considering the parity of the total wave function under the interchange of two subsystems. The subsystems are bosons if the wave function is symmetric (has even parity) under transpositions of the subsystems. The system consists of fermions if the parity of the wave function is odd (wave function obtains a minus sign) under transposition of the identical subsystems.

In order to show that composite subsystems consisting solely of fermions may behave as bosons, we recall that the Pauli statistics postulate requires fermionic wave functions to be antisymmetric (to have odd parity) under interchange of space-spin coordinates of any two identical fermions. Consider as an example a system consisting of N = 2 subsystems A and B, each itself consisting of two fermions. The fermions are labeled 1,..., 4. We assume that fermion 1 of A is identical to 3 of B and likewise that 2 and 4 are identical. A system wave function is

If both fermions of A (1 and 2) and of B (3 and 4) are identical one may expect the intra-subsystem transpositions 1 ↔ 2 and 3 ↔ 4 to enter the discussion as well. However, when there is strong coupling within the subsystems (for instance by nuclear forces), permutations that effectively interchange 1 and 2 and/or 3 and 4 are not feasible and do not have to be considered. The only relevant permutation is then the transposition of A and B,

Since the interchange of A and B consists of two transpositions of fermions, it gives the sign (−1)×(−1) = 1 and hence the identical systems A and B are bosons (i.e., their wave function is symmetric under transposition). The wave function must be symmetrized,

in order to obey the Pauli postulate.

In this way one can explain why an ideal gas, which by definition consists of non-interacting particles, is sometimes bosonic and sometimes fermionic. For instance, an ideal gas of H-atoms is bosonic (an H-atom consists of two fermions: one electron and one proton), while an ideal gas consisting of D (= 2H) atoms is fermionic. Deuterium consists of three fermions: a proton, a neutron, and an electron. Simultaneous transpositions within the three pairs of fermions, which is equivalent to the interchange of two D-atoms, gives (−1)3 = −1.

Finally, it must be noted that for non-ideal gases the conclusions can be different. If, for example, we turn on the chemical interaction between the D-atoms in the fermionic ideal gas of D-atoms, D2 molecules will be formed. A D2 molecule consists of an even number of fermions and hence a gas of D2 molecules is a boson system.

Complete sets of orbital products

Consider a linear space VM spanned by the M boson spin-orbitals φ1, φ2, ..., φM. We assume these orbitals to be orthonormal. Consider the tensor power

The tensor power is of dimension MN and the symmetrized states are in the subspace , which is of dimension . An unnormalized basis (maximum linearly independent set) for W is the following set

Example. M = 3 and N = 3, the number of symmetrized states is

Here we exemplified the "occupation number notation" in which an arbitrary element is written as

where occupation numbers nk, k=1, ..., M, may be zero. A complete basis for the symmetrized space W is labeled by all possible decompositions of N

The symmetrized products are orthogonal but not normalized. In this article it is shown that their normalization coefficient is the square root of a multinomial coefficient:

In conclusion, the following set of symmetrized states forms an orthonormal basis for W:

where the M occupation numbers are natural numbers (positive or zero) with the constraint that they sum up to N.

Application

The trace of a linear operator A can be expressed in terms of an orthonormal basis. The canonical partition function Q is the trace of exp(−βH), where H is an energy operator and β ≡ 1/(kT). (Here k is Boltzmann's constant and T is the absolute temperature.)

As an example of a statistical physics trace calculation we consider a Hamiltonian of a system of N non-interacting bosons:

and assume that the one-particle space VM is spanned by eigenstates:

The operator H commutes with so that

From this follows

because the symmetrized states are normalized. Compare this with the multinomial expansion

where q is the one-boson partition function. In statistical physics M is usually infinite, the space VM is a one-boson Hilbert space. When the temperature T is fairly high, not many states are occupied more than once. For most terms in the sum we have occupation numbers 1 or 0. If we make the approximation that this is true for all terms, the multinomial coefficient becomes equal to N! and we obtain

This approximation is known as Boltzmann statistics.