Wick ordering
   HOME

TheInfoList



OR:

In quantum field theory a product of quantum fields, or equivalently their
creation and annihilation operators Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator (usually d ...
, is usually said to be normal ordered (also called Wick order) when all creation operators are to the left of all annihilation operators in the product. The process of putting a product into normal order is called normal ordering (also called Wick ordering). The terms antinormal order and antinormal ordering are analogously defined, where the annihilation operators are placed to the left of the creation operators. Normal ordering of a product quantum fields or
creation and annihilation operators Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator (usually d ...
can also be defined in many other ways. Which definition is most appropriate depends on the expectation values needed for a given calculation. Most of this article uses the most common definition of normal ordering as given above, which is appropriate when taking
expectation value In probability theory, the expected value (also called expectation, expectancy, mathematical expectation, mean, average, or first moment) is a generalization of the weighted average. Informally, the expected value is the arithmetic mean of a l ...
s using the vacuum state of the
creation and annihilation operators Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator (usually d ...
. The process of normal ordering is particularly important for a
quantum mechanical Quantum mechanics is a fundamental theory in physics that provides a description of the physical properties of nature at the scale of atoms and subatomic particles. It is the foundation of all quantum physics including quantum chemistry, qua ...
Hamiltonian Hamiltonian may refer to: * Hamiltonian mechanics, a function that represents the total energy of a system * Hamiltonian (quantum mechanics), an operator corresponding to the total energy of that system ** Dyall Hamiltonian, a modified Hamiltonian ...
. When quantizing a classical Hamiltonian there is some freedom when choosing the operator order, and these choices lead to differences in the ground state energy.


Notation

If \hat denotes an arbitrary product of creation and/or annihilation operators (or equivalently, quantum fields), then the normal ordered form of \hat is denoted by \mathopen \hat \mathclose. An alternative notation is \mathcal(\hat). Note that normal ordering is a concept that only makes sense for products of operators. Attempting to apply normal ordering to a sum of operators is not useful as normal ordering is not a linear operation.


Bosons

Bosons In particle physics, a boson ( ) is a subatomic particle whose spin quantum number has an integer value (0,1,2 ...). Bosons form one of the two fundamental classes of subatomic particle, the other being fermions, which have odd half-integer ...
are particles which satisfy
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 ...
. We will now examine the normal ordering of bosonic creation and annihilation operator products.


Single bosons

If we start with only one type of boson there are two operators of interest: * \hat^\dagger: the boson's creation operator. * \hat: the boson's annihilation operator. These satisfy the commutator relationship :\left hat^\dagger, \hat^\dagger \right- = 0 :\left hat, \hat \right- = 0 :\left hat, \hat^\dagger \right- = 1 where \left A, B \right- \equiv AB - BA denotes the commutator. We may rewrite the last one as: \hat\, \hat^\dagger = \hat^\dagger\, \hat + 1.


Examples

1. We'll consider the simplest case first. This is the normal ordering of \hat^\dagger \hat: : \hat^\dagger \, \hat = \hat^\dagger \, \hat. The expression \hat^\dagger \, \hat has not been changed because it is ''already'' in normal order - the creation operator (\hat^\dagger) is already to the left of the annihilation operator (\hat). 2. A more interesting example is the normal ordering of \hat \, \hat^\dagger : : \hat \, \hat^\dagger = \hat^\dagger \, \hat. Here the normal ordering operation has ''reordered'' the terms by placing \hat^\dagger to the left of \hat. These two results can be combined with the commutation relation obeyed by \hat and \hat^\dagger to get : \hat \, \hat^\dagger = \hat^\dagger \, \hat + 1 = \hat \, \hat^\dagger \; + 1. or : \hat \, \hat^\dagger - \hat \, \hat^\dagger = 1. This equation is used in defining the contractions used in
Wick's theorem Wick's theorem is a method of reducing high-order derivatives to a combinatorics problem. It is named after Italian physicist Gian-Carlo Wick. It is used extensively in quantum field theory to reduce arbitrary products of creation and annihila ...
. 3. An example with multiple operators is: : \hat^\dagger \, \hat \, \hat \, \hat^\dagger \, \hat \, \hat^\dagger \, \hat = \hat^\dagger \, \hat^\dagger \, \hat^\dagger \, \hat \, \hat \, \hat \, \hat = (\hat^\dagger)^3 \, \hat^4. 4. A simple example shows that normal ordering cannot be extended by linearity from the monomials to all operators in a self-consistent way: : \hat \hat^\dagger = 1 + \hat^\dagger \hat = 1 + \hat^\dagger \hat = 1 + \hat^\dagger \hat \ne \hat^\dagger \hat=\hat \hat^\dagger The implication is that normal ordering is not a linear function on operators.


Multiple bosons

If we now consider N different bosons there are 2 N operators: * \hat_i^\dagger: the i^ boson's creation operator. * \hat_i: the i^ boson's annihilation operator. Here i = 1,\ldots,N. These satisfy the commutation relations: :\left hat_i^\dagger, \hat_j^\dagger \right- = 0 :\left hat_i, \hat_j \right- = 0 :\left hat_i, \hat_j^\dagger \right- = \delta_ where i,j = 1,\ldots,N and \delta_ denotes the
Kronecker delta In mathematics, the Kronecker delta (named after Leopold Kronecker) is a function of two variables, usually just non-negative integers. The function is 1 if the variables are equal, and 0 otherwise: \delta_ = \begin 0 &\text i \neq j, \\ 1 & ...
. These may be rewritten as: :\hat_i^\dagger \, \hat_j^\dagger = \hat_j^\dagger \, \hat_i^\dagger :\hat_i \, \hat_j = \hat_j \, \hat_i :\hat_i \,\hat_j^\dagger = \hat_j^\dagger \,\hat_i + \delta_.


Examples

1. For two different bosons (N=2) we have : : \hat_1^\dagger \,\hat_2 : \,= \hat_1^\dagger \,\hat_2 : : \hat_2 \, \hat_1^\dagger : \,= \hat_1^\dagger \,\hat_2 2. For three different bosons (N=3) we have : : \hat_1^\dagger \,\hat_2 \,\hat_3 : \,= \hat_1^\dagger \,\hat_2 \,\hat_3 Notice that since (by the commutation relations) \hat_2 \,\hat_3 = \hat_3 \,\hat_2 the order in which we write the annihilation operators does not matter. : : \hat_2 \, \hat_1^\dagger \, \hat_3 : \,= \hat_1^\dagger \,\hat_2 \, \hat_3 : : \hat_3 \hat_2 \, \hat_1^\dagger : \,= \hat_1^\dagger \,\hat_2 \, \hat_3


Bosonic operator functions

Normal ordering of bosonic operator functions f(\hat n), with occupation number operator \hat n=\hat b\vphantom^\dagger \hat b, can be accomplished using (falling) factorial powers \hat n^=\hat n(\hat n-1)\cdots(\hat n-k+1) and
Newton series A finite difference is a mathematical expression of the form . If a finite difference is divided by , one gets a difference quotient. The approximation of derivatives by finite differences plays a central role in finite difference methods for the ...
instead of
Taylor series In mathematics, the Taylor series or Taylor expansion of a function is an infinite sum of terms that are expressed in terms of the function's derivatives at a single point. For most common functions, the function and the sum of its Taylor ser ...
: It is easy to show that factorial powers \hat n^ are equal to normal-ordered (raw) powers \hat n^ and are therefore normal ordered by construction, : \hat^ = \hat b\vphantom^ \hat b\vphantom^k = \hat n^k, such that the Newton series expansion : \tilde f(\hat n) = \sum_^\infty \Delta_n^k \tilde f(0) \, \frac of an operator function \tilde f(\hat n), with k-th
forward difference A finite difference is a mathematical expression of the form . If a finite difference is divided by , one gets a difference quotient. The approximation of derivatives by finite differences plays a central role in finite difference methods for the ...
\Delta_n^k \tilde f(0) at n=0, is always normal ordered. Here, the
eigenvalue equation In linear algebra, linear transformations can be represented by matrices. If T is a linear transformation mapping \mathbb^n to \mathbb^m and \mathbf x is a column vector with n entries, then T( \mathbf x ) = A \mathbf x for some m \times n matrix ...
\hat n , n\rangle = n , n\rangle relates \hat n and n. As a consequence, the normal-ordered Taylor series of an arbitrary function f(\hat n) is equal to the Newton series of an associated function \tilde f(\hat n), fulfilling : \tilde f(\hat n) = f(\hat n) , if the series coefficients of the Taylor series of f(x), with continuous x, match the coefficients of the Newton series of \tilde f(n), with integer n, : \begin f(x) &= \sum_^\infty F_k \, \frac, \\ \tilde f(n) &= \sum_^\infty F_k \, \frac, \\ F_k &= \partial_x^k f(0) = \Delta_n^k \tilde f(0), \end with k-th partial derivative \partial_x^k f(0) at x=0. The functions f and \tilde f are related through the so-called normal-order transform \mathcal N[f] according to : \begin \tilde f(n) &= \mathcal N_x[f(x)](n) \\ &= \frac \int_^0 \mathrm d x \, e^x \, f(x) \, (-x)^ \\ &= \frac\mathcal M_[e^ f(x)](-n), \end which can be expressed in terms of the Mellin transform \mathcal M, see for details.


Fermions

Fermions are particles which satisfy Fermi–Dirac statistics. We will now examine the normal ordering of fermionic creation and annihilation operator products.


Single fermions

For a single fermion there are two operators of interest: * \hat^\dagger: the fermion's creation operator. * \hat: the fermion's annihilation operator. These satisfy the anticommutator relationships :\left hat^\dagger, \hat^\dagger \right+ = 0 :\left hat, \hat \right+ = 0 :\left hat, \hat^\dagger \right+ = 1 where \left[A, B \right]_+ \equiv AB + BA denotes the anticommutator. These may be rewritten as :\hat^\dagger\, \hat^\dagger = 0 :\hat \,\hat = 0 :\hat \,\hat^\dagger = 1 - \hat^\dagger \,\hat . To define the normal ordering of a product of fermionic creation and annihilation operators we must take into account the number of Transposition (mathematics), interchanges between neighbouring operators. We get a minus sign for each such interchange.


Examples

1. We again start with the simplest cases: : : \hat^\dagger \, \hat : \,= \hat^\dagger \, \hat This expression is already in normal order so nothing is changed. In the reverse case, we introduce a minus sign because we have to change the order of two operators: : : \hat \, \hat^\dagger : \,= -\hat^\dagger \, \hat These can be combined, along with the anticommutation relations, to show : \hat \, \hat^\dagger \,= 1 - \hat^\dagger \, \hat = 1 + :\hat \,\hat^\dagger : or : \hat \, \hat^\dagger - : \hat \, \hat^\dagger : = 1. This equation, which is in the same form as the bosonic case above, is used in defining the contractions used in
Wick's theorem Wick's theorem is a method of reducing high-order derivatives to a combinatorics problem. It is named after Italian physicist Gian-Carlo Wick. It is used extensively in quantum field theory to reduce arbitrary products of creation and annihila ...
. 2. The normal order of any more complicated cases gives zero because there will be at least one creation or annihilation operator appearing twice. For example: : : \hat\,\hat^\dagger \, \hat \hat^\dagger : \,= -\hat^\dagger \,\hat^\dagger \,\hat\,\hat = 0


Multiple fermions

For N different fermions there are 2 N operators: * \hat_i^\dagger: the i^ fermion's creation operator. * \hat_i: the i^ fermion's annihilation operator. Here i = 1,\ldots,N. These satisfy the anti-commutation relations: :\left hat_i^\dagger, \hat_j^\dagger \right+ = 0 :\left hat_i, \hat_j \right+ = 0 :\left hat_i, \hat_j^\dagger \right+ = \delta_ where i,j = 1,\ldots,N and \delta_ denotes the
Kronecker delta In mathematics, the Kronecker delta (named after Leopold Kronecker) is a function of two variables, usually just non-negative integers. The function is 1 if the variables are equal, and 0 otherwise: \delta_ = \begin 0 &\text i \neq j, \\ 1 & ...
. These may be rewritten as: :\hat_i^\dagger \, \hat_j^\dagger = -\hat_j^\dagger \, \hat_i^\dagger :\hat_i \, \hat_j = -\hat_j \, \hat_i :\hat_i \,\hat_j^\dagger = \delta_ - \hat_j^\dagger \,\hat_i . When calculating the normal order of products of fermion operators we must take into account the number of Transposition (mathematics), interchanges of neighbouring operators required to rearrange the expression. It is as if we pretend the creation and annihilation operators anticommute and then we reorder the expression to ensure the creation operators are on the left and the annihilation operators are on the right - all the time taking account of the anticommutation relations.


Examples

1. For two different fermions (N=2) we have : : \hat_1^\dagger \,\hat_2 : \,= \hat_1^\dagger \,\hat_2 Here the expression is already normal ordered so nothing changes. : : \hat_2 \, \hat_1^\dagger : \,= -\hat_1^\dagger \,\hat_2 Here we introduce a minus sign because we have interchanged the order of two operators. : : \hat_2 \, \hat_1^\dagger \, \hat^\dagger_2 : \,= \hat_1^\dagger \, \hat_2^\dagger \,\hat_2 = -\hat_2^\dagger \, \hat_1^\dagger \,\hat_2 Note that the order in which we write the operators here, unlike in the bosonic case, ''does matter''. 2. For three different fermions (N=3) we have : : \hat_1^\dagger \, \hat_2 \, \hat_3 : \,= \hat_1^\dagger \,\hat_2 \,\hat_3 = -\hat_1^\dagger \,\hat_3 \,\hat_2 Notice that since (by the anticommutation relations) \hat_2 \,\hat_3 = -\hat_3 \,\hat_2 the order in which we write the operators ''does matter'' in this case. Similarly we have : : \hat_2 \, \hat_1^\dagger \, \hat_3 : \,= -\hat_1^\dagger \,\hat_2 \, \hat_3 = \hat_1^\dagger \,\hat_3 \, \hat_2 : : \hat_3 \hat_2 \, \hat_1^\dagger : \,= \hat_1^\dagger \,\hat_3 \, \hat_2 = -\hat_1^\dagger \,\hat_2 \, \hat_3


Uses in quantum field theory

The vacuum expectation value of a normal ordered product of creation and annihilation operators is zero. This is because, denoting the vacuum state by , 0\rangle, the creation and annihilation operators satisfy :\langle 0 , \hat^\dagger = 0 \qquad \textrm \qquad \hat , 0\rangle = 0 (here \hat^\dagger and \hat are creation and annihilation operators (either bosonic or fermionic)). Let \hat denote a non-empty product of creation and annihilation operators. Although this may satisfy :\langle 0 , \hat , 0 \rangle \neq 0, we have :\langle 0 , :\hat: , 0 \rangle = 0. Normal ordered operators are particularly useful when defining a quantum mechanical
Hamiltonian Hamiltonian may refer to: * Hamiltonian mechanics, a function that represents the total energy of a system * Hamiltonian (quantum mechanics), an operator corresponding to the total energy of that system ** Dyall Hamiltonian, a modified Hamiltonian ...
. If the Hamiltonian of a theory is in normal order then the ground state energy will be zero: \langle 0 , \hat, 0\rangle = 0.


Free fields

With two free fields φ and χ, ::\phi(x)\chi(y):\,\,=\phi(x)\chi(y)-\langle 0, \phi(x)\chi(y), 0\rangle where , 0\rangle is again the vacuum state. Each of the two terms on the right hand side typically blows up in the limit as y approaches x but the difference between them has a well-defined limit. This allows us to define :φ(x)χ(x):.


Wick's theorem

Wick's theorem states the relationship between the time ordered product of n fields and a sum of normal ordered products. This may be expressed for n even as :\begin T\left[\phi(x_1)\cdots \phi(x_n)\right]=&:\phi(x_1)\cdots \phi(x_n): +\sum_\textrm\langle 0 , T\left[\phi(x_1)\phi(x_2)\right], 0\rangle :\phi(x_3)\cdots \phi(x_n):\\ &+\sum_\textrm\langle 0 , T\left[\phi(x_1)\phi(x_2)\right], 0\rangle \langle 0 , T\left[\phi(x_3)\phi(x_4)\right], 0\rangle:\phi(x_5)\cdots \phi(x_n):\\ \vdots \\ &+\sum_\textrm\langle 0 , T\left[\phi(x_1)\phi(x_2)\right], 0\rangle\cdots \langle 0 , T\left[\phi(x_)\phi(x_n)\right], 0\rangle \end where the summation is over all the distinct ways in which one may pair up fields. The result for n odd looks the same except for the last line which reads : \sum_\text\langle 0 , T\left[\phi(x_1)\phi(x_2)\right], 0\rangle\cdots\langle 0 , T\left[\phi(x_)\phi(x_)\right], 0\rangle\phi(x_n). This theorem provides a simple method for computing vacuum expectation values of time ordered products of operators and was the motivation behind the introduction of normal ordering.


Alternative definitions

The most general definition of normal ordering involves splitting all quantum fields into two parts (for example see Evans and Steer 1996) \phi_i(x)=\phi^+_i(x)+\phi^-_i(x). In a product of fields, the fields are split into the two parts and the \phi^+(x) parts are moved so as to be always to the left of all the \phi^-(x) parts. In the usual case considered in the rest of the article, the \phi^+(x) contains only creation operators, while the \phi^-(x) contains only annihilation operators. As this is a mathematical identity, one can split fields in any way one likes. However, for this to be a useful procedure one demands that the normal ordered product of ''any'' combination of fields has zero expectation value :\langle :\phi_1(x_1)\phi_2(x_2)\ldots\phi_n(x_n):\rangle=0 It is also important for practical calculations that all the commutators (anti-commutator for fermionic fields) of all \phi^+_i and \phi^-_j are all c-numbers. These two properties means that we can apply Wick's theorem in the usual way, turning expectation values of time-ordered products of fields into products of c-number pairs, the contractions. In this generalised setting, the contraction is defined to be the difference between the time-ordered product and the normal ordered product of a pair of fields. The simplest example is found in the context of Thermal quantum field theory (Evans and Steer 1996). In this case the expectation values of interest are statistical ensembles, traces over all states weighted by \exp (-\beta \hat). For instance, for a single bosonic quantum harmonic oscillator we have that the thermal expectation value of the number operator is simply the Bose–Einstein statistics, Bose–Einstein distribution :\langle\hat^\dagger \hat\rangle = \frac = \frac So here the number operator \hat^\dagger \hat is normal ordered in the usual sense used in the rest of the article yet its thermal expectation values are non-zero. Applying Wick's theorem and doing calculation with the usual normal ordering in this thermal context is possible but computationally impractical. The solution is to define a different ordering, such that the \phi^+_i and \phi^-_j are ''linear combinations'' of the original annihilation and creations operators. The combinations are chosen to ensure that the thermal expectation values of normal ordered products are always zero so the split chosen will depend on the temperature.


References

* F. Mandl, G. Shaw, Quantum Field Theory, John Wiley & Sons, 1984. * S. Weinberg, The Quantum Theory of Fields (Volume I) Cambridge University Press (1995) * T.S. Evans, D.A. Steer
Wick's theorem at finite temperature
Nucl. Phys B 474, 481-496 (1996
arXiv:hep-ph/9601268
{{DEFAULTSORT:Normal Order Quantum field theory