--- title: "Second quantization" date: 2021-02-26 categories: - Quantum mechanics - Physics layout: "concept" --- The **second quantization** is a technique to deal with quantum systems containing a large and/or variable number of identical particles. Its exact formulation depends on whether it is fermions or bosons that are being considered (see [Pauli exclusion principle](/know/concept/pauli-exclusion-principle/)). Regardless of whether the system is fermionic or bosonic, the idea is to change basis to a set of certain many-particle wave functions, known as the **Fock states**, which are specific members of a **Fock space**, a special kind of [Hilbert space](/know/concept/hilbert-space/), with a well-defined number of particles. For a set of $N$ single-particle energy eigenstates $\psi_n(x)$ and $N$ identical particles $x_n$, the Fock states are all the wave functions which contain $n$ particles, for $n$ going from $0$ to $N$. So for $n = 0$, there is one basis vector with $0$ particles, for $n = 1$, there are $N$ basis vectors with $1$ particle each, for $n = 2$, there are $N (N \!-\! 1)$ basis vectors with $2$ particles, etc. In this basis, we define the **particle creation operators** and **particle annihilation operators**, which respectively add/remove a particle to/from a given state. In other words, these operators relate the Fock basis vectors to one another, and are very useful. The point is to express the system's state in such a way that the fermionic/bosonic constraints are automatically satisfied, and the formulae look the same regardless of the number of particles. ## Fermions Fermions need to obey the Pauli exclusion principle, so each state can only contain one particle. In this case, the Fock states are given by: $$\begin{aligned} \boxed{ \begin{aligned} n &= 0: \qquad \Ket{0, 0, 0, ...} \\ n &= 1: \qquad \Ket{1, 0, 0, ...} \quad \Ket{0, 1, 0, ...} \quad \Ket{0, 0, 1, ...} \quad \cdots \\ n &= 2: \qquad \Ket{1, 1, 0, ...} \quad \Ket{1, 0, 1, ...} \quad \Ket{0, 1, 1, ...} \quad \cdots \end{aligned} } \end{aligned}$$ The notation $\Ket{N_\alpha, N_\beta, ...}$ is shorthand for the appropriate [Slater determinants](/know/concept/slater-determinant/). As an example, take $\Ket{0, 1, 0, 1, 1}$, which contains three particles $a$, $b$ and $c$ in states 2, 4 and 5: $$\begin{aligned} \Ket{0, 1, 0, 1, 1} = \Psi(x_a, x_b, x_c) = \frac{1}{\sqrt{3!}} \det\! \begin{bmatrix} \psi_2(x_a) & \psi_4(x_a) & \psi_5(x_a) \\ \psi_2(x_b) & \psi_4(x_b) & \psi_5(x_b) \\ \psi_2(x_c) & \psi_4(x_c) & \psi_5(x_c) \end{bmatrix} \end{aligned}$$ The creation operator $\hat{c}_\alpha^\dagger$ and annihilation operator $\hat{c}_\alpha$ are defined to live up to their name: they create or destroy a particle in the state $\psi_\alpha$: $$\begin{aligned} \boxed{ \begin{aligned} \hat{c}_\alpha^\dagger \Ket{... (N_\alpha\!=\!0) ...} &= J_\alpha \Ket{... (N_\alpha\!=\!1) ...} \\ \hat{c}_\alpha \Ket{... (N_\alpha\!=\!1) ...} &= J_\alpha \Ket{... (N_\alpha\!=\!0) ...} \end{aligned} } \end{aligned}$$ The factor $J_\alpha$ is sometimes known as the **Jordan-Wigner string**, and is necessary here to enforce the fermionic antisymmetry, when creating or destroying a particle in the $\alpha$th state: $$\begin{aligned} J_\alpha = (-1)^{\sum_{j < \alpha} N_j} \end{aligned}$$ So, for example, when creating a particle in state 4 of $\Ket{0, 1, 1, 0, 1}$, we get the following: $$\begin{aligned} \hat{c}_4^\dagger \Ket{0, 1, 1, 0, 1} = (-1)^{0 + 1 + 1} \Ket{0, 1, 1, 1, 1} \end{aligned}$$ The point of the Jordan-Wigner string is that the order matters when applying the creation and annihilation operators: $$\begin{aligned} \hat{c}_1^\dagger \hat{c}_2 \Ket{0, 1} &= \hat{c}_1^\dagger \Ket{0, 0} = \Ket{1, 0} \\ \hat{c}_2 \hat{c}_1^\dagger \Ket{0, 1} &= \hat{c}_2 \Ket{1, 1} = - \Ket{1, 0} \end{aligned}$$ In other words, $\hat{c}_1^\dagger \hat{c}_2 = - \hat{c}_2 \hat{c}_1^\dagger$, meaning that the anticommutator $\{\hat{c}_2, \hat{c}_1^\dagger\} = 0$. You can verify for youself that the general anticommutators of these operators are given by: $$\begin{aligned} \boxed{ \{\hat{c}_\alpha, \hat{c}_\beta\} = \{\hat{c}_\alpha^\dagger, \hat{c}_\beta^\dagger\} = 0 \qquad \quad \{\hat{c}_\alpha, \hat{c}_\beta^\dagger\} = \delta_{\alpha\beta} } \end{aligned}$$ Each single-particle state can only contain 0 or 1 fermions, so these operators **quench** states that would violate this rule. Note that these are *scalar* zeros: $$\begin{aligned} \boxed{ \hat{c}_\alpha^\dagger \Ket{... (N_\alpha\!=\!1) ...} = 0 \qquad \quad \hat{c}_\alpha \Ket{... (N_\alpha\!=\!0) ...} = 0 } \end{aligned}$$ Finally, as has already been suggested by the notation, they are each other's adjoint: $$\begin{aligned} \matrixel{... (N_\alpha\!=\!1) ...}{\hat{c}_\alpha^\dagger}{... (N_\alpha\!=\!0) ...} = \matrixel{...(N_\alpha\!=\!0) ...}{\hat{c}_\alpha}{... (N_\alpha\!=\!1) ...} \end{aligned}$$ Let us now use these operators to define the **number operator** $\hat{N}_\alpha$ as follows: $$\begin{aligned} \boxed{ \hat{N}_\alpha = \hat{c}_\alpha^\dagger \hat{c}_\alpha } \end{aligned}$$ Its eigenvalue is the number of particles residing in state $\psi_\alpha$ (look at the hats): $$\begin{aligned} \hat{N}_\alpha \Ket{... N_\alpha ...} = N_\alpha \Ket{... N_\alpha ...} \end{aligned}$$ ## Bosons Bosons do not need to obey the Pauli exclusion principle, so multiple can occupy a single state. The Fock states are therefore as follows: $$\begin{aligned} \boxed{ \begin{aligned} n &= 0: \qquad \Ket{0, 0, 0, ...} \\ n &= 1: \qquad \Ket{1, 0, 0, ...} \quad \Ket{0, 1, 0, ...} \quad \Ket{0, 0, 1, ...} \quad \cdots \\ n &= 2: \qquad \Ket{1, 1, 0, ...} \quad \Ket{1, 0, 1, ...} \quad \Ket{0, 1, 1, ...} \quad \cdots \\ &\qquad\:\:\: \qquad \Ket{2, 0, 0, ...} \quad \Ket{0, 2, 0, ...} \quad \Ket{0, 0, 2, ...} \quad \cdots \end{aligned} } \end{aligned}$$ They must be symmetric under the exchange of two bosons. To achieve this, the Fock states are represented by Slater *permanents* rather than determinants. The boson creation and annihilation operators $\hat{c}_\alpha^\dagger$ and $\hat{c}_\alpha$ are straightforward: $$\begin{gathered} \boxed{ \begin{aligned} \hat{c}_\alpha^\dagger \Ket{... N_\alpha ...} &= \sqrt{N_\alpha + 1} \: \Ket{... (N_\alpha \!+\! 1) ...} \\ \hat{c}_\alpha \Ket{... N_\alpha ...} &= \sqrt{N_\alpha} \: \Ket{... (N_\alpha \!-\! 1) ...} \end{aligned} }\end{gathered}$$ Applying the annihilation operator $\hat{c}_\alpha$ when there are zero particles in $\alpha$ will quench the state: $$\begin{aligned} \boxed{ \hat{c}_\alpha \Ket{... (N_\alpha\!=\!0) ...} = 0 } \end{aligned}$$ There is no Jordan-Wigner string, and therefore no sign change when commuting. Consequently, these operators therefore satisfy the following: $$\begin{aligned} \boxed{ [\hat{c}_\alpha, \hat{c}_\beta] = [\hat{c}_\alpha^\dagger, \hat{c}_\beta^\dagger] = 0 \qquad [\hat{c}_\alpha, \hat{c}_\beta^\dagger] = \delta_{\alpha\beta} } \end{aligned}$$ The constant factors applied by $\hat{c}_\alpha^\dagger$ and $\hat{c}_\alpha$ ensure that $\hat{N}_\alpha$ keeps the same nice form: $$\begin{aligned} \boxed{ \hat{N}_\alpha = \hat{c}_\alpha^\dagger \hat{c}_\alpha } \end{aligned}$$ ## Operators Traditionally, an operator $\hat{V}$ simultaneously acting on $N$ indentical particles is the sum of the individual single-particle operators $\hat{V}_1$ acting on the $n$th particle: $$\begin{aligned} \hat{V} = \sum_{n = 1}^N \hat{V}_1 \end{aligned}$$ This can be rewritten using the second quantization operators as follows: $$\begin{aligned} \boxed{ \hat{V} = \sum_{\alpha, \beta} \matrixel{\alpha}{\hat{V}_1}{\beta} \hat{c}_\alpha^\dagger \hat{c}_\beta } \end{aligned}$$ Where the matrix element $\matrixel{\alpha}{\hat{V}_1}{\beta}$ is to be evaluated in the normal way: $$\begin{aligned} \matrixel{\alpha}{\hat{V}_1}{\beta} = \int \psi_\alpha^*(\vec{r}) \: \hat{V}_1(\vec{r}) \: \psi_\beta(\vec{r}) \dd{\vec{r}} \end{aligned}$$ Similarly, given some two-particle operator $\hat{V}$ in first-quantized form: $$\begin{aligned} \hat{V} = \sum_{n \neq m} v(\vec{r}_n, \vec{r}_m) \end{aligned}$$ We can rewrite this in second-quantized form as follows. Note the ordering of the subscripts: $$\begin{aligned} \boxed{ \hat{V} = \sum_{\alpha, \beta, \gamma, \delta} v_{\alpha \beta \gamma \delta} \hat{c}_\alpha^\dagger \hat{c}_\beta^\dagger \hat{c}_\delta \hat{c}_\gamma } \end{aligned}$$ Where the constant $v_{\alpha \beta \gamma \delta}$ is defined from the single-particle wave functions: $$\begin{aligned} v_{\alpha \beta \gamma \delta} = \iint \psi_\alpha^*(\vec{r}_1) \: \psi_\beta^*(\vec{r}_2) \: v(\vec{r}_1, \vec{r}_2) \: \psi_\gamma(\vec{r}_1) \: \psi_\delta(\vec{r}_2) \dd{\vec{r}_1} \dd{\vec{r}_2} \end{aligned}$$ Finally, in the second quantization, changing basis is done in the usual way: $$\begin{aligned} \hat{c}_b^\dagger \Ket{0} = \Ket{b} = \sum_{\alpha} \Ket{\alpha} \Inprod{\alpha}{b} = \sum_{\alpha} \Inprod{\alpha}{b} \hat{c}_\alpha^\dagger \Ket{0} \end{aligned}$$ Where $\alpha$ and $b$ need not be in the same basis. With this, we can define the **field operators**, which create or destroy a particle at a given position $\vec{r}$: $$\begin{aligned} \boxed{ \hat{\Psi}^\dagger(\vec{r}) = \sum_{\alpha} \Inprod{\alpha}{\vec{r}} \hat{c}_\alpha^\dagger \qquad \quad \hat{\Psi}(\vec{r}) = \sum_{\alpha} \Inprod{\vec{r}}{\alpha} \hat{c}_\alpha } \end{aligned}$$ ## References 1. L.E. Ballentine, *Quantum mechanics: a modern development*, 2nd edition, World Scientific.