Formalism of quantum statistical mechanics#
(Also see Kardar 6.4 (Quantum microstates) and Kardar 6.5 (Quantum macrostates))
Quantum Analog of a Distribution Over Microstates: The Density Matrix#
In classical statistical mechanics, we describe a system by a probability distribution over microstates. In quantum mechanics, the analogous object is the density matrix.
Suppose a quantum system is in a pure state \(\left|\psi_{\alpha}\right\rangle\) with probability \(p_{\alpha}\). The system’s statistical state is then described by:
This matrix satisfies:
The expectation value of an observable \(\hat{O}\) in this state is given by:
To derive the trace form, consider:
The density matrix can be represented in any orthonormal basis \({|i\rangle}\):
Note
Conditions for a valid density matrix:
Trace one: \(\operatorname{Tr}(\hat{\rho}) = 1\)
Hermitian: \(\hat{\rho}^\dagger = \hat{\rho}\)
Positive semi-definite: For all \(|v\rangle\),
In a general basis, the density matrix can be understood as having two distinct components: populations and coherences.
The populations are represented by the diagonal elements of the density matrix, \(p_i\equiv \rho_{ii}\). These elements behave like probabilities, satisfying \(0 \leq p_i \leq 1\) and \(\sum_i p_i = 1\). They describe the likelihood of the system being in a particular quantum state associated with the chosen basis. But note that the probabilistic interpretation of the density matrix depends on the chosen basis.
The coherences are represented by the off-diagonal elements, \(\rho_{ij}\) for \(i \neq j\). These elements capture the quantum interference effects between different basis states, reflecting the phase relationships and superposition properties of the quantum system.
The structure of the density matrix in a general basis can be visualized as:
Here, the diagonal elements \(p_i=\rho_{ii}\) correspond to the populations, while the off-diagonal elements \(\rho_{ij}\) (\(i \neq j\)) represent the coherences. The populations are basis-dependent, and in certain special bases (e.g., the eigenbasis of \(\hat{\rho}\)), the coherences vanish, leaving a diagonal matrix.
This decomposition into populations and coherences provides a useful framework for understanding the physical meaning of the density matrix in quantum mechanics.
Note
Example: Single Spin System
Consider a single spin system. Let us examine how the density matrix transforms when we change the basis from the \(|\uparrow\rangle, |\downarrow\rangle\) (up/down) basis to the \(|\rightarrow\rangle, |\leftarrow\rangle\) (left/right) basis.
In the up/down basis, the density matrix is given by:
This represents a completely mixed state, where the probabilities of being in the \(|\uparrow\rangle\) and \(|\downarrow\rangle\) states are equal.
To change to the left/right basis, we use the fact that the \(|\rightarrow\rangle\) and \(|\leftarrow\rangle\) states are related to the \(|\uparrow\rangle\) and \(|\downarrow\rangle\) states by the following transformations:
The transformation matrix \(U\) that relates the two bases is:
To express the density matrix in the right/left basis, we perform the basis transformation:
Substituting \(\hat{\rho}_{\text{up/down}}\) and \(U\):
Carrying out the matrix multiplication:
This result shows that in the right/left basis, the density matrix becomes diagonal, with the system entirely in the \(|\leftarrow\rangle\) state. The coherence terms present in the up/down basis vanish in this new basis.
There always exists a special basis, namely the eigenbasis of \(\hat{\rho}\), in which the density matrix becomes diagonal:
In this eigenbasis, the diagonal elements \(p_i\) can be interpreted as the “populations” of the corresponding quantum states, such as energy levels or other relevant states.
Dynamics#
Exercise: Show that the Schrödinger equation
leads to the von Neumann equation:
From this, we can deduce:
(since \(\operatorname{Tr}([A, B]) = 0\) for any operators \(A\) and \(B\)).
Thus, the trace condition \(\operatorname{Tr}(\hat{\rho}) = 1\) is preserved under unitary time evolution.
For a time-independent Hamiltonian, the state evolves as \(\psi(t) = \hat{U}(t) \psi(0)\), where:
The corresponding evolution of the density matrix is:
For systems interacting with an environment, the evolution is no longer unitary, and the dynamics are described by the Lindblad equation, which incorporates aspects of a master equation.
Density matrices in equilbrium.#
The various equilibrium ensembles are then taken to be diagonal in the \(\hat{H}\) basis:
(Note that \([\hat{H}, \hat{N}]=0\)). The density matrices can be conveniently summarized as operator expressions, eg.
This points to an interesting observation: with
Thus the thermal density matrix is formally equivalent to “imaginary time evolution”,
This mathematical coincidence (28) gives rise to a number of technical tricks (Wick rotation | Euclidean path integral) important to quantum field and many body theory.
For an ensemble like \(\hat{H}-\mu \hat{N}\), where \([\hat{H}, \hat{N}]=0\), it is easy to show that
Note that \(Z\) is just a number (as in classical stat mech) in contrast to \(\hat{H}\) and \(\hat{\rho}\).
So, fortunately, all our thermo results like \(F=-\frac{1}{\beta} \ln (Z), \mu=-\frac{\partial F}{\partial N}\), etc and Maxwell’s relations / thermodynamic inequalities are unchanged by quantum effects. It’s just harder to compute \(Z\) because we need eigenspectrom \(\hat{H} \longrightarrow E_{n}\), which is itself hard!
The Many-Body Hilbert Space#
The sum \(\sum_{n} e^{-\beta E_{n}}\left|E_{n}\right\rangle\left\langle E_{n}\right|\) spans the Hilbert space \(\left|E_{n}\right\rangle \in \mathcal{H}\). To understand this better in the context of many-body systems, let us explore the structure of \(\mathcal{H}\) in three common cases:
Spins/qubits: States like \(|\uparrow \uparrow \downarrow \uparrow \cdots\rangle\).
Bosons: Symmetric wavefunctions, \(\psi(x_1, x_2) = \psi(x_2, x_1)\).
Fermions: Antisymmetric wavefunctions, \(\psi(x_1, x_2) = -\psi(x_2, x_1)\).
We will focus on (1) for now, with (2) and (3) covered in future lectures.
Spins and Qubits#
The Hilbert space for a single \(s = 1/2\) spin is:
A general state is written as:
In quantum information, this is referred to as a “qubit,” with basis states \(\{|0\rangle, |1\rangle\}\). Operators acting on this space are spanned by the \(2 \times 2\) Pauli matrices \(\sigma^{x/y/z}\), often denoted as \(X, Y, Z\) in quantum information.
For a spin-\(S\) system, the Hilbert space generalizes to:
This is referred to as a “qudit,” where \(d = 2S + 1\).
Two Spins#
For two spins, the Hilbert space is:
This is a tensor product space:
\(N\) Spins#
For \(N\) spins, the Hilbert space extends as:
The dimension of \(\mathcal{H}^{(N)}\) is \(2^N\), not \(2 \times N\). A general state is:
This requires \(2^N\) complex coefficients \(\psi_{\{\sigma\}}\), which grows exponentially with \(N\).
Classical vs. Quantum Storage#
In contrast, a classical Ising model microstate \(\mu = \sigma_1 \sigma_2 \cdots \sigma_N\) requires only \(N\) classical bits. For example, simulating a \(20 \times 20 = 400\) spin system in classical Monte Carlo requires less than 1 KB of RAM. However, storing the quantum state \(\psi_{\{\sigma\}}\) with 128-bit precision would require:
This demonstrates the exponential difficulty of simulating quantum systems with classical resources.
Many-Body Operators#
The tensor product structure also applies to operators:
For the Pauli matrix \(\hat{X} = \begin{bmatrix} 0 & 1 \\ 1 & 0 \end{bmatrix}\), with \(\sigma \in \{1, -1\}\), we have:
The operators \(\hat{Z}_{i}\) and \(\hat{X}_{i}\) are \(2^N \times 2^N\) matrices. For example, for \(N = 2\), the Hilbert space is \(\{|1, 1\rangle, |1, -1\rangle, |-1, 1\rangle, |-1, -1\rangle\}\), and the operators are:
Generic \(2^N \times 2^N\) operators can be constructed by combining \(\hat{Z}_{i}\) and \(\hat{X}_{i}\) through addition or multiplication.
Example: Transverse-Field Ising Model
The Hamiltonian for the transverse-field Ising model is:
and the magnetization operator is:
Typical Problem:
Given the thermal density matrix \(\hat{\rho}_{\beta} = \frac{e^{-\beta \hat{H}_{\text{TFI}}}}{Z}\), compute the expectation value \(\operatorname{Tr}(\hat{m} \hat{\rho}_{\beta})\).
This is challenging because calculating \(\hat{\rho}_{\beta}\) requires diagonalizing the \(2^N \times 2^N\) Hamiltonian matrix \(\hat{H}_{\text{TFI}}\). Current computational techniques can handle systems with \(N \sim 20\) spins, as discussed in the context of “quantum supremacy.”
Reduced Density Matrices and Entanglement#
In classical probability theory, a joint distribution over two random variables, \(\mu = \{\mu_1, \mu_2\}\), is given by \(p(\mu_1, \mu_2)\). From this, we can define a marginal distribution for one variable:
If we are only interested in observables \(O_1(\mu_1)\) (rather than \(O(\mu_1, \mu_2)\)), we can compute their expectation value using the marginal distribution:
Thus, the marginal \(p_1\) is sufficient to recover \(\langle O_1 \rangle\). However, in general:
This indicates that \(\mu_1\) and \(\mu_2\) are “correlated,” leading to \(\langle O_1 O_2 \rangle \neq \langle O_1 \rangle \langle O_2 \rangle\).
For example, the distribution:
does not factorize, whereas:
does factorize as \(p_1(\sigma_1) p_2(\sigma_2) \ldots\).
In quantum mechanics, for a system on \(\mathcal{H} = \mathcal{H}_1 \otimes \mathcal{H}_2\), the equivalent of a marginal distribution is the Reduced Density Matrix.
Let \(\operatorname{span}\{\left|i_1\right\rangle\} = \mathcal{H}_1 = \mathbb{C}^{D_1}\) and \(\operatorname{span}\{\left|i_2\right\rangle\} = \mathcal{H}_2 = \mathbb{C}^{D_2}\). Then:
A general quantum state can be expressed as:
The corresponding density matrix is:
For an observable \(\hat{O} = \hat{O}_1 \otimes \hat{\mathbb{I}}\), which acts only on subsystem 1, we have:
This effectively traces out subsystem 2.
Note
For instance, in a two-spin system \(\mathcal{H} = \mathbb{C}^2 \otimes \mathbb{C}^2\), we might physically separate them to different regions of the lab, and then use a Stern-Gerlach aparatus to measure only spin 1. This is equivalent to projecting onto an operator \(\hat{G} = \hat{Z}_1 \otimes \hat{\mathbb{I}}_2\).
The expectation value of the observable is:
Defining a \(D_1 \times D_1\) matrix:
and the corresponding operator:
we find:
This shows that the “partial trace” \(\hat{\rho}^{(1)} = \operatorname{Tr}_2(\hat{\rho})\) serves as the quantum analog of a marginal distribution, known as the Reduced Density Matrix for subsystem 1.
Note
Example: EPR State Consider the EPR state:
The density matrix is:
The reduced density matrix for subsystem 1 is:
Note:
\(\hat{\rho}_1\) is mixed, even though \(\hat{\rho}\) is pure!!
\(\langle X_1 \rangle = \langle Y_1 \rangle = \langle Z_1 \rangle = 0\) despite the fact that \(\langle Z_1 Z_2 \rangle = -1\). This shows that even though the subsystem for themselves look completely random, the two subsystems are strongly correlated (entangled) when considered together. the two EPR spins are (anti-)correlated, even though the overall system is in a definite (pure) state. The randomness of the individual subsystems does not contradict the fact that the overall system is in a definite (pure) state. “Local randomness with global order” is a hallmark of quantum entanglement.
Entropy#
The entropy of a quantum state is defined as:
For a density matrix \(\hat{\rho} = \sum_\alpha p_\alpha |\alpha\rangle \langle\alpha|\), where \(\langle\alpha | \beta\rangle = \delta_{\alpha \beta}\), this simplifies to:
A state \(\hat{\rho}\) is called pure if:
Otherwise, if \(S[\hat{\rho}] > 0\), the state \(\hat{\rho}\) is considered mixed.
Entanglement#
Consider a composite system with Hilbert space \(\mathcal{H} = \mathcal{H}_1 \otimes \mathcal{H}_2\), and let \(|\psi\rangle \in \mathcal{H}\) be a pure state. While the total state \(\hat{\rho} = |\psi\rangle \langle\psi|\) has \(S[\hat{\rho}] = 0\), the reduced density matrix (RDM) for a subsystem, \(\hat{\rho}_1 = \operatorname{Tr}_2(\hat{\rho})\), can be mixed.
The entropy of the RDM, \(S[\hat{\rho}_1]\), is referred to as the entanglement entropy. This measures the degree of quantum entanglement between the two subsystems.
Even though the overall system is in a pure state, measurements on subsystems \(1\) and \(2\) can exhibit correlations:
For example, in the EPR state \(|\psi\rangle = |EPR\rangle\), the reduced density matrix for one subsystem is:
This gives an entanglement entropy of \(S = \ln(2)\), corresponding to one bit of entanglement.
Note
For more insights, check out Matthew Fisher’s recent colloquium on phase transitions in entanglement entropy.
Measures of Quantum Information#
Many concepts from classical information theory extend to the quantum domain, though not all do.
For a composite Hilbert space \(\mathcal{H}_{A} \otimes \mathcal{H}_{B} = \mathcal{H}_{AB}\), with the reduced density matrix \(\hat{\rho}_{A} = \operatorname{Tr}_{B}(\hat{\rho}_{AB})\), the mutual information is defined as a measure of total correlation between subsystems \(A\) and \(B\):
The mutual information is zero if and only if the joint state factorizes:
An important property of mutual information is its relationship to the strong subadditivity of quantum entropy:
The difference \(I(A: BC) - I(A: B)\) can be expressed as:
This inequality, known as the strong subadditivity of quantum entropy, is a fundamental result in quantum information theory, though its proof is non-trivial. For more details, see the Wikipedia article on strong subadditivity of quantum entropy.