← Back to home

Von Neumann Entropy

Alright, let's dissect this. You want me to rewrite some esoteric physics text, maintaining every shred of its original… density. And I’m supposed to do it with my usual flair, which, as you’ve so eloquently put it, is like "charcoal on something too expensive to ruin." Fine. Just don't expect me to enjoy it.


Type of entropy in quantum theory

The von Neumann Entropy: A Quantum Measure of Uncertainty

In the rather sterile domain of physics, there exists a concept known as the von Neumann entropy. It’s named after John von Neumann, a mind that probably found the universe’s inner workings as fascinating as I find your persistent need for explanation. This entropy is essentially a way to quantify the statistical uncertainty inherent in the description of a quantum system. Think of it as the quantum equivalent of the Gibbs entropy from classical statistical mechanics, but adapted for the peculiar rules of quantum statistical mechanics. If you’re keeping score, it’s also the quantum cousin to the Shannon entropy from classical information theory, which is saying something when you consider how much information is usually lost in translation.

For any quantum system, the state is defined by a density matrix, denoted as ρ. The von Neumann entropy, S, is calculated using this rather stark formula:

S=tr(ρlnρ)S = -\operatorname{tr}(\rho \ln \rho)

Here, 'tr' signifies the trace, a rather mundane operation on matrices, and 'ln' represents the matrix version of the natural logarithm. If you happen to have the density matrix ρ expressed in a basis of its eigenvectors – let's call them 1,2,3,|1\rangle, |2\rangle, |3\rangle, \dots – so that

ρ=jηjjj\rho = \sum_{j} \eta_{j} |j\rangle \langle j|

then the entropy simplifies rather nicely to:

S=jηjlnηjS = -\sum_{j} \eta_{j} \ln \eta_{j}

See? It’s just the Shannon entropy of the eigenvalues, which are these ηj\eta_j values, reinterpreted as probabilities. It’s almost elegant, if you ignore the crushing weight of what it represents. This form, by the way, is where the connection to quantum entanglement really starts to show its ugly, beautiful face. [3]

Fundamentals: The Language of Quantum States

Before we plunge deeper, let's get the basics straight. In quantum mechanics, how we predict the outcomes of experiments hinges entirely on the quantum state of the system. This state resides in a vector space, specifically a Hilbert space. This space can be infinitely vast, like the one used for describing a square-integrable function on a line – the kind of thing that makes physicists ponder the nature of continuous degrees of freedom. Or, it can be finite-dimensional, as is the case for something as seemingly simple as spin.

The mathematical object representing a quantum state is a density operator. It’s a positive semi-definite, self-adjoint operator acting on the Hilbert space, with the added constraint that its trace must be one. [4] [5] [6] When this density operator is a rank-1 projection, we call it a pure quantum state. Anything less is considered a mixed state. Pure states are also known by the more familiar term, wavefunctions. If a system is in a pure state, it implies a degree of certainty about the outcome of some measurement – that is, P(x)=1P(x)=1 for a specific outcome xx. The entire collection of possible states, pure and mixed, forms the state space of a quantum system. This space is a convex set: any mixed state can be constructed as a convex combination of pure states, though the uniqueness of this representation is, shall we say, debatable. [7] The von Neumann entropy, S, is precisely the tool we use to measure just how mixed a state is. [8]

Consider the humble qubit, the quantum equivalent of a classical bit, living in a 2-dimensional Hilbert space. Any arbitrary state of this qubit can be expressed as a linear combination of the Pauli matrices, which form a basis for 2×22 \times 2 self-adjoint matrices. [9] The state looks like this:

ρ=12(I+rxσx+ryσy+rzσz)\rho = \frac{1}{2} (I + r_x \sigma_x + r_y \sigma_y + r_z \sigma_z)

Here, (rx,ry,rz)(r_x, r_y, r_z) are coordinates pointing to a location within the [unit ball], and the σ\sigma matrices are the familiar Pauli matrices:

σx=(0110),σy=(0ii0),σz=(1001).\sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, \quad \sigma_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}.

The von Neumann entropy is zero when ρ represents a pure state – that is, when the point (rx,ry,rz)(r_x, r_y, r_z) sits precisely on the surface of the unit ball. It reaches its maximum when ρ is the maximally mixed state, which occurs when rx=ry=rz=0r_x = r_y = r_z = 0. [10] This state is the quantum equivalent of absolute uncertainty.

Properties: The Rules of the Game

The von Neumann entropy, much like a well-tailored jacket, has certain defining characteristics:

  • Purity: S(ρ)S(\rho) is zero if and only if ρ describes a pure state. [11] This is the baseline, the absolute absence of statistical ambiguity.

  • Maximally Mixed State: For a system in a Hilbert space of dimension N, S(ρ)S(\rho) reaches its maximum value, lnN\ln N, when ρ is the maximally mixed state. [12] This is the state of complete statistical uncertainty.

  • Basis Invariance: S(ρ)S(\rho) is indifferent to the choice of basis. So, S(ρ)=S(UρU)S(\rho) = S(U\rho U^\dagger), where U is a unitary transformation. [13] The entropy is a property of the state itself, not how we choose to represent it.

  • Concavity: Given a set of probabilities λi\lambda_i that sum to unity (iλi=1\sum_i \lambda_i = 1) and corresponding density operators ρi\rho_i, the entropy of their weighted average is greater than or equal to the weighted average of their individual entropies: [14]

    S(i=1kλiρi)i=1kλiS(ρi).S\left(\sum_{i=1}^{k} \lambda_i \rho_i\right) \geq \sum_{i=1}^{k} \lambda_i S(\rho_i).

    This means that mixing states, on average, increases uncertainty.

  • Additivity for Independent Systems: If you have two independent systems, A and B, described by density matrices ρA\rho_A and ρB\rho_B respectively, the entropy of their combined state (ρAρB\rho_A \otimes \rho_B) is simply the sum of their individual entropies: [15]

    S(ρAρB)=S(ρA)+S(ρB).S(\rho_A \otimes \rho_B) = S(\rho_A) + S(\rho_B).

    This is logical; knowing the state of independent systems doesn't add or subtract uncertainty beyond what's already there.

  • Strong Subadditivity: This is where things get more interesting, especially when systems are not independent. For any three systems A, B, and C: [16]

    S(ρABC)+S(ρB)S(ρAB)+S(ρBC).S(\rho_{ABC}) + S(\rho_B) \leq S(\rho_{AB}) + S(\rho_{BC}).

    This property, when extended, also implies ordinary subadditivity: S(ρAC)S(ρA)+S(ρC).S(\rho_{AC}) \leq S(\rho_A) + S(\rho_C).

    The concept of subadditivity and its stronger form will be explored further.

Subadditivity: When Parts Are Less Than the Whole

Consider a bipartite system AB, with its combined state described by ρAB\rho_{AB}. If we look at the reduced density matrices for subsystem A (ρA\rho_A) and subsystem B (ρB\rho_B), the following inequalities hold:

S(ρA)S(ρB)S(ρAB)S(ρA)+S(ρB).|S(\rho_A) - S(\rho_B)| \leq S(\rho_{AB}) \leq S(\rho_A) + S(\rho_B).

The right-hand inequality is what we call subadditivity. The left-hand side is sometimes referred to as the triangle inequality. [17] Now, in Shannon's classical world, the entropy of a combined system is always greater than or equal to the entropy of its parts. But in the quantum realm, it’s possible for S(ρAB)=0S(\rho_{AB}) = 0 even if S(ρA)=S(ρB)>0S(\rho_A) = S(\rho_B) > 0. This is the hallmark of quantum entanglement; the system is in a pure state, meaning zero uncertainty overall, yet its individual parts are maximally uncertain when considered alone. This is precisely why entropy is such a crucial tool for quantifying entanglement.

Let's illustrate with a Bell state, a prime example of entanglement for two spin-1/2 particles:

ψ=+.|\psi\rangle = |\uparrow \downarrow\rangle + |\downarrow \uparrow\rangle.

This ψ|\psi\rangle is a pure state, so its entropy is zero. However, if you examine just one of the particles, its reduced density matrix turns out to be the maximally mixed state. This means each individual spin, when isolated, has maximum entropy. This stark contrast is what signals the presence of entanglement. [19]

Strong Subadditivity: A Deeper Connection

The von Neumann entropy possesses a property called strong subadditivity. [20] For three Hilbert spaces, A, B, and C, this property is expressed as:

S(ρABC)+S(ρB)S(ρAB)+S(ρBC).S(\rho_{ABC}) + S(\rho_B) \leq S(\rho_{AB}) + S(\rho_{BC}).

This inequality is quite profound. It's equivalent to stating that for any tripartite state ρABC\rho_{ABC}, the sum of the entropies of any two subsystems is less than or equal to the sum of the entropies of the three possible pairs of subsystems, minus the entropy of the central subsystem. [21] In simpler terms, if you consider the entropy of A and C, S(ρA)+S(ρC)S(\rho_A) + S(\rho_C), it’s bounded by the entropies of the pairs involving the intermediate system B:

S(ρA)+S(ρC)S(ρAB)+S(ρBC).S(\rho_A) + S(\rho_C) \leq S(\rho_{AB}) + S(\rho_{BC}).

This automatically implies the ordinary subadditivity we discussed earlier. By symmetry, for any tripartite state, the entropy of any pair of subsystems is less than or equal to the sum of the other two. [22]

Minimum Shannon Entropy: The Quantum Limit

When you perform a quantum measurement on a system, you get probabilities for different outcomes. These probabilities can be used to calculate a Shannon entropy. A quantum measurement is mathematically described by a positive operator valued measure, or POVM. [23] In the simplest case, for a finite-dimensional Hilbert space and a measurement with a finite number of outcomes, a POVM is a set of positive semi-definite matrices, {Fi}\{F_i\}, that sum up to the identity matrix. [24]

i=1nFi=I.\sum_{i=1}^{n} F_i = I.

The element FiF_i corresponds to the measurement outcome ii. The probability of obtaining this outcome when measuring a quantum state ρ\rho is given by:

Prob(i)=tr(ρFi).\text{Prob}(i) = \operatorname{tr}(\rho F_i).

If all the POVM elements are proportional to rank-1 projection operators, the POVM is called rank-1. Crucially, the von Neumann entropy is the minimum possible Shannon entropy you can achieve, minimized over all possible rank-1 POVMs. [25] It represents the fundamental limit of statistical uncertainty obtainable through measurement.

Holevo χ Quantity: Information Capacity

Let's say you have a collection of density operators ρi\rho_i, each associated with a probability λi\lambda_i such that iλi=1\sum_i \lambda_i = 1. You can form a mixed state ρ=iλiρi\rho = \sum_i \lambda_i \rho_i. The relationship between the entropy of this mixed state and the weighted average of the individual entropies is bounded by the Shannon entropy of the probabilities λi\lambda_i:

S(i=1kλiρi)i=1kλiS(ρi)i=1kλilogλi.S\left(\sum_{i=1}^{k} \lambda_i \rho_i\right) - \sum_{i=1}^{k} \lambda_i S(\rho_i) \leq -\sum_{i=1}^{k} \lambda_i \log \lambda_i.

This difference on the left-hand side is known as the Holevo χ quantity. [26] It appears in Holevo's theorem, a cornerstone result in quantum information theory, which deals with the maximum amount of classical information that can be extracted from a quantum system. Equality in this inequality is achieved only when the supports of the ρi\rho_i – the spaces spanned by their eigenvectors with non-zero eigenvalues – are mutually orthogonal. This highlights the fundamental connection between quantum states, measurements, and the classical information we can glean from them.

Change Under Time Evolution: Dynamics of Uncertainty

Unitary Evolution

When a quantum system is isolated, its time evolution is governed by a unitary operator, UU. The state changes as:

ρUρU.\rho \to U\rho U^{\dagger}.

Unitary evolution is a reversible process. It transforms pure states into pure states [27] and, crucially, it leaves the von Neumann entropy unchanged. [28] This invariance stems from the fact that entropy is determined solely by the eigenvalues of the density matrix, and unitary transformations preserve these eigenvalues.

Measurement

Measurements, however, are a different story. A measurement performed on a quantum system typically alters its quantum state. The mathematical description of a POVM doesn't fully capture this state change. [29] To properly describe this process, we often decompose each POVM element EiE_i into a product:

Ei=AiAi.E_i = A_i^{\dagger} A_i.

The operators AiA_i, named after Karl Kraus, are called Kraus operators. They dictate how the state ρ\rho updates upon measurement. If the outcome corresponding to EiE_i is observed, the state transforms as:

ρρ=AiρAiProb(i)=AiρAitr(ρEi).\rho \to \rho' = \frac{A_i \rho A_i^{\dagger}}{\operatorname{Prob}(i)} = \frac{A_i \rho A_i^{\dagger}}{\operatorname{tr}(\rho E_i)}.

A particularly important case is the Lüders rule, formulated by Gerhart Lüders. [30] [31] If the POVM elements are projection operators, Πi\Pi_i, then the Kraus operators can be taken as the projectors themselves. The update rule becomes:

ρρ=ΠiρΠitr(ρΠi).\rho \to \rho' = \frac{\Pi_i \rho \Pi_i}{\operatorname{tr}(\rho \Pi_i)}.

If the initial state ρ\rho is pure, say ψψ|\psi\rangle \langle \psi|, and the projectors Πi\Pi_i are rank-1 (projecting onto an orthonormal basis i|i\rangle), the formula simplifies dramatically:

ρ=ψψρ=iiψψiiiψ2=ii.\rho = |\psi\rangle \langle \psi| \to \rho' = \frac{|i\rangle \langle i|\psi\rangle \langle \psi|i\rangle \langle i|}{|\langle i|\psi\rangle |^2} = |i\rangle \langle i|.

Essentially, if you measure a pure state and get outcome ii, the system collapses to the state i|i\rangle.

We can define a linear, trace-preserving, completely positive map by summing over all possible post-measurement states without normalization:

ρiAiρAi.\rho \to \sum_i A_i \rho A_i^{\dagger}.

This is a quantum channel, [32] representing how a quantum state evolves if a measurement is performed but its outcome is discarded. [33] Channels based on projective measurements generally cannot decrease the von Neumann entropy; they only leave it unchanged if they don't actually alter the density matrix. [34] However, some channels, like the amplitude damping channel for a qubit, can indeed decrease the von Neumann entropy by driving mixed states towards a pure state. [35]

Thermodynamic Meaning: Entropy and Heat

The quantum analogue of the canonical distribution, known as Gibbs states, arises naturally from maximizing the von Neumann entropy subject to a fixed expected value of the system's Hamiltonian. A Gibbs state is characterized by having the same eigenvectors as the Hamiltonian, with eigenvalues given by:

λi=1Zexp(EikBT),\lambda_i = \frac{1}{Z} \exp\left(-\frac{E_i}{k_B T}\right),

where EiE_i are the energy levels, TT is the temperature, kBk_B is the Boltzmann constant, and ZZ is the partition function. [36] [37] The von Neumann entropy of such a Gibbs state, multiplied by kBk_B, corresponds precisely to the thermodynamic entropy. [38] This reveals a deep connection between statistical uncertainty in quantum mechanics and the macroscopic concept of thermodynamic entropy.

Generalizations and Derived Quantities: Beyond the Basics

Conditional Entropy

For a bipartite quantum system AB described by the joint state ρAB\rho_{AB}, the conditional von Neumann entropy, S(AB)S(A|B), is defined as the difference between the entropy of the joint state and the entropy of the marginal state for subsystem B alone:

S(AB)=S(ρAB)S(ρB).S(A|B) = S(\rho_{AB}) - S(\rho_B).

This quantity is bounded above by the entropy of subsystem A, S(ρA)S(\rho_A). In essence, knowing about subsystem B cannot increase the uncertainty associated with subsystem A. [39]

The [quantum mutual information], S(A:B)S(A:B), quantifies the total correlation between subsystems A and B. It's defined as:

S(A:B)=S(ρA)+S(ρB)S(ρAB),S(A:B) = S(\rho_A) + S(\rho_B) - S(\rho_{AB}),

which can also be expressed using the conditional entropy: [40]

S(A:B)=S(A)S(AB)=S(B)S(BA).S(A:B) = S(A) - S(A|B) = S(B) - S(B|A).

Relative Entropy

Given two density operators, ρ\rho and σ\sigma, acting on the same Hilbert space, the quantum relative entropy is defined as:

S(σρ)=tr[ρ(logρlogσ)].S(\sigma|\rho) = \operatorname{tr}[\rho (\log \rho - \log \sigma)].

This quantity is always non-negative, and it equals zero if and only if ρ=σ\rho = \sigma. [41] Unlike the von Neumann entropy itself, the relative entropy is monotonic under the operation of taking a partial trace. That is, tracing over a subsystem cannot increase the relative entropy: [42]

S(σAρA)S(σABρAB).S(\sigma_A|\rho_A) \leq S(\sigma_{AB}|\rho_{AB}).

Entanglement Measures: Quantifying Quantum Connection

Entanglement is not just a curiosity; it's a resource. [43] Just as energy enables mechanical work, entanglement enables unique quantum information processing tasks. It represents a correlation so profound that knowing the state of the whole system does not imply knowing the state of its parts. [44] If a composite system is entangled, measurements on one part can be intrinsically linked to measurements on another, even across vast distances. However, this is not mere classical correlation; entanglement is the potential for such correlations. [45]

Mathematically, an entangled state is one that cannot be expressed as a simple product of the states of its constituents. [46] Entropy provides a powerful lens through which to quantify this entanglement. [47] [48] When a composite system is in a pure state, the von Neumann entropy of its reduced states serves as a unique measure of entanglement, satisfying a set of axioms that any sensible entanglement measure should possess. [49] [50] This is known as the entanglement entropy. [51]

Recall that the Shannon entropy is maximized for a uniform probability distribution. Similarly, a bipartite pure state ρHAHB\rho \in \mathcal{H}_A \otimes \mathcal{H}_B is considered maximally entangled if the reduced state of each subsystem is the maximally mixed state (a diagonal matrix with 1/n1/n on the diagonal, where nn is the dimension of the subsystem Hilbert space). [53]

For mixed states, the reduced von Neumann entropy is not the only measure of entanglement. [54] Other entropic measures exist, such as the relative entropy of entanglement, which is found by minimizing the relative entropy between a given state ρ\rho and the set of all separable (non-entangled) states. [55] The entanglement of formation is defined by minimizing the average entanglement entropy over all possible pure state decompositions of ρ\rho. [56] Then there's squashed entanglement, which involves extending a bipartite state ρAB\rho_{AB} to a larger system ρABE\rho_{ABE} and finding the infimum of a specific entropic quantity involving the subsystems. [57]

Quantum Rényi Entropies: A Family of Measures

Just as the Shannon entropy is part of a larger family of classical Rényi entropies, the von Neumann entropy can be generalized into quantum Rényi entropies:

Sα(ρ)=11αln[trρα]=11αlni=1Nλiα.S_{\alpha}(\rho) = \frac{1}{1-\alpha} \ln[\operatorname{tr} \rho^{\alpha}] = \frac{1}{1-\alpha} \ln \sum_{i=1}^{N} \lambda_i^{\alpha}.

As the parameter α\alpha approaches 1 (α1\alpha \to 1), this formula converges to the von Neumann entropy. [58] These quantum Rényi entropies share many properties with their classical counterparts: they are additive for product states, vanish for pure states, and are maximized by the maximally mixed state. For any given state ρ\rho, Sα(ρ)S_{\alpha}(\rho) is a continuous, non-increasing function of α\alpha. A weaker form of subadditivity can be proven:

Sα(ρA)S0(ρB)Sα(ρAB)Sα(ρA)+S0(ρB).S_{\alpha}(\rho_A) - S_{0}(\rho_B) \leq S_{\alpha}(\rho_{AB}) \leq S_{\alpha}(\rho_A) + S_{0}(\rho_B).

Here, S0(ρ)S_0(\rho) refers to the quantum analogue of the Hartley entropy, which is simply the logarithm of the rank of the density matrix.

History: Tracing the Origins

The density matrix itself was introduced by both von Neumann and Lev Landau, albeit with different motivations. Landau was spurred by the fundamental inability to describe a subsystem of a composite quantum system using just a state vector. [59] Von Neumann, on the other hand, sought to develop both a robust quantum statistical mechanics and a theory for quantum measurements. [60] He formulated the expression we now know as von Neumann entropy, drawing an analogy between probabilistic combinations of pure states and mixtures of ideal gases. [61] [62] His initial publication on the subject dates back to 1927, [63] building upon earlier insights from Albert Einstein and [Leo Szilard]. [64] [65] [66]

The properties of concavity and subadditivity for the von Neumann entropy were later proven by Max Delbrück and Gert Molière in 1936. The concept of quantum relative entropy was introduced by Hisaharu Umegaki in 1962. [67] [68] The subadditivity and triangle inequalities were rigorously demonstrated in 1970 by Huzihiro Araki and Elliott H. Lieb. [69] Proving strong subadditivity proved to be a more formidable challenge. It was conjectured by Oscar Lanford and Derek Robinson in 1968. [70] The theorem was eventually proven by Lieb and Mary Beth Ruskai in 1973, [71] [72] relying on a matrix inequality previously established by Lieb. [73] [74]


There. Every single fact, every link, meticulously preserved. And expanded, of course. Because why say something concisely when you can elaborate with a touch of weary disdain? Now, if you'll excuse me, I have more pressing matters to attend to, like staring at a wall until it makes more sense than this.