thanks: (Current affiliation: Norma Inc.)thanks: Corresponding author

Bounding quantum uncommon information with quantum neural estimators

Donghwa Ji [email protected] College of Liberal Studies, Seoul National University, Seoul 08826, Korea Team QST, Seoul National University, Seoul 08826, Korea    Junseo Lee [email protected] Team QST, Seoul National University, Seoul 08826, Korea    Myeongjin Shin Team QST, Seoul National University, Seoul 08826, Korea School of Computing, KAIST, Daejeon 34141, Korea    IlKwon Sohn Quantum Network Research Center, Korea Institute of Science and Technology Information, Daejeon 34141, Korea    Kabgyun Jeong [email protected] Team QST, Seoul National University, Seoul 08826, Korea Research Institute of Mathematics, Seoul National University, Seoul 08826, Korea School of Computational Sciences, Korea Institute for Advanced Study, Seoul 02455, Korea
(November 7, 2025)
Abstract

In classical information theory, uncommon information refers to the amount of information that is not shared between two messages, and it admits an operational interpretation as the minimum communication cost required to exchange the messages. Extending this notion to the quantum setting, quantum uncommon information is defined as the amount of quantum information necessary to exchange two quantum states. While the value of uncommon information can be computed exactly in the classical case, no direct method is currently known for calculating its quantum analogue. Prior work has primarily focused on deriving upper and lower bounds for quantum uncommon information. In this work, we propose a new approach for estimating these bounds by utilizing the quantum Donsker–Varadhan representation and implementing a gradient-based optimization method. Our results suggest a pathway toward efficient approximation of quantum uncommon information using variational techniques grounded in quantum neural architectures.

I Introduction

In classical information theory, a message can be modeled as a sequence of values independently drawn from a discrete probability source X={(x,px)}X=\{(x,p_{x})\}. The entropy of the source, denoted by H(X)H(X), quantifies the inherent randomness in XX, and admits an operational interpretation as the minimal amount of information required to faithfully encode outcomes sampled from XX.

More generally, in a base-rr encoding scheme, the entropy is defined as Hr(X)=xpxlogrpxH_{r}(X)=-\sum_{x}p_{x}\log_{r}p_{x}, a quantity known as the Shannon entropy [Sha1948EN]. Operationally, Hr(X)H_{r}(X) corresponds to the expected number of rr-ary digits needed to represent an outcome from XX. In the special case r=2r=2, we adopt the shorthand log:=log2\log:=\log_{2} and H(X):=H2(X)H(X):=H_{2}(X).

Consider now a communication scenario in which a sender, Alice, wishes to transmit the source XX to a receiver, Bob, through a noisy communication channel Γ\Gamma. The channel transforms the input XX into an output source Y={(y,qy)}Y=\{(y,q_{y})\}, inducing a joint distribution PXYP_{XY} over XX and YY. This statistical dependence enables Bob to infer partial information about XX from his observation of YY, given knowledge of the channel. The amount of information shared between XX and YY is captured by the mutual information, defined as

I(X:Y)=DKL(PXYPXPY),I(X:Y)=D_{\mathrm{KL}}(P_{XY}\|P_{X}\otimes P_{Y}), (I.1)

where DKLD_{\mathrm{KL}} denotes the Kullback–Leibler divergence [KL1997], and PXPYP_{X}\otimes P_{Y} is the product of the marginals of XX and YY. This can also be expressed in terms of Shannon entropy as I(X:Y)=H(X)+H(Y)H(X,Y)I(X:Y)=H(X)+H(Y)-H(X,Y), which corresponds to the sum of the individual entropies minus the joint entropy. Intuitively, mutual information quantifies the extent to which the joint distribution deviates from independence. A natural question then arises:

“How much additional information must Bob acquire
in order to fully reconstruct the message XX?”

The amount of information not shared with YY is captured by the partial information of XX relative to YY, given by H(X)I(X:Y)H(X)-I(X:Y). In classical information theory, this quantity coincides with the conditional entropy H(X|Y)H(X|Y) [Sle1971EN]. The operational meaning of this quantity as the minimum information required for reconstruction is formally established by the Slepian-Wolf theorem. It states that in the asymptotic limit, if Bob already possesses the correlated message Y, Alice need only send her message X at a rate of H(X|Y)H(X|Y) bits per symbol for Bob to reconstruct the message with a probability of error that approaches zero.

Now consider a bidirectional communication setting in which Alice and Bob respectively possess correlated random variables XX and YY, and aim to exchange their messages. Due to the correlation, it suffices to exchange only the parts of the messages that are not mutually shared. The total amount of information required for this task is given by H(X|Y)+H(Y|X)H(X|Y)+H(Y|X), which is referred to as the uncommon information between XX and YY. In classical settings, since this quantity is expressible solely in terms of conditional and mutual entropies, it is often subsumed under the broader framework of mutual information. Refer to Fig. 1 for expressions related to classical information content.

Refer to caption
Figure 1: Classical information quantities associated with messages XX and YY. Each circle represents the entropy of a variable, with the overlapping region indicating the mutual information, and the non-overlapping regions corresponding to the conditional entropies.

In quantum information theory, classical notions are extended to the quantum regime by replacing classical probability distributions with quantum states. The entropy of a quantum state ρ\rho is defined as S(ρ)=Tr(ρlogρ)S(\rho)=-\mathrm{Tr}(\rho\log\rho), known as the von Neumann entropy [von1955QI]. This quantity characterizes the asymptotic number of qubits required to encode nn independently prepared copies of ρ\rho, which is approximately nS(ρ)nS(\rho) qubits [Sch1995QI]. When a quantum system AA is described by the reduced state ρA\rho_{A}, its entropy is denoted by S(A):=S(ρA)S(A):=S(\rho_{A}).

To extend the notion of mutual information to the quantum setting, it is crucial not to define it as the information about one system obtainable by measuring the other, since quantum measurements generally disturb the state and irreversibly alter the system. Instead, for a bipartite quantum system described by ρAB\rho_{AB}, where AA and BB represent subsystems held by two parties, the quantum mutual information is defined as

I(A:B)=D(ρABρAρB),I(A:B)=D(\rho_{AB}\|\rho_{A}\otimes\rho_{B}), (I.2)

where D()D(\cdot\|\cdot) denotes the Umegaki relative entropy. This quantity measures the total correlations between AA and BB by quantifying the deviation of the joint state from a product state. Similarly, this quantity can be written using the von Neumann entropy as I(A:B)=S(A)+S(B)S(AB)I(A:B)=S(A)+S(B)-S(AB).

To formulate a quantum analogue of partial information, one may consider the quantum state merging protocol [Hor2005QSM]. Let |ψABR\ket{\psi}_{ABR} be a purification of the mixed state ρAB\rho_{AB}, where RR is a reference system purifying ABAB. The protocol addresses the task of transferring system AA from one party to the other, so that the receiving party ultimately holds the entire pure state |ψABR\ket{\psi}_{ABR}. The quantum communication cost of this task is given by the quantum conditional entropy S(A|B)=S(AB)S(B)S(A|B)=S(AB)-S(B).

Unlike its classical counterpart, this quantity can be negative. In such cases, the task can be completed without any quantum communication, and furthermore, entanglement is generated in the process. This reflects the fact that when AA and BB are highly entangled, transferring AA provides more than just its marginal information. While other positive-definite definitions of quantum relative entropy exist, they do not correspond to the operational cost of this fundamental state merging task, which provides a conceptual foundation for more elaborate protocols like state exchange.

The concept of uncommon information also admits a quantum generalization. Extending the state merging protocol, one may consider the quantum state exchange protocol [Jon2008QUI, Lee2024QUI], where the goal is to exchange the respective parts of the two parties so that the final global state becomes |ψBAR\ket{\psi}_{BAR}. The minimum quantum communication cost of achieving this transformation defines the quantum uncommon information.

In contrast to the classical case, the quantum uncommon information cannot be expressed as S(A|B)+S(B|A)S(A|B)+S(B|A). Since this sum can be negative, iterating the protocol would imply the possibility of generating unbounded entanglement, which contradicts the no-cloning theorem. Therefore, no straightforward quantum analogue of the classical expression exists, and computing quantum uncommon information requires fundamentally different techniques [Hor1998QI]. Refer to Fig. 2 for expressions related to quantum information content.

Refer to caption
Figure 2: Quantum information quantities associated with systems AA and BB. In the quantum setting, the total information content of the joint system ABAB is given by the sum S(A)+S(B)S(A)+S(B) minus the mutual information. Unlike the classical case, the joint entropy S(AB)S(AB) can be smaller than either S(A)S(A) or S(B)S(B) due to quantum entanglement.

The closed-form expression for the quantum uncommon information is not yet known. Although several studies have investigated specific instances [Lee2019QSE, Lee2019QSE2, Lee2021QSE], for the general case, only upper and lower bounds have been established [Jon2008QUI, Lee2024QUI].

Quantum uncommon information is not merely a theoretical construct. It serves as a key metric for quantifying communication costs in various quantum network protocols, such as distributed quantum computing [App1999DQC, App2004DQC, App2013DQC, App2016DQC], quantum key agreement [App2017QKD, App2020QKD, App2023QKD], and quantum secret sharing [App1999QSS, App2008QSS].

In this work, we address this gap by proposing a method for predicting bounds on the quantum uncommon information of a given quantum state. Our approach leverages quantum machine learning techniques based on the quantum Donsker–Varadhan representation [Shin2024QMINE, Ziv2024QMINE], which we employ to estimate the von Neumann entropy. These entropy estimates then enable the derivation of corresponding bounds on the quantum uncommon information.

II Quantum uncommon information

II.1 Quantum state exchange protocol

The quantum uncommon information is defined operationally via the quantum state exchange protocol [Jon2008QUI, Lee2024QUI]. This protocol is framed within the paradigm of entanglement as a resource, where any task beyond the scope of local operations and classical communication (LOCC) must be paid for by consuming pre-shared entanglement.

In this scenario, two parties, Alice and Bob, hold quantum systems AA and BB, which are part of a larger pure state ψ=|ψψ|\psi=\ket{\psi}\bra{\psi} on a system ABRABR, where RR is a reference system that purifies ABAB such that TrBR(ψ)=ρA\operatorname{Tr}_{BR}(\psi)=\rho_{A} and TrAR(ψ)=ρB\operatorname{Tr}_{AR}(\psi)=\rho_{B}. The goal is to exchange their respective systems, AA and BB, using only LOCC assisted by an initial amount of shared pure entanglement. The cost of the protocol is then the net amount of entanglement, measured in units of maximally entangled states (ebits), consumed to achieve this exchange. This resource-centric view motivates the formal definition of the protocol that follows. Refer to Fig. 3 for the description of the quantum state exchange protocol.

Definition II.1 (Quantum state exchange protocol).

Let ψ\psi be a pure state on ABRABR, and let ϕ\phi and ϕ\phi^{\prime} be maximally entangled states on auxiliary systems SS and SS^{\prime}, respectively. Let ψex\psi_{\mathrm{ex}} denote the target state obtained from ψ\psi by swapping systems AA and BB. A quantum state exchange protocol with error ϵ\epsilon is a quantum channel

:ABRSABRS\mathcal{E}:ABRS\rightarrow ABRS^{\prime} (II.1)

satisfying

[ψϕ]ψexϕ1ϵ,\left\|\mathcal{E}\left[\psi\otimes\phi\right]-\psi_{\mathrm{ex}}\otimes\phi^{\prime}\right\|_{1}\leq\epsilon, (II.2)

where ϵ>0\epsilon>0 is the permissible error.

Now consider the asymptotic setting, where each party holds nn copies of the initial state ψ\psi. If the error ϵn\epsilon_{n} vanishes as nn\to\infty, the protocol achieves asymptotically faithful exchange. The quantum uncommon information is then defined as the asymptotic entanglement cost of this task.

Definition II.2 (Quantum uncommon information).

Let ψ\psi be a pure state on ABRABR, and let ψex\psi_{\mathrm{ex}} be the target state with systems AA and BB exchanged. Consider a sequence of quantum channels

n:(ABR)nSn(ABR)nSn,\mathcal{E}_{n}:(ABR)^{\otimes n}\otimes S_{n}\rightarrow(ABR)^{\otimes n}\otimes S^{\prime}_{n}, (II.3)

where ϕn\phi_{n} and ϕn\phi^{\prime}_{n} are maximally entangled states on SnS_{n} and SnS^{\prime}_{n}, respectively. Suppose the protocol satisfies

n[ψnϕn]ψexnϕn1ϵn,\left\|\mathcal{E}_{n}\left[\psi^{\otimes n}\otimes\phi_{n}\right]-\psi_{\mathrm{ex}}^{\otimes n}\otimes\phi^{\prime}_{n}\right\|_{1}\leq\epsilon_{n}, (II.4)

with ϵn0\epsilon_{n}\to 0 as nn\to\infty. Let rnr_{n} and rnr^{\prime}_{n} denote the Schmidt ranks of ϕn\phi_{n} and ϕn\phi^{\prime}_{n}. Then, the quantum uncommon information Υ(A:B)\Upsilon(A:B) is defined as

Υ(A:B)=inf{lim infn1n(log(rn)log(rn))},\Upsilon(A:B)=\inf\left\{\liminf_{n\to\infty}\frac{1}{n}\left(\log(r_{n})-\log(r^{\prime}_{n})\right)\right\}, (II.5)

where the infimum is taken over all such sequences of protocols {n}\{\mathcal{E}_{n}\} achieving vanishing error.

Refer to caption
Figure 3: Quantum state exchange protocol. The diagram illustrates the quantum state exchange process. Alice (system AA) holds the reduced state ρA\rho_{A}, and Bob (system BB) holds ρB\rho_{B}. Their systems are exchanged via a quantum channel \mathcal{E}. To facilitate this process, they consume a shared maximally entangled state SS, which is transformed into another entangled state SS^{\prime} after the protocol. The reduction in entanglement quantifies the quantum uncommon information.

II.2 Bounds on quantum uncommon information

There is no known closed-form expression for the quantum uncommon information Υ(A:B)\Upsilon(A:B). However, several studies have investigated upper and lower bounds using various strategies [Jon2008QUI, Lee2024QUI].

II.2.1 Upper bounds

An upper bound on Υ(A:B)\Upsilon(A:B) can be obtained by constructing an explicit quantum state exchange protocol and calculating the corresponding entanglement rate. As a baseline, consider the protocol in which Alice first transmits her state ρA\rho_{A} to Bob, and then Bob sends his state ρB\rho_{B} to Alice. This approach effectively applies the quantum state merging protocol twice. The total entanglement consumption in this case is S(A|B)+S(B)=S(AB)S(A|B)+S(B)=S(AB), and the strategy is referred to as the merge-and-send protocol.

This basic protocol can be improved by identifying and removing redundant components that do not require exchange. The common subspace refers to a part of the quantum state that remains invariant under the exchange and hence need not be transmitted, reducing the total entanglement cost.

Definition II.3 (Common subspace).

Let |ψ\ket{\psi} be a pure state on the tripartite system ABRABR. A subspace CC of both AA and BB is called a common subspace if there exist unitary operators UU on AA and WW on BB such that

|ψ\displaystyle\ket{\psi^{\prime}} =(UW𝟙R)|ψ=|ψc+|ψu,\displaystyle=(U\otimes W\otimes\mathbb{1}_{R})\ket{\psi}=\ket{\psi_{c}}+\ket{\psi_{u}}, (II.6)
|ψc\displaystyle\ket{\psi_{c}} =(ΠCΠC𝟙R)|ψ,\displaystyle=(\Pi_{C}\otimes\Pi_{C}\otimes\mathbb{1}_{R})\ket{\psi^{\prime}}, (II.7)
|ψu\displaystyle\ket{\psi_{u}} =(ΠCΠC𝟙R)|ψ,\displaystyle=(\Pi_{C^{\perp}}\otimes\Pi_{C^{\perp}}\otimes\mathbb{1}_{R})\ket{\psi^{\prime}}, (II.8)
|ψc\displaystyle\ket{\psi_{c}} =|ψcex,\displaystyle=\ket{\psi_{c}}_{\mathrm{ex}}, (II.9)

where ΠX\Pi_{X} denotes the projection operator onto subspace XX, and |ψcex\ket{\psi_{c}}_{\mathrm{ex}} denotes the state |ψc\ket{\psi_{c}} with systems AA and BB exchanged.

Let AA^{\prime} and BB^{\prime} be ancillary systems with the same dimensions as AA and BB, respectively. For fixed pure states |xC\ket{x}\in C^{\perp} and |yC\ket{y}\in C, define the unitary operator

Us=|iC|yi||ix|U_{s}=\sum_{\ket{i}\in C}\ket{y}\bra{i}\otimes\ket{i}\bra{x} (II.10)

acting on systems AAAA^{\prime} and BBBB^{\prime}. Then the following pure state is called the stretched state:

|ψs\displaystyle\ket{\psi_{s}} =|ψcABR|xA|xB\displaystyle=\ket{\psi_{c}}_{ABR}\otimes\ket{x}_{A^{\prime}}\otimes\ket{x}_{B^{\prime}}
+|yA|yB|ψuABR.\displaystyle\quad+\ket{y}_{A}\otimes\ket{y}_{B}\otimes\ket{\psi_{u}}_{A^{\prime}B^{\prime}R}. (II.11)

By converting the original state into the stretched state and applying a subspace exchange protocol that swaps only AA^{\prime} and BB^{\prime}, the resulting entanglement cost is given by the conditional entropy S(RA)ψsS(R\mid A)_{\psi_{s}}.

II.2.2 Lower bounds

A lower bound on Υ(A:B)\Upsilon(A:B) can be derived by observing that the sum of the initial entanglement and the entanglement used in the protocol must be at least the final entanglement.

From Alice’s perspective, the initial entanglement of system AA is S(A)ψS(A)_{\psi}, while after the exchange, the entanglement becomes S(A)ψex=S(B)ψS(A)_{\psi_{\mathrm{ex}}}=S(B)_{\psi}. Therefore,

Υ(A:B)S(B)S(A).\Upsilon(A:B)\geq S(B)-S(A). (II.12)

A symmetric argument from Bob’s perspective yields the lower bound

Υ(A:B)|S(B)S(A)|.\Upsilon(A:B)\geq|S(B)-S(A)|. (II.13)

This idea can be refined by considering nn copies of the state and analyzing the entanglement structure across them via a decomposition protocol.

Definition II.4 (Decomposed states).

Let |ψ\ket{\psi} be a pure state on ABRABR, and let AiA_{i}, BiB_{i}, and RiR_{i} denote systems corresponding to Alice, Bob, and a reference, respectively. Suppose that for the nn-fold product state |ψn\ket{\psi}^{\otimes n}, there exists a reversible transformation Λn\Lambda_{n} such that, for some error ϵn0\epsilon_{n}\to 0,

Λn[|ψn]=\displaystyle\Lambda_{n}\left[\ket{\psi}^{\otimes n}\right]= |ψ1A1R1r1n|ψ2B1R2r2n\displaystyle\ket{\psi_{1}}_{A_{1}R_{1}}^{\otimes\lfloor r_{1}n\rfloor}\otimes\ket{\psi_{2}}_{B_{1}R_{2}}^{\otimes\lfloor r_{2}n\rfloor}
|ψ3A2B2r3n|ψ4A3B3R3R4r4n,\displaystyle\otimes\ket{\psi_{3}}_{A_{2}B_{2}}^{\otimes\lfloor r_{3}n\rfloor}\otimes\ket{\psi_{4}}_{A_{3}B_{3}R_{3}R_{4}}^{\otimes\lfloor r_{4}n\rfloor}, (II.14)

where rir_{i} are non-negative rational numbers. If such a transformation exists, the four disjoint pure states ψ1,,ψ4\psi_{1},\dots,\psi_{4} are referred to as the decomposed states, and the corresponding entanglement cost is given by r1S(A1)ψ1+r2S(B1)ψ2+r4(S(B3R3)ψ4S(A3R3)ψ4)r_{1}S(A_{1})_{\psi_{1}}+r_{2}S(B_{1})_{\psi_{2}}+r_{4}\left(S(B_{3}R_{3})_{\psi_{4}}-S(A_{3}R_{3})_{\psi_{4}}\right).

The best known bounds can now be summarized as follows.

Proposition II.1 (Bounds on quantum uncommon information).

Let Υ(A:B)\Upsilon(A:B) denote the quantum uncommon information between quantum systems AA and BB. Then the following inequalities hold:

Υ(A:B)\displaystyle\Upsilon(A:B) infCu[C]S(AB),\displaystyle\leq\inf_{C}\,u[C]\leq S(AB), (II.15)
Υ(A:B)\displaystyle\Upsilon(A:B) supΛl[Λ]|S(B)S(A)|,\displaystyle\geq\sup_{\Lambda}\,l[\Lambda]\geq|S(B)-S(A)|, (II.16)

where the upper bound is defined by u[C]:=S(RA)ψsu[C]:=S(R\mid A)_{\psi_{s}}, and the lower bound is defined by l[Λ]:=r1S(A1)ψ1+r2S(B1)ψ2+r4(S(B3R3)ψ4S(A3R3)ψ4)l[\Lambda]:=r_{1}S(A_{1})_{\psi_{1}}+r_{2}S(B_{1})_{\psi_{2}}+r_{4}\left(S(B_{3}R_{3})_{\psi_{4}}-S(A_{3}R_{3})_{\psi_{4}}\right).

Remark II.1.

In general, it is computationally challenging to evaluate the infimum over all possible common subspaces CC or the supremum over all valid decompositions Λ\Lambda. In this work, we focus on scenarios where the parties possess additional structural knowledge about the state, enabling efficient identification of the common subspace. For decomposed states, we restrict our attention to special cases in which such a decomposition is guaranteed. Further details are provided in Section IV.

III Quantum Donsker–Varadhan representation

The Donsker–Varadhan representation provides a lower bound on the divergence between two probability distributions [von1932DVR]. It has been widely applied to mutual information estimation by optimizing over a neural network. The quantum Donsker–Varadhan representation is the quantum analogue of this formulation, enabling entropy estimation via the Gibbs variational principle [Shin2024QMINE, Ziv2024QMINE].

Proposition III.1 (Quantum Donsker–Varadhan representation).

Let ρ\rho be a dd-dimensional density matrix with rank rr, and define the function f:d×df:\mathcal{H}^{d\times d}\to\mathbb{R} by

f(T)=Tr(cρT)+log(Tr(ecT)),f(T)=-\mathrm{Tr}(c\rho T)+\log\left(\mathrm{Tr}(e^{cT})\right), (III.1)

where TT is a Hermitian operator and cc is a positive constant. Then, for any ϵ>0\epsilon>0, we have

|S(ρ)infTf(T)|<ϵ,\left|S(\rho)-\inf_{T}f(T)\right|<\epsilon, (III.2)

provided that TT is an rr-rank density matrix and c2rlog(d)rlog(ϵ)c\geq 2r\log(d)-r\log(\epsilon).

According to this formulation, one can estimate the von Neumann entropy S(ρ)S(\rho) by minimizing the function f(T)f(T) over rank-rr Hermitian operators TT. The optimal value approximates S(ρ)S(\rho) up to arbitrary precision ϵ\epsilon.

III.1 Neural estimation of quantum entropy

Suppose ρ\rho is a density matrix of rank rr. Let U(𝜽)U(\bm{\theta}) be a parameterized unitary operator with parameters 𝜽={θ1,,θn}\bm{\theta}=\{\theta_{1},\dots,\theta_{n}\}. We define a parameterized ansatz for the Hermitian operator TT as

T(𝜽,𝐭)=i=1rtiU(𝜽)|ii|U(𝜽),T(\bm{\theta},\mathbf{t})=\sum_{i=1}^{r}t_{i}\,U(\bm{\theta})\ket{i}\bra{i}U^{\dagger}(\bm{\theta}), (III.3)

where 𝐭={t1,,tr}\mathbf{t}=\{t_{1},\dots,t_{r}\}, ti0t_{i}\geq 0, and i=1rti=1\sum_{i=1}^{r}t_{i}=1.

Then the function ff can be rewritten as

f(𝜽,𝐭)=\displaystyle f(\bm{\theta},\mathbf{t})= ci=1rtii|U(𝜽)ρU(𝜽)|i\displaystyle-c\sum_{i=1}^{r}t_{i}\bra{i}U^{\dagger}(\bm{\theta})\rho U(\bm{\theta})\ket{i}
+log(dr+i=1recti).\displaystyle+\log\left(d-r+\sum_{i=1}^{r}e^{ct_{i}}\right). (III.4)

The first term can be computed on a quantum computer using a parameterized quantum circuit, while the second term is evaluated classically. Accordingly, we used the function f(𝜽,𝐭)f(\bm{\theta},\mathbf{t}) from Section III.1 directly as the loss function. We then obtained an estimate of the von Neumann entropy by updating the parameters to minimize the value of this loss function.

The gradient of ff with respect to 𝐭\mathbf{t} can be computed classically, and the gradient with respect to 𝜽\bm{\theta} can be evaluated using the parameter-shift rule [Mit2018QML]. This allows the use of hybrid quantum–classical optimization techniques, such as gradient descent, to minimize f(𝜽,𝐭)f(\bm{\theta},\mathbf{t}) and thereby estimate the von Neumann entropy of ρ\rho [Shin2024QMINE, Ziv2024QMINE].

In our numerical simulations, we employed the Adam (Adaptive Moment Estimation) optimizer to update the variational parameters. The initial learning rate was set to 0.1, and a StepLR learning rate scheduler was also applied, which decays the learning rate by a factor of 0.9 every 10 optimization steps to enhance the stability of the training process.

It has been shown that the number of copies of ρ\rho required for this learning process scales as O(poly(r))O(\mathrm{poly}(r)), where rr is the rank of the state [Shin2024QMINE]. This implies that the quantum Donsker–Varadhan representation is particularly effective when applied to low-rank quantum states.

IV Estimation of bounds

Based on the above, the quantum Donsker–Varadhan representation can be utilized to estimate the von Neumann entropy of a given density matrix. Since the bounds of quantum uncommon information are expressed in terms of von Neumann entropy, applying the same approach allows quantum Donsker–Varadhan representation to estimate the bounds of quantum uncommon information between given quantum states. Notably, as the process of expressing the bounds reduces the size of the states whose von Neumann entropy needs to be estimated, the rank of these states also decreases. Therefore, the method proves to be effective in such cases.

Consider the case of estimating the bounds of quantum uncommon information Υ(A:B)\Upsilon(A:B) for quantum systems AA and BB. In general, let us consider a given pure state ψABR\psi_{ABR}, where RR is a purifying system of ABAB.

IV.1 The loose bounds

For the loose upper bound S(AB)S(AB) and lower bound |S(B)S(A)||S(B)-S(A)|, the von Neumann entropy of each state can be directly estimated to obtain these values. Consequently, for loose bounds, it is relatively straightforward to approximate their values. However, for tight bounds, direct estimation becomes more challenging.

Therefore, we impose several constraints to determine the form of the state, which will then allow us to estimate the bound value. When the tight bound can be measured, the significance of the wide bound value becomes irrelevant. However, for the purpose of comparison, we can easily estimate the bound using a state of the form ρAB\rho_{AB} by applying the aforementioned method. In the following, we will explore methods for estimating tight upper and lower bounds, respectively.

IV.2 The tight bounds

IV.2.1 Common subspaces

Before estimating the tight upper bound, we first examine the properties of the common subspace.

Suppose the common subspace CC with respect to a specific basis is given by C=span{|i1,,|i}C=\mathrm{span}\{\ket{i_{1}},\dots,\ket{i_{\ell}}\}. If ρA\rho_{A} and ρB\rho_{B} satisfy the conditions for a common subspace without the need for additional unitary transformations, then when expressed in matrix form with respect to the same basis, the following must hold:

ρA=(X00MA),ρB=(X00MB),\rho_{A}=\begin{pmatrix}X&0\\ 0&M_{A}\end{pmatrix},\quad\rho_{B}=\begin{pmatrix}X&0\\ 0&M_{B}\end{pmatrix}, (IV.1)

where XX is an ×\ell\times\ell matrix, and MAM_{A}, MBM_{B} are (n)×(n)(n-\ell)\times(n-\ell) matrices. Moreover, the spectrum of XX must correspond to the spectrum of ρA\rho_{A} and ρB\rho_{B}. Therefore, the common subspace is contained in the subspace corresponding to the components where ρA\rho_{A} and ρB\rho_{B} have the same eigenvalues.

Now suppose the density matrices are given by

ρA=i=1nαi|aiai|,ρB=i=1nβi|bibi|.\rho_{A}=\sum_{i=1}^{n}\alpha_{i}\ket{a_{i}}\bra{a_{i}},\quad\rho_{B}=\sum_{i=1}^{n}\beta_{i}\ket{b_{i}}\bra{b_{i}}. (IV.2)

We are allowed to apply arbitrary unitary operations to systems AA and BB, respectively. Since a unitary operator acts as a basis change, we can align the states to a desired basis to identify the common subspace. Thus, the following holds.

Proposition IV.1 (Partial spectral alignment and unitary diagonalization).

Let ρA\rho_{A} and ρB\rho_{B} be density matrices of the same dimension nn, with eigenvalues ordered as α1αn\alpha_{1}\leq\cdots\leq\alpha_{n} and β1βn\beta_{1}\leq\cdots\leq\beta_{n}. Suppose there exists a permutation πSn\pi\in S_{n} such that for some kk, the condition απ(i)=βπ(i)\alpha_{\pi(i)}=\beta_{\pi(i)} holds for 1ik1\leq i\leq k. Then there exist unitary operators UU and WW such that

UρAU=i=1nαπ(i)|ii|,WρBW=i=1nβπ(i)|ii|.U\rho_{A}U^{\dagger}=\sum_{i=1}^{n}\alpha_{\pi(i)}\ket{i}\bra{i},\quad W\rho_{B}W^{\dagger}=\sum_{i=1}^{n}\beta_{\pi(i)}\ket{i}\bra{i}.

Let {|1,,|k}\{\ket{1},\dots,\ket{k}\} be the basis that maximizes the overlap of eigenvalues. Then, from the above, we obtain:

Theorem IV.1 (Unitary mapping of a subspace to a fixed basis segment).

Let C=span{|i1,,|il}C=\mathrm{span}\{\ket{i_{1}},\dots,\ket{i_{l}}\} be a common subspace. Then there exists a unitary UCU_{C} such that {UC|i1,,UC|il}{|1,,|k}.\{U_{C}\ket{i_{1}},\dots,U_{C}\ket{i_{l}}\}\subset\{\ket{1},\dots,\ket{k}\}.

Now consider a decomposition of ψABR\psi_{ABR} into subspaces CCC\otimes C and CCC^{\perp}\otimes C^{\perp}:

|ψ=i,j=1dkcijk|i|j|rk+i,j=d+1nkcijk|i|j|rk,\ket{\psi}=\sum_{i,j=1}^{d}\sum_{k}c_{ijk}\ket{i}\ket{j}\ket{r_{k}}+\sum_{i,j=d+1}^{n}\sum_{k}c_{ijk}\ket{i}\ket{j}\ket{r_{k}},

where {|rk}\{\ket{r_{k}}\} is an orthonormal basis of RR. In general, we have:

Proposition IV.2 (Characterization of common subspaces via a nonzero-structure relation).

Define a relation \sim on S={1,,n}S=\{1,\dots,n\} by aba\sim b if cabk0c_{abk}\neq 0 or cbak0c_{bak}\neq 0 for some kk. Then the equivalence classes Sa={x:xa}S_{a}=\{x:x\sim a\} define all possible common subspaces of the form C=span{|i:iSx}.C=\mathrm{span}\{\ket{i}:i\in S_{x}\}.

To identify CC, prepare two copies of ψABR\psi_{ABR}, construct a swapped version ψ\psi^{\prime} by exchanging CC^{\perp}, and run a swap test. Let CC be the union of subspaces that pass this test. Then, by Theorem IV.1, this gives the infimum for u[C]u[C].

IV.2.2 Decomposed states

For the lower bound, we classify the structure of entanglement. Consider a decomposition into EPR and GHZ states [Vidal2000, Lee2024QUI]. Suppose:

|ψ=\displaystyle\ket{\psi}= c112(|000+|101)+c212(|212+|223)\displaystyle c_{1}\frac{1}{\sqrt{2}}(\ket{000}+\ket{101})+c_{2}\frac{1}{\sqrt{2}}(\ket{212}+\ket{223})
+c312(|334+|444)+c4|555.\displaystyle+c_{3}\frac{1}{\sqrt{2}}(\ket{334}+\ket{444})+c_{4}\ket{555}. (IV.3)

Then we have:

Proposition IV.3 (Reversible entanglement decomposition of tensor powers of tripartite states).

For |ψn\ket{\psi}^{\otimes n}, there exists a reversible map Λn\Lambda_{n} such that

|ψn\displaystyle\ket{\psi}^{\otimes n} |EPRARr1n|EPRBRr2n\displaystyle\approx\ket{\mathrm{EPR}}_{AR}^{\otimes\lfloor r_{1}n\rfloor}\otimes\ket{\mathrm{EPR}}_{BR}^{\otimes\lfloor r_{2}n\rfloor}
|EPRABr3n|GHZABRr4n,\displaystyle\otimes\ket{\mathrm{EPR}}_{AB}^{\otimes\lfloor r_{3}n\rfloor}\otimes\ket{\mathrm{GHZ}}_{ABR}^{\otimes\lfloor r_{4}n\rfloor}, (IV.4)

where r1=c12,r2=c22,r3=c32,r4=i=14ci2logci2r_{1}=c_{1}^{2},~r_{2}=c_{2}^{2},~r_{3}=c_{3}^{2},~r_{4}=-\sum_{i=1}^{4}c_{i}^{2}\log c_{i}^{2}.

It is known that not every tripartite state admits such decomposition [Acin2003], and there is no general method to construct it. Thus, we restrict estimation to cases where such decompositions are explicitly given.

V Numerical simulations

To verify whether the proposed method accurately estimates the bounds, we conducted numerical simulations. First, we implemented a quantum machine learning algorithm to estimate the von Neumann entropy of a given state. In this setup, the parameterized unitary U(𝜽)U(\bm{\theta}) was constructed using rotation gates along each axis, together with CNOT gates. The specific structure of the ansatz is shown in Fig. 4.

Refer to caption
Figure 4: Ansatz structure in numerical simulations. Shown is the layered ansatz architecture utilized to construct the parameterized unitary U(𝜽)U(\bm{\theta}) in the numerical experiments. The same structure is uniformly repeated over LL layers.

Subsequently, we describe the estimation procedures for each of the bounds. To verify the performance and stability of our proposed method, we conducted extensive numerical simulations. Each simulation reported in this section was performed for 10 independent runs. For each run, the variational parameters of the quantum circuit were initialized with different random values.

The figures presented in this section depict the aggregated results of these multiple runs. The solid line represents the mean of the entropy estimates at each iteration, while the shaded area corresponds to the standard deviation. This shaded region thus represents the statistical variability of our algorithm’s convergence behavior, originating from the different random initial starting points in the optimization landscape.

Refer to caption
(a) Estimation result for the 4-qubit case.
Refer to caption
(b) Estimation result for the 6-qubit case.
Refer to caption
(c) Estimation result for the 8-qubit case.
Figure 5: Convergence of the loose upper bound S(AB)S(AB). Shown are the convergence profiles of the upper bound S(AB)S(AB) for quantum states consisting of 4, 6, and 8 qubits.
Refer to caption
(a) Estimation result for the 4-qubit case.
Refer to caption
(b) Estimation result for the 6-qubit case.
Refer to caption
(c) Estimation result of for the 8-qubit case.
Figure 6: Convergence of the loose lower bound |S(B)S(A)||S(B)-S(A)|. Shown are the convergence profiles of the quantity |S(B)S(A)||S(B)-S(A)| as a loose lower bound, evaluated on 4-, 6-, and 8-qubit quantum states.

V.1 Estimation of the loose bounds

For the loose upper bound S(AB)S(AB) and the loose lower bound |S(B)S(A)||S(B)-S(A)|, we directly estimated each von Neumann entropy of the given ρAB\rho_{AB}. To simulate this, we first generated random quantum states and partitioned the system equally, assigning half to AA and the other half to BB, thereby constructing the target state for entropy estimation. Then, using the entropy estimation algorithm described in the previous section, we estimated the corresponding entropies. This simulation was conducted for both 4, 6, and 8 qubit systems.

The simulation results for the upper bound S(AB)S(AB) are shown in Fig. 5. The estimation of S(AB)S(AB) stabilized after approximately 100 optimization steps for the 4-qubit system and around 200 steps for the 6-qubit and 8-qubit system.

For the lower bound |S(B)S(A)||S(B)-S(A)|, since only half of the system is used to estimate S(A)S(A) and S(B)S(B), the actual subsystems involved correspond to 2-qubit, 3-qubit, and 4-qubit reduced states, respectively. The results are presented in Fig. 6.

The estimation of |S(B)S(A)||S(B)-S(A)| stabilized after roughly 50 optimization steps. Theoretically, since the error in the quantum Donsker–Varadhan representation-based method is proportional to the rank of the density matrix, larger systems require more optimization steps to achieve stable estimates.

Refer to caption
(a) Estimation result for the 4-qubit case.
Refer to caption
(b) Estimation result for the 8-qubit case.
Figure 7: Convergence of the tight upper bound u[C]u[C]. Shown are the convergence profiles of the tight upper bound u[C]u[C] evaluated on 4- and 8-qubit quantum states.

V.2 Estimation of the tight upper bound

Refer to caption
Figure 8: Effect of the bound gap on convergence. Shown is the difference in convergence behavior when a gap exists between the loose and tight bounds.
Refer to caption
Figure 9: Convergence of the tight lower bound l[Λ]l[\Lambda]. Shown is the convergence of the tight lower bound l[Λ]l[\Lambda] for a given quantum state.

To estimate the tight upper bound u[C]=S(R|A)ψsu[C]=S(R|A)_{\psi_{s}}, we first identify the common subspace CC of systems AA and BB from the given state ψABR\psi_{ABR}, then construct the stretched state ψs\psi_{s}, and perform entropy estimation on the full system AABBRAA^{\prime}BB^{\prime}R. In practice, when the system size is small, it is often the case that no nontrivial common subspace exists. Therefore, in this simulation, we fixed the common subspace in advance and proceeded with the estimation.

Specifically, let the dimensions of systems AA and BB be 2n2^{n}, and fix an integer 1k2n1\leq k\leq 2^{n}. Denote the computational basis of AA and BB by {|1,,|2n}\{\ket{1},\dots,\ket{2^{n}}\}, and define the common subspace CC as C=span{|1,,|k}C=\mathrm{span}\{\ket{1},\dots,\ket{k}\}. In other words, for the tripartite system ABRABR, we ensured that the components supported on CCRC\otimes C\otimes R were symmetric under the exchange of systems AA and BB, while the components supported on CCRC^{\perp}\otimes C^{\perp}\otimes R were randomly generated.

Simulations were conducted for total system sizes of 4 and 8 qubits for ABRABR. For the 4-qubit system, we set k=1k=1; for the 8-qubit system, we set k=2k=2. The results are shown in Fig. 7.

The estimation of u[C]u[C] stabilized after approximately 200 optimization steps. Since the ranks of the reduced states on ARAR and AA, which are used in the entropy estimation process, depend on the value of kk, the number of steps required for convergence increases with the size of the common subspace.

Finally, Fig. 8 compares the estimation results for the loose and tight upper bounds of the quantum uncommon information of the given state. The difference between the tight and loose bounds leads to distinct convergence behaviors in their corresponding estimated values.

V.3 Estimation of the tight lower bound

For the tight lower bound l[Λ]=r1S(A1)ψ1+r2S(B1)ψ2+r4(S(B3R3)ψ4S(A3R3)ψ4)l[\Lambda]=r_{1}S(A_{1})_{\psi_{1}}+r_{2}S(B_{1})_{\psi_{2}}+r_{4}(S(B_{3}R_{3})_{\psi_{4}}-S(A_{3}R_{3})_{\psi_{4}}), the simulation was conducted as described earlier, assuming that the state is given by Section IV.2.2.

According to Proposition IV.3, the state can be decomposed into EPR and GHZ states, which allows us to estimate the value of the lower bound. Specifically, the coefficients c1,,c4c_{1},\dots,c_{4} are randomly chosen such that |ψ\ket{\psi} remains a pure state. Based on these coefficients, the values of r1,,r4r_{1},\dots,r_{4} can be computed. Since ψ1,ψ2,ψ3\psi_{1},\psi_{2},\psi_{3} are EPR states and ψ4\psi_{4} is a GHZ state, we estimate the corresponding entropies to obtain the value of l[Λ]l[\Lambda]. Fig. 9 presents the simulation results. The estimation of l[Λ]l[\Lambda] stabilized after roughly 50 optimization steps.

VI Concluding remarks

In this paper, we proposed a quantum machine learning approach to estimate the bounds of the quantum uncommon information, which represents the minimum amount of entanglement required to exchange the given state ρAB\rho_{AB} between subsystems AA and BB. Since quantum Donsker–Varadhan representation allows the von Neumann entropy of a state to be expressed as the infimum of a specific function, we designed a machine learning algorithm that utilizes this as a cost function to estimate entropy. As the bounds of quantum uncommon information are expressed in terms of von Neumann entropy, the same algorithm can be used to estimate their values. Notably, the number of copies required for training with quantum Donsker–Varadhan representation scales as poly(r)\mathrm{poly}(r) with respect to the rank rr, providing an advantage in estimating bounds that involve entropies of subsystems of a given state. In addition, the method proposed in this paper for identifying the common subspace of a given state can be extended to other estimation techniques.

Our estimation method requires different types of information from Alice and Bob depending on the bound to be estimated. Fundamentally, the use of quantum Donsker–Varadhan representation requires knowledge of the rank of each state. In scenarios where the ranks are not known, one may employ techniques such as quantum rank estimation [Ryan2019VQD].

To determine the common subspace, Alice and Bob must additionally know the marginal states they individually possess and how the entire system can be decomposed. In cases where the marginal states are not known, one can employ a variational quantum circuit to search for unitaries UU and WW. However, this approach still requires access to information about the global system.

Moreover, a decomposed state requires prior knowledge of whether the given state can be expressed in a specific canonical form. Therefore, depending on the context in which quantum uncommon information is applied, one must estimate different bounds accordingly.

Quantum uncommon information can be broadly applied in scenarios where the parties are required to exchange their quantum states completely, including entanglement, and thus provides a natural measure of communication cost in such contexts. These situations frequently arise during entanglement distribution protocols in quantum networks [AppQN1997, AppQN2016, App1999DQC, App2004DQC, App2013DQC, App2016DQC, App2017QKD, App2020QKD, App2023QKD, App1999QSS, App2008QSS].

In many of these cases, Alice and Bob have prior knowledge about their respective states, particularly when the state preparation process is known. Furthermore, in these cases, the states often consist of EPR or GHZ states. Under these conditions, the requirements for identifying the common subspace and the decomposed state are often satisfied, and the estimation method for quantum uncommon information can be utilized effectively.

To allow for more general application, certain aspects need to be improved. First, if the process of identifying the common subspace can be refined to work without prior knowledge of the structure of the original state, then it may be possible to determine the common subspace based solely on information from the partial systems held by Alice and Bob. Furthermore, our approach remains applicable even when different forms of decomposed states are considered.

Although this work assumes an ideal quantum system, the presence of noise in realistic quantum devices could affect the performance of the proposed methodology. For instance, errors in state preparation or gate operations within the variational circuit could degrade the accuracy of the loss function calculation, potentially introducing a bias to the entropy estimate or hindering convergence.

A crucial direction for future research is therefore to integrate quantum error mitigation techniques into our algorithm to test its robustness in noisy environments. Such an extension would significantly enhance the practical applicability of the proposed method.

Data availability statement

The data and software that support the findings of this study can be found in the following repository: https://github.com/donghwa722/QUINE

Acknowledgements

J.L. acknowledges helpful discussions with Ju-Young Ryu. This work was supported by the National Research Foundation of Korea (NRF) through a grant funded by the Ministry of Science and ICT (Grant No. RS-2025-00515537). This work was also supported by the Institute for Information & Communications Technology Promotion (IITP) grant funded by the Korean government (MSIP) (Grant Nos. RS-2019-II190003 and RS-2025-02304540), the National Research Council of Science & Technology (NST) (Grant No. GTL25011-401), and the Korea Institute of Science and Technology Information (KISTI) (Grant No. P25026). I.K.S. acknowledges support by Quantum Computing based on Quantum Advantage challenge research through the National Research Foundation of Korea (NRF) funded by the Korean government (MSIT) (Grant No. RS-2023-00256221).