Loading web-font TeX/Main/Regular
Shemesh Theorem and Its Relation With the Zero-Error Quantum Information Theory | IEEE Journals & Magazine | IEEE Xplore

Shemesh Theorem and Its Relation With the Zero-Error Quantum Information Theory


The relationship between quantum channel fixed points, Shemesh subspace and zero-error capacity

Abstract:

In this paper, our objective is to establish results that connect Shemesh theorem with quantum zero-error information theory. To achieve this, we will first introduce def...Show More

Abstract:

In this paper, our objective is to establish results that connect Shemesh theorem with quantum zero-error information theory. To achieve this, we will first introduce definitions and properties related to the zero-error capacity of a quantum channel, with particular emphasis on demonstrating that the eigenstates common to the Kraus operators representing the quantum channel are fixed points. This result has previously been employed to develop a test for determining whether a quantum channel possesses positive zero-error capacity. In this work, we demonstrate that the eigenstates common to the Kraus operators representing the channel reside within the Shemesh subspace. Consequently, the fixed points of the quantum channel act as a bridge between Shemesh theorem and quantum zero-error information theory, thereby enabling the derivation of new results on the zero-error capacity of quantum channels.
The relationship between quantum channel fixed points, Shemesh subspace and zero-error capacity
Published in: IEEE Access ( Volume: 12)
Page(s): 186153 - 186159
Date of Publication: 09 December 2024
Electronic ISSN: 2169-3536

Funding Agency:


CCBY - IEEE is not the copyright holder of this material. Please follow the instructions via https://creativecommons.org/licenses/by/4.0/ to obtain full-text articles and stipulations in the API documentation.
SECTION I.

Introduction

Information theory can be defined as a science whose fundamental elements are the quantification, loading and transmission of information. Information theory is studied according to two scenarios, the classical and the quantum. In the classical scenario, basic elements of classical information theory are bits, while the transmission, processing and load of information obey the laws of classical mechanics [1]. In the quantum context, quantum information theory introduces a new paradigm for processing and transmitting information through quantum channels. Unlike classical systems, quantum information theory is governed by the principles of quantum mechanics, leveraging phenomena such as superposition and entanglement. This implies that information can be represented not only in the form of classical bits but also as quantum bits, commonly known as qubits. Whereas bits can only take on two values, qubits can exist in a superposition of both states simultaneously, providing greater flexibility and computational power in quantum information processing (see Fig. 1). However, the approach to concepts of information theory in the quantum setting is carried out with an analogous strategy to the classical setting [2], [3].

FIGURE 1. - (a) The classic bit has two well-defined values, either 0 or 1. (b) The qubit can be a superposition of the states 
$\left \vert {{0}}\right \rangle $
 and 
$\left \vert {{1}}\right \rangle $
 according to the following form 
$\left \vert {{\psi }}\right \rangle = \alpha \left \vert {{0}}\right \rangle + \beta \left \vert {{1}}\right \rangle $
, where 
$\alpha $
 and 
$\beta $
 are complex numbers and 
$|\alpha |^{2} + |\beta |^{2} = 1$
.
FIGURE 1.

(a) The classic bit has two well-defined values, either 0 or 1. (b) The qubit can be a superposition of the states \left \vert {{0}}\right \rangle and \left \vert {{1}}\right \rangle according to the following form \left \vert {{\psi }}\right \rangle = \alpha \left \vert {{0}}\right \rangle + \beta \left \vert {{1}}\right \rangle , where \alpha and \beta are complex numbers and |\alpha |^{2} + |\beta |^{2} = 1 .

The study of channel capacity is one of the applications of information theory and has a number of practical implications. In conceptual terms, the capacity of classical channels is a number that indicates the asymptotic rate whose information can be reliably transmitted through the channel. Due to features of quantum mechanics such as entanglement and superposition, it is possible to transmit information over quantum channels in a variety of ways. Naturally, the capacity of these channels can be defined in different ways, each depending on the type of information being transmitted, making it possible to transmit classical information or quantum states [4], [5]. Regarding the transmission of classical information, there are two examples of quantum channel capacity definitions. The quantum channel capacity was defined by Holevo-Schumacher-Westmoreland (HSW), as the maximum asymptotic rate in which classical information can be reliably transmitted using quantum encoding and decoding [6], [7]. Another example of definition for quantum channel capacity is the classical entanglement capacity C_{E} , which is defined as the supreme of transmission rates of classical information through a quantum channel when an infinite amount of entanglement is available between transmitter and receiver [8].

The capacities mentioned previously are defined considering that the probability decreases exponentially as the code size increases. Shannon defined the classical zero-error capacity of channels as the maximum rate achieved when the probability of error is equal to zero [9]. Since its definition, zero-error capacity has been studied in various scenarios [10]. The zero-error capacity of quantum channels was defined by Medeiros and Assis [11] as the supreme of rates in which classical information can be transmitted by a quantum channel with an error probability exactly equal to zero. Thus, using quantum block coding, classical information is mapped into quantum states and these states are transmitted through a noisy quantum channel. As part of the process, quantum states are measured on reception and, in this case, the transmission needs to be carried out with exactly zero-error rate.

An important aspect in the study of zero-error capacity of quantum channels is to verify whether a quantum channel has positive zero-error capacity or not. It is well known that determining the HSW capacity of a quantum channel is NP-Complete, while determining the zero-error capacity of a quantum channel is QMA-Complete [12]. However, even determining whether a channel’s capacity is positive can be a challenging problem [13]. Three conditions are known in the literature to verify the zero-error capacity condition of quantum channels, which are found in Medeiros and Assis [11], [14] and Gupta et al. [15]. In addition to these zero-error capacity conditions, Oliveira et al. [16], proves another condition to verify the zero-error capacity of quantum channels, which takes into account whether Kraus operators representing the quantum channel have at least two common eigenstates.

The relation between eigenstates common to the Kraus operators and the zero-error capacity of quantum channels is useful to relate quantum zero-error information theory and Shemesh theorem. The Shemesh theorem establishes a criterion that guarantees the existence of a common subspace for two square matrices. In other words, the theorem states that two square matrices A_{1} and A_{2} in M_{d} have common eigenstates if, and only if, the subspace\begin{equation*}\mathcal {M}= \cap _{r,s=1}^{d-1}\text {ker}\ [A_{1}^{r}, A_{2}^{s}]\end{equation*}

View SourceRight-click on figure for MathML and additional features.is a nontrivial [17]. Shemesh theorem is generalized in [18].

The main objective of this paper is to establish results concerning the relationship between Shemesh theorem and zero-error quantum information theory. The motivation behind this study stems from Shemesh theorem, which provides a criterion for linear operators to share a common eigenstate. Notably, every eigenstate common to the Kraus operators of a quantum channel is also a fixed point of the channel.

This paper is organized as follows: Section II presents the definition of a quantum channel and some basic properties. In addition, it discusses the zero-error capacity of a quantum channel, which is fundamental to understanding the main objective of this work. Section III presents the statement of Shemesh theorem along with its corresponding generalization. Section IV demonstrates results related to the connection between zero-error quantum information theory and Shemesh theorem. Finally, Section V presents the conclusions.

A. Related Work

Shemesh theorem presents a criterion for two square matrices to have a common eigenstate [17]. After the publication of the original idea by Dan Shemesh [17], other researchers proposed studies related to the Shemesh criterion. The criterion was generalized to a finite number of square matrices in [18].

One of the resulting problems related to the result proposed by Shemesh is the investigation of invariant common subspace for matrices. When it comes to investigating conditions for two square matrices A_{1} and A_{2} in M_{d} to have invariant common subspace, we can highlight the work of Alan and Khakim [19], which presents a procedure based on Shemesh’s theorem to check whether A_{1} and A_{2} have invariant common subspace, where the only restriction is that at least one of the eigenvalues A_{1} and A_{2} must be distinct. Yurii et al. proposed an algorithm for finding common invariant subspaces of dimension k=2 , 2 \leq k \lt d for matrices A_{1} and A_{2} [20]. Moreover, they proved that if the matrices A_{1} and A_{2} generate a semisimple algebra, the algorithm can solve the problem of finding a common invariant subspace for any k. Tsatsomeros [21] proposed an algorithm ensuring that a subspace of any dimension k is a common invariant subspace for A_{1} and A_{2} , if and only if, for some scalar \tau , A_{1}+\tau \mathbb {I} and A_{2}+\tau \mathbb {I} are invertible, and their k-th compounds have a common eigenstate, which is a Grassmann representative of the subspace.

Jamiołkowski and Pastuszak [18] presented a criterion guaranteeing the existence of a common invariant subspace for a finite number of matrices in M_{d} , based on the work of [19] and [21]. Another approach taken as a consequence of the Shemesh theorem is by Arapura and Peterson [22] who, using Gröbner’s bases [23], presented a criteria for determining the existence and quantity of a one-dimensional subspace common to a set of matrices. Grzegorz [24] presented a criterion for a finite number of matrices to have a common invariant subspace the criterion is based on a constructive proof of Tarski’s theorem [25].

The concept of a common invariant subspace plays a significant role in Quantum Information Theory. Jamiołkowski [26] used the Shemesh theorem and the generalized Shemesh theorem to analyze the properties of quantum control systems. Farenick [27] related the concept of invariant common subspace to irreducible quantum channels, showing that a quantum channel represented by Kraus operators is irreducible if, and only if, the Kraus operators have no common invariant subspace. Another relevant question is the relationship between common invariant subspace and the concept of decoherence-free subspace (DFS) [28], [29]. To understand this relationship, we should consider that quantum channels are used to transmit information, but the information can be corrupted by decoherence [30]. To reduce the effect of decoherence, one can hide the information in a decoherence-free subspace (DFS). Using the generalized Shemesh theorem, one can decide when the algebra generated by the Kraus operators of the quantum channel has a DFS with dimension k \geq 2 .

It is also important to highlight the work of Białonczyk et al. [31], which uses the Shemesh theorem to analyze the spectral properties of the Kraus operators associated with a quantum channel, as well as the connection that these spectral properties have with the algebra generated by the Kraus operators of the quantum channel. The spectral properties of the Kraus operators of a quantum channel are essential for determining the asymptotic dynamics of quantum systems subject to multiple interactions described by the same quantum channel.

In the context of Quantum Zero-Error Information Theory, we found no existing literature that explores a connection between Shemesh theorem and this field. More specifically, our research indicates that there are no published works linking Shemesh theorem to the concept of zero-error capacity in quantum channels, highlighting the novelty of this investigation.

B. Notations and Definitions

In this paper, we always refer to Hilbert spaces of finite dimension d and denote them by \mathcal {H} . By \mathcal {B}(\mathcal {H}) we will denote the set of all linear continuous operators on \mathcal {H} . By M_{d}(\mathbb {C}) we denote the set of all complex square matrices of order d. By assuming that \mathcal {H}\cong \mathbb {C}^{d} and \mathcal {B}(\mathcal {H})\cong M_{d}(\mathbb {C}) , then the quantum channel \mathcal {E} can be viewed as a trace-preserving quantum superoperator \mathcal {E}: \mathcal {B}(\mathcal {H})\longrightarrow \mathcal {B}(\mathcal {H}) , or \mathcal {E}: M_{d}(\mathbb {C})\longrightarrow M_{d}(\mathbb {C}) .

SECTION II.

Zero-Error Capacity of Quantum Channels

In this section we discuss the main definitions and known results on zero-error capacity of quantum channels is presented, which are important to facilitate the reading and understanding of this paper [11], [32].

A quantum channel \mathcal {E} defined on \mathcal {H} can be modeled by a completely positive and trace-preserving linear map of the density operators, \mathcal {E} \equiv \{A_{i}\} , where A_{i} are Kraus operators on \mathcal {H} and satisfy the condition \sum _{i}^{\kappa }A_{i}^{\dagger } A_{i}=\mathbb {I} and \kappa \leq d^{2} .

Let \mathcal {X} \subset \mathcal {H} be a set of possible input states for the quantum channel \mathcal {E} . If \rho \in \mathcal {X} , which is denoted by \sigma = \mathcal {E}(\rho) , the quantum state received when \rho is transmitted over the quantum channel can be written as\begin{equation*} \mathcal {E}(\rho)=\sum _{i=1}^{\kappa }A_{i}\rho A_{i}^{\dagger }. \tag {1}\end{equation*}

View SourceRight-click on figure for MathML and additional features.

When considering a quantum channel \mathcal {E} , the communication protocol associated with the zero-error capacity can be summarized as follows. Define the finite subset \mathcal {S}=\{\rho _{1},\ldots,\rho _{\ell } \} \subset \mathcal {X} of input states as the alphabet of the zero-error quantum code. The tensor product of any n states of \mathcal {S} forms the quantum code words of length n, \overline {\rho }_{i}=\rho _{i_{1}}\otimes \ldots \otimes \rho _{i_{n}} . Denote by \mathcal {S}^{\otimes n} the set of all quantum code words of length n. Codewords are used to encode classic messages before they are sent over the channel. If Bob performs measurements using a POVM (Positive Operator-Valued Measurements) \{M_{j}\} , where \sum _{j}^{}M_{j}= \mathbb {I} , then p(j|i) is defined as the probability of Bob measuring j given that Alice sent the state \rho _{i} . Thus,\begin{equation*} p(j|i)=\text {tr}[\sigma _{i}M_{j}]=\text {tr} [\mathcal {E}(\rho _{i})M_{j}]. \tag {2}\end{equation*}

View SourceRight-click on figure for MathML and additional features.

A quantum (m,n) zero-error code for \mathcal {E} is composed of:

  1. a set of indices \{1,\ldots,m\} , where each index is associated with a classical message;

  2. a coding function\begin{equation*} f_{n}:\{1,\ldots,m\}\longrightarrow \mathcal {S}^{\otimes n} \tag {3}\end{equation*}

    View SourceRight-click on figure for MathML and additional features.leading to codewords f_{n}(1)=\overline {\rho _{1}},\ldots, f_{n}(m)=\overline {\rho _{m}} , where each \overline {\rho _{i}} \in \mathcal {S}^{\otimes n} ;

  3. A decoding function\begin{equation*} g:\{1,\ldots,k\}\longrightarrow \{1,\ldots,m\} \tag {4}\end{equation*}

    View SourceRight-click on figure for MathML and additional features.which deterministically associates a message with one of the possible measurement y \in \{1,\ldots,k\} performed by POVM \{M_{i}\}_{i=1}^{k} . Furthermore, the decoding function has the following property:\begin{equation*} \text { Pr}[g(\mathcal {E}(f_{n}(i)))\neq i]=0 \tag {5}\end{equation*}
    View SourceRight-click on figure for MathML and additional features.
    for all i \in \{1,\ldots,m\} .

The rate of a (m,n) code is given by\begin{equation*} R=\frac {1}{n} \log \ m\ \text { bits/use.} \tag {6}\end{equation*}
View SourceRight-click on figure for MathML and additional features.

Definition 1 Zero-Error Capacity of Quantum Channels[11]):

The quantum zero-error capacity of a quantum channel \mathcal {E}(\cdot) , denoted by C^{(0)}(\mathcal {E}) , is the supreme of the achievable rates with decoding error probability equal to zero,\begin{equation*} C^{(0)}(\mathcal {E})=\text {sup}_{\mathcal {S}}\text {sup}_{n}\ \frac {1}{n} \text { log}\ m \tag {7}\end{equation*}

View SourceRight-click on figure for MathML and additional features.where m is the maximum number of classical messages the system can transmit without error when a zero-error quantum block code (m,n) is used and the input alphabet is \mathcal {S} .

The decoding error is related to the ability to distinguish between two states. Two quantum states are distinguishable if, and only if, the Hilbert spaces generated by the supports of these quantum states are orthogonal. Thus, two quantum states \rho _{i},\rho _{j} \in \mathcal {S} , i\neq j , then \rho _{i} and \rho _{j} are said to be non-adjacent (or distinguishable) at the output of the quantum channel \mathcal {E} if \mathcal {E}(\rho _{i}) and \mathcal {E}(\rho _{j}) belong to orthogonal Hilbert subspaces. Otherwise, \rho _{i} and \rho _{j} are said to be adjacent (or indistinguishable) in the output of \mathcal {E} .

Consequently, certain tests for determining whether a channel has positive zero-error capacity (e.g., [11, Proposition 1] and [14, Proposition 1]) assert that a quantum channel \mathcal {E} possesses positive zero-error capacity if, and only if, there exist at least two non-adjacent quantum states. The zero-error capacity condition presented above is equivalently proved in [15, Lemma 4.4.1] since the zero-error capacity of the channel \mathcal {E} is positive if, and only if, | {\rho _{m}}{\rho _{m'}}| is orthogonal to the subspace\begin{equation*} S:=\text {span}\{A_{i}^{\dagger } A_{j}\} \tag {8}\end{equation*}

View SourceRight-click on figure for MathML and additional features.where \left \vert {{\rho _{m}}}\right \rangle and \left \vert {{\rho _{m'}}}\right \rangle with m\neq m' are input states in the channel, and A_{i} and A_{j} are the Kraus operators representing the quantum channel \mathcal {E} .

There is a relation between zero-error capacity and fixed point of the quantum channel \mathcal {E} . Schauder’s fixed point theorem [33, Theorem 2.3.7] guarantees a quantum channel has at least one quantum state \rho that is fixed point for the quantum channel \mathcal {E} , i.e., \mathcal {E} (\rho)=\rho .

Proposition 2[11]:

Let \mathcal {E} be a quantum channel with N_{f} fixed points, then the zero-error capacity of \mathcal {E} is at least log\ N_{f} .

The proof of this result can be found in [11, Proposition 2].

Next, it is proved every eigenstate common to all Kraus operators representing the quantum channel is a fixed point of this channel. Based on this result, one can conclude that zero-error capacity of quantum channel is positive if all the Kraus operators have at least two common eigenstates.

Lemma 3[16]:

Let a quantum channel be \mathcal {E}: M_{d}\longrightarrow M_{d} with Kraus operators A_{1},\ldots,A_{\kappa } . If \left \vert {{\psi }}\right \rangle is a common eigenstate of the operators A_{i} , then it is a fixed point to \mathcal {E} .

Proof:

We just need to show that \mathcal{E}(|\psi\rangle\langle\psi|)=|\psi\rangle\langle\psi| . In general, consider that A_{i}=\lambda _{i} \left \vert {{\psi }}\right \rangle can be associated with different eigenvalues to different Kraus operators. Then,\begin{align*} \mathcal{E}(|\psi\rangle\langle\psi|) & \stackrel{(\mathrm{a})}{=} \sum_{i=1}^\kappa A_i|\psi\rangle\langle\psi| A_i^{\dagger} \stackrel{(\mathrm{b})}{=} \sum_{i=1}^\kappa \lambda_i|\psi\rangle \lambda_i^*\langle\psi| \tag{9}\\ & \stackrel{(\mathrm{c})}{=}|\psi\rangle\langle\psi| \sum_{i=1}^\kappa\left|\lambda_i\right|^2 \stackrel{(\mathrm{~d})}{=}|\psi\rangle\langle\psi|, \tag{10}\end{align*}

View SourceRight-click on figure for MathML and additional features.where in (a) we used Kraus representation of the quantum channel, in (b) the fact that \left \vert {{\psi }}\right \rangle is an eigenstate of every A_{i} (by hypothesis) and (d) because \mathcal {E} is trace preserving.□

Define N_{\mathcal {E}} = \{ \left \vert {{\psi }}\right \rangle \left \langle {{\psi }}\right \vert \in \mathcal {S}\ :\ A_{i}\left \vert {{\psi }}\right \rangle =\lambda _{i}\left \vert {{\psi }}\right \rangle,\ i=1,\ldots,\kappa \} as the set of eigenstates the are common to every Kraus operator for the quantum channel \mathcal {E} . Denote by |N_{\mathcal {E}}| the cardinality of N_{\mathcal {E}} .

Theorem 4[16]:

Let N_{f} be the number of fixed points of a quantum channel \mathcal {E} . Then, N_{f} \geq |N_{\mathcal {E}}| and C^{(0)}(\mathcal {E}) \geq \text { log}\ |N_{\mathcal {E}}| .

Proof:

The Lemma 3 showed that states in N_{\mathcal {E}} are fixed points of the channel. Since a fixed point does not need to be a common state of each Kraus operator, the inequality N_{f} \geq |N_{\mathcal {E}}| is valid. For the zero-error capacity lower bound, one can construct a trivial quantum block code with zero-error probability by encoding classical information in the states in N_{\mathcal {E}} and then it is possible to transmit at least \text {log}\ |N_{\mathcal {E}}| bits through the quantum channel without error. So, we have C^{(0)}(\mathcal {E}) \geq \text { log}\ |N_{\mathcal {E}}| .□

The condition for positive zero-error capacity of a quantum channel established in Theorem 4 is highly practical. Given the d possible distinct eigenstates in the Hilbert space, it is sufficient to verify that at least two of these eigenstates are common to the Kraus operators representing the channel.

SECTION III.

The Shemesh Theorem

In order to show the relation between Shemesh theorem and quantum zero-error information theory, in this section we will present Shemesh theorem in the non-generalized version with proof, and state the generalized theorem.

Theorem 5 (Shemesh Theorem[17]):

Let A_{1} and A_{2} be M_{d} . Then A_{1} and A_{2} have common eigenstates if, and only if, the subspace\begin{equation*} \mathcal {M}= \cap _{r,s=1}^{d-1}\text {ker}\ [A_{1}^{r}, A_{2}^{s}] \tag {11}\end{equation*}

View SourceRight-click on figure for MathML and additional features.is a nontrivial subspace, that is, some \left \vert {{\varphi }}\right \rangle \in \mathcal {M} with 0 \neq \left \vert {{\varphi }}\right \rangle \in \mathbb {C}^{d} . The symbol [\cdot, \cdot] denotes the matrix commutator and r,s \in \{1,2,\ldots,d-1\} , while A_{1}^{r} and A_{2}^{s} denote the r-th and s-th powers of A_{1} and A_{2} , respectively.

The demonstration of Shemesh theorem can be found in [17].

Observation 6:

The subspace of Shemesh theorem is contained in \mathbb {C}^{d} , i.e.,\begin{equation*} \mathcal {M}= \cap _{r,s=1}^{d-1}\text {ker}[A_{1}^{r}, A_{2}^{s}] \subset \mathbb {C}^{d}. \tag {12}\end{equation*}

View SourceRight-click on figure for MathML and additional features.

The Theorem 5 can be generalized to a quantity s of matrices and give conditions for them to have common eigenstates. This generalization occurs due to the use of the standard polynomial concept and the Amitsur-Levitzki theorem. The generalized Shemesh Theorem is stated below and it can be found in [18, Theorem 2.4].

Theorem 7 (Generalized Shemesh Theorem[18]):

The matrices A_{1},\ldots, A_{\kappa } \in M_{d} have common eigenstates if, and only if, the subspace\begin{equation*} \mathcal {M}= \cap _{\alpha _{i},l_{i}=1}^{d-1}\text {ker}\ [A_{1}^{\alpha _{1}}\ldots \ A_{\kappa }^{\alpha _{\kappa }}, A_{1}^{l_{1}}\ldots \ A_{\kappa }^{l_{\kappa }}] \tag {13}\end{equation*}

View SourceRight-click on figure for MathML and additional features.is a nontrivial subspace, or in other words, for some x \in \mathcal {M} with 0\neq x \in \mathbb {C}^{d} and \sum _{i}^{}\alpha _{i}\neq 0 and \sum _{i}^{}l_{i}\neq 0 .

Note that the sufficient and necessary condition for matrices A_{1},\ldots, A_{\kappa } \in M_{d} to have a common eigenstate is that Shemesh subspace \mathcal {M} is nontrivial. Initially, the nonzero vectors belonging to \mathcal {M} are not necessarily common eigenstates of A_{i} . In the following section it is proved that if \mathcal {M} is nontrivial, then the eigenstates common to A_{i} belong to \mathcal {M} .

SECTION IV.

The Zero-Error Capacity of Quantum Channels and Shemesh Theorem

In this section, results are presented so that we can construct the relation between Shemesh theorem and quantum zero-error information theory.

Lemma 8:

Let A \in M_{d} and an associated eigenstate \left \vert {{\varphi }}\right \rangle . Then for m\gt 0 , \left \vert {{\varphi }}\right \rangle is eigenstate of A^{m} associated with the eigenvalue \lambda ^{m} . In other words, A^{m}\left \vert {{\varphi }}\right \rangle =\lambda ^{m}\left \vert {{\varphi }}\right \rangle .

Proof:

By hypothesis \left \vert {{\varphi }}\right \rangle is eigenstate associated to \alpha , so for m=2 we have\begin{equation*} A^{2}\left \vert {{\varphi }}\right \rangle = A(A \left \vert {{\varphi }}\right \rangle)= A(\lambda \left \vert {{\varphi }}\right \rangle)=\lambda A\left \vert {{\varphi }}\right \rangle =\lambda ^{2}\left \vert {{\varphi }}\right \rangle. \tag {14}\end{equation*}

View SourceRight-click on figure for MathML and additional features.By induction hypothesis suppose that the result holds for k, then\begin{equation*} A^{k}\left \vert {{\varphi }}\right \rangle =\lambda ^{k}\left \vert {{\varphi }}\right \rangle. \tag {15}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
Thus,\begin{align*} A^{k+1}\left \vert {{\varphi }}\right \rangle =A(A^{k}\left \vert {{\varphi }}\right \rangle)=A(\lambda ^{k}\left \vert {{\varphi }}\right \rangle)=\lambda ^{k}\lambda \left \vert {{\varphi }}\right \rangle =\lambda ^{k+1}\left \vert {{\varphi }}\right \rangle. \tag {16}\end{align*}
View SourceRight-click on figure for MathML and additional features.
Therefore, the result holds for all m\gt 0 .□

Proposition 9:

Let be a quantum channel \mathcal {E}: M_{d}\longrightarrow M_{d} with Kraus operators A_{1},\ldots,A_{\kappa } . Then, every eigenstate of N_{\mathcal {E}} belongs to the subspace\begin{equation*} \mathcal {M}= \cap _{\alpha _{i},l_{i}=1}^{d-1}\text {ker}\ [A_{1}^{\alpha _{1}}\ldots A_{\kappa }^{\alpha _{\kappa }}, A_{1}^{l_{1}}\ldots A_{\kappa }^{l_{\kappa }}] \tag {17}\end{equation*}

View SourceRight-click on figure for MathML and additional features.with \sum _{i}^{}\alpha _{i}\neq 0 and \sum _{i}^{}l_{i}\neq 0 .

Proof:

Suppose that \left \vert {{\varphi }}\right \rangle \in N_{\mathcal {E}} , then \left \vert {{\varphi }}\right \rangle is common eigenstate to A_{1},\ldots,A_{\kappa } , that is,\begin{align*} A_{1} \left \vert {{\varphi }}\right \rangle = \lambda _{1} \left \vert {{\varphi }}\right \rangle,\ A_{2} \left \vert {{\varphi }}\right \rangle = \lambda _{2} \left \vert {{\varphi }}\right \rangle,\ldots,A_{\kappa }\left \vert {{\varphi }}\right \rangle = \lambda _{\kappa }\left \vert {{\varphi }}\right \rangle. \tag {18}\end{align*}

View SourceRight-click on figure for MathML and additional features.By using the Theorem 7 we have that the subspace \mathcal {M} \neq 0 , so suppose a_{1},\ldots,a_{\kappa } \in \{1,2,\ldots,d-1\} and b_{1},\ldots,b_{\kappa } \in \{1,2,\ldots,d-1\} and consider the commutator\begin{align*}& [A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }},A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }}] \\ & = (A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})-(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }}). \tag {19}\end{align*}
View SourceRight-click on figure for MathML and additional features.

Note that\begin{align*}& ((A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})-(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }}))\left \vert {{\varphi }}\right \rangle \tag {20}\\ & =(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa -1}^{b_{\kappa -1}})A_{\kappa }^{b_{\kappa }}\left \vert {{\varphi }}\right \rangle \\ & \quad -(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa -1}^{a_{\kappa -1}})A_{\kappa }^{a_{\kappa }}\left \vert {{\varphi }}\right \rangle \tag {21}\\ & =(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa -1}^{b_{\kappa -1}})\lambda _{\kappa }^{b_{\kappa }}\left \vert {{\varphi }}\right \rangle \\ & \quad -(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa -1}^{a_{\kappa -1}})\lambda _{\kappa }^{a_{\kappa }}\left \vert {{\varphi }}\right \rangle \tag {22}\\ & =\lambda _{\kappa }^{b_{\kappa }}(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa -1}^{b_{\kappa -1}})\left \vert {{\varphi }}\right \rangle \\ & \quad -\lambda _{\kappa }^{a_{\kappa }}(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa -1}^{a_{\kappa -1}})\left \vert {{\varphi }}\right \rangle \tag {23}\\ & =\lambda _{\kappa }^{b_{\kappa }}(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa -2}^{b_{\kappa -2}})A_{\kappa -1}^{b_{\kappa -1}}\left \vert {{\varphi }}\right \rangle \\ & \quad -\lambda _{\kappa }^{a_{\kappa }}(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa -2}^{a_{\kappa -2}})A_{\kappa -1}^{a_{\kappa -1}}\left \vert {{\varphi }}\right \rangle \tag {24}\\ & =\lambda _{\kappa }^{b_{\kappa }}\lambda _{\kappa -1}^{b_{\kappa -1}}(A_{1}^{a_{1}}\ldots A_{\kappa }^{a_{\kappa }})(A_{1}^{b_{1}}\ldots A_{\kappa -2}^{b_{\kappa -2}})\left \vert {{\varphi }}\right \rangle \\ & \quad -\lambda _{\kappa }^{a_{\kappa }}\lambda _{\kappa -1}^{a_{\kappa -1}}(A_{1}^{b_{1}}\ldots A_{\kappa }^{b_{\kappa }})(A_{1}^{a_{1}}\ldots A_{\kappa -2}^{a_{\kappa -2}})\left \vert {{\varphi }}\right \rangle \\ & \quad \vdots \hspace {2cm} \vdots \hspace {2cm} \vdots \hspace {2cm} \vdots \tag {25}\\ & =\lambda _{\kappa }^{b_{\kappa }}\lambda _{\kappa -1}^{b_{\kappa -1}}\ldots \lambda _{1}^{b_{1}} \lambda _{\kappa }^{a_{\kappa }}\lambda _{\kappa -1}^{a_{\kappa -1}}\ldots \lambda _{1}^{a_{1}}\left \vert {{\varphi }}\right \rangle \\ & \quad -\lambda _{\kappa }^{a_{\kappa }}\lambda _{\kappa -1}^{a_{\kappa -1}}\ldots \lambda _{1}^{a_{1}}\lambda _{\kappa }^{b_{\kappa }}\lambda _{\kappa -1}^{b_{\kappa -1}}\ldots \lambda _{1}^{b_{1}}\left \vert {{\varphi }}\right \rangle \tag {26}\\ & =(\lambda _{\kappa }^{b_{\kappa }+a_{s}}\lambda _{\kappa -1}^{b_{\kappa -1}+a_{\kappa -1}}\ldots \lambda _{1}^{b_{1}+a_{1}} \\ & \quad -\lambda _{\kappa }^{a_{\kappa }+b_{\kappa }}\lambda _{\kappa -1}^{a_{\kappa -1}+b_{\kappa -1}}\ldots \lambda _{1}^{a_{1}+b_{1}})\left \vert {{\varphi }}\right \rangle \tag {27}\\ & = 0 \cdot \left \vert {{\varphi }}\right \rangle =0. \hspace {5.74cm} \tag {28}\end{align*}

View SourceRight-click on figure for MathML and additional features.In other words, the eigenstate \left \vert {{\varphi }}\right \rangle belongs to the subspace\begin{equation*} \mathcal {M}= \cap _{\alpha _{i},l_{i}=1}^{d-1}\text {ker}\ [A_{1}^{\alpha _{1}}\ldots A_{\kappa }^{\alpha _{\kappa }}, A_{1}^{l_{1}}\ldots A_{\kappa }^{l_{\kappa }}]. \tag {29}\end{equation*}
View SourceRight-click on figure for MathML and additional features.
Therefore, every eigenstate of N_{\mathcal {E}} belongs to the subspace \mathcal {M} .□

The Proposition 9 shows that if \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in N_{\mathcal {E}} , then \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in \mathcal {M} , i.e., N_{\mathcal {E}} \subseteq \mathcal {M} . In addition, consider the following set\begin{equation*}{{\mathcal {M}}_{\mathcal {E}}=\{ \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in \mathcal {M}: \mathcal {E}(\left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert)=\left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \}},\end{equation*}

View SourceRight-click on figure for MathML and additional features.with at least one \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in N_{\mathcal {E}} . Denote by |{\mathcal {M}}_{\mathcal {E}}| the cardinality of {\mathcal {M}}_{\mathcal {E}} . Therefore, based on this idea, we have the following corollary.

Corollary 10:

|{\mathcal {M}}_{\mathcal {E}}| \geq |N_{\mathcal {E}}| .

Proof:

Let \mathcal {E} be a quantum channel with Kraus operators A_{1},\ldots,A_{\kappa } . Then it is possible to have a \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in {\mathcal {M}}_{\mathcal {E}} which is not a common eigenstate of the operators A_{1},\ldots,A_{\kappa } , i.e., \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in {\mathcal {M}}_{\mathcal {E}} and \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \notin N_{\mathcal {E}} . Therefore, |{\mathcal {M}}_{\mathcal {E}}| \geq |N_{\mathcal {E}}| .□

Proposition 11:

If |{\mathcal {M}}_{\mathcal {E}}| \geq 2 then C^{(0)}(\mathcal {E})\gt 0 .

Proof:

By hypothesis the states belongs to {\mathcal {M}}_{\mathcal {E}} are fixed points of \mathcal {E} and we have at least two states in {\mathcal {M}}_{\mathcal {E}} , so it follows that C^{(0)}(\mathcal {E}) \geq \text { log}\ |{\mathcal {M}}_{\mathcal {E}}| .□

The result of Proposition 11 establishes a connection between Shemesh theorem and the zero-error capacity of the quantum channel.

Example 12:

Let \mathcal {E} be the quantum channel with Kraus operators A_{1} and A_{2} , where p \in (0,1) :\begin{align*}A_{1}& = \sqrt {p}\left ({{ \begin{array}{ccccc} 1 & \quad 0 & \quad 0 & \quad 0 \\[7pt] 0 & \quad \dfrac {1}{2} & \quad \dfrac {\sqrt {3}}{2} & \quad 0 \\[7pt] 0 & \quad \dfrac {\sqrt {3}}{2} & \quad -\dfrac {1}{2} & \quad 0 \\[7pt] 0 & \quad 0 & \quad 0 & \quad 1 \\ \end{array} }}\right), \\ A_{2}& = \sqrt {1-p}\left ({{ \begin{array}{ccccc} 1 & \quad 0 & \quad 0 & \quad 0 \\[7pt] 0 & \quad \dfrac {1}{2} & \quad - \dfrac {\sqrt {3}}{2} & \quad 0 \\[7pt] 0 & \quad -\dfrac {\sqrt {3}}{2} & \quad -\dfrac {1}{2} & \quad 0 \\[7pt] 0 & \quad 0 & \quad 0 & \quad 1 \\ \end{array} }}\right).\end{align*}

View SourceRight-click on figure for MathML and additional features.For p=\frac {1}{4} , we have that A_{1} and A_{2} have the form:\begin{align*}A_{1}= \left ({{ \begin{array}{ccccc} \dfrac {1}{2} & \quad 0 & \quad 0 & \quad 0 \\[7pt] 0 & \quad \dfrac {1}{4} & \quad \dfrac {\sqrt {3}}{4} & \quad 0 \\[7pt] 0 & \quad \dfrac {\sqrt {3}}{4} & \quad -\dfrac {1}{4} & \quad 0 \\[7pt] 0 & \quad 0 & \quad 0 & \quad \dfrac {1}{2} \\ \end{array} }}\right), \\ A_{2}= \left ({{ \begin{array}{ccccc} \dfrac {\sqrt {3}}{2} & \quad 0 & \quad 0 & \quad 0 \\[7pt] 0 & \quad \dfrac {\sqrt {3}}{4} & \quad - \dfrac {3}{4} & \quad 0 \\[7pt] 0 & \quad -\dfrac {3}{4} & \quad -\dfrac {\sqrt {3}}{4} & \quad 0 \\[7pt] 0 & \quad 0 & \quad 0 & \quad \dfrac {\sqrt {3}}{2} \\ \end{array} }}\right).\end{align*}
View SourceRight-click on figure for MathML and additional features.
The eigenstates \left \vert {{\psi }}\right \rangle =(1,0,0,0) and \left \vert {{\varphi }}\right \rangle =(0,0,0,1) are common to A_{1} and A_{2} , so by the lemma 3 the eigenstates \left \vert {{\psi }}\right \rangle and \left \vert {{\varphi }}\right \rangle are fixed points of the quantum channel. Therefore, \left \vert {{\psi }}\right \rangle \left \langle {{\psi }}\right \vert, \left \vert {{\varphi }}\right \rangle \left \langle {{\varphi }}\right \vert \in {\mathcal {M}}_{\mathcal {E}} , i.e., |{\mathcal {M}}_{\mathcal {E}}| \geq 2 and by the Proposition 11 we have that C^{(0)}(\mathcal {E})\geq \log 2 .

SECTION V.

Conclusion

In this paper, we established a relationship between Shemesh theorem and zero-error quantum information theory. We first defined a quantum channel using Kraus operators on a finite-dimensional Hilbert space and introduced the concept of zero-error capacity. We particularly emphasized a test for zero-error capacity based on the eigenstates common to the Kraus operators. We then analyzed Shemesh theorem and demonstrated its connection to zero-error quantum information theory by proving that every eigenstate common to the Kraus operators, which serves as a fixed point of the channel, belongs to the Shemesh subspace. It is also important to recognize that there may be additional fixed points within the Shemesh subspace beyond those eigenstates common to the Kraus operators. Given that a lower bound for the zero-error capacity of a quantum channel can be associated with the number of its fixed points, we highlight the connection between Shemesh theorem and zero-error quantum information theory, particularly in the context of zero-error capacity.

References

References is not available for this document.