[table]capposition=top \floatsetup[figure]capposition=bottom \newfloatcommandcapbtabboxtable[][\FBwidth]
A two-stage solution to quantum process tomography: error analysis and optimal design
Abstract
Quantum process tomography is a critical task for characterizing the dynamics of quantum systems and achieving precise quantum control. In this paper, we propose a two-stage solution for both trace-preserving and non-trace-preserving quantum process tomography. Utilizing a tensor structure, our algorithm exhibits a computational complexity of where is the dimension of the quantum system and , represent the numbers of different input states and measurement operators, respectively. We establish an analytical error upper bound and then design the optimal input states and the optimal measurement operators, which are both based on minimizing the error upper bound and maximizing the robustness characterized by the condition number. Numerical examples and testing on IBM quantum devices are presented to demonstrate the performance and efficiency of our algorithm.
Index Terms:
Quantum process tomography, quantum system identification, error analysis, quantum system.I Introduction
In the past decades, quantum science and technologies have significantly advanced in many fields such as quantum computation [1, 2], quantum communication [3, 4], quantum sensing [5] and quantum control [6, 7, 8]. To fully develop and realize these technologies, a fundamental challenge to overcome is to characterize unknown quantum dynamics. A typical framework to formulate this problem is quantum process tomography (QPT), where the parameters determining the map from input quantum states to output states need to be estimated.
For a closed quantum system, QPT is reduced to Hamiltonian identification and many results have been obtained in this scenario [9, 10, 11, 12, 13, 14]. For an open quantum system characterized by time-independent parameters, the process is completely positive (CP) and attention is usually restricted to the trace-preserving (TP) case. Using direct and convex optimization methods, Zorzi et al. [15, 16] gave the minimum positive operator-valued measure (POVM) resources required to estimate a CPTP quantum process, and Knee et al. [17] and Surawy-Stepney et al. [18] proposed iterative projection algorithms to identify such a process. The efficiency of parameter estimation for quantum processes was studied in [19]. Zhang and Sarovar [20] utilized the time traces of observable measurements to identify the parameters in a master equation. The identifiability and identification for the passive quantum systems were investigated in [21]. When the quantum process is non-trace-preserving (non-TP), Huang et al. [22] proposed a convex optimization method and Bongioanni et al. [23] demonstrated non-TP QPT based on a maximum likelihood approach in an experiment. In practice, a quantum process may be affected by non-Markovian noise. White et al. [24] developed a framework to characterize non-Markovian dynamics and the work in [25] proposed process tensor tomography for non-Markovian QPT. Moreover, adaptive strategies for QPT were proposed in [26, 27, 28].
In this paper, we focus on Standard Quantum Process Tomography (SQPT) [2, 14, 29, 30, 31] where states with the same dimension as the process are inputted, and the corresponding outputs are measured to reconstruct the process. We extend the framework in [2, 13] to allow for general input states and non-TP processes, and also directly connect the process with measurement data in a unified system equation. Using a mathematical representation, we deduce the special structural properties of the system equation, allowing for more efficient analysis, solution and optimal design. We give an analytical two-stage solution (TSS) to a general QPT problem, as a closed-form estimation which is not common among the existing QPT solutions. For -dimensional quantum systems, our algorithm has computational complexity and storage requirement , where is the number of different informationally complete or overcomplete input states and is the number of different measurement operators. We also establish an analytical error upper bound for our algorithm. Then we study the optimization of input states and measurement operators, based on this error upper bound and the condition number of the system equation, which reflects the robustness of our algorithm. We give theorems to characterize lower bounds on the error and condition number for both input states and measurement operators. We prove that SIC (symmetric informationally complete) states and MUB (mutually unbiased bases) states [32] are two examples achieving the lower bounds for our algorithm. We also prove MUB measurement achieves the lower bounds among measurement operators. Numerical examples and testing on IBM Quantum devices demonstrate the effectiveness of our algorithm and validate the theoretical error analysis. We compare our TSS algorithm with the convex optimization method in [14] and the results show that our algorithm is more efficient in both time and space costs during calculation. The main contributions of this paper are summarized as follows.
-
(i)
We propose an analytical two-stage solution (TSS) to a general QPT problem. Our TSS algorithm does not require specific prior knowledge about the unknown process and can be applied in both TP and non-TP quantum processes.
-
(ii)
Utilizing a tensor structure, our TSS algorithm has the computational complexity and a clear storage requirement , resulting from our closed-form estimation formula.
-
(iii)
Our TSS algorithm can give an analytical error upper bound which can be further utilized to optimize the input states and the measurement operators.
-
(iv)
Numerical examples and testing on IBM Quantum devices demonstrate the theoretical results and the effectiveness of our method. Compared to the convex optimization method in [14], our algorithm is more efficient in both time and space costs.
The organization of this paper is as follows. Section II introduces some preliminary knowledge and establishes our framework for QPT. Section III studies the tensor structure and presents our TSS algorithm for QPT. In Section IV, we analyze the computational complexity and the storage requirement. In Section V, we show the error analysis and in Section VI, we study the optimization of the input states and the measurement operators. Numerical examples are presented in Section VII and Section VIII concludes this paper.
Notation: The -th row and -th column of a matrix is . The -th column of is . The transpose of is . The conjugate and transpose of is . The sets of real and complex numbers are and , respectively. The sets of -dimension complex vectors and complex matrices are and , respectively. The identity matrix is . . The trace of is . The Frobenius norm of a matrix is denoted as and the 2-norm of a vector is . We use density matrix to represent a quantum state where , and and use a unit complex vector to represent a pure state. The estimate of is . The inner product of two matrices and is defined as . The inner product of two vectors and is defined as . The tensor product of and is denoted . Hilbert space is . The Kronecker delta function is . The denotes a diagonal matrix with the -th diagonal element being the -th element of the vector . For any positive semidefinite with spectral decomposition , we define or as .
II Preliminaries and quantum process tomography
II-A Vectorization function
We introduce the (column-)vectorization function . For a matrix ,
(1) | ||||
Thus, is a linear and basis dependent map. We also define the inverse which maps a vector into a square matrix. The common properties of are listed as follows [33, 34]:
(2) |
(3) |
(4) |
where denotes the partial trace on the space with belonging to the space .
II-B Quantum process tomography
According to the system architecture, QPT generally can be divided into three classes: Standard Quantum Process Tomography (SQPT) [2, 14, 29, 30, 31], Ancilla-Assisted Process Tomography (AAPT) [35, 36, 37, 38] and Direct Characterization of Quantum Dynamics (DCQD) [39, 40, 41]. In SQPT, one inputs quantum states with the same dimension as the process, and then we reconstruct the process from quantum state tomography (QST) on the output states. In AAPT and DCQD, an auxiliary system (ancilla) is attached to the principal system, and the input states and the measurement of the outputs are both carried out on the extended Hilbert space. AAPT transforms QPT to QST in a larger Hilbert space and the input state can be separable [42]. DCQD requires entangled input states and measurements such that the measured probability distributions are directly related to the elements of the process representation [42]. Since AAPT needs high dimensional states in the extended Hilbert space and DCQD usually requires entanglement which are both more technologically demanding, we focus on the SQPT scenario in this paper.
Here, we briefly introduce the SQPT framework as given in [2, 13] and extend it to a more general framework based on the following two aspects:
-
(i)
Input Fock states are extended to be arbitrary states.
-
(ii)
The unknown process is extended from the TP case to both the TP and non-TP cases.
For a -dimensional quantum system, its dynamics can be described by a completely-positive (CP) linear map . If we input a quantum state (input state), using the Kraus operator-sum representation [2], the output state is given by
(5) |
where and satisfy
(6) |
When the equality in (6) holds, the map is said to be trace-preserving (TP). Otherwise, it is non-trace-preserving (non-TP). Let be a fixed basis of . From , we can obtain process matrix [2, 13] which is a Hermitian and positive semidefinite matrix such that
(7) |
In this paper, we consider both TP and non-TP quantum processes and (6) becomes
(8) |
Since the relationship between and is one-to-one [2, 13], the identification of is equivalent to identifying . The number of independent elements in is for the TP case and for the non-TP case.
Let be a complete basis set of . Denote the set of all the input states as . We can expand each output state uniquely in as
(9) |
We also define such that
(10) |
From the linear independence of , the relationship between and is [2, 13]
(11) |
To guarantee that (11) has a unique solution, the maximal linear independent subset of must have elements. If this is achieved with , we call the input state set informationally complete. If this is achieved with , we call it informationally overcomplete. The work in [2, 13] is restricted to a special complete case, where (after a linear combination of the practical experiment results). Here we extend the framework of [2, 13] to the general informationally complete/overcomplete case with . We define the matrix where and arrange the elements into a matrix :
(12) |
which is an matrix. We rewrite (11) into a compact form as
(13) |
where is determined once the bases and are chosen.
In an experiment, assume that the number of copies for each input state is . Assume is the -th POVM element (i.e., measurement operator) in the -th POVM set where and . Therefore, the total number of different measurement operators is and we have
(14) |
for , which is the completeness constraint for each POVM set. We then vectorize these POVM elements as where, e.g., . These are the measurement operators on the output states. To extract all of the information in the output state, these measurement operators should be informationally complete or overcomplete, and thus . Some efficient methods for QST with fewer measurements are also discussed in [43, 44, 45]. Then, we can also uniquely expand in as
(15) |
and the ideal measurement probability of the -th output state and the -th measurement operator is
(16) |
Define the matrix with and let
(17) |
which is determined by the measurement operators . Thus we have
(18) |
Define such that and thus is a commutation matrix. Using (13) and (18), we have
(19) |
which directly connects the unknown quantum process with measurement data in the experiment.
Assume that the practical measurement result is and the measurement error is . According to the central limit theorem, converges in distribution to a normal distribution with mean zero and variance [46, 47]. We consider the general scenario where there is no prior knowledge about whether the process is TP or non-TP, and the QPT can be formulated as the following optimization problem:
Problem 1
Given the parameterization matrix for the input states, the parameterization matrix for the measurement operators, the permutation matrix and the experimental data , find a Hermitian and positive semidefinite estimate minimizing and satisfying the constraint (8).
III Tensor structure and two-stage quantum process tomography solution
In this section, we discuss how to simplify the structure of under a suitably chosen representation, which can reduce the computational complexity and storage requirement. Then we propose a two-stage solution (TSS) to obtain analytical tomography formulas for Problem 1. When there is prior knowledge that the process is TP, most of the following analysis still applies, unless otherwise specified.
III-A Tensor structure of QPT
Firstly, we consider the basis sets and , which, if properly chosen, can simplify the QPT problem and greatly benefit the time and space costs of our algorithm. We choose these basis sets as the natural basis where . The advantages of this choice can be demonstrated as follows.
Lemma 1 fellows from the Choi-Jamiołkowski isomorphism and the proof for in the TP case can be found in [13, 15], which can be straightforwardly extended to non-TP cases. Define , and represents the success probability of the quantum process [23]. Furthermore, we have the following benefit.
Proposition 1
Define the collection of all the vectorized input states as
(20) |
If and are chosen as the natural basis where , then we have
(21) |
where is a permutation matrix.
Proof:
(22) |
Since the is natural basis, using (20), we have
(23) |
and is the -th column of . Let , . Then
(24) |
where only the element at is non-zero. Then
(25) |
and therefore,
(26) |
where only the -th column is non-zero. Thus the -th column of is
(27) |
The matrix is block diagonal. Thus the -th column of is the same as the -th column of . Therefore,
(28) |
where is a permutation matrix. ∎
Since we choose as the natural basis, we have
(29) |
Using Proposition 1 and (19), we have
(30) |
Define . Since we assume the input states and measurement operators are both informationally complete or overcomplete, the unique least squares solution to is
(31) |
However, the computational complexity for this least squares solution is usually [17] which is quite high. To reduce the computational complexity, we do not actually utilize (31). Noting the special structure property of (30), we firstly reconstruct and then obtain , which is more efficient in computation. To reconstruct from (18), we have
(32) |
and then the least squares solution is given by
(33) | ||||
Therefore, we can directly obtain the least squares solution as
(34) |
Remark 1
The procedure to obtain in (32) is usually known as QST for output states [2, 13] but our method is different from such a procedure. Here we do not consider any constraints, i.e., Hermitian, positive semidefinite and unit trace, on the estimate and we directly reconstruct the process from the data vector by linear inversion. Of course, it is possible to obtain an alternative physical estimate from as in [48] or via Maximum Likelihood Estimation (MLE). However, here is unnecessary to be a physical estimate because it is only an intermediate product, and our direct method can already guarantee the ultimate process estimation to be physical by later solving Problem 2 and Problem 2 in Sec. III-B.
Considering the positive semidefinite requirement and the constraint , we convert QPT into an optimization problem as follows.
Problem 2
Given the parameterization matrix of all the input states, the permutation matrix and reconstructed , find a Hermitian and positive semidefinite estimate minimizing
such that .
III-B Two-stage solution of QPT
The direct solution to Problem 2 is usually difficult and we split it into two optimization sub-problems (i.e., a two-stage solution):
Problem 2
Let
be a given matrix. Find a Hermitian and positive semidefinite matrix minimizing .
Problem 2
Let be given. Find a Hermitian and positive semidefinite matrix minimizing , such that .
For Problem 2, note that if , we have . Therefore,
(35) |
We perform the spectral decomposition as where is a diagonal matrix. We define . Since is positive semidefinite, we have . Therefore,
(36) |
Therefore, the unique optimal solution is where
(37) |
and .
After solving Problem 2, we rewrite the spectral decomposition of as
(38) |
where each . Define . Using (4), we have
(39) | ||||
Then, we consider the partial trace constraint by correcting . Assume that the spectral decomposition of is
(40) |
where and the spectral decomposition of is
(41) |
where and , i.e., . Then, we define
(42) |
where for , for , and is the number of copies for each input states. Since is invertible, is well defined and we also define
(43) |
where for . Thus, for . To solve Problem 2 and satisfy the partial trace constraint, let . We then obtain as
(44) |
where might not be an orthogonal basis. From (44), is Hermitian, positive semidefinite and also satisfies
(45) |
The Stage 1 solution is a projection under the Frobenius norm satisfying the CP constraint, which was also discussed in [17, 18]. The TP constraint makes QPT a nontrivial extension of state tomography [17]. The work in [17, 18] introduced two iterative algorithms designed to ensure that the estimate satisfies the CPTP constraint. While these algorithms may offer good accuracy, our two-stage algorithm can achieve a closed-form solution. We can also efficiently optimize the input states and the measurement operators using the TSS algorithm. It is worth highlighting that our algorithm is versatile and can be applied to non-TP cases. Additionally, considering Hamiltonian identification in [13] where and , our Stage 2 solution is optimal for Problem 2 which has been proved in [13]. For general QPT, the estimated process matrix is a sub-optimal solution.
Remark 2
If we have the prior knowledge that the quantum process is TP, we can assume that is non-singular (i.e., ) and it is true when the number of copies is sufficiently large, because converges to as tends to infinity. Thus, we correct as and where .
IV Computational complexity and storage requirements

We summarize our procedure for the QPT framework with the TSS algorithm in Fig. 1. In this paper, we do not consider the time spent on experiments. Here we discuss the computational complexity and storage requirement in each step in the red box of Fig. 1 and give the total time and space complexity in Table I.
Step 1. We reconstruct using (32). The computational complexity is for , for and for where is the number of different input states and is the number of different measurement operators. Thus, the total computational complexity is . For storage requirement, the storage is for the parameterization matrix of measurement operators, for measurement data and for . Since is a permutation matrix, it can be stored with an expense .
Step 2. Using (33), the computational complexity is for , for
and for . Thus, the total computational complexity is . We need to store and . Since is a matrix and is a permutation matrix, the storage requirements are and , respectively.
Step 3. To solve Problem 2, the computational complexity is determined by the spectral decomposition of , which is . The storage requirements are for , for the spectral decomposition of , and for .
Step 4. To solve Problem 2, the computational complexity for calculating , and are , and , respectively. Then the computational complexity for calculating is . The storage requirements for and are both . Then the storage for , and are all , and for is .
Computational complexity | Storage requirement | |
---|---|---|
Step 1 | ||
Step 2 | ||
Step 3 | ||
Step 4 | ||
Total |
The total computational complexity and storage requirements for our TSS algorithm are presented in Table I. Since and , the total computational complexity is . From [2, 42, 17], the computational complexity for a direct least squares solution with different input states is , which is significantly higher than . The reason we obtain a lower computational complexity here is that we utilize the special structure of described in Proposition 1. As for the total storage requirement, our TSS algorithm is . Without employing the tensor structure as (33), the storage space will be because the storage requirement of will be .
V Error analysis
From Fig. 1, our TSS algorithm has four steps. In this section, we analyze the error upper bound and give the following theorem to characterize the error upper bound analytically.
Theorem 1
If and are chosen as the natural basis and the input quantum states are , then the estimation error of the TSS algorithm scales as
where is the number of copies for each output state, is the number of POVM sets, is defined as (29), is defined as (20), and denotes expectation with respect to all possible measurement results.
Proof:
V-A Error in Step 1
V-B Error in Step 2
V-C Error in Step 3
Since minimizes , we have and thus
(51) |
because the true value .
V-D Error in Step 4
For , if , i.e., , we have and thus . Therefore, using (48), (50) and (51), we obtain the final error bound as
(52) |
Then we consider the case where at least one eigenvalue of is larger than . Assume for and for in (41). From Lemma 2 and (39), we have
(53) | ||||
Using (40), (41), and (131) in Appendix A, we can obtain
(54) |
and therefore we have
(55) |
Using (42) and (132) in Appendix A, we have
(56) |
Since , we have , and thus (56) indicates
(57) |
Then we consider as
(58) |
For , . For , , and we have
(59) |
Thus, using (59) and (132) in Appendix A, we have
(60) | ||||
Then using (45) and (57), we have
(61) | ||||
Using (44), we thus have
(62) | ||||
Similarly, using (38) and (39), we have
(63) | ||||
Thus, from (3) and the definition of , we know , and we have
(64) |
where we use Proposition 3 in the inequality. Therefore, using (62)–(64), we have
(65) |
where we use the Cauchy-Schwarz inequality in the third inequality and AM-QM inequality [49] in the fourth inequality. Thus, using (48), (50), (51) and (65), the final error bound is
(66) | ||||
Comparing (52) and (66), we obtain the final error bound for any quantum process in the approximation sense as
(67) |
where represents the success probability of the quantum process [23]. ∎
Note that when the quantum process is TP, and the error analysis is similar to non-TP case. Then the final error bound is
(68) |
VI Optimization of input states and measurement operators
To minimize the tomography error, the optimal input states and optimal measurement operators are usually dependent on the specific process. One advantage of our TSS algorithm is that we can give an analytical error upper bound as (67) or (68) which depends on the input states and measurement operators instead of the unknown process. Therefore, we can optimize the input states and measurement operators based on the upper bound. In addition, we also consider how robust the estimated process is w.r.t. the measurement error, which is determined by (32) and (33). Therefore, we use the condition numbers of and to describe the robustness, which reflects the sensitivity of the solution to the perturbations in data. Moreover, we consider the relationship between and the final estimation error with randomly generated input states.
VI-A Optimal input quantum states
Define the set with different types of input states as
(69) |
Similarly to the optimal probe states in quantum detector tomography [32, 50], we consider the optimal input states for QPT. The part of (67) or (68) dependent on the input states is , which will be our first cost function. Therefore, we define the set of all the optimal input states to minimize this bound as
(70) |
For general QPT, we have no prior knowledge of the unknown process. From (33) with the reconstructed , the following conditions are equivalent: (i) the process cannot be uniquely identified; (ii) the input states do not span the space of all -dimensional Hermitian matrices; (iii) is singular; (iv) is infinite. From (i) and (iv), it is thus reasonable to take as a cost function.
In addition, maximizing the robustness of the reconstructed process with respect to measurement noise amounts to minimizing the condition number. Therefore, from (33), the second cost function is the condition number of ; i.e., , defined as
and is the maximum/minimum singular value of . Using (21), we have which also only depends on the input states. Similarly, we also define the set of all the optimal input states to minimize the condition number as
(71) |
We then define two sets and to characterize two lower bounds of these two cost functions as
(72) |
(73) |
and the proof is as given in Theorem 2. In addition, we define the set of the optimal input states () and the set of the optimal input states achieving both of the lower bounds () as
(74) |
and
(75) |
Thus when is non-empty, it should minimize and minimize simultaneously. Additionally, when is non-empty, it can achieve the lower bounds of and simultaneously. Then, we present the following theorem for .
Theorem 2
For a -dimensional quantum process, with different types of input states and the parameterization matrix as in (20), we have and . These lower bounds are achieved simultaneously; i.e., if and only if the eigenvalues of are and .
Proof:
Denote the eigenvalues of as . Since , we can obtain two constraints and coming from the purity requirement and unit trace of quantum states, respectively, which have been proved in [32]. To minimize and , we relax the optimization problems as follows:
min | (76) | |||
s.t. |
and
min | (77) | |||
s.t. |
Using Lagrange multiplier method, we have
(78) |
and
(79) |
These lower bounds hold if and only if and . ∎
Since is a closed space, and are always non-empty for any . However, it is not clear whether this also holds for . For the lower bounds of these two cost functions, if one of them is achieved, we have and thus
(80) |
If the two lower bounds cannot be achieved, the above three sets are empty. Therefore, (80) always holds for any . However, we still have not fully determined what values of make (80) non-empty. In addition, when is not empty, we have . If , all of them must be pure states from the purity requirement. This indicates that pure states may be better than mixed states in QPT for reducing the estimation error. We provide the following two examples in Appendix B which belong to and achieve the lower bounds: SIC (symmetric informationally complete) states with the smallest and MUB (mutually unbiased bases) states for . Moreover, we consider that the system is restricted to a multi-qubit system where and each input state is an -qubit tensor product state as where and there are different types of single qubits for the -th qubit of the product states. Thus, the total number of these -qubit tensor product states is thus . We have the following theorem.
Theorem 3
For -qubit product input states, we have and . These lower bounds are achieved simultaneously if and only if for each , .
Proof:
Assuming that the parameterization matrix in (20) of all the -th single-qubit states is . We thus have . Therefore,
(81) |
and
(82) |
The above two equalities hold if and only if for each , . ∎
A similar proof for quantum detector tomography can also be found in [32]. For one-qubit input states, they are in the Bloch sphere and [32] has proved that platonic solids–tetrahedron, cube, octahedron, dodecahedron and icosahedron construct five examples, which belong to with and achieve one-qubit lower bounds. and are one-qubit SIC states and one-qubit MUB states, respectively. For two-qubit product input states, Cube states (product states of two one-qubit MUB states) can achieve the lower bounds and .
VI-B Optimal measurement operators
For our error bound (67) or (68), the part depending on the measurement operators is and we need to minimize it. Since , we also consider the minimum of .
Assume is the -th POVM element (i.e., a measurement operator) in the -th POVM set where and . Define the set size . Therefore, the total number of measurement operators . Let be a complete set of -dimensional traceless Hermitian matrices except , and they satisfy where is the Kronecker function. Then we can parameterize the POVM element as
(83) |
where is real. Thus we use as the parameterization of and the completeness constraint (14) becomes
(84) |
and thus
(85) |
Conversely, using (85), we can also obtain (84). Define
(86) |
which is the parameterization matrix of all the measurement operators in the basis of and is a real matrix. Since and natural basis are both orthonormal basis, the relationship between and is where is a unitary matrix. Therefore, the eigenvalues of and are the same, and , .
Similarly, we define the set of all -dimensional measurement operators with set number and set size as
(87) |
For general QPT, we have no prior knowledge of the output states. From (32), the following conditions are equivalent: (i) the output states cannot be uniquely identified; (ii) the measurement operators do not span the space of all -dimensional Hermitian matrices; (iii) is singular; (iv) is infinite. From (i) and (iv), it is also reasonable to choose as the first cost function. Thus we define the set of all the optimal measurement operators to minimize () as
(88) |
Since maximizing the robustness of the reconstructed output states with respect to measurement noise is equivalent to minimizing the condition number, from (32), we also choose as the second cost function. Thus, we define the set of all the optimal measurement operators to minimize () as
(89) |
Then we define and two sets and to characterize the lower bounds of these two cost functions as
(90) |
(91) |
where the proof is given as in Theorem 4. We also define the set of all the optimal measurement operators () as
(92) |
and the set of all the optimal measurement operators achieving both of the lower bounds () as
(93) |
Thus, , when it is non-empty, should minimize and minimize simultaneously. When is non-empty, it can achieve the lower bounds of and simultaneously. Then we present the following theorem for .
Theorem 4
For a -dimensional quantum process, let the total number of POVM sets be and the number of POVM elements in the -th POVM set be . Then
and
These lower bounds are achieved simultaneously; i.e., , if and only if the eigenvalues of are and .
Proof:
Denote the eigenvalues of or as . Since all the measurement operators are positive semidefinite, we have
(94) |
Using (84), we have
(95) | ||||
Therefore,
(96) |
Since
(97) |
the first diagonal element of is
(98) |
and thus
(99) |
To minimize and , the optimization problems can be relaxed as follows
min | (100) | |||
s.t. |
and
min | (101) | |||
s.t. |
Using the Lagrange multiplier method, we have
(102) |
and
(103) |
These lower bounds are achieved simultaneously if and only if and . ∎
Since is a closed space, there always exist non-empty and for arbitrary set number and set size . However, it is not clear whether this also holds for . For the lower bounds of these two cost functions, if one of them is achieved, we have and thus
(104) |
If the two lower bounds can not be achieved, the above three sets are empty. Therefore, (104) always holds for any and . Our results show that if , then in each POVM set, these POVM elements are orthogonal; i.e., from the first constraint. To achieve these lower bounds, we must have because the maximum number of orthogonal POVM elements in one set is . It is an open problem as to whether is non-empty for arbitrary set number and set sizes (). When it is non-empty, we have .
Here, we give one example of : MUB (mutually unbiased bases) measurement with and for , which can achieve these lower bounds. Two sets of orthogonal bases and are called mutually unbiased if and only if [51]
(105) |
For a prime and a positive integer , there exist maximally sets of mutually unbiased bases in Hilbert spaces of prime-power dimension [52]. For other values, it is still an open problem as to whether there exist sets of mutually unbiased bases. The corresponding MUB measurements are . Then we give the following proposition with a similar proof to that of Proposition 2 in [32].
Proposition 2
-dimensional MUB measurements (when they exist) belong to with and for .
Proof:
Remark 3
SIC-POVM are usually thought as optimal (in certain senses) measurements in quantum physics. The simplest mathematical definition of SIC-POVM is a set of normalized vectors in satisfying [53]
(108) |
The corresponding POVM elements are for and with , . However, here SIC-POVM does not belong to because these POVM elements are not orthogonal with each other. Therefore, the first constraint is not satisfied. We can calculate and for SIC-POVM. These values are the same as MUB measurement which has and for . Thus we conjecture that SIC-POVM belongs to for , .
Remark 4
Our analysis for optimal measurement operators can also be applied to QST. A similar problem for QST has been discussed in [46]. However, their work is restricted to optimal projective measurement, while our result considers general POVM. Miranowicz et al. [54] also considered the minimum condition number for the optimal measurement operators in QST and proposed generalized Pauli operators whose condition number is . However, generalized Pauli operators are realized from projective measurements and their condition number does not directly link to the experimental data.
With input states like SIC states or MUB states belonging to and measurement operators like MUB measurements belonging to , the error upper bound of (67) becomes . The limitation of our error upper bound is that it might be loose with respect to the system dimension and we leave it an open problem to characterize more accurately. We can obtain a tighter upper bound if we use the natural basis states and with MUB measurements. Note that in experiments, all of the quantum states are Hermitian. Hence, we cannot actually use () as input quantum states. According to [2], since we have
(109) |
where , can be obtained as
(110) |
Using natural basis states which we define as all the states needed by the RHS of (109), Wang et al. [13] have proved that is a permutation matrix. Thus, the error analysis in Step 2 becomes
(111) |
and the error analysis in other steps are the same as our analysis. Therefore, we obtain a tighter upper bound as .
VI-C On the different types of input quantum states and measurement operators
In practice, the different types of input quantum states and measurement operators may vary and here we discuss the MSE scaling with respect to and . We first allow to change, with the dimension , the measurement operators and the copy number of each input state given. The different input states are randomly generated as [55, 56] or as [57] for truncated coherent states. Coherent states are more straightforward to be prepared than Fock states in quantum optics and have been used in QPT in [58, 59]. Each scenario with copies generates the input states according to a certain probability distribution and the total copies are , based on which denotes the expectation, different from in Theorem 1. Define , which is i.i.d. with respect to the subscript . Thus we have
(112) |
According to the central limit theorem, as tends to infinity, converges to a fixed matrix [57]. Therefore the expectation of is
(113) |
In Step 1, since is given, using (46), we have
(114) |
In Step 2, assume that the SVD of is
(115) |
where is a diagonal matrix, and and are unitary. Using (49) and (113), . Let
(116) |
where is a vector, is a vector and each element in scales as from (114). Then for the error in Step 2, we have
(117) |
Since Steps 3–4 are not related to the scaling on , the final error scaling on is
(118) |
From our error upper bound in (67), using (113), we have
(119) |
and thus the scaling of (67) is with given and . But the accurate scaling on is . Therefore, our error upper bound (67) is not always tight with respect to .
Then we consider the MSE scaling of our TSS on . With given input states, and , we can also generate different measurement operators according to certain probability distributions. Similarly, we can prove
(120) |
and
(121) |
Therefore, our error upper bound (67) is tight with respect to .
VII Numerical examples
To perform measurement on the output states, there are different measurement bases such as SIC-POVM [53], MUB measurement [52], and Cube bases [60]. In this section, we use Cube bases, as it is relatively easy to be implemented in experiments. For one-qubit systems, the Cube bases are where , , are Pauli matrices. For multi-qubit systems, the Cube bases are the tensor products of one-qubit Cube bases.
VII-A Performance illustration
We consider a -dimensional quantum system and the TP process matrix is determined by where
(122) | ||||
We use four different sets of probe states. The first set is SIC state which is given in (133) in Appendix B and . The second set is MUB states which is given in (135) in Appendix B and . The third one is random states where we randomly generate input states using the algorithm in [55, 56]. The fourth set is natural basis states in the RHS of (109). The result of MSE (mean squared error) versus the total number of copies is shown in Fig. 2(a), where and we make all of the four sets have the same . For each number of copies , we repeat our algorithm times and obtain the average MSE and error bars. For all the four sets of input states, the scaling of MSE is basically which satisfies Theorem 1. Moreover, the MSEs of SIC states, MUB states, natural basis states are all smaller than random states. The MSEs of SIC states and MUB states are quite similar and both smaller than natural basis states. For non-TP processes, we consider the same and different as
(123) |
where is a random unitary matrix. The result for non-TP processes is shown in Fig. 2(b) which is similar to Fig. 2(a). The MSE of non-TP processes is a little smaller than that of TP processes because the constraint for TP processes is stricter than that for non-TP processes.


We also test our algorithm on the IBM Quantum device [61]. Here we perform QPT on the CNOT gate defined as
(124) |
in the IBM Quantum Composer. For the input states, we use states which can be generated by applying Hadamard and gates to qubits initialized in the state [62]. In addition, we use Cube measurements. Fig. 3 is an example of IBM Quantum Composer where the input state is and the measurement operators are .

If we have prior knowledge that the process is unitary, some references have proposed effective identification algorithms [63, 64]. Here we assume that we do not have prior knowledge that the unknown process is in essence unitary. We perform these numerical examples using the MATLAB on a classical computer and the . The total numbers of copies are and the experiments are repeated 5 times. Then we apply our TSS algorithm and the results are shown in Fig. 4. Besides MSE, we also consider another common fidelity metric, defined as [23]
(125) |
and the correspond infidelity is defined as . From Fig. 4, the scalings of the MSEs from and simulation are also both which satisfies Theorem 1. The scalings of the infidelities are both because the CNOT process is rank-deficient. A similar infidelity scaling has also been studied in QST [51, 65]. In addition, the errors between simulation and are close. Then we apply our algorithm on the 5-qubit system. The MSE and fidelity are presented in Fig. 5 where the total number of copies are , respectively, and we repeat times. These results are both worse than and simulation because there exist state preparation error and readout error in [61]. The average CNOT error is and average readout error is for . Thus, the approximate fidelity is .


VII-B Performance comparison
Here, we compare the performance between our TSS algorithm and the convex optimization method in [14]. Baldwin et al. [14] formulated the TP QPT problem as a convex optimization problem
(126) | ||||
s.t. | ||||
where is measurement data, is a constant matrix which is the tensor product between the transpose of the -th input states and -th measurement bases where , , and is the estimated process matrix. The detailed description can be found in equation (13) in [14]. This problem can be solved as a convex semidefinite program, which is realized by CVX [66, 67].

We randomly generate input quantum states as [55, 56] and . With given number of copies for each output state and randomly generated process as in (122), a comparison of results is shown in Fig. 6 where the horizontal axis is the number of qubits , the left vertical axis is the MSE on a logarithmic scale and the right vertical axis is the running time (in seconds) on a logarithmic scale. We repeat the simulation times and obtain the average MSE, the average running time and error bars.
When the two algorithms have a similar MSE (we set the error of convex optimization a little larger than TSS by changing CVX precision), the running time of TSS algorithm is smaller than convex optimization by approximately three orders of magnitude for three-qubit systems, indicating that our TSS algorithm is more efficient than the convex optimization method. Since and we choose Cube bases where , the computational complexity for our algorithm is . For the simulated running time of our algorithm, since our computational complexity analysis is asymptotical on , we fit the rightmost three points and the slope is , which is close to the theoretical value . The slight difference might be attributed to fact that the qubit number is not large enough. The storage requirement is also labeled by the black numbers in Fig. 6. The storage requirement for our algorithm is . Since we use Cube bases of measurement, and thus . As increases from to , are , , , and , respectively. For the convex optimization method, the main storage requirement is the cost of storing , which is a matrix and the total number of these matrices is . Thus, the storage requirement is . As increases from to , are , and , respectively. Thus, our algorithm needs less storage than the convex optimization method.

We then compare the two algorithms by changing the types of input states with , where we randomly generate input states as in [55, 56] and the number of copies for each output state is . The comparison results are shown in Fig. 7 where the horizontal axis is the different types of input states , the left vertical axis is the MSE and the right vertical axis is the running time , all on a logarithmic scale. From Fig. 7, the scaling between MSE and is , which is close to and consistent with (118). This scaling is different from the scaling on in Fig. 2 because here is given and increases, while in Fig. 2, is given and increases. For the simulated running time of our algorithm, we fit the rightmost three points to obtain the slope , which is close to the theoretical value . For the convex optimization method, the slope of the fitted line of the rightmost three points is , which is larger than that of our TSS algorithm, indicating that our TSS algorithm is more efficient than the convex optimization method.
VIII Conclusion
In this paper, we have introduced an analytical two-stage solution (TSS) applicable to both trace-preserving and non-trace-preserving QPT. Leveraging the natural basis, we have utilized the tensor structure of the coefficient matrix and provided insights into the computational complexity and storage requirements for our algorithm. Our contributions include the establishment of an analytical upper bound for errors and the optimization of input quantum states and measurement operators based on this error upper bound and condition number. The effectiveness and theoretical soundness of our algorithm have been demonstrated through numerical examples and testing on an IBM Quantum device. Furthermore, we have benchmarked our TSS algorithm against the convex optimization method outlined in [14], and the results show that our algorithm is more efficient in terms of both time and space costs. Further work will focus on extending our algorithm to AAPT.
Acknowledgments
The authors would like to thank Prof. Lloyd C. L. Hollenberg for the helpful discussion. This research was supported by the University of Melbourne through the establishment of the IBM Quantum Network Hub at the University. Shuixin Xiao would like to thank the support from the IEEE Control Systems Society Graduate Collaboration Fellowship.
Appendix A Several propositions and lemmas
Here, we give some propositions and lemmas which are utilized in proving Theorem 1.
Proposition 3
For any and , we have
(127) |
Proof:
Let
(128) |
where is a vector. Therefore,
(129) |
∎
Lemma 2
[68] Let and be finite-dimensional Hilbert spaces of dimensions and , respectively, and let . Then for any unitarily invariant norm that is multiplicative over tensor products, the partial trace satisfies the norm inequality
(130) |
where is the identity operator.
Lemma 3
([69] Theorem 8.1 and Theorem 28.3) Let , be Hermitian matrices with eigenvalues and , respectively. Then
(131) |
and
(132) |
Appendix B SIC states and MUB states
For input states achieving the lower bounds of and , we give two examples, SIC states and MUB states, as the following propositions.
Proposition 4
-dimensional SIC states (when they exist) belong to with the smallest as .
Proposition 5
-dimensional MUB states (when they exist) belong to for .
The proofs are similar to the optimal probe states in quantum detector tomography and can be found in [32].
For SIC-POVM in , Bengtsson [70] gave one expression ignoring overall phases and normalization in the natural basis as
(133) |
where . Let be the -th column of (133). In this paper, we call the set of
as SIC states and as SIC-POVM.
For , five MUB measurement sets are
(134) |
and
(135) |
where , , and in the natural basis. Adamson et al. [51] utilized these Mutually Unbiased Bases in the QST. In this paper, we call a MUB state and a MUB measurement operator.
References
- [1] D. P. DiVincenzo, “Quantum computation,” Science, vol. 270, no. 5234, pp. 255–261, 1995.
- [2] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information. Cambridge University Press, 2010.
- [3] N. Gisin and R. Thew, “Quantum communication,” Nature Photonics, vol. 1, no. 3, p. 165, 2007.
- [4] M.-H. Hsieh and M. M. Wilde, “Trading classical communication, quantum communication, and entanglement in quantum Shannon theory,” IEEE Transactions on Information Theory, vol. 56, no. 9, pp. 4705–4730, 2010.
- [5] C. L. Degen, F. Reinhard, and P. Cappellaro, “Quantum sensing,” Reviews of Modern Physics, vol. 89, p. 035002, 2017.
- [6] D. Dong and I. R. Petersen, “Quantum control theory and applications: a survey,” IET Control Theory & Applications, vol. 4, no. 12, pp. 2651–2671, 2010.
- [7] D. Dong and I. R. Petersen, “Quantum estimation, control and learning: Opportunities and challenges,” Annual Reviews in Control, vol. 54, pp. 243–251, 2022.
- [8] D. Dong and I. R. Petersen, Learning and Robust Control in Quantum Technology. Springer Nature Switzerland AG, 2023.
- [9] D. Burgarth and K. Yuasa, “Quantum system identification,” Physical Review Letters, vol. 108, p. 080502, 2012.
- [10] A. Sone and P. Cappellaro, “Hamiltonian identifiability assisted by a single-probe measurement,” Physical Review A, vol. 95, p. 022335, 2017.
- [11] Y. Wang, D. Dong, A. Sone, I. R. Petersen, H. Yonezawa, and P. Cappellaro, “Quantum Hamiltonian identifiability via a similarity transformation approach and beyond,” IEEE Transactions on Automatic Control, vol. 65, no. 11, pp. 4632–4647, 2020.
- [12] J. Zhang and M. Sarovar, “Quantum Hamiltonian identification from measurement time traces,” Physical Review Letters, vol. 113, p. 080401, 2014.
- [13] Y. Wang, D. Dong, B. Qi, J. Zhang, I. R. Petersen, and H. Yonezawa, “A quantum Hamiltonian identification algorithm: Computational complexity and error analysis,” IEEE Transactions on Automatic Control, vol. 63, no. 5, pp. 1388–1403, 2018.
- [14] C. H. Baldwin, A. Kalev, and I. H. Deutsch, “Quantum process tomography of unitary and near-unitary maps,” Physical Review A, vol. 90, p. 012110, 2014.
- [15] M. Zorzi, F. Ticozzi, and A. Ferrante, “Minimal resources identifiability and estimation of quantum channels,” Quantum Information Processing, vol. 13, no. 3, pp. 683–707, 2014.
- [16] M. Zorzi, F. Ticozzi, and A. Ferrante, “Estimation of quantum channels: Identifiability and ML methods,” in 2012 51st IEEE Conference on Decision and Control (CDC), pp. 1674–1679, 2012.
- [17] G. C. Knee, E. Bolduc, J. Leach, and E. M. Gauger, “Quantum process tomography via completely positive and trace-preserving projection,” Physical Review A, vol. 98, p. 062336, 2018.
- [18] T. Surawy-Stepney, J. Kahn, R. Kueng, and M. Guţă, “Projected least-squares quantum process tomography,” Quantum, vol. 6, p. 844, 2022.
- [19] Z. Ji, G. Wang, R. Duan, Y. Feng, and M. Ying, “Parameter estimation of quantum channels,” IEEE Transactions on Information Theory, vol. 54, no. 11, pp. 5172–5185, 2008.
- [20] J. Zhang and M. Sarovar, “Identification of open quantum systems from observable time traces,” Physical Review A, vol. 91, no. 5, p. 052121, 2015.
- [21] M. Guţă and N. Yamamoto, “System identification for passive linear quantum systems,” IEEE Transactions on Automatic Control, vol. 61, no. 4, pp. 921–936, 2016.
- [22] X.-L. Huang, J. Gao, Z.-Q. Jiao, Z.-Q. Yan, Z.-Y. Zhang, D.-Y. Chen, X. Zhang, L. Ji, and X.-M. Jin, “Reconstruction of quantum channel via convex optimization,” Science Bulletin, vol. 65, no. 4, pp. 286–292, 2020.
- [23] I. Bongioanni, L. Sansoni, F. Sciarrino, G. Vallone, and P. Mataloni, “Experimental quantum process tomography of non-trace-preserving maps,” Physical Review A, vol. 82, p. 042307, 2010.
- [24] G. A. White, C. D. Hill, F. A. Pollock, L. C. Hollenberg, and K. Modi, “Demonstration of non-Markovian process characterisation and control on a quantum processor,” Nature Communications, vol. 11, p. 6301, 2020.
- [25] G. A. White, F. A. Pollock, L. C. Hollenberg, K. Modi, and C. D. Hill, “Non-markovian quantum process tomography,” PRX Quantum, vol. 3, p. 020344, 2022.
- [26] H. Wang, W. Zheng, N. Yu, K. Li, D. Lu, T. Xin, C. Li, Z. Ji, D. Kribs, B. Zeng, X. Peng, and J. Du, “Quantum state and process tomography via adaptive measurements,” Science China Physics, Mechanics & Astronomy, vol. 59, no. 10, p. 100313, 2016.
- [27] Y. S. Teo, B.-G. Englert, J. Řeháček, and Z. Hradil, “Adaptive schemes for incomplete quantum process tomography,” Physical Review A, vol. 84, p. 062125, 2011.
- [28] I. A. Pogorelov, G. I. Struchalin, S. S. Straupe, I. V. Radchenko, K. S. Kravtsov, and S. P. Kulik, “Experimental adaptive process tomography,” Physical Review A, vol. 95, p. 012302, 2017.
- [29] I. L. Chuang and M. A. Nielsen, “Prescription for experimental determination of the dynamics of a quantum black box,” Journal of Modern Optics, vol. 44, no. 11-12, pp. 2455–2467, 1997.
- [30] J. F. Poyatos, J. I. Cirac, and P. Zoller, “Complete characterization of a quantum process: The two-bit quantum gate,” Physical Review Letters, vol. 78, pp. 390–393, 1997.
- [31] J. Fiurášek and Z. Hradil, “Maximum-likelihood estimation of quantum processes,” Physical Review A, vol. 63, p. 020101, 2001.
- [32] S. Xiao, Y. Wang, D. Dong, and J. Zhang, “Optimal and two-step adaptive quantum detector tomography,” Automatica, vol. 141, p. 110296, 2022.
- [33] R. A. Horn and C. R. Johnson, Matrix Analysis. Cambridge University Press, 2012.
- [34] J. Watrous, The Theory of Quantum Information. Cambridge University Press, 2018.
- [35] D. W. Leung, “Choi’s proof as a recipe for quantum process tomography,” Journal of Mathematical Physics, vol. 44, no. 2, pp. 528–533, 2003.
- [36] G. M. D’Ariano and P. Lo Presti, “Quantum tomography for measuring experimentally the matrix elements of an arbitrary quantum operation,” Physical Review Letters, vol. 86, pp. 4195–4198, 2001.
- [37] J. B. Altepeter, D. Branning, E. Jeffrey, T. C. Wei, P. G. Kwiat, R. T. Thew, J. L. O’Brien, M. A. Nielsen, and A. G. White, “Ancilla-assisted quantum process tomography,” Physical Review Letters, vol. 90, p. 193601, 2003.
- [38] G. M. D’Ariano and P. Lo Presti, “Imprinting complete information about a quantum channel on its output state,” Physical Review Letters, vol. 91, p. 047902, 2003.
- [39] M. Mohseni and D. A. Lidar, “Direct characterization of quantum dynamics,” Physical Review Letters, vol. 97, p. 170501, 2006.
- [40] M. Mohseni and D. A. Lidar, “Direct characterization of quantum dynamics: General theory,” Physical Review A, vol. 75, p. 062331, 2007.
- [41] Z.-W. Wang, Y.-S. Zhang, Y.-F. Huang, X.-F. Ren, and G.-C. Guo, “Experimental realization of direct characterization of quantum dynamics,” Physical Review A, vol. 75, p. 044304, 2007.
- [42] M. Mohseni, A. T. Rezakhani, and D. A. Lidar, “Quantum-process tomography: Resource analysis of different strategies,” Physical Review A, vol. 77, p. 032322, 2008.
- [43] M. Zorzi, F. Ticozzi, and A. Ferrante, “Minimum relative entropy for quantum estimation: Feasibility and general solution,” IEEE Transactions on Information Theory, vol. 60, no. 1, pp. 357–367, 2014.
- [44] J. Haah, A. W. Harrow, Z. Ji, X. Wu, and N. Yu, “Sample-optimal tomography of quantum states,” IEEE Transactions on Information Theory, vol. 63, no. 9, pp. 5628–5641, 2017.
- [45] M. Berta, J. M. Renes, and M. M. Wilde, “Identifying the information gain of a quantum measurement,” IEEE Transactions on Information Theory, vol. 60, no. 12, pp. 7987–8006, 2014.
- [46] B. Qi, Z. Hou, L. Li, D. Dong, G.-Y. Xiang, and G.-C. Guo, “Quantum state tomography via linear regression estimation,” Scientific Reports, vol. 3, p. 3496, 2013.
- [47] B. Mu, H. Qi, I. R. Petersen, and G. Shi, “Quantum tomography by regularized linear regressions,” Automatica, vol. 114, p. 108837, 2020.
- [48] J. A. Smolin, J. M. Gambetta, and G. Smith, “Efficient method for computing the maximum-likelihood quantum state from measurements with additive Gaussian noise,” Physical Review Letters, vol. 108, p. 070502, 2012.
- [49] H. Sedrakyan and N. Sedrakyan, Algebraic Inequalities. Springer, 2018.
- [50] S. Xiao, Y. Wang, D. Dong, and J. Zhang, “Optimal quantum detector tomography via linear regression estimation,” in 2021 60th IEEE Conference on Decision and Control (CDC), pp. 4140–4145, 2021.
- [51] R. B. A. Adamson and A. M. Steinberg, “Improving quantum state estimation with mutually unbiased bases,” Physical Review Letters, vol. 105, p. 030406, 2010.
- [52] T. Durt, B.-G. Englert, I. Bengtsson, and K. Życzkowski, “On mutually unbiased bases,” International Journal of Quantum Information, vol. 08, no. 04, pp. 535–640, 2010.
- [53] J. M. Renes, R. Blume-Kohout, A. J. Scott, and C. M. Caves, “Symmetric informationally complete quantum measurements,” Journal of Mathematical Physics, vol. 45, no. 6, pp. 2171–2180, 2004.
- [54] A. Miranowicz, K. Bartkiewicz, J. Peřina, M. Koashi, N. Imoto, and F. Nori, “Optimal two-qubit tomography based on local and global measurements: Maximal robustness against errors as described by condition numbers,” Physical Review A, vol. 90, p. 062123, 2014.
- [55] J. A. Miszczak, “Generating and using truly random quantum states in Mathematica,” Computer Physics Communications, vol. 183, no. 1, pp. 118–124, 2012.
- [56] N. Johnston, “QETLAB: A MATLAB toolbox for quantum entanglement, version 0.9,” Jan. 2016.
- [57] Y. Wang, S. Yokoyama, D. Dong, I. R. Petersen, E. H. Huntington, and H. Yonezawa, “Two-stage estimation for quantum detector tomography: Error analysis, numerical and experimental results,” IEEE Transactions on Information Theory, vol. 67, no. 4, pp. 2293–2307, 2021.
- [58] M. Lobino, D. Korystov, C. Kupchak, E. Figueroa, B. C. Sanders, and A. I. Lvovsky, “Complete characterization of quantum-optical processes,” Science, vol. 322, no. 5901, pp. 563–566, 2008.
- [59] S. Rahimi-Keshari, A. Scherer, A. Mann, A. T. Rezakhani, A. I. Lvovsky, and B. C. Sanders, “Quantum process tomography with coherent states,” New Journal of Physics, vol. 13, no. 1, p. 013006, 2011.
- [60] M. D. de Burgh, N. K. Langford, A. C. Doherty, and A. Gilchrist, “Choice of measurement sets in qubit tomography,” Physical Review A, vol. 78, p. 052122, 2008.
- [61] “IBM Quantum.” https://quantum-computing.ibm.com, 2021.
- [62] A. Dang, G. A. White, L. C. Hollenberg, and C. D. Hill, “Process tomography on a 7-qubit quantum processor via tensor network contraction path finding,” arXiv preprint arXiv:2112.06364, 2021.
- [63] G. Gutoski and N. Johnston, “Process tomography for unitary quantum channels,” Journal of Mathematical Physics, vol. 55, no. 3, p. 032201, 2014.
- [64] Y. Wang, Q. Yin, D. Dong, B. Qi, I. R. Petersen, Z. Hou, H. Yonezawa, and G.-Y. Xiang, “Quantum gate identification: Error analysis, numerical results and optical experiment,” Automatica, vol. 101, pp. 269–279, 2019.
- [65] B. Qi, Z. Hou, Y. Wang, D. Dong, H.-S. Zhong, L. Li, G.-Y. Xiang, H. M. Wiseman, C.-F. Li, and G.-C. Guo, “Adaptive quantum state tomography via linear regression estimation: Theory and two-qubit experiment,” npj Quantum Information, vol. 3, p. 19, 2017.
- [66] M. Grant and S. Boyd, “CVX: Matlab software for disciplined convex programming, version 2.1.” http://cvxr.com/cvx, Mar. 2014.
- [67] M. Grant and S. Boyd, “Graph implementations for nonsmooth convex programs,” in Recent Advances in Learning and Control (V. Blondel, S. Boyd, and H. Kimura, eds.), Lecture Notes in Control and Information Sciences, pp. 95–110, Springer-Verlag Limited, 2008. http://stanford.edu/~boyd/graph_dcp.html.
- [68] D. A. Lidar, P. Zanardi, and K. Khodjasteh, “Distance bounds on quantum dynamics,” Physical Review A, vol. 78, p. 012308, 2008.
- [69] R. Bhatia, Perturbation Bounds for Matrix Eigenvalues. Society for Industrial and Applied Mathematics, 2007.
- [70] I. Bengtsson, “From SICs and MUBs to Eddington,” Journal of Physics: Conference Series, vol. 254, p. 012007, 2010.