On Estimating the Quantum Distance
Abstract
We study the computational complexity of estimating the quantum distance , defined via the Schatten -norm , given -size state-preparation circuits of -qubit quantum states and . This quantity serves as a lower bound on the trace distance for . For any constant , we develop an efficient rank-independent quantum estimator for with time complexity , achieving an exponential speedup over the prior best results of due to Wang, Guan, Liu, Zhang, and Ying (IEEE Trans. Inf. Theory 2024). Our improvement leverages efficiently computable uniform polynomial approximations of signed positive power functions within quantum singular value transformation, thereby eliminating the dependence on the rank of the states.
Our quantum algorithm reveals a dichotomy in the computational complexity of the Quantum State Distinguishability Problem with Schatten -norm (), which involves deciding whether is at least or at most . This dichotomy arises between the cases of constant and :
-
For any , is -complete.
-
For any , is -complete, implying that no efficient quantum estimator for exists unless .
The hardness results follow from reductions based on new rank-dependent inequalities for the quantum distance with , which are of independent interest.
Keywords and phrases:
quantum algorithms, quantum state testing, trace distance, Schatten normFunding:
Yupan Liu: Supported in part by the Ministry of Education, Culture, Sports, Science and Technology (MEXT) Quantum Leap Flagship Program (Q-LEAP) under Grant JPMXS0120319794, in part by Japan Science and Technology Agency (JST) Support for Pioneering Research Initiated by the Next Generation (SPRING) under Grant JPMJSP2125 and “THERS Make New Standards Program for the Next Generation Researchers”, and in part by the Japan Society for the Promotion of Science (JSPS) Grants-in-Aid for Scientific Research (KAKENHI) under Grant 24H00071.Copyright and License:
2012 ACM Subject Classification:
Theory of computation Design and analysis of algorithms ; Theory of computation Quantum query complexity ; Theory of computation Quantum information theory ; Theory of computation Quantum complexity theoryEditors:
Anne Benoit, Haim Kaplan, Sebastian Wild, and Grzegorz HermanSeries and Publisher:
Leibniz International Proceedings in Informatics, Schloss Dagstuhl – Leibniz-Zentrum für Informatik
1 Introduction
Closeness testing of quantum states is a central topic in quantum property testing [58], which aims to develop (efficient) quantum testers for properties of quantum objects. This problem is also closely related to verifying the functionality of quantum devices, such as and , which are commonly designed to prepare the respective -qubit (mixed) quantum states and . The goal of (tolerant) quantum state testing is to design efficient quantum algorithms that test whether is -far from or -close to with respect to a given closeness measure. Notably, this problem generalizes classical (tolerant) distribution testing (see [20] and [35, Chapter 11]) from a non-commutative perspective.
When the “source codes” of distribution- or state-preparation circuits are given, a surprising correspondence was established between such closeness testing problems – measured by the norm distance [66, 37] or entropy difference [36] – and interactive proof systems that admit statistical zero-knowledge (SZK). This correspondence links closeness testing problems to both complexity theory and cryptography. A similar correspondence was later identified in the quantum world: closeness testing of quantum states with respect to the trace distance (given by Schatten -norm) [77, 78], denoted by QSD, or the von Neumann entropy difference [10] was shown to be QSZK-complete.111The QSZK containment of the closeness testing problem with respect to the trace distance, denoted by , holds only in the polarizing regime , as shown in [77, 78]. A recent work [51] slightly improved the parameter regime for this containment.
In contrast, when the closeness measure follows an -norm-like definition, such as the Hilbert-Schmidt distance or the quantum linear entropy, the corresponding closeness testing problems are in BQP using the SWAP test [17, 27]. Taken together, these results reveal a dichotomy in the complexity of closeness testing: when the measure is -norm-like, the problems are QSZK-hard and their query or sample complexities have polynomial dependence on the dimension or rank of the states; whereas for -norm-like measures, the problems are contained in BQP and their query or sample complexities are rank-independent.
What about the closeness testing problems with respect to generalizations that approximates the trace distance or the von Neumann entropy? The quantum distance, defined as via the Schatten -norm, generalizes both the trace distance () and the Hilbert-Schmidt distance (). Similarly, the quantum -Tsallis entropy extends both von Neumann entropy () and quantum linear entropy ().
Interestingly, prior results show a divergence in behavior for closeness measures looser than the norm: The closeness testing problem with respect to , denoted by QSD (see Definition 17), is in BQP only for even integer via the Shift test [27]; while for odd integers , the query and sample complexities generally depend on the rank [71]. However, the techniques in [27] yield BQP algorithms for estimating for all integer . A recent work [53] further explored the closeness testing problem with respect to , and extended the observed dichotomy from integers – where the transition occurs between and – to a continuous setting, showing a sharp distinction between and any constant . These results naturally lead to an intriguing question:
Problem 1.
What is the computational complexity of the closeness testing problem with respect to ? Does a similar dichotomy hold between and constants , or does the complexity vary largely depending on whether is even or odd?
Why study problems for possibly non-integer ? The trace distance () is a fundamental closeness measure of quantum states, capturing the maximum success probability of quantum state discrimination [41, 40] and playing a key role in applications such as the security of quantum key distribution [11, 63]. For , such as , the quantum distance provides a natural lower bound on the trace distance, and addressing Problem 1 could make this bound efficiently computable. Moreover, insights from problems have previously contributed to progress on well-studied problems, as seen in [49].
Beyond their connections to problems, problems for are of independent interest. In classical scenarios, they have applications in machine learning (e.g., [46]), as well as in streaming and sketching algorithms (e.g. [43]). In quantum scenarios, the Hilbert- Schmidt distance () is widely used in quantum information theory (e.g., [42, 61]), and more recently, has been leveraged in designing near-term (variational) quantum algorithms (e.g., [5, 28]). Consequently, positive answers to Problem 1 may offer new opportunities to refine, extend, or develop techniques relevant to these areas.
A classical counterpart to Problem 1 was investigated in [69] nearly a decade ago. The main takeaway aligns with [53]: For constant , the sample complexity for distinguishing whether is at least or at most is independent of the dimension of the probability distributions and ,222The closeness measure represents the classical distance based on the norm and generalizes the total variation distance, which is recovered at . fewer samples are needed as increases. Classically, these upper bounds can be achieved by drawing a polynomial number of samples and computing the norm distance between the resulting empirical distributions. However, this approach does not directly extend to the quantum world: (1) quantum states and are generally not simultaneously diagonalizable; and (2) even when they are, estimating their eigenvalues associated with the unknown common eigenbasis remains challenging. Addressing these challenges is central to resolving Problem 1, which is the focus of our work.
1.1 Main results
We begin by stating our first main theorem, which provides a positive answer to Problem 1 when lies in the range :
Theorem 2 (Quantum estimator for , informal).
Given quantum query access to the state-preparation circuits of -qubit quantum states and , for any constant , there is a quantum algorithm for estimating within additive error and query complexity . If the state-preparation circuits have -size descriptions, then the algorithm’ time complexity is . Thus, for any constant , QSD is in BQP.
More precisely, for a given additive error , the explicit query complexity of Theorem 2 is (see Theorem 14). In combination with the samplizer [74, 75], estimating can be done using samples of and (see Theorem 16). Both upper bounds can be expressed as for the regime . In addition, if the state-preparation circuits of and have size , then Theorem 2 implies a quantum algorithm with time complexity , or equivalently, .
Previous quantum algorithms for estimating the quantum distance for constant have all relied on its powered variant, specifically the powered quantum distance:
Thus, for , the estimates of and are polynomially related.
When is an even integer, estimating follows from a folklore result via the Shift test [27], using queries or samples.333The sample complexity was noted in [62, Equations (83) and (84)]. However, for odd integer , no efficient quantum algorithm is known in general. Closeness testing of quantum states with respect to for , with query complexity , has been noted only in [32]. For general non-integer constants , the quantum query complexity of estimating was studied in [71], with polynomial dependence on the maximum rank of and . A technical comparison of our approach with this result is provided in Section 1.2.
By combining our efficient quantum estimator for in the regime (Theorem 2) with our hardness results for QSD (Theorem 3), we identify a sharp phase transition between the case of and constant , addressing Problem 1. For clarity, we summarize our main theorems and the quantitative bounds on quantum query and sample complexities, derived from both our results and prior work, in Table 1.
| QSD | QSZK-complete(*) | QSZK-complete(*) | BQP-complete | ||
| [77, 78] | Theorem˜3(2) | Theorems˜2 and 3(1) | |||
| Query | Upper | ||||
| Bound | [72] | [71] | Theorem˜14 | ||
| Compl. | Lower | ||||
| Bound | [18] | Theorem˜27(2) | Theorem˜27(1) | Theorem˜22(1) | |
| Sample | Upper | ||||
| Bound | [72] | Noted in [71, Footnote 2] | Theorem˜16 | ||
| Compl. | Lower | ||||
| Bound | [60] | Theorem˜28 | Theorem˜22(2) | ||
-
(*)
For any , the promise problem is contained in QSZK only under the polarizing regime , which can be slightly improved when (see Footnote˜1). However, establishing containment in a complexity class typically requires the natural regime , as in Theorem˜2.
Finally, we present our second main theorem, which addresses the computational hardness of QSD, as outlined in Theorem 3. In this context, PureQSD refers to a restricted variant of QSD (see also Definition 17), where the states of interest are pure.
Theorem 3 (Computational hardness of QSD).
The promise problem QSD captures the computational power of the respective complexity classes in the corresponding regimes of :
-
(1)
Easy regimes: For any , PureQSD is BQP-hard. As a corollary, QSD is BQP-complete for .
-
(2)
Hard regimes: For any , QSD is QSZK-complete, where the QSZK containment of only holds for the polarizing regime .
1.2 Proof techniques: BQP containment for constantly above
At a high level, Quantum Singular Value Transformation (QSVT) [34] implies that the main challenge in designing a quantum algorithm based on a smooth function – e.g., Grover search [38] and the OR function, or the HHL algorithm [39] and the multiplicative inverse function (see [57] for more examples) – reduces to finding an efficiently computable polynomial approximation. Once such an approximation is obtained, the algorithm follows directly using techniques from [34], with its efficiency determined entirely by the polynomial’s properties.
Now we focus on quantum algorithms for estimating the powered quantum distance. We begin by reviewing [71] and then provide a high-level overview of our approach.
The quantum query complexity of estimating the quantum distance for non-integer was first considered in [71, Theorem IV.1]. Their approach begins with the identity
where and denotes the projector onto the support subspace of . According to this identity, they aim to prepare a quantum state that is a block-encoding of (normalized) .444See Definition 10 for the formal definition of block-encoding. To this end, they first prepare a quantum state that is a block-encoding of , and then perform a unitary operator that is a block-encoding of on it.555This is because of the evolution of subnormalized density operators [71, Lemma II.2]. Finally, the (unnormalized) powered quantum distance, , can be obtained by estimating the trace of using quantum amplitude estimation [16]. After the error analysis, their approach was shown to have query complexity .666Here, denotes the fractional part of . The dependence on the rank is inherent in the approach of [71], as they have to prepare a rank-dependent quantum state that is a block-encoding of , making the rank parameters unavoidable in the error analysis.
To overcome this technical issue, we utilize an identity different from theirs:
The idea is to estimate the terms for individually, and then combine them to obtain an estimate of . Our algorithm is sketched as follows:
-
1.
Find a good approximation polynomial for .
- 2.
-
3.
Perform the Hadamard test [3] on and with outcome for each .
-
4.
Estimate by computing the expected value of .
Our algorithm is actually inspired by the trace distance estimation in [72], which corresponds to the case of . However, the approach in [72] still has a rank-dependent query complexity of , compared to the in [71].777Some readers may wonder if our approach applies to trace distance estimation () to remove the rank dependence in quantum query and sample complexities in [72]. However, the answer is generally no, as the rank dependence is intrinsic to trace distance estimation due to the polynomial dependence of the rank in the quantum query and sample complexities lower bounds (see [52, Section 2.2.2] in the full version). Nevertheless, we discover an approach for estimating the quantum distance with a rank-independent complexity as long as is constantly greater than . Specifically, we use the best uniform approximation polynomial (of degree ) for the function , as given in [31, Theorem 8.1.1]:
Our use of the best uniform approximation by polynomials is inspired by the recent work [53] on estimating the -Tsallis entropy of quantum states for non-integer , where they used the best uniform approximation polynomial for in the non-negative range (given in [67]). The difference is that in our case, we have to further consider the sign of , thereby requiring the polynomial approximation to behave well in the negative part. It turns out that the polynomial approximation given in [31] is suitable for our purpose. Having noticed this, we then use the now standard techniques (used in [48, 53]) such as Chebyshev truncations and the de La Vallée Poussin partial sum (cf. [65]) to construct efficiently computable asymptotically best approximation polynomials such that
Using this efficiently computable polynomial (with ) and with further analysis, we can then estimate the quantum distance to within additive error with the desired query upper bound in Theorem 2. Moreover, using the (multi-)samplizer [74, 75], a quantum query-to-sample simulation, we can also achieve the desired sample upper bound.
1.3 Proof techniques: QSZK completeness for near
To establish the BQP- and QSZK-hardness results in Theorem 3, we reduce the promise problems QSD and PureQSD () to the corresponding promise problems QSD and PureQSD for appropriate ranges of . The key technique underlying these reductions is the following rank-dependent inequalities that generalize the case of from [24, 25]:
Theorem 4 ( vs. , informal).
For any states and and , it holds that:
| (1) |
For , the inequalities hold in the limit as .
The proof of Theorem 4 follows from considering orthogonal positive semi-definite matrices and satisfying , and analyzing their properties carefully.
We then illuminate the hardness results in Theorem 3:
-
For the easy regime, Equation 1 becomes an equality when both and are pure states. This equality implies the BQP-hardness of PureQSD, along with the query and sample complexity lower bounds for all , establishing Theorem 3(1).
-
For the hard regime, Equation 1 is sensitive to . Particularly, for , if quantum states and are -far, meaning , it follows only that . However, when for any arbitrarily small constant , the same trace distance condition ensures that as , leading to the QSZK hardness result in Theorem 3(2) and distinct query complexity lower bounds in Table 1.
Lastly, we explain the QSZK containment in the hard regime. Simply combining Theorem 4 and the QSZK containment of QSD from [77, 78] does not work, as the resulting QSZK containment of holds only for , which is even weaker than the polarizing regime defined in Footnote 1. To address this, we establish a partial polarization lemma for (Lemma 25), which ensures that for quantum states and where is either at least or at most , we can construct new quantum states and such that is either at least or at most , as long as the parameters and are in the polarizing regime. Theorem 3(2) follows by combining this partial polarization lemma for with the polarization lemma for in [77].
1.4 Discussion and open problems
While the quantum distance and its powered version are almost computationally interchangeable for , their behavior differs greatly when :
-
The quantity corresponds to the largest eigenvalue of . The associated promise problem is BQP-hard and contains in QMA.888The verification circuit in the QMA containment simply follows from phase estimation [45], where a (normalized) eigenvector corresponding to serves as a witness state. However, establishing a BQP containment appears challenging, as does not directly admit an efficiently computable basis – unlike its classical counterpart in [69], which does.
-
The quantity takes values in for any states and , and is nonzero if and only if the states are orthogonal, with at least one being pure. Thus, even the pure-state-restricted variant of the associated promise problem, , is -hard (see [52, Appendix A] in the full version). Here, [2, 80], a subclass of PP that provides a precise variant of BQP, ensuring acceptance for all yes instances.
This fundamental difference between these quantities raises an intriguing question on :
-
(a)
What is the computational complexity of the promise problem , defined by ? Can we show that is also in BQP, or is it inherently more difficult?
1.5 Related works
Schatten -norm estimation of -local Hermitian on qubits to within additive error for and real was shown to be -complete in [19]. Given a unitary block-encoding of a matrix , in [56], they presented a quantum algorithm that estimates the Schatten -norm to relative error for integer , where a condition number satisfying is required for the case of odd .
The query complexity of -dimensional quantum state certification (i.e., determine whether two quantum states are identical or -far) with respect to trace distance was shown to be in [32]. The query complexity of trace distance estimation was shown to be in [71] and later improved to in [72], where is the rank of the quantum states, confirming a conjecture in [25] that low-rank trace distance estimation is in BQP. Both Low-rank trace distance and fidelity estimations are known to be BQP-complete [1, 72]. Based on the approach of [72], space-bounded quantum state discrimination with respect to trace distance was shown to be -complete in [48]. In addition to trace distance, fidelity is another important measure of the closeness between quantum states. The query complexity of fidelity estimation was shown to be in [76] and later improved to in [71] and to in [33]. Recently, the query complexity of pure-state trace distance and fidelity estimations was shown to be in [70] and was recently extended in [29] to estimating fidelity of a mixed state to a pure state.
In addition to the query complexity, the sample complexity has also been studied in the literature. In [7], the sample complexity of -dimensional quantum state certification was shown to be with respect to trace distance and with respect to fidelity. The sample complexity of trace distance estimation is known to be in [72] and that of fidelity estimation is known to be , where is the rank of quantum states. The sample complexity of pure-state squared fidelity estimation is known to be via the SWAP test [17], where the matching lower bound was given in [4]. Recently, the sample complexity of pure-state trace distance and fidelity estimations was shown to be in [73], which was achieved by using the samplizer in [75].
2 Preliminaries
We assume a fundamental knowledge of quantum computation and quantum information theory. For an introduction, we refer the reader to the textbook [59].
We adopt the following notations throughout the paper: (1) ; (2) denotes , while denotes ; and (3) represents for integer . In addition, the Schatten -norm of a matrix is defined as , where . For simplicity, we use the notation to denote the operator norm (equivalently, the Schatten -norm) of a matrix .
2.1 Closeness measures for quantum states
We start by defining the trace distance and providing useful properties of this distance:
Definition 5 (Trace distance).
Let and be two quantum states that are mixed in general. The trace distance between and is defined by
Notably, the trace distance is a distance metric (e.g., [79, Lemma 9.1.8]).
Next, we define the quantum distance and its powered version, which generalize the trace distance () using the Schatten norm. Notably, the quantum distance coincides with the Hilbert-Schmidt distance when :
Definition 6 (Quantum distance and its powered version).
Let and be two quantum states that are mixed in general. The quantum distance and its powered version between and are defined as follows:
Here, the Schatten -norm of is given by .
By the monotonicity of the Schatten norm, e.g., [6, Equation (1.31)], it holds that:
As a corollary of the Davis convexity theorem [26], the quantum distance also serves as a distance metric, whereas its powered version does not. Further inequalities for the trace distance and the quantum distance are given in the full version, particularly in [52, Section 2.1].
Lastly, we require the following relationship for additive error estimation between the quantum distance and its powered version. The proof is provided in the full version, specifically in [52, Proposition 2.5].
Proposition 7 ( vs. powered ).
The quantum distance and its powered version are related through the equality . Accordingly, if is an estimate of to within additive error , then serves as an estimate of to within additive error .
2.2 Closeness testing of quantum states via state-preparation circuits
We begin by defining the closeness testing of quantum states with respect to the trace distance, denoted as QSD,999While Definition 8 aligns with the classical counterpart of QSD defined in [66, Section 2.2], it is slightly less general than the definition in [77, Section 3.3]. Specifically, Definition 8 assumes that the input length and the output length are polynomially equivalent, whereas [77, Section 3.3] allows for cases where the output length (e.g., a single qubit) is much smaller than the input length. and two variants of this problem:
Definition 8 (Quantum State Distinguishability Problem, QSD, adapted from [77, Section 3.3]).
Let and be quantum circuits acting on qubits (“input length”) and having specified output qubits (“output length”), where is a polynomial function of . Let denote the quantum state obtained by running on state and tracing out the non-output qubits. Let and be efficiently computable functions. Decide whether:
-
Yes: A pair of quantum circuits such that ;
-
No: A pair of quantum circuits such that .
Furthermore, we denoted the restricted version, where and are pure states, as PureQSD.
In this work, we consider the purified quantum access input model, as defined in [77], in both white-box and black-box scenarios:
-
White-box input model: The input of the problem QSD consists of descriptions of polynomial-size quantum circuits and . Specifically, for , the description of includes a sequence of polynomially many - and -qubit gates.
-
Black-box input model: In this model, instead of providing the descriptions of the quantum circuits and , only query access to is allowed, denoted as for . For convenience, we also allow query access to and controlled-, denoted by and controlled-, respectively.
In addition to query complexity, defined within the black-box input model, sample complexity refers to the number of copies of quantum states and needed to accomplish a specific closeness testing task. Useful lemmas on computational hardness and quantitative lower bounds for query and sample complexities are available in the full version, specifically in [52, Section 2.2].
2.3 Polynomial approximations
We now present useful tools on best uniform polynomial approximations. In addition, definitions and lemmas on Chebyshev expansion and truncations can be found in the full version, particularly in [52, Section 2.3.2].
Let be a continuous function defined on the interval that we aim to approximate using a polynomial of degree at most . We define as a best uniform approximation on to of degree if, for any degree- polynomial approximation of , it holds that .
The best uniform (polynomial) approximation of positive (constant) powers was first established by Serge Bernstein [14, 13]. However, the focus here is on the best uniform approximation of signed positive powers , as stated in Lemma 9. This result is often attributed to Bernstein’s work (see, e.g., [68, Equation (10.2)]), and a proof of a more general version can be found in [31, Theorem 8.1.1].
Lemma 9 (Best uniform approximation of signed positive powers, adapted from [31, Theorem 8.1.1]).
For any positive real (constantly large) order , let be the best uniform polynomial approximation for of degree , where is a constant depending on . Then, for sufficiently small , .
2.4 Quantum algorithmic toolkit
In this subsection, we recap the quantum singular value transformation. Additional quantum algorithmic tools, including useful quantum algorithmic subroutines and the quantum samplizer can be found in the full version, specifically in [52, Sections 2.4.2 and 2.4.3].
2.4.1 Quantum singular value transformation
We start by defining block-encoding:
Definition 10 (Block-encoding).
A linear operator on an -qubit Hilbert space is said to be an -block-encoding of an -qubit linear operator , if , where is the -qubit identity operator and is the operator norm.
Then, we state the quantum singular value transformation:
Lemma 11 (Quantum singular value transformation, [34, Theorem 31]).
Suppose that unitary operator is a -block-encoding of Hermitian operator , and is a polynomial of degree with for . Then, we can implement a quantum circuit that is a -block-encoding of , by using queries to and one- and two-qubit quantum gates. Moreover, the classical description of can be computed in deterministic time .
3 Efficient quantum algorithms for estimating quantum distance
In this section, we present efficient quantum algorithms for estimating the quantum distance when . These algorithms utilize either queries to state-preparation circuits or samples of the states and . The core of our approach is an efficient uniform approximation to signed positive constant power functions (Lemma 12), which provides a uniform error bound over the entire interval .
This uniform polynomial approximation enables a query-efficient quantum algorithm for estimating through its powered version , as shown in Theorem 14. As a result, we establish a BQP containment of the promise problem QSD defined in Section 4. Additionally, by leveraging the multi-samplizer in [73], we devise a sample-efficient quantum algorithm for estimating , detailed in Theorem 16.
3.1 Efficient uniform approximations of signed positive powers
Using the averaged Chebyshev truncation specified in [52, Section 2.3.2] in the full version, we provide an efficiently computable uniform polynomial approximation of signed positive constant powers (see also [52, Lemma 3.1]):
Lemma 12 (Efficient uniform polynomial approximation of signed positive powers).
Let be a positive real (constantly large) number. For any , there is a degree- polynomial , where and is a constant depending on , that can be deterministically computed in time. For sufficiently small , it holds that:
More specifically, the degree- uniform approximation of signed positive powers (see Lemma 9) ensures that the degree- averaged Chebyshev truncation (given in [52, Equation 2.3]) achieves almost the same approximation error bound, with efficiently computable Chebyshev coefficients. The complete proof is available in the full version, specifically in [52, Section 3.1].
3.2 Quantum distance estimation for constantly large
3.2.1 Query-efficient quantum algorithm for estimating powered
We now present efficient quantum query algorithms for estimating and , as presented in [52, Lemma 3.2] in the full version, and the complete proof is provided in [52, Section 3.2.1]:
Lemma 13 (Powered quantum distance estimation via queries).
Suppose that and are unitary operators that prepare purifications of mixed quantum states and , respectively. For every constantly large , there is a quantum query algorithm that estimates to within additive error by using queries to and .
By combining Proposition 7 with Lemma 13 for additive error , we obtain a quantum query algorithm for estimating when is constantly large, as also stated in [52, Theorem 3.3] in the full version:
Theorem 14 (Quantum distance estimation via queries).
Suppose that and are unitary operators that prepare purifications of mixed quantum states and , respectively. For every constantly large , there is a quantum query algorithm that estimates to within additive error by using queries to and .
3.2.2 Sample-efficient quantum algorithm for estimating powered
We proceed by describing efficient quantum sample algorithms for and , as stated in Lemma 15 and in [52, Lemma 3.4] in the full version. Our sample algorithms are obtained by combining the quantum query algorithm from Lemma 13 with the samplizer [74, 75]. The corresponding explanatory framework is presented in in [52, Algorithm 1], and the detailed proof is given in [52, Section 3.2.2].
Lemma 15 (Powered quantum distance estimation via samples).
For every constantly large , can be estimated to within additive error on a quantum computer by using samples of and .
By combining Proposition 7 with Lemma 15 for additive error , we obtain a quantum sample algorithm for estimating when is constantly large, as also stated in [52, Theorem 3.5] in the full version:
Theorem 16 (Quantum distance estimation via samples).
For every constantly large , there is a quantum sample algorithm that estimates the quantum distance to within additive error by using samples of and .
4 Hardness and lower bounds for constantly above
We begin by introducing a generalization of QSD from [77], where the trace distance is replaced with the quantum distance as the closeness measure:
Definition 17 (Quantum State Distinguishability Problem with Schatten -norm, QSD).
Let and be quantum circuits acting on qubits (“input length”) and having specified output qubits (“output length”), where is a polynomial function of . Let denote the quantum state obtained by running on state and tracing out the non-output qubits. Let and be efficiently computable functions. Decide whether:
-
Yes: A pair of quantum circuits such that ;
-
No: A pair of quantum circuits such that .
Moreover, we denoted the restricted version, where and are pure states, as PureQSD.
In the remainder of this section, we establish rank-dependent inequalities between the quantum distance and the trace distance (Theorem 18) in Section 4.1. These inequalities facilitate reductions that demonstrate the BQP hardness (Theorem 21) and derive quantitative lower bounds on queries and samples (Theorem 22) for PureQSD in Section 4.2.
4.1 Rank-dependent inequalities between and the trace distance
We generalize the rank-dependent inequalities between the (squared) Hilbert-Schmidt distance and the trace distance, as demonstrated in [24, Appendix G] and [25, Theorem 1] for the case of , to all :
Theorem 18 ( vs. ).
Let and be quantum states. The following holds:
-
(1)
For any in the range ,
-
(2)
For ,
It is worth noting that Item 1 and Item 2 in Theorem 18 are consistent, specifically
Additionally, the inequalities in Theorem 18 sharpen the inequalities between the trace norm and the Schatten norm (see, e.g., [6, Equation (1.31)]):
| (2) |
By considering the maximum rank of and , we can derive a simplified form of Theorem 18 for convenience:
Corollary 19 ( vs. , simplified).
For any quantum states and , the following holds:
Moreover, for pure quantum states, Theorem 18 yields the following equality:
Corollary 20 ( for pure states).
For any pure states and , we have:
The proof of Theorem 18 (see also [52, Theorem 4.2] in the full version) relies on the positive semi-definite matrices and , defined for each as . These matrices are orthogonal and satisfy the relation . The complete proof is provided in [52, Section 4.1].
4.2 Computational hardness and lower bounds
We first state the computational hardness result of PureQSD with , obtained by a reduction from PureQSD, which is BQP hard [64], as stated in Theorem 21 and in [52, Theorem 4.5]. The proof can be found in the full version, specifically in [52, Section 4.2].
Theorem 21 (PureQSD is BQP-hard).
For any and , it holds that:
Next, we provide lower bounds on the query and sample complexities for PureQSD by reducing to those for PureQSD in [70], as presented in Theorem 22 and in [52, Theorem 4.6]. The proof is provided in the full version, particularly in [52, Section 4.2].
Theorem 22 (Quantitative lower bounds for PureQSD).
For any and , there exist -qubit pure states and such that deciding whether is at least or exactly requires:
-
(1)
Queries: In the purified quantum access model, the quantum query complexity is .
-
(2)
Samples: The quantum sample complexity is .
5 Quantum distance estimation for near
In this section, we establish that QSD is QSZK-complete for (see also [52, Theorem 5.1] in the full version), extending the prior result that QSD () is QSZK-complete, as shown in [77]:
Theorem 23 (QSD is QSZK-complete for near ).
Let and be efficiently computable functions such that . Then, for any , it holds that:
Moreover, is QSZK-hard if and for every constant and sufficiently large integer .
The main challenge in proving Theorem 23 is to establish a QSZK containment of QSD under the polarizing regime .101010Notably, similar to the classical cases [12], by reducing to the Quantum Jensen-Shannon Divergence Problem (QJSP) or the Measured Quantum Triangular Discrimination Problem (measQTDP) introduced in [51], the QSZK containment of QSD holds slightly beyond the polarizing regime. A direct approach, combining the inequalities between and (Corollary 19) with the QSZK containment of QSD from [77, 78], only yields a QSZK containment of under a weaker regime, . To circumvent this, we provide a (partial) polarization lemma for (Lemma 25), which enables us to achieve the desired QSZK containment in Theorem 23.
The remainder of this section establishes the QSZK containment of QSD in Section 5.1. We then show the QSZK hardness of QSD (Theorem 26) and derive quantitative lower bounds on query complexity (Theorem 27) and sample complexity (Theorem 28) in Section 5.2.
5.1 QSZK containment via a partial polarization lemma for
Theorem 24 (QSD is in QSZK).
Let and be efficiently computable functions satisfying . Then, the following holds:
The proof of Theorem 24 (see also [52, Theorem 5.2] in the full version) can be found in [52, Section 5.1]. The main technical tool is a partial polarization lemma for (see also [52, Lemma 5.3]), which ensures that any pair within the polarizing regime can be made constantly separated:
Lemma 25 (A partial polarization lemma for ).
Let and be quantum circuits that prepare quantum states and , respectively. There is a deterministic procedure that, given an input where , outputs new quantum circuits and that prepare the states and , respectively. The resulting states satisfy that:
Here, the states and are defined over qubits. Moreover, when or , the time complexity of the procedure is .
In analogy with polarization lemmas for various classical [12, 21, 66] and quantum [51, 77] closeness measures, the proof of Lemma 25 proceeds by separately reducing the errors on both sides of the problem QSD. This is achieved using the XOR lemma ([52, Lemma 5.4]) and the direct product lemma ([52, Lemma 5.5]) for . The full statements of these two lemmas, along with the full proofs of all three, are given in the full version, particularly in [52, Section 5.1].
5.2 Computational hardness and lower bounds for near
Theorem 26 (QSD is QSZK-hard).
For any positive constant that can be made arbitrarily small, the following holds for sufficiently large :
-
(1)
For any , , is QSZK-hard, where and .
-
(2)
For any , , is QSZK-hard.
The proof of Theorem 26 (see also [52, Theorem 5.6]), based on reductions to QSD that is QSZK-hard [77, 78], is provided in the full version, particularly in [52, Section 5.2]. For any -qubit state of rank , let be the corresponding -qubit state whose eigenvalues are uniformly distributed over the support of . Next, we can establish the following lower bounds (see also [52, Theorem 5.7]), and the proof also can be found in [52, Section 5.2]:
Theorem 27 (Query complexity lower bounds for QSD).
The following query complexity lower bounds hold in the purified quantum query access model, depending on the range of , where is a constant that can be made arbitrarily small:
-
(1)
For any and , there exist an -qubit state of rank and the corresponding state such that deciding whether is at least or exactly requires queries.
-
(2)
For any , there exist a constant such that, for some -qubit state of rank and the corresponding state , estimating to within additive error requires queries.
By leveraging the same reduction used to prove Theorem 27(1), the rank-dependent sample complexity lower bound (given in [52, Lemma 2.10(2)] in the full version) for estimating the trace distance can be extended to the quantum distance with , as also stated in [52, Theorem 5.8]:
Theorem 28 (Sample complexity lower bound for QSD).
For any and , there exists an -qubit state of rank and the corresponding state such that deciding whether is at least or exactly requires samples of .
References
- [1] Jayadev Acharya, Ibrahim Issa, Nirmal V. Shende, and Aaron B. Wagner. Estimating quantum entropy. IEEE Journal on Selected Areas in Information Theory, 1(2):454–468, 2020. Preliminary version in ISIT 2019. doi:10.1109/JSAIT.2020.3015235.
- [2] Leonard M. Adleman, Jonathan Demarrais, and Ming-Deh A. Huang. Quantum computability. SIAM Journal on Computing, 26(5):1524–1540, 1997. doi:10.1137/S0097539795293639.
- [3] Dorit Aharonov, Vaughan Jones, and Zeph Landau. A polynomial quantum algorithm for approximating the Jones polynomial. Algorithmica, 55(3):395–421, 2009. Preliminary version in STOC 2006. doi:10.1007/s00453-008-9168-0.
- [4] Anurag Anshu, Zeph Landau, and Yunchao Liu. Distributed quantum inner product estimation. In Proceedings of the 54th Annual ACM SIGACT Symposium on Theory of Computing, pages 44–51, 2022. doi:10.1145/3519935.3519974.
- [5] Andrew Arrasmith, Lukasz Cincio, Andrew T. Sornborger, Wojciech H. Zurek, and Patrick J. Coles. Variational consistent histories as a hybrid algorithm for quantum foundations. Nature communications, 10(1):3438, 2019. doi:10.1038/s41467-019-11417-0.
- [6] Guillaume Aubrun and Stanisław J. Szarek. Alice and Bob Meet Banach: The Interface of Asymptotic Geometric Analysis and Quantum Information Theory, volume 223 of Mathematical Surveys and Monographs. American Mathematical Society, 2017. doi:10.1090/surv/223.
- [7] Costin Bădescu, Ryan O’Donnell, and John Wright. Quantum state certification. In Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pages 503–514, 2019. doi:10.1145/3313276.3316344.
- [8] Bernhard Baumgartner. An inequality for the trace of matrix products, using absolute values. ArXiv e-prints, 2011. arXiv:1106.6189.
- [9] Aleksandrs Belovs. Quantum algorithms for classical probability distributions. In Proceedings of the 27th Annual European Symposium on Algorithms, ESA 2019, volume 144 of LIPIcs, pages 16:1–16:11. Schloss Dagstuhl – Leibniz-Zentrum für Informatik, 2019. doi:10.4230/LIPICS.ESA.2019.16.
- [10] Avraham Ben-Aroya, Oded Schwartz, and Amnon Ta-Shma. Quantum expanders: motivation and construction. Theory of Computing, 6(3):47–79, 2010. Preliminary version in CCC 2008. doi:10.4086/toc.2010.v006a003.
- [11] Michael Ben-Or, Michał Horodecki, Debbie W. Leung, Dominic Mayers, and Jonathan Oppenheim. The universal composable security of quantum key distribution. In Theory of Cryptography, Second Theory of Cryptography Conference, TCC 2005, Cambridge, MA, USA, February 10-12, 2005, Proceedings, pages 386–406. Springer, 2005. doi:10.1007/978-3-540-30576-7_21.
- [12] Itay Berman, Akshay Degwekar, Ron D. Rothblum, and Prashant Nalini Vasudevan. Statistical difference beyond the polarizing regime. In Proceedings of the 17th International Conference on Theory of Cryptography Conference, pages 311–332. Springer, 2019. ECCC:TR19-038. doi:10.1007/978-3-030-36033-7_12.
- [13] Serge Bernstein. Sur la meilleure approximation de . Doklady Akademii Nauk SSSR, 18:379–384, 1938.
- [14] Serge Bernstein. Sur la meilleure approximation de par des polynômes de degrés très élevés. Izvestiya Akademii Nauk SSSR. Seriya Matematicheskaya, 2(2):169–190, 1938. URL: https://www.mathnet.ru/eng/im3513.
- [15] Dominic W. Berry, Andrew M. Childs, Richard Cleve, Robin Kothari, and Rolando D. Somma. Simulating Hamiltonian dynamics with a truncated Taylor series. Physical Review Letters, 114(9):090502, 2015. doi:10.1103/PhysRevLett.114.090502.
- [16] Gilles Brassard, Peter Høyer, Michele Mosca, and Alain Tapp. Quantum amplitude amplification and estimation. In Samuel J. Lomonaco, Jr. and Howard E. Brandt, editors, Quantum Computation and Information, volume 305 of Contemporary Mathematics, pages 53–74. AMS, 2002. doi:10.1090/conm/305/05215.
- [17] Harry Buhrman, Richard Cleve, John Watrous, and Ronald de Wolf. Quantum fingerprinting. Physical Review Letters, 87(16):167902, 2001. doi:10.1103/PhysRevLett.87.167902.
- [18] Mark Bun, Robin Kothari, and Justin Thaler. The polynomial method strikes back: tight quantum query bounds via dual polynomials. Theory of Computing, 16(10):1–71, 2020. Preliminary version in STOC 2018. doi:10.4086/toc.2020.v016a010.
- [19] Chris Cade and Ashley Montanaro. The quantum complexity of computing Schatten -norms. In Proceedings of the 13th Conference on the Theory of Quantum Computation, Communication and Cryptography, volume 111 of LIPIcs, pages 4:1–4:20. Schloss Dagstuhl – Leibniz-Zentrum für Informatik, 2018. doi:10.4230/LIPICS.TQC.2018.4.
- [20] Clément L. Canonne. A survey on distribution testing: your data is big. but is it blue? In Theory of Computing Library, number 9 in Graduate Surveys, pages 1–100. University of Chicago, 2020. ECCC:TR15-063. doi:10.4086/toc.gs.2020.009.
- [21] André Chailloux, Dragos Florin Ciocan, Iordanis Kerenidis, and Salil Vadhan. Interactive and noninteractive zero knowledge are equivalent in the help model. In Proceedings of the Fifth Theory of Cryptography Conference, pages 501–534. Springer, 2008. IACR ePrint:2007/467. doi:10.1007/978-3-540-78524-8_28.
- [22] Sourav Chakraborty, Eldar Fischer, Arie Matsliah, and Ronald de Wolf. New results on quantum property testing. In 30th Annual Conference on Foundations of Software Technology and Theoretical Computer Science, FSTTCS 2010, volume 8 of LIPIcs, pages 145–156. Schloss Dagstuhl – Leibniz-Zentrum für Informatik, 2010. doi:10.4230/LIPICS.FSTTCS.2010.145.
- [23] Andrew M. Childs and Nathan Wiebe. Hamiltonian simulation using linear combinations of unitary operations. Quantum Information and Computation, 12(11–12):901–924, 2012. doi:10.26421/QIC12.11-12-1.
- [24] Patrick J. Coles. Unification of different views of decoherence and discord. Physical Review A, 85(4), 2012. doi:10.1103/physreva.85.042103.
- [25] Patrick J. Coles, M. Cerezo, and Lukasz Cincio. Strong bound between trace distance and Hilbert-Schmidt distance for low-rank states. Physical Review A, 100(2):022103, 2019. doi:10.1103/physreva.100.022103.
- [26] Chandler Davis. All convex invariant functions of Hermitian matrices. Archiv der Mathematik, 8(4):276–278, 1957. doi:10.1007/bf01898787.
- [27] Artur K. Ekert, Carolina Moura Alves, Daniel K. L. Oi, Michał Horodecki, Paweł Horodecki, and Leong Chuan Kwek. Direct estimations of linear and nonlinear functionals of a quantum state. Physical Review Letters, 88(21):217901, 2002. doi:10.1103/physrevlett.88.217901.
- [28] Nic Ezzell, Elliott M. Ball, Aliza U. Siddiqui, Mark M. Wilde, Andrew T. Sornborger, Patrick J. Coles, and Zoë Holmes. Quantum mixed state compiling. Quantum Science and Technology, 8(3):035001, 2023. doi:10.1088/2058-9565/acc4e3.
- [29] Wang Fang and Qisheng Wang. Optimal quantum algorithm for estimating fidelity to a pure state. In Proceedings of the 33rd Annual European Symposium on Algorithms (ESA), 2025. arXiv:2506.23650.
- [30] Christopher A. Fuchs and Jeroen van de Graaf. Cryptographic distinguishability measures for quantum-mechanical states. IEEE Transactions on Information Theory, 45(4):1216–1227, 1999. doi:10.1109/18.761271.
- [31] Michael I. Ganzburg. Limit theorems of polynomial approximation with exponential weights, volume 192 of Memoirs of the American Mathematical Society. American Mathematical Society, 2008. doi:10.1090/memo/0897.
- [32] András Gilyén and Tongyang Li. Distributional property testing in a quantum world. In Proceedings of the 11th Innovations in Theoretical Computer Science Conference, pages 25:1–25:19, 2020. doi:10.4230/LIPIcs.ITCS.2020.25.
- [33] András Gilyén and Alexander Poremba. Improved quantum algorithms for fidelity estimation. ArXiv e-prints, 2022. arXiv:2203.15993.
- [34] András Gilyén, Yuan Su, Guang Hao Low, and Nathan Wiebe. Quantum singular value transformation and beyond: exponential improvements for quantum matrix arithmetics. In Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pages 193–204, 2019. doi:10.1145/3313276.3316366.
- [35] Oded Goldreich. Introduction to Property Testing. Cambridge University Press, 2017. doi:10.1017/9781108135252.
- [36] Oded Goldreich, Amit Sahai, and Salil Vadhan. Honest-verifier statistical zero-knowledge equals general statistical zero-knowledge. In Proceedings of the 30th Annual ACM Symposium on Theory of Computing, pages 399–408, 1998. doi:10.1145/276698.276852.
- [37] Oded Goldreich and Salil Vadhan. Comparing entropies in statistical zero knowledge with applications to the structure of SZK. In Proceedings of the Fourteenth Annual IEEE Conference on Computational Complexity, pages 54–73. IEEE, 1999. ECCC:TR98-063. doi:10.1109/CCC.1999.766262.
- [38] Lov K. Grover. A fast quantum mechanical algorithm for database search. In Proceedings of the 28th Annual ACM Symposium on Theory of Computing, pages 212–219, 1996. doi:10.1145/237814.237866.
- [39] Aram W. Harrow, Avinatan Hassidim, and Seth Lloyd. Quantum algorithm for linear systems of equations. Physical Review Letters, 103(15):150502, 2009. doi:10.1103/PhysRevLett.103.150502.
- [40] Carl W. Helstrom. Detection theory and quantum mechanics. Information and Control, 10(3):254–291, 1967. doi:10.1016/S0019-9958(67)90302-6.
- [41] Alexander S. Holevo. Statistical decision theory for quantum systems. Journal of Multivariate Analysis, 3(4):337–394, 1973. doi:10.1016/0047-259X(73)90028-6.
- [42] Zdeněk Hradil, Jaroslav Řeháček, Jaromír Fiurášek, and Miroslav Ježek. Maximum-likelihood methods in quantum mechanics. Quantum State Estimation, pages 59–112, 2004. doi:10.1007/978-3-540-44481-7_3.
- [43] Piotr Indyk. Stable distributions, pseudorandom generators, embeddings, and data stream computation. Journal of the ACM (JACM), 53(3):307–323, 2006. Preliminary version in FOCS 2000. doi:10.1145/1147954.1147955.
- [44] Shelby Kimmel, Cedric Yen-Yu Lin, Guang Hao Low, Maris Ozols, and Theodore J. Yoder. Hamiltonian simulation with optimal sample complexity. npj Quantum Information, 3(1):1–7, 2017. doi:10.1038/s41534-017-0013-7.
- [45] A. Yu. Kitaev. Quantum measurements and the Abelian stabilizer problem. ArXiv e-prints, 1995. arXiv:quant-ph/9511026.
- [46] Marius Kloft, Ulf Brefeld, Sören Sonnenburg, and Alexander Zien. -norm multiple kernel learning. The Journal of Machine Learning Research, 12:953–997, 2011. doi:10.5555/1953048.2021033.
- [47] Hirotada Kobayashi. Non-interactive quantum perfect and statistical zero-knowledge. In Proceedings of the 14th International Symposium on Algorithms and Computation, pages 178–188. Springer, 2003. doi:10.1007/978-3-540-24587-2_20.
- [48] François Le Gall, Yupan Liu, and Qisheng Wang. Space-bounded quantum state testing via space-efficient quantum singular value transformation. ArXiv e-prints, 2023. doi:10.48550/arXiv.2308.05079.
- [49] James R Lee and Assaf Naor. Embedding the diamond graph in and dimension reduction in . Geometric & Functional Analysis GAFA, 14(4):745–747, 2004. doi:10.1007/s00039-004-0473-8.
- [50] Nana Liu, Qisheng Wang, Mark M. Wilde, and Zhicheng Zhang. Quantum algorithms for matrix geometric means. npj Quantum Information, 11:101, 2025. doi:10.1038/s41534-025-00973-7.
- [51] Yupan Liu. Quantum state testing beyond the polarizing regime and quantum triangular discrimination, 2025. To appear in computational complexity. arXiv:2303.01952.
- [52] Yupan Liu and Qisheng Wang. On estimating the quantum distance. ArXiv e-prints, 2025. The full version of this paper also includes references [8, 9, 21, 22, 30, 44, 47, 50, 54, 55]. arXiv:2505.00457v2.
- [53] Yupan Liu and Qisheng Wang. On estimating the trace of quantum state powers. In Proceedings of the 2025 Annual ACM-SIAM Symposium on Discrete Algorithms (SODA), pages 947–993. SIAM, 2025. doi:10.1137/1.9781611978322.28.
- [54] Seth Lloyd, Masoud Mohseni, and Patrick Rebentrost. Quantum principal component analysis. Nature Physics, 10(9):631–633, 2014. doi:10.1038/nphys3029.
- [55] Guang Hao Low and Isaac L. Chuang. Hamiltonian simulation by qubitization. Quantum, 3:163, 2019. doi:10.22331/q-2019-07-12-163.
- [56] Alessandro Luongo and Changpeng Shao. Quantum algorithms for spectral sums. ArXiv e-prints, 2020. arXiv:2011.06475.
- [57] John M. Martyn, Zane M. Rossi, Andrew K. Tan, and Isaac L. Chuang. Grand unification of quantum algorithms. PRX quantum, 2(4):040203, 2021. doi:10.1103/PRXQuantum.2.040203.
- [58] Ashley Montanaro and Ronald de Wolf. A survey of quantum property testing. In Theory of Computing Library, number 7 in Graduate Surveys, pages 1–81. University of Chicago, 2016. doi:10.4086/toc.gs.2016.007.
- [59] Michael A. Nielsen and Isaac L. Chuang. Quantum Computation and Quantum Information. Cambridge University Press, 2010. doi:10.1017/CBO9780511976667.
- [60] Ryan O’Donnell and John Wright. Quantum spectrum testing. Communications in Mathematical Physics, 387(1):1–75, 2021. Preliminary version in STOC 2015. doi:10.1007/s00220-021-04180-1.
- [61] Palash Pandya, Omer Sakarya, and Marcin Wieśniak. Hilbert-Schmidt distance and entanglement witnessing. Physical Review A, 102(1):012409, 2020. doi:10.1103/PhysRevA.102.012409.
- [62] Yihui Quek, Eneet Kaur, and Mark M. Wilde. Multivariate trace estimation in constant quantum depth. Quantum, 8:1220, 2024. doi:10.22331/Q-2024-01-10-1220.
- [63] Renato Renner and Robert König. Universally composable privacy amplification against quantum adversaries. In Theory of Cryptography, Second Theory of Cryptography Conference, TCC 2005, Cambridge, MA, USA, February 10-12, 2005, Proceedings, pages 407–425. Springer, 2005. doi:10.1007/978-3-540-30576-7_22.
- [64] Soorya Rethinasamy, Rochisha Agarwal, Kunal Sharma, and Mark M. Wilde. Estimating distinguishability measures on quantum computers. Physical Review A, 108(1):012409, 2023. doi:10.1103/PhysRevA.108.012409.
- [65] Theodore J. Rivlin. Chebyshev Polynomials: From Approximation Theory to Algebra and Number Theory. Courier Dover Publications, 1990.
- [66] Amit Sahai and Salil Vadhan. A complete problem for statistical zero knowledge. Journal of the ACM, 50(2):196–249, 2003. Preliminary version in FOCS 1997. ECCC:TR00-084. doi:10.1145/636865.636868.
- [67] Aleksandr F. Timan. Theory of Approximation of Functions of a Real Variable, volume 34 of International Series of Monographs on Pure and Applied Mathematics. Pergamon Press, 1963. doi:10.1016/c2013-0-05307-8.
- [68] Vilmos Totik. Metric properties of harmonic measures, volume 184 of Memoirs of the American Mathematical Society. American Mathematical Society, 2006. doi:10.1090/memo/0867.
- [69] Bo Waggoner. testing and learning of discrete distributions. In Proceedings of the 2015 Conference on Innovations in Theoretical Computer Science, pages 347–356, 2015. doi:10.1145/2688073.2688095.
- [70] Qisheng Wang. Optimal trace distance and fidelity estimations for pure quantum states. IEEE Transactions on Information Theory, 70(12):8791–8805, 2024. doi:10.1109/TIT.2024.3447915.
- [71] Qisheng Wang, Ji Guan, Junyi Liu, Zhicheng Zhang, and Mingsheng Ying. New quantum algorithms for computing quantum entropies and distances. IEEE Transactions on Information Theory, 70(8):5653–5680, 2024. doi:10.1109/TIT.2024.3399014.
- [72] Qisheng Wang and Zhicheng Zhang. Fast quantum algorithms for trace distance estimation. IEEE Transactions on Information Theory, 70(4):2720–2733, 2024. doi:10.1109/TIT.2023.3321121.
- [73] Qisheng Wang and Zhicheng Zhang. Sample-optimal quantum estimators for pure-state trace distance and fidelity via samplizer. ArXiv e-prints, 2024. doi:10.48550/arXiv.2410.21201.
- [74] Qisheng Wang and Zhicheng Zhang. Quantum lower bounds by sample-to-query lifting, 2025. To appear in SIAM Journal on Computing. arXiv:2308.01794.
- [75] Qisheng Wang and Zhicheng Zhang. Time-efficient quantum entropy estimator via samplizer. IEEE Transactions on Information Theory, 2025. Preliminary version in ESA 2024. doi:10.1109/TIT.2025.3576137.
- [76] Qisheng Wang, Zhicheng Zhang, Kean Chen, Ji Guan, Wang Fang, Junyi Liu, and Mingsheng Ying. Quantum algorithm for fidelity estimation. IEEE Transactions on Information Theory, 69(1):273–282, 2023. doi:10.1109/TIT.2022.3203985.
- [77] John Watrous. Limits on the power of quantum statistical zero-knowledge. In Proceedings of the 43rd Annual IEEE Symposium on Foundations of Computer Science, pages 459–468. IEEE, 2002. doi:10.1109/SFCS.2002.1181970.
- [78] John Watrous. Zero-knowledge against quantum attacks. SIAM Journal on Computing, 39(1):25–58, 2009. Preliminary version in STOC 2006. doi:10.1137/060670997.
- [79] Mark M. Wilde. Quantum Information Theory. Cambridge University Press, 2013. doi:10.1017/9781316809976.
- [80] Tomoyuki Yamakami and Andrew C. Yao. . Information Processing Letters, 71(2):63–69, 1999. doi:10.1016/s0020-0190(99)00084-8.
