Passive verification protocol for thermal graph states
Abstract
Graph states are entangled resource states for universal measurement-based quantum computation. Although matter qubits such as superconducting circuits and trapped ions are promising candidates to generate graph states, it is technologically hard to entangle a large number of them due to several types of noise. Since they must be sufficiently cooled to maintain their quantum properties, thermal noise is one of major ones. In this paper, we show that for any temperature , the fidelity between an ideal graph state at zero temperature and a thermal graph state , which is a graph state at temperature , can be efficiently estimated by using only one measurement setting. A remarkable property of our protocol is that it is passive, while existing protocols are active, namely they switch between at least two measurement settings. Since thermal noise is equivalent to an independent phase-flip error, our estimation protocol also works for that error. By generalizing our protocol to hypergraph states, we apply our protocol to the quantum-computational-supremacy demonstration with instantaneous quantum polynomial time circuits. Our results should make the characterization of entangled matter qubits extremely feasible under thermal noise.
I Introduction
Quantum computation is expected to outperform classical counterparts M16 . Driven by this expectation, several models for achieving it have been proposed such as the quantum circuit model NC00 , adiabatic quantum computation FGGS00 , measurement-based quantum computation (MBQC) RB01 , and topological quantum computation K03 . MBQC is one of promising universal quantum computing models due to its wide application range, from secure cloud quantum computing BFK09 ; MF13 to quantum error correction RHG06 . In this model, the quantum computation proceeds via adaptive single-qubit measurements on an entangled resource state (e.g., graph states (see Sec. II for its definition)). Here, “adaptive” means that each measurement basis can depend on all previous measurement outcomes.
So far, several physical systems have been proposed as candidates for realizing MBQC such as trapped ions LJZHMDBBR13 and superconducting qubits ABSRLRS18 . However, at least with current technology, it is hard to generate a large-scale graph state by using them. An obstacle to its generation is the necessity of cooling. To maintain their quantum properties, such as superposition and entanglement, they must be cooled below a few tens of millikelvins. If the cooling is not sufficient, the graph state is affected by thermal noise and consequently becomes a thermal graph state (see Sec. II for the definition). Several effects of thermal noise on graph states are explored in Ref. JDBBR09 . In particular, a critical temperature at which the graph state becomes a quantum state with only classical correlations is discussed in Ref. HV10 .
As the size of MBQC increases, it becomes more important to efficiently check whether a target graph state is precisely prepared, namely to efficiently estimate the fidelity between and an actually prepared state that suffers from any noise . This task is called the verification of graph states. Although the fidelity can be estimated by using quantum state tomography SBRF93 ; H97 ; BDPS99 or direct fidelity estimation FL11 , these protocols require an exponential number of copies of . Therefore, several efficient verification protocols tailored for graph states have been proposed HM15 ; MNS16 ; FH17 ; TM18 ; TMMMF19 ; ZH19 ; ZH19A ; MK20 . In brief, these protocols can be formalized as follows: first, a quantum computer generates some copies of and sends each qubit one by one to a verifier. Then the verifier randomly chooses a measurement basis and measures the received state in the basis. He/she repeats the same procedure for all copies of . Finally, by processing all measurement outcomes with a classical computer, he/she outputs an estimated value of (or a lower bound on) the fidelity. Note that in most cases, only (non-adaptive) single-qubit measurements and efficient classical operations are allowed for the verifier because multi-qubit operations are technologically demanding. In this paper, we also consider the same restriction on the verifier to simplify the requirement for the verifier as much as possible.
The above existing protocols are active, that is the verifier must switch between some kinds of measurement settings. In many practical scenarios, the switching of measurement settings could be slow, and in some cases, it may be impossible (e.g., see Ref. LZLZ21 ). For example, it would be somewhat demanding to change measurement bases in the IBM quantum cloud service. It is theoretically and experimentally important to clarify how many measurement settings are required for the verification of graph states.
If any -qubit unitary operator is allowed for the verifier, the necessary number of measurement settings can be trivially reduced to . This is because by using the unitary operator such that , the verifier can perform the measurement , where is the two-dimensional identity operator. However, as we have mentioned above, we would prefer not to allow the verifier to perform such multi-qubit operations. Furthermore, if the verifier can perform such , then he/she can generate the ideal graph state by his/herself, that is he/she can completely remove the thermal noise, which is unrealistic. Therefore, we should consider the necessary number of measurement settings under the assumption that the verifier can perform only non-adaptive single-qubit projective measurements. Unfortunately, under this assumption, at least two measurement settings are required for the verification of any bipartite pure entangled state that includes a subclass of graph states LZLZ21 . Even if adaptive single-qubit projective measurements are allowed for the verifier, at least two measurement settings are still necessary ZH19A .
In this paper, we circumvent the no-go result by assuming that the noise is thermal noise. More precisely, we propose a passive verification protocol for thermal graph states that requires only one measurement setting. Since the thermal noise is a major obstacle hindering the generation of large-scale graph states with matter qubits, our protocol is still sufficiently practical under this assumption. Since the thermal noise can be rephrased as an independent phase-flip error, our protocol also works for MBQC with photonic graph states. Photons are another promising candidate for qubits, and several experiments have been performed to generate photonic graph states WRRSWVAZ05 ; LZGGZYGYP07 and photonic thermal graph states AKCACWR14 . Furthermore, by generalizing our protocol to thermal hypergraph states that are generalizations of thermal graph states (see Sec. II for the definition), we apply our protocol to the demonstration of quantum computational supremacy with instantaneous quantum polynomial time (IQP) circuits BMS16 . The demonstration of quantum computational supremacy is to generate a probability distribution that cannot be efficiently generated with classical computers (under plausible complexity-theoretic assumptions) HM17 . Although its proof-of-principle experiments have already been performed by using random unitary circuits with superconducting qubits google ; ZC5D2FG4H4L7Q2RS2W4XY7Z8LPZP21 and a boson sampler with squeezed states ZWDCPLQWDHHYZLLJGYYWLLLP20 , an experiment with IQP circuits has not yet been performed. Our protocol should facilitate the realization of large-scale IQP circuits.
The rest of this paper is organized as follows: in Sec. II, we review graph and hypergraph states and their thermal analogues. In Sec. III, as a main result, we present our passive verification protocol for thermal graph states and show that it is optimal in a sense. In Sec. IV, we compare our protocol with existing protocols. In Sec. V, we generalize our protocol to thermal hypergraph states and apply it to the demonstration of quantum computational supremacy with IQP circuits. Section VI concludes the paper with a brief discussion. In Appendices A and B, we provide a proof of Theorem 1.
II Thermal graph and hypergraph states
In this section, we formally define thermal graph and hypergraph states. To this end, we first review graph BR01 and hypergraph states RHBM13 . A graph is a pair of the set of vertices and the set of edges. An -qubit graph state corresponding to the graph is defined as
(1) |
where , and is the controlled- () gate applied on the th and th qubits. It is known that Eq. (1) is a unique common eigenstate with eigenvalue of stabilizer operators , where
(2) |
for all . Here, and are the Pauli- and operators acting on the th and th qubits, respectively, and the product is taken over all vertices such that .
Hypergraph states are defined by generalizing graphs to hypergraphs. A hypergraph is a pair of the set of vertices and the set of hyperedges that can connect more than two vertices, while edges can connect only two vertices. Let be the union of and , which are sets of hyperedges connecting two and three vertices, respectively FN1 . An -qubit hypergraph state corresponding to the hypergraph is defined as
(3) |
where is the controlled-controlled- () gate applied on the th, th, and th qubits. From Eq. (3), we notice that graph states are special cases of hypergraph states. When the set is empty, the hypergraph state becomes a graph state. For the hypergraph state , we can define generalized stabilizer operators as follows:
(4) |
The equality holds for any , and no other state can satisfy it for all .
Let with the Boltzmann constant and temperature . For a Hamiltonian ,
(5) |
where is the partition function, is called the thermal state at temperature . When
(6) |
and , we call the thermal graph state, which is a graph state at non-zero temperature. This is because when the temperature is zero, . An experiment has been performed toward the ground-state cooling for the Hamiltonians that are sums of stabilizer operators BMSNMCHRZB11 . In a similar way, we call the thermal hypergraph state when
(7) |
and . In this case, becomes when .
These thermal states can also be represented as graph and hypergraph states with an independent phase-flip error (i.e., a Pauli- error). Note that in general, the replacement of the thermal noise with a Pauli- error is not possible. By restring ideal quantum states to graph and hypergraph states, we make the replacement possible. Let be the superoperator realizing the error on the th qubit with error probability . From Ref. F15 , when ,
(8) |
where . For the thermal hypergraph states, the same expression holds by replacing with in Eq. (8). These expressions are useful in evaluating our passive verification protocols.


III Passive verification protocol
In this section, we propose our passive verification protocol for thermal graph states. Note that since we assume that the thermal graph states are generated due to thermal noise, the temperature (i.e., the value of ) is unknown for the verifier. As a remarkable property, no switching between measurement bases is required for the verifier in our protocol, while previous verification protocols require it, namely they are active. For simplicity, we assume that the size of graph states is for a natural number . Let with . Our protocol runs as follows (see also Fig. 1):
-
1.
A quantum computer generates thermal graph states and sends them to the veirifier.
-
2.
The verifier measures on each received state . Let be the th outcome for .
-
3.
The verifier outputs
(9) as an estimated value of the fidelity.
The measurement of in step 2 can be realized by single-qubit Pauli measurements because is a tensor product of Pauli operators. Furthermore, by sequentially sending qubits one by one in step 1, no quantum memory is required for the verifier. To clarify these properties, we give concrete examples of our protocol in Fig. 2.
For our protocol, the following theorem holds:
Theorem 1
Let , be the value defined in Eq. (9), and and be the -qubit target graph state and the -qubit thermal graph state at unknown temperature , respectively. When and is even and at least , the estimated value satisfies
(10) | |||||
(11) |
with probability at least . Here, is the ceiling function. Particularly, in the limit of large , the inequality holds with unit probability.
The proof is given in Appendix A. Theorem 1 implies that the more the number of qubits increases, the more the precision of our protocol improves. In other words, our protocol is efficient for any temperature . Although Theorem 1 is shown for only even , our protocol can be applied when is odd by simply replacing the ideal state with another graph state . By adding , we can always make the number of vertices even.
In step 2, we choose among kinds of . This choice is optimal in the sense that it maximally improves the dependence on of the upper bound in Eq. (10). More formally, we show the following theorem:
Theorem 2
Let be the Hamming weight for any . Suppose that we replace with any in step 2. In the limit of large , the upper bound in Eq. (10) is
(12) |
with unit probability for any fixed natural number .
(Proof) In the large limit of , the values and become zero, i.e., the estimated value becomes . Therefore, from Eqs. (26) and (27) in Appendix A, we immediately obtain Eq. (12).
Since Eq. (12) is asymptotically minimized when the term of vanishes, Theorem 2 implies that our protocol is optimized when . Thus is one of optimal choices.

IV Comparison with previous protocols
There exist several verification protocols for graph states. For example, the protocol in Ref. MNS16 works for any type of error even if it is a time-correlated one. However, it requires all of , that is it needs kinds of measurement settings FN2 . Although the necessary number of measurement settings can be improved to by using the union bound as shown in Ref. TM18 , the obtained lower bound on the fidelity becomes loose. When the protocol is applied to the case of thermal noise with unknown temperature , the lower bound is
(13) | |||||
(14) |
in the limit of large , where we have used Eq. (27) to derive the second equality. Therefore, by increasing , the lower bound becomes loose. In contrast, in the limit of large , our lower bound becomes tight by increasing [see Eq. (11)]. Fig. 3 gives a comparison between and in the limit of large . It shows that our estimated value is close to the actual value even when the number of qubits and (i.e., ) are large. On the other hand, becomes distinct from when is large (i.e., is large).
The number of required measurement settings was further improved by restricting the target states. In particular, only two measurement settings are sufficient when the target state is any bipartite graph state HM15 . Recently, Li et al. generalized this result to any bipartite pure entangled state and showed that two settings cannot be improved to one LZLZ21 . They also showed that any bipartite pure product state can be verified with a single measurement setting if local projective measurements are allowed for the verifier LZLZ21 . Our result implies that a single measurement setting is sufficient even for graph states if the type of noise is restricted. By assuming thermal noise (i.e., the independent error), we circumvent the no-go result in Ref. LZLZ21 .
V Generalizations to thermal hypergraph states
In this section, we generalize our protocol to a restricted class of -qubit thermal hypergraph states that become, at the temperature , the hypergraph states in Eq. (3) such that is the union of , , , and . Note that is arbitrary, and we assume that is even. For clarity, we depict such a hypergraph state with and in Fig. 4.
Let with . In the case of hypergraph states mentioned above, by setting , we can obtain
(15) |
When , the equality holds. Otherwise, . Therefore, is a tensor product of Pauli operators (up to ), and just half of them are the Pauli- or . As a result, from the argument in Appendix A, by replacing with in step 2 of our protocol, we can also show that Theorem 1 holds for the thermal hypergraph states.
Hypergraph states in the above class have only two qubits in the column direction while they can have any number of qubits in the row direction (see Fig. 4). The above argument can also be applied even if we increase the number of qubits in the column direction to more than two. As an example, let us consider the -qubit hypergraph state in Fig. 5. Since , we can show that Theorem 1 holds for this hypergraph state. Therefore, our verification protocol is applicable to hypergraph states from which we can obtain the Union Jack states MM16 , which are universal resource states for MBQC, under postselection.
Our verification protocol can be applied to the demonstration of quantum computational supremacy with IQP circuits BMS16 under conjectures. An -qubit IQP circuit is a quantum circuit such that the initial state and measurements are and -basis measurements, respectively, and the applied quantum gate is any diagonal unitary in the basis SB09 . For IQP circuits with consisting of , , and gates, Bremner, Montanaro, and Shepherd showed that it is hard to efficiently simulate classically any probability distribution obtained from IQP circuits BMS16 . More precisely, it is hard to generate in classical polynomial time such that . Therefore, the generation of a probability distribution such that is called the demonstration of quantum computational supremacy. Since IQP circuits just measure hypergraph states in the basis when consists of , , and gates, their result shows that if appropriate hypergraph states can be prepared with sufficiently high fidelity, then quantum computational supremacy is successfully demonstrated.
Their hardness proof is based on the anticoncentration lemma and two plausible complexity-theoretic conjectures (i.e., the average-case -hardness of the approximation of output probabilities for random ’s and infiniteness of the polynomial hierarchy). The lemma implies that for any , when is randomly chosen, the probability of being larger than a certain value is at least a constant. It holds even if qubits on which gates are applied are fixed as long as qubits on which and gates are applied are randomly chosen. In other words, hypergraph states with a fixed are sufficient to show the anticoncetration lemma. This is why our hypergraph states mentioned above can be used to demonstrate quantum computational supremacy under the assumption that the two conjectures hold for our class of hypergraph states with the fixed . A similar assumption was used in Ref. MTH17 .

Although several certification protocols to check whether the demonstration of quantum computational supremacy is correctly achieved have been proposed for IQP circuits HKSE17 ; MTH17 ; TM18 ; KD19 , they are active; that is, some switching between quantum operations is required for a certifier. In contrast, we propose a passive certification protocol under the assumption that the noise is thermal noise or an independent phase-flip error. For simplicity, we set , , and in Theorem 1. Our certified quantum computational supremacy protocol proceeds as follows:
-
1.
An experimentalist runs a quantum computer to generate an -qubit thermal hypergraph state and measures . He/she repeats the same procedure times. Let be the th outcome for .
-
2.
The experimentalist calculates
(16) -
3.
If , the experimentalist generates by generating ’s again and then measuring them in the basis. Otherwise, he/she declares that the precision of the quantum computer is not enough to demonstrate quantum computational supremacy.

We show that when , the generated probability distribution satisfies with probability of , that is the experimentalist succeeds in demonstrating quantum computation supremacy. Note that if the temperature is , the inequality can be satisfied with unit probability because and . From Ref. NC00 and the Fuchs-van de Graaf inequality FG99 ,
(17) | |||||
(18) |
where is the trace norm. Therefore, from Theorem 1 with and , the required number of samples is at most , and
(19) |
with probability at least . To satisfy that the right-hand side of Eq. (19) is at most , it is sufficient to satisfy
(20) | |||||
(21) |
VI Conclusion and discussion
We have proposed a passive verification protocol for thermal graph and hypergraph states. As a remarkable property, our verification protocol requires only one measurement setting. This passiveness cannot be obtained when the noise is general, even if any single-qubit projective measurements are allowed for the verifier ZH19A ; LZLZ21 . We circumvent their no-go result by assuming thermal noise (or independent phase-flip error), which is a major obstacle to the generation of large-scale entangled states. Note that for Fock-basis photonic quantum states, a passive verification protocol has already been proposed CGKM21 . Particularly, under the two conjectures, our verification protocol for thermal hypergraph states can be used to demonstrate quantum computational supremacy in a certifiable manner. As another application, our verification protocol can also be used as a quantum sensing protocol to estimate unknown temperature . This is because the fidelity is uniquely determined by and [see Eq. (26) and Figs. 3 (a) and (b)]. Note that in this application, the product state is sufficient as an input.
It would be interesting to consider whether our protocol can be generalized to other classes of quantum states such as weighted graph states HDERNB05 ; HCDB07 and/or other types of errors such as depolarizing noises and noises induced by the finite lifetime of matter qubits. Recently, Ref. YS22 proposed a one-shot protocol to decide whether an error rate is lower than a constant value or larger than another constant value for any bounded-degree periodic graph states with depolarizing noise. Their idea may be useful in generalizing our passive verification protocol, but we leave that as a future work.
For simplicity, we have assumed that the verifier’s measurements are perfect. However, independent phase-flip errors on the measurement apparatuses can be allowed if the error probabilities are the same for all apparatuses. Since measurement errors affect the performance of our verification protocol, it should be important to also take other measurement errors into account. Although an active error-tolerant verification protocol has already been proposed FH17 , it is still open whether it can be combined with our protocol to construct a passive error-tolerant one.
ACKNOWLEDGMENTS
We thank Shion Yamashika for fruitful discussions about Appendix A. ST is supported by the Japan Society for the Promotion of Science Grant-in-Aid for Scientific Research (KAKENHI Grant No. 19K03691). RY is supported by JSPS Grant-in-Aid for Scientific Research (KAKENHI Grant No. 19K14616 and 20H01838). YT is supported by MEXT Quantum Leap Flagship Program (MEXT Q-LEAP) Grant Number JPMXS0118067394 and JPMXS0120319794 and JST [Moonshot R&D – MILLENNIA Program] Grant Number JPMJMS2061.
APPENDIX A: Proof of Theorem 1
In this appendix, we prove Theorem 1. First, we derive a value to which the estimated value in Eq. (9) converges. Let and be the numbers of and outcomes, respectively. From Eq. (9),
(22) |
Therefore, it converges to , and thus the Hoeffding inequality H63 guarantees that when , the inequality
(23) |
holds with probability at least . By using the triangle inequality and Eq. (23),
(25) | |||||
with probability at least .
The remaining task is to upper bound . To this end, we calculate and in order. By substituting Eq. (8) for , we obtain
(26) |
This is because if at least one error occurs on the graph state , it becomes orthogonal to the ideal state .
To calculate , we derive a general expression for with any . Recall that is a tensor product of Pauli operators, i.e., with and . Here, is the Pauli- operator. Let us call the th operator in . From Eq. (2), we notice that the th operator in is or if and only if . In other words, when , the th operator is or . The thermal noise can be considered as an independent phase-flip error as shown in Eq. (8), and the phase-flip error is detected by and measurements. Since the phase-flip errors occur with probability , using Eq. (8), we obtain a general expression for as
(27) | |||
(28) |
where is the Hamming weight, , and . We assumed . is the summation of the outcomes of the measurements of for the cases of phase-flip errors. The term is the number of patterns where phase-flip errors among phase-flip errors occur on the qubits such that , and the other phase-flip errors occur on the other qubits such that . The factor corresponds to the outcome of the measurement of that is if the number of errors among the qubits is even (odd). To make our argument clearer, we give the concrete derivation of Eq. (27) for the ideal state in Appendix B.
represents that the outcome of the measurement of is always when no error occurs. This is because is stabilized by any . Note that since does not imply , it does not mean the ideal case. We here just claim that when the error probability is , no error occurs on all qubits with probability .
We find that is equivalent to the coefficient of the term in the expansion of the polynomial . Therefore, substituting and in , we finally obtain
(29) | |||||
(30) |
At zero temperature, i.e., the ideal case , using , we obtain from Eq. (30) as it should be.
We confirm that Eqs. (27), (28), and (30) include the cases of and . Setting in them, we obtain
(31) | |||||
(32) |
where we have used the binomial theorem. This is the expected result. Analogously, setting in them, we obtain
(33) | |||||
(34) |
This is consistent with the fact that the outcome of the measurement of is always unity.
APPENDIX B: Derivation of Eq. (27) for
In this appendix, we derive the value of when the ideal state is . From Eq. (8), the thermal graph state is
(41) |
where is the Pauli- operator on the th qubit for . Regardless of the value of , the ideal state satisfies . Therefore,
(42) |
which corresponds to the term of in Eq. (27).
Next, we consider the second and third terms in Eq. (41), which correspond to terms of and in Eq. (27), respectively. The values of , , and depend on . For example, when ,
(43) | |||||
(44) |
Therefore,
(45) |
where and , and
(46) | |||||
where . By combining Eqs. (42), (45), and (46), we actually obtain Eq. (27). For the cases of and , a similar argument holds.
References
- (1) A. Montanaro, Quantum algorithms: an overview, npj Quantum Information 2, 15023 (2016).
- (2) M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University Press, Cambridge, 2000).
- (3) E. Farhi, J. Goldstone, S. Gutmann, and M. Sipser, Quantum Computation by Adiabatic Evolution, arXiv:quant-ph/0001106.
- (4) R. Raussendorf and H. J. Briegel, A One-Way Quantum Computer, Phys. Rev. Lett. 86, 5188 (2001).
- (5) A. Kitaev, Fault-tolerant quantum computation by anyons, Ann. Phys. 303, 2 (2003).
- (6) A. Broadbent, J. Fitzsimons, and E. Kashefi, Universal Blind Quantum Computation, in Proceedings of the 50th Annual Symposium on Foundations of Computer Science (IEEE, Los Alamitos, 2009), p. 517.
- (7) T. Morimae and K. Fujii, Blind quantum computation protocol in which Alice only makes measurements, Phys. Rev. A 87, 050301(R) (2013).
- (8) R. Raussendorf, J. Harrington, and K. Goyal, A fault-tolerant one-way quantum computer, Ann. Phys. 321, 2242 (2006).
- (9) B. P. Lanyon, P. Jurcevic, M. Zwerger, C. Hempel, E. A. Martinez, W. Dür, H. J. Briegel, R. Blatt, and C. F. Roos, Measurement-Based Quantum Computation with Trapped Ions, Phys. Rev. Lett. 111, 210501 (2013).
- (10) F. Albarrán-Arriagada, G. Alvarado Barrios, M. Sanz, G. Romero, L. Lamata, J. C. Retamal, and E. Solano, One-way quantum computing in superconducting circuits, Phys. Rev. A 97, 032320 (2018).
- (11) D. Jennings, A. Dragan, S. D. Barrett, S. D. Bartlett, and T. Rudolph, Quantum computation via measurements on the low-temperature state of a many-body system, Phys. Rev. A 80, 032328 (2009).
- (12) M. Hajdušek and V. Vedral, Entanglement in pure and thermal cluster states, New J. Phys. 12, 053015 (2010).
- (13) D. T. Smithey, M. Beck, M. G. Raymer, and A. Faridani, Phys. Rev. Lett. 70, 1244 (1993).
- (14) Z. Hradil, Phys. Rev. A 55, R1561 (1997).
- (15) K. Banaszek, G. M. D’Ariano, M. G. A. Paris, and M. F. Sacchi, Phys. Rev. A 61, 010304 (1999).
- (16) S. T. Flammia and Y.-K. Liu, Direct Fidelity Estimation from Few Pauli Measurements, Phys. Rev. Lett. 106, 230501 (2011).
- (17) M. Hayashi and T. Morimae, Verifiable Measurement-Only Blind Quantum Computing with Stabilizer Testing, Phys. Rev. Lett. 115, 220502 (2015).
- (18) T. Morimae, D. Nagaj, and N. Schuch, Quantum proofs can be verified using only single-qubit measurements, Phys. Rev. A 93, 022326 (2016).
- (19) K. Fujii and M. Hayashi, Verifiable fault tolerance in measurement-based quantum computation, Phys. Rev. A 96, 030301(R) (2017).
- (20) Y. Takeuchi and T. Morimae, Verification of Many-Qubit States, Phys. Rev. X 8, 021060 (2018).
- (21) Y. Takeuchi, A. Mantri, T. Morimae, A. Mizutani, and J. F. Fitzsimons, Resource-efficient verification of quantum computing using Serfling’s bound, npj Quantum Information 5, 27 (2019).
- (22) H. Zhu and M. Hayashi, Efficient Verification of Pure Quantum States in the Adversarial Scenario, Phys. Rev. Lett. 123, 260504 (2019).
- (23) H. Zhu and M. Hayashi, General framework for verifying pure quantum states in the adversarial scenario, Phys. Rev. A 100, 062335 (2019).
- (24) D. Markham and A. Krause, A Simple Protocol for Certifying Graph States and Applications in Quantum Networks, Cryptography 4, 3 (2020).
- (25) Y. Li, H. Zhang, Z. Li, and H. Zhu, Minimum number of experimental settings required to verify bipartite pure states and unitaries, Phys. Rev. A 104, 062439 (2021).
- (26) P. Walther, K. J. Resch, T. Rudolph, E. Schenck, H. Weinfurter, V. Vedral, M. Aspelmeyer, and A. Zeilinger, Experimental one-way quantum computing, Nature (London) 434, 169 (2005).
- (27) C.-Y. Lu, X.-Q. Zhou, O. Gühne, W.-B. Gao, J. Zhang, Z.-S. Yuan, A. Goebel, T. Yang, and J.-W. Pan, Experimental entanglement of six photons in graph states, Nat. Phys. 3, 91 (2007).
- (28) G. H. Aguilar, T. Kolb, D. Cavalcanti, L. Aolita, R. Chaves, S. P. Walborn, and P. H. Souto Ribeiro, Linear-Optical Simulation of the Cooling of a Cluster-State Hamiltonian System, Phys. Rev. Lett. 112, 160501 (2014).
- (29) M. J. Bremner, A. Montanaro, and D. J. Shepherd, Average-Case Complexity Versus Approximate Simulation of Commuting Quantum Computations, Phys. Rev. Lett. 117, 080501 (2016).
- (30) A. W. Harrow and A. Montanaro, Quantum computational supremacy, Nature (London) 549, 203 (2017).
- (31) F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. S. L. Brandao, D. A. Buell, B. Burkett, Y. Chen, Z. Chen, B. Chiaro, R. Collins, W. Courtney, A. Dunsworth, E. Farhi, B. Foxen, A. Fowler, C. Gidney, M. Giustina, R. Graff, K. Guerin, S. Habegger, M. P. Harrigan, M. J. Hartmann, A. Ho, M. Hoffmann, T. Huang, T. S. Humble, S. V. Isakov, E. Jeffrey, Z. Jiang, D. Kafri, K. Kechedzhi, J. Kelly, P. V. Klimov, S. Knysh, A. Korotkov, F. Kostritsa, D. Landhuis, M. Lindmark, E. Lucero, D. Lyakh, S. Mandr¥‘a, J. R. McClean, M. McEwen, A. Megrant, X. Mi, K. Michielsen, M. Mohseni, J. Mutus, O. Naaman, M. Neeley, C. Neill, M. Y. Niu, E. Ostby, A. Petukhov, J. C. Platt, C. Quintana, E. G. Rieffel, P. Roushan, N. C. Rubin, D. Sank, K. J. Satzinger, V. Smelyanskiy, K. J. Sung, M. D. Trevithick, A. Vainsencher, B. Villalonga, T. White, Z. J. Yao, P. Yeh, A. Zalcman, H. Neven, and J. M. Martinis, Quantum supremacy using a programmable superconducting processor, Nature (London) 574, 505 (2019).
- (32) Q. Zhu, S. Cao, F. Chen, M.-C. Chen, X. Chen, T.-H. Chung, H. Deng, Y. Du, D. Fan, M. Gong, C. Guo, C. Guo, S. Guo, L. Han, L. Hong, H.-L. Huang, Y.-H. Huo, L. Li, N. Li, S. Li, Y. Li, F. Liang, C. Lin, J. Lin, H. Qian, D. Qiao, H. Rong, H. Su, L. Sun, L. Wang, S. Wang, D. Wu, Y. Wu, Y. Xu, K. Yan, W. Yang, Y. Yang, Y. Ye, J. Yin, C. Ying, J. Yu, C. Zha, C. Zhang, H. Zhang, K. Zhang, Y. Zhang, H. Zhao, Y. Zhao, L. Zhou, C.-Y. Lu, C.-Z. Peng, X. Zhu, and J.-W. Pan, Quantum Computational Advantage via 60-Qubit 24-Cycle Random Circuit Sampling, arXiv:2109.03494.
- (33) H.-S. Zhong, H. Wang, Y.-H. Deng, M.-C. Chen, L.-C. Peng, Y.-H. Luo, J. Qin, D. Wu, X. Ding, Y. Hu, P. Hu, X.-Y. Yang, W.-J. Zhang, H. Li, Y. Li, X. Jiang, L. Gan, G. Yang, L. You, Z. Wang, L. Li, N.-L. Liu, C.-Y. Lu, and J.-W. Pan, Quantum computational advantage using photons, Science 370, 6523 (2020).
- (34) H. J. Briegel and R. Raussendorf, Persistent Entanglement in Arrays of Interacting Particles, Phys. Rev. Lett. 86, 910 (2001).
- (35) M. Rossi, M. Huber, D. Bruß, and C. Macchiavello, Quantum hypergraph states, New J. Phys. 15, 113022 (2013).
- (36) In the definition of hypergraphs, hyperedges that connect more than three vertices are also allowed. However, for our purpose, it is unnecessary to consider such hyperedges.
- (37) J. T. Barreiro, M. Müller, P. Schindler, D. Nigg, T. Monz, M. Chwalla, M. Hennrich, C. F. Roos, P. Zoller, and R. Blatt, An open-system quantum simulator with trapped ions, Nature (London) 470, 486 (2011).
- (38) K. Fujii, Quantum Computation with Topological Codes: From Qubit to Topological Fault-Tolerance (Springer, Berlin, 2015).
- (39) Note that their protocol finishes in a polynomial time in because it is sufficient to randomly choose a polynomial number of among kinds of .
- (40) J. Miller and A. Miyake, Hierarchy of universal entanglement in 2D measurement-based quantum computation, npj Quantum Information 2, 16036 (2016).
- (41) D. Shepherd and M. J. Bremner, Temporally unstructured quantum computation, Proc. R. Soc. A 465, 1413 (2009).
- (42) T. Morimae, Y. Takeuchi, and M. Hayashi, Verification of hypergraph states, Phys. Rev. A 96, 062321 (2017).
- (43) D. Hangleiter, M. Kliesch, M. Schwarz, and J. Eisert, Direct certification of a class of quantum simulations, Quantum Sci. Tech. 2, 015004 (2017).
- (44) T. Kapourniotis and A. Datta, Nonadaptive fault-tolerant verification of quantum supremacy with noise, Quantum 3, 164 (2019).
- (45) C. Fuchs and J. van de Graaf, Cryptographic distinguishability measures for quantum-mechanical states, IEEE Transactions on Information Theory 45, 1216 (1999).
- (46) U. Chabaud, F. Grosshans, E. Kashefi, and D. Markham, Efficient verification of Boson Sampling, Quantum 5, 578 (2021).
- (47) M. Hein, W. Dür, J. Eisert, R. Raussendorf, M. Van den Nest, and H.-J. Briegel, Entanglement in graph states and its applications, in Proceedings of the International School of Physics “Enrico Fermi” on “Quantum Computers, Algorithms and Chaos” (IOS Press, Varenna, 2005), p. 115.
- (48) L. Hartmann, J. Calsamiglia, W. Dür, and H. J. Briegel, Weighted graph states and applications to spin chains, lattices and gases, J. Phys. B: At. Mol. Opt. Phys. 40, S1 (2007).
- (49) H. Yamasaki and S. Subramanian, Constant-time one-shot testing of large-scale graph states, arXiv:2201.11127.
- (50) W. Hoeffding, Probability Inequalities for Sums of Bounded Random Variables, Journal of the American Statistical Association 58, 13 (1963).