Skip to main content
Log in

Szegedy Walk Unitaries for Quantum Maps

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

Szegedy developed a generic method for quantizing classical algorithms based on random walks (Szegedy, in: 45th annual IEEE Symposium on Foundations of Computer Science, pp 32–41, 2004. https://doi.org/10.1109/FOCS.2004.53). A major contribution of his work was the construction of a walk unitary for any reversible random walk. Such unitary posses two crucial properties: its eigenvector with eigenphase 0 is a quantum sample of the limiting distribution of the random walk and its eigenphase gap is quadratically larger than the spectral gap of the random walk. It was an open question if it is possible to generalize Szegedy’s quantization method for stochastic maps to quantum maps. We answer this in the affirmative by presenting an explicit construction of a Szegedy walk unitary for detailed balanced Lindbladians—generators of quantum Markov semigroups—and detailed balanced quantum channels. We prove that our Szegedy walk unitary has a purification of the fixed point of the Lindbladian as eigenvector with eigenphase 0 and that its eigenphase gap is quadratically larger than the spectral gap of the Lindbladian. To construct the walk unitary we leverage a canonical form for detailed balanced Lindbladians showing that they are structurally related to Davies generators. We also explain how the quantization method for Lindbladians can be applied to quantum channels. We give an efficient quantum algorithm for quantizing Davies generators that describe many important dynamics of open quantum systems, for instance, the relaxation of a quantum system coupled to a bath. Our algorithm extends known techniques for simulating dynamics of quantum systems on a quantum computer.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. We will use the terms random walk, Markov chain, stochastic matrix, and stochastic map synonymously throughout the manuscript.

  2. More precisely, the walk unitary W(P) acts on the Hilbert space \(\mathbbm {C}^{\vert \Omega \vert }\otimes \mathbbm {C}^{\vert \Omega \vert }\) and the corresponding eigenvector with eigenvalue 1 is the state \(\vert \pi {\rangle } = \sum _{x\in \Omega } \sqrt{\pi _x} \vert x{\rangle } \otimes \sum _{y\in \Omega } \sqrt{p_{yx}} \vert y{\rangle }\).

  3. It is not possible to obtain a coherent version of the Hamiltonian part, just as it is impossible to quantize a permuation matrix in the classical setting.

  4. We need that the matrix \(\sigma \) has full rank because otherwise the corresponding weighted inner product would not satisfy positive-definiteness. If \(\sigma \) were a multiple of the identity matrix, then the corresponding weighted inner product would reduce to the usual HS inner product.

  5. The property for \(G^\alpha (\omega )\) in (35) is a statement about the ratio of \(G^\alpha (\pm \omega )\). It is clear that all bounded functions \(G^\alpha (\omega )\) can be rescaled to ensure that \(1\ge G^\alpha (\omega )\) holds. We point out that the widely used Metropolis-Hastings filter \(\min \{1,e^{-\omega }\}\) and Glauber filter \(e^{-\omega }/(1+e^{-\omega })\) are both bounded from above by 1.

  6. It is important that a small number of generators will suffice to generate the entire matrix algebra and, thus, guarantee that the fixed point \(\sigma \) is unique.

  7. Note that action of U on states \(\vert \psi {\rangle } \otimes \vert z{\rangle }\) with \(z\ne 0^n\) can be arbitrary.

  8. We omit the energy register B whose value is alway reversibly reset to the initial state \(\vert 0^r{\rangle }\).

References

  1. Szegedy, M.: Quantum speed-up of Markov chain based algorithms. In: 45th Annual IEEE Symposium on Foundations of Computer Science, pp. 32–41 (2004). https://doi.org/10.1109/FOCS.2004.53

  2. Gilyén, A., Su, Y., Low, G.H., Wiebe, N.: 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. STOC 2019, pp. 193–204. Association for Computing Machinery, New York, NY, USA (2019). https://doi.org/10.1145/3313276.3316366

  3. Martyn, J.M., Rossi, Z.M., Tan, A.K., Chuang, I.L.: A grand unification of quantum algorithms (2021). arXiv:2105.02859

  4. Rall, P.: Faster coherent quantum algorithms for phase, energy, and amplitude estimation. arXiv:2103.09717 (2021)

  5. Somma, R.D., Boixo, S., Barnum, H., Knill, E.: Quantum simulations of classical annealing processes. Phys. Rev. Lett. 101(13), 130504 (2008)

    Article  ADS  Google Scholar 

  6. Poulin, D., Wocjan, P.: Sampling from the thermal quantum Gibbs state and evaluating partition functions with a quantum computer. Phys. Rev. Lett. 103, 220502 (2009). https://doi.org/10.1103/PhysRevLett.103.220502

    Article  ADS  MathSciNet  Google Scholar 

  7. Harrow, A.W., Wei, A.Y.: Adaptive quantum simulated annealing for Bayesian inference and estimating partition functions. In: Proceedings of the Fourteenth Annual ACM-SIAM Symposium on Discrete Algorithms, pp. 193–212 (2020)

  8. Arunachalam, S., Havlicek, V., Nannicini, G., Temme, K., Wocjan, P.: Simpler (classical) and faster (quantum) algorithms for Gibbs partition functions. Quantum 6, 789 (2022). https://doi.org/10.22331/q-2022-09-01-789

    Article  Google Scholar 

  9. Olkiewicz, R., Zegarlinski, B.: Hypercontractivity in noncommutative Lp spaces. J. Funct. Anal. 161(1), 246–285 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  10. Temme, K., Kastoryano, M.J., Ruskai, M.B., Wolf, M.M., Verstraete, F.: The \(\chi \)-divergence and mixing times of quantum Markov processes. J. Math. Phys. 51(12), 122201 (2010). https://doi.org/10.1063/1.3511335

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Kastoryano, M.J., Temme, K.: Quantum logarithmic Sobolev inequalities and rapid mixing. J. Math. Phys. 54(5), 052202 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Bardet, I.: Estimating the decoherence time using non-commutative functional inequalities. arXiv preprint arXiv:1710.01039 (2017)

  13. Bardet, I., Rouzé, C.: Hypercontractivity and logarithmic Sobolev inequality for non-primitive quantum Markov semigroups and estimation of decoherence rates. arXiv preprint arXiv:1803.05379 (2018)

  14. Datta, N., Rouzé, C.: Relating relative entropy, optimal transport and Fisher information: a quantum HWI inequality. In: Annales Henri Poincaré, vol. 21, pp. 2115–2150 (2020). Springer

  15. Temme, K.: Lower bounds to the spectral gap of Davies generators. J. Math. Phys. 54(12), 122110 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Kastoryano, M.J., Brandao, F.G.S.L.: Quantum Gibbs samplers: the commuting case. Commun. Math. Phys. 344(3), 915–957 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Temme, K.: Thermalization time bounds for Pauli stabilizer Hamiltonians. Commun. Math. Phys. 350(2), 603–637 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Capel, A., Rouzé, C., França, D.S.: The modified logarithmic Sobolev inequality for quantum spin systems: classical and commuting nearest neighbour interactions. arXiv preprint arXiv:2009.11817 (2020)

  19. Bardet, I., Junge, M., LaRacuente, N., Rouzé, C., França, D.S.: Group transference techniques for the estimation of the decoherence times and capacities of quantum Markov semigroups. IEEE Trans. Inf. Theory 67(5), 2878–2909 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  20. Bardet, I., Capel, A., Lucia, A., Pérez-García, D., Rouzé, C.: On the modified logarithmic Sobolev inequality for the heat-bath dynamics for 1D systems. J. Math. Phys. 62(6), 061901 (2021)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Lucia, A., Pérez-García, D., Pérez-Hernández, A.: Thermalization in Kitaev’s quantum double models via tensor network techniques (2021)

  22. Temme, K., Osborne, T.J., Vollbrecht, K.G., Poulin, D., Verstraete, F.: Quantum Metropolis sampling. Nature 471(7336), 87–90 (2011)

    Article  ADS  Google Scholar 

  23. Yung, M.-H., Aspuru-Guzik, A.: A quantum–quantum Metropolis algorithm. Proc. Natl. Acad. Sci. 109(3), 754–759 (2012). https://doi.org/10.1073/pnas.1111758109

    Article  ADS  Google Scholar 

  24. Carlen, E.A., Maas, J.: Non-commutative calculus, optimal transport and functional inequalities in dissipative quantum systems. J. Stat. Phys. 178, 319–378 (2020). https://doi.org/10.1007/s10955-019-02434-w

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Alicki, R.: On the detailed balance condition for non-Hamiltonian systems. Rep. Math. Phys. 10(2), 249–258 (1976). https://doi.org/10.1016/0034-4877(76)90046-X

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Kossakowski, A., Frigerio, A., Gorini, V., Verri, M.: Quantum detailed balance and KMS condition. Commun. Math. Phys. 57, 97–110 (1977). https://doi.org/10.1007/BF01625769

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Lindblad, G.: On the generators of quantum dynamical semigroups. Commun. Math. Phys. 48, 119–130 (1976). https://doi.org/10.1007/BF01608499

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Breuer, H.-P., Petruccione, F.: The Theory of Open Quantum Systems. Oxford University Press, Oxford (2002)

    MATH  Google Scholar 

  29. Davies, E.B.: Quantum Theory of Open Systems. Academic Press, London (1976)

    MATH  Google Scholar 

  30. Davies, E.B.: Generators of dynamical semigroups. J. Funct. Anal. 34(3), 249–258 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  31. Spohn, H., Lebowitz, J.L.: Irreversible thermodynamics for quantum systems weakly coupled to thermal reservoirs. Adv. Chem. Phys. Ilya Prigogine 38, 109–142 (1979)

    Google Scholar 

  32. Wocjan, P., Janzing, D., Decker, T., Beth, T.: Measuring 4-local \(n\)-qubit observables could probabilistically solve PSPACE (2003). arXiv:quant-ph/0308011

  33. Childs, A.M.: Lecture notes on quantum algorithms. Technical report, University of Maryland (2021). https://www.cs.umd.edu/~amchilds/qa/qa.pdf

Download references

Acknowledgements

P. W. and K. T. would like to thank Patrick Rall for discussions on energy estimation and Vojtěch Havlíček for comments on the manuscript. We would also like to thank Srinivasan Arunachalam, Vojtěch Havlíček, Giacomo Nannicini, and Mario Szegedy for discussions on classical and quantum walks. Both authors acknowledge support from the IBM Research Frontiers Institute and the ARO Grant W911NF-20-1-0014.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kristan Temme.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

Additional information

Communicated by A. Childs.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Quantum–Quantum Metropolis Algorithm and Its Limitations

We mentioned in the introduction that the quantum-quantum Metropolis algorithm (QQMA) as presented in [23] only works when the spectrum of the Hamiltonian is non-degenerate. The goal of this appendix is to discuss the reasons for this limitation.

One of the problems of QQMA in the degenerate case is the implementation of the projector \(\Lambda _1\) defined below equation (6) in [23]. Before explaining this issue in detail, let us state some facts to put everything better into context.

It is stated that QQMA prepares a coherent encoding of the thermal state

$$\begin{aligned} \vert \alpha _0\rangle = \sum _{i=0}^{N-1} \sqrt{e^{-\beta E_i}/Z_\beta } \, \vert i \rangle \quad \text{ with } \quad \vert i \rangle \equiv \vert \varphi _i \rangle \otimes \vert {\tilde{\varphi }}_i \rangle , \end{aligned}$$
(A1)

where \(\vert \varphi _i\rangle \) denote the eigenstates of the Hamiltonian with corresponding eigenvalues \(E_i\). The tilde in \(\vert {\tilde{\varphi }}_i\rangle \) denotes complex conjugation of the entries of the vectors \(\vert \varphi _i\rangle \).

When the spectrum of the Hamiltonian is degenerate (and \(\beta >0)\), it may seem at first that there could exist nonequivalent coherent encodings that depend on the particular chosen eigenbasis \(\vert \varphi _i\rangle \). More precisely, let U be any unitary that commutes with the Hamiltonian H and set \(\vert \psi _i\rangle = U\vert \varphi _i\rangle \) for \(i=0,\ldots ,N-1\). Define the corresponding “alternative” coherent encoding

$$\begin{aligned} \vert \beta _0\rangle = \sum _{i=0}^{N-1} \sqrt{e^{-\beta E_i}/Z_\beta } \, \vert \psi _i \rangle \otimes \vert {\tilde{\psi }}_i \rangle . \end{aligned}$$
(A2)

Clearly, we have \((U\otimes {\tilde{U}}) \vert \alpha _0\rangle = \vert \beta _0\rangle \). However, it already holds that \(\vert \alpha _0\rangle = \vert \beta _0\rangle \) so that the coherent encoding does not depend on the chosen eigenbasis.

The central problem of QQMA in the degenerate case is the implementation of the projector \(\Lambda _1\) defined below equation (6) in [23]. The operator \(\Lambda _1\) is defined there as

$$\begin{aligned} \Lambda _1 = \sum _{i=0}^{N-1} \vert i{\rangle }{\langle } i \vert \otimes \vert 0{\rangle }{\langle } 0 \vert = \sum _{i=0}^{N-1} \vert \varphi _i{\rangle }{\langle }\varphi _i \vert \otimes \vert {\tilde{\varphi }}_i{\rangle }{\langle }{\tilde{\varphi }}_i \vert \otimes \vert 0{\rangle }{\langle }0 \vert . \end{aligned}$$
(A3)

It is obvious that only energy measurements (or measurements of energy differences) are performed in QQMA. It is essential to observe that it is impossible to realize this projector with energy measurements alone if the spectrum is degenerate. The issue is that the above projector does depend on the chosen eigenbasis \(\vert \varphi _i{\rangle }\). More precisely, let

$$\begin{aligned} \Lambda _1' = \sum _{i=0}^{N-1} \vert \psi _i{\rangle }{\langle }\psi _i\vert \otimes \vert {\tilde{\psi }}_i{\rangle }{\langle }{\tilde{\psi }}_i\vert \otimes \vert 0{\rangle }{\langle }0 \vert . \end{aligned}$$
(A4)

The problem is now that these two projectors are not equal, that is \(\Lambda _1 \ne \Lambda _1'\). (A simple example that demonstrates inequality of the projectors \(\Lambda _1\) and \(\Lambda '_1\) is \(\vert \phi _0{\rangle } = \vert 0{\rangle }, \vert \phi _1{\rangle } = \vert 1{\rangle }\) and \(\vert \psi _0{\rangle } = \vert +{\rangle }, \vert \psi _1{\rangle } = \vert -{\rangle }\).) This is in contrast to the situation for the states \(\vert \alpha _0{\rangle }\) and \(\vert \beta _0{\rangle }\). Clearly, if one can only compare energies, then one cannot distinguish between different bases of a degenerate Hamiltonian. Thus, the authors implicitly assume that the Hamiltonian has non-degenerate spectrum in [23].

While projectors of the form \(\Lambda _1\) are unphysical, it is possible to implement a projector (assuming perfect energy estimation) of the form

$$\begin{aligned} \sum _E P_E \otimes {\tilde{P}}_E, \end{aligned}$$
(A5)

where E ranges over the eigenvalues of the Hamiltonian and \(P_E\) over the projectors onto the corresponding eigensubspaces.

Observe that in the classical setting there is a way out. Although some energies of the classical Hamiltonian are equal, there is always a unique preferred basis—namely, the computational basis. But such preferred eigenbasis does not exist in the quantum case for degenerate Hamiltonians unless some additional assumptions are made.

Appendix B: Gap Amplification

The appendix is an adaptation of [33, 17.2 How to quantize a Markov chain].

Definition 6

(Quantization). Let \(Q\in \mathbbm {C}^{M\times M}\) be a Hermitian matrix, \( \vert \varphi _1{\rangle },\ldots , \vert \varphi _M{\rangle }\) an orthonormal basis of \(\mathbbm {C}^M\) consisting of eigenvectors of Q with eigenvalues \(\lambda _1=1\gneqq \lambda _2 = 1 - \Delta \ge \lambda _3 \ge \ldots \ge \lambda _M \ge -1\). We refer to \(\Delta \) as the classical gap.

Assume that \(T\in \mathbbm {C}^{N\times M}\) is an isometry from \(\mathbbm {C}^M\) to \(\mathbbm {C}^N\) and \(S\in \mathbbm {C}^{N\times N}\) is a reflection acting on \(\mathbbm {C}^N\) such that

$$\begin{aligned} T^\dagger S T = Q. \end{aligned}$$
(B6)

Then, we define the quantization of Q to be

$$\begin{aligned} U=S(2\Pi - I)\in \mathbbm {C}^{N\times N}, \end{aligned}$$
(B7)

where \(\Pi =TT^\dagger \) is an orthogonal projector in \(\mathbbm {C}^{N\times N}\). Let \(\mathcal {A}\) denote the image of T (or equivalently the image of \(\Pi \)). We refer to the value \(\theta =\arccos (1-\Delta )\) as the phase gap of the quantization U of Q.

For \(j\in \{1,\ldots ,M\}\), define the states

$$\begin{aligned} \vert \chi _j{\rangle }=T \vert \varphi _j{\rangle } \in \mathbbm {C}^N. \end{aligned}$$
(B8)

Let \({\mathcal {B}}\) be the subspace \({\mathcal {A}}+S{\mathcal {A}}\), where \(S{\mathcal {A}}=\{ S \vert \chi {\rangle }: \vert \chi {\rangle }\in {\mathcal {A}}\}\), and \({\mathcal {B}}^\perp \) the orthogonal complement of \({\mathcal {B}}\).

Lemma 9

The subspace spanned \( \vert \chi _1{\rangle },\ldots , \vert \chi _M{\rangle }\) coincides with the subspace \(\mathcal {A}\).

Proof

We have \(\sum _{j\in [M]} \vert \chi _j{\rangle }{\langle }\chi _j \vert = \sum _{j\in [M]} T \vert \varphi _j{\rangle }{\langle }\varphi _j \vert T^\dagger = T T^\dagger = \Pi \). \(\square \)

Theorem 5

(Spectrum of quantization). The subspace \({\mathcal {B}}\) and its orthogonal complement \({\mathcal {B}}^\perp \) are invariant under U. The spectrum of U restricted to \({\mathcal {B}}\) is as follows:

  1. 1.

    For \(j=1\), the one-dimensional subspace \(\mathcal {V}_1\) spanned by \( \vert \chi _1{\rangle }\) is invariant under U and the eigenvector of U in \(\mathcal {V}_1\) is

    $$\begin{aligned} \vert \psi _1{\rangle } = \vert \chi _1{\rangle } \in {\mathcal {B}} \end{aligned}$$
    (B9)

    with eigenvalue 1.

  2. 2.

    For \(j\ge 2\), the two-dimensional subspace \(\mathcal {V}_j\) spanned by \( \vert \chi _j{\rangle }\) and \(S \vert \chi _j{\rangle }\) is invariant under U and the two orthogonal eigenvectors of U in \(\mathcal {V}_j\) are

    $$\begin{aligned} \vert \psi ^{\pm }_j{\rangle } = \vert \chi _j{\rangle } - \mu ^{\pm }_j S \vert \chi _j{\rangle } \in {\mathcal {B}} \end{aligned}$$
    (B10)

    with corresponding eigenvalues \(\mu ^{\pm }_j\), where

    $$\begin{aligned} \mu ^{\pm }_j = \lambda _j \pm i \sqrt{1-\lambda _j^2} = e^{\pm i \, \textrm{arccos}(\lambda _j)}. \end{aligned}$$
    (B11)

Proof

Let j be arbitrary. We have

$$\begin{aligned} U \vert \chi _j{\rangle }&= S (2\Pi - I) \vert \chi _j{\rangle } \end{aligned}$$
(B12)
$$\begin{aligned}&= S (T T^\dagger - I) T \vert \varphi _j{\rangle } \end{aligned}$$
(B13)
$$\begin{aligned}&= 2 S T \vert \varphi _j{\rangle } - ST \vert \varphi _j{\rangle } \end{aligned}$$
(B14)
$$\begin{aligned}&= S \vert \chi _j{\rangle } \end{aligned}$$
(B15)

and

$$\begin{aligned} U S \vert \chi _j{\rangle }&= S(2\Pi - I) S T \vert \varphi _j{\rangle } \end{aligned}$$
(B16)
$$\begin{aligned}&= S(2 T T^\dagger - I) ST \vert \varphi _j{\rangle } \end{aligned}$$
(B17)
$$\begin{aligned}&= (2S T T^\dagger S T - T) \vert \varphi _j{\rangle } \end{aligned}$$
(B18)
$$\begin{aligned}&= (2S T Q - T) \vert \varphi _j{\rangle } \end{aligned}$$
(B19)
$$\begin{aligned}&= (2\lambda _j S T - T) \vert \varphi _j{\rangle } \end{aligned}$$
(B20)
$$\begin{aligned}&= (2\lambda _j S - I) \vert \chi _j{\rangle }. \end{aligned}$$
(B21)

We see that the subspace spanned by \(\vert \chi _j{\rangle }\) and \(S\vert \chi _j{\rangle }\) is invariant under U for each j. Thus, we can find eigenvectors of U within this subspace.

First, we consider the case \(j=1\). We have

$$\begin{aligned} 1 = {\langle }\varphi _1 \vert Q \vert \varphi _1{\rangle } = {\langle }\varphi _1 \vert T^\dagger S T \vert \varphi _1{\rangle } = {\langle }\chi _1 \vert S \vert \chi _1{\rangle }, \end{aligned}$$
(B22)

implying that \( \vert \chi _1{\rangle }\) is an eigenvector of S with eigenvalue 1. Therefore, the subspace \(\mathcal {V}_1\) is one-dimensional. Moreover, we have

$$\begin{aligned} U \vert \chi _1{\rangle } = S(2\Pi - I) \vert \chi _1{\rangle } = S \vert \chi _1{\rangle } = \vert \chi _1{\rangle }, \end{aligned}$$
(B23)

where we used \(\Pi \vert \chi _1{\rangle }=T T^\dagger T \vert \varphi _1{\rangle }= T \vert \varphi _1{\rangle }= \vert \chi _1{\rangle }\), implying that \(\vert \chi _1{\rangle }\) is an eigenvector of U with eigenvalue 1.

Second, we consider the case \(j\ge 2\), that is, \(\lambda _j\) is bounded away from 1 by at least the spectral gap \(\Delta \). Let

$$\begin{aligned} \vert \psi ^{\pm }_j{\rangle } = \vert \chi _j{\rangle } - \mu ^{\pm }_j S \vert \chi _j{\rangle } \end{aligned}$$
(B24)

be an ansatz for the eigenvectors of U supported on \(\mathcal {V}_j\). We have

$$\begin{aligned} U \vert \psi ^\pm _j{\rangle }&= S \vert \chi _j{\rangle } - \mu ^\pm _j (2\lambda _j S - I) \vert \chi _j{\rangle } \end{aligned}$$
(B25)
$$\begin{aligned}&= \mu ^\pm _j \vert \chi _j{\rangle } - (2\lambda _j \mu ^\pm _j - 1) S \vert \chi _j{\rangle }. \end{aligned}$$
(B26)

Therefore, \( \vert \psi ^\pm _j{\rangle }\) is an eigenvector of U with eigenvalue \(\mu ^\pm _j\) provided that

$$\begin{aligned} (\mu ^\pm _j)^2 - 2 \lambda _j\mu ^\pm _j + 1 = 0, \end{aligned}$$
(B27)

that is,

$$\begin{aligned} \mu ^\pm _j = \lambda _j \pm i \sqrt{1 - \lambda _j^2} = e^{\pm i \, \textrm{arccos}(\lambda _j)}. \end{aligned}$$
(B28)

This implies that the subspaces \(V_j\) with \(j \ge 2\) are two-dimensional.

The union of all the subspaces \(\mathcal {V}_j\) is \(\mathcal {B}\), implying that \(\mathcal {B}\) and its orthogonal complement \(\mathcal {B}^\perp \) are invariant under U. \(\square \)

Lemma 10

(Lower bound on quantum gap). Let \(\Delta \) denote the classical gap of Q. The quantum gap of the corresponding unitary U is \(\theta =\arccos (1-\Delta )\) and is quadratically larger than the classical gap because

$$\begin{aligned} \theta \ge \sqrt{2\Delta }. \end{aligned}$$
(B29)

Proof

Write the second largest eigenvalue as \(\lambda _2 = 1 - \Delta = \cos (\theta )\) for a suitable \(\theta \). We have

$$\begin{aligned} 1 - \Delta = \cos (\theta ) \ge 1 - \frac{\theta ^2}{2}, \end{aligned}$$
(B30)

which implies the bound \(\theta \ge \sqrt{2\Delta }\). \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wocjan, P., Temme, K. Szegedy Walk Unitaries for Quantum Maps. Commun. Math. Phys. 402, 3201–3231 (2023). https://doi.org/10.1007/s00220-023-04797-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-023-04797-4

Navigation