Higher-order quantum transformations of Hamiltonian dynamics
Tatsuki Odake
Department of Physics, Graduate School of Science, The University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033, Japan
Hlér Kristjánsson
Current address: Perimeter Institute for Theoretical Physics, 31 Caroline Street North, Waterloo, Ontario, N2L 2Y5, Canada, and Institute for Quantum Computing, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, N2L 3G1, Canada
Department of Physics, Graduate School of Science, The University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033, Japan
Akihito Soeda
National Institute of Informatics,
Hitotsubashi 2-1-2, Chiyoda-ku, Tokyo 101-8430, Japan
Department of Physics, Graduate School of Science, The University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033, Japan
Mio Murao
Department of Physics, Graduate School of Science, The University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033, Japan
Trans-Scale Quantum Science Institute, The University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033, Japan
Abstract
We present a quantum algorithm to achieve higher-order transformations of Hamiltonian dynamics. Namely, the algorithm takes as input a finite number of queries to a black-box seed Hamiltonian dynamics to simulate a desired Hamiltonian.
Our algorithm efficiently simulates linear transformations of any seed Hamiltonian with a bounded energy range consisting of a polynomial number of terms in system size,
making use of only controlled-Pauli gates
and time-correlated randomness.
This algorithm is an instance of quantum functional programming, where the desired function is specified as a concatenation of higher-order quantum transformations.
By way of example, we demonstrate the simulation of negative time-evolution and time-reversal, and perform a Hamiltonian learning task.
Efficiently simulating the dynamics of complex quantum systems is often stated as one of the main motivations of quantum computing. While such simulation is considered hard on classical computers, a range of efficient quantum algorithms have been developed for simulating Hamiltonian dynamics [1, 2, 3, 4, 5, 6, 7].
The core principle behind the standard Hamiltonian simulation algorithms is that the desired Hamiltonian dynamics can be well-approximated by a series of (arguably) simpler quantum operations. These algorithms rely on having a classical description of the desired Hamiltonian, which can often be used for obtaining a decomposition into a sum of easily implementable terms. This limits the way we can develop large-scale, complex quantum programs for dynamics simulation. Quantum algorithms which do not require detailed descriptions of quantum resources have a higher flexibility in quantum software development.
This is related to the fundamental problem of understanding how much quantum algorithms need to rely on the classical description of their inputs in order to achieve quantum advantages in information processing.
In this work, we study Hamiltonian dynamics that can be implemented given a seed Hamiltonian without using a classical description of .
That is, we study transformations of black-box Hamiltonians.
We present a quantum algorithm that simulates the dynamics of , where is any physically realizable linear function of , given a description of and using a
black-box Hamiltonian with a bounded energy range. This algorithm is an instance of a higher-order quantum transformation on the unitary operation realized by the seed Hamiltonian dynamics. The functions that the algorithm can implement include
both the negative time-evolution and the time-reversal of an unknown Hamiltonian evolution
by considering and (transposition of in terms of the computational basis), respectively.
Such general transformations have applications ranging from fundamental physics simulations to potential improvements in state-of-the-art algorithms, such as the Hamiltonian singular value transformation [8]. We also show an application of our algorithm for Hamiltonian learning [9], in particular, a task of efficiently estimating a parameter of a multi-parameter Hamiltonian using Hamiltonian dynamics by appropriately choosing .
Our work constitutes the first systematic study of higher-order quantum transformations in the context of Hamiltonian dynamics. Higher-order quantum transformations have attracted significant attention in recent years in the context of quantum circuit transformations, and are also known as superchannels, supermaps, quantum combs and process matrices [10, 11, 12, 13, 14, 15].
Higher-order algorithms for quantum computation can be seen as an analogue of functional programming in classical computing, where the possible inputs to an algorithm are quantum channels (for example, unitaries) specified “operationally” by their input-output description only (i.e. as black boxes).
Previous works on this topic have focused on the possible transformations that can be achieved when the input channels are taken to be a finite sequence of quantum gates [10, 16, 17, 18, 19, 20, 21, 22, 23, 15, 24]. Yet, the resources available in a given computation are not always best described by a finite sequence of gates, but rather by a continuously parameterized Hamiltonian evolution. In fact, it is known that certain functions such as controllization, which cannot be implemented on black box unitaries [25, 26, 27, 28], can in fact be implemented if access to the underlying Hamiltonian evolution is given [29, 17]. This is because it is possible to apply an arbitrary fractional power of an unknown Hamiltonian evolution by changing the evolution time, whereas applying a fractional power is not possible for black box unitaries.
Summary of algorithm.—We now present our algorithm in detail (see Algorithm 1). We represent Hilbert spaces of an -qubit quantum system and a single-qubit auxiliary system by and , respectively.
We assume that we can invoke the Hamiltonian evolution of a seed Hamiltonian with an upper bound of the difference between the maximum and the minimum energy eigenvalues is given, for any time .
Algorithm 1 Simulating
1:Input:
•
A finite number of queries to a black box Hamiltonian dynamics of a seed Hamiltonian with on an -qubit system
•
An upper bound of the difference between the maximum and the minimum energy eigenvalues
•
Hermitian-preserving linear map satisfying , which can always be represented by the Pauli transfer matrix elements as
(1)
for some and functions defined by
(2)
for any tensor products of Pauli operators , where and are Pauli index vectors
•
Input state , Allowed error , Time
2:Output:
A state approximating with an error less than (measured by the 1-norm)
3:
4:Runtime:
for
5:Used Resources:
6: System: and one auxiliary qubit
7: Gates: and controlled-Pauli gates on
8:
9:Procedure:
10:Compute
11:Initialize:
12:
13:fordo
14: Randomly choose
•
with prob.
•
with prob.
15: Prepare the gate sequence [with ]
where (all gates other than are independent of ) and refers to the Hadamard gate
16:
17:endfor
18:Trace out of
19:Return
We assume that
, which ensures that the resulting evolution preserves the invariance under the global phase of . This class of covers all physically realizable linear transformations of as shown in Appendix C.
In our setting, we are given the Pauli transfer matrices [30] as in Eq. (1) of a hermitian-preserving linear map .
Our algorithm simulates the Hamiltonian evolution for any representing the time for the transformed Hamiltonian dynamics up to an error and variance (see proof in Appendix A, which relies on more general results proven in Appendix B. A similar analysis of variance is obtained in probabilistic state synthesis [31]).
The runtime of our algorithm is upper-bounded as in terms of , which is a function of elements of the Pauli transfer matrix. The total evolution time of the input dynamics is which can be shown from step 3 and step 6 of the Algorithm 1.
In Algorithm 1, the gate sequence is constructed only from controlled-Pauli gates, which are Clifford gates. The only element which may be non-Clifford is the black box dynamics . Dependence on the transformation is specified only through the probability distribution in choosing in Step 4 and through the gate in Step 5. The total runtime is calculated by multiplying the number of iterations with the runtime for implementing the controlled-Pauli gates in using CNOT gates and single-qubit Clifford gates.
Note that is independent of , even though the set of parameters has exponentially many terms.
The procedure of Algorithm 1 is summarized in Figure 1.
Figure 1: A circuit representation of Algorithm 1 implementing the transformation for an arbitrary hermitian-preserving linear map satisfying . The unitary is simulated deterministically and approximately, for an arbitrary input state and the auxiliary qubit initialized in the state . The number on the top-right of the bracket refers to the number of iterations while is the Hamiltonian evolution time of each iteration. For each iteration, an index is randomly chosen from the probability distribution , to perform the -dependent circuit inside the square brackets.Figure 2: A description of how a seed Hamiltonian is transformed after each pair of gates in Algorithm 1, for a fixed choice of . The labels \scriptsize1⃝ to \scriptsize7⃝ correspond to the Processes defined in the text.
To understand how the gate sequence transforms the Hamiltonian at each iteration, Figure 2 shows the explicit evolution of an arbitrary seed Hamiltonian after pre- and post-processing with each successive gate in the (random) sequence averaged over and , namely, .
For simplicity, is assumed to be traceless (any trace-full part is proportional to the identity and is therefore invariant under the overall transformation , by construction).
The gate sequence of is constructed in a functional programming approach, namely, by concatenations of a series of higher-order transformations, here called Processes \scriptsize1⃝ to \scriptsize7⃝.
Each of these processes is designed to implement a Hamiltonian dynamics whose Hamiltonian is given by
(9)
(16)
(21)
Applying the first controlled- gate before and after the seed Hamiltonian evolution with chosen independently from the uniform distribution in each iteration but perfectly correlated between the pre- and post-processing within each iteration (Process \scriptsize1⃝) implements Hamiltonian controllization [17]. That is, the effective evolution averaged over simulates a Hamiltonian of the form .
Process \scriptsize4⃝ is based on the identity
(24)
(27)
where are arbitrary operators, noting that for all , and . The two gates are chosen independently from the uniform distribution at each iteration.
Algorithm 1 is universal in the sense that it transforms the dynamics of any seed Hamiltonian to that of the Hamiltonian for any choice of a physically realizable linear transformation , even if is only given as a black box.
Therefore the algorithm is an instance of higher-order quantum transformations of Hamiltonian dynamics.
The algorithm makes use of a general approximation technique for simulating Hamiltonians of the form , where is a set of unitaries, is a set of positive numbers and is a seed Hamiltonian. This approximation is represented by the following circuit
(28)
where and are defined as and . The approximation is based on the randomized Hamiltonian simulation of Ref. [3] and the identity for any unitary , time and hermitian operator . This technique is also known as Hamiltonian reshaping [32]. Our algorithm can be seen as a special case of the approximation (28) with and , where the seed Hamiltonian has the form .
Applications of the algorithm.—We describe three applications of our algorithm: the negative time-evolution of Hamiltonian dynamics , the time-reversal of Hamiltonian dynamics , and a Hamiltonian single-parameter learning task of estimating one of the parameters represented by a Pauli coefficient of a Hamiltonian with Heisenberg-limited precision scaling using its dynamics .
In general, all three applications can be performed even if the dynamics is given as a black box, apart from knowledge of . However, given the knowledge that belongs to a subspace of spanned by the set for some , negative time-evolution and time-reversal can be performed in a runtime of . This property is useful when the Hamiltonian is known to be -local for some constant , in which case satisfies , so that the overall runtime is polynomial in the system size , based on the fact that is also .
In quantum algorithms that make direct use of Hamiltonian dynamics, both the positive and negative time-evolution are often assumed to be readily accessible. For example, this is required in the recent Hamiltonian singular value transformation [8]. However, in practice, a Hamiltonian evolution being native to a given hardware does not automatically guarantee that the same is true for the corresponding negative time-evolution. Therefore, the ability to efficiently simulate the negative time-evolution of any Hamiltonian given as a black box can decrease the resources required for such algorithms. On the more foundational side, given access to a black box Hamiltonian evolution, one might be interested in simulating the corresponding time-reversed evolution. For example, the evolution of an antiparticle can be described by the time-reversal of the corresponding particle evolution [33, 34].
The simulations of both negative time-evolution and time-reversal are performed by choosing the function as and , respectively, which are specified by
(29)
where .
In the definition of , the fact that , , , and are used.
In both of these cases, , thus the runtime is exponential in in general. However, when is in a subspace of spanned by the set ,
we can define
(30)
(31)
since does not depend on values of for .
In this case, so the runtime scales as , which is for a realistic Hamiltonian whose number of terms is polynomial in the system size .
For a general Hamiltonian linear transformation , if both the seed Hamiltonian and the transformed Hamiltonian have a polynomial number of terms in , then the non-zero elements of can be truncated so that the runtime has a polynomial dependence on .
We note that the runtime scales as , meaning that in order to perform the time-reversal or negative time-evolution by this algorithm, the dynamics is slowed down quadratically. An application of simulating the negative time-evolution to Hamiltonian block encoding [8] is described in Appendix D.
Finally, we consider an application of our algorithm to Hamiltonian single-parameter learning. Estimation techniques of parameters of unknown Hamiltonians for Hamiltonian learning have many applications in quantum sensing [35], analyzing properties of quantum many-body physics [36], and quantum device calibration [37]. Recently, an estimation technique achieving the Heisenberg limit for the precision scaling in the estimation of parameters of a low-interaction Hamiltonian utilizing transformations of Hamiltonian dynamics has been proposed [32]. Our algorithm can be used to extend similar techniques to a more general class of -qubit Hamiltonians.
Our estimation algorithm consists of two steps. The first step simulates using the Hamiltonian dynamics where specifies that we want to estimate and is a hermitian-preserving linear map chosen as
.
This function “filters” to keep only the coefficient and changes all other coefficients to be zero, and then sends the coefficient to the coefficient of , which is chosen for the convenience of the second step. The corresponding is given by .
The second step performs robust phase estimation [38] using similarly to the technique in [32] to obtain an estimate for , by measuring only the first qubit in our algorithm. The total evolution time is where is precision and is the failure probability, which achieves the Heisenberg-limited precision scaling. The detailed procedure and analysis of the total evolution time are given in Appendix E.
For parameter estimation of low-interaction Hamiltonians, the method of [32] can perform the full-parameter estimation in a single run with total evolution time , while our method requires polynomially longer total evolution time as we need to repeat the single parameter estimation for every parameter to perform the same task. However, the method of [32] requires exponential total evolution time for estimating a high-interaction coefficient (a coefficient of a -local Hamiltonian term with ), while our algorithm requires the same total evolution time for any coefficient. Therefore, our algorithm is suitable for estimating a single parameter of non-local Hamiltonians.
Summary and outlook.—
We presented a universal algorithm that can simulate any linear physically realizable hermitian-preserving transformation of any Hamiltonian dynamics given as a black box.
Our algorithm requires only a finite number of calls to the black box Hamiltonian dynamics and random pairs of correlated controlled-Pauli gates. We showed how our algorithm can simulate both the time-reversal and negative time-evolution of any unknown Hamiltonian dynamics, as well as an application to Hamiltonian single-parameter learning, efficiently estimating a single parameter of a multi-parameter Hamiltonian.
In our algorithm, the probability distribution for choosing multiple gates at different time steps are correlated in the sense that the gate is always used together with its adjoint , and the probabilities for picking its component controlled-Pauli gates are correlated via a joint probability distribution. This algorithm demonstrates how multiply correlated randomness can be leveraged to construct unitary operators without introducing decoherence.
Our algorithm is a starting point for the emerging field of black box Hamiltonian simulation. One possible future direction is to extend higher-order quantum transformations of Hamiltonian dynamics to Hamiltonian transformations beyond hermitian-preserving linear transformations.
We would like to thank Seiseki Akibue, Mile Gu, Toshinori Itoko, Antonio Mezzacapo, Kunal Sharma, Philip Taranto, and Satoshi Yoshida for fruitful discussions. This work was supported by the MEXT Quantum Leap Flagship Program (MEXT QLEAP) Grant No. JPMXS0118069605, the Japan Society for the Promotion of Science (JSPS) KAKENHI Grants No. 21H03394 and No. 18K13467, and partly by IBM-UTokyo lab. Research at Perimeter Institute is supported in part by the Government of Canada through the Department of Innovation, Science and Economic Development and by the Province of Ontario through the Ministry of Colleges and Universities.
References
Suzuki [1990]M. Suzuki, Fractal decomposition of
exponential operators with applications to many-body theories and Monte
Carlo simulations, Physics Letters A 146, 319 (1990).
Suzuki [1991]M. Suzuki, General theory of fractal
path integrals with applications to many-body theories and statistical
physics, Journal
of Mathematical Physics 32, 400 (1991).
Campbell [2019]E. Campbell, Random compiler for fast
Hamiltonian simulation, Physical Review Letters 123, 070503 (2019).
Berry et al. [2015]D. W. Berry, A. M. Childs,
R. Cleve, R. Kothari, and R. D. Somma, Simulating Hamiltonian dynamics with a truncated Taylor
series, Physical
Review Letters 114, 090502 (2015).
Low and Chuang [2017]G. H. Low and I. L. Chuang, Optimal Hamiltonian
simulation by quantum signal processing, Physical Review Letters 118, 010501 (2017).
Low and Chuang [2019]G. H. Low and I. L. Chuang, Hamiltonian simulation by
qubitization, Quantum 3, 163
(2019).
Childs et al. [2018]A. M. Childs, D. Maslov,
Y. Nam, N. J. Ross, and Y. Su, Toward the first quantum simulation with quantum speedup, Proceedings of the
National Academy of Sciences 115, 9456 (2018).
Lloyd et al. [2021]S. Lloyd, B. T. Kiani,
D. R. Arvidsson-Shukur,
S. Bosch, G. De Palma, W. M. Kaminsky, Z.-W. Liu, and M. Marvian, Hamiltonian singular value transformation and inverse block
encoding, arXiv
preprint arXiv:2104.01410 (2021).
Granade et al. [2012]C. E. Granade, C. Ferrie,
N. Wiebe, and D. G. Cory, Robust online hamiltonian learning, New Journal of Physics 14, 103013 (2012).
Chiribella et al. [2008a]G. Chiribella, G. M. D’Ariano, and P. Perinotti, Quantum circuit
architecture, Physical Review Letters 101, 060401 (2008a).
Chiribella et al. [2008b]G. Chiribella, G. M. D’Ariano, and P. Perinotti, Transforming quantum
operations: Quantum supermaps, Europhysics Letters 83, 30004 (2008b).
Chiribella et al. [2009]G. Chiribella, G. M. D’Ariano, and P. Perinotti, Theoretical framework
for quantum networks, Physical Review A 80, 022339 (2009).
Bisio and Perinotti [2019]A. Bisio and P. Perinotti, Theoretical framework
for higher-order quantum theory, Proceedings of the Royal Society A 475, 20180706 (2019).
Chitambar and Gour [2019]E. Chitambar and G. Gour, Quantum resource theories, Reviews of Modern
Physics 91, 025001
(2019).
Oreshkov et al. [2012]O. Oreshkov, F. Costa, and Č. Brukner, Quantum correlations with no causal
order, Nature
Communications 3, 1092
(2012).
Miyazaki et al. [2019]J. Miyazaki, A. Soeda, and M. Murao, Complex conjugation supermap of unitary quantum
maps and its universal implementation protocol, Physical Review Research 1, 013007 (2019).
Dong et al. [2019]Q. Dong, S. Nakayama,
A. Soeda, and M. Murao, Controlled quantum operations and combs, and their
applications to universal controllization of divisible unitary operations, arXiv preprint
arXiv:1911.01645 (2019).
Quintino et al. [2019a]M. T. Quintino, Q. Dong,
A. Shimbo, A. Soeda, and M. Murao, Probabilistic exact universal quantum circuits for transforming
unitary operations, Physical Review A 100, 062339 (2019a).
Quintino et al. [2019b]M. T. Quintino, Q. Dong,
A. Shimbo, A. Soeda, and M. Murao, Reversing unknown quantum transformations: Universal quantum circuit
for inverting general unitary operations, Physical Review Letters 123, 210502 (2019b).
Yoshida et al. [2022]S. Yoshida, A. Soeda, and M. Murao, Reversing unknown qubit-unitary operation,
deterministically and exactly, arXiv preprint arXiv:2209.02907 (2022).
Chiribella and Kristjánsson [2019]G. Chiribella and H. Kristjánsson, Quantum
Shannon theory with superpositions of trajectories, Proceedings of the Royal Society
A 475, 20180903
(2019).
Chiribella et al. [2013]G. Chiribella, G. M. D’Ariano, P. Perinotti, and B. Valiron, Quantum computations
without definite causal structure, Physical Review A 88, 022318 (2013).
Pollock et al. [2018]F. A. Pollock, C. Rodríguez-Rosario, T. Frauenheim, M. Paternostro, and K. Modi, Non-markovian quantum
processes: Complete framework and efficient characterization, Physical Review A 97, 012127 (2018).
Bai et al. [2020]G. Bai, Y.-D. Wu,
Y. Zhu, M. Hayashi, and G. Chiribella, Efficient algorithms for causal order discovery in quantum
networks, arXiv
preprint arXiv:2012.01731 (2020).
Gavorová et al. [2020]Z. Gavorová, M. Seidel, and Y. Touati, Topological obstructions
to implementing controlled unknown unitaries, arXiv preprint arXiv:2011.10031 (2020).
Araújo et al. [2014]M. Araújo, A. Feix,
F. Costa, and Č. Brukner, Quantum circuits cannot control unknown
operations, New
Journal of Physics 16, 093026 (2014).
Soeda [2013]A. Soeda, Limitations on quantum
subroutine designing due to the linear structure of quantum operators (Talk at the International Conference on Quantum
Information and Technology (IC- QIT), 2013).
Friis et al. [2014]N. Friis, V. Dunjko,
W. Dür, and H. J. Briegel, Implementing quantum control for unknown
subroutines, Physical Review A 89, 030303 (2014).
Nakayama et al. [2015]S. Nakayama, A. Soeda, and M. Murao, Quantum algorithm for universal implementation of
the projective measurement of energy, Physical Review Letters 114, 190501 (2015).
Chow et al. [2012]J. M. Chow, J. M. Gambetta,
A. D. Corcoles, S. T. Merkel, J. A. Smolin, C. Rigetti, S. Poletto, G. A. Keefe, M. B. Rothwell, J. R. Rozen,
et al., Universal quantum
gate set approaching fault-tolerant thresholds with superconducting qubits, Physical review
letters 109, 060501
(2012).
Akibue et al. [2023]S. Akibue, G. Kato, and S. Tani, Probabilistic state synthesis based on optimal
convex approximation, arXiv preprint arXiv:2303.10860 (2023).
Huang et al. [2023]H.-Y. Huang, Y. Tong,
D. Fang, and Y. Su, Learning many-body hamiltonians with heisenberg-limited
scaling, Physical Review Letters 130, 200403 (2023).
Weinberg [2005]S. Weinberg, The Quantum Theory of
Fields, Volume 1: Foundations (Cambridge
University Press, 2005).
de Burgh and Bartlett [2005]M. de Burgh and S. D. Bartlett, Quantum methods for
clock synchronization: Beating the standard quantum limit without
entanglement, Physical Review A 72, 042301 (2005).
Wiebe et al. [2014]N. Wiebe, C. Granade,
C. Ferrie, and D. Cory, Quantum hamiltonian learning using imperfect quantum
resources, Physical Review A 89, 042314 (2014).
Boulant et al. [2003]N. Boulant, T. F. Havel,
M. A. Pravia, and D. G. Cory, Robust method for estimating the lindblad
operators of a dissipative quantum process from measurements of the density
operator at multiple time points, Physical Review A 67, 042322 (2003).
Kimmel et al. [2015]S. Kimmel, G. H. Low, and T. J. Yoder, Robust calibration of a universal
single-qubit gate set via robust phase estimation, Physical Review A 92, 062315 (2015).
Watrous [2018]J. Watrous, The theory of quantum
information (Cambridge University Press, 2018).
Fuchs and Van
De Graaf [1999]C. A. Fuchs and J. Van
De Graaf, Cryptographic
distinguishability measures for quantum-mechanical states, IEEE Transactions on Information
Theory 45, 1216
(1999).
Gilyén et al. [2019]A. Gilyén, Y. Su,
G. H. Low, and N. 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 (2019) pp. 193–204.
Martyn et al. [2021]J. M. Martyn, Z. M. Rossi,
A. K. Tan, and I. L. Chuang, Grand unification of quantum algorithms, PRX Quantum 2, 040203 (2021).
Haah et al. [2023]J. Haah, R. Kothari,
R. O’Donnell, and E. Tang, Query-optimal estimation of unitary channels in
diamond distance, arXiv preprint arXiv:2302.14066 (2023).
Appendix A Appendix A: Error and variance analysis on Algorithm 1
In this section, we prove a theorem (Theorem 1) evaluating the error and variance of Algorithm 1 in simulating . We use the same notation and symbols appearing in Algorithm 1 in this appendix and the following appendices.
For evaluating the error of the simulated quantum operation from the target quantum operation, we use the diamond norm of a quantum operation defined as [39]
(32)
where the identity operation acts on a Hilbert space isomorphic to , is an arbitrary linear operator on , and is the trace norm defined by .
Theorem 1.
.
1.
The approximation error of Algorithm 1 is given by
(33)
where is a map defined by and is the quantum operation averaged over all random instances of operations performed in Algorithm 1.
2.
The variance of Algorithm 1 is given by
(34)
where refers to the set of all indices chosen in iterations, is the probability that is chosen, and is the unitary performed when is chosen.
Without loss of generality, we can limit the proof of Theorem 1 to the case where , where is the traceless part of . When , then will also be greater than 1, where where is the set of eigenvalues of . By noticing that the procedure of Algorithm 1 for a Hamiltonian and time is the same as that for a Hamiltonian (whose traceless part has operator norm of at most 1) and time , we can always assume that the traceless part of the seed Hamiltonian has operator norm at most 1.
Before presenting the sketch of the proof, we describe the error and variance of a simulation technique shown in the circuit of Fig. 4, which is a key element in our algorithm. Similar randomization techniques are also used in [3] (QDrift) and [32] (Hamiltonian learning). The simulation error is evaluated based on [3]. In addition, the variance is also evaluated for our algorithm.
Lemma 1.
The quantum operation represented by the following circuit
where is a Hamiltonian normalized as , where is the traceless part of , , are constants, is a unitary applied with probability , and is the iteration number given by for ,
simulates the dynamics of a transformed Hamiltonian defined as
(35)
with
in an error smaller than or equal to and the variance smaller than or equal to .
Here, the error and the variance of the simulation are defined as
(36)
(37)
respectively, where is defined by
for a density operator , is the averaged quantum operation simulated in the above circuit, refers to the set of all the indices chosen in iterations, is the probability is chosen, is the unitary operation performed when is chosen, and the identity operation belongs to a Hilbert space with an arbitrary finite dimension.
Proof.
Both and stay unchanged when the seed Hamiltonian is changed to since the two dynamics and for the same time differ only by a global phase. Therefore, giving a proof for is sufficient.
Error:
The set of hermitian operators for satisfies , thus the protocol given by
with , , and simulates in an error smaller than .
Here, the error is measured in terms of Eq. (36) with .
This is proved using the fact that the error is smaller than or equal to as shown in [3] and .
Variance: The variance is proven by using the upper bound shown by Theorem 3 of Appendix B given by
(38)
∎
We refer to linear maps expressed by Eq. (35) as linear maps in Class T.
Definition 1.
Class T of linear maps is defined as a set of linear maps which can be specified by a set of positive numbers and unitaries on through the equation
where is an hermitian operator .
Class T is closed to concatenation, namely, the following lemma holds.
Lemma 2.
A finite concatenations of Class T transformations also belong to class T. In particular, if for is specified by , then is specified by
. Here, the index sets can be individually chosen for .
Proof. It can be proved from the equation given by
∎
In the following lemma, follow the definitions in the description of Algorithm 1; .
Lemma 3.
1.
The Class T transformation specified by for is expressed as a sum of concatenations of Class T transformations as
(39)
where are defined as
(44)
(49)
(54)
for an input operator .
2.
The transformation transforms an input operator for a hermitian operator as
(59)
Note that the quantum operation performed in step 3 to step 7 of Algorithm 1 approximates the unitary . Indeed, the circuit shown in Lemma 2 for defined above and defined inside Algorithm 1 coincides with the procedures in step 3 to step 7 due to and .
The functions correspond to the transformation of the effective Hamiltonian by Processes \scriptsize1⃝ to \scriptsize7⃝ of Fig. 2.
Proof.
1. Class T transformations , , , , , , are specified by
,
,
,
,
,
,
, thus the right hand side of Eq. (39) consists of Class T transformations specified by
2. We describe how the effective Hamiltonians are transformed in each Class T transformation according to Processes \scriptsize1⃝ to \scriptsize7⃝. By defining
for a seed Hamiltonian .
Using defined by where is the traceless part of (), Eq. (69) is proved by calculating the effective Hamiltonian after each process as
(76)
(81)
(86)
(91)
(96)
(101)
(106)
(111)
(116)
(121)
(126)
(131)
(136)
as the final effective Hamiltonian coincides with the right-hand side of Eq. (69).
The transformations in and are obtained due to the equality
where is an arbitrary linear operator,
which is proved using the fact that gives a basis of , , and
∎
Proof of Theorem 1.
As we mentioned above, we limit without loss of generality the proof of Theorem 1 to the case where , where is the traceless part of .
Error
We denote three types of quantum operations, , , depending on the domain and range of the operations. In Algorithm A, is performed in step 2, is performed in step 3-7, and is performed in step 8 where and are density operators.
For representing the dynamics of transformed Hamiltonian defined by
where is the linear transformation defined in the first statement of Lemma 3, it can be shown using Lemma 2 and Lemma 3 that
Therefore, Algorithm 1 approximates in an error bounded above as
which can be shown using the relationship for arbitrary CPTP maps and where , , are Hilbert spaces [39].
Also, satisfies
which completes the proof.
Variance
Denoting the quantum operation performed in steps 3-7 of Algorithm 1 when is chosen as , we obtain
(137)
The second inequality in Eq. (137) is shown by the fact that the partial trace is a 1-norm non-increasing map, which can be proved by
where are Hilbert spaces, is an operator in , and are unitaries. The third inequality in Eq. (137) follows from Eq. (37) because is a Class T transformation.
∎
Appendix B Appendix B: General relationship between error and variance
In this appendix, we provide a theorem on the relationship between the error and variance of a general random protocol for approximating a unitary operation.
Theorem 2.
For an arbitrary unitary operation defined by with a unitary and a density operator on a Hilbert space , if a set of deterministic quantum operations (completely-positive trace-preserving maps) and a probability distribution satisfies
(138)
for some , the inequality
(139)
holds for any pure state .
Note that the equation
follows from the properties of the diamond norm [39] given by Eq. (138), thus Theorem 2 can be strengthened to
Theorem 3.
For an arbitrary unitary operation defined by with an unitary operator and a density operator on a Hilbert space , if a set of deterministic quantum operations (completely-positive trace-preserving maps) and a probability distribution satisfies
for some , the variance of the averaged operation is upper bounded by
(140)
where is the identity operation on a Hilbert space
and is a pure state on .
As will be shown in the proof, Equation (139) still holds when the condition for the bound of the error in Eq. (138) is weakened to only hold for pure states as
(141)
instead of the case of the diamond norm as in Eq. (138).
First, we rewrite the statement of Theorem 2 by introducing a set of deterministic quantum operations as
The following Lemma for the quantum operation implemented by a random protocol applying with probability provides a sufficient condition for Theorem 2.
Lemma 4.
If the error of a random protocol simulating an identity operation on any pure state satisfies
(142)
then the bound of the variance satisfies
(143)
The sufficiency of this statement is shown at the end of this appendix. We also define a complex coefficient to decompose the action of on any pure state as
(144)
where
is an operator satisfying that is
perpendicular to , namely, .
According to the Fuchs-van de Graaf inequalities[40], we have
Proof of Lemma 4:
First, we find a lower bound to the left hand side of Eq. (142).
Using the decomposition of given by Eq. (144), Eq. (142) can be bounded as
where is a set of states (density operators) perpendicular to and the set forms an orthonormal basis.
In the above evaluation, the second inequality holds due to the definition
namely, this is based on the fact that for a unitary
(147)
the term
(148)
is equal to the right hand side of the second inequality.
Next, we find a upper bound of the left hand side of Eq. (143).
Using Eq. (145), the left hand side of Eq. (143) can be bound as
Proof of Theorem 2.
The sufficiency of the statement (Eq. (142) Eq. (143)) for proving Theorem 2 is based on the equivalence of Eq. (141) (the weaker version of Eq. (138)) and Eq. (142), and the equivalence of Eq. (139) and Eq. (143). The one-norm is invariant under unitary transformations, namely, for an arbitrary linear operator and a unitary operator , holds. This is because
holds. Therefore, the left hand side of Eq. (141) and the left hand side of Eq. (142) can be shown to be equal as
The equality of the left-hand side of Eq. (139) and the left-hand side of Eq. (143) can be shown in the same way. Thus the equivalence of Eq. (141) and Eq. (142), and the equivalence of Eq. (139) and Eq. (143) can be shown.∎
Appendix C Appendix C: Universality of Algorithm 1
In this section, we show that our algorithm simulating is a universal algorithm to simulate physically realizable linear transformation on a Hamiltonian. Under the assumption that can also be seen as a Hamiltonian if an input is a Hamiltonian, we can assume that is a hermitian-preserving linear map. The universality of our algorithm can be shown in the following lemma.
Lemma 5.
The following two classes of linear maps are equal:
(a)
The class of hermitian-preserving linear maps such that the transformation is physically realizable with an arbitrarily small error
(b)
The class of hermitian-preserving linear maps such that
Since our algorithm can simulate an arbitrary in class (b), it is shown to be able to simulate arbitrary “physically realizable” .
Proof. The Algorithm 1 can simulate for such that , thus (b)(a). Assuming that is physically realizable for an such that is not proportional to , the output unitaries , for are physically distinguishable. However, the inputs and are only different up to the global phase and thus physically indistinguishable, which leads to a contradiction. Thus (a)(b) is proved.
∎
Appendix D Appendix D: The application of simulating the negative time-evolution to Hamiltonian block encoding
One application of simulating the negative time-evolution is the block-encoding of an unknown operator given as a block of an unknown Hamiltonian. In this appendix, we present an algorithm for Hamiltonian block encoding utilizing our algorithm and analyze the approximation errors of the algorithm.
D.0.1 Algorithm for Hamiltonian block encoding
Assume we are given access to the Hamiltonian dynamics of a Hamiltonian with an upper bound of the maximum difference in energy eigenvalues represented as
(152)
where diagonal blocks can be of arbitrary operators, and the smallest singular value of the operator on the off-diagonal part is positive. Then, we can construct a quantum operation approximating the operation of a unitary operator , giving a block-encoding of defined as
(155)
in a runtime of with an allowed error
(156)
If we know that belongs to a subspace of spanned by , then .
This construction is realized by combining the algorithm presented in [8] with our algorithm simulating the negative time-evolution. The algorithm presented in [8] requires the use of for both positive and negative , but for only positive is used in our algorithm. Therefore our algorithm broadens the applicability of the quantum singular value transformation [41, 42] to the case where the classical description of the target operator is unknown, but is given as the dynamics of a Hamiltonian whose off-diagonal block is guaranteed to be given by .
D.0.2 Analysis of the runtime
Ref. [8] gives an algorithm which simulates from dynamics where is defined as
(159)
where is a linear operator satisfying ,
in an error smaller than using queries in total to . Moreover, can be fixed to . This is achieved by constructing a unitary
(164)
by performing the quantum singular value transformation using
(169)
and its dagger ,
where is a singular value decomposition of , and polynomial functions , are chosen in such a way that approximates and approximates both with an error smaller than or equal to for all (equivalently, approximates and approximates both with an error smaller than or equal to for all ). The measure of the approximation error is based on the error of approximating and is different from the measure used in Eq. (156), but it can be easily shown that the error in terms of Eq. (156) is bounded by
(172)
thus the total number of queries to for the case where the allowed error in terms of Eq. (156) is chosen as is also .
Suppose that we approximate using this algorithm for the case where the allowed error in terms of Eq. (156) is set as using and a quantum operation which approximates the quantum operation of using instead of preparing . For implementing , we choose an allowed error in terms of Eq. (33) to be .
In this situation, the overall procedure only requires as input. Because the error measured by Eq. (33) times two is larger than the error measured by Eq. (156) based on the fact the the diamond norm of an operation is greater than or equal to the operator norm for an arbitrary input density operator , the overall error of simulating in terms of Eq. (156) is upper bounded by
where the second term of the left hand side corresponds to the total error arising from the approximation error for calculated by (error of approximating ) (upper bound of the number of queries to ). Because can be constructed from by simulating with the negative time-evolution for with an allowed error in terms of Eq. (33) being chosen as in a runtime
(see the main text), the total runtime of constructing will be
(173)
calculated by (runtime of simulating ) (upper bound of the number of queries to ).
We have described the runtime analysis of constructing using so far. This procedure can be extended to the case where the input dynamics is instead of because can be constructed from and , and is further constructed from by negative time-evolution based on the equation
(174)
Since is simulated by a Class T transformation (see Appendix A), the above equation also provides the description of Class T transformations which transform to . Denoting this transformation as , the sum is calculated as which is . This technique is also introduced in [8]. Using this technique, quantum operations approximating with an allowed error can be constructed from also in time , thus can be constructed in runtime as well.
Appendix E Appendix E: Runtime and total evolution time analysis of Hamiltonian single-parameter learning
In this appendix, we present the runtime and total evolution time analysis of our algorithm for the Hamiltonian single-parameter learning task of efficiently estimating a single parameter from the dynamics of a -qubit unknown Hamiltonian satisfying with the Pauli vectors presented in the main text.
In the Hamiltonian transformation algorithm (Algorithm 1) applied for the Hamiltonian single-parameter learning task, we choose the transformation of the Hamiltonian given by
(175)
The set of linear operators is
characterized by defined as
E.0.1 Probability distribution obtained by Algorithm 1
We consider performing a projective measurement on the basis of or on the first qubit, namely, the qubit on which the operator appears in Eq. (175) of the state . The probability of obtaining the outcome for the projective measurement in the basis of is given by
(176)
and the probability of obtaining the outcome for the projective measurement in the basis of is given by
(177)
Suppose that we construct a quantum operation that approximates with an allowed error in terms of Eq. (33).
The runtime for simulating is for , while the total evolution time of will only be , as can be seen from the procedure of Algorithm 1.
In this case, the probability of obtaining the outcome in the basis measurement and in the basis measurement will be close by to values in Eq. (176) and Eq. (177), respectively.
This can be proved for the case (and similarly for the case) as
In the above equations, is an orthonormal basis of the Hilbert space to which belongs.
E.0.2 Learning using robust quantum phase estimation
Theorem I.1. in [38] can be rephrased to the following theorem.
Theorem 4.
Suppose that we can perform two families of measurements, -measurements and -measurements, where , whose success probabilities obtaining the outcome and outcome are given in terms of as
respectively, where and satisfy
Then for any allowed standard deviation , an estimate of can be obtained with a standard deviation smaller than or equal to by a classical computation with runtime of a function of the numbers of success of -measurements and -measurements both among times of the measurements. Here, and are defined as
(178)
We can perform -measurements and -measurements for for using
simulation of for and (or any positive number smaller than can be used instead).
The runtime of -measurements and -measurements in this case is bounded above using a constant independent of by , and the total evolution time of the dynamics is . Therefore, the total time of running the quantum circuit is bounded above as
while the total evolution time of in the overall experiment is
In the above calculation, we use an equation
which holds for an arbitrary .
The runtime of the classical calculation can be ignored in the evaluation of the total runtime of estimating , since it depends on as and tends to infinity more slowly than .
To calculate total evolution time in terms of the error upper bound and the maximum failure probability , we can modify the learning procedure to repeating estimation of with standard deviation smaller than or equal to for times and adopting the median of estimates. In this case, the total evolution time will be . The failure probability is shown to be smaller than or equal to using the fact that the probability that an estimate of with standard deviation satisfies is strictly smaller than .
E.0.3 Comparison with the Hamiltonian simulation method in [32]
Recently, another Hamiltonian learning technique achieving the Heisenberg limit for precision scaling in parameter estimation of a low-interaction Hamiltonian has been proposed [32]. Given an -qubit Hamiltonian for a known set of vectors satisfying the low-interaction Hamiltonian condition, this technique estimates all values of in a single run with the total evolution time where is precision and is the failure probability.
Our method can learn any single parameter of a general Hamiltonian for with total evolution time , which also achieves the Heisenberg-limited precision scaling.
Note that the total evolution time has no dependence on or more generally on .
In this sense, we give a partial answer to the open problem proposed by [32] regarding the learning of Hamiltonians without any structure. We still note that in terms of the total runtime instead of the evolution time, our algorithm is only efficient when is small.
For a low-interaction Hamiltonian, the estimation of all ’s using our algorithm requires to run our algorithm for every parameter. Therefore, the total evolution time is (), which is longer than that of the algorithm in [32] by a polynomial factor (the number of parameters of the -qubit low-interaction Hamiltonian). However, for estimation of a single parameter for representing the high-interaction Hamiltonian, namely, the term consisting of multiplications of non-identity Pauli operators, the algorithm in [32] requires a total evolution time exponential in , which leads to an exponential overall runtime, as can be seen from section C.2 of [32] (the algorithm in [32] can be applied to a general -qubit Hamiltonian and restricted to estimation of only a single parameter). On the other hand, our algorithm requires a constant total evolution time , and a runtime independent of .
For these reasons, our algorithm is suited to estimate a single parameter of non-local Hamiltonians. With our method, a parameter of an experimentally simulated Hamiltonian with a not-too-large norm, which does not necessarily have a simple structure, becomes obtainable.
Finally, we note that our algorithm also runs in a shorter runtime than the methods based on unitary tomography for the estimation of a single parameter. Recently, a unitary tomography method for estimating only a small number of entries of a unitary operation in a short time has been proposed [43]. However, this is not equivalent to the estimation of a small number of entries of a
Hamiltonian. In order to obtain the value of of a Hamiltonian by tomography of the unitary evolution , the full-tomography of is required, which requires a runtime in [43].