Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Single-ancilla ground state preparation via Lindbladians

Zhiyan Ding,1 Chi-Fang (Anthony) Chen,2 and Lin Lin1, 3, 4, ∗


1
Department of Mathematics, University of California, Berkeley, CA 94720, USA
2
Institute for Quantum Information and Matter, California Institute of Technology, Pasadena, CA 91125, USA
3
Applied Mathematics and Computational Research Division,
Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
4
Challenge Institute of Quantum Computation, University of California, Berkeley, CA 94720, USA
We design an early fault-tolerant quantum algorithm for ground state preparation. As a Monte
Carlo-style quantum algorithm, our method features a Lindbladian where the target state is sta-
tionary, and its evolution can be efficiently implemented using just one ancilla qubit. Our algorithm
can prepare the ground state even when the initial state has zero overlap with the ground state,
bypassing the most significant limitation of methods like quantum phase estimation. As a variant,
arXiv:2308.15676v1 [quant-ph] 30 Aug 2023

we also propose a discrete-time algorithm, which demonstrates even better efficiency, providing a
near-optimal simulation cost for the simulation time and precision. Numerical simulation using Ising
models and Hubbard models demonstrates the efficacy and applicability of our method.

Introduction— A promising application of quantum low-energy state. Nonetheless, these methods often re-
computers is to simulate ground state properties of quan- quire a large bath and lack the theoretical framework
tum many-body systems [3, 14, 22, 25, 33, 43]. To con- to rigorously study their efficacy. Guided by thermody-
cretely evaluate the end-to-end algorithmic cost, how- namic intuition, a new wave of Monte-Carlo style quan-
ever, the state preparation problem rises as a major con- tum algorithm has been proposed to extract the precise
ceptual bottleneck [23, 32]. working principle behind cooling [5, 7, 11, 38, 40, 44].
From a complexity theory standpoint, few-body The idea is to algorithmically emulate a bath by design-
Hamiltonian ground states can be QMA-hard to pre- ing a Lindbladian L whose fixed point is provably the
pare [1, 2, 20, 21, 31]. Thus, we do not expect quan- Gibbs state ρβ = e−βH / Tr[e−βH ]. That is, we evolve
tum computers to efficiently prepare ground states for ev- the density operator by
ery few-body Hamiltonian. While this worst-case hard-
ness may not apply to practically relevant systems, it dρ
= L[ρ] where L[ρβ ] ≈ 0. (1)
explains the theoretical obstacles towards proving algo- dt
rithmic guarantees. Indeed, justifying the efficacy of
This quantum algorithm can be regarded as a cousin
most existing ground-state algorithms requires additional
of the classical Markov chain Monte-Carlo (MCMC)
assumptions. The most common and transparent as-
method. As long as we prescribe the fixed point to be
sumption is the existence of an easy-to-prepare (pure or
our target state, the algorithmic cost per sample is
mixed) quantum state ρ0 with a good overlap with the
ground state |ψ0 ⟩, i.e., p0 = ⟨ψ0 |ρ0 |ψ0 ⟩ | = Ω(1/poly(n)) (simulation cost per unit time) × (mixing time). (2)
where n is the system size. This assumption allows
us to provably solve ground state preparation problems Thus, the hardness of low-energy problems, in this frame-
with a cost scaling with poly(1/p0 ) (among other depen- work, reduces to the mixing time of certain “quantum
dences). For instance, the cost associated with quantum Markov chains”. The mixing time can be very much case-
phase estimation (QPE) scales as O(1/p20 ) or O(1/p0 ) dependent for classical Markov chains; we expect similar
depending on the implementation [21, 26, 30], while rich behavior to transfer to the quantum case, and in
nearly optimal post-QPE algorithms exhibit a scaling of particular, we do not expect this algorithm to efficiently

O(1/ p0 ) [13, 25]. Unfortunately, practically running solve QMA-hard problems (just as we do not expect clas-
this algorithm demands an instance-dependent choice of sical Monte Carlo to solve NP-hard problems efficiently).
good trial states and can be nontrivial to justify in some While the efficacy of the QPE and the Monte Carlo ap-
systems of interest [23]. proaches both rely on additional assumptions, a key dif-
From a thermodynamics standpoint, however, ground ference is that the latter can be implemented without a
state (and more generally, low-energy states) should case-by-case recipe for the ansatz state. Furthermore,
be easy to prepare: just put the sample into a low- based on the experimental success of cooling, we might
temperature fridge. Drawing from this intuition, several expect the mixing time for physically relevant systems to
studies [4, 19, 28, 29, 34, 36] explore the idea of creat- be reasonably short, i.e., polynomial with system size.
ing an appropriate cold bath and establishing an effective
P
For a Hamiltonian H = i λ i |ψi ⟩ ⟨ψi | with spectral
system-bath coupling to efficiently cool the system to a gap ∆ = λ1 −λ0 > 0, the ground state is very close to the
Gibbs state ρβ with an inverse temperature β = Ω(∆−1 ).
However, Gibbs state preparation requires imposing the
quantum detailed balance condition, a delicate balance of
∗ linlin@math.berkeley.edu the rates between the heating and cooling transitions. To
2

accommodate specific energy measurements, energy la- Correctness— The right-hand side of (3) consists of
bel storage, and certain coherent arithmetic operations, two terms: The coherent part features a Hamiltonian
the algorithm uses techniques such as Quantum Fourier [H, ρ] generating unitary dynamics; the dissipative part
Transform and controlled Hamiltonian simulation, along LK is parameterized by a jump operator K. Since the
with other complex control logic. [7, 35, 40, 44]. Us- coherent part trivially fixes the ground state, it suffices
ing all these techniques might seem excessive, considering to focus on the dissipative part. The fixed point prop-
that the ground state is simply the lowest energy state erty remains valid with or without the coherent part1 .
where the detailed balance condition becomes singular In the frequency domain, K is expressed as the opera-
(i.e., all heating processes are forbidden). tor A in the energy basis weighted by the filter function
The approach involving Gibbs state preparation fˆ(ω) that depends on the energy difference ω = λi − λj .
presents another algorithmic challenge. While simulating This function should be real, non-negative, and satisfy
the Lindblad dynamics up to time T may achieve a com- the following condition (detailed assumptions are given
plexity of O(T log(T /ϵ)/ log log(T /ϵ)) [10, 24] – achiev- in Assumption 7 in Supplemental Materials):
ing near-optimal complexity in T and ϵ – this method
relies heavily on intricate block encodings and precise fˆ(ω) = 0 for each ω ≥ 0. (6)
control circuits. As a result, the near optimal algorithms
are only suitable on fully fault-tolerant quantum com- Under Eq. (6), we obtain that ⟨ψi |K|ψj ⟩ = 0 when
puters. For early fault-tolerant quantum computers with λi ≥ λj . Intuitively, this implies the jump operator K
constrained quantum resources, such as a limited num- forbids energy increments and favors a decrease in energy.
ber of logical qubits, a significant simplification would be Remarkably, such a simple condition guarantees that the
essential prior to the algorithm’s deployment. The above ground state is an exact fixed point of the Lindbladian.
line of thought leads to our guiding question: To see this, we first notice that
Can we devise a Monte-Carlo style quantum algorithm
for ground states in the early fault-tolerant regime while
X
K |ψ0 ⟩ = fˆ(λi − λ0 ) |ψi ⟩ ⟨ψi |A|ψ0 ⟩ = 0 , (7)
maintaining near-optimal complexity in the runtime T i
and precision ϵ?
Main result— In this work, we develop algorithms which implies LK [|ψ0 ⟩ ⟨ψ0 | = 0. Together with
that satisfy the following salient features: (1) The ground LH [|ψ0 ⟩ ⟨ψ0 |] = 0, we conclude that |ψ0 ⟩ ⟨ψ0 | is a fixed
state is a fixed point of a Lindblad evolution defined by point of the Lindblad dynamics (3). That is,
a single jump operator. (2) The algorithmic cost per
ground state sample scales with the mixing time and may e(LH +LK )t [|ψ0 ⟩ ⟨ψ0 |] = |ψ0 ⟩ ⟨ψ0 | , t ≥ 0. (8)
not depend on the initial state. Thus, even with zero ini-
tial overlap p0 = 0, the algorithm still converges to the Of course, there might exist other fixed points (i.e., the
ground state after the mixing time. (3) The continuous- map is not ergodic), and the proper choice of A that
time Lindblad evolution can be simulated using one an- ensures ergodicity should be case-dependent.
cilla qubit and minimal control logic. (4) With a discrete- In a simplified scenario where A is represented by a
time reformulation of the Lindblad evolution, the cost random matrix with independent entries (written in the
attains near-optimal complexity in both the runtime and energy eigenbasis), we give a partial argument for ergod-
precision. icity.
The main Lindbladian is defined as follows
Theorem 1 (Random coupling matrix and ergodicity,
d 1 informal). Let ρ(t) be the solution to the Lindblad dy-
ρ = L[ρ] = −i[H, ρ] + KρK † − {K † K, ρ} , (3)
dt | {z } | 2
{z } namics (3) and Ai,j = ⟨ψi | A |ψj ⟩. Assume for any t ≥ 0,
=:LH [ρ] A is independently drawn from a probability distribution
=:LK [ρ]
on the set of Hermitian matrices2 such that E(Ai,j ) = 0
with the only jump operator and ρ(0) is a diagonal matrix in the basis of |ψj ⟩, then
X ρ⋆ = |ψ0 ⟩ ⟨ψ0 | is the unique fixed point of the Lindblad
K := fˆ(λi − λj ) |ψi ⟩ ⟨ψi |A|ψj ⟩ ⟨ψj | (4) dynamics in the expectation sense. In particular, given
i,j∈[N ] any observable O, limt→∞ E(Tr(Oρ(t))) = ⟨ψ0 | O |ψ0 ⟩,
Z ∞
where the expectation is taken on the randomness of A.
= f (s)A(s) ds. (time domain) (5)
−∞

Here A is a simple operator that acts on the system


with its Heisenberg evolution A(s) = eiHs Ae−iHs . The 1 Nevertheless, our numerical findings suggest that incorporating
1
fˆ(ω)e−iωs dω is the
R
time-domain function f (s) := 2π R the coherent part often leads to a significant reduction in the
inverse Fourier transform of the filter function. In the mixing time.
2 To accurately represent its dependence on time, A should be de-
following, we elaborate on the desirable features of the
noted as At . However, for the sake of simplicity and consistency
proposed algorithm. with other notations, we will omit the subscript t in this theorem.
3

Figure 1. Left: Sketch of the circuit structure for one iteration of the main algorithm. The circuit utilizes only one ancilla qubit,
and the detailed derivation of this circuit is given in Appendix B. In this graph, C1 , · · · , CM are local interaction operators
that only act on one system qubit (or a small number of system qubits). Right: a visual representation of how the dissipative
part exp(LK τ ) transforms the density matrix in each iteration towards the ground state.
.

See the complete result in the Supplemental Material simulating the continuous Lindblad dynamics Eq. (3) us-
(Theorem 20). Strictly speaking, the expected opera- ing one ancilla qubit. For simulation time T and preci-
tor E(ρ(t)) is not the density operator ρ(t) we store in sion ϵ, the total cost in terms of Hamiltonian simulation
e 1 + ∥H∥)∆−1 T 2+o(1) ϵ−1−o(1) ).

the quantum memory but still gives some optimistic in- time is TH,total = Θ(
tuition regarding ergodicity (see Supplemental Section
D.2). Technically, taking expectations over independent A more comprehensive version of the above theorem
entries of A substantially simplifies the transition matrix. (Theorem 13), including the cost based on the number
This independent assumption can also be seen as a ver- of controlled-A gate queries, can be found in the Supple-
sion of the Eigenstate Thermalization Hypothesis (ETH) mental material (Appendix C.1). The cost also depends
[12, 39], which incorporates additional assumptions re- on the specific form fˆ, f that we choose. More specif-
garding σi,j . Indeed, ETH has been employed to explain ically, when the function f have bounded N -th deriva-
finite-time thermalization in chaotic open quantum sys- tives, the term o(1) = N −1 , which diminishes as the
tems [5, 38]. Under stronger ETH-typed assumptions, smoothness of f increases (as N approaches infinity).
one may be able to prove convergence for ρ(t) instead of Additionally, the constant in
TH,total also depends on the
E(ρ(t)) as in [5], but we merely focus on the much sim- Nˆ N
upper bound of d f /dω .
pler object E(ρ(t)) without distracting the presentation We expect the most resource-intensive part of the algo-
of the algorithm. rithm to be Hamiltonian simulation. Thus, we quantify
Convergence— The convergence rate of a Markov the cost through the total Hamiltonian simulation time,
chain process is characterized by the mixing time, which represented as TH,total . For an end-to-end cost analysis,
describes the time scale at which any two input states one may further Trotterize the Hamiltonian simulation
become close to each other (see detail in Appendix A). subroutine and analyze its discretization error, and there
Analogous to classical Monte Carlo sampling, we do is no hidden block-encoding or extra ancilla involved (i.e.,
not know a priori the mixing time associated with the no controlled Hamiltonian simulation).
jump operator K and the Hamiltonian H; we expect this The complexity of our algorithm approximately scales
to be system dependent. In fact, proving rapid mixing quadratically with the simulation time T (a reasonable
for Lindbladians is widely acknowledged as a highly chal- choice is the mixing time tmix ) and inversely with the
lenging problem, and it is not the primary objective of error ϵ. This scaling is reminiscent of that of the first-
our work. However, as potential evidence supporting the order product formula, which has a second-order error
convergence of our method (in Appendix D), we show
exp((LH + LK )τ )) = exp(LH τ ) exp(LK τ ) + O(τ 2 ). (9)
that under proper assumptions on A, the Lindblad dy-
namics discussed in our study may exhibit a reasonably We emphasize that the Lindbladian simulation qualita-
fast mixing time. tively differs from the Hamiltonian simulation: even to
Simulating continuous-time dynamics— A main ad- arrive at the ∼ T 2 /ϵ scaling requires a nontrivial argu-
vantage of our algorithm is the use of merely one ancilla ment to approximately implement exp(LK τ ).
qubit and minimal control logic. Indeed, in the context of Simulating discrete-time dynamics— Is it possible to
Hamiltonian simulation, we know that product formulas reduce the “first-order” scaling T 2 /ϵ of the cost? For
(or Trotterization) have these desirable features. How- Hamiltonian simulation we can use higher-order formu-
ever, the case of Lindbladian simulation has been less las, but here we cannot invert dissipation (i.e., simulating
studied, which we address as follows. exp(−LK τ )), which leads to fundamental constraints on
the permissible product formula sequences. A shift in
Theorem 2 (Single ancilla simulation of continuous-time perspective is necessary. What if we interpret
Lindblad dynamics, informal). For any Hamiltonian H
and simple operator A, there exists a quantum algorithm ρn+1 = exp(LH τ ) exp(LK τ )ρn = Nτ (ρn ) (10)
4

not as an approximation for the continuous dynamics ex- transverse field Ising model (TFIM) model defined on
pressed by Eq. (3), but by itself a discrete-time dynam- 6 sites (see detailed setup, including a concrete example
ics? Indeed, our primary goal is not the precise simula- of fˆ and f , and additional numerical results on TFIM
tion of Lindbladian dynamics, but rather, to prepare the and Hubbard models, in Appendix E). In our numeri-
ground state. A key observation is that the ground state cal simulations, we set τ = 1 and r = 2 for discrete-time
is a stationary state of the discretized dynamics for any Lindblad simulation (r is the number of segments that we
value of τ . Unlike the continuous dynamics simulation, use to approximately simulate exp(LK τ ) in the discrete-
which requires τ = O(ϵ/ ∥H∥ T ), we could now poten- time Lindblad simulation algorithm, see Appendix B.5),
tially choose τ as any arbitrary value. In such a scenario, while for continuous-time Lindblad simulation, we use
Eq. (10) might not approximate the continuous Lindblad τ = 0.1. We start with an initial state with zero overlap
dynamics due to the discretization error, yet it still fixes (⟨ψ0 | ρ0 |ψ0 ⟩ ≈ 10−17 ), repeat each Lindblad dynamics
the ground state. Simulating this discrete-time dynam- simulation 100 times, and compute the average energy
ics for M discrete steps corresponds to a simulation for and overlap with the ground state. The results are de-
an effective time T = M τ . By choosing τ = O(1), we picted in Figure 2. Based on our results, Lindblad dy-
obtain a scheme with significantly improved scaling with namics demonstrates a high level of effectiveness in re-
respect to both the simulation time T and precision ϵ. ducing energy to the ground state and generating a state
Remarkably, the total Hamiltonian simulation time of the that closely aligns with the ground state. The trajecto-
discrete-time dynamics no longer depends on ∆/ ∥H∥ as ries of the continuous and discrete-time dynamics notice-
in Theorem 2, but only on ∆. This can be significantly ably differ from each other, yet they can both prepare
faster when simulating large-scale systems, or Hamiltoni- the ground state, and the mixing times are compara-
ans involving unbounded operators (such as differential ble. Additionally, the discrete-time Lindblad dynamics
operators in first quantization). (τ = 1, r = 2) demands almost one-tenth of the total
Hamiltonian simulation time required by the approxi-
Theorem 3 (Discrete-time algorithm for ground state,
mated continuous-time Lindblad dynamics (τ = 0.1).
informal). For any Hamiltonian H and simple oper-
Key ideas— The design of our specific Monte Carlo
ator A, there exists a quantum algorithm simulating
style quantum algorithms draws inspiration from the
the continuous Lindblad dynamics Eq. (10) using one
recent work for preparing thermal states [7], which
ancilla qubit. For simulation time T and precision
⋆ considers a set of jump operators labeled by a, ω as
ϵ, the total Hamiltonian simulation time is TH,total = p
{ γ(ω)Aa (ω)}a,ω . Intuitively, {Aa (ω)}a,ω corresponds
−1 1+o(1) −o(1)
Θ(∆ T
e ϵ ). to the transitions of Aa with an energy difference of ap-
proximately ω. The weights γ(ω) depend on temperature
See the Supplemental Materials (Corollary 15) for
and are designed to satisfy the approximate quantum de-
the formal statement. Compared with the continuous-
tailed balance condition. A key observation of this work
time Lindblad simulation, this algorithm exhibits a total
is that the “detailed balance condition” for ground state
Hamiltonian simulation time that scales nearly linearly
preparation can be achieved using merely a single jump
with T and sub-polynomially with 1/ϵ.3 In Theorem 3,
operator.
the appearance of o(1) stems from adopting a high-order
To reduce the end-to-end ancilla usage, we first define
Trotter formulation of order p, which contributes an ex-
a dilation K̃ := |1⟩ ⟨0| ⊗ K + |0⟩ ⟨1| ⊗ K † , and use the
ponent of the form 1/p = o(1) as p → ∞. Similar to the
fact
Hamiltonian simulation problem, adopting a high-order √ √
Trotter scheme results in asymptotic improvement, but eLK t [ρ] ≈ Tra e−iK t
[|0⟩ ⟨0| ⊗ ρ] eiK t
+ O(t2 ). (11)
e e

this also comes with a significant constant overhead [8].


The algorithm in Theorem 3 approximates the Here Tra traces out the ancilla qubit. Since Ke is a Hermi-
discrete-time Lindblad dynamics (10). Theoretically, the tian matrix, the above formulation reduces the Lindbla-
effective “mixing time” t′mix = Mmix τ of the discrete dy- dian simulation to the√dilated Hamiltonian simulation.
To implement e−iK t , we observe that K is expressed
e
namics may not be the same as the continuous-time Lind-
blad dynamics tmix . On the other hand, numerical obser- in integral form, and its integrand also involves Hamilto-
vations indicate that the mixing time of both dynamics nian simulation. Consequently, the direct discretization
can be comparable, but the discrete-time version requires and Trotterization of the right-hand side of (5) pose chal-
significantly less total Hamiltonian simulation time and lenges since the exponent in each Trotter step contains
therefore reduces the total cost. another Hamiltonian simulation e−iHs .
Numerics results— We numerically illustrate the per- The technical challenges rise due to the currently for-
formance of our algorithms using the one-dimensional bidding cost of implementing general block encodings and
a limited number of control ancillas. This is analogous
to the story between the two very different algorithms
for Hamiltonian simulation: linear combinations of uni-
3 While adopting a high-order Trotter formula does not reduce the tary (LCU) [9] / quantum singular value transformation
cost of continuous-time simulation, it does benefit the discrete- (QSVT) [16], and product formulas [8, 42]. The former
time dynamics. is theoretically transparent and helps us understand the
5

0 Lindblad ( = 0.1) 1.0 Lindblad ( = 0.1) 1.0


Lindblad ( = 1, r = 2) Lindblad ( = 1, r = 2)
0 0.8 0.8
2
0.6 0.6

< p0 >

< p0 >
4
<E>

0.4 0.4
6
0.2 0.2 Lindblad (exact)
Lindblad ( = 0.1)
8 0.0 0.0 Lindblad ( = 1, r = 2)
101 102 103 104 101 102 103 104 0 25 50 75 100 125 150 175 200
total Hamiltonian simulation time total Hamiltonian simulation time time

(a) Hamiltonian simulation time vs energy (b) Hamiltonian simulation time vs overlap (c) Lindblad simulation time vs overlap

Figure 2. Performance of continuous versus discrete-time Lindblad dynamics for preparing the ground state for the TFIM with
6 sites. The continuous-time simulation uses a small time step τ = 0.1. The trajectory of the discrete-time dynamics (using a
large time step τ = 1) deviates from that of the continuous-time Lindblad dynamics, but it successfully prepares the ground
state and is more efficient than the continuous-time dynamics. Here the Hamiltonian simulation time refers to the sum of the
Hamiltonian simulation t in all e±iHt subroutines used in the circuit.

asymptotic complexity of the Hamiltonian simulation, tonian using a single ancilla qubit. Merely a single jump
but practitioners often turn to the latter because of its operator suffices to ensure that the system’s ground state
simple implementation. is a fixed point of the Lindbaldian dynamics. Unlike
Our method to overcome this difficulty is motivated by many ground-state preparation algorithms, such as QPE,
the following observation. When τ is sufficiently small, the performance of this algorithm is mainly determined
RS RS by the mixing time and may work even with zero ini-
eiτ 0
f (s)A(s) ds
= T ei 0
τ f (s)A(s) ds
+ O(τ 2 ) tial overlap. Similar to classical Monte Carlo methods,
RS (12) the relationship between mixing time and system size,
=eiHS T ei 0
(−H+τ f (s)A) ds
+ O(τ 2 ). the practical selection of the coupling matrix A, and the
optimal choice of the filtering function f may be system-
The first line introduces a time-ordered integral. More
Rt dependent. For instance, optimizing these choices to pre-
precisely, we use u(t) = T ei 0 τ f (s)A(s) ds to denote the pare the ground state of strongly correlated chemical sys-
formal solution to the differential equation du/ dt = tems like FeMoCo [23], which are currently inaccessible
iτ f (t)A(t)u(t); this can be also understood as the con- via adiabatic state preparation, might offer a compelling
tinuum limit of first-order Trotter formula. We recognize avenue of research.
the time-ordered exponential as the interaction picture
over Hamiltonian H, which leads to the second expres-
sion. Even in cases where f (s) is not real, the applica- Acknowledgments— This material is based upon
tion of the dilatation and trace-out technique allows us work supported by the U.S. Department of Energy, Of-
to transform f (s)A − H into a time-dependent Hamil- fice of Science, National Quantum Information Science
tonian simulation. Then, we can further discretize and Research Centers, Quantum Systems Accelerator (Z.D.).
Trotterize it into simple gates and only one ancilla qubit Additional support is acknowledged from the Challenge
is needed. More carefully, our actual implementation uti- Institute for Quantum Computation (CIQC) funded by
lizes a second-order Trotter formula (not the first-order National Science Foundation (NSF) through grant num-
formula in (12)), which ensures that the simulation error ber OMA-2016245, the Applied Mathematics Program of
√ the US Department of Energy (DOE) Office of Advanced
of exp(−iK e t) is of the order O(t2 ) rather than O(t).
Scientific Computing Research under contract number
For a more comprehensive description of the implemen- DE-AC02-05CH1123, and a Google Quantum Research
tation, see Appendix B.4. Award (L.L.). L.L. is a Simons investigator. We thank
Discussion— This paper presents a Monte Carlo ap- Garnet Chan, Toby Cubitt, Yulong Dong, Xiantao Li,
proach for ground state preparation of a quantum Hamil- Subhayan Roy Moulik, Yu Tong for helpful discussions.

[1] D. Aharonov, D. Gottesman, S. Irani, and J. Kempe. [2] D. Aharonov and T. Naveh. Quantum NP-a survey.
The power of quantum systems on a line. Comm. Math. arXiv preprint quant-ph/0210077, 2002.
Phys., 287(1):41–65, 2009. [3] A. Aspuru-Guzik, A. D. Dutoi, P. J. Love, and M. Head-
Gordon. Simulated quantum computation of molecular
6

energies. Science, 309(5741):1704–1707, 2005. [22] B. P. Lanyon, J. D. Whitfield, G. G. Gillett, M. E.


[4] M. Cattaneo, M. A. Rossi, G. Garcı́a-Pérez, R. Zam- Goggin, M. P. Almeida, I. Kassal, J. D. Biamonte,
brini, and S. Maniscalco. Quantum simulation of dis- M. Mohseni, B. J. Powell, M. Barbieri, A. Aspuru-Guzik,
sipative collective effects on noisy quantum computers. and A. G. White. Towards quantum chemistry on a quan-
PRX Quantum, 4:010324, 2023. tum computer. Nature Chemistry, 2(2):106–111, 2010.
[5] C.-F. Chen and F. G. S. L. Brandão. Fast thermal- [23] S. Lee, J. Lee, H. Zhai, Y. Tong, A. M. Dalzell, A. Ku-
ization from the eigenstate thermalization hypothesis. mar, P. Helms, J. Gray, Z.-H. Cui, W. Liu, et al. Evalu-
arxiv/2112.07646, 2023. ating the evidence for exponential quantum advantage
[6] C.-F. Chen, H.-Y. Huang, R. Kueng, and J. A. Tropp. in ground-state quantum chemistry. Nature Comm.,
Concentration for random product formulas. PRX Quan- 14(1):1952, 2023.
tum, 2:040305, 2021. [24] X. Li and C. Wang. Simulating Markovian open
[7] C.-F. Chen, M. J. Kastoryano, F. G. S. L. Brandão, quantum systems using higher-order series expansion.
and A. Gilyén. Quantum thermal state preparation. arXiv/2212.02051, 2022.
arXiv/2303.18224, 2023. [25] L. Lin and Y. Tong. Near-optimal ground state prepara-
[8] A. M. Childs, Y. Su, M. C. Tran, N. Wiebe, and S. Zhu. tion. Quantum, 4:372, 2020.
Theory of trotter error with commutator scaling. Phys. [26] L. Lin and Y. Tong. Heisenberg-limited ground state
Rev. X, 11(1):011020, 2021. energy estimation for early fault-tolerant quantum com-
[9] A. M. Childs and N. Wiebe. Hamiltonian simula- puters. PRX Quantum, 3:010318, 2022.
tion using linear combinations of unitary operations. [27] G. Lindblad. On the generators of quantum dynamical
arXiv/1202.5822, 2012. semigroups. Communications in Mathematical Physics,
[10] R. Cleve and C. Wang. Efficient quantum algorithms 48(2):119 – 130, 1976.
for simulating Lindblad evolution. In 44th International [28] M. Metcalf, E. Stone, K. Klymko, A. F. Kemper,
Colloquium on Automata, Languages, and Programming, M. Sarovar, and W. A. de Jong. Quantum Markov
ICALP 2017, July 2017. chain Monte Carlo with digital dissipative dynamics on
[11] T. S. Cubitt. Dissipative ground state preparation and quantum computers. Quantum Science and Technology,
the dissipative quantum eigensolver. arXiv/2303.11962, 7(2):025017, 2022.
2023. [29] X. Mi, A. A. Michailidis, S. Shabani, K. C. Miao, P. V.
[12] L. D’Alessio, Y. Kafri, A. Polkovnikov, and M. Rigol. Klimov, J. Lloyd, E. Rosenberg, R. Acharya, I. Aleiner,
From quantum chaos and eigenstate thermalization to T. I. Andersen, M. Ansmann, F. Arute, K. Arya, A. As-
statistical mechanics and thermodynamics. Advances in faw, J. Atalaya, J. C. Bardin, A. Bengtsson, G. Bortoli,
Physics, 65(3):239–362, 2016. A. Bourassa, J. Bovaird, L. Brill, M. Broughton, B. B.
[13] Y. Dong, L. Lin, and Y. Tong. Ground-state prepara- Buckley, D. A. Buell, T. Burger, B. Burkett, N. Bush-
tion and energy estimation on early fault-tolerant quan- nell, Z. Chen, B. Chiaro, D. Chik, C. Chou, J. Co-
tum computers via quantum eigenvalue transformation gan, R. Collins, P. Conner, W. Courtney, A. L. Crook,
of unitary matrices. PRX Quantum, 3:040305, 2022. B. Curtin, A. G. Dau, D. M. Debroy, A. D. T. Barba,
[14] Y. Ge, J. Tura, and J. I. Cirac. Faster ground state S. Demura, A. D. Paolo, I. K. Drozdov, A. Dunsworth,
preparation and high-precision ground energy estima- C. Erickson, L. Faoro, E. Farhi, R. Fatemi, V. S. Fer-
tion with fewer qubits. Journal of Mathematical Physics, reira, L. F. B. E. Forati, A. G. Fowler, B. Foxen,
60(2):022202, 2019. E. Genois, W. Giang, C. Gidney, D. Gilboa, M. Giustina,
[15] M. Gevrey. Sur la nature analytique des solutions des R. Gosula, J. A. Gross, S. Habegger, M. C. Hamilton,
équations aux dérivées partielles. Premier mémoire. An- M. Hansen, M. P. Harrigan, S. D. Harrington, P. Heu,
nales scientifiques de l’École Normale Supérieure, 3e M. R. Hoffmann, S. Hong, T. Huang, A. Huff, W. J.
série, 35:129–190, 1918. Huggins, L. B. Ioffe, S. V. Isakov, J. Iveland, E. Jeffrey,
[16] A. Gilyén, Y. Su, G. H. Low, and N. Wiebe. Quantum Z. Jiang, C. Jones, P. Juhas, D. Kafri, K. Kechedzhi,
singular value transformation and beyond: exponential T. Khattar, M. Khezri, M. Kieferova, S. Kim, A. Kitaev,
improvements for quantum matrix arithmetics. In Pro- A. R. Klots, A. N. Korotkov, F. Kostritsa, J. M. Kreike-
ceedings of the 51st Annual ACM SIGACT Symposium baum, D. Landhuis, P. Laptev, K. M. Lau, L. Laws,
on Theory of Computing, pages 193–204, 2019. J. Lee, K. W. Lee, Y. D. Lensky, B. J. Lester, A. T.
[17] V. Gorini, A. Kossakowski, and E. C. G. Sudarshan. Lill, W. Liu, A. Locharla, F. D. Malone, O. Martin,
Completely positive dynamical semigroups of N-level sys- J. R. McClean, M. McEwen, A. Mieszala, S. Montazeri,
tems. Journal of Mathematical Physics, 17(5):821–825, A. Morvan, R. Movassagh, W. Mruczkiewicz, M. Neeley,
2008. C. Neill, A. Nersisyan, M. Newman, J. H. Ng, A. Nguyen,
[18] L. Hormander. Analysis of linear partial differential op- M. Nguyen, M. Y. Niu, T. E. OBrien, A. Opremcak,
erators I: distribuction theory and fourier analysis. New A. Petukhov, R. Potter, L. P. Pryadko, C. Quintana,
York: Springer-Verlag, 1990. C. Rocque, N. C. Rubin, N. Saei, D. Sank, K. Sankarago-
[19] D. B. Kaplan, N. Klco, and A. Roggero. Ground mathi, K. J. Satzinger, H. F. Schurkus, C. Schuster, M. J.
states via spectral combing on a quantum computer. Shearn, A. Shorter, N. Shutty, V. Shvarts, J. Skruzny,
arxiv/1709.08250, 2017. W. C. Smith, R. Somma, G. Sterling, D. Strain, M. Sza-
[20] J. Kempe, A. Kitaev, and O. Regev. The complexity lay, A. Torres, G. Vidal, B. Villalonga, C. V. Heid-
of the local Hamiltonian problem. SIAM J. Comput., weiller, T. White, B. W. K. Woo, C. Xing, Z. J. Yao,
35(5):1070–1097, 2006. P. Yeh, J. Yoo, G. Young, A. Zalcman, Y. Zhang, N. Zhu,
[21] A. Y. Kitaev, A. Shen, and M. N. Vyalyi. Classical N. Zobrist, H. Neven, R. Babbush, D. Bacon, S. Boixo,
and quantum computation. American Mathematical Soc., J. Hilton, E. Lucero, A. Megrant, J. Kelly, Y. Chen,
2002. P. Roushan, V. Smelyanskiy, and D. A. Abanin. Sta-
ble quantum-correlated many body states via engineered
dissipation. arxiv/2304.13878, 2023.
[30] M. A. Nielsen and I. Chuang. Quantum computation and
quantum information. Cambridge Univ. Pr., 2000.
[31] R. Oliveira and B. M. Terhal. The complexity of quantum
spin systems on a two-dimensional square lattice. arXiv
preprint quant-ph/0504050, 2005.
[32] P. J. J. O’Malley, R. Babbush, I. D. Kivlichan,
J. Romero, J. R. McClean, R. Barends, J. Kelly,
P. Roushan, A. Tranter, N. Ding, B. Campbell, Y. Chen,
Z. Chen, B. Chiaro, A. Dunsworth, A. G. Fowler, E. Jef-
frey, E. Lucero, A. Megrant, J. Y. Mutus, M. Neeley,
C. Neill, C. Quintana, D. Sank, A. Vainsencher, J. Wen-
ner, T. C. White, P. V. Coveney, P. J. Love, H. Neven,
A. Aspuru-Guzik, and J. M. Martinis. Scalable quantum
simulation of molecular energies. Phys. Rev. X, 6:031007,
2016.
[33] T. E. O’Brien, B. Tarasinski, and B. M. Terhal. Quan-
tum phase estimation of multiple eigenvalues for small-
scale (noisy) experiments. New Journal of Physics,
21(2):023022, 2019.
[34] S. Polla, Y. Herasymenko, and T. E. O’Brien. Quantum
digital cooling. Phys. Rev. A, 104:012414, Jul 2021.
[35] P. Rall, C. Wang, and P. Wocjan. Thermal state prepa-
ration via rounding promises. arxiv/2210.01670, 2022.
[36] B. Rost, L. D. Re, N. Earnest, A. F. Kemper, B. Jones,
and J. K. Freericks. Demonstrating robust simulation of
driven-dissipative problems on near-term quantum com-
puters. arXiv/2108.01183, 2021.
[37] M. B. RUSKAI. Beyond strong subadditivity? improved
bounds on the contraction of generalized relative en-
tropy. Reviews in Mathematical Physics, 06(05a):1147–
1161, 1994.
[38] O. Shtanko and R. Movassagh. Preparing thermal states
on noiseless and noisy programmable quantum proces-
sors. arXiv/2112.14688, 2023.
[39] M. Srednicki. The approach to thermal equilibrium in
quantized chaotic systems. Journal of Physics A: Math-
ematical and General, 32(7):1163, 1999.
[40] K. Temme, T. J. Osborne, K. G. Vollbrecht, D. Poulin,
and F. Verstraete. Quantum Metropolis sampling. Na-
ture, 471(7336):87–90, 2011.
[41] L. N. Trefethen and J. A. C. Weideman. The expo-
nentially convergent trapezoidal rule. SIAM Review,
56(3):385–458, 2014.
[42] H. Trotter. On the product of semi-groups of operators.
Proc. Amer. Math. Soc., 10:545, 1959.
[43] L. Veis and J. Pittner. Quantum computing applied to
calculations of molecular energies: CH2 benchmark. The
Journal of Chemical Physics, 133(19):194106, 11 2010.
[44] M.-H. Yung and A. Aspuru-Guzik. A quantum–quantum
Metropolis algorithm. Proceedings of the National
Academy of Sciences, 109(3):754–759, 2012.
8

Supplemental Material

Appendix A: Notation, facts, and organization

In the supplementary material, we use capital letters for matrices and curly font for superoperators. Besides the
usual O notation, we use the following asymptotic notations: we write f = Ω(g) if g = O(f ); f = Θ(g) if f = O(g)
and g = O(f ); f = O(g)
e if f = O(g polylog(g)). We use ∥ · ∥ to denote vector or matrix 2-norm: when v is a vector
we denote by ∥v∥ its 2-norm, and when A is matrix we denote by ∥A∥ its operator norm (or the Schatten ∞-norm).
Given any Hermitian matrix H and a positive integer N , there exists a constant CN such that

N
X (−iH)n n
exp(−iHt) − t ≤ CN ∥H∥N +1 tN +1 . (A1)

n!
n=0

h√ i
The trace norm (or the Schatten 1-norm) of a matrix A is ∥A∥1 = Tr A† A . A useful inequality that we often
use in our proof is that

∥AB∥1 ≤ ∥A∥∥B∥1 . (A2)

for any two matrices A, B such that AB is well defined.


Given a superoperator L that acts on operators, the induced 1-norm is

∥L∥1 := sup ∥L(ρ)∥1 . (A3)


∥ρ∥1 ≤1

According to GKLS theorem [17, 27], if L is a Lindbladian, which has the form
M
X
† 1 †
L(ρ) = −i[H, ρ] + Km ρKm − {Km Km , ρ} ,
m=1
2

then exp(Lt) is a quantum channel, i.e., it is a completely positive trace-preserving (CPTP) map. It is also contractive
under trace distance [37]. For any two density operators ρ1 , ρ2 , and any t > 0,

∥ exp(Lt)ρ1 − exp(Lt)ρ2 ∥1 ≤ ∥ρ1 − ρ2 ∥1 . (A4)

The following lemma provides an upper bound on the first-order Trotter error between two superoperators:
Lemma 4. Given two superoperators L1 , L2 such that exp(L1 t), exp(L2 t), and exp((L1 + L2 )t) are quantum channels
for all t > 0. Then for all t > 0,

∥exp((L1 + L2 )t) − exp(L1 t) exp(L2 t)∥1 = O ∥[L1 , L2 ]∥1 t2 .



(A5)

Proof. This proof is similar to the proof of [8, Theorem 6]. Define

E(t) = exp(L1 t) exp(L2 t) .

Then, we obtain
dE(t)
= (L1 + L2 )E(t) + [exp(L1 t), L2 ] exp(L2 t) ,
dt
which implies
Z t
E(t) = exp((L1 + L2 )t) + exp((L1 + L2 )(t − τ ))[exp(L1 τ ), L2 ] exp(L2 τ ) dτ .
0

Because ∥[exp(L1 τ ), L2 ]∥1 = O ([L1 , L2 ]τ ) and ∥ exp((L1 + L2 )(t − τ ))∥1 = ∥ exp(L2 τ )∥1 = 1, we obtain that

∥E(t) − exp((L1 + L2 )t)∥1 = O [L1 , L2 ]t2 ,




which proves (A5).


9

Next, we bound the 1-norm of LK corresponding to the jump operator K in Eq. (5) of the Lindblad dynamics in
Eq. (3):
R∞
Lemma 5. Assume −∞ |f (t)| dt = O(1), then

∥LK ∥1 = O(∥K∥2 ) = O(∥A∥2 ) . (A6)

Proof. To prove (A6), we first use (A2) to obtain ∥LK ∥1 = O(∥K∥2 ). Then
Z ∞
∥K∥ ≤ ∥A∥ |f (s)| ds = O(∥A∥).
−∞

Finally, to characterize the convergence of the Lindblad dynamics, we define the mixing time as follows:

Definition 6 (Mixing time of Lindbladians and channels). We define the continuous and discrete mixing time as
follows.

• For any Lindbladian L, the mixing time tmix is the smallest time for which

e mix [ρ − ρ′ ] ≤ 1 ∥ρ − ρ′ ∥
Lt
ρ, ρ′ .

1 1 for any states
2

• For any quantum channel N , the “discrete” mixing time Mmix is the smallest integer M for which

N [ρ − ρ′ ] ≤ 1 ∥ρ − ρ′ ∥ ρ, ρ′ .
M
1 1 for any states
2

In particular, if the Lindblad L has mixing time tmix , any initial state must be ϵ close to the ground state after
time tmix · log2 (2/ϵ) since the trace distance between any two quantum states is at most 2.
The rest of the supplementary material is organized as follows:

• Section B introduces our algorithms for the approximate preparation of the ground state.

• Section C analyzes the costs associated with these algorithms.

• Section D discusses the convergence properties of the continuous-time Lindblad dynamics under the Eigenstate
Thermalization Hypothesis (ETH) type ansatz.

• Section E presents details of numerical results, including additional simulations with the TFIM and Hubbard
models.

Appendix B: The simulation algorithm

In this section, we first introduce the filter function needed to define the jump operator (Section B.1) and present our
algorithm for the approximate preparation of the ground state by simulating the continuous-time Lindblad dynamics
(Section B.2). Subsequently, we introduce the discrete-time Lindblad dynamics (Section B.5). These algorithms offer
the advantage of requiring only a single ancilla qubit and can be implemented without the need for constructing a
block encoding of Ks .

B.1. Choice of filter function

As mentioned earlier, ensuring that the ground state remains a fixed point of the Lindblad dynamics requires fˆ
to satisfy the condition stated in Eq. (6). Furthermore, as discussed
R∞ in Key ideas, in order to numerically simulate
exp(LK (t)), we need to discretize and truncate the integral −∞ f (s)A(s) ds. To ensure that this truncation does
not introduce significant errors, additional assumptions must be imposed on fˆ such that f (t) decays rapidly as |t|
approaches infinity. In summary, we adopt the following assumption regarding fˆ:
10

Assumption 7 (Filter function in the frequency domain). Given parameter Sω > 2∆, where ∆ = λ1 − λ0 . We
assume the filter function fˆ(ω) ∈ C∞ (R) satisfies

(
Ω(1) ω = −∆
0 ≤ fˆ(ω) ≤ 1 and fˆ(ω) = for all ω ∈ R. (B1)
0 ω∈/ [−Sω , 0]

• Given any N > 0, there exists a constant CN such that



dN fˆ(ω)
≤ CN ∆−N , for all ω ∈ R.

dω N

The choice of Sω significantly impacts the ease of discretizing the time integral (5) and the mixing time of the
Lindblad dynamics. Striking the right balance is crucial. Choosing a larger Sω facilitates the transition of weight
from high energy to low energy, leading to a reduction in the mixing time. However, it also introduces challenges in
discretizing the time integral (5), increasing the complexity of simulating the Lindblad dynamics. An example of such
filter function that satisfies Assumption 7 can be found in Section E, where we set Sω = O(∥H∥).
We present two results on the Fourier transform of fˆ(ω)
Z
1
f (s) := fˆ(ω)e−iωs dω (B2)
2π R
This is slightly different from the standard convention of the Fourier transform, but it agrees with the convention
used in Ref. [7].
Lemma 8. Let fˆ(ω) satisfy Assumption 7. Then f (s) in Eq. (B2) can be extended into an entire function denoted
by f (z). Given N > 0,
|f (s)| = O CN Sω ∆−N (1 + |s|)−N , for all s ∈ R,

(B3)

Z ∞
|f (s + iζ) exp(iω(s + iζ))| ds = O CN Sω ∆−N exp(2∥H∥ + Sω ) .

sup (B4)
|ω|≤2∥H∥, |ζ|≤1 ∞

Here CN is the constant from Assumption 7.


Proof. According to the Paley–Wiener–Schwartz theorem [18, Theorem 7.3.1], f (s) can be extended to become an
entire function in the complex plane. First, when |z| ≤ 1, we have
Z 0 Z 0
1 1
ˆ −iωz ˆ
f (ω)e−iωz dω = O (CN Sω exp(Sω |Im(z)|)) .

|f (z)| = f (ω)e dω ≤

2π −Sω 2π −Sω
When |z| > 1,
Z

1
Z 0
1 0 dN fˆ(ω) 1
fˆ(ω)e −iωz −iωz

|f (z)| = dω = e dω

2π 2π

dω N N
−Sω −Sω (−iz)

1 dN fˆ(ω) −iωz
Z 0
1
dω = O CN Sω ∆−N |z|−N exp(Sω |Im(z)|) .

≤ N N e
2π −Sω |z| dω



For any N ∈ N, there exists a constant CN > 0 such that

|f (z)| = O CN Sω ∆−N (1 + |z|)−N exp(Sω |Im z|) ,



for all z ∈ C, (B5)
which proves (B3) by choosing z to be a real number.
Plugging (B5) into left-hand side (B4), we obtain that
Z ∞  Z ∞ 
|f (s + iζ) exp(iω(s + iζ))| ds ≤ O CN Sω ∆−N exp(2∥H∥ + Sω ) (1 + |s|)−N ds
∞ ∞

for any |ω| ≤ 2∥H∥, |ζ| ≤ 1. This proves (B4).


11

The conclusion of Lemma 8 helps us control the error of the trapezoid rule:
Theorem 9 (Error bound for the infinite trapezoidal rule [41, Theorem 5.1]). Suppose h is analytic in the strip
|Im(z)| < 1. Suppose

lim sup |h(s + iζ)| → 0 ,


|s|→∞ |ζ|<1

and
Z ∞
|h(s + iζ)| ds ≤ M
−∞

for any complex numbers s and |ζ| < 1. Then, for any τs > 0, we have
Z ∞

∞ X 2M
h(s) ds − τs h(nτs ) ≤ .

exp(2π/τs) − 1

−∞ n=−∞

B.2. Simulating continuous-time Lindblad dynamics with one ancilla qubit

The Lindblad consists of the coherent part and the dissipative part. For simplicity, we use a first-order Trotter
splitting

eLt = e(LH +LK )t (B6)


LH τ LK τ t/τ
≈ (e e ) for time step τ. (B7)

This allows us to focus on simulating eLK τ for a short time τ .

B.3. Simulation of Lindblad dynamics

We define the dilated Hermitian jump operator using one ancilla qubit as

0 K†
 
K :=
e . (B8)
K 0
P 
1 P1
Define the partial trace Tra i,j=0 |i⟩ ⟨j| ⊗ ρi,j = i=0 ρi,i , which traces out the ancilla qubit. Our simulation
approach is based on the following lemma:
Lemma 10 (Lindbladian simulation using one ancilla qubit). Let
√ √
σ(t) := Tra e−iK t
[|0⟩ ⟨0| ⊗ ρ] eiK t
. (B9)
e e

Then for a short time t ≥ 0,

∥σ(t) − ρ(t)∥1 = O(∥K∥4 t2 ) (B10)

where ρ(t) is the solution to the Lindblad dynamics

∂t ρ(t) = LK [ρ(t)] with initial condition ρ(0) = ρ. (B11)

Proof of Lemma 10. According to Eq. (A1), we first obtain


√ √
∥ exp(−iK e t) − (1 − iK e t+K e 2 t/2 + K
e 3 t3/2 /6)∥ = O(∥K∥
e 4 t2 ) = O(∥K∥4 t2 ) .

Let T3,K = (1−iK
e t+ K e 2 t/2+ K
e 3 t3/2 /6) denote the truncated Taylor expansion, and 1 denotes the identity operator.
We have


σ(t) − Tra T3,K [|0⟩ ⟨0| ⊗ ρ] T3,K = O(∥K∥4 t2 ) . (B12)

1
12

Because Tra (|i⟩ ⟨j| ⊗ ρ) = δi,j ρ, We note that terms with an odd power of O(t1/2 ) vanish after applying the partial
trace operation. This implies
 
Tra T3,K [|0⟩ ⟨0| ⊗ ρ] T † − 1 + KρK † t − 1 (K † Kρt + ρK † Kt)

3,K
2
1
(B13)


= Tra T3,K [|0⟩ ⟨0| ⊗ ρ] T3,K − (1 + LK t)ρ

1
=O(∥K∥4 t2 )

By combining equations (B12) and (B13), we arrive at the following expression:

∥σ(t) − (1 + LK t)ρ∥1 = O(∥K∥4 t2 ) .

Finally, since

∥ρ(t) − (1 + LK t)ρ∥1 = ∥ exp(LK t) − (1 + LK t)ρ∥1 = O ∥LK ∥21 t2 = O(∥K∥4 t2 ) ,




we conclude the proof.

According to Lemma 10, we can approximate eLK τ [ρ] by employing a dilated Hamiltonian simulation, tracing out
the ancilla qubit, and resetting the ancilla qubit to the state |0⟩.

B.4. Trotter-based simulation without block encoding the jump operator

Simulating Eq. (B9) still requires a block encoding for the dilation K, e which can require many ancilla qubits. Now

we demonstrate that it is possible to use Trotter splitting to implement exp(−iK e t) directly using one ancilla qubit
in total and without the block encoding of K or K. e
 R 

Discretizing the time integral. As discussed before, to implement exp i −∞ f (s)A(s)ds , we need to truncate
and discretize the integral.
First, since fˆ(ω) is assumed to be compactly supported in Assumption 7, according to Appendix A Lemma 8, f (s)
decays super-polynomially as |s| → ∞. Thus, we can choose a suitable Ss > 0, restrict the integration range to
[−Ss , Ss ], and discretize this interval using a uniform grid sl = lτs and l = −Ms , . . . , Ms , where τs = Ss /Ms . We can
then approximate the integral using a trapezoidal rule:
Ms
X
Ks := f (sl )eiHsl Ae−iHsl wl , (B14)
l=−Ms

where wl = τs /2 for l = ±Ms , and wl = τs for −Ms < l < Ms .

Lemma 11 (Convergence of the quadrature error). Let fˆ satisfy Assumption 7. Then, given any N, ϵ′ > 0, if
  1 !   −1 !
1 CN Sω ∥A∥ (N −1) CN Sω ∥A∥
Ss = Ω , τs = O ∥H∥ + Sω + log ,
∆ ∆ϵ′ ∆ϵ′

we have

∥K − Ks ∥ = O(ϵ′ ) . (B15)

Here the constant CN comes from Assumption 7.

The proof is given in Appendix C.3. By taking N → ∞ in Eq. (B5), we obtain Ss = O(ϵ′−o(1) ), the number
of quadrature points Ms = Ss /τs = O(ϵ e ′−o(1) ). Therefore the accuracy of the trapezoidal rule improves super
polynomially with respect to the increase of Ss , Ms . We remark that by imposing stronger regularity conditions
on fˆ(ω) (e.g., fˆ(ω) is compactly supported and is in the Gevrey class [15], which is stronger than fˆ ∈ C ∞ (R)), the
aforementioned result can be enhanced to Ss = Ms = O(poly log(ϵ′−1 )). For simplicity, this work does not incorporate
such an improvement.
13

According to Lemma 11, a small number of point Ms already ensures a good approximation
 †  Ms
P
Ms
0 l=−M f (sl )A(sl )wl
X
K ≈ Ks = P
e e  s  =: H
el. (B16)
Ms
l=−Ms f (sl )A(sl )wl 0 l=−Ms

Since the Heisenberg evolution A(s) = eiHs Ae−iHs is Hermitian, the term further factorizes as
f ∗ (sl )A(sl )wl
 
0
Hl =
e = σl ⊗ A(sl ) where σl := wl (σx Re f (sl ) + σy Im f (sl )) (B17)
f (sl )A(sl )wl 0
and σx and σy are Pauli matrices.
Second-order Trotter splitting. After discretizing the time labels, the next step is to Trotterize the Hamiltonian
evolution
√ e √ P e
e−i τ Ks
= e−i τ l Hl . (B18)
Specifically, we employ the second-order Trotter formula to balance between efficiency and accuracy. The second-order
√ e
Trotter formula for exp(−i τ K) can be expressed as:
→ √ ← √ √ e
Y τ Y τ X
e−i H
e−i H
− e−i τ Ks
= τ 3/2 al1 ,l2 ,l3 H
el Hel He l + O(τ 2 ), (B19)
el el
2 2
1 1 3
l l l1 ,l2 ,l3

where the coefficients al1 ,l2 ,l3 can be calculated from Taylor expansion, and the left and right-ordered products are
defined by

Y √ e √ e √ e →
Y √ e √ e √ e
e−iτ Hl
:= e−i τ H Ms
· · · e−i τ H−Ms
and e−i τ Hl
:= e−i τ H−Ms
· · · e−i τ H Ms
. (B20)
l l

The Trotter error bounds (B19) and Lemma 10 together give an approximation scheme for simulating the Lindblad
dynamics in (3). For any initial state ρ, we have that
→ ← → ←
!
√ √ √ √
−i 2τ H −i 2τ H i 2τ H i 2τ H
Y Y Y Y
Tra e e |0⟩ ⟨0| ⊗ ρ e e
el el el el

l l l l
 √ √ e  (B21)
=Tra e−i τ Ks [|0⟩ ⟨0| ⊗ ρ]ei τ Ks + O(τ 2 )
e

=eLKs τ [ρ] + O(τ 2 ) ≈ eLK τ [ρ] + O(τ 2 ) ,


   
where the second equality is obtained by using Tra H el Hel H e† H
e l |0⟩ ⟨0| ⊗ ρ = 0 and Tra |0⟩ ⟨0| ⊗ ρH e† H
e † = 0.
1 2 3 l3 l2 l1
We also use Lemma 10 in the last equality.
From the analysis above, we find that replacing the second-order formula with the first-order formula results in a
local truncation error of O(τ ). This substitution leads to an excessively large global error of O(1). While higher-order
Trotter can further suppress the error in the first equality, the main error in the last equality comes from Lindbladian
simulation by means of the method in Lemma 10 and is not easily reducible. Therefore, the second-order Trotter
seems to be adequate for the purposes of Lindbladian simulation4 . √
τ e
Canceling out back-and-forth Hamiltonian evolution. Finally, to efficiently implement the products e−i 2 Hl ,
we notice
 √  √
τ τ
exp −i σl ⊗ A(sl ) = (I ⊗ eiHsl )e|−i 2{zσl ⊗A}(I ⊗ e−iHsl ). (B22)
2 √
=:A
el ( τ )

el (√τ ) is mostly negligible compared to that of the


Since A is a simple operator (e.g., a local Pauli), the cost of A
simulation of the system Hamiltonian.
What about the eiHsl terms? A moment of thought reveals that we may rewrite the consecutive product in a
form that efficiently cancels out the back-and-forth Hamiltonian evolution. We present the most abstract form to
emphasize the simplicity of this observation.

4 There exists another method to symmetrize the first-order for- not be discussed here.
mula to reduce the local truncation error to O(τ 2 ), which will
14

Proposition 12 (Cancellations in time-order products). Consider a time-order product at discretized times sl = lτs
for l = −Ms , · · · , Ms . Then, for any Hamiltonian H and a set of matrices Al depending on l,
→ →
!
Y Y
−iHSs iHτs
Al (sl ) = e Al e eiH(Ss +τs ) (B23)
l l
← ←
!
Y Y
Al (sl ) = e−iH(Ss +τs ) e−iHτs Al eiHSs (B24)
l l

where Al (s) = eiHs Al e−iHs denotes the Heisenberg evolution of Al with H.


Proof. The Hamiltonian evolution from two consecutive steps nearly cancels (for the right-ordered product for exam-
ple)

Al (sl )Al (sl+1 ) = eiHsl Al eiHτs Al e−iHsl+1 since e−iHsl eiHsl+1 = eiHτs . (B25)

For our usage, our second-order formula consists of both the left and right products, which is
→ ←
Y √ Y √
el ( τ )(I ⊗ e−iHsl ) (I ⊗ eiHsl )A
(I ⊗ eiHsl )A el ( τ )(I ⊗ e−iHsl )
l l
→ ←
! !
−iHSs
Y √ iHτs
Y
−iHτs
√ (B26)
=(I ⊗ e ) el ( τ )(I ⊗ e
A ) (I ⊗ e el ( τ ) (I ⊗ eiHSs ).
)A
l l
| {z√ }
=:W ( τ )

where W ( τ ) is a product of short-time Hamiltonian simulation.
Even nicer, the long-time simulation (I ⊗ e−iHSs ) from the previous time step exactly cancels with (I ⊗ eiHSs ) from
the subsequent time step. Therefore, we may remove both long-time evolution steps and define the quantum channel
√ √
W(τ )[ρ] := Tra W ( τ ) [|0⟩ ⟨0| ⊗ ρ] W † ( τ ). (B27)

The total simulation time of the system Hamiltonian for implementing the quantum channel W(τ ) now becomes
Ms
X
τs = O(Ss ). (B28)
l=−Ms

In summary, the single ancilla simulation of the Lindblad dynamics takes the form

ρm+1 = W(τ )[ρm ] (purely dissipative) (B29)


LH τ
or ρm+1 = e W(τ )[ρm ] (Trotterizing the coherent part) (B30)

The second line displays the option


√ to include the coherence term via Trotter, which may introduce additional errors 5 .
The quantum circuits for W ( τ ) and (B29) are drawn in Figure 3.
Finally, let T = Mt τ and allow τ to approach zero. We expect that ρMt from (B30) converges to
exp(−LH Ss )[ρ(T )] = eiHSs ρ(T )e−iHSs , where ρ(t) is the solution of the modified Lindblad dynamics

∂t ρ(t) = LH [ρ(t)] + LK [ρ(t)], where ρ(0) = exp(LH Ss )[ρI ]. (B31)

Compared to the original Lindblad dynamics in Eq. (3), the initial state is changed to exp(LH Ss )[ρI ] = e−iHSs ρI eiHSs .
In particular, if ρI commute with H (e.g., ρI is the density operator corresponding to an eigenstate of H, or the
maximally mixed state), then exp(LH Ss )[ρI ] = ρI . At the end of the simulation, if ρ(T ) is the ground state ρg , then
exp(−LH Ss )[ρg ] = ρg . We, therefore, expect the behavior of the modified Lindblad dynamics to be very similar to
that of the original dynamics.

5 The dissipative part itself already fixes the ground state. How- the algorithm stops at other fixed points.
ever, numerically, without this coherent term, we observe that
15

|0⟩ ···
e−Ms (√τ )
A e−Ms +1 (√τ )
A eMs −1 (√τ )
A eMs (√τ )
A
···
|ψ⟩ e−iHτs eiHτs
···

Figure 3. Quantum
√ circuits for the algorithm to simulate the continuous-time Lindblad dynamics. Upper: Quantum circuit
used for W ( τ ); Lower: Quantum circuit used for continuous-time Lindblad simulation. The time step should satisfy τ = O(ϵ).
The measurement result of the ancilla qubit is discarded and the ancilla qubit is reset to |0⟩ after each measurement. Here we
assume A is a local operator that only acts on one system qubit (or a small number of system qubits).

B.5. Discrete-time Lindblad dynamics with one ancilla qubit

The short time propagator for simulating the continuous-time Lindblad dynamics in Eq. (B30) is limited to first-
order accuracy, which is mainly limited by the inexactness of the propagator according to Lemma 10 (also discussed
after Theorem 13). However, our goal is not to simulate the continuous-time Lindblad dynamics but to prepare the
ground state. If we apply the dilated jump operator K e to |0⟩ ⟨0| ⊗ ρg , then according to Eq. (7),
   
0 0
K |0⟩ ⟨0| ⊗ ρg =
e = . (B32)
Kρg 0

This implies
   
exp −iKt
e [|0⟩ ⟨0| ⊗ ρg ] exp iKt
e = |0⟩ ⟨0| ⊗ ρg ,

for any t > 0. Thus, when the quadrature error and the Trotter error are properly controlled, we have

exp(LH τ )Wa (τ )[ρg ] ≈ ρg , for all τ > 0. (B33)

e √τ [|0⟩ ⟨0| ⊗ ρg ] exp iK


    √ 
where Wa (τ )[ρg ] = Tra exp −iK e τ .
Therefore, the simulation scheme outlined in Eq. (B30) could conceivably be used for ground state preparation
with an arbitrarily large time step τ . By choosing a large time step, the quantum state ρm may not necessarily
approximate the exact dynamics ρ(mτ ). However, we expect that after the mixing time t′mix = Mmix τ , ρMmix
becomes a good approximation to ρg because of the fixed point argument. Here, the “discrete” mixing time Mmix
is given in Definition 6. If t′mix is not much larger than tmix , we can significantly reduce the simulation cost thanks
to the increase of τ . This gives rise to the discrete-time Lindblad dynamics, which is defined as the continuous-time
Lindblad simulation with a large time step.
e√
When using a large τ , additional attention must be paid to controlling the Trotter error. Specifically, e−iK τ should
be simulated as
e√

e√
r √ r
e−iK τ = e−iK τ /r ≈ W ( τ /r) , (B34)

where r denotes the number of segments and should be properly chosen.


In summary, one single step of the discrete-time Lindblad dynamics is

ρm+1 = exp(LH τ )W(τ, r)[ρm ]. (B35)


16

Figure 4. Quantum circuit for discrete-time Lindblad dynamics simulation. The measurement result of the ancilla qubit is
discarded and the ancilla qubit is reset to |0⟩ after each measurement. The circuit for W is given in Fig. 3. The time step τ
can be chosen to be independent of the precision ϵ, and the number of segments r should be properly chosen to control the
Trotter error.

where
√ √
W(τ, r)[ρ] = Tra [W ( τ /r)]r (|0⟩ ⟨0| ⊗ ρ) [(W ( τ /r))† ]r . (B36)

The quantum circuit for (B35) is drawn in Figure 4.


Since we also cancel the long-time simulation when formulating W (τ, r) in this case, if we increase r and fix T, τ ,
we expect that ρMt from (B35) converges to the solution ρaMt of the following discrete dynamics:
e√ e√
 
ρam+1 = exp(LH τ )Tra e−iK τ [|0⟩ ⟨0| ⊗ ρam ]eiK τ , where ρa0 = exp(LH Ss )[ρ0 ] . (B37)

Appendix C: Analysis of the algorithm

In this section, we investigate the complexity of simulating the Lindblad dynamics as described by equations (B31)
and (B37) using the schemes (B30) and (B35). The rigorous version of Theorems 2 and 3 are presented in Theorems
13 and 14, respectively.

C.1. Simulation cost of continuous-time Lindblad dynamics

Given a stopping time T > 0 and the time step τ , we assume the number of iterations Mt = T /τ is an integer. We
now prove that ρMt from (B30) converges to ρ(T ) from (B31) with first order accuracy in τ .
Theorem 13 (Simulation cost of the continuous-time Lindblad dynamics with one ancilla qubit). Given a
stopping time T > 0 and ˆ
iHS the accuracy ϵ > 0, we assume that f satisfies Assumption 7. Then, to obtain
e s ρ(T )e−iHSs − ρMt ≤ ϵ, we can choose
1

  (N1−1) !   −1 !  
1 CN Sω ∥A∥T CN Sω ∥A∥ T ϵ
Ss = Θ , τs = Θ ∥H∥ + Sω + log , τ =Θ ,
∆ ∆ϵ ∆ϵ (∥A∥ + ∥A∥2 ∥H∥)T
4

where CN is the constant that comes from Assumption 7. In particular, the total Hamiltonian simulation time for H

TH,total = Θ(T τ −1 Ss + T ) = Θ((1 + ∥H∥)∆−1 T 2+o(1) ϵ−1−o(1) ) .

In addition, the total number of controlled-A gates is

NA,gate = Θ(T τ −1 Ss τs−1 ) = Θ((1


e + Sω + ∥H∥)2 ∆−1 T 2+o(1) ϵ−1−o(1) ) .

Note that the Hamiltonian simulation time is independent of the step size τs for discretizing the integral in forming
the jump operator K.
The proof of Theorem 13 is in Appendix C.3. We emphasize that the bottleneck leading to the first-order accuracy
is not due to the Trotter splitting scheme used in Eq. (B26), but the first-order accuracy of the Lindblad simulation
method in Lemma 10.
17

C.2. Simulation cost of discrete-time Lindblad dynamics

The simulation cost of (B35) is shown in the following theorem:

Theorem 14 (Simulation cost of the discrete-time Lindblad dynamics). Assume that fˆ satisfies Assumption 7. Given
a stopping time T > 0, a time step τ = O(∥A∥−2 ) such that Mt = T /τ ∈ N, we generate ρaMt and ρMt using (B37)
and (B35) correspondingly.
Given the accuracy ϵ > 0, to obtain ∥eiHSs ρaMt e−iHSs − ρMt ∥1 ≤ ϵ, we can choose
 (N1−1) ! −1 !
∥A∥T 1/2
    
1 CN Sω ∥A∥T CN Sω ∥A∥ T
Ss = Θ , τs = Θ ∥H∥ + Sω + log , r=Θ ,
∆ ∆ϵ ∆ϵ ϵ1/2
(C1)
where CN is a constant that only depends on N, fˆ(ω). In particular, letting N → ∞, the total Hamiltonian simulation
time for H

TH,total = Θ(T τ −1 Ss r + T ) = Θ(∆−1 T 3/2+o(1) ϵ−1/2−o(1) ) ,

and the total number of controlled-A evolution gates


 
NA,gate = Θ(T τ −1 Ss τs−1 r) = Θ
e (1 + Sω + ∥H∥)∆−1 T 3/2+o(1) ϵ−1/2−o(1) .

We put the proof in Appendix C.4.


Unlike the simulation of the continuous-time Lindblad dynamics in Theorem 13, the bottleneck of the
√ second order
accuracy in Theorem 14 is due to the second order Trotter formula for the short time
√ propagator W ( τ ). Replacing
the second-order Trotter formula with a p-th order Trotter formula in defining W ( τ ), we can further improve the
asymptotic scaling to be nearly linear in T . This is shown in Corollary 15.
Corollary 15 (Simulation cost of the discrete-time Lindblad dynamics
√ with high order splitting for W ). Under the
same assumptions of Theorem 14, but assume that the propagator W ( τ /r) is constructed using a p-th order Trotter
method, we may choose

r = O((∥A∥2 T /ϵ)1/p ).

Then with the same choice of Ss , τs as in Theorem 14, and letting N, p → ∞, the total Hamiltonian simulation time
for H

TH,total = Θ(T τ −1 Ss r + T ) = Θ(∆


e −1 T 1+o(1) ϵ−o(1) ) ,

In addition, the total number of controlled-A gates is


 
NA,gate = Θ(T τ −1 Ss τs−1 r) = Θ
e (1 + Sω + ∥H∥)∆−1 T 1+o(1) ϵ−o(1) .

Remark 16. While Eq. (B30) partitions the coherent and dissipative components of the continuous-time Lindblad
dynamics in a manner similar to a first-order Trotter method, this scheme still maintains the ground state. Therefore
this first-order-like splitting may influence the mixing time, but may not√contribute to the error of the ground state.

On the other hand, the accuracy of W ( τ ), which approximates e−iK τ up to a change of frame, plays a vital role
e

in the fixed point argument and needs to be implemented accurately. In practical scenarios (Sec. Numerics results
and Appendix E), we observe that employing a second-order Trotter method with τ = O(1) and r = O(1) is often
sufficient.

C.3. Proof of Theorem 13

To prove Theorem 13, we first show the following proposition:


Proposition 17. Under the assumptions of Theorem 13, we assume τ = O(∥A∥−2 ), then
e s ρ(T )e−iHSs − ρMt = O (∥K − Ks ∥∥A∥T ) + O (∥A∥4 + ∥A∥2 ∥H∥)T τ .
iHS 
1
(C2)
18

Proof of Proposition 17. We first notice that

e s ρ(T )e−iHSs − ρM = ρ(T ) − e−iHSs ρM eiHSs .


iHS
t 1 t 1

Define ρ̄m = e−iHSs ρm eiHSs . Then

√ √ †
ρ̄m+1 = e−iHτ Tra e−iHSs W ( τ )eiHSs (|0⟩ ⟨0| ⊗ ρ̄m ) e−iHSs W ( τ )eiHSs eiHτ


with ρ̄0 = e−iHSs ρI eiHSs .


Because the Lindblad dynamics is contractive in trace distance according to Eq. (A4), to prove (C2), it suffices to
prove

4
∥A∥4 + ∥H∥2 )τ 2

∥ρ(τ ) − ρ̄1 ∥1 = O (∥K − Ks ∥∥A∥τ ) + O (CN (C3)

Define
 √ e √ e
ρe1 = e−iHτ Tra e−i τ K |0⟩ ⟨0| ⊗ ρ(0)ei τ K eiHτ ,

and
 √ e √ e 
ρe1,s = e−iHτ Tra e−i τ Ks |0⟩ ⟨0| ⊗ ρ(0)ei τ Ks eiHτ .

Then

∥ρ(τ ) − ρ1 ∥1 ≤ ∥ρ(τ ) − ρe1 ∥1 + ∥e


ρ1 − ρe1,s ∥1 + ∥e
ρ1,s − ρ̄1 ∥1 . (C4)

The first term contains the Lindblad simulation error, the second term contains the error from the numerical integra-
tion, and the third term contains the error from Trotter splitting. According to Lemma 10 and Lemma 4 (A5), we
first bound the Lindblad simulation error:

∥ρ(τ ) − ρe1 ∥1
≤ ∥e
ρ1 − exp(LH τ ) exp(LK τ )ρ(0)∥1 + ∥exp((LH + LK )τ )ρ(0) − exp(LH τ ) exp(LK τ )ρ(0)∥1
(C5)
=O ∥K∥4 τ 2 + O ∥ [LH , LK ] ∥1 τ 2
 

=O ∥A∥2 (∥A∥2 + ∥H∥)τ 2




where we use Lemma 10 in the first equality and Lemma 5 in the last equality.
Next, in order to bound the error of the numerical integration, we observe that
 √ √ e
eiHτ ρe1 e−iHτ = Tra e−i τ K |0⟩ ⟨0| ⊗ ρ(0)ei τ K ,
e

and
 √ √ e 
eiHτ ρe1,s e−iHτ = Tra e−i τ Ks |0⟩ ⟨0| ⊗ ρ(0)ei τ Ks .
e

Define

ρe1,a (t) = e−itK |0⟩ ⟨0| ⊗ ρ(0)eitK , ρe1,s,a (t) = e−itKs |0⟩ ⟨0| ⊗ ρ(0)eitKs .
e e e e

Noticing

∥e
ρ1,s,a (t) − ρe1,s,a (0)∥1 = O(∥K
e s ∥t) ,
19

we have
√ √ 
∥eρ1 − ρe1,s ∥1 = Tra ρe1,a ( τ ) − ρe1,s,a ( τ )
Z √τ !
 √   √ 
= Tra exp −iK( e τ − t) [K e −K e s , ρe1,s,a (t)] exp iK( e τ − t) dt


0
1
Z τ √ !
 √   √ 
≤ Tra exp −iK( e τ − t) [K e −K e s , ρe1,s,a (t) − ρe1,s,a (0)] exp iK(
e τ − t) dt


0
1

, (C6)
!
Z τ  √   √ 
+ Tra exp −iK(
e τ − t) [K e −K e s , ρe1,s,a (0)] exp iK( e τ − t) dt

0
1
√ √ e √ e √ e
−i τ Ke i τK −i τ K i τK
≤ e |0⟩ ⟨0| ⊗ ρ(0)e −e s
|0⟩ ⟨0| ⊗ ρ(0)e s

1
   
=O K − Ks Ks + K τ
e e e e

=O (∥K − Ks ∥ ∥A∥τ )
e −K
where we use the Duhamel principle in the second equality, Tra ([K e s , ρe1,s,a (0)]) = 0 in the second inequality, and
∥Ks ∥ = O(∥A∥) in the last equality.  √ 
Finally, we bound the error of Trotter splitting. Using (A1) to expand each exp −iH̃l τ /2 up to N = 3, we
obtain that:
 !4   !4 
√ X X
∥Hl ∥  = O τ 2 ∥A∥4 τs |f (sl )|  = O τ 2 ∥A∥4

∥e
ρ1,s − ρ̄1 ∥1 =O  τ (C7)
l l

where we use Lemma 8 (B3) with N = 1 in the third equality. Here all τ 3/2 terms disappear because of the partial
trace. Plugging (C5), (C6), and (C7) into (C4), we prove (C3).
To bound ∥K − Ks ∥ in (C2), we need to use Lemma 11. We first give its proof there.
Proof of Lemma 11. We separate the error into two parts:
∥K − Ks ∥ ≤ ∥K − K∞ ∥ + ∥K∞ − Ks ∥ , (C8)
P∞
where K∞ = l=−∞ f (sl )eiHsl Ae−iHsl τs . For a given N , using Lemma 8 (B3), the second part of error can be
bounded as
 
X  Z ∞ 
∥K∞ − Ks ∥ = O CN Sω ∆−N ∥A∥ (1 + |sl |)−N τs  = O CN Sω ∆−N ∥A∥ (1 + s)−N ds = O (ϵ′ ) . (C9)
|l|≥Ms Ss

To bound the discretization error ∥K − K∞ ∥, first define



X
fˆs,∞ (ω) = f (sl ) exp(iωsl )τs .
l=−∞

Use the fact that |λi − λj | ≤ 2 ∥H∥, we have



∥K − K∞ ∥ ≤ ∥A∥ sup fˆ(λi − λj ) − fˆs (λi − λj )

i,j
Z ∞

∞ X
=∥A∥ sup f (s) exp(iωs) ds − f (sl ) exp(iωsl )τs

|ω|≤2∥H∥ −∞ l=−∞

Fixing an arbitrary ω ∈ [−2∥H∥, 2∥H∥], we define h(z) = f (z) exp(iωz). According to Lemma 8 (B4), we obtain that
h(z) is an analytic function in z and
Z ∞
|h(x + iy)| dx = O CN Sω ∆−N exp(2∥H∥ + Sω )

sup
|ω|≤2∥H∥, |y|≤1 −∞
20

We then use Theorem 9 to obtain


Z ∞

∞ X
h(sl )τs = O CN Sω ∆−N exp(2∥H∥ + Sω ) exp(−2π/τs ) .

h(s) ds −


−∞
l=−∞

This gives

∥K − K∞ ∥ = O CN Sω ∆−N ∥A∥ exp(2∥H∥ + Sω ) exp(−2π/τs ) = O (ϵ′ ) .



(C10)
  −1 
CN Sω ∥A∥
where we use τs = O ∥H∥ + Sω + log ∆ϵ′ in the last inequality. Plugging (C9) and (C10) into (C8),

we obtain ∥K − Ks ∥ = O(ϵ ).

Now, we are ready to prove Theorem 13:

Proof of Theorem 13. According to Proposition 17 (C2) and Lemma 11, we set ϵ′ = ϵ
∥A∥T in (B15) to obtain that

ϵ
+ O (∥A∥4 + ∥A∥2 ∥H∥)T τ ≤ ϵ ,

∥ρ(T ) − ρMt ∥1 =
2
 
ϵ
where we use τ = O (∥A∥4 +∥A∥2 ∥H∥)T in the inequality.
Next, we calculate the total Hamiltonian simulation time for H. According to (B28), each implementation of W (τ )
needs to simulate the system Hamiltonian for time Θ(Ss ). Thus, the total Hamiltonian simulation time:

TH,total = number of steps × (τ + Θ(Ss )) = Mt (τ + Θ(Ss )) = Θ(T + T Ss τ −1 ) = Θ(T


e 2+o(1) ϵ−1−o(1) ) .

Finally, to calculate the number of controlled-A evolution gates, we notice that each trotter splitting step in W (τ )
needs to implement Θ(1) controlled-A evolution gates. This implies

NA,gate = number of steps × Θ(Ss τs−1 ) = Θ(T


e 2+o(1) ϵ−1−o(1) ) .

C.4. Proof of Theorem 14 and Corollary 15

To prove Theorem 14, we first show the following proposition:


Proposition 18. Under the conditions of Theorem 14, we further assume r = Ω(1), then
iHS a −iHS
− ρMt 1 = O (∥K − Ks ∥∥A∥T ) + O ∥A∥4 T τ /r2 .

e s ρM e s
(C11)
t

Proof of Proposition 18. For simplicity, we ignore the effect of eiHSs . Because the Lindblad dynamics is contractive
in trace distance according to Eq. (A4), to prove (C11), it suffices to prove

∥ρa1 − ρ1 ∥1 = O (∥K − Ks ∥∥A∥τ ) + O ∥A∥4 τ 2 /r2 .



(C12)

Define
 √ e √ e 
ρe1,s = e−iHτ Tra e−i τ Ks |0⟩ ⟨0| ⊗ ρ(0)ei τ Ks eiHτ .

Then

∥ρa1 − ρ1 ∥1 ≤ ∥ρa1 − ρe1,s ∥1 + ∥e


ρ1,s − ρ1 ∥1 , (C13)

where the first term contains the error from the numerical integration, and the third term contains the error from
Trotter splitting. Similar to Eq. (C6) in the proof of Theorem 13, for the first term, we can bound is by:

∥ρa1 − ρe1,s ∥1 ≤ O(∥K − Ks ∥∥A∥τ ) . (C14)


21

To bound the second term, we notice


∥e
ρ1,s − ρ1 ∥1

e √ e √ √ √  (C15)
≤ Tra e−iKs τ |0⟩ ⟨0| ⊗ ρ(0)eiKs τ − [W ( τ /r)]r (|0⟩ ⟨0| ⊗ ρ(0)) [(W ( τ /r))† ]r .

1

Using the second-order Trotter splitting formula and Eq. (A1) with N = 3, we can rewrite W ( τ /r) as

√ √ e τ 3/2
W ( τ /r) = e−i τ Ks /r + 3 E1 + E2 (τ, r) . (C16)
r
Here the derivation of the expression
X
E1 = al1 ,l2 ,l3 H
el H
1
el H
1
el
3
l1 ,l2 ,l3

is similar to that in Eq. (B20). Moreover,


 !4   !4 
2
τ2 ∥A∥4 τ 2
 
τ X X
∥E2 (τ, r)∥ = O  4 ∥Hl ∥  = O  4 ∥A∥4 τs |f (sl )| =O .
r r r4
l l


Here, in order to obtain the correct leading order error, we need to expand W ( τ /r) up to O(τ 2 ).
Plugging (C16) into (C15), we have

∥e
ρ1,s − ρ1 ∥1
r  3/2 √ e √ e √ e

X τ −i(k−1) τ K s /r −i(r−k) τ K s /r i τK

≤ Tra
e E1 e |0⟩ ⟨0| ⊗ ρ(0)e s
r3
1
k=1
r  3/2 √ e √ e √ e

X τ −i τ K i(k−1) τ K s /r † i(r−k) τ K s /r

+ Tra e s
|0⟩ ⟨0| ⊗ ρ(0)e E1 e
r3
1
k=1
2 4
 
τ ∥A∥
+O
r4
 
|τ |p/2 ∥A∥p
Here, we use the relationship r = Ω(1) and τ = O(∥A∥−2 ) to incorporate all terms of the form rp (for p ≥ 4)
 2 4
into the asymptotic notation O τ ∥A∥r4 .
√ e
Next, since Tra (E1 |0⟩ ⟨0| ⊗ ρ(0)) = 0, we can expand ei τ Ks to first order and obtain the refined estimate
 3/2 √ √ √
  2 4

Tra τ e−i(k−1) τ Ke s /r E1 e−i(r−k) τ Ke s /r |0⟩ ⟨0| ⊗ ρ(0)ei τ Ke s = O τ ∥A∥

r3
1 r3
3
for all 1 ≤ k ≤ r. Here we use ∥E1 ∥ = O(∥A∥ ), and ∥K∥ = O(∥A∥). Therefore
r  3/2 √ √ √
  2 4

Tra τ e−i(k−1) τ Ke s /r E1 e−i(r−k) τ Ke s /r |0⟩ ⟨0| ⊗ ρ(0)ei τ Ke s = O τ ∥A∥ .
X
(C17)
r3
1 r2
k=1
 
Similarly using Tra |0⟩ ⟨0| ⊗ ρ(0)E1† = 0, we have

r  3/2 √ √ √
  2 4

Tra τ e−i τ Ke s |0⟩ ⟨0| ⊗ ρ(0)ei(k−1) τ Ke s /r E † ei(r−k) τ Ke s /r = O τ ∥A∥
X
1 (C18)
r3 r2 1
k=1

This gives
τ 2 ∥A∥4
 
∥e
ρ1,s − ρ1 ∥1 = O .
r2

Plugging this equality and (C14) into (C13), we prove (C12).


22

Now, we are ready to prove Theorem 14:


Proof of Theorem 14. We note that Lemma 11 also holds for this case. Thus, we apply Proposition 18 (C11) and
Lemma 11 by setting ϵ′ = ∥A∥T
ϵ
in (B15) to obtain that

ϵ
+ O ∥A∥4 T τ /r2 ≤ ϵ ,

∥ρ(T ) − ρMt ∥1 =
2
where we use r = Θ(∥A∥T 1/2 ϵ−1/2 ) and τ = O(∥A∥−2 ) in the inequality.
Next, we calculate the total Hamiltonian simulation time for H. According to (B28), the implementation of each
W (τ /r) needs to simulate the system Hamiltonian for time Θ(Ss ). Thus, the total Hamiltonian simulation time:

TH,total = number of steps × (τ + rΘ(Ss )) = Θ(T + rT Ss τ −1 ) = Θ(T


e 3/2+o(1) ϵ−1/2−o(1) ) .

Finally, each step within the Trotter-splitting process of W (τ /r) requires the implementation of Θ(1) controlled-A
evolution gates. This implies

NA,gate = number of steps × Θ(Ss τs−1 r) = Θ(T


e 3/2+o(1) ϵ−1/2−o(1) ) .

Proof of Corollary 15. We note that Lemma 11 also holds for this case. In addition, when applying p-th order Trotter
scheme, we obtain
iHS a −iHS  
p+2 p/2 p

e s ρM e
t
s
− ρ Mt 1
= O (∥K − K s ∥∥A∥T ) + O ∥A∥ T τ /r . (C19)

Then, we apply (C19) and Lemma 11 by setting ϵ′ = ϵ


∥A∥T in (B15) to obtain that

ϵ  
∥ρ(T ) − ρMt ∥1 = + O ∥A∥p+2 T τ p/2 /rp ≤ ϵ ,
2
where we use r = Θ(∥A∥2/p T 1/p ϵ−1/p ) and τ = O(∥A∥−2 ) in the inequality.
Similar to the previous proof, the total Hamiltonian simulation time is

TH,total = number of steps × (τ + rΘ(Ss )) = Θ(T + rT Ss τ −1 ) = Θ(T


e 1+1/p ϵ−1/p )

and the total number of controlled-A gates is

NA,gate = number of steps × Θ(Ss τs−1 r) = Θ(T


e 1+1/p ϵ−1/p ) .

This concludes the proof.

Appendix D: Convergence under ETH type ansatz

In this section, we will investigate the convergence of our algorithm to the ground state based on appropriate
assumptions. Specifically, we focus on the continuous-time Lindblad dynamics (3).
Firstly, we study the convergence property. Define Ai,j = ⟨ψi | A |ψj ⟩ and fˆi,j = fˆ(λi − λj ). In order to study the
convergence of the continuous-time Lindblad dynamics (3), we need to make some independent assumptions for A.
Assumption 19. • (Random matrix elements) Assume for any t ≥ 0, A is independently drawn from random
probability distribution ΞA on the set of Hermitian matrices such that Ai,j are independent and E(Ai,j ) = 0
when i ̸= j. Denote σi,j = E(|Ai,j |2 ) > 0.
−2
• (Support of filter) Sω > minN
j=0 (λj+1 − λj ).

N −1
• (Diagonal initial state) ρ(0) is a diagonal matrix in the basis of {|ψi ⟩}i=0 .
Strictly speaking, in the above assumption, A should be a function of time and denoted as At . However, for the
sake of simplicity and consistency with other notations, we will omit the subscript t.
Based on Assumption 19, we demonstrate the convergence in the following theorem:
23

Theorem 20. • Under Assumption 19, ρ⋆ = |ψ0 ⟩ ⟨ψ0 | is the unique fixed point of (3) in the expectation sense.
In particular,
lim E(ρ(t)) = |ψ0 ⟩ ⟨ψ0 | .
t→∞

We put the proof of Theorem 20 in Appendix D.1. It is important to highlight that while the expected operator
E(ρ(t)) shares a conceptual connection with ρ(t), they are not identical. Nonetheless, this connection provides valuable
insights into ergodicity, offering a promising perspective. For more details, please refer to Section D.2.
Next, we study the convergence speed (or mixing time) of the continuous-time Lindblad dynamics (3). According
to the proof of Theorem 20 in Appendix D.1, we define p(t) = (p0 (t), p1 (t), . . . , pN −1 (t))⊤ , then the solution of (3)
satisfies
N −1
dp(t) X
= Tp(t), E(ρ(t)) = pi (t) |ψi ⟩ ⟨ψi | . (D1)
dt i=0

Here the transition matrix elements are


N
X −1
Tj,i = fˆj,i
2
σj,i , i ̸= j, and Ti,i = − fˆj,i
2
σj,i . (D2)
j̸=i

We are now prepared to demonstrate that the solution of the continuous-time Lindblad dynamics (3) rapidly
converges to low-energy states, as proven by the following theorem:
Theorem 21 (Polynomial mixing time). Together with Assumption 19, we also assume that there exists a decreasing
sequence {Rl }L
l=1 with L = O(poly(n)) such that

• R1 = N − 1, and RL = O(poly(n)).
• For each 1 ≤ l ≤ L − 1
Rl+1
X
fˆi,j
2
σi,j = Ω(1/poly(n)), for all j ∈ (Rl+1 , Rl ]. (D3)
i=0

Then, there exists T ⋆ = O(poly(n)) such that


RL
! RL
X X
E ⟨ψi | ρ(t) |ψi ⟩ = pi (t) = Ω(1), for all t > T ⋆ . (D4)
i=0 i=0
PN
for any initial condition p(0) such that pi (0) ≥ 0 and i=1 pi (0) = 1.
We would like to emphasize that condition (D3) is crucial in ensuring that the weight in pRl+1 +1≤i≤Rl can be
efficiently transferred to pi≤Rl +1 within O(poly(n)) time. Since this transfer only needs to occur L = O(poly(n))
times, the total time required for the weight translation from pRL +1≤i to pi≤RL remains O(poly(n)). To provide a
clear visual representation of the transition matrix’s structure, we have included its graph in (5).
The theorem above implies that, given appropriate assumptions, the continuous-time Lindblad dynamics (3) can
rapidly drive the quantum state towards low-energy eigenstates, resulting in an increased overlap with them. Now,
if we assume that the mixing time of the sub-transition matrix T(1:RL ,1:RL ) ∈ RRL ×RL is tsub
mix = O(1/poly(n)), then
the mixing time for preparing the ground state is tmix = O(poly(n)). In other words, when t > tmix , the overlap with
the ground state E (⟨ψ0 | ρ(t) |ψ0 ⟩) = a0 (s) = Ω(1).

D.1. Proof of Theorem 20 and Theorem 21

Proof of Theorem 20. Since fˆ(ω) = 0 for ω ≥ 0, we have LK [|ψ0 ⟩ ⟨ψ0 |] = 0. Plugging this into (3), we prove that
ρ⋆ = |ψ0 ⟩ ⟨ψ0 | is a fixed point of (3).
Now, we assume A, ρ(0) satisfy Assumption 19. To track the evolution of E(ρ(t)), we first observe,
N
X −1 N
X −1
K |ψi ⟩ = fˆj,i Aj,i |ψj ⟩ , K † |ψi ⟩ = fˆi,j Aj,i |ψj ⟩ ,
j=0 j=0
24

Figure 5. Structure of the transition matrix T in the eigenbasis of the Hamiltonian.

where we use A as a Hermitian matrix in the second equality. Taking the expectation on the randomness of A, we
obtain
−1
 NX
E K |ψi ⟩ ⟨ψi | K † = fˆj,i
2
σj,i |ψj ⟩ ⟨ψj | ,
j=0

and
 
N
X −1
E K † K |ψi ⟩ ⟨ψi | = E  fˆj,i Aj,i K † |ψj ⟩ ⟨ψi |

j=0
 
N
X −1
= E fˆj,i Aj,i fj,k Ak,j |ψk ⟩ ⟨ψi |
j,k=0
N
X −1
= fˆj,i
2
σj,i |ψi ⟩ ⟨ψi | .
j=0

These two equalities imply that


N
X −1
E (LK (|ψi ⟩ ⟨ψi |)) = fˆj,i
2
σj,i (|ψj ⟩ ⟨ψj | − |ψi ⟩ ⟨ψi |) . (D5)
j̸=i

−1
Since ρ(0) is a diagonal matrix in the basis of {|ψi ⟩}N
i=0 , plugging (D5) into (3) and taking expectation on both
−1
sides, we find that E(ρ(t)) is always a diagonal matrix in the basis of {|ψi ⟩}N
i=0 and

dE(ρ(t))
= E (LK (ρ(t))) ,
dt
where we use [H, E(ρ(t))] = 0 for any t. According to (D5), we can rewrite the above equation as
N
X −1
E(ρ(t)) = pi (t) |ψi ⟩ ⟨ψi |
i=0

where pi (t) solves


N −1 N −1
dpi (t) X X
= fˆi,j
2
σi,j pj (t) − fˆj,i
2
σj,i pi (t) . (D6)
dt
j̸=i j̸=i
25

In particular,
N −1
da0 (t) X
= fˆ0,j
2
σ0,j pj (t) ,
dt j=1

which implies p0 (t) = 1, pi≥1 (t) = 0 is the unique fixed point. Since σi,j > 0 for all i, j, we conclude that |ψ0 ⟩ ⟨ψ0 | is
the unique fixed point of the Lindblad dynamics (3), and the solution of (3) converges to |ψ0 ⟩ ⟨ψ0 | as t → ∞.

Next, we prove Theorem 21.


Proof of Theorem 21. To show (D4), it suffices to prove that for any 1 ≤ l ≤ L − 1, there exists Tl = O(poly(n)) such
that
N −1
X 1 1
pi (t) ≤ − , for all t > Tl . (D7)
2 l+3
i=Rl+1 +1

Since fˆ(x) = 0 when x ≥ 0, we first notice that T is an upper triangular matrix, which implies the weight can only
move from high energy states to low energy states. Then, for l = 1, using (D3), we obtain
PN −1 R2
! N −1
! N −1
!!
d i=R2 +1 pi (t) X X 1 X
≤− min fˆi,j σi,j
2
pi (t) = O − pi (t) ,
dt j∈(R2 ,R1 ]
i=0
poly(n)
i=R2 +1 i=R2 +1
PN −1
Since i=R2 +1 pi (0) ≤ 1, there exists T1 = O(poly(n)) such that
N −1
X 1
pi (t) < , for all t > T1 . (D8)
4
i=R2 +1

Next, for l = 2, according (D3), we obtain


PN −1 R3
! R2
! R2
!!
d i=R3 +1 pi (t) X X 1 X
≤− min fˆi,j
2
σi,j pi (t) =O − pi (t) ,
dt j∈(R3 ,R2 ]
i=0
poly(n)
i=R3 +1 i=R3 +1
PR3 ˆ2 1
where we use minj∈(R3 ,R2 ] i=0 fi,j σi,j = Ω( poly(n) ) in the inequality. This implies that when T > T1 , we have
   R2
PN −1  1 1 X 1
O − , if pi (t) >

d i=R3 +1 pi (t) 
≤ 20 poly(n) 20 .
dt  i=R3 +1

0, otherwise

PR2 PN −1
Because i=R3 +1 pi (t) ≤ i=R3 +1 pi (t), there exists T2 = max {T1 , O(poly(n))} such that
R2
X 1 1 1
pi (t) < − = , for all t > T2 .
1+3 2+3 20
i=R3 +1

Combining this with (D8), we can obtain that


N −1
X 3 1 1
pi (t) < = − , for all t > T2 .
10 2 2+3
i=R3 +1

We can continue this process to any l ≤ L − 1. Specifically, for any l, we obtain that
   Rl
PN −1  1 1 X 1
d i=Rl+1 +1 pi (t) O − , if pi (t) >

≤ (l − 1 + 3)(l + 3) poly(n) (l − 1 + 3)(l + 3) .
dt  i=Rl+1 +1

0, otherwise

26

Thus, there exists Tl = max {Tl−1 , O(poly(n))} such that

Rl
X 1 1 1
pi (t) < − =
l−1+3 l+3 20
i=Rl+1 +1

and
N −1
X 1 1
pi (t) < −
2 l+3
i=Rl+1 +1

for any t > Tl . This proves (D7).

D.2. Concentration around deterministic dynamics

In this section, we demonstrate that, when the coupling operator A satisfies Assumption 19, our continuous Lindblad
dynamics simulation in Appendix B approximates E (exp(−LH Ss )[ρ(T )]) when τ → 0, where ρ(T ) the solution of the
modified Lindblad dynamics (B31).
For simplicity, we omit the phase shift exp(−LH Ss ) and assume our simulation scheme can fit into the following
classical setting: Given a stopping time T > 0 and a small time step τ > 0 such that M = T /τ ∈ N, we approximate
ρ(nτ ) using ρn , which is defined by

ρn = F(τ, An )ρn−1 , where ρ0 = ρ(0) . (D9)

Here, F(τ, An ) is a linear operator that depends on the random operator An and {An }M
n=1 are independently drawn
from a probability distribution ΞA .
Next, we introduce the Frobenius norm of a matrix, which is defined as
q
∥A∥F := Tr (A† A) .

In addition, given a superoperator L that acts on matrices, the induced F -norm is

∥L∥F := sup ∥L(A)∥F .


∥A∥F ≤1

Now, we are ready to introduce the approximation result, which is summarized in the following theorem:

Theorem 22 (Concentration of a single trajectory around deterministic dynamics). Assuming our scheme possesses
at least first-order accuracy, meaning that there exists a constant CF such that

∥F(τ, A) − exp(LA τ )∥1 ≤ CF τ 2


 n o
almost surely for A drawn from ΞA . Define CΞ = supA∈supp(ΞA ) max {∥LA ∥1 , ∥LA ∥F }. If τ = O min T1 , C1Ξ , we
have
 √ 
E ∥ρM − E(ρ(T ))∥F = O exp(CΞ T )CΞ T τ + CF T τ (D10)

Remark 23. The relationship between the F -norm and the trace norm in (D10) can be expressed as follows:

1
√ ∥ρM − E(ρ(T ))∥1 ≤ ∥ρM − E(ρ(T ))∥F ≤ ∥ρM − E(ρ(T ))∥1 .
N

It is important to highlight that these inequalities are sharp. Therefore, in the worst-case scenario, the upper bound of
∥ρM − E(ρ(T ))∥1 depends exponentially on the number of qubits n. It remains an intriguing question whether there
exists potential for further improvement in the bound of ∥ρM − E(ρ(T ))∥1 .
27

Proof. Define
ρen = (1 + E(LAn )τ ) ρen−1 , where ρe0 = ρ(0) .
Here LAn is defined in (3), and the subscript An indicates its dependence on the random matrix An . We notice that
dE(ρ(t))
= E(LA )E(ρ(t)) .
dt
Because exp(E(LA )t) is a completely positive trace-preserving map for any t ≥ 0,
∥E(ρ(T )) − ρeM ∥F ≤ ∥E(ρ(T )) − ρeM ∥1 = O CΞ2 T τ .

(D11)
Next, we rewrite (D9) as
ρn = (1 + LAn τ ) ρn−1
+ (F(τ, An ) − exp(LAn τ )) ρn−1 (D12)
+ (exp(LAn τ ) − (1 + LAn τ )) ρn−1 .
The second and third terms can be bounded by:
∥(F(τ, An ) − exp(LAn τ )) ρn−1 ∥1 ≤ CF τ 2 , (D13)
and
∥(exp(LAn τ ) − (1 + LAn τ )) ρn−1 ∥1 = O(CΞ2 τ 2 ) . (D14)

Plugging (D13), (D14) into (D12), when τ < ∥LAn ∥−11 , (1 + LAn τ ) are completely positive trace-preserving maps,
and we obtain
ρM − ΠN N 2

n=1 (1 + LAn τ ) ρ0 F ≤ ρM − Πn=1 (1 + LAn τ ) ρ0 1 = O((CF + CΞ )T τ ) . (D15)

N 2
Finally, we borrow idea from [6, Lemma 3.7] to bound E Πn=1 (1 + LAn τ ) ρ0 − ρeM F . First, we rewrite

ΠN
n=1 (1 + LAn τ ) ρ0 − ρ
eM
N −1
= (LAN τ − E (LAN ) τ ) Πn=1 (1 + LAn τ ) ρ0
| {z }
(I)
−1 N −1
+ (1 + E (LAN ) τ ) ΠN

n=1 (1 + LAn τ ) ρ0 − Πn=1 (1 + E (LAn ) τ ) ρ0
| {z }
(II)

The key observation is that only (I) contains the random operator LAN . Thus,
    
† †
E (I) (II) = E E (I) (II) A1:n−1 = 0.

 

and E (II) (I) = 0. These further imply
2
E ΠN eM F = ETr ((I) + (II))† ((I) + (II))
 
n=1 (1 + LAn τ ) ρ0 − ρ
2 2 2 N −1 N −1
 2
=E ∥(I)∥F + E ∥(II)∥F ≤ E ∥(I)∥1 + (1 + M τ )2 E Πn=1 (1 + LAn τ ) ρ0 − Πn=1 (1 + E (LAn ) τ ) ρ0 F
−1 N −1
 2
=4M 2 τ 2 + (1 + M τ )2 E ΠN

n=1 (1 + LAn τ ) ρ0 − Πn=1 (1 + E (LAn ) τ ) ρ0

F
−1 N −1 2
≤4M 2 τ 2 + exp(2M τ )E ΠN

n=1 (1 + LAn τ ) ρ 0 − Π n=1 (1 + E (L An ) τ ) ρ0

F
.
Repeating the above calculation iteratively, we obtain
2
E ΠN eM F = O exp(2CΞ T )CΞ2 T τ .

n=1 (1 + LAn τ ) ρ0 − ρ

Therefore,
 2 1/2  √ 
E ∥ρM − ρeM ∥F ≤ E ΠN
n=1 (1 + LAn τ ) ρ0 − ρ
eM F = O exp(CΞ T )CΞ T τ . (D16)

Combining (D11), (D15), and (D16), we prove (D10).


28

1.0 1.5
f( ) |f|
0.8
Sw 1.0 ±Ss
0.6 0 0.5

|f|
f

0.4 0.0

0.2 0.5

0.0 1.0
20 10 0 10 20 6 4 2 0 2 4 6
s
(a) fˆ(ω) (b) f (s)

Figure 6. Illustration of fˆ(ω) and its Fourier transform f (s) following Eq. (E2) and Eq. (E3).

Appendix E: Details of numerical experiments and additional experiments

In our numerical simulations, we consider the following TFIM models with L sites:
L−1
! L
X X
H=− Zi Zi+1 + ZL Z1 − g Xi , (E1)
i=1 i=1

where g is the coupling coefficient, Zi , Xi are Pauli operators for the i-th site and the dimension of H is 2L . In the
numerical test in Numerics results, we set L = 6 and the coupling constant g = 1.2. For the jump operator K, we
utilize a local Hermitian operator A = Z ⊗ I2L−1 and choose fˆ(ω) as follows:
    
ˆ 1 ω+a ω+b
f (ω) := erf − erf , (E2)
2 δa δb
Rω 2
where erf(ω) = √2π 0 e−x dx denotes the error function. The parameters a and δa are chosen to be of the order Sω ,
while b and δb are of the order ∆. The inverse Fourier transform f (s) = R f (Ω)e−iωs dω is given by
R

2 s2
δa 2 s2
δb
e− 4 eias − e− 4 eibs
f (s) = ,. (E3)
2πis
It is worth noting that limt→0 f (s) = a−b
2π is well-defined, and f (s) is a smooth complex function that is approximately
supported on the interval [−Ss , Ss ], where Ss = Θ(δb−1 ). Although (E2) does not strictly satisfy Assumption 7, as
fˆ(ω) is approximately supported in [−2a, 0] and fˆ(0) ≈ 0, our numerical tests have shown that this choice of f still
yields satisfactory results. The graph of f can be found in Figure 6.
For simulating the continuous/discrete-time Lindblad dynamics, we always choose a = 2.5∥H∥, δa = 0.5∥H∥,
b = δb = ∆, Ss = 5/δb , and τs = π/(2a).

E.1. TFIM-4 model

To simulate the Lindblad dynamics with TFIM-4 model, we choose L = 4, g = 1.2 in (E1), and the Hermitian
operator A = Z ⊗ I2L−1 . In our numerical experiment, we set τ = 1 and r = 1 for the discrete-time Lindblad
dynamics simulation, and τ = 0.1 for continuous-time Lindblad dynamics simulation. The stopping times are set
to T = 80 for TFIM-4. Each Lindblad dynamics simulation starts from the initial state with zero overlaps between
the ground state and is repeated 100 times, and we calculate the average energy and overlap with the ground state.
29

0 1.0
Lindblad (exact)
1 Lindblad ( = 0.1)
Lindblad ( = 1, r = 1) 0.8
0
2
0.6

< p0 >
<E>
3
0.4
4
0.2 Lindblad (exact)
5
Lindblad ( = 0.1)
0.0 Lindblad ( = 1, r = 1)
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
time time

(a) Lindblad simulation time vs energy (b) Lindblad simulation time vs overlap

0 1.0
Lindblad ( = 0.1) Lindblad ( = 0.1)
1 Lindblad ( = 1, r = 1) Lindblad ( = 1, r = 1)
0 0.8
2
0.6

< p0 >
<E>

3
0.4
4
0.2
5
0.0
101 102 103 104 101 102 103 104
total Hamiltonian simulation time total Hamiltonian simulation time

(c) Hamiltonian simulation time vs energy (d) Hamiltonian simulation time vs overlap

Figure 7. Continuous vs discrete-time Lindblad dynamics for TFIM-4.

The findings are depicted in Figure 7. Our observations indicate that both Lindblad dynamics exhibit efficient
convergence to the ground state starting from the initial state with zero overlaps. Furthermore, the discrete-time
Lindblad dynamics (τ = 1) can achieve a one-order reduction in total Hamiltonian simulation time compared to the
approximate continuous-time Lindblad dynamics (τ = 0.1).

E.2. Hubbard model

Consider the one-dimensional Hubbard model defined on L spinful sites with open boundary conditions
L−1 L   
1 1
c†j,σ cj+1,σ + U
X X X
H = −t nj,↑ − nj,↓ − .
j=1 σ∈{↑,↓} j=1
2 2

Here cj,σ (c†j,σ ) denotes the fermionic annihilation (creation) operator on the site j with spin σ. ⟨·, ·⟩ denotes sites that
are adjacent to each other. nj,σ = c†j,σ cj,σ is the number operator.
We choose L = 4, t = 1, U = 4, and A = X ⊗ I2L−1 . We set τ = 0.5 and r = 2 for discrete-time Lindblad dynamics
simulation and τ = 0.025 for continuous-time Lindblad dynamics simulation. The stopping times are set to T = 100
and each Lindblad dynamics simulation is repeated 100 times starting from an initial state with zero overlap between
the ground state. The results are presented in Figures 8. In our observations, we find that in both dynamics, the
energy decreases to λ0 and the overlap with the ground state increases to 1. We observe that both the continuous-time
and discrete-time Lindblad dynamics exhibit a fast convergence rate over time and achieve an accurate ground state
construction.
30

Lindblad (exact) 1.0


3
Lindblad ( = 0.05)
2 Lindblad ( = 0.5, r = 2) 0.8
0
1 0.6

< p0 >
<E>
0 0.4

1 0.2 Lindblad (exact)


Lindblad ( = 0.05)
2 0.0 Lindblad ( = 0.5, r = 2)
0 20 40 60 80 100 0 20 40 60 80 100
time time

(a) Lindblad simulation time vs energy (b) Lindblad simulation time vs overlap

1.0 Lindblad ( = 0.05)


3
Lindblad ( = 0.5, r = 2)
0.8
2

1 0.6

< p0 >
<E>

0 0.4

1 Lindblad ( = 0.05) 0.2


Lindblad ( = 0.5, r = 2)
0 0.0
2
102 103 104 105 102 103 104 105
total Hamiltonian simulation time total Hamiltonian simulation time

(c) Hamiltonian simulation time vs energy (d) Hamiltonian simulation time vs overlap

Figure 8. Continuous vs discrete-time Lindblad dynamics for Hubbard-4.

You might also like