Liu Dissertation
Liu Dissertation
Jin-Peng Liu
Doctor of Philosophy, 2022
for certain computational problems. Originally developed for simulating quantum physics,
challenges.
There has been extensive previous work for linear dynamics and discrete models,
including Hamiltonian simulations and systems of linear equations. However, for more
computing is far from well understood. One fundamental challenge is the substantial
difference between the linear dynamics of a system of qubits and real-world systems with
this dissertation, I focus mainly on the design and analysis of quantum algorithms for
differential equations. Systems of linear ordinary differential equations (ODEs) and linear
elliptic partial differential equations (PDEs) appear ubiquitous in natural and social science,
engineering, and medicine. I propose a variety of quantum algorithms based on finite
difference methods and spectral methods for producing the quantum encoding of the
solutions, with an exponential improvement in the precision over previous quantum algorithms.
Nonlinear differential equations exhibit rich phenomena in many domains but are
notoriously difficult to solve. Whereas previous quantum algorithms for general nonlinear
equations have been severely limited due to the linearity of quantum mechanics, I give
the first efficient quantum algorithm for nonlinear differential equations with sufficiently
strong dissipation. I also establish a lower bound, showing that nonlinear differential
time, giving an almost tight classification of the quantum complexity of simulating nonlinear
dynamics.
Overall, utilizing advanced linear algebra techniques and nonlinear analysis, I attempt
to build a bridge between classical and quantum mechanics, understand and optimize
the power of quantum computation, and discover new quantum speedups over classical
by
Jin-Peng Liu
Advisory Committee:
Professor Andrew M. Childs, Chair/Advisor
Professor Alexey V. Gorshkov
Professor Carl A. Miller
Professor Konstantina Trivisa
Professor Xiaodi Wu
© Copyright by
Jin-Peng Liu
2022
Acknowledgments
well as an amazing advisor from whom I can constantly learn a lot. Without Andrew,
I cannot imagine how I would get through the challenges of my studies. He taught me
all I know about becoming an independent researcher, from solving research problems,
writing papers, to completing a professional presentation. I will never forget the days
when we worked on problems, talked about my career path, and had lunch together on
studies and during my post-doctoral applications. I would like to thank Furong Huang for
serving as my preliminary exam committee member and for discussing machine learning
topics.
The work described in this dissertation is the product of collaborations with many
people. I cherish the opportunities to work with Dong An, Di Fang, Herman Kolden, Hari
Krovi, Tongyang Li, Noah Linden, Nuno Loureiro, Ashley Montanaro, Aaron Ostrander,
Changpeng Shao, Chunhao Wang, Chenyi Zhang, and Ruizhe Zhang. Despite the fact
ii
that most of the collaborations were virtual, I benefited greatly from their diverse research
backgrounds and areas of expertise. I expect to meet everyone in person in the future.
At the University of Maryland, I had the privilege of learning from many fantastic
faculties, including Gorjan Alagic, John Baras, Alexander Barg, Jacob Bedrossian, James
Jabin, Richard La, Brad Lackey, David Levermore, Yi-Kai Liu, Ricardo Nochetto, and
Eitan Tadmor. I appreciate the instructors’ willingness to share knowledge and skills with
At the Joint Center for Quantum Information and Computer Science (QuICS), I had
in-depth interactions with many quantum information colleagues: Chen Bai, Aniruddha
Bapat, Charles Cao, Shouvanik Chakrabarti, Nai-Hui Chia, Ze-Pei Cian, Abhinav Deshpande,
Bill Fefferman, Hong Hao Fu, Andrew Guo, Kaixin Huang, Shih-Han Hung, Jiaqi Leng,
Yuxiang Peng, Eddie Schoute, Yuan Su, Minh Tran, Chiao-Hsuan Wang, Daochen Wang,
Guoming Wang, Xin Wang, Yidan Wang, Xinyao Wu, Penghui Yao, Qi Zhao, Daiwei
Zhu, Guanyu Zhu, and Shaopeng Zhu. We had wonderful days at QuICS. I would also like
to thank Andrea Svejda, the QuICS coordinator. Whenever I was looking for assistance,
I appreciated the support from the QISE-NET Award, which allows me to collaborate
with Microsoft Quantum researchers: Guang Hao Low and Stephen Jordan. Special
thanks to Stephen for his generous assistance during my graduate studies, regardless of
at Amazon Web Sciences (AWS) Center for Quantum Computing in the summer of 2021,
and I am grateful to Fernando Brandao, Earl Campbell, Michael Kastoryano, and Nicola
iii
Pancotti for their mentorship. Over that summer, I had a lot of fun speaking with AWS
scientists: Steve Flammia, Sam McArdle, Ash Milstead, and Martin Schuetz, as well as
talking with my fellow internship students: Alexander Delzell, Hsin-Yuan Huang, Noah
Corolado, and Shenzhen. In the winter of 2019 and the spring of 2020, I appreciated
the hosts from the University of California, Berkeley, the Lawrence Berkeley National
Laboratory, and the Simons Institute for the Theory of Computing. I would like to thank
Lin Lin and Chao Yang for the numerous discussions about many valuable ideas, as well
as the many challenging questions. I greatly appreciated Lin’s strong support during
Simons programs and for inviting me to participate. In the winter of 2021, I had the
pleasure of visiting the MIT Center for Theoretical Physics and Plasma Science and
Fusion Center in Boston. I was delighted to interact with Soonwon Choi, Issac Chuang,
Edward Farhi, Aram Harrow, and Seth Lloyd. Indeed, my graduate studies relied heavily
on Harrow and Lloyd’s results. It is an honor for me to collaborate with and compete
against MIT folks. In the winter of 2021, I was also thrilled to interview the Harvard
Quantum Initiative, where I had the opportunity to speak with Arthur Jaffe, Mikhail
Lukin, and Susannie Yelin. In these years, I also enjoyed attending conferences hosted by
Caltech, Chicago, Corolado, and Shenzhen. I would like to express my gratitude to many
researchers apart from those mentioned above: Scott Aaronson, Anurag Anshu, Ryan
iv
Babbush, Dominic Berry, Adam Bouland, Sergey Bravyi, Paola Cappellaro, Steve Flammia,
David Gosset, Stuart Hadfield, Matthew Hastings, Patrick Hayden, Zhengfeng Ji, Robin
Kothari, Debbie Liang, Shunlong Luo, Jarrod McClean, Christopher Monroe, Michele
Mosca, John Preskill, Yun Shang, Fang Song, Nikitas Stamatopoulos, Nathan Wiebe,
Thomas Vidick, Beni Yoshida, Henry Yuan, Bei Zeng, William Zeng, and Shenggeng
I would also like to thank my colleagues and friends in the broader quantum information
community: Kaifeng Bu, Chenfeng Cao, Ningping Cao, Lijie Chen, Mo Chen, Yifang
Chen, Rui Chao, Andrea Coladangelo, David Ding, Yulong Dong, Xun Gao, András
Gilyén, Cupjin Huang, Yichen Huang, Hong-Ye Hu, Jiala Ji, Jiaqing Jiang, Ce Jin, Gushu
Li, Jianqiang Li, Yinan Li, Jiahui Liu, Jinguo Liu, Junyu Liu, Mengke Liu, Qipeng Liu,
Yunchao Liu, Yupan Liu, Ziwen Liu, Chuhan Lu, Di Luo, Chinmay Nirkhe, Luowen
Qian, Yihui Quek, Yixin Shen, Qichen Song, Ewin Tang, Yuanjia Wang, Yadong Wu,
Zhujing Xu, Yuxiang Yang, Jiahao Yao, Cong Yu, Zhan Yu, Haimeng Zhang, Jiayu Zhang,
Yuxuan Zhang, Zhendong Zhang, Zijian Zhang, Chen Zhao, Chunlu Zhou, Hengyun
Zhou, Shangnai Zhou, Sisi Zhou, and Jiamin Zhu. I wish you all the best in your future
endeavors.
with fellow graduate students at the Applied Mathematics & Statistics, and Scientific
Christopher Dock, Alexis Boleda, Muhammed Elgebali, Luke Evans, Blake Fritz, Siming
He, Gareth Johnson, Sophie Kessler, Wenbo Li, Ying Li, Yiran Li, Jiaxing Liang, Yuchen
Luo, Jingcheng Lu, Michael Rawson, Tengfei Su, Manyuan Tao, Cem Unsal, Peng Wan,
v
Qiong Wu, Shuo Yang, Anqi Ye, Yiran Zhang, and Yi Zhou. These five years could not
be so enjoyable without you. I would like to thank Jessica Sadler, the AMSC program
coordinator, for offering help during my graduate years. I also enjoyed insightful interactions
with many fellow graduate students at the Department of Computer Science: Jingling Li,
Yanchao Sun, Jiahao Su, and Xuchen You. I would like to thank Jingling in particular for
her thoughtfulness and heartwarming encouragement. Thank you for being there, and I
I would especially thank Yanwen Zhang for giving me the courage to believe in
myself, so that I can complete my Ph.D. and pursue my dream of becoming a successful
Finally, I would like to thank my parents for their endless love, faith, and support.
vi
Table of Contents
Acknowledgements ii
List of Tables ix
List of Figures x
vii
3.3.5 Boundary conditions via the method of images . . . . . . . . . . 83
3.4 Multi-dimensional spectral method . . . . . . . . . . . . . . . . . . . . . 86
3.4.1 Quantum shifted Fourier transform and quantum cosine transform 90
3.4.2 Linear system . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
3.4.3 Condition number . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.4.4 State preparation . . . . . . . . . . . . . . . . . . . . . . . . . . 117
3.4.5 Main result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
3.5 Discussion and open problems . . . . . . . . . . . . . . . . . . . . . . . 122
Bibliography 192
viii
List of Tables
ix
List of Figures
4.1 Integration of the forced viscous Burgers equation using Carleman linearization
on a classical computer (source code available at https://github.
com/hermankolden/CarlemanBurgers). The viscosity is set so
that the Reynolds number Re = U0 L0 /ν = 20. The parameters nx = 16
and nt = 4000 are the number of spatial and temporal discretization
intervals, respectively. The corresponding Carleman convergence parameter
is R = 43.59. Top: Initial condition and solution plotted at a third of
L0
the nonlinear time 13 Tnl = 3U 0
. Bottom: l2 norm of the absolute error
between the Carleman solutions at various truncation levels N (left), and
the convergence of the corresponding time-maximum error (right). . . . . 182
x
Chapter 1: Introduction: quantum scientific computation
classical numerical analysis with advanced quantum technologies to model, analyze, and
solve complex problems arising in physics, chemistry, biology, engineering, social sciences,
and so on. Originally developed for simulating quantum physics, various quantum algorithms
algebra in Hilbert space. For such problems, quantum algorithms are supposed to provide
The basic notations and terminologies of quantum computing are covered in this
the notations and terminologies of quantum computing. More details are available in
1
standard textbooks, such as Nielsen and Chuang, Quantum Computation and Quantum
inner product (the Hilbert space). We usually take N = 2n for some non-negative integer
n. We use the Dirac notation |ψ⟩ to represent a quantum state, and ⟨ϕ| = |ϕ⟩† to represent
its Hermitian conjugation. The scalar ⟨ϕ|ψ⟩ gives the inner product of |ψ⟩ and |ϕ⟩. For
the quantum state, we always assume ⟨ψ|ψ⟩ = 1, i.e. |ψ⟩ is a unit complex vector. We
⟨j|ψ⟩.
Given two quantum states, |ψ1 ⟩ ∈ CN1 , |ψ2 ⟩ ∈ CN2 , their tensor product can be
written as |ψ1 ⟩ |ψ2 ⟩ = |ψ1 ⟩ ⊗ |ψ2 ⟩ ∈ CN1 ×N2 , where ⊗ denotes the Kronecker product.
One quantum bit (one qubit) is a quantum state in C2 , and the tensor product of one-qubit
n
state |ψj ⟩ forms an n-qubit state |ψ1 ⟩ ⊗ . . . ⊗ |ψn ⟩ ∈ C2 .
n ×2n
The n-qubit quantum gate U ∈ C2 is a unitary matrix, i.e. U † U = I, where I
is the identity matrix. It is used to map one n-qubit state to the other, i.e. U : |ψ⟩ → |ψ ′ ⟩.
The universality of two-qubit gates means that every n-qubit gate can be written as
projection operator onto the eigenspace, i.e. Pm2 = Pm , associated the eigenvalue λm . We
2
P
usually assume m Pm = I. When the quantum state |ψ⟩ is measured by M , the outcome
P
is λm with probability pm = ⟨ψ| Pm |ψ⟩, with m pm = 1. After the measurement, the
√
post-state is Pm |ψ⟩ / pm .
n
!1/p
X
∥a∥p := |ak |p . (1.1)
k=1
For a matrix A ∈ Rn×n , we denote the operator norm ∥·∥p,q induced by the vector ℓp and
ℓq norms as
∥Ax∥q
∥A∥p,q := sup , ∥A∥p := ∥A∥p,p . (1.2)
x̸=0 ∥x∥p
Z 1/p
p
∥u(·, t)∥Lp (Ω) := |u(x, t)| dx . (1.4)
Ω
In particular, when no subscript is used in the vector, matrix and function norms, we mean
For real functions f, g : R → R, we write f = O(g) if there exists c > 0, such that
3
|f (τ )| ≤ c|g(τ )| for all τ ∈ R. We write f = Ω(g) if g = O(f ), and f = Θ(g) if both
[3]. The first explicit quantum simulation algorithm was proposed by Lloyd [4] using
product formulas, and numerous quantum algorithms for quantum simulations have been
developed since then[5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23,
24, 25, 26, 27], with various applications ranging from quantum field theory [28, 29] to
quantum chemistry [18, 30, 31] and condensed matter physics [32].
We then assume there is a sparse matrix oracle OH that computes the locations and values
gives the location of the l-th nonzero entry in row j, denoted as k. Then the oracle OH
The quantum oracle OH is reversible, and it allows access to different input elements (j, l)
4
in superposition, which is essential for quantum computing. Here we require the number
of nonzero entries of this matrix in every row and column is at most s, where s is much
We count the number of querying these oracles as the query complexity. If such
oracles can be implemented by one- and two-qubit gates, we usually count the number of
these one- and two-qubit gates as the gate complexity of quantum algorithms.
Hamiltonian system
d
i |ψ(t)⟩ = H(t) |ψ(t)⟩ , |ψ(0)⟩ = |ψin ⟩ . (1.7)
dt
Given the ability to prepare a quantum state |ψ(0)⟩, a sparse matrix oracle to provide the
locations and values of nonzero entries of a Hamiltonian H, and an evolution time T , the
|ψ(T )⟩ = e−iHT |ψ(0)⟩, for which a computation process attempts to produce an evolution
cannot even represent the quantum state efficiently. All classical algorithms require
time complexity at least Ω(N ) to explicitly store all the entries of the quantum state.
circuit) to process |ψ(t)⟩ from |ψ(0)⟩ to |ψ(T )⟩ while lowering the cost dramatically.
5
In Problem 1.1, we aim to produce the final state |ψ(T )⟩ with query and gate
Hj and each Hj acts on k = O(1) qubits, then the evolution operator e−iHt for
P
H= j
Y
e−iHt = e−iHj t + O(t2 ). (1.8)
j
Each operator e−iHj t can be efficiently implemented on a quantum computer. For a long
time t, we can divide the time interval onto r subintervals, in which we simulate each
Y
e−iHt = ( e−iHj t/r )r + O(t2 /r). (1.9)
j
It takes r = O(∥H∥2 t2 /ϵ) to ensure the error tolerance of exact and approximated normalized
states in ℓ2 norm at most ϵ. High-order product formulas are known for a better approximation.
Using the 2kth-order Suzuki formula, Berry, Ahokas, Cleve, and Sanders showed that the
52k ∥H∥1+1/2k t1+1/2k /ϵ1/2k [33]. A large body of work has substantially developed fast
quantum algorithms based on product formulas [33, 34, 35, 36, 37, 38, 39, 40].
have emerged diversely. Berry, Childs, Cleve, Kothari, and Somma proposed high-precision
6
Hamiltonian simulations by implementing the linear combinations of unitaries (LCU)
based Hamiltonian simulations with respect to ϵ. Low and Chuang developed Hamiltonian
simulations based on the quantum signal processing and qubitization [41, 42] with complexity
log(1/ϵ)
O(t + log log(1/ϵ)
), realizing an optimal tradeoff in terms of t and ϵ.
Note that the Hamiltonian simulation problem can be regarded as a particular initial
value problem of the differential equation. A more general problem will be introduced
later.
Harrow, Hassidim, and Lloyd [43], advanced quantum linear system algorithms [43, 44,
45, 46, 47, 48, 49, 50, 51] have been well developed to provide a quantum state encoding
the solution with complexity poly(log N ). Such algorithms have been considerably applied
We introduce the oracles for the quantum linear system problem. Given an N -
that produces an N -dimensional quantum state |b⟩, where |b⟩ is proportional to the vector
defined in (1.6). In detail, on input (j, l), OA gives the location of the l-th nonzero entry
7
in row j, denoted as k. Then the oracle OA gives the value Aj,k . Here we require the
number of nonzero entries of this matrix in every row and column is at most s, where s is
linear system
Ax = b. (1.10)
Given the ability to prepare a quantum state proportional to the vector b, and a sparse
matrix oracle to provide the locations and values of nonzero entries of a matrix A, the
goal is to produce a quantum state that is ϵ-close to the normalized A−1 b in ℓ2 norm.
It takes at least Ω(N ) for a classical computer (and even a quantum computer) to
explicitly write down every entry of an N -dimensional vector. In Problem 1.2, we aim to
design a quantum circuit to provide a quantum state as an encoding of the inverse solution
The first quantum linear system algorithm (QLSA), known as the HHL algorithm,
was proposed by Harrow, Hassidim, and Lloyd [43]. The algorithm requires that A be
Hermitian so that it can be converted into a unitary operator. If A is not Hermitian, the
0 A 0 b
= . (1.11)
A† 0 x 0
The algorithm requires that a quantum state |b⟩ proportional to the vector b is given.
8
Let {λj , |νj ⟩}N
j=1 be eigenvalues and eigenvectors of A, and let |b⟩ be expanded on the
P
eigenbasis of A as |b⟩ = j βj |νj ⟩.
The HHL algorithm queries the unitary operator eiAt with the sparse matrix oracle.
different time t onto |b⟩. The HHL algorithm then estimates the corresponding λj for each
X X
βj |0⟩|νj ⟩ → βj |λj ⟩|νj ⟩, (1.12)
j j
where each λj is stored in the ancilla register. Next, the HHL algorithm performs a
X X
βj |λj ⟩|νj ⟩ → βj λ−1
j |λj ⟩|νj ⟩. (1.13)
j j
This algorithm achieves query and gate complexity κ2 poly(log N )/ϵ, where where
N is the dimension, ϵ measures the error tolerance of exact and approximated normalized
Subsequent work has been widely developed to improve the performance of that
improved the scale of ϵ from poly(1/ϵ) to poly(log(1/ϵ)) by utilizing the linear combinations
1
The HHL algorithm uses the quantum phase estimation (QPE) [62] to estimate the eigenvalues. More
details can be found in [1].
9
of unitaries (LCU) [46]. The LCU approach approximates the inverse of A by employing
of unitary operations with cost poly(log(1/ϵ)). Gilyén, Su, Low, and Wiebe provided an
alternative encoding to approximate the inverse of A based on the quantum singular value
Although the best-known classical linear system solver based on conjugate gradient descent
√
can produce explicit N entries of the solution with cost scaling as κ, it is known that a
quantum algorithm must make at least κ queries for encoding the solution of Problem 1.2
in poly(log N ) [43]. Compared to the HHL algorithm, Ambainis adopted the variable
time amplitude amplification (VTAA) to improve the κ2 scaling to linear scaling[44], but
it resulted in a worse 1/ϵ3 scaling. Childs, Kothari, and Somma combined VTAA with
adiabatic approaches [45, 48, 49, 51] have been recently investigated to reach the same
Models governed by both ordinary differential equations (ODEs) and partial differential
equations (PDEs) arise extensively in natural and social science, medicine, and engineering.
Such equations characterize physical and biological systems that exhibit a wide variety
10
systems of linear ODEs [52, 53, 54] and PDEs [55, 56, 57, 58, 59, 60, 61].
We introduce the oracles for the quantum linear ODE problem. Given an N -
(1.5) that produces an N -dimensional quantum state |u(0)⟩, where |u(0)⟩ is proportional
to the vector u(0). Given an N -dimensional vector f (t) with a specific t, we assume there
N × N sparse matrix A(t) with a specific t, we assume there is a sparse matrix oracle OA
as defined in (1.6). In detail, on input (t, j, l), OA gives the location of the l-th nonzero
entry in row j, denoted as k. Then the oracle OA gives the value Aj,k (t), i.e.
Here we require the number of nonzero entries of this matrix in every row and column is
at most s for any time t, where s is much smaller than the dimension N .
Problem 1.3. In the quantum linear ODE problem, we are given a system of d-dimensional
differential equations
du(t)
= A(t)u(t) + f (t). (1.16)
dt
Given the ability to prepare a quantum state proportional to the initial condition u(0) and
11
the inhomogeneity f (t) with a specific t, a sparse matrix oracle to provide the locations
and values of nonzero entries of a matrix A(t) with a specific t, and an evolution time T ,
the goal is to produce a quantum state that is ϵ-close to the normalized u(T ) in ℓ2 norm.
Berry presented the first quantum algorithm for general linear ODEs [52]. This
work explicitly considered the time-independent case, assuming that real parts of the
into small time intervals as a system of linear equations using the Euler method or high-
order linear multistep methods [63, 64], for which QLSAs can be applied to produce
an approximated encoded solution. For instance, the first-order forward Euler method
The kth-order linear multistep methods can reduce the error to O(hk ). This approach
equation over the interval [0, T ], the number of iterations is T /h = Θ(ϵ−1/k ) for fixed k,
system that encodes a truncated Taylor series. However, this approach assumes that A(t)
and f (t) are time-independent so that the solution of the ODE can be written as an explicit
12
series, and it is unclear how to generalize the algorithm to time-dependent ODEs. Prior to
the work [54], it was an open problem regarding whether there was a quantum algorithm
[54]. Our main contribution is to implement a method that uses a global approximation of
the solution instead of locally discretizing the ODEs into small time intervals. We do this
linear combinations of basis functions ϕj (t) expressing the time dependence. Specifically,
This approach approximates the solution by a truncated Chebyshev series with undetermined
coefficients and solves for those coefficients using a linear system that interpolates the
solution error decreases exponentially provided the solution is sufficiently smooth [67,
68]. We use the LCU-based QLSA to solve this linear system with high precision [46].
To analyze the algorithm, we upper bound the solution error and condition number of the
linear system and lower bound the success probability of the final measurement. Overall,
we show that the total complexity of this approach is poly(log(1/ϵ)) for general time-
dependent ODEs. We give a formal statement of the main result in Theorem 2.1.
13
For solving PDEs on a digital computer, it is common to consider a system of linear
equations that approximates the PDE on the grid space, and produce the solution on those
We introduce the oracles for the quantum linear PDE problem. Given an N -dimensional
vector f (x) with a specific x, we assume there is a state preparation oracle Ob as defined
in (1.14) that produces an N -dimensional quantum state |f (x)⟩, where |f (x)⟩ is proportional
to the vector f (x). Given N × N sparse matrices Aj1 j2 (x), Aj (x), and A0 (x) with
a specific x, we then assume there are sparse matrix oracles as defined in (1.15). For
instance, Aj1 j2 (x) is modeled by a sparse matrix oracle that, on input (m, l), gives the
location of the l-th nonzero entry in row m, denoted as n, and gives the value Aj1 j2 (x)m,n .
Problem 1.4. In the quantum linear PDE problem, we are given a system of second-order
d-dimensional equations
d d
X ∂ 2 u(x) X ∂u(x)
Aj1 j2 (x) + Aj (x) + A0 (x)u(x) = f (x). (1.18)
j1 ,j2 =1
∂xj1 ∂xj2 j=1 ∂xj
Given the ability to prepare a quantum state proportional to the inhomogeneity f (x),
sparse matrix oracles to provide the locations and values of nonzero entries of a matrix
Aj1 j2 (x), Aj (x), and A0 (x) on a set of interpolation nodes x, the goal is to produce
a quantum state |u(x)⟩ that is ϵ-close to the normalized u(x) on a set of interpolation
nodes x in ℓ2 norm.
Problem 3.1 gives a formal statement of Problem 1.4, for which we consider additional
14
The discretized solution u(x) on a set of interpolation nodes x is a multi-dimensional
vector function. If each spatial coordinate has n discrete values, then nd points are needed
to discretize a d-dimensional problem. Simply producing the solution on these grid points
state proportional to the solution on the grid, which requires only poly(d, log n) space.
There are a large variety of quantum algorithms using different numerical schemes and
the QLSA to reach poly(d, 1/ϵ) [55, 56, 57, 58, 59, 60, 61], because of the additional
PDEs with such an improvement [59]. Our first algorithm is based on a quantum version
of linear equations and then solve that system using the QLSA. We analyze our FDM
algorithm as applied to Poisson’s equation under periodic, Dirichlet, and Neumann boundary
conditions. Whereas previous FDM approaches [69, 70] considered fixed orders of truncation,
we adapt the order of truncation depending on ϵ, inspired by the classical adaptive FDM
[71]. As the order increases, the eigenvalues of the FDM matrix approach the eigenvalues
of the continuous Laplacian, allowing for more precise approximations. The main algorithm
we present uses the quantum Fourier transform (QFT) and takes advantage of the high-
precision LCU-based QLSA [46]. This quantum adaptive FDM approach produces a
quantum state that approximates the solution of the Poisson’s equation with complexity
15
We also propose a quantum algorithm for more general second-order elliptic PDEs
the quantum spectral method [54] that globally approximates the solution of a PDE by a
truncated Fourier or Chebyshev series with undetermined coefficients, and then finds the
apply the QLSA efficiently, we show how to make the system sparse using variants of the
quantum Fourier transform. Our bound on the condition number of the linear system uses
global strict diagonal dominance, and introduces a factor in the complexity that measures
the extent to which this condition holds. We give a complexity of d2 poly(log(1/ϵ)) for
here on differential equations with nonlinearities that can be expressed with quadratic
polynomials. Note that polynomials of degree higher than two, and even more general
[72, 73]. The quadratic case also directly includes many archetypal models, such as the
16
logistic equation in biology, the Lorenz system in atmospheric dynamics, and the Navier–
We introduce the oracles for the quantum quadratic PDE problem. Given an N -
(1.5) that produces an N -dimensional quantum state |u(0)⟩, where |u(0)⟩ is proportional
to the vector u(0). Given N × N sparse matrices F2 , F1 , and F0 (t) with a specific t, we
then assume there are sparse matrix oracles as defined in (1.6) and (1.15). For instance,
F2 is modeled by a sparse matrix oracle OF2 that, on input (j, l), gives the location of the
l-th nonzero entry in row j, denoted as k, and gives the value (F2 )j,k .
Problem 1.5. In the quantum quadratic ODE problem, we are given a system of N -
du(t)
= F2 u⊗2 (t) + F1 u(t) + F0 (t). (1.19)
dt
Given the ability to prepare a quantum state proportional to the initial condition u(0) and
sparse matrix oracles to provide the locations and values of nonzero entries of matrices
F2 , F1 , and F0 (t) for any specified t, and an evolution time T , the goal is to produce a
Early work on quantum algorithms for differential equations already considered the
nonlinear case by Leyton and Osborne [74]. It gave a quantum algorithm that simulates
the nonlinear ODE by storing and maintaining multiple copies of the solution. In each
17
iteration from t → t + ∆t, it costs multiple copies |x(t)⟩ to represent the nonlinearity
F (x(t)) to obtain one copy of |x(t + ∆t)⟩. The complexity of this approach is polynomial
in the logarithm of the dimension but exponential in the evolution time, scaling as O(1/ϵT )
due to exponentially increasing resources used to maintain sufficiently many copies of the
Recently, heuristic quantum algorithms for nonlinear ODEs have been studied.
method that might be amenable to the quantum linear system algorithm. In [76], the
authors provide a high-level description of how linearization can help solve nonlinear
preprint [77] also describes a quantum algorithm to solve a nonlinear ODE by linearizing
it using a different approach from the one taken here. However, neither paper makes
algorithms. The authors also do not describe how barriers such as those of [78] could be
of the theory have been widely studied. Generically, such modifications enable quickly
solving hard computational problems (e.g., solving unstructured search among n items
general [78, 79, 80]. Therefore, constructing efficient quantum algorithms for general
Prior to the work [81], it was a long-standing and celebrated open problem regarding
18
whether quantum computing can efficiently characterize nonlinear differential equations.
limitation using Carleman linearization [73, 82, 83]. This approach embeds polynomial
in time using the forward Euler method and solve the resulting linear equations with the
quantum linear system algorithm [46, 84]. We control the approximation error of this
approach by combining a novel convergence theorem with a bound for the global error
of the Euler method. Furthermore, we upper bound the condition number of the linear
system and lower bound the success probability of the final measurement. Subject to the
condition R < 1, where the quantity R characterizes the relative strength of the nonlinear
and dissipative linear terms, we show that the total complexity of this quantum Carleman
is the evolution time, q quantifies the decay of the final solution relative to the initial
condition, n is the dimension, and ϵ is the allowed error (see Theorem 4.1). In the regime
R < 1, this is an exponential improvement over [74], which has complexity exponential
We also provide a quantum lower bound for the worst-case complexity of simulating
strongly nonlinear dynamics, demonstrating that the algorithm’s condition R < 1 cannot
a protocol for distinguishing two states of a qubit driven by a certain quadratic ODE.
√
Provided R ≥ 2, this procedure distinguishes states with overlap 1−ϵ in time poly(log(1/ϵ)).
Since nonorthogonal quantum states are hard to distinguish, this implies a lower bound
19
on the complexity of the quantum ODE problem. We state our main lower bound result
in Theorem 4.2.
by quadratic ODEs arising in biology and epidemiology as well as in fluid and plasma
which describes many physical phenomena, can be treated by our approach provided
the Reynolds number is sufficiently small. We also note that while the formal validity of
our arguments assumes R < 1, we find in one numerical experiment that our proposed
precision quantum algorithms for linear ordinary differential equations. Chapter 3 presents
high-precision quantum algorithms for linear elliptic partial differential equations. Chapter 4
Finally, we conclude the results of the dissertation and discuss future work in Chapter 5.
20
Chapter 2: High-precision quantum algorithms for linear ordinary differential
equations
2.1 Introduction
(ODEs)1 . As earlier introduced in Problem 1.3, such equations can be written in the form
dx(t)
= A(t)x(t) + f (t) (2.1)
dt
where t ∈ [0, T ] for some T > 0, the solution x(t) ∈ Cd is a d-dimensional vector, and
inhomogeneity f (t) ∈ Cd . Provided A(t) and f (t) are continuous functions of t, the
initial value problem (i.e., the problem of determining x(t) for a given initial condition
Recent work has developed quantum algorithms with the potential to extract information
about solutions of systems of differential equations even faster than is possible classically.
This body of work grew from the quantum linear systems algorithm (QLSA) [84], which
21
equations in time poly(log d). We have introduced the quantum linear system algorithm
in Chapter 1.
Berry presented the first efficient quantum algorithm for general linear ODEs [52].
His algorithm represents the system of differential equations as a system of linear equations
using a linear multistep method and solves that system using the QLSA. This approach
close to quadratic in the evolution time T . While this method could in principle be applied
to handle time-dependent equations, the analysis of [52] only explicitly considers the
as a function of the solution error ϵ is poly(1/ϵ) [52]. Reference [53] improved this to
[46] to solve a linear system that encodes a truncated Taylor series. However, this approach
assumes that A(t) and f (t) are time-independent so that the solution of the ODE can be
written as an explicit series, and it is unclear how to generalize the algorithm to time-
dependent ODEs.
the differential equations into small time intervals to obtain a system of linear equations or
linear differential equations that can be solved by the QLSA or Hamiltonian simulation.
For example, the central difference scheme approximates the time derivative at the point
x(t) as
dx(t) x(t + h) − x(t − h)
= + O(h2 ). (2.2)
dt 2h
22
High-order finite difference or finite element methods can reduce the error to O(hk ),
where k − 1 is the order of the approximation. However, when solving an equation over
the interval [0, T ], the number of iterations is T /h = Θ(ϵ−1/k ) for fixed k, giving a
total complexity that is poly(1/ϵ) even using high-precision methods for the QLSA or
Hamiltonian simulation.
For ODEs with special structure, some prior results already show how to avoid
A and f are time-independent, then [53] uses a Taylor series to achieve complexity
poly(log(1/ϵ)). However, the case of general time-dependent linear ODEs had remained
elusive.
dependence on ϵ over previous work, it does not necessarily give an exponential runtime
subroutine k times, we should ensure error O(1/k) for each subroutine to give an overall
algorithm with bounded error. A subroutine with complexity poly(log(1/ϵ)) can potentially
23
science and engineering. Examples include the wave equation and the Stokes equation
(i.e., creeping flow) in fluid dynamics [86], the heat equation and the Boltzmann equation
in thermodynamics [87, 88], the Poisson equation and Maxwell’s equations in electromagnetism
[89, 90], and of course Schrödinger’s equation in quantum mechanics. Moreover, some
linear equations (e.g., the linearized advection equation in fluid dynamics [91]).
PDEs with both time and space dependence can be regarded as sparse linear systems
of time-dependent ODEs. Thus we focus on an equation of the form (2.1) with initial
condition
x(0) = γ (2.3)
for some specified γ ∈ Cd . We assume that A(t) is s-sparse (i.e., has at most s nonzero
entries in any row or column) for any t ∈ [0, T ]. Furthermore, we assume that A(t), f (t),
computations). In particular, following essentially the same model as in [53] (see also
Section 1.1 of [46]), suppose we have an oracle OA (t) that, for any t ∈ [0, T ] and any
given row or column specified as input, computes the locations and values of the nonzero
entries of A(t) in that row or column. We also assume oracles Ox and Of (t) that, for
any t ∈ [0, T ], prepare normalized states |γ⟩ and |f (t)⟩ proportional to γ and f (t), and
that also compute ∥γ∥ and ∥f (t)∥, respectively. Given such a description of the instance,
the goal is to produce a quantum state ϵ-close to |x(T )⟩ (a normalized quantum state
24
proportional to x(T )).
called spectral methods, a technique from classical numerical analysis that (approximately)
P
represents the components of the solution x(t)i ≈ j cij ϕj (t) as linear combinations
Chebyshev pseudospectral method [65, 66] using the QLSA. This approach approximates
the solution by a truncated Chebyshev series with undetermined coefficients and solves
for those coefficients using a linear system that interpolates the differential equations.
According to the convergence theory of spectral methods, the solution error decreases
exponentially provided the solution is sufficiently smooth [67, 68]. We use the LCU-based
QLSA to solve this linear system with high precision [46]. To analyze the algorithm,
we upper bound the solution error and condition number of the linear system and lower
bound the success probability of the final measurement. Overall, we show that the total
Theorem 2.1 (Informal). Consider a linear ODE (2.1) with given initial conditions.
Assume A(t) is s-sparse and diagonalizable, and Re(λi (t)) ≤ 0 for all eigenvalues of
A(t). Then there exists a quantum algorithm that produces a state ϵ-close in l2 norm
to the exact solution, succeeding with probability Ω(1), with query and gate complexity
In addition to initial value problems (IVPs), our approach can also address boundary
25
value problems (BVPs). Given an oracle for preparing a state α|x(0)⟩+β|x(T )⟩ expressing
a general boundary condition, the goal of the quantum BVP is to produce a quantum state
ϵ-close to |x(t)⟩ (a normalized state proportional to x(t)) for any desired t ∈ [0, T ].
We also give a quantum algorithm for this problem with complexity poly(log(1/ϵ)), as
follows:
Theorem 2.2 (Informal). Consider a linear ODE (2.1) with given boundary conditions.
Assume A(t) is s-sparse and diagonalizable, and Re(λi (t)) ≤ 0 for all eigenvalues of
A(t). Then there exists a quantum algorithm that produces a state ϵ-close in l2 norm
to the exact solution, succeeding with probability Ω(1), with query and gate complexity
We give formal statements of Theorem 2.1 and Theorem 2.2 in Section 2.8 and
Section 2.9, respectively. Note that the dependence of the complexity on ∥A∥ and T is
worse for BVPs than for IVPs. This is because a rescaling approach that we apply for
The remainder of this chapter is organized as follows. Section 2.2 introduces the
spectral method and Section 2.3 shows how to encode it into a quantum linear system.
Then Section 2.4 analyzes the exponential decrease of the solution error, Section 2.5
bounds the condition number of the linear system, Section 2.6 lower bounds the success
probability of the final measurement, and Section 2.7 describes how to prepare the initial
quantum state. We combine these bounds in Section 2.8 to establish the main result.
We then extend the analysis for initial value problems to boundary value problems in
Section 2.9. Finally, we conclude in Section 2.10 with a discussion of the results and
26
some open problems.
approximations [67, 68]. The main idea of the approach is as follows. First, express an
solution should satisfy. Finally, solve the linear system to determine the coefficients of
Spectral methods offer a flexible approach that can be adapted to different settings
by careful choice of the basis functions and the linear system. A Fourier series provides an
appropriate basis for periodic problems, whereas Chebyshev polynomials can be applied
more generally. The linear system can be specified using Gaussian quadrature (giving a
spectral element method or Tau method), or one can simply interpolate the differential
equations using quadrature nodes (giving a pseudo-spectral method) [68]. Since general
n
X
xi (t) = ci,k Tk (t), i ∈ [d]0 := {0, 1, . . . , d − 1} (2.4)
k=0
27
for any n ∈ Z+ . Here Tk (t) = cos(k arccos x) is the Chebyshev polynomial of the first
kind. (See [54, Appendix A] for its properties.) The coefficients ci,k ∈ C for all i ∈ [d]0
and k ∈ [n + 1]0 are determined by demanding that x(t) satisfies the ODE and initial
conditions at a set of interpolation nodes {tl }nl=0 (with 1 = t0 > t1 > · · · > tn = −1),
where x(t0 ) and x(tn ) are the initial and final states, respectively. In other words, we
require
dx(tl )
= A(tl )x(tl ) + f (tl ), ∀ l ∈ [n + 1], t ∈ [−1, 1], (2.5)
dt
and
We choose the domain [−1, 1] in (2.5) because this is the natural domain for Chebyshev
value problems to be [−1, 1]. We would like to be able to increase the accuracy of the
There are many possible choices for the interpolation nodes. Here we use the
nodes achieve the highest convergence rate among all schemes with the same number
of nodes [63, 64]. These nodes also have the convenient property that Tk (tl ) = cos klπ
n
,
To evaluate the condition (2.5), it is convenient to define coefficients c′i,k for i ∈ [d]0
28
and k ∈ [n + 1]0 such that
n
dxi (t) X ′
= ci,k Tk (t). (2.8)
dt k=0
′ ′
Tk+1 (t) Tk−1 (t)
2Tk (t) = − , (2.9)
k+1 k−1
to determine the transformation between ci,k and c′i,k . Based on the property of derivatives
n
X
c′i,k = [Dn ]kj ci,j , i ∈ [d]0 , k ∈ [n + 1]0 , (2.10)
j=0
2j
[Dn ]kj = , k + j odd, j > k, (2.11)
σk
where
2 k = 0
σk := (2.12)
1 k ∈ [n] := {1, 2, . . . , n}.
Using this expression in (2.5), (2.10), and (2.11), we obtain the following linear
equations:
n
X d−1
X n
X
Tk (tl )c′i,k = Aij (tl ) Tk (tl )cj,k + f (tl )i , i ∈ [d]0 , l ∈ [n + 1]0 . (2.13)
k=0 j=0 k=0
29
We also demand that the Chebyshev series satisfies the initial condition xi (1) = γi for all
i ∈ [d]0 . This system of linear equations gives a global approximation of the underlying
times, these linear equations use the behavior of the differential equations at the n + 1
times {tl }nl=0 to capture their behavior over the entire interval [−1, 1].
Our algorithm solves this linear system using the high-precision QLSA [46]. Given
an encoding of the Chebyshev coefficients cik , we can obtain the approximate solution
x(t) as a suitable linear combination of the cik , a computation that can also be captured
within a linear system. The resulting approximate solution x(t) is close to the exact
solution x̂(t):
Lemma 2.1 (Lemma 19 of [67]). Let x̂(t) ∈ C r+1 (−1, 1) be the solution of the differential
equations (2.1) and let x(t) satisfy (2.5) and (2.6) for {tl = cos lπ }n . Then there is a
n l=0
∥x̂(n+1) (t)∥
max ∥x̂(t) − x(t)∥ ≤ C max . (2.14)
t∈[−1,1] t∈[−1,1] nr−2
This shows that the convergence behavior of the spectral method is related to the
smoothness of the solution. For a solution in C r+1 , the spectral method approximates the
tighter bound:
Lemma 2.2 (Eq. (1.8.28) of [68]). Let x̂(t) ∈ C ∞ (−1, 1) be the solution of the differential
30
equations (2.1) and let x(t) satisfy (2.5) and (2.6) for {tl = cos lπ }n . Then
n l=0
r
2 e n
max ∥x̂(t) − x(t)∥ ≤ max ∥x̂(n+1) (t)∥ . (2.15)
t∈[−1,1] π t∈[−1,1] 2n
p
For simplicity, we replace the value 2/π by the upper bound of 1 in the following
analysis.
This result implies that if the solution is in C ∞ , the spectral method approximates
the solution to within ϵ using only n = poly(log(1/ϵ)) terms in the Chebyshev series.
In this section we construct a linear system that encodes the solution of a system of
differential equations via the Chebyshev pseudospectral method introduced in Section 2.2.
and focus on the following initial value problem. This is a formal statement of Problem 1.3.
Problem 2.1. In the quantum ODE problem, we are given a system of equations
dx(t)
= A(t)x(t) + f (t) (2.16)
dt
where x(t) ∈ Cd , A(t) ∈ Cd×d is s-sparse, and f (t) ∈ Cd for all t ∈ [0, T ]. We
assume that Aij , fi ∈ C ∞ (0, T ) for all i, j ∈ [d]. We are also given an initial condition
x(0) = γ ∈ Cd . Given oracles that compute the locations and values of nonzero entries
31
of A(t) for any t2 , and that prepare normalized states |γ⟩ proportional to γ and |f (t)⟩
proportional to f (t) for any t ∈ [0, T ], the goal is to output a quantum state |x(T )⟩ that
Without loss of generality, we rescale the interval [0, T ] onto [−1, 1] by the linear
d
map t 7→ 1 − 2t/T . Under this rescaling, we have dt
7→ − T2 dt
d
, so A 7→ − T2 A, which
can increase the spectral norm. To reduce the dependence on T —specifically, to give
an algorithm with complexity close to linear in T —we divide the interval [0, T ] into
[Γh , Γh+1 ] for h ∈ [m]0 is then rescaled onto [−1, 1] with the linear map Kh : [Γh , Γh+1 ] →
[−1, 1] defined by
2(t − Γh )
Kh : t 7→ 1 − , (2.17)
Γh+1 − Γh
which satisfies Kh (Γh ) = 1 and Kh (Γh+1 ) = −1. To solve the overall initial value
problem, we simply solve the differential equations for each successive interval (as encoded
τh
Ah (t) := − A(Kh (t)) (2.18)
2
xh (t) := x(Kh (t)) (2.19)
τh
fh (t) := − f (Kh (t)). (2.20)
2
2
A(t) is modeled by a sparse matrix oracle OA that, on input (j, l), gives the location of the l-th nonzero
entry in row j, denoted as k, and gives the value A(t)j,k .
32
Then, for each h ∈ [m]0 , we have the rescaled differential equations
dxh
= Ah (t)xh (t) + fh (t) (2.21)
dt
γ h=0
xh (1) = (2.22)
xh−1 (−1) h ∈ [m].
By taking
2
τh ≤ (2.23)
maxt∈[Γh ,Γh+1 ] ∥A(t)∥
where ∥·∥ denotes the spectral norm, we can ensure that ∥Ah (t)∥ ≤ 1 for all t ∈ [−1, 1].
2
τ := max τh ≤ . (2.24)
h∈{0,1,...,m−1} maxt∈[0,T ] ∥A(t)∥
Having rescaled the equations to use the domain [−1, 1], we now apply the Chebyshev
series of x(t) into the differential equations with interpolating nodes {tl = cos lπ
n
:l ∈
dx(tl )
= Ah (tl )x(tl ) + fh (tl ), h ∈ [m]0 , l ∈ [n + 1] (2.25)
dt
33
with initial condition
x(t0 ) = γ. (2.26)
Note that in the following, terms with l = 0 refer to this initial condition.
The vector |X⟩ ∈ Cm+p ⊗ Cd ⊗ Cn+1 represents the solution in the form
m−1 d−1 X
n m+p d−1 n
XX X XX
|X⟩ = ci,l (Γh+1 )|hil⟩ + xi |hil⟩ (2.28)
h=0 i=0 l=0 h=m i=0 l=0
where ci,l (Γh+1 ) are the Chebyshev series coefficients of x(Γh+1 ) and xi := x(Γm )i is the
The right-hand-side vector |B⟩ represents the input terms in the form
m−1
X
|B⟩ = |h⟩|B(fh )⟩ (2.29)
h=0
where
d−1
X d−1 X
X n
|B(fh )⟩ = γi |i0⟩ + fh (cos lπ ) |il⟩,
n i
h ∈ [m − 1]. (2.30)
i=0 i=0 l=1
Here γ is the initial condition and fh (cos lπ ) is ith component of fh at the interpolation
n i
point tl = cos lπ
n
.
34
We decompose the matrix L in the form
dx
The matrix L1 is a discrete representation of dt
, satisfying
d−1 X
X n d−1
X n
X
|h⟩⟨h| ⊗ L1 |X⟩ = Tk (t0 )ci,k |hi0⟩ + Tk (tl )[Dn ]kr ci,r |hil⟩ (2.32)
i=0 k=0 i=0 l=1,k,r=0
(recall from (2.8) and (2.10) that Dn encodes the action of the time derivative on a
d−1 X
n d−1 n
X X X klπ
L1 = Tk (t0 )|i0⟩⟨ik| + cos [Dn ]kr |il⟩⟨ir| (2.33)
i=0 k=0 i=0 l=1,k,r=0
n
n
X
= Id ⊗ (|0⟩⟨0|Pn + |l⟩⟨l|Pn Dn ) (2.34)
l=1
n
X klπ
Pn := cos |l⟩⟨k|. (2.35)
l,k=0
n
d−1 X
X n
|h⟩⟨h| ⊗ L2 (Ah )|X⟩ = − Ah (tl )ij Tk (tl )cj,k |hil⟩. (2.36)
i,j=0 l=1,k=0
35
Thus
d−1 X
n
X klπ
L2 (Ah ) = − Ah (tl )ij cos |il⟩⟨jk| (2.37)
i,j=0 l=1,k=0
n
n
X
=− Ah (tl ) ⊗ |l⟩⟨l|Pn . (2.38)
l=1
The matrix L3 combines the Chebyshev series coefficients ci,l to produce xi for
each i ∈ [d]0 . To express the final state x(−1), L3 represents the linear combination
Pn Pn k
xi (−1) = k=0 ci,k Tk (−1) = k=0 (−1) ci,k . Thus we take
d−1 X
X n
L3 = (−1)k |i0⟩⟨ik|. (2.40)
i=0 k=0
repeat xi n times for l ∈ [n]. When m+1 ≤ h ≤ m+p, both L4 and L5 are used to repeat
xi (n + 1)p times for l ∈ [n]. This repetition serves to increase the success probability of
d−1 X
X n d−1 X
X n
L4 = − |il⟩⟨il − 1| + |il⟩⟨il| (2.41)
i=0 l=1 i=0 l=0
36
and
d−1
X
L5 = − |i0⟩⟨in|. (2.42)
i=0
In summary, the linear system is as follows. For each h ∈ [m]0 , (L1 +L2 (Ah ))|X⟩ =
|Bh ⟩ solves the differential equations over [Γh , Γh+1 ], and the coefficients ci,l (Γh+1 ) are
combined by L3 into the (h + 1)st block as initial conditions. When h = m, the final
coefficients ci,l (Γm ) are combined by L3 and L4 into the final state with coefficients xi ,
In this section, we bound how well the solution of the linear system defined above
Lemma 2.3. For the linear system L|X⟩ = |B⟩ defined in (2.27), let x be the approximate
ODE solution specified by the linear system and let x̂ be the exact ODE solution. Then
(n+1) en+1
∥x̂(T ) − x(T )∥ ≤ m max ∥x̂ (t)∥ . (2.43)
t∈[0,T ] (2n)n
e log(ω)
n≥ (2.44)
2 log(log(ω))
37
where
∥x̂(n+1) (t)∥
ω := max (m + 1) (2.45)
t∈[0,T ] ∥γ∥
to ensure that
∥x̂(n+1) (t)∥ e n 1
max ≤ . (2.46)
t∈[0,T ] ∥γ∥ 2n m+1
dx
= Ah (t)x(t) + fh (t), h ∈ [m]0 . (2.47)
dt
Pd Pn n
We denote the exact solution by x̂(Γh+1 ), and we let x(Γh+1 ) = i=0 l=0 (−1) ci,l (Γh+1 ),
e n
∆1 = ∥x̂(Γ1 ) − x(Γ1 )∥ ≤ max ∥x̂(n+1) (t)∥ . (2.49)
t∈[0,T ] 2n
contributions: the error from the linear system and the error in the initial condition. We
e(Γh+1 ) denote the solution of the linear system L1 + L2 (Ah ) |e
let x x(Γh+1 )⟩ = |B(fh )⟩
∆h+1 ≤ ∥x̂(Γh+1 ) − x
e(Γh+1 )∥ + ∥e
x(Γh+1 ) − x(Γh+1 )∥. (2.50)
38
The first term can be bounded using Lemma 2.2, giving
e n
e(Γh+1 )∥ ≤ max ∥x̂(n+1) (t)∥
∥x̂(Γh+1 ) − x . (2.51)
t∈[0,T ] 2n
The second term comes from the initial error ∆h , which is transported through the linear
system. Let
where Eh+1 is the solution of the linear system with input ∆h and Êh+1 is the exact
dx
solution of dt
= Ah+1 (t)x(t) + fh+1 (t) with initial condition x(Γh ) = ∆h . Then by
Lemma 2.2,
∆h (n+1)
e n
∥δh+1 ∥ = ∥Êh+1 − Eh+1 ∥ ≤ max ∥x̂ (t)∥ , (2.53)
∥γ∥ t∈[0,T ] 2n
so
∆h e n
∥e
x(Γh+1 ) − x(Γh+1 )∥ ≤ ∆h + max ∥x̂(n+1) (t)∥ . (2.54)
∥γ∥ t∈[0,T ] 2n
∥x̂(n+1) (t)∥ e n 1 1
max ≤ ≤ , (2.56)
t∈[0,T ] ∥γ∥ 2n m+1 m
39
so
∥x̂(n+1) (t)∥ e n m−1 1 m
1 + max ≤ 1+ ≤ e. (2.57)
t∈[0,T ] ∥γ∥ 2n m
Therefore
which shows that the solution error decreases exponentially with n. In other words, the
Note that for time-independent differential equations, we can directly estimate ∥x̂(n+1) (t)∥
using
Thus the exact solution of time-independent equation with initial condition x̂(1) = γ is
40
Since Re(λi ) ≤ 0 for all eigenvalues λi of Ah for i ∈ [d]0 , we have ∥eΛh ∥ ≤ 1. Therefore
≤ κV (∥γ∥ + 2τ ∥f ∥).
en+1
∥x̂(T ) − x(T )∥ ≤ mκV (∥γ∥ + 2τ ∥f ∥) . (2.63)
(2n)n
Note that, although we represent the solution differently, this bound is similar to the
Lemma 2.4. Consider an instance of the quantum ODE problem as defined in Problem 2.1.
For all t ∈ [0, T ], assume A(t) can be diagonalized as A(t) = V (t)Λ(t)V −1 (t) for
some Λ(t) = diag(λ0 (t), . . . , λd (t)), with Re(λi (t)) ≤ 0 for all i ∈ [d]0 . Let κV :=
maxt∈[0,T ] κV (t) be an upper bound on the condition number of V (t). Then for m, p ∈ Z+
41
and n sufficiently large, the condition number of L in the linear system (2.27) satisfies
Proof. We begin by bounding the norms of some operators that appear in the definition
n(n+2) n even,
n
X 2
∥Dn ∥∞ := max |[Dn ]ij | = (2.65)
1≤i≤n
j=0
(n+1)2 − 2 n odd.
2
√ (n + 1)2.5
∥Dn ∥ ≤ n + 1∥Dn ∥∞ ≤ . (2.66)
2
Next we upper bound the spectral norm of the discrete cosine transform matrix Pn :
n
X klπ
∥Pn ∥2 ≤ max cos2 ≤ max {n + 1} = n + 1. (2.67)
0≤l≤n
k=0
n 0≤l≤n
Therefore
√
∥Pn ∥ ≤ n + 1. (2.68)
(n + 1)3
∥L1 ∥ ≤ ∥Dn ∥∥Pn ∥ ≤ . (2.69)
2
42
Next we consider the spectral norm of L2 (Ah ) for any h ∈ [m]0 . We have
n
X
L2 (Ah ) = − Ah (tl ) ⊗ |l⟩⟨l|Pn . (2.70)
l=1
Since the eigenvalues of each Ah (tl ) for l ∈ [n + 1]0 are all eigenvalues of
n
X
Ah (tl ) ⊗ |l⟩⟨l|, (2.71)
l=0
we have
n
X n
X
Ah (tl ) ⊗ |l⟩⟨l| ≤ Ah (tl ) ⊗ |l⟩⟨l| ≤ max ∥Ah (t)∥ ≤ 1 (2.72)
t∈[−1,1]
l=1 l=0
by (2.23). Therefore
√
∥L2 (Ah )∥ ≤ ∥Pn ∥ ≤ n + 1. (2.73)
√
∥L3 ∥ = n + 1, (2.74)
∥L4 ∥ ≤ 2, (2.75)
∥L5 ∥ = 1. (2.76)
(n + 1)3 √ √
∥L∥ ≤ + n + 1 + n + 1 + 2 + 1 ≤ (n + 1)3 . (2.77)
2
43
Next we upper bound ∥L−1 ∥. By definition,
We express |B⟩ as
m+p
n X
d−1 m+p
n
XX XX
|B⟩ = βhil |hil⟩ = |bhl ⟩ (2.79)
h=0 l=0 i=0 h=0 l=0
Pd−1 Pd−1
where |bhl ⟩ := i=0 βhil |hil⟩ satisfies ∥|bhl ⟩∥2 = i=0 |βhil |2 ≤ 1. For any fixed
h ∈ [m + p + 1]0 and l ∈ [n + 1]0 , we first upper bound ∥L−1 |bhl ⟩∥ and use this to
upper bound the norm of L−1 applied to linear combinations of such vectors.
Recall that the linear system comes from (2.13), which is equivalent to
n
X d−1
X n
X
Tk′ (tr )ci,k (Γh ) = Ah (tr )ij Tk (tr )cj,k (Γh ) + fh (tr )i , i ∈ [d]0 , r ∈ [n + 1]0 .
k=0 j=0 k=0
(2.80)
For fixed h ∈ [m + p + 1]0 and r ∈ [n + 1]0 , define vectors xhr , x′hr ∈ Cd with
n
X n
X
(xhr )i := Tk (tr )ci,k (Γh ), (x′hr )i := Tk′ (tr )ci,k (Γh ) (2.81)
k=0 k=0
for i ∈ [d]0 . We claim that xhr = x′hr = 0 for any r ̸= l. Combining only the equations
44
Consider a corresponding system of differential equations
dx̂hr (t)
= Ah (tr )x̂(t) + b (2.83)
dt
where x̂hr (t) ∈ Cd for all t ∈ [−1, 1]. The solution of this system with b = 0 and initial
condition x̂hr (1) = 0 is clearly x̂hr (t) = 0 for all t ∈ [−1, 1]. Then the nth-order
truncated Chebyshev approximation of (2.83), which should satisfy the linear system
(2.82) by (2.4) and (2.5), is exactly xhr . Using Lemma 2.3 and observing that x̂(n+1) (t) =
0, we have
When t = tl , we let |B⟩ = |bhl ⟩ denote the first nonzero vector. Combining only
dx̂hr (t)
= Ah (tr )x̂(t) + b, (2.86)
dt
Using the diagonalization Ah (tl ) = V (tl )Λh (tl )V −1 (tl ), we have eA = V (tl )eΛh (tl ) V −1 (tl ).
Thus the exact solution of the differential equations (2.83) with r = l and initial condition
45
x̂hr (1) = γ is
x̂hr (t) = eAh (tl )(1−t) γ + (eAh (tl )(1−t) − I)Ah (tl )−1 b
(2.87)
Λh (tl )(1−t) −1 Λh (tl )(1−t) −1 −1
= V (tl )e V (tl )γ + V (e − I)Λh (tl ) V b.
where
(n+1) en+1 e∥γ∥
∥δhl ∥ ≤ max ∥x̂hl (t)∥ n
≤ . (2.89)
t∈[0,T ] (2n) m+1
Now for h ∈ [m + 1]0 , we take xhl to be the initial condition γ for the next
subinterval to obtain x(h+1)l . Using (2.87) and (2.88), starting from γ = bh0 , b = 0
for l = 0, we find
m−h+1
Y m−h
X k
Y
−1
xml = V (tl ) e 2Λh (tl )
V (tl )γ+ V (tl ) e 2Λh (tl )
V −1 (tl )δ(m−k)l . (2.90)
j=1 k=0 j=1
Since ∥Λh (tl )∥ ≤ ∥Λ∥ ≤ 1 and Λh (tl ) = diag(λ0 , . . . , λd−1 ) with Re(λi ) ≤ 0 for
∥xhl ∥ ≤ ∥xml ∥ ≤ κV (tl )∥bhl ∥ + (m − h + 1)κV (tl )∥δhl ∥ ≤ κV (tl ) + e∥γ∥ ≤ κV + e∥γ∥.
(2.91)
46
On the other hand, with γ = 0, b = bhl for l ∈ [n], we have
m−h
Y
2Λh (tl )
(e2Λh (tl ) − I)Λh (tl )−1 V −1 (tl )b
xml = V (tl ) e
j=1
m−h k
Y (2.92)
X
+ V (tl ) e2Λh (tl )
V −1 (tl )δ(m−k)l ,
k=0 j=1
so
∥xhl ∥ ≤ 2κV (tl )∥bhl ∥+(m−h+1)κV (tl )∥δhl ∥ ≤ 2κV (tl )+e∥γ∥ ≤ 2κV +e∥γ∥. (2.93)
have
According to (2.87), x̂hl (t) is a monotonic function of t ∈ [−1, 1], which implies
we have
Z 1 Z 1
2 dt dt
∥x̂hl (t)∥ √ ≤ (2κV + e∥γ∥)2 √ = π(2κV + e∥γ∥)2 . (2.97)
1 − t 2 1 − t2
−1 −1
47
Consider the Chebyshev expansion of x̂hl (t) as in (2.4):
∞
d−1 X
X
x̂hl (t) = ci,l (Γh+1 )Tl (t). (2.98)
i=0 l=0
have
Z 1 Z 1 X ∞
d−1 X 2
2 dt dt
∥x̂hl (t)∥ √ = ci,l (Γh+1 )Tl (t) √
−1 1 − t2 −1 i=0 l=0
1 − t2
∞
d−1 X d−1 d−1 X
n d−1
(2.99)
X X X X
= c2i,l (Γh+1 ) + 2 c2i,0 (Γh+1 ) ≥ c2i,l (Γh+1 ) + 2 c2i,0 (Γh+1 ).
i=0 l=1 i=0 i=0 l=1 i=0
d−1 X
n Z 1
X dt
c2i,l (Γh+1 ) ≤ x̂2hl (t) √ ≤ π(2κV + e∥γ∥)2 . (2.100)
1−t 2
i=0 l=0 −1
Now we compute ∥|X⟩∥, summing the contributions from all ci,r (Γh ) and xmr , and
m−1
XX d−1
∥|X⟩∥2 = c2i,l (Γh+1 ) + (p + 1)(xml )2
h=0 i=0
(2.101)
≤ πm(2κV + e∥γ∥)2 + (p + 1)(κV + e∥γ∥)2
m+p
n
XX
2
∥|B⟩∥ = ∥|bhl ⟩∥2 ≤ 1, (2.102)
h=0 l=0
48
so
m+p
n
XX
−1 2 −1
∥L ∥ = sup ∥L |B⟩∥ = sup 2
∥L−1 |bhl ⟩∥2
∥|B⟩∥≤1 ∥|B⟩∥≤1 h=0 l=0
(2.103)
≤ (πm + p + 1)(m + p + 1)(n + 1)(2κV + e∥γ∥)2
and therefore
as claimed.
We now evaluate the success probability of our approach to the quantum ODE
problem.
Lemma 2.5. Consider an instance of the quantum ODE problem as defined in Problem 2.1
with the exact solution x̂(t) for t ∈ [0, T ], and its corresponding linear system (2.27)
with m, p ∈ Z+ and n sufficiently large. When applying the QLSA to this system, the
Pd−1
probability of measuring a state proportional to |x(T )⟩ = i=0 xi |i⟩ is
(p + 1)(n + 1)
Pmeasure ≥ , (2.106)
πmq 2 + (p + 1)(n + 1)
49
where xi is defined in (2.28), τ is defined in (2.24), and
∥x̂(t)∥
q := max . (2.107)
t∈[0,T ] ∥x(T )∥
Proof. After solving the linear system (2.27) using the QLSA, we measure the first and
where
m−1
XX d−1 X
n
|Xbad ⟩ = ci,l (Γh+1 )|hil⟩, (2.109)
h=0 i=0 l=0
m+p d−1 n
X XX
|Xgood ⟩ = xi |hil⟩. (2.110)
h=m i=0 l=0
what outcome is seen for the third register), we output the second register, which is then
|x(T )⟩
|Xmeasure ⟩ = , (2.111)
∥|x(T )⟩∥
with
d−1
X d−1 X
X n
|x(T )⟩ = xi |i⟩ = ci,k Tk (t)|i⟩. (2.112)
i=0 i=0 k=0
50
Notice that
d−1
X
2
∥|x(T )⟩∥ = x2i (2.113)
i=0
and
d−1
X
2
∥|Xgood ⟩∥ = (p + 1)(n + 1) x2i = (p + 1)(n + 1)∥|x(T )⟩∥2 . (2.114)
i=0
Considering the definition of q, the contribution from time interval h under the
1 1
Z
2 2 2 dτ
q ∥x(T )∥ = max ∥x̂(t)∥ = √ max ∥x̂(t)∥2
t∈[0,T ] π −1 1 − τ 2 t∈[0,T ]
1 1
Z
dτ
≥ √ max ∥x̂(t)∥2
π −1 1 − τ 2 t∈[Γh ,Γh+1 ]
(2.115)
1 1
Z
dτ
= √ max ∥x̂h (t)∥2
π −1 1 − τ 2 t∈[−1,1]
1 1
Z
dt
≥ ∥x̂h (t)∥2 p ,
π −1 1−t
2
where x̂h (t) is the solution of (2.47) with the rescaling in (2.19). By the orthogonality of
1 Z d−1 ∞
1 1 XX
Z
2 1 2 2 dt 2 dt
q ∥x(T )∥ ≥ ∥x̂h (t)∥ p = ( c i,k (Γ h+1 )Tk (t)) p
π −1 1−t
2 π −1 i=0 k=0 1−t
2
51
For all h ∈ [m]0 , we have
m−1 d−1 n
2 2
X 1 XX 2 1
mq ∥x(T )∥ ≥ ci,k (Γh+1 ) = ∥|Xbad ⟩∥2 , (2.117)
h=0
π i=0 k=0 π
and therefore
(p + 1)(n + 1)
∥|Xgood ⟩∥2 = (p + 1)(n + 1)∥x(T )∥2 ≥ ∥|Xbad ⟩∥2 . (2.118)
πmq 2
(p + 1)(n + 1)
Pmeasure ≥ (2.119)
πmq 2 + (p + 1)(n + 1)
as claimed.
We now describe a procedure for preparing the vector |B⟩ in the linear system
(2.27) (defined in (2.29) and (2.30)) using the given ability to prepare the initial state of
the system of differential equations. We also evaluate the complexity of this procedure.
Lemma 2.6. Consider state preparation oracles acting on a state space with basis vectors
l ∈ [n], let Ox be a unitary oracle that maps |0⟩|0⟩|0⟩ to a state proportional to |0⟩|γ⟩|0⟩
and |h⟩|ϕ⟩|l⟩ to |h⟩|ϕ⟩|l⟩ for any |ϕ⟩ orthogonal to |0⟩; let Of (h, l) be a unitary that maps
52
|h⟩|0⟩|l⟩ to a state proportional to |h⟩|fh (cos lπ
n
)⟩|l⟩ and maps |0⟩|ϕ⟩|0⟩ to |0⟩|ϕ⟩|0⟩ for
quantum state
m−1
XX n
|B⟩ ∝ |0⟩|γ⟩|0⟩ + |h⟩|fh (cos lπ
n
)⟩|l⟩ (2.120)
h=0 l=1
∥γ∥
b00 = q Pn ,
2 lπ 2
∥γ∥ + l=1 ∥fh (cos n )∥
(2.121)
∥fh (cos lπ
n
)∥
bhl = q , h ∈ [m]0 , l ∈ [n]
∥γ∥2 + nl=1 ∥fh (cos lπ
P 2
n
)∥
so that
m−1
XX n
b2hl = 1. (2.122)
h=0 l=0
This can be done in time complexity O(mn) by standard techniques [92]. Then we
m−1
XX n
|0⟩|γ⟩|0⟩ + |h⟩|fh (cos lπ
n
)⟩|l⟩ (2.124)
h=0 l=1
53
2.8 Main result
Having analyzed the solution error, condition number, success probability, and state
preparation procedure for our approach, we are now ready to establish the main result.
Theorem 2.1. Consider an instance of the quantum ODE problem as defined in Problem 2.1.
Assume A(t) can be diagonalized as the form A(t) = V (t)Λ(t)V −1 (t) where Λ(t) =
diag(λ1 (t), . . . , λd (t)) with Re(λi (t)) ≤ 0 for each i ∈ [d]0 and t ∈ [0, T ]. Then there
exists a quantum algorithm that produces a state x(T )/∥x(T )∥ ϵ-close to x̂(T )/∥x̂(T )∥
in l2 norm, succeeding with probability Ω(1), with a flag indicating success, using
queries to oracles OA (h, l) (a sparse matrix oracle for Ah (tl ) as defined in (2.18)) and
Ox and Of (h, l) (as defined in Lemma 2.6). Here ∥A∥ := maxt∈[0,T ] ∥A(t)∥; κV :=
∥x̂(t)∥
g := ∥x̂(T )∥, g ′ := max max ∥x̂(n+1) (t)∥, q := max . (2.126)
t∈[0,T ] n∈N t∈[0,T ] ∥x(T )∥
The gate complexity is larger than the query complexity by a factor of poly(log(κV ds∥A∥g ′ T /ϵ)).
Proof. We first present the algorithm and then analyze its complexity.
∥A∥T
≤ 1. (2.127)
2m
54
Then, as in Section 2.3, we divide the interval [0, T ] into small subintervals
T
τ := max {τh }, τh := |Γh+1 − Γh | = . (2.128)
0≤h≤m−1 m
Each subinterval [Γh , Γh+1 ] for h ∈ [m − 1] is mapped onto [−1, 1] with a linear
2(t − Γh )
Kh : t 7→ t = 1 − . (2.129)
Γh+1 − Γh
We choose
e log(Ω) log(ω)
n = max , (2.130)
2 log(log(Ω)) log(log(ω))
where
g ′ em g ′ em(1 + ϵ)
Ω := = (2.131)
δ gϵ
and
g′
ω := (m + 1). (2.132)
∥γ∥
en+1
∥x̂(T ) − x(T )∥ ≤ m max ∥x̂(n+1) (t)∥ ≤δ (2.133)
t∈[0,T ] (2n)n
and
∥x̂(n+1) (t)∥ e n 1
max ≤ . (2.134)
t∈[0,T ] ∥γ∥ 2n m+1
55
Now ∥x̂(T ) − x(T )∥ ≤ δ implies
x̂(T ) x(T ) δ δ
− ≤ ≤ =: ϵ, (2.135)
∥x̂(T )∥ ∥x(T )∥ min{∥x̂(T )∥, ∥x(T )∥} g−δ
so we can choose such n to ensure that the normalized output state is ϵ-close to x̂(T )/∥x̂(T )∥.
Following Section 2.3, we build the linear system L|X⟩ = |B⟩ (see (2.27)) that
encodes the quantum spectral method. By Lemma 2.4, the condition number of this linear
system is at most (πm + p + 1)(n + 1)3.5 (2κV + e∥γ∥∞ ). Then we use the QLSA from
reference [46] to obtain a normalized state |X⟩ and measure the first and third register of
|X⟩ in the standard basis. If the measurement outcome for the first register belongs to
we output the state of the second register, which is a normalized state |x(T )⟩/∥|x(T )⟩∥
satisfying (2.135). By Lemma 2.5, the probability of this event happening is at least
(p+1)(n+1)
πmq 2 +(p+1)(n+1)
. To ensure m + p = O(∥A∥T ), we can choose
√
so we can achieve success probability Ω(1) with O(q/ n) repetitions of the above procedure.
matrix with O(ns) nonzero entries in any row or column. By Lemma 2.4 and our choice of
parameters, the condition number of L is O κV (m + p)n3.5 . Consequently, by Theorem
56
5 of [46], the QLSA produces the state |x(T )⟩ with
(2.138)
queries to the oracles OA (h, l), Ox , and Of (h, l), and its gate complexity is larger by
√
a factor of poly(log(κV mnds/δ)). Using O(q/ n) steps of amplitude amplification to
achieve success probability Ω(1), the overall query complexity of our algorithm is
(2.139)
as claimed.
useful bound in such a case by solving the implicit equations (2.133) and (2.134).
57
and
∥γ∥ + 2τ ∥f ∥
ω := (m + 1)κV . (2.142)
∥γ∥
en+1
∥x̂(T ) − x(T )∥ ≤ max ∥x̂(t) − x(t)∥ ≤ mκV (∥γ∥ + 2τ ∥f ∥) ≤δ (2.143)
t∈[−1,1] (2n)n
and
Corollary 2.1. For time-independent differential equations, under the same assumptions
O κV s∥A∥T q poly(log(κV sγ∥A∥∥f ∥T /ϵg)) (2.145)
queries to OA (h, l), Ox , and Of (h, l). The gate complexity of this algorithm is larger than
which characterizes the decay of the final state relative to the initial state. As discussed
procedure for performing this would have the unlikely consequence BQP = PP.
58
We also require the real parts of the eigenvalues of A(t) to be non-positive for all t ∈
[0, T ] so that the solution cannot grow exponentially. This requirement is essentially the
same as in the time-independent case considered in [53] and improves upon the analogous
condition in [52] (which requires an additional stability condition). Also as in [53], our
algorithm can produce approximate solutions for non-diagonalizable A(t), although the
So far we have focused on initial value problems (IVPs). Boundary value problems
(BVPs) are another widely studied class of differential equations that appear in many
Problem 2.1 but with a constraint on some linear combination of the initial and final
states:
dx(t)
= A(t)x(t) + f (t), (2.146)
dt
where x(t) ∈ Cd , A(t) ∈ Cd×d is s-sparse, and f (t) ∈ Cd for all t ∈ [0, T ], and a
solution x̂ ∈ C ∞ (0, T ) of this boundary value problem. Given oracles that compute the
59
locations and values of nonzero entries of A(t) for any t, and that prepare quantum states
α|x(0)⟩+β|x(T )⟩ = |γ⟩ and |f (t)⟩ for any t, the goal is to output a quantum state |x(t∗ )⟩
As before, we can rescale [0, T ] onto [−1, 1] by a linear mapping. However, since
subintervals. Instead, we directly map [0, T ] onto [−1, 1] with a linear map K satisfying
dx T
= − A(t)x + f (t) . (2.148)
dt 2
dx
= AK (t)x(t) + fK (t) (2.149)
dt
for t ∈ [−1, 1]. Now the boundary condition takes the form
Since we only have one solution interval, we need to choose a larger order n of the
60
Chebyshev series to reduce the solution error. In particular, we take
e log(Ω) log(ω)
n = ∥A∥T max , (2.151)
2 log(log(Ω)) log(log(ω))
dx(tl )
= AK (tl )x(tl ) + f (tl ), l ∈ [n] (2.152)
dt
Observe that the linear equations have the same form as in (2.25). Instead of (2.26),
the term with l = 0 encodes the condition (2.153) expanded in a Chebyshev series, namely
n
X n
X
αi ci,k Tk (t0 ) + βi ci,k Tk (tn ) = γi (2.154)
k=0 k=0
for each i ∈ [d]0 . Since Tk (t0 ) = 1 and Tk (tn ) = (−1)k , this can be simplified as
n
X
(αi + (−1)k βi )ci,k = γi . (2.155)
k=0
without loss of generality, the two sides of this equality can be divided by αi + (−1)k βi
61
to guarantee that the terms with l = 0 can be encoded as in (2.26).
Now this system can be written in the form of equation (2.27) with m = 1. Here
L, |X⟩, and |B⟩ are the same as in (2.31), (2.28), and (2.29), respectively, with m = 1,
we take
d−1 X
X n
L3 = Tk (t∗ )|i0⟩⟨ik|. (2.156)
i=0 k=0
∥L3 ∥ ≤ n + 1, (2.157)
and it follows that Lemma 2.3 also holds for boundary value problems. Similarly, Lemma 2.4
We are now ready to analyze the complexity of the quantum BVP algorithm. The
entries in any row or column, with condition number O(κV pn3.5 ). By Lemma 2.5 with
√
p = O(1), O(q/ n) repetitions suffice to ensure success probability Ω(1). By (2.151), n
Theorem 2.2. Consider an instance of the quantum BVP as defined in Problem 2.2.
Assume A(t) can be diagonalized as the form A(t) = V (t)Λ(t)V −1 (t) where Λ(t) =
diag(λ1 (t), . . . , λd (t)) with Re(λi (t)) ≤ 0 for each i ∈ [d]0 and t ∈ [0, T ]. Then there
exists a quantum algorithm that produces a state x(t∗ )/∥x(t∗ )∥ ϵ-close to x̂(t∗ )/∥x̂(t∗ )∥
62
in l2 norm, succeeding with probability Ω(1), with a flag indicating success, using
queries to OA (h, l), Ox , and Of (h, l). Here ∥A∥, κV , g, g ′ and q are defined as in
Theorem 2.1. The gate complexity is larger than the query complexity by a factor of
As for initial value problems, we can simplify this result in the time-independent
case.
Corollary 2.2. For a time-independent boundary value problem, under the same assumptions
queries to OA (h, l), Ox , and Of (h, l). The gate complexity of this algorithm is larger than
2.10 Discussion
based on the spectral method as an alternative to the more straightforward finite difference
method. Our algorithm handles time-independent differential equations with almost the
same complexity as [53], but unlike that approach, can also handle time-dependent differential
63
equations. Compared to [52], our algorithm improves the complexity of solving time-
This work raises several natural open problems. First, our algorithm must assume
1
that the solution is smooth. If the solution is in C r , the solution error is O( nr−2 ) by
Lemma 2.1. Can we improve the complexity to poly(log(1/ϵ)) under such weaker smoothness
assumptions?
35]. Can we develop quantum algorithms for general differential equations with query
complexity independent of g ′ ?
Third, our algorithm has nearly optimal dependence on T , scaling as O(T poly(log T )).
According to the no-fast-forwarding theorem [33], the complexity must be at least linear
in T , and indeed linear complexity is achievable for the case of Hamiltonian simulation
[93]. Can we handle general differential equations with complexity linear in T ? Furthermore,
can we achieve an optimal tradeoff between T and ϵ as shown for Hamiltonian simulation
in [42]?
64
Chapter 3: High-precision quantum algorithms for linear elliptic partial
differential equations
3.1 Introduction
In this chapter, we study high-precision quantum algorithms for linear elliptic partial
address can be stated as follows: Given a linear PDE with boundary conditions and an
error parameter ϵ, output a quantum state that is ϵ-close to one whose amplitudes are
proportional to the solution of the PDE at a set of grid points in the domain of the PDE.
We focus on elliptic PDEs, and we assume a technical condition that we call global strict
Our first algorithm is based on a quantum version of the FDM approach: we use
that system using the QLSA. We analyze our FDM algorithm as applied to Poisson’s
equation (which automatically satisfies global strict diagonal dominance) under periodic,
Dirichlet, and Neumann boundary conditions. Whereas previous FDM approaches [69,
70] considered fixed orders of truncation, we adapt the order of truncation depending on
ϵ, inspired by the classical adaptive FDM [71]. As the order increases, the eigenvalues
1
This chapter is based on the paper [59].
65
of the FDM matrix approach the eigenvalues of the continuous Laplacian, allowing for
more precise approximations. The main algorithm we present uses the quantum Fourier
transform (QFT) and takes advantage of the high-precision LCU-based QLSA [46]. We
this approach can also be applied to homogeneous Dirichlet and Neumann boundary
conditions. We state our result in Theorem 3.1, which (informally) says that this quantum
adaptive FDM approach produces a quantum state approximating the solution of Poisson’s
We also propose a quantum algorithm for more general second-order elliptic PDEs
quantum spectral methods [54]. The spectral method globally approximates the solution
for smooth functions) with undetermined coefficients, and then finds the coefficients by
for classical algorithms but feasible in a quantum context. To be able to apply the QLSA
efficiently, we show how to make the system sparse using variants of the quantum Fourier
transform. Our bound on the condition number of the linear system uses global strict
diagonal dominance, and introduces a factor in the complexity that measures the extent
to which this condition holds. We state our result in Theorem 3.2, which (informally)
66
the spatial dimension d. Bounding the complexities of these algorithms requires analyzing
how d and ϵ affect the condition numbers of the relevant linear systems (finite difference
matrices and matrices relating the spectral coefficients) and accounting for errors in the
approximate solution provided by the QLSA. Furthermore, the complexities of both approaches
scale logarithmically with high-order derivatives of the solution and the inhomogeneity.
The detailed complexity dependence is presented in Theorem 3.1 and Theorem 3.2, and
Table 3.1 compares the performance of our approaches to other classical and quantum
algorithms for PDEs. Compared to classical algorithms, quantum algorithms improve the
caveat that they produce a different representation of the solution). Compared to previous
quantum FDM/FEM/FVM algorithms [57, 69, 70, 94], the quantum adaptive FDM and
quantum spectral method improve the error dependence from poly(1/ϵ) to poly(log(1/ϵ)).
Our approaches achieve the best known dependence on the parameter ϵ for the Poisson
method approach not only achieves the best known dependence on d and ϵ for elliptic
PDEs with inhomogeneous Dirichlet boundary conditions, but also improves the dependence
The remainder of the chapter is structured as follows. Section 3.2 introduces technical
details about linear PDEs and formally states the problem we solve. Section 3.3 covers
our FDM algorithm for Poisson’s equation. Section 3.4 details the spectral algorithm for
elliptic PDEs. Finally, Section 3.5 concludes with a brief discussion of the results, their
67
Algorithm Equation Boundary conditions Complexity
FDM/FEM/FVM general general poly((1/ϵ)d )
Classical
Adaptive FDM/FEM [71] general general poly((log (1/ϵ))d )
Spectral method [67, 68] general general poly((log (1/ϵ))d )
Sparse grid FDM/FEM [95, 96] general general poly((1/ϵ)(log(1/ϵ))d )
Sparse grid spectral method [97, 98] elliptic general poly(log (1/ϵ)(log log(1/ϵ))d )
FEM [57] Poisson homogeneous poly(d, 1/ϵ)
FDM [69] Poisson homogeneous Dirichlet d poly(log d, 1/ϵ)
Quantum
Table 3.1: Summary of the time complexities of classical and quantum algorithms for d-
dimensional PDEs with error tolerance ϵ. Portions of the complexity in bold represent
best known dependence on that parameter.
In this chapter, we focus on systems of linear PDEs. Such equations can be written
in the form
C and the inhomogeneity f (x) ∈ C are scalar functions, and L is a linear differential
and its derivatives. A linear differential operator L of order h has the form
X ∂j
L (u(x)) = Aj (x) u(x), (3.2)
∂xj
∥j∥1 ≤h
68
where j = (j1 , . . . , jd ) is a d-dimensional non-negative vector with ∥j∥1 = j1 +· · ·+jd ≤
h, Aj (x) ∈ C, and
∂j ∂ j1 ∂ jd
u(x) = · · · u(x). (3.3)
∂xj ∂xj11 ∂xjdd
The problem reduces to a system of linear ordinary differential equations (ODEs) when
For example, systems of first-order linear PDEs can be written in the form
d
X ∂u(x)
Aj (x) + A0 (x)u(x) = f (x), (3.4)
j=1
∂xj
where Aj (x), A0 (x), f (x) ∈ C for j ∈ [d] := {1, . . . , d}. Similarly, systems of second-
d d
X ∂ 2 u(x) X ∂u(x)
Aj1 j2 (x) + Aj (x) + A0 (x)u(x) = f (x), (3.5)
j1 ,j2 =1
∂xj1 ∂xj2 j=1 ∂xj
where Aj1 ,j2 (x), Aj (x), A0 (x), f (x) ∈ C for j1 , j2 , j ∈ [d]. A well-known second-order
d
X ∂2
∆u(x) := u(x) = f (x). (3.6)
j=1
∂x2j
A linear PDE of order h is called elliptic if its differential operator (3.2) satisfies
X
Aj (x)ξ j ̸= 0, (3.7)
∥j∥1 =h
69
for all nonzero ξ j = ξ1j1 . . . ξdjd with ξ1 , . . . , ξd ∈ Rm and all x. Note that ellipticity only
depends on the highest-order terms. When h = 2, the linear PDE (3.5) is called a second-
order elliptic PDE if and only if Aj1 j2 (x) is positive-definite or negative-definite for any
d
X 1 X
C := 1 − |Aj1 ,j2 (x)| > 0 (3.8)
j1 =1
|Aj1 ,j1 (x)|
j2 ∈[d]\{j1 }
for all x. We call this condition global strict diagonal dominance, since it is a strengthening
d
X 1 X
d− |Aj1 ,j2 (x)| > 0. (3.9)
j1 =1
|Aj1 ,j1 (x)|
j2 ∈[d]\{j1 }
Observe that (3.8) holds for the Poisson equation (3.6) with C = 1.
In this chapter, we focus on the following boundary value problem. This is a formal
Problem 3.1. In the quantum PDE problem, we are given a system of second-order
elliptic equations
d
X ∂j X ∂ 2 u(x)
L (u(x)) = Aj j u(x) = Aj1 j2 = f (x) (3.10)
∂x j ,j =1
∂x j 1 ∂x j 2
∥j∥1 =2 1 2
satisfying the global strict diagonal dominance condition (3.8), where the variable x =
70
scalar function of x satisfying f (x) ∈ C ∞ , and the linear coefficients Aj ∈ C. We are
∂u(x)
also given boundary conditions u(x) = γ(x) ∈ ∂D or ∂xj xj =±1
= γ(x)|xj =±1 ∈ ∂D
where γ(x) ∈ C ∞ . We assume there exists a weak solution û(x) ∈ C for the boundary
value problem (see Reference [99, Section 6.1.2]). Given sparse matrix oracles to provide
the locations and values of nonzero entries of a matrix Aj1 j2 (x), Aj (x), and A0 (x) on
a set of interpolation nodes x2 , and that prepare normalized states |γ(x)⟩ and |f (x)⟩
whose amplitudes are proportional to γ(x) and f (x) on a set of interpolation nodes x,
the goal is to output a quantum state |u(x)⟩ that is ϵ-close to the normalized u(x) on a
We now describe our first approach to quantum algorithms for linear PDEs, based
on the finite difference method (FDM). Using this approach, we show the following.
Theorem 3.1. There exists a quantum algorithm that outputs a state ϵ-close to |u⟩ that
runs in time
r
d2k+1 u h d2k+1 u i
6.5 4.5 3
Õ d log /ϵ log d4 log /ϵ /ϵ (3.11)
dx2k+1 dx2k+1
and makes
r
d2k+1 u h d2k+1 u i
4 3 4 3
Õ d log /ϵ log d log /ϵ /ϵ (3.12)
dx2k+1 dx2k+1
2
For instance, Aj1 j2 (x) is modeled by a sparse matrix oracle that, on input (m, l), gives the location of
the l-th nonzero entry in row m, denoted as n, and gives the value Aj1 j2 (x)m,n .
71
queries to the oracle for f .
To show this, we first construct a linear system corresponding to the finite difference
approximation of Poisson’s equation with periodic boundary conditions and bound the
error of this high-order FDM in Section 3.3.1 (Lemma 3.1). Then we bound the condition
number of this system in Section 3.3.2 (Lemma 3.2 and Lemma 3.3) and bound the
error of approximation in Section 3.3.3 (Lemma 3.4). We use these results to give an
discussing how to use the method of images to apply this algorithm for Neumann and
values of f on a finite set of points near x. Generally there are no restrictions on where
these points are located relative to x, but they are typically taken to be uniformly spaced
points with respect to a certain coordinate. This corresponds to discretizing [−1, 1]d
(or [0, 2π)d ) to a d-dimensional rectangular lattice (where we use periodic boundary
conditions).
For a scalar field, in which u(x) ∈ C, the canonical elliptic PDE is Poisson’s
equation (3.6), which we consider solving on [0, 2π)d with periodic boundary conditions.
This also implies results for the domain Ω = [−1, 1]d under Dirichlet (u(∂Ω) = 0) and
Neumann (n̂·∇u(∂Ω) = 0 where n̂ denotes the normal direction to ∂Ω, which for domain
∂u
Ω = [−1, 1]d is equivalent to ∂xj xj =±1
= 0 for j ∈ [d]) boundary conditions.
72
3.3.1 Linear system
the central finite difference formula of order 2k. Taking xj = jh for a lattice with spacing
k
′′ 1 X
f (0) ≈ 2 rj f (jh) (3.13)
h j=−k
2(−1)j+1 (k!)2
j ∈ [k]
j 2 (k−j)!(k+j)!
rj := −2 Pk rj j=0 (3.14)
j=1
r
−j j ∈ −[k].
We leave the dependence on k implicit in this notation. The following lemma characterizes
Lemma 3.1 ([6, Theorem 7]). Let k ≥ 1 and suppose f (x) ∈ C 2k+1 for x ∈ R. Define
k
d2 u(x0 ) 1 X d2k+1 u eh 2k−1
= r j f (x 0 + jh) + O (3.15)
dx2 h2 j=−k dx2k+1 2
73
where
d2k+1 u d2k+1 u
:= max (y) . (3.16)
dx2k+1 y∈[x0 −kh,x0 +kh] dx2k+1
Since we assume periodic boundary conditions and apply the same FDM formula
at each lattice site, the matrices we consider are circulant. Define the 2n × 2n matrix
S to have entries Si,j = δi,j+1 mod 2n . If we represent the solution u(x) as a vector u =
P2n
j=1 u(πj/n)ej , then we can approximate Poisson’s equation using a central difference
formula as
k
1 1 X
j −j
Lu = r0 I + rj (S + S ) u=f (3.17)
h2 h2 j=1
P2n
where f = j=1 f (πj/n)ej . The solution u corresponds exactly with the quantum
in Reference [70] and other quantum differential equation algorithms. The matrix in
this linear system is just the finite difference matrix, so it suffices to bound its condition
number and approximation error (whereas previous quantum algorithms involved more
on 2n points.
Pk
Lemma 3.2. For k < (6/π 2 )1/3 n2/3 , the matrix L = r0 I + j=1 rj (S
j
+ S −j ) with rj as
74
in (3.14) has condition number κ(L) = O(n2 ).
Proof. We first upper bound ∥L∥ using Gershgorin’s circle theorem [101] (a similar
2(k!)2 2
|rj | = 2
≤ 2 (3.18)
j (k − j)!(k + j)! j
since
(k!)2 k(k − 1) · · · (k − j + 1)
= < 1. (3.19)
(k − j)!(k + j)! (k + j)(k + j − 1) · · · (k + 1)
k k
X X 2 2π 2
2 |rj | ≤ 2 ≤ . (3.20)
j=1 j=1
j2 3
k
X 2π 2
|r0 | ≤ 2 |rj | ≤ , (3.21)
j=1
3
4π 2
so ∥L∥ ≤ 3
.
To lower bound ∥L−1 ∥ we lower bound the (absolute value of the) smallest non-
zero eigenvalue of L (since by construction the all-ones vector is a zero eigenvector). Let
75
ω := exp(πi/n). Since L is circulant, its eigenvalues are
k
X
λl = r0 + rj (ω lj + ω −lj ) (3.22)
j=1
k
X πlj
= r0 + 2rj cos (3.23)
j=1
n
k
π 2 l2 j 2 (πcj )4
X πc
j
= r0 + 2rj 1 − 2
+ 4
cos (3.24)
j=1
2n 4!n n
k 22 2
X π l j (πcj )4 πc
j
= 2rj − 2
+ 4
cos (3.25)
j=1
2n 4!n n
where the cj ∈ [0, lj] arise from the Taylor remainder theorem. Using (3.18), we have
k
π2 X 2 π4k3
λ1 + 2 rj j ≤ . (3.26)
n j=1 6n4
76
We now compute the sum
k k
X
2
X 2(−1)j (k!)2
− rj j = j2 (3.27)
j=1 j=1
j 2 (k + j)!(k − j)!
k
X (−1)j
= 2(k!)2 (3.28)
j=1
(k + j)!(k − j)!
k
2(k!)2 X
j 2k
= (−1) (3.29)
(2k)! j=1 k+j
2k
2(k!)2 X
j+k 2k
= (−1) (3.30)
(2k)! j=k+1 j
2 2k
k (k!) j 2k
X
= (−1) (−1) (3.31)
(2k)! j=0, j̸=k j
2
k (k!) 2k k 2k
= (−1) (1 − 1) − (−1) (3.32)
(2k)! k
= −1. (3.33)
Therefore, we have
π2 π4k3
λ1 ≤ − + . (3.34)
n2 6n4
77
In d dimensions, a similar analysis holds.
Pk
Lemma 3.3. For k < (6/π 2 )1/3 n2/3 , let L := r0 I + j=1 rj (S
j
+ S −j ) with rj as in
Proof. By the triangle inequality for spectral norms, ∥L′ ∥ ≤ d∥L∥. Since L has zero-sum
rows by construction, the all-ones vector lies in its kernel, and thus the smallest non-zero
4 π 2 k 3 −1
κ(L′ ) ≤ dn2 1 − (3.38)
3 6n2
There are two types of error relevant to our analysis: the FDM error and the QLSA
error. We assume that we are able to perfectly generate states proportional to f . The FDM
errors arise from the remainder terms in the finite difference formulas and from inexact
We introduce several states for the purpose of error analysis. Let |u⟩ be the quantum
P Nd
state that is proportional to u = j∈Zd2n u(πj/n) i=1 eji for the exact solution of the
differential equation. Let |ū⟩ be the state output by a QLSA that exactly solves the
linear system. Let |ũ⟩ be the state output by a QLSA with error. Then the total error
78
of approximating |u⟩ by |ũ⟩ is bounded by
and without loss of generality we can take ϵFDM and ϵQLSA to be of the same order of
magnitude.
d2
Pd d
Lemma 3.4. Let u(x) be the exact solution of ( i=1 dx2i )u(x) = f (x). Let u ∈ R(2n)
Nd d
Let ū ∈ R(2n)
P
encode the exact solution in the sense that u = j∈Zd2n u(πj/n) i=1 eji .
1 ′
be the exact solution of the FDM linear system h2
L ū = f , where L′ is a d-dimensional
P2n
(2k)th-order Laplacian as above with k < (6/π 2 )1/3 n3/2 , and f = j=1 f (πj/n)ej .
d2k+1 u
Then ∥u − ū∥ ≤ O(2d/2 n(d/2)−2k+1 dx2k+1
(e2 /4)k ).
d2k+1 u
Proof. The remainder term of the central difference formula is O( dx2k+1
h2k−1 (e/2)2k ),
so
1 ′ d2k+1 u
2k−1
L u = f + O (eh/2) ϵ (3.41)
h2 dx2k+1
where ϵ is a (2n)d dimensional vector whose entries are O(1). This implies
1 ′ d2k+1 u
2k−1
L (u − ū) = O (eh/2) ϵ (3.42)
h2 dx2k+1
79
and therefore
d2k+1 u
∥u − ū∥ = O (eh/2) 2k+1
∥(L′ )−1 ϵ∥ (3.43)
dx2k+1
d2k+1 u
= O (2n)d/2 (eh/2) 2k+1
/λ 1 . (3.44)
dx2k+1
d2k+1 u
∥u − ū∥ = O 2d/2 n(d/2)−2k+1 (e/2)2k
(3.45)
dx2k+1
as claimed.
that correspond to Laplacian FDM operators. For periodic boundary conditions, the
Laplacians are circulant, so they can be diagonalized by the QFT F (or a tensor product of
the simplest way to simulate exp(iLt) is to perform the inverse QFT, apply controlled
phase rotations to implement exp(iDt), and perform the QFT. Reference [102] shows how
to exactly implement arbitrary diagonal unitaries on m qubits using O(2m ) gates. Since
we consider Laplacians on n lattice sites, simulating exp(iLt) takes O(n) gates with
the dominant contribution coming from the phase rotations (alternatively, the methods
of Reference [103] or Reference [104] could also be used). Using this Hamiltonian
simulation algorithm in a QLSA for the FDM linear system gives us the following theorem.
80
We restate Theorem 3.1 as follows.
Theorem 3.1. There exists a quantum algorithm that outputs a state ϵ-close to |u⟩ that
runs in time
r
d2k+1 u h d2k+1 u i
6.5 4.5 3
Õ d log /ϵ log d4 log /ϵ /ϵ (3.11)
dx2k+1 dx2k+1
and makes
r
d2k+1 u h d2k+1 u i
4 3 4 3
Õ d log /ϵ log d log /ϵ /ϵ (3.12)
dx2k+1 dx2k+1
Proof. We use the Fourier series based QLSA from Reference [46]. By Theorem 3
p
of that work, the QLSA makes O(κ log(κ/ϵQLSA )) uses of a Hamiltonian simulation
algorithm and uses of the oracle for the inhomogeneity. For Hamiltonian simulation we
use d parallel QFTs and phase rotations as described in Reference [102], for a total of
p
O(dnκ log(κ/ϵQLSA )) gates. The condition number for the d-dimensional Laplacian
scales as κ = O(dn2 ).
We take ϵFDM and ϵQLSA to be of the same order and just write ϵ. Then the QLSA has
p
time complexity O(d2 n3 log(dn2 /ϵ)) and query complexity O(dn2 log(dn2 /ϵ)). The
adjustable parameters are the number of lattice sites n and the order 2k of the finite
difference formula. To keep the error below the target error of ϵ we require
81
or equivalently,
d2k+1 u
(−d/2) + (2k − 1 − (d/2)) log(n) − 2k log(e/2) = Ω log /ϵ . (3.47)
dx2k+1
Now we focus on the choice of adjustable n and k relying on ϵ. This procedure is inspired
by the classical adaptive FDM [71], so we call it the adaptive FDM approach. We must
have 2k − 1 > d/2 for the left-hand side of (3.47) to be positive for large n. Indeed,
we find the best performance by taking k as large as possible subject to the assumption
of Lemma 3.2, i.e., k = cn2/3 where c := (6/π 2 )1/3 . For this choice of k and for n
d2k+1 u
k log(n) = cn2/3 log(n) = Ω log /ϵ . (3.48)
dx2k+1
To satisfy the condition 2cn2/3 − 1 > d/2, we must have n = Ω(d3/2 ). Combining this
d2k+1 u
n = Θ d3/2 log3/2 /ϵ (3.49)
dx2k+1
so that
d2k+1 u
k = cn2/3 = Θ d log /ϵ . (3.50)
dx2k+1
82
The QLSA then has the stated time complexity
r
p
d2k+1 u h d2k+1 u i
2 3 2 6.5 4.5 4 3
Õ(d n log(dn /ϵ)) = O d log /ϵ) log d log /ϵ /ϵ ,
dx2k+1 dx2k+1
(3.51)
and makes
r
d2k+1 u
h d2k+1 u i
2 2 4 3 3
Õ(dn log(dn /ϵ)) = O d log /ϵ log d4 log /ϵ)/ϵ .
dx2k+1 dx2k+1
(3.52)
This can be compared to the cost of using the conjugate gradient method to solve
the same linear system classically. The sparse conjugate gradient algorithm for an N × N
√
matrix has time complexity O(N s κ log(1/ϵ)). For arbitrary dimension N = Θ(nd ), we
have s = dk = cdn2/3 and κ = O(dn2 ), so that the time complexity is O(d4+3d/2 log(1/ϵ)
d2k+1 u
log5/2+3d/2 ( dx2k+1
/ϵ)). Alternatively, d fast Fourier transforms could be used, although
d2k+1 u
this will generally take Ω(nd ) = Ω(d3d/2 log3d/2 ( dx2k+1
/ϵ)) time.
We can apply the method of images to deal with homogeneous Neumann and
Dirichlet boundary conditions using the algorithm for periodic boundary conditions described
above. In the method of images, the domain [−1, 1] is extended to include all of R, and
the boundary conditions are related to symmetries of the solutions. For a pair of Dirichlet
83
boundary conditions there are two symmetries: the solutions are anti-symmetric about −1
Continuity and anti-symmetry about −1 and 1 imply f (−1) = f (1) = 0, and furthermore
that f (x) = 0 for all odd x ∈ Z and that f (x + 4) = f (x) for all x ∈ R. For Neumann
boundary conditions, the solutions are instead symmetric about −1 and 1, which also
We would like to combine the method of images with the FDM to arrive at finite
difference formulas for this special case. In both cases, the method of images implies
that the solutions are periodic, so without loss of generality we can consider a lattice on
[0, 2π) instead of a lattice on R. It is useful to think of this lattice in terms of the cycle
graph on 2n vertices, i.e., (V, E) = (Z2n , {(i, i + 1) | i ∈ Z2n }), which means that the
vectors encoding the solution u(x) will lie in R2n . Let each vector ej correspond to the
vertex j. Then we divide R2n into a symmetric and an anti-symmetric subspace, namely
span{ej + e2n+1−j }nj=1 and span{ej − e2n+1−j }nj=1 , respectively. Vectors lying in the
symmetric subspace correspond to solutions that are symmetric about 0 and π, so they
Restricting to a subspace of vectors reduces the size of the FDM vectors and matrices
we consider, and the symmetry of that subspace indicates how to adjust the coefficients.
84
If the FDM linear system is L′′ u′′ = f ′′ then L′′ has entries
r|i−j| ± ri+j−1 i≤k
L′′i,j = r|i−j| k <i≤n−k (3.53)
r
|i−j|±r 2n−i−j+1 n−k ≤i
where + (−) is chosen for Neumann (Dirichlet) boundary conditions and due to the
truncation order k, rj = 0 for any j > k. This is similar to how Laplacian coefficients are
For the purpose of solving the new linear systems using quantum algorithms, we
still treat these cases as obeying periodic boundary conditions. We assume access to an
oracle that produces states |f ′′ ⟩ proportional to the inhomogeneity f ′′ (x). Then we apply
the QLSA for periodic boundary conditions using |f ′′ ⟩|±⟩ to encode the inhomogeneity,
which will output solutions of the form |u′′ ⟩|±⟩. Here the ancillary state is chosen to be
Typically, the (second-order) graph Laplacian for the path graph with Dirichlet
boundary conditions has diagonal entries that are all equal to 2; however, using the above
specification for the entries of L leads to the (1, 1) and (n, n) entries being 3 while the
Reference [106] to diagonalize the Dirichlet graph Laplacian. In this case it is easiest to
consider the lattice of a cycle graph on 2n + 2 vertices, where the vertices 0 and n + 1 are
selected as boundary points where the field takes the value 0. The relevant antisymmetric
85
subspace is now span({ej − e2n+2−j }nj=1 ) (which has no support on e0 and en+1 ).
If we again write the linear system as L′′ u′′ = f ′′ , then the Laplacian has entries
r|i−j| − ri+j i≤k
L′′i,j = r|i−j| k <i≤n−k
r
|i−j|−r 2n−i−j+2 n − k ≤ i.
we assume that this oracle operates in a Hilbert space with one additional dimension
|f ′′ ⟩|−⟩ and solve Poisson’s equation for periodic boundary conditions to output a state
|u′′ ⟩|−⟩ (where |u′′ ⟩ lies in an (n + 1)-dimensional Hilbert space but has no support on
We now turn our attention to the spectral method for multi-dimensional PDEs.
a quantum algorithm based on the pseudo-spectral method [67, 68, 107] for second-
order elliptic equations with global strict diagonal dominance, under various boundary
Theorem 3.2. Consider an instance of the quantum PDE problem as defined in Problem 3.1
with Dirichlet boundary conditions (3.81). Then there exists a quantum algorithm that
86
produces a state in the form of (3.82) whose amplitudes are proportional to u(x) on a set
of interpolation nodes x (with respect to the uniform grid nodes for periodic boundary
norm for all nodes x, succeeding with probability Ω(1), with a flag indicating success,
using
d∥A∥Σ
+ qd poly(log(g ′ /gϵ))
2
(3.54)
C∥A∥∗
P
queries to oracles as defined in Section 3.4.4. Here ∥A∥Σ := ∥j∥1 ≤h ∥Aj ∥, ∥A∥∗ :=
Pd
j=1 |Aj,j |, C > 0 is defined in (3.8), and
The gate complexity is larger than the query complexity by a factor of poly(log(d∥A∥Σ /ϵ)).
After introducing the method, we discuss the complexity of the quantum shifted
Fourier transform (Lemma 3.5) and the quantum cosine transform (Lemma 3.6) in Section 3.4.1.
These transforms are used as subroutines in our algorithm. Then we construct a linear
system whose solution encodes the solution of the PDE in Section 3.4.2 , analyze its
condition number in Section 3.4.3 (Lemma 3.10, established using Lemma 3.7, Lemma 3.8,
and Lemma 3.9), and consider the complexity of state preparation in Section 3.4.4 (Lemma 3.11).
In the spectral approach, we approximate the exact solution û(x) by a linear combination
87
of basis functions
X
u(x) = ck ϕk (x) (3.57)
∥k∥∞ ≤n
and
d
Y
ϕk (x) = ϕkj (xj ), j ∈ [d]. (3.58)
j=1
We choose different basis functions for the case of periodic boundary conditions
and for the more general case of non-periodic boundary conditions. When the boundary
some cases (e.g., for the Poisson equation), can be faster. Specifically, for any kj ∈
ei(kj −⌊n/2⌋)πxj , periodic conditions,
ϕkj (xj ) = (3.59)
Tkj (xj ) := cos(kj arccos xj ), non-periodic conditions.
The coefficients ck are determined by demanding that u(x) satisfies the ODE and
boundary conditions at a set of interpolation nodes {χl = (χl1 , . . . , χld )}∥l∥∞≤n with
lj ∈ [n + 1]0 , where
2lj
− 1, periodic conditions,
n+1
χlj = (3.60)
cos πlj ,
non-periodic conditions.
n
88
2l
Here { n+1 − 1 : l ∈ [n + 1]0 } are called the uniform grid nodes, and {cos πl
n
: l ∈
so that
The convergence behavior of the spectral method is related to the smoothness of the
solution. For a solution in C r+1 , the spectral method approximates the solution with n =
for all j ∈ [d], we have (n + 1)d terms in total. Consequently, a classical pseudo-spectral
method solves multi-dimensional PDEs with complexity poly(logd (1/ϵ)). Such classical
spectral methods rapidly become infeasible since the number of coefficients (n + 1)d
applies techniques from the quantum spectral method for ODEs [54]. However, in the
case of PDEs, the linear system to be solved is non-sparse. We address this difficulty
89
3.4.1 Quantum shifted Fourier transform and quantum cosine transform
of the discrete Fourier transform (DFT) acting on the amplitudes of a quantum state. The
n
1 X 2πikl
v̂l = √ exp vk , l ∈ [n + 1]0 . (3.63)
n + 1 k=0 n+1
n
1 X 2πikl
Fn := √ exp |l⟩⟨k|. (3.64)
n + 1 k,l=0 n+1
Here we also consider the quantum shifted Fourier transform (QSFT), an analogue
of the classical shifted discrete Fourier transform, which maps v ∈ Cn+1 to v̂ ∈ Cn+1
with
90
We define the multi-dimensional QSFT by the tensor product, namely
d
1 X Y 2πi(kj −⌊n/2⌋)(lj −(n+1)/2)
Fns := p exp n+1
|l1 ⟩ . . . |ld ⟩⟨k1 | . . . ⟨kd |,
(n + 1)d ∥k∥∞ ,∥l∥∞ ≤n j=1
(3.67)
Lemma 3.5. The QSFT Fns defined by (3.66) can be performed with gate complexity
O(log n log log n). More generally, the d-dimensional QSFT Fns defined by (3.67) can be
Proof. The unitary matrix Fns can be written as the product of three unitary matrices
Fns = Sn Fn Rn , (3.68)
where
n
X 2πik(n + 1)/2
Rn = exp − |k⟩⟨k| (3.69)
k=0
n + 1
and
n
X 2πi ⌊n/2⌋ (l − (n + 1)/2)
Sn = exp − |l⟩⟨l|. (3.70)
l=0
n+1
It is well known that Fn can be implemented with gate complexity O(log n log log n), and
it is straightforward to implement Rn and Sn with gate complexity O(log n). Thus the
91
We rewrite v in the form
X
v= vk |k1 ⟩ . . . |kd ⟩, (3.71)
∥k∥∞ ≤n
where vk ∈ C with k = (k1 , . . . , kd ), and each kj ∈ [n]0 for j ∈ [d]. The unitary matrix
d
O
Fns = Fns . (3.72)
j=1
QSFT on each register. Thus, the gate complexity of performing Fns is O(d log n log log n).
[108, 109]. The QCT can be regarded as an analogue of the discrete cosine transform
r n
2X klπ
v̂l = δk δl cos vk , l ∈ [n + 1]0 , (3.73)
n k=0 n
where
√12 l = 0, n
δl := (3.74)
1 l ∈ [n − 1].
92
In other words, the QCT is the orthogonal transform
r n
2 X klπ
Cn := δl δk cos |l⟩⟨k|. (3.75)
n k,l=0 n
d
r
2 d X Y kj lj π
Cn := δkj δlj cos |l1 ⟩ . . . |ld ⟩⟨k1 | . . . ⟨kd |, (3.76)
n n
∥k∥∞ ,∥l∥∞ ≤n j=1
[n + 1]0 .
The classical DCT on (n + 1)-dimensional vectors takes Θ(n log n) gates, while
poly(log n). According to Theorem 1 of Reference [108], the gate complexity of performing
Lemma 3.6. The quantum cosine transform Cn defined by (3.75) can be performed
with gate complexity O(log n log log n). More generally, the multi-dimensional QCT Cn
defined by (3.76) can be performed with gate complexity O(d log n log log n).
93
decomposed into a QFT Fn+1 , a permutation
1
1
Pn =
1
,
(3.77)
..
.
1
and additional operations with O(1) cost. The QFT Fn+1 has gate complexity O(log n
log log n). We then consider an alternative way to implement Pn that improves over the
approach in [110].
Pn = Fn Tn Fn−1 , (3.78)
Pn 2πik
where Fn is the Fourier transform (3.64) and Tn = k=0 e− n+1 |k⟩⟨k| is diagonal. The
gate complexities of performing Fn and Tn are O(log n log log n) and O(log n), respectively.
It follows that Cn can be implemented with circuit complexity O(log n log log n).
d
O
Cn = Cn . (3.79)
j=1
QCT on each register. Thus, the gate complexity of performing Cn is O(d log n log log n).
94
3.4.2 Linear system
In this section we introduce the quantum PDE solver for the problem (3.1). We
construct a linear system that encodes the solution of (3.1) according to the pseudo-
spectral method introduced above, using the QSFT/QCT introduced in Section 3.4.1 to
ensure sparsity.
We consider a linear PDE problem (Problem 3.1) with periodic boundary conditions
According to the elliptic regularity theorem (Theorem 6 in Section 6.3 of Reference [99]),
We now show how to apply the Fourier and Chebyshev pseudo-spectral methods to
X
|u⟩ ∝ ck ϕk (χl )|l1 ⟩ . . . |ld ⟩, (3.82)
∥k∥∞ ,∥l∥∞ ≤n
where ϕk (χl ) is defined by (3.58) using (3.59) for the appropriate boundary conditions
95
(periodic or non-periodic). This state corresponds to a truncated Fourier/Chebyshev
approximation and is ϵ-close to the exact solution û(χl ) with n = poly(log(1/ϵ)) [68].
Note that this state encodes the values of the solution at the interpolation nodes (3.60)
appropriate to the boundary conditions (the uniform grid nodes in the Fourier approach,
Instead of developing our algorithm for the standard basis, we aim to produce a
state
X
|c⟩ ∝ ck |k1 ⟩ . . . |kd ⟩ (3.83)
∥k∥∞ ≤n
that is the inverse QSFT/QCT of |u⟩. We then apply the QSFT/QCT to transform back
The truncated spectral series of the inhomogeneity f (x) and the boundary conditions
and
X
γ(x) = γ̂k ϕk (x), (3.85)
∥k∥∞ ≤n
respectively. We define quantum states |f ⟩ and |γ⟩ by interpolating the nodes {χl }
defined by (3.60) as
X
|f ⟩ ∝ ϕkj (χl )fˆk |l1 ⟩ . . . |ld ⟩, (3.86)
∥k∥∞ ,∥l∥∞ ≤n
96
and
X
|γ⟩ ∝ ϕkj (χl )γ̂k |l1 ⟩ . . . |ld ⟩, (3.87)
∥k∥∞ ,∥l∥∞ ≤n
respectively. These are the states that we assume we can produce using oracles. We
X
|fˆ⟩ ∝ fˆk |k1 ⟩ . . . |kd ⟩, (3.88)
∥k∥∞ ≤n
and
X
|γ̂⟩ ∝ γ̂k |k1 ⟩ . . . |kd ⟩. (3.89)
∥k∥∞ ≤n
Having defined these states, we now detail the construction of the linear system. At
a high level, we construct two linear systems: one system Ax = f (where x corresponds
to (3.83)) describes the differential equation, and another system Bx = g describes the
boundary conditions. We combine these into a linear system with the form
Lx = (A + B)x = f + g. (3.90)
Even though we do not impose the two linear systems separately, we show that there exists
a unique solution of (3.90) (which is therefore the solution of the simultaneous equations
Ax = f and Bx = g), since we show that L has full rank, and indeed we upper bound its
97
Part of this linear system will correspond to just the differential equation
X ∂j
L (u(χl )) = Aj u(χl ) = f (χl ), (3.91)
∂xj
∥j∥1 =2
S
while another part will come from imposing the boundary conditions on ∂D = j∈[d] ∂Dj ,
∂j X (j)
u(x) = ck ϕk (x) (3.94)
∂xj
∥k∥∞ ≤n
for some fixed j ∈ Nd (as we explain below, such a decomposition exists for the choices of
basis functions in (3.59)). Using this expression, we obtain the following linear equations
98
(j)
for ck :
X X (j)
X
Aj ϕk (χl )ck |l1 ⟩ . . . |ld ⟩ = ϕk (χl )fˆk |l1 ⟩ . . . |ld ⟩. (3.95)
∥j∥1 =2 ∥k∥∞ ,∥l∥∞ ≤n ∥k∥∞ ,∥l∥∞ ≤n
(j)
To determine the transformation between ck and ck , we can make use of the differential
d ikπx
e = ikπeikπx (3.96)
dx
and
′ ′
Tk+1 (t) Tk−1 (t)
2Tk (t) = − , (3.97)
k+1 k−1
respectively. We have
(j)
X
ck = [Dn(j) ]kr cr. , ∥k∥∞ ≤ n, (3.98)
∥r∥∞ ≤n
(j)
where Dn can be expressed as the tensor product
99
with j = (j1 , . . . , jd ). The matrix Dn for the Fourier basis functions in (3.59) can be
As detailed in Appendix A of Reference [54], the matrix Dn for the Chebyshev polynomials
entries
2r
[Dn ]kr = , k + r odd, r > k, (3.101)
σk
where
2 k = 0
σk := (3.102)
1 k ∈ [n].
Substituting (3.99) into (3.95), with Dn defined by (3.100) in the periodic case or
(3.101) in the non-periodic case, and performing the multi-dimensional inverse QSFT/QCT
(for a reason that will be explained in the next section), we obtain the following linear
equations for cr :
X X X
Aj [Dn(j) ]kr cr |l1 ⟩ . . . |ld ⟩ = fˆk |l1 ⟩ . . . |ld ⟩. (3.103)
∥j∥1 =2 ∥k∥∞ ,∥l∥∞ ,∥r∥∞ ≤n ∥k∥∞ ,∥l∥∞ ≤n
Notice that the matrices (3.100) and (3.101) are not full rank. More specifically,
there exists at least one zero row in the matrix of (3.103) when using either (3.100) (k =
⌊n/2⌋) or (3.101) (k = n). To obtain an invertible linear system, we next introduce the
boundary conditions.
100
3.4.2.2 Adding the linear system from the boundary conditions
When we use the form (3.82) of u(x) to write linear equations describing the
boundary conditions (3.93), we obtain a non-sparse linear system. Thus, for each x ∈
X X
ck |k1 ⟩ . . . |kd ⟩ = γ̂k1+ |k1 ⟩ . . . |kd ⟩, γ̂kj+ ∈ ∂Dj ,
∥k∥∞ ≤n ∥k∥∞ ≤n
kj =n kj =n
(3.104)
X X
kj
(−1) ck |k1 ⟩ . . . |kd ⟩ = γ̂k1− |k1 ⟩ . . . |kd ⟩, γ̂kj− ∈ ∂Dj
∥k∥∞ ≤n ∥k∥∞ ≤n
kj =n−1 kj =n−1
for all j ∈ [d], where the values of kj indicate that we place these constraints in the last
two rows with respect to the jth coordinate. We combine these equations with (3.103) to
d
X X (j) X X
Aj [D n ]kr cr |k1 ⟩ . . . |kd ⟩ = (Aj,j γ̂kj+ +Aj,j γ̂kj− +fˆk )|k1 ⟩ . . . |kd ⟩,
∥j∥1 =2 ∥k∥∞ ,∥r∥∞ ≤n ∥k∥∞ ≤n j=1
(3.105)
where
Dn(j) + G(j)
n , ∥j∥1 = 2, ∥j∥∞ = 2;
(j)
Dn = (3.106)
Dn(j) ,
∥j∥1 = 2, ∥j∥∞ = 1
101
tensor product
⊗r−1
G(j)
n = I ⊗ Gn ⊗ I ⊗d−r (3.107)
where the rth entry of j is 2 and all other entries are 0. For the Fourier case in (3.59) used
for periodic boundary conditions, Dn comes from (3.100), and the nonzero entries of Gn
are
Alternatively, for the Chebyshev case in (3.59) used for non-periodic boundary conditions,
The system (3.105) has the form of (3.90). For instance, the matrix in (3.90) for
(3.111)
For periodic boundary conditions, using (3.98), (3.100), and (3.108), the second-order
102
(2)
differential matrix Dn has nonzero entries
(2)
[Dn ]k,k = −((k − ⌊n/2⌋)π)2 , k ∈ [n + 1]0 \{⌊n/2⌋},
(3.112)
(2)
[Dn ]⌊n/2⌋,k = 1, k ∈ [n + 1]0 .
(2)
For non-periodic boundary conditions, using (3.98), (3.101), and (3.109), Dn has nonzero
entries
r−1 r−1
(2) X 2r X 2l r(r2 − k 2 )
[Dn ]kr = [Dn ]kl [Dn ]lr = = , k + r even, r > k + 1,
l=k+1
σk l=k+1 σl σk
k + l odd k + l odd
l + r odd l + r odd
(2)
[Dn ]n,k = 1, k ∈ [n + 1]0 ,
(2)
[Dn ]n−1,k = (−1)k , k ∈ [n + 1]0 .
(3.113)
We discuss the invertible linear system (3.105) and upper bound its condition number
We now analyze the condition number of the linear system. We begin with two
lemmas bounding the singular values of the matrices (3.112) and (3.113) that appear in
Lemma 3.7. Consider the case of periodic boundary conditions. Then for n ≥ 4, the
103
(2)
largest and smallest singular values of Dn defined in (3.112) satisfy
(2)
σmax (Dn ) ≤ (2n)2.5 ,
(3.114)
(2) 1
σmin (Dn ) ≥√ .
2
Proof. By direct calculation of the l∞ norm (i.e., the maximum absolute column sum) of
2
(2) (n + 1)π
∥Dn ∥∞ ≤ ≤ (2n)2 . (3.115)
2
(2) 1
[(Dn )−1 ]k,k = − , k ∈ [n + 1]0 \{⌊n/2⌋},
((k − ⌊n/2⌋)π)2
(3.116)
(2) 1
[(Dn )−1 ]⌊n/2⌋,k = , k ∈ [n + 1]0
((k − ⌊n/2⌋)π)2
∞
2 X 1 2 π4
∥(Dn )−1 ∥2F ≤1+2 = 1 + ≤ 2. (3.117)
k=1
k4π4 π 4 90
(2) √ 2
σmax (Dn ) ≤ n + 1∥Dn ∥∞ ≤ (2n)2.5 ,
(3.118)
(2) 1
1
σmin (Dn ) ≥ 2 −1 ≥√
∥(Dn ) ∥F 2
104
as claimed.
Lemma 3.8. Consider the case of non-periodic boundary conditions. Then the largest
(2)
and smallest singular values of Dn defined in (3.113) satisfy
(2)
σmax (Dn ) ≤ n4 ,
(3.119)
(2) 1
σmin (Dn ) ≥ .
16
2
r(r2 − k 2 )
(2)
∥Dn ∥2F 2
≤ n max ≤ n2 · n6 = n8 . (3.120)
k,r σk
(2)
Next we upper bound ∥(Dn )−1 ∥. By definition,
(2) 2
∥(Dn )−1 ∥ = sup ∥(Dn )−1 b∥. (3.121)
∥b∥≤1
Given any vector b satisfying ∥b∥ ≤ 1, we estimate ∥x∥ defined by the full-rank linear
system
(2)
Dn x = b. (3.122)
(2)
Notice that Dn is the sum of the upper triangular matrix Dn2 and (3.109), the coordinates
system
105
Given the same b, we also define the vector y by
n−1 n−1
X X 2l
bk = [Dn ]kl yl = yl , k + l odd, l > k, k ∈ [n − 1]0 . (3.125)
l=1 l=1
σ k
Using this equation with k = l − 1 and k = l + 1, we can express yl in terms of bl−1 and
bl+1 :
2l 1
yl = bl−1 − bl+1 , l ∈ [n − 1], (3.126)
σl−1 σl−1
106
where each coordinate of y can be expressed by
n n
X X 2r
yl = [Dn ]lr xr = xr , l + r odd, r > l, l ∈ [n − 1]. (3.129)
r=1 r=1
σl
Substituting the (r − 1)st and the (r + 1)st equations of (3.129), we can express x in terms
of y:
2r 1
xr = yr−1 − yr+1 , r ∈ [n]\{1}, (3.130)
σr−1 σr−1
n
X n−1
X
x2l ≤2 yl2 . (3.131)
l=2 l=1
Then we calculate x20 +x21 based on the last two equations of (3.122), (3.127), and (3.130),
107
giving
1
x20 + x21 = [(x0 + x1 )2 + (x0 − x1 )2 ]
2
n
!2 n
!2
1 X X
= bn − xl + bn−1 − (−1)l xl
2 l=2 l=2
n !2
1 X σl−1 1
= bn − yl−1 − yl+1
2 l=2
2l σl−1
n !2
X σl−1 1
+ bn−1 − (−1)l yl−1 − yl+1
l=2
2l σ l−1
n
!" n 2
2
1 X σl−1 2
X 1 (3.132)
≤ 1+ 2
bn + yl−1 − yl+1 + b2n−1
2 l=2
4l l=2
σ l−1
n
#
2
X 1
+ yl−1 − yl+1
l=2
σl−1
∞
! " n #
1 1X 1 X 1
≤ 1+ b2n + b2n−1 + 1+ 2 2
(yl−1 2
+ yl+1 )
2 4 l=2 l2 l=2
σ l−1
2
" n−1
#
1 π X
≤ 1+ b2n + b2n−1 + 4 yl2
2 24 l=1
n−2
X
≤ b2n + b2n−1 +8 b2l .
l=0
n
X n
X
x2l = x20 + x21 + x2l
l=0 l=2
n−2
X n−2
X
≤ b2n + b2n−1 +8 b2l +4 b2l (3.133)
l=0 l=0
n−2
X
≤ b2n + b2n−1 + 12 b2l ≤ 12
l=0
108
holds for any vectors b satisfying ∥b∥ ≤ 1. Thus
(2)
∥(Dn )−1 ∥ = sup ∥x∥ ≤ 12 < 16. (3.134)
∥b∥≤1
Altogether, we have
(2) 2
σmax (Dn ) ≤ ∥Dn ∥F ≤ n4 ,
(3.135)
(2) 11
σmin (Dn ) ≥ 2 −1 ≥
∥(Dn ) ∥ 16
as claimed in (3.119).
Using these two lemmas, we first upper bound the condition number of the linear
system for Poisson’s equation, and then extend the result to general elliptic PDEs.
For the case of the Poisson equation, we use the following simple bounds on the
d
M
L= Mj = M1 ⊗ I ⊗d−1 + I ⊗ M2 ⊗ I ⊗d−2 + · · · + I ⊗d−1 ⊗ Md , (3.136)
j=1
where {Mj }dj=1 are square matrices. If the largest and smallest singular values of Mj
satisfy
σmax (Mj ) ≤ smax
j ,
(3.137)
σmin (Mj ) ≥ smin
j ,
109
respectively, then the condition number of L satisfies
Pd max
j=1 sj
κL ≤ Pd min
. (3.138)
j=1 sj
max
σmax (exp(Mj )) ≤ esj ,
(3.139)
smin
σmin (exp(Mj )) ≥ e j
Nd
using (3.137). The singular values of the Kronecker product j=1 exp(Mj ) are
d
O Yd
σk1 ,...,kd exp(Mj ) = σkj (exp(Mj )) (3.140)
j=1 j=1
where σkj (exp(Mj )) are the singular values of the matrix exp(Mj ) for each j ∈ [d], where
kj runs from 1 to the dimension of Mj . Using the property of the Kronecker sum that
d
M d
O
exp(L) = exp Mj = exp(Mj ), (3.141)
j=1 j=1
Pd max
σmax (exp(L)) ≤ e j=1 sj ,
(3.142)
Pd min
σmin (exp(L)) ≥ e j=1 sj .
110
Finally, we bound the singular values of the matrix logarithm of (3.142) by
d
X
σmax (L) ≤ smax
j ,
j=1
d
(3.143)
X
σmin (L) ≥ smin
j .
j=1
Pd
j=1 smax
j
κL ≤ Pd min
(3.144)
j=1 sj
as claimed.
This lemma easily implies a bound on the condition number of the linear system for
Poisson’s equation:
Corollary 3.1. Consider an instance of the quantum PDE problem as defined in Problem 3.1
for Poisson’s equation (3.6) with Dirichlet boundary conditions (3.81). Then for n ≥ 4,
Proof. The matrix in (3.90) for Poisson’s equation (3.6) is LPoisson defined in (3.111). For
(2)
σmax (Dn ) ≤ n4 ,
(3.146)
(2) 1
σmin (Dn ) ≥
16
111
(2)
by Lemma 3.7 and Lemma 3.8. Let Mj = Dn for j ∈ [d] in (3.136), and apply
of LPoisson is bounded by
(2)
σmax (Dn )
κLPoisson ≤ (2)
≤ (2n)4 (3.147)
σmin (Dn )
as claimed.
We now consider the condition number of the linear system for general elliptic
PDEs.
Lemma 3.10. Consider an instance of the quantum PDE problem as defined in Problem 3.1
with Dirichlet boundary conditions (3.81). Then for n ≥ 4, the condition number of L in
∥A∥Σ
κL ≤ (2n)4 , (3.148)
C∥A∥∗
P Pd Pd
where ∥A∥Σ := ∥j∥1 ≤2 |Aj | = j1 ,j2 =1 |Aj1 ,j2 |, ∥A∥∗ := j=1 |Aj,j |, and C > 0 is
defined in (3.8).
Recall that C quantifies the extent to which the global strict diagonal dominance
condition holds.
X (j)
L= Aj D n . (3.149)
∥j∥1 =2
112
We upper bound the spectral norm of the matrix L by
X (j)
∥L∥ ≤ |Aj |∥D n ∥. (3.150)
∥j∥1 =2
(j)
For the matrix D n defined by (3.106), Lemma 3.7 (in the periodic case) and Lemma 3.8
(j)
∥D n ∥ ≤ n4 , (3.151)
so we have
X
∥L∥ ≤ |Aj |n4 = ∥A∥Σ n4 . (3.152)
∥j∥1 =2
similar to (3.111) that can be bounded by Lemma 3.9, and L2 is a sum of tensor products
(2) (2)
L1 = A1,1 Dn ⊗ I ⊗d−1 + · · · + Ad,d I ⊗d−1 ⊗ Dn
(3.153)
L2 = L − L1 .
j ∈ [d] are either all positive or all negative; we consider Aj,j > 0 without loss of
113
generality, so
d
X d
X
∥A∥∗ = |Aj,j | = Aj,j . (3.154)
j=1 j=1
d
X 1 X
C =1− |Aj1 ,j2 | > 0, (3.155)
j1 =1
Aj1 ,j1
j2 ∈[d]\{j1 }
where 0 < C ≤ 1.
(j)
We now upper bound ∥L2 L−1 −1
1 ∥ by bounding ∥Dn L1 ∥ for each j = (j1 , . . . , jd )
that has exactly two entries equal to 1 and all other entries 0. Specifically, consider jr1 =
(j) (j)
We first upper bound ∥Dn ∥ by 21 ∥L(j) ∥. Notice the matrices Dn and L(j) share the
same singular vectors. For k ∈ [n + 1]0 , we let vk and λk denote the right singular
vectors and corresponding singular values of Dn , respectively. Then the right singular
(j) Nd
vectors of Dn and L(j) are vk := j=1 vkj , where k = (k1 , . . . , kd ) with kj ∈ [n + 1]0
P
for j ∈ [d]. For any vector v = ∥k∥∞ ≤n αk vk , we have
X X
∥Dn(j) v∥2 = |αk |2 ∥Dn(j) vk ∥2 = |αk |2 (λkjr1 λkjr2 )2 , (3.157)
∥k∥∞ ≤n ∥k∥∞ ≤n
X X
∥L(j) v∥2 = |αk |2 ∥L(j) vk ∥2 = |αk |2 (λ2kjr + λ2kjr )2 , (3.158)
1 2
∥k∥∞ ≤n ∥k∥∞ ≤n
(j)
which implies ∥Dn v∥ ≤ 21 ∥L(j) v∥ by the inequality of arithmetic and geometric means
114
(also known as the AM-GM inequality). Since this holds for any vector v, we have
1 (j) −1
∥Dn(j) L−1
1 ∥ ≤ ∥L L1 ∥. (3.159)
2
(2)
Next we upper bound ∥Dn2 ∥ by ∥Dn ∥. For any vector u = [u0 , . . . , un ]T , define
and
2
Dn [u0 , . . . , un ]T = [w0 , . . . , wn ]T . (3.161)
Notice that w⌊n/2⌋ = 0 and wk = wk for k ∈ [n + 1]0 \{⌊n/2⌋} for periodic conditions,
n n
X X (2)
∥Dn2 v∥2 = ∥w∥ = 2
wk2 ≤ w2k = ∥w∥2 = ∥Dn v∥2 . (3.162)
k=0 k=0
Therefore,
2 2
X X (2)
∥L(j) L−1
1 ∥ ≤ ∥I ⊗rs −1 ⊗ Dn2 ⊗ I ⊗d−rs L−1
1 ∥ ≤ ∥I ⊗rs −1 ⊗ Dn ⊗ I ⊗d−rs L−1
1 ∥.
s=1 s=1
(3.163)
We also have
2
1 X ⊗rs −1 (2)
∥Dn(j) L−1
1 ∥ ≤ ∥I ⊗ Dn ⊗ I ⊗d−rs L−1
1 ∥. (3.164)
2 s=1
115
(2)
We can rewrite I ⊗rs −1 ⊗ Dn ⊗ I ⊗d−rs L−1
1 in the form
d
!−1
(2) X (2)
I ⊗rs −1 ⊗ Dn ⊗ I ⊗d−rs Ah,h I ⊗rh −1 ⊗ Dn ⊗ I ⊗d−rh . (3.165)
h=1
(2)
The matrices I ⊗rh −1 ⊗ Dn ⊗ I ⊗d−rh share the same singular values and singular vectors,
so
(2) λkrs 1
∥I ⊗rs −1 ⊗ Dn ⊗ I ⊗d−rs L−1
1 ∥ = max Pd < , (3.166)
λk r
h=1 Ah,h λkh Ars ,rs
(2)
where λkh are singular values of I ⊗rh −1 ⊗ Dn ⊗ I ⊗d−rh for kh ∈ [n]0 , h ∈ [d]. This
implies
1 1 1
∥Dn(j) L−1
1 ∥ ≤ ( + ). (3.167)
2 Ar1 ,r1 Ar2 ,r2
(j)
Using (3.155), considering each instance of Dn in L2 , we have
d
X X 1 X
∥L2 L−1
1 ∥ ≤ |Aj1 ,j2 |∥Dn(j) L−1
1 ∥ ≤ |Aj1 ,j2 | ≤ 1 − C. (3.168)
j1 ̸=j2 j1 =1
Aj1 ,j1
j2 ∈[d]\{j1 }
∥L−1
1 ∥, we have
1
−1 −1 ∥L−1
1 ∥ 1/ 16 ∥A∥∗ 16
∥L ∥ = ∥(L1 +L2 ) ∥ ≤ ∥(I+L2 L−1 −1 −1
1 ) ∥∥L1 ∥ ≤ −1 ≤ = .
1 − ∥L2 L1 ∥ C C∥A∥∗
(3.169)
Thus we have
∥A∥Σ
κL = ∥L∥∥L−1 ∥ ≤ (2n)4 (3.170)
C∥A∥∗
116
as claimed.
We now describe a state preparation procedure for the vector f + g in the linear
system (3.90).
Lemma 3.11. Let Of be a unitary oracle that maps |0⟩|0⟩ to a state proportional to
|0⟩|f ⟩, and |ϕ⟩|0⟩ to |ϕ⟩|0⟩ for any |ϕ⟩ orthogonal to |0⟩; let Ox be a unitary oracle that
maps |0⟩|0⟩ to |0⟩|0⟩, |j⟩|0⟩ to a state proportional to |j⟩|γ j+ ⟩ for j ∈ [d], and |j + d⟩|0⟩
to a state proportional to |j + d⟩|γ j− ⟩ for j ∈ [d]. Suppose ∥|f ⟩∥, ∥|γ j+ ⟩∥, ∥|γ j− ⟩∥ and
v
u ∥k∥∞ ≤n dj=1 [fˆk2 + (Aj,j γ̂kj+ )2 + (Aj,j γ̂kj− )2 ]
uP P
q := t . (3.171)
P Pd
|fˆk + Aj,j γ̂ j+ + Aj,j γ̂ j− |2
∥k∥∞ ≤n j=1 k k
d
X X
|B⟩ ∝ (fˆk + Aj,j γ̂kj+ + Aj,j γ̂kj− )|k1 ⟩ . . . |kd ⟩, (3.172)
∥k∥∞ ≤n j=1
with coefficients defined as in (3.88) and (3.89), can be prepared with gate and query
Proof. Starting from the initial state |0⟩|0⟩, we first perform a unitary transformation U
117
satisfying
∥|f ⟩∥
U |0⟩ = q Pd |0⟩
∥|f ⟩∥2 + j=1 A2j,j ∥|γ j+ ⟩∥2 + A2j,j ∥|γ j− ⟩∥2
d
X Aj,j ∥|γ j+ ⟩∥
+ q Pd 2
|j⟩ (3.173)
j=1 ∥|f ⟩∥2 + j+ 2
j=1 Aj,j ∥|γ ⟩∥ + A2j,j ∥|γ j− ⟩∥2
d
X Aj,j ∥|γ j− ⟩∥
+ q Pd |j + d⟩
j=1 ∥|f ⟩∥2 + j=1 A2j,j ∥|γ j+ ⟩∥2 + A2j,j ∥|γ j− ⟩∥2
This can be done in time O(2d + 1) by standard techniques [92]. Then we apply Ox and
Of to obtain
according to (3.86) and (3.87). We then perform the d-dimensional inverse QSFT (for
X
(fˆk |0⟩ + A1,1 γ̂k1+ |1⟩ + · · · + Ad,d γ̂kd− |2d⟩)|k1 ⟩ . . . |kd ⟩. (3.176)
∥k∥∞ ≤n
Finally, observe that if we measure the first register in a basis containing the uniform
118
superposition |0⟩ + |1⟩ + · · · + |2d⟩ (say, the Fourier basis) and obtain the outcome
d
X X
(fˆk + Aj,j γ̂kj+ + Aj,j γ̂kj− )|k1 ⟩ . . . |kd ⟩. (3.177)
∥k∥∞ ≤n j=1
Since this outcome occurs with probability 1/q 2 , we can prepare this state with probability
close to 1 using O(q) steps of amplitude amplification. According to Lemma 3.5 and
Lemma 3.6, the d-dimensional (inverse) QSFT or QCT can be performed with gate
complexity O(d log n log log n). Thus the total gate and query complexity is
may be able to avoid the factor of q in the complexity of the overall algorithm.
Having analyzed the condition number and state preparation procedure for our
approach, we are now ready to establish the main result. Theorem 3.2 as follows.
Theorem 3.2. Consider an instance of the quantum PDE problem as defined in Problem 3.1
with Dirichlet boundary conditions (3.81). Then there exists a quantum algorithm that
produces a state in the form of (3.82) whose amplitudes are proportional to u(x) on a set
of interpolation nodes x (with respect to the uniform grid nodes for periodic boundary
119
norm for all nodes x, succeeding with probability Ω(1), with a flag indicating success,
using
d∥A∥Σ
+ qd poly(log(g ′ /gϵ))
2
(3.54)
C∥A∥∗
P
queries to oracles as defined in Section 3.4.4. Here ∥A∥Σ := ∥j∥1 ≤h ∥Aj ∥, ∥A∥∗ :=
Pd
j=1 |Aj,j |, C > 0 is defined in (3.8), and
The gate complexity is larger than the query complexity by a factor of poly(log(d∥A∥Σ /ϵ)).
First we choose
log(Ω)
n := , (3.178)
log(log(Ω))
where
g ′ (1 + ϵ)
Ω= . (3.179)
gϵ
(n+1) en g′ gϵ
∥û(x) − u(x)∥ ≤ max ∥û (x)∥ n
≤ = =: δ. (3.180)
x (2n) Ω 1+ϵ
120
Now ∥û(x) − u(x)∥ ≤ δ implies
û(x) u(x) δ δ
− ≤ ≤ = ϵ, (3.181)
∥û(x)∥ ∥u(x)∥ min{∥û(x)∥, ∥u(x)∥} g−δ
so we can choose n to ensure that the normalized output state is ϵ-close to û(x)/∥û(x)∥.
As described in Section 3.4.2, the algorithm uses the high-precision QLSA from
Reference [46] and the multi-dimensional QSFT/QCT (and its inverse). According to
Lemma 3.5 and Lemma 3.6, the d-dimensional (inverse) QSFT or QCT can be performed
with gate complexity O(d log n log log n). According to Lemma 3.11, the query and gate
matrix with (n + 1) or (n + 1)d nonzero entries in any row or column for periodic or
and its gate complexity is larger by a factor of poly(log(d∥A∥Σ n)). Using the value of n
d∥A∥Σ
+ qd poly(log(g ′ /gϵ)),
2
(3.182)
C∥A∥∗
d∥A∥Σ
poly(log(d∥A∥Σ /ϵ)) + qd poly(log(g ′ /gϵ))
2
(3.183)
C∥A∥∗
121
which is larger by a factor of poly(log(d∥A∥Σ /ϵ)), as claimed.
Note that we can establish a more efficient algorithm in the special case of the
Poisson equation with homogeneous boundary conditions. In this case, ∥A∥Σ = ∥A∥∗ =
can be reduced to d poly(log(g ′ /gϵ)), since we can remove 2d applications of the QSFT
or QCT for preparing a state depending on the boundary conditions, and since γ = 0
conditions. In summary, the query complexity of the Poisson equation with homogeneous
boundary conditions is
using the FDM and spectral methods. These algorithms use high-precision QLSAs to
solve Poisson’s equation and other second-order elliptic equations. Whereas previous
This work raises several natural open problems. First, for the quantum adaptive
FDM, we only deal with Poisson’s equation with homogeneous boundary conditions.
Can we apply the adaptive FDM to other linear equations or to inhomogeneous boundary
conditions? The quantum spectral algorithm applies to second-order elliptic PDEs with
Dirichlet boundary conditions. Can we generalize it to other linear PDEs with Neumann
122
or mixed boundary conditions? Also, can we develop algorithms for space- and time-
dependent PDEs? These cases are more challenging since the quantum Fourier transform
cannot be directly applied to ensure sparsity. Finally, can we improve the dependence on
d?
Second, the complexity scales logarithmically with high-order derivatives (of the
inhomogeneity or solution) for both the adaptive FDM and the spectral method. In
particular, Theorem 3.1 shows that the complexity of the quantum adaptive FDM scales
d2k+1 u
logarithmically with dx2k+1
, and Theorem 3.2 shows that the complexity of the quantum
spectral method is poly(log g ′ ), where g ′ upper bounds ∥û(n+1) (x)∥ (see (3.55)). Such a
algorithms, including the classical adaptive FDM (see for example Theorem 7 of Reference [6])
and spectral methods (see for example Eq. (1.8.28) of Reference [107]), both of which
d2k+1 u
have the same logarithmic dependence on dx2k+1
and g ′ . This logarithmic dependence
means that the algorithm is efficient even when faced with a highly oscillatory solution
depends on the first-order derivatives of the Hamiltonian [34, 35]. Can we develop
quantum algorithms for PDEs with query complexity independent of high-order derivatives,
Third, can we use quantum algorithms for PDEs as a subroutine of other quantum
algorithms? For example, some PDE algorithms have state preparation steps that require
inverting finite difference matrices (such as Reference [70] using certain oracles for the
initial conditions); are there other scenarios in which state preparation can be done using
123
the solution of another system of PDEs? Can quantum algorithms for PDEs be applied to
Finally, how should these algorithms be applied? While PDEs have broad applications,
much more work remains to understand the extent to which quantum algorithms can be
of practical value. Answering this question will require careful consideration of various
technical aspects of the algorithms. In particular: What measurements give useful information
about the solutions, and how can those measurements be efficiently implemented? How
should the oracles encoding the equations and boundary conditions be implemented in
practice? And with these aspects taken into account, what are the resource requirements
124
Chapter 4: Efficient quantum algorithms for dissipative nonlinear differential
equations
4.1 Introduction
In this chapter, we study efficient quantum algorithm for dissipative nonlinear differential
(4.1). Note that polynomials of degree higher than two, and even more general nonlinearities,
can be reduced to the quadratic case by introducing additional variables [72, 73]. The
quadratic case also directly includes many archetypal models, such as the logistic equation
in biology, the Lorenz system in atmospheric dynamics, and the Navier–Stokes equations
in fluid dynamics.
the solutions of high-dimensional systems of linear ODEs [52, 53, 54] and PDEs [55,
56, 57, 58, 59, 60, 61]. Such algorithms can produce a quantum state proportional to
equations in time poly(log n) using the quantum linear system algorithm [84].
Early work on quantum algorithms for differential equations already considered the
1
This chapter is based on the paper [81].
125
nonlinear case [74]. It gave a quantum algorithm for ODEs that simulates polynomial
nonlinearities by storing multiple copies of the solution. The complexity of this approach
is polynomial in the logarithm of the dimension but exponential in the evolution time,
many copies of the solution to represent the nonlinearity throughout the evolution.
Recently, heuristic quantum algorithms for nonlinear ODEs have been studied.
method that might be amenable to the quantum linear system algorithm. In [76], the
authors provide a high-level description of how linearization can help solve nonlinear
preprint [77] also describes a quantum algorithm to solve a nonlinear ODE by linearizing
it using a different approach from the one taken here. However, a proof of correctness of
their algorithm involving a bound on the condition number and probability of success is
not given. The authors also do not describe how barriers such as those of [78] could be
of the theory have been widely studied. Generically, such modifications enable quickly
solving hard computational problems (e.g., solving unstructured search among n items
general [78, 79, 80]. Therefore, constructing efficient quantum algorithms for general
We design and analyze a quantum algorithm that overcomes this limitation using
126
Carleman linearization [73, 82, 83]. This approach embeds polynomial nonlinearities into
dimensional linear approximation. The Carleman method has previously been used in
the analysis of dynamical systems [111, 112, 113] and the design of control systems
[114, 115, 116], but to the best of our knowledge it has not been employed in the context of
quantum algorithms. We discretize the finite ODE system in time using the forward Euler
method and solve the resulting linear equations with the quantum linear system algorithm
[46, 84]. We control the approximation error of this approach by combining a novel
convergence theorem with a bound for the global error of the Euler method. Furthermore,
we provide an upper bound for the condition number of the linear system and lower
bound the success probability of the final measurement. Subject to the condition R < 1,
where the quantity R (defined in Problem 4.1 below) characterizes the relative strength
of the nonlinear and dissipative linear terms, we show that the total complexity of this
is the sparsity, T is the evolution time, q quantifies the decay of the final solution relative
to the initial condition, n is the dimension, and ϵ is the allowed error (see Theorem 4.1).
In the regime R < 1, this is an exponential improvement over [74], which has complexity
exponential in T .
Note that the solution cannot decay exponentially in T for the algorithm to be
of quantum ODE algorithms [53]. For homogeneous ODEs with R < 1, the solution
necessarily decays exponentially in time (see equation (4.30)), so the algorithm is not
asymptotically efficient. However, even for solutions with exponential decay, we still
127
find an improvement over the best previous result O(1/ϵT ) [74] for sufficiently small ϵ.
Thus our algorithm might provide an advantage over classical computation for studying
evolution for short times. More significantly, our algorithm can handle inhomogeneous
quadratic ODEs, for which it can remain efficient in the long-time limit since the solution
can remain asymptotically nonzero (for an explicit example, see the discussion just before
the proof of Lemma 4.2), or can decay slowly (i.e., q can be poly(T )). Inhomogeneous
equations arise in many applications, including for example the discretization of PDEs
We also provide a quantum lower bound for the worst-case complexity of simulating
strongly nonlinear dynamics, showing that the algorithm’s condition R < 1 cannot be
significantly improved in general (Theorem 4.2). Following the approach of [78, 79], we
construct a protocol for distinguishing two states of a qubit driven by a certain quadratic
√
ODE. Provided R ≥ 2, this procedure distinguishes states with overlap 1 − ϵ in time
poly(log(1/ϵ)). Since nonorthogonal quantum states are hard to distinguish, this implies
by quadratic ODEs arising in biology and epidemiology as well as in fluid and plasma
which describes many physical phenomena, can be treated by our approach provided
the Reynolds number is sufficiently small. We also note that while the formal validity of
our arguments assumes R < 1, we find in one numerical experiment that our proposed
We emphasize that, as in related quantum algorithms for linear algebra and differential
128
equations, instantiating our approach requires an implicit description of the problem that
allows for efficient preparation of the initial state and implementation of the dynamics.
features that can be revealed by efficient quantum measurements. More work remains
to understand how these methods might be applied, as we discuss further in Section 4.7.
the quantum quadratic ODE problem. Section 4.3 presents the Carleman linearization
procedure and describes its performance. Section 4.4 gives a detailed analysis of the
bound for simulating quadratic ODEs. Section 4.6 describes how our approach could
be applied to several well-known ODEs and PDEs and presents numerical results for the
case of the viscous Burgers equation. Finally, we conclude with a discussion of the results
ODE
du
= F2 u⊗2 + F1 u + F0 (t), u(0) = uin . (4.1)
dt
2
Here u = [u1 , . . . , un ]T ∈ Rn , u⊗2 = [u21 , u1 u2 , . . . , u1 un , u2 u1 , . . . , un un−1 , u2n ]T ∈ Rn ,
each uj = uj (t) is a function of t on the interval [0, T ] for j ∈ [n] := {1, . . . , n}, F2 ∈
2
Rn×n , F1 ∈ Rn×n are time-independent matrices, and the inhomogeneity F0 (t) ∈ Rn is
129
The main computational problem we consider is as follows.
quadratic ODE as in (4.1). We assume F2 , F1 , and F0 are s-sparse (i.e., have at most s
nonzero entries in each row and column), F1 is diagonalizable, and that the eigenvalues
the quantity
1 ∥F0 ∥
R := ∥uin ∥∥F2 ∥ + . (4.2)
| Re (λ1 )| ∥uin ∥
For some given T > 0, we assume the values Re (λ1 ), ∥F2 ∥, ∥F1 ∥, ∥F0 (t)∥ for each
t ∈ [0, T ], and ∥F0 ∥ := maxt∈[0,T ] ∥F0 (t)∥, ∥F0′ ∥ := maxt∈[0,T ] ∥F0′ (t)∥ are known, and
that we are given oracles OF2 , OF1 , and OF0 that provide the locations and values of the
nonzero entries of F2 , F1 , and F0 (t) for any specified t, respectively, for any desired row
or column2 . We are also given the value ∥uin ∥ and an oracle Ox that maps |00 . . . 0⟩ ∈ Cn
to a quantum state proportional to uin . Our goal is to produce a quantum state |u(T )⟩
that is ϵ-close to the normalized u(T ) for some given T > 0 in ℓ2 norm.
∥uin ∥∥F2 ∥
When F0 (t) = 0 (i.e., the ODE is homogeneous), the quantity R = | Re (λ1 )|
is
qualitatively similar to the Reynolds number, which characterizes the ratio of the (nonlinear)
convective forces to the (linear) viscous forces within a fluid [117, 118]. More generally,
R quantifies the combined strength of the nonlinearity and the inhomogeneity relative to
dissipation.
Note that without loss of generality, given a quadratic ODE satisfying (4.1) with
2
For instance, F1 is modeled by a sparse matrix oracle OF1 that, on input (j, l), gives the location of the
l-th nonzero entry in row j, denoted as k, and gives the value (F1 )j,k .
130
R < 1, we can modify it by rescaling u → γu with a suitable constant γ to satisfy
and
with R left unchanged by the rescaling. We use this rescaling in our algorithm and
its analysis. With this rescaling, a small R implies both small ∥uin ∥∥F2 ∥ and small
and indeed the complexity of the best known quantum algorithm is exponential in the
evolution time T [74]. However, for a dissipative nonlinear ODE without a source,
any quadratic nonlinear effect will only be significant for a finite time because of the
dissipation. To exploit this, we can create a linear system that approximates the initial
nonlinear evolution within some controllable error. After the nonlinear effects are no
longer important, the linear system properly captures the almost-linear evolution from
then on.
131
achieved by introducing powers of the variables into the system, allowing it to be written
system to N equations, where the truncation level N depends on the allowed approximation
Let us describe the Carleman linearization procedure in more detail. Given a system
of quadratic ODEs (4.1), we apply the Carleman procedure to obtain the system of linear
ODEs
dŷ
= A(t)ŷ + b(t), ŷ(0) = ŷin (4.5)
dt
1 1
ŷ1 A1 A2 ŷ1 F0 (t)
ŷ A2 A2 A2 ŷ 0
2 1 2 3 2
ŷ 3 3 3
d 3
A2 A 3 A 4
ŷ 0
3
= . . , (4.6)
+
dt .
..
... ... ... .. ..
N −1 N −1 N −1
ŷN −1 AN −2 AN −1 AN ŷN −1 0
ŷN AN
N −1 AN
N ŷN 0
j j j ×nj+1 j ×nj
where ŷj = u⊗j ∈ Rn , ŷin = [uin ; u⊗2 ⊗N
in ; . . . ; uin ], and Aj+1 ∈ R
n
, Ajj ∈ Rn ,
j ×nj−1
Ajj−1 ∈ Rn for j ∈ [N ] satisfying
132
Note that A is a (3N s)-sparse matrix. The dimension of (4.5) is
nN +1 − n
∆ := n + n2 + · · · + nN = = O(nN ). (4.10)
n−1
yin := ŷ(0) = ŷin , and letting all y k be equal for k ∈ [m + p + 1]0 \ [m + 1]0 , for some
sufficiently large integer p. (It is unclear whether another discretization could improve
that encodes (4.11) and uses it to produce a numerical solution at time T , where
m+p m m+p
X X X
L= |k⟩⟨k| ⊗ I − |k⟩⟨k − 1| ⊗ [I + A((k − 1)h)h] − |k⟩⟨k − 1| ⊗ I (4.13)
k=0 k=1 h=m+1
and
m
1 X
|B⟩ = √ ∥yin ∥|0⟩ ⊗ |yin ⟩ + ∥b((k − 1)h)∥|k⟩ ⊗ |b((k − 1)h)⟩ (4.14)
Bm k=1
133
with a normalizing factor Bm . Observe that the system (4.12) has the lower triangular
structure
I y0 yin
1
−[I +A(0)h] I y b(0)
.. ..
.. ..
. .
. .
=
y b((m − 1)h) .
m
−[I +A((m − 1)h)h] I
−I I y m+1 0
.. .. .
..
.
..
. .
−I I y m+p 0
(4.15)
In the above system, the first n components of y k for k ∈ [m + p + 1]0 (i.e., y1k )
quantum linear system algorithm (QLSA) [46] to (4.12) and postselect on k to produce
y1k /∥y1k ∥ for some k ∈ [m + p + 1]0 \ [m]0 . The resulting error is at most
u(T ) yk
ϵ := max − 1k . (4.16)
k∈[m+p+1]0 \[m]0 ∥u(T )∥ ∥y1 ∥
This error includes contributions from both Carleman linearization and the forward Euler
method. (The QLSA also introduces error, which we bound separately. Note that we
could instead apply the original QLSA [84] instead of its subsequent improvement [46],
but this would slightly complicate the error analysis and might perform worse in practice.)
quantum quadratic ODE problem as defined in Problem 4.1, with its Carleman linearization
134
as defined in (4.5). Assume R < 1. Let
∥uin ∥
g := ∥u(T )∥, q := . (4.17)
∥u(T )∥
There exists a quantum algorithm producing a state that approximates u(T )/∥u(T )∥ with
error at most ϵ ≤ 1, succeeding with probability Ω(1), with a flag indicating success,
using at most
queries to the oracles OF2 , OF1 , OF0 and Ox . The gate complexity is larger than the query
complexity by a factor of poly log(nsT ∥F2 ∥∥F1 ∥∥F0 ∥∥F0′ ∥/(1−∥uin ∥)gϵ)/ log(1/∥uin ∥) .
and the gate complexity is larger by poly log(nsT ∥F2 ∥∥F1 ∥∥F0 ∥∥F0′ ∥/gϵ)/ log(1/∥uin ∥) .
In this section we establish several lemmas and use them to prove Theorem 4.1.
135
4.4.1 Solution error
The solution error has three contributions: the error from applying Carleman linearization
to (4.1), the error in the time discretization of (4.5) by the forward Euler method, and the
error from the QLSA. Since the QLSA produces a solution with error at most ϵ with
First, we provide an upper bound for the error from Carleman linearization for
arbitrary evolution time T . To the best of our knowledge, the first and only explicit
bound on the error of Carleman linearization appears in [73]. However, they only consider
homogeneous quadratic ODEs; and furthermore, to bound the error for arbitrary T , they
assume the logarithmic norm of F1 is negative (see Theorems 4.2 and 4.3 of [73]), which
is too strong for our case. Instead, we give a novel analysis under milder conditions,
providing the first convergence guarantee for general inhomogeneous quadratic ODEs.
We begin with a lemma that describes the decay of the solution of (4.1).
Lemma 4.1. Consider an instance of the quadratic ODE (4.1), and assume R < 1 as
p
− Re(λ1 ) ± Re(λ1 )2 − 4∥F2 ∥∥F0 ∥
r± := . (4.20)
2∥F2 ∥
Then r± are distinct real numbers with 0 ≤ r− < r+ , and the solution u(t) of (4.1)
136
Proof. Consider the derivative of ∥u(t)∥. We have
d∥u∥2
= u† F2 (u ⊗ u) + (u† ⊗ u† )F2† u + u† (F1 + F1† )u + u† F0 (t) + F0 (t)† u,
dt
≤ 2∥F2 ∥∥u∥3 + 2 Re(λ1 )∥u∥2 + 2∥F0 ∥∥u∥. (4.21)
If ∥u∥ =
̸ 0, then
d∥u∥
≤ ∥F2 ∥∥u∥2 + Re(λ1 )∥u∥ + ∥F0 ∥. (4.22)
dt
Letting a = ∥F2 ∥ > 0, b = Re(λ1 ) < 0, and c = ∥F0 ∥ > 0 with a, b, c > 0, we consider
dx
= ax2 + bx + c, x(0) = ∥uin ∥. (4.23)
dt
c
Since R < 1 ⇔ b > a∥uin ∥ + ∥uin ∥
, the discriminant satisfies
2 2
2 c c c
b − 4ac > a∥uin ∥ + − 4a∥uin ∥ · ≥ a∥uin ∥ − ≥ 0. (4.24)
∥uin ∥ ∥uin ∥ ∥uin ∥
Thus, r± defined in (4.20) are distinct real roots of ax2 + bx + c. Since r− + r+ = − ab > 0
c
and r− r+ = a
≥ 0, we have 0 ≤ r− < r+ . We can rewrite the ODE as
dx
= ax2 + bx + c = a(x − r− )(x − r+ ), x(0) = ∥uin ∥. (4.25)
dt
137
Letting y = x − r− , we obtain an associated homogeneous quadratic ODE
dy
= −a(r+ − r− )y + ay 2 = ay[y − (r+ − r− )], y(0) = ∥uin ∥ − r− . (4.26)
dt
r+ − r−
y(t) = , (4.27)
1− ea(r+ −r− )t [1 − (r+ − r− )/(∥uin ∥ − r− )]
r+ − r−
x(t) = + r− . (4.28)
1− ea(r+ −r− )t [1 − (r+ − r− )/(∥uin ∥ − r− )]
Therefore we have
r+ − r−
∥u(t)∥ ≤ + r− . (4.29)
1− ea(r+ −r− )t [1 − (r+ − r− )/(∥uin ∥ − r− )]
c
Since R < 1 ⇔ a∥uin ∥ + ∥uin ∥
< −b ⇔ a∥uin ∥2 + b∥uin ∥ + c < 0, ∥uin ∥ is located
between the two roots r− and r+ , and thus 1 − (r+ − r− )/(∥uin ∥ − r− ) < 0. This implies
∥u(t)∥ in (4.29) decreases from u(0) = ∥uin ∥, so we have ∥u(t)∥ < ∥uin ∥ < r+ for any
t > 0.
d∥u∥
We remark that limt→∞ dt
= 0 since d∥u∥
dt
< 0 and ∥u(t)∥ ≥ 0, so u(t) approaches
to a stationary point of the right-hand side of (4.1) (called an attractor in the theory of
dynamical systems).
Note that for a homogeneous equation (i.e., ∥F0 ∥ = 0), this shows that the dissipation
138
inevitably leads to exponential decay. In this case we have r− = 0, so (4.29) gives
∥uin ∥r+
∥u(t)∥ = , (4.30)
ear+ t (r+ − ∥uin ∥) + ∥uin ∥
f1 uj + f0 , j ∈ [n], with uj (0) = x0 > 0, f2 > 0, f1 < 0, f0 > 0, and R < 1. We see that
√
−f1 − f12 −4f2 f0
each uj (t) decreases from x0 to x1 := 2f2
> 0, with 0 < x1 < uj (t) < x0 .
√ √
Hence, the norm of u(t) is bounded as 0 < nx1 < ∥u(t)∥ < nx0 for any t > 0. In
general, it is hard to lower bound ∥u(t)∥, but the above example shows that a nonzero
inhomogeneity can prevent the norm of the solution from decreasing to zero.
Lemma 4.2. Consider an instance of the quadratic ODE (4.1), with its corresponding
Carleman linearization as defined in (4.5). As in Problem 4.1, assume that the eigenvalues
R < 1. Then for any j ∈ [N ], the error ηj (t) := u⊗j (t) − ŷj (t) satisfies
139
Proof. The exact solution u(t) of the original quadratic ODE (4.1) satisfies
1 1
u A1 A2 u F0 (t)
u⊗2 A2 A2 A2 u⊗2 0
1 2 3
u⊗3 A3
A 3
A 3 u⊗3 0
2 3 4
d .
.. .. .. . .
.
. = . . . .. + .. ,
dt
u⊗(N −1) N −1 N −1 N −1 ⊗(N −1)
AN −2 AN −1 AN u
0
.
⊗N
u AN A N . . u⊗N 0
N −1 N
. . .
.. . .. . .. .. ..
(4.32)
and the approximated solution ŷj (t) satisfies (4.6). Comparing these equations, we have
dη
= A(t)η + b̂(t), η(0) = 0 (4.33)
dt
1 1
η1 A1 A2 η1 0
η A2 A2 A2 η 0
2 1 2 3 2
η 3 3 3
d 3 A2 A3 A4 η3
0
= + ,
dt ..
.. .. ..
..
..
. . . . .
.
−1 N −1 N −1
ηN −1 AN A A η 0
N −2 N −1 N N −1
⊗(N +1)
ηN AN
N −1 A N
N η N AN
N +1 u
(4.34)
140
Consider the derivative of ∥η(t)∥. We have
d∥η∥2
= η † (A(t) + A† (t))η + η † b̂(t) + b̂(t)† η. (4.35)
dt
Define a matrix G ∈ Rn×n with nonzero entries Gj−1,j = j∥F0 ∥, Gj,j = j Re(λ1 ),
Since ∥F2 ∥+∥F0 ∥ < |Re (λ1 )|, G is strictly diagonally dominant and thus the eigenvalues
141
⊗(N +1)
Since ∥AN
N +1 ∥ = N ∥F2 ∥, and ∥u ∥ = ∥u∥N +1 ≤ ∥uin ∥N +1 , we have
d∥η∥2
≤ 2 Re(ν1 )∥η∥2 + 2N ∥F2 ∥∥uin ∥N +1 ∥η∥, (4.41)
dt
d∥η∥ 1 d∥η∥2
= ≤ Re(ν1 )∥η∥ + N ∥F2 ∥∥uin ∥N +1 . (4.42)
dt 2∥η∥ dt
Solving the differential inequality as an equation with η(0) = 0 gives us a bound on ∥η∥:
Z t Z t
Re(ν1 )(t−s) N +1 N +1
∥η(t)∥ ≤ e N ∥F2 ∥∥uin ∥ ds ≤ N ∥F2 ∥∥uin ∥ eRe(ν1 )(t−s) ds.
0 0
(4.43)
Finally, using
t
1 − eRe(ν1 )t
Z
eRe(ν1 )(t−s) ds = ≤t (4.44)
0 |Re(ν1 )|
(where we used the inequality 1 − eat ≤ −at with a < 0), (4.43) gives the bound
as claimed.
142
Note that (4.44) can be bounded alternatively by
t
1 − eRe(ν1 )t
Z
1
eRe(ν1 )(t−s) ds = ≤ , (4.46)
0 |Re(ν1 )| |Re(ν1 )|
1
and thus ∥ηj (t)∥ ≤ ∥η(t)∥ ≤ |Re(ν1 )|
N ∥F2 ∥∥uin ∥N +1 . We select (4.44) because it avoids
We also give an improved analysis that works for homogeneous quadratic ODEs
(F0 (t) = 0) under milder conditions. This analysis follows the proof in [73] closely.
Corollary 4.1. Under the same setting of Lemma 4.2, assume F0 (t) = 0 in (4.1). Then
for any j ∈ [N ], the error ηj (t) := u⊗j (t) − ŷj (t) satisfies
N
∥η1 (t)∥ ≤ ∥uin ∥ RN 1 − eRe(λ1 )t . (4.48)
Proof. We again consider η satisfying (4.33). Since F0 (t) = 0, (4.33) reduces to a time-
dηj
= Ajj ηj + Ajj+1 ηj+1 , j ∈ [N − 1] (4.49)
dt
and
dηN ⊗(N +1)
= AN N
N ηN + AN +1 u . (4.50)
dt
143
We proceed by backward substitution. Since ηN (0) = 0, we have
Z t
N ⊗(N +1)
ηN (t) = eAN (t−s0 ) AN
N +1 u (s0 ) ds0 . (4.51)
0
j
For j ∈ [N ], (4.8) gives ∥eAj t ∥ = ej Re(λ1 )t and (4.7) gives ∥Ajj+1 ∥ = j∥F2 ∥. By
Lemma 4.1, ∥u⊗(N +1) ∥ = ∥u∥N +1 ≤ ∥uin ∥N +1 . We can therefore upper bound (4.51)
by
Z t
N ⊗(N +1)
∥ηN (t)∥ ≤ ∥eAN (t−s0 ) ∥ · ∥AN
N +1 u (s0 )∥ ds0
0
(4.52)
Z t
≤ N ∥F2 ∥∥uin ∥N +1 eN Re(λ1 )(t−s0 ) ds0 .
0
dηN −1 −1 N −1
= AN
N −1 ηN −1 + AN ηN . (4.53)
dt
Z t
N −1
−1
ηN −1 (t) = eAN −1 (t−s1 ) AN
N ηN (s1 ) ds1 (4.54)
0
Z t
N −1
−1
∥ηN −1 (t)∥ ≤ ∥eAN −1 (t−s1 ) ∥ · ∥AN
N ηN (s1 )∥ ds1
0
Z t
≤ (N − 1)∥F2 ∥ e(N −1) Re(λ1 )(t−s1 ) ∥ηN (s1 )∥ ds1
0
Z t Z s1
(N −1) Re(λ1 )(t−s1 )
≤ N (N − 1)∥F2 ∥ ∥uin ∥ 2 N +1
e eN Re(λ1 )(s1 −s) ds0 ds1 ,
0 0
(4.55)
where we used (4.52) in the last step. Iterating this procedure for j = N − 2, . . . , 1, we
144
find
Z t Z sN −j
N!
∥ηj (t)∥ ≤ ∥F2 ∥ N +1−j
∥uin ∥ N +1
ej Re(λ1 )(t−sN −j )
e(j+1) Re(λ1 )(sN −j −sN −1−j )
(j − 1)! 0 0
Z s2 Z s1
··· e(N −1) Re(λ1 )(s2 −s1 ) eN Re(λ1 )(s1 −s0 ) ds0 · · · dsN −j
0 0
Z sN +1−j Z s2Z s1
N! PN −j+1
= ∥F2 ∥N +1−j ∥uin ∥N +1 ··· eRe(λ1 )(−N s0 + k=1 sk )
ds0 · · · dsN −j .
(j − 1)! 0 0 0
(4.56)
Finally, using
sk+1
1 − e(N −k) Re(λ1 )sk+1
Z
1
e(N −k) Re(λ1 )(sk+1 −sk ) dsk = ≤ (4.57)
0 (N − k)|Re(λ1 )| (N − k)|Re(λ1 )|
N! (j − 1)!
∥ηj (t)∥ ≤ ∥F2 ∥N +1−j ∥uin ∥N +1
(j − 1)! N !|Re(λ1 )|N +1−j
(4.58)
∥uin ∥N +1 ∥F2 ∥N +1−j
= = ∥uin ∥j RN +1−j .
|Re(λ1 )|N +1−j
sN s2 s1
(easN − 1)N
Z Z Z PN
··· ea(−N s0 + k=1 sk ) ds0 ds1 · · · dsN −1 = . (4.59)
0 0 0 N !aN
145
With sN = t and a = Re(λ1 ), we find
Z sN Z s2 Z s1 PN
N N +1
∥η1 (t)∥ ≤ N !∥F2 ∥ ∥uin ∥ ··· eRe(λ1 )(−N s0 + k=1 sk ) ds0 ds1 · · · dsN −1
0 0 0
Re(λ1 )t N
(e − 1)
≤ N !∥F2 ∥N ∥uin ∥N +1
N ! Re(λ1 )N
N
= ∥uin ∥ RN 1 − eat ,
(4.60)
While Problem 4.1 makes some strong assumptions about the system of differential
equations, they appear to be necessary for our analysis. Specifically, the conditions
Since the Euler method for (4.5) is unstable if Re (λ1 ) > 0 [52, 119], we only
sN s2 s1
tN
Z Z Z PN
a(−N s0 + k=1 sk )
··· e ds0 ds1 · · · dsN −1 = , (4.61)
0 0 0 N!
instead of (4.60). Then the error bound can be made arbitrarily small for a finite time by
increasing N , but after t > 1/∥uin ∥∥F2 ∥, the error bound diverges.
∥uin ∥(1 − N eRe(λ1 )t ) ≤ ∥uin ∥(1 − N eRe(λ1 )t ) RN ≤ ∥uin ∥(1 − eRe(λ1 )t )N RN , (4.63)
146
where the right-hand side upper bounds ∥η1 (t)∥ as in (4.60). Assuming ∥η1 (t)∥ =
In other words, to apply (4.48) for the Carleman linearization procedure, the truncation
based on a technique other than Carleman linearization) can solve Problem 4.1 efficiently.
√
It remains open to understand the complexity of the problem for 1 ≤ R < 2.
On the other hand, if R < 1, both (4.47) and (4.48) decrease exponentially with N ,
making the truncation efficient. We discuss the specific choice of N in (4.149) below.
Next, we provide an upper bound for the error incurred by approximating (4.5) with
the forward Euler method. This problem has been well studied for general ODEs. Given
dz
an ODE dt
= f (z) on [0, T ] with an inhomogenity f that is an L-Lipschitz continuous
function of z, the global error of the solution is upper bounded by eLT , although in
most cases this bound overestimates the actual error [120]. To remove the exponential
dependence on T in our case, we derive a tighter bound for time discretization of (4.5) in
Lemma 4.3 below. This lemma is potentially useful for other ODEs as well and can be
Lemma 4.3. Consider an instance of the quantum quadratic ODE problem as defined in
147
Problem 4.1, with R < 1 as defined in (4.2). Choose a time step
1 2(| Re(λ1 )| − ∥F2 ∥ − ∥F0 ∥)
h ≤ min , (4.65)
N ∥F1 ∥ N (| Re(λ1 )|2 − (∥F2 ∥ + ∥F0 ∥)2 + ∥F1 ∥2 )
in general, or
1
h≤ (4.66)
N ∥F1 ∥
if the eigenvalues of F1 are all real. Suppose the error from Carleman linearization η(t)
g
∥η(t)∥ ≤ , (4.67)
4
where g is defined in (4.17). Then the global error from the forward Euler method (4.11)
∥ŷ1 (T ) − y1m ∥ ≤ ∥ŷ(T ) − y m ∥ ≤ 3N 2.5 T h[(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥]. (4.68)
Proof. We define a linear system that locally approximates the initial value problem (4.5)
where ŷ(t) is the exact solution of (4.5). For k ∈ [m], we denote the local truncation error
148
by
ek := ∥ŷ(kh) − z k ∥ (4.70)
g k := ∥ŷ(kh) − y k ∥, (4.71)
For the local truncation error, we Taylor expand ŷ((k − 1)h) with a second-order
ŷ ′′ (ξ)h2
ŷ(kh) = ŷ((k − 1)h) + ŷ ′ ((k − 1)h)h + (4.72)
2
for some ξ ∈ [(k − 1)h, kh]. Since ŷ ′ ((k − 1)h) = A((k − 1)h)ŷ((k − 1)h) + b((k − 1)h)
by (4.5), we have
ŷ ′′ (ξ)h2 ŷ ′′ (ξ)h2
ŷ(kh) = [I +A((k −1)h)h]ŷ((k −1)h)+b((k −1)h)+ = zk + , (4.73)
2 2
h2 M h2
ek = ∥ŷ(kh) − z k ∥ = ∥ŷ ′′ (ξ)h∥ ≤ , (4.74)
2 2
By the triangle inequality, the global error (4.71) can therefore be bounded as
149
Since y k and z k are obtained by the same linear system with different right-hand sides,
In order to provide an upper bound for ∥I + A(t)h∥ for all t ∈ [0, T ], we write
where
N
X −1
H2 = |j⟩⟨j + 1| ⊗ Ajj+1 h, (4.78)
j=1
N
X
H1 = I + |j⟩⟨j| ⊗ Ajj h, (4.79)
j=1
N
X
H0 (t) = |j⟩⟨j − 1| ⊗ Ajj−1 h. (4.80)
j=2
We provide upper bounds separately for ∥H2 ∥, ∥H1 ∥, and ∥H0 ∥ := maxt∈[0,T ] ∥H0 (t)∥,
The eigenvalues of Ajj consist of all j-term sums of the eigenvalues of F1 . More
150
With J := maxℓ∈[n] | Im(λℓ )|, we have
X 2 X 2 X 2
1+h λI j = 1+h Re(λI j ) + h Im(λI j )
ℓ ℓ ℓ
ℓ∈[j] ℓ∈[j] ℓ∈[j] (4.81)
≤ 1 − 2N h| Re(λ1 )| + N 2 h2 (| Re(λ1 )|2 + J 2 )
X p
∥H1 ∥ = max max
j j
1+h λI j ≤ 1 − 2N h| Re(λ1 )| + N 2 h2 (| Re(λ1 )|2 + J 2 ).
j∈[N ] I ∈[n] ℓ
ℓ∈[j]
(4.82)
We also have
N
X −1
∥H2 ∥ = |j⟩⟨j + 1| ⊗ Ajj+1 h ≤ max ∥Ajj+1 ∥h ≤ N ∥F2 ∥h (4.83)
j∈[N ]
j=1
and
N
X −1
∥H0 ∥ = max |j⟩⟨j01|⊗Ajj−1 h ≤ max max ∥Ajj−1 ∥h ≤ N max ∥F0 (t)∥h ≤ N ∥F0 ∥h.
t∈[0,T ] t∈[0,T ] j∈[N ] t∈[0,T ]
j=1
(4.84)
Using the bounds (4.82) and (4.83), we aim to select the value of h to ensure
151
(note that the denominator is non-zero due to (4.3)). Then we have
if J = 0, (4.86) simplifies to
2
h≤ , (4.88)
N (|λ1 | + ∥F2 ∥ + ∥F0 ∥)
k
X
k k−1 k k−2 k−1 k
g ≤g +e ≤g +e + e ≤ ··· ≤ ej , k ∈ [m + 1]0 . (4.90)
j=1
Using (4.74), this shows that the global error from the forward Euler method is bounded
by
k
X M kh2
∥ŷ 1 (kh) − y1k ∥ ≤ ∥ŷ(kh) − y k ∥ = g k ≤ ej ≤ , (4.91)
j=1
2
152
and when k = m, mh = T ,
M h2 MT h
∥ŷ 1 (T ) − y1m ∥ ≤ g m ≤ m = . (4.92)
2 2
we have
∥ŷ ′′ (t)∥ = ∥A(t)∥2 ∥ŷ(t)∥ + ∥A(t)∥∥b(t)∥ + ∥A′ (t)∥∥ŷ(t)∥ + ∥b′ (t)∥. (4.94)
N
X −1 N
X N
X
max ∥A(t)∥ = |j⟩⟨j + 1| ⊗ Ajj+1 + |j⟩⟨j| ⊗ Ajj + |j⟩⟨j − 1| ⊗ Ajj−1
t∈[0,T ]
j=1 j=1 j=2
and using ∥u∥ ≤ ∥uin ∥ < 1, ∥ηj (t)∥ ≤ ∥η(t)∥ ≤ g/4 < ∥uin ∥/4 by (4.67), and R < 1,
153
we have
N
X N
X N
X
∥ŷ(t)∥2 ≤ ∥ŷj (t)∥2 = ∥u⊗j (t) − ηj (t)∥2 ≤ 2 (∥u⊗j (t)∥2 + ∥ηj (t)∥2 )
j=1 j=1 j=1
N N
∥uin ∥2
X
2j
X 1
≤2 ∥uin ∥ + <2 (1 + )∥uin ∥2 < 4N ∥uin ∥2 < 4N
j=1
16 j=1
16
(4.99)
for all t ∈ [0, T ]. Therefore, substituting the bounds (4.95)–(4.99) into (4.94), we find
≤ 2N 2.5 (∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + N (∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)∥F0 ∥ + 2N 1.5 ∥F0′ ∥ + ∥F0′ ∥
∥ŷ 1 (kh) − y1m ∥ ≤ 3N 2.5 kh2 [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥], (4.101)
and when k = m, mh = T ,
as claimed.
154
Lemma 4.4. Consider an instance of the quantum quadratic ODE problem as defined in
Problem 4.1. Apply the forward Euler method (4.11) with time step (4.65) to the Carleman
linearization (4.5). Then the condition number of the matrix L defined in (4.12) satisfies
L = L1 + L2 + L3 , (4.104)
where
m+p
X
L1 = |k⟩⟨k| ⊗ I, (4.105)
k=0
Xm
L2 = − |k⟩⟨k − 1| ⊗ [I + A((k − 1)h)h], (4.106)
k=1
m+p
X
L3 = − |k⟩⟨k − 1| ⊗ I. (4.107)
k=m+1
155
Next we upper bound
We express |B⟩ as
m+p m+p
X X
|B⟩ = βk |k⟩ = |bk ⟩, (4.110)
k=0 k=0
m+p
X
∥|bk ⟩∥2 = ∥|B⟩∥2 ≤ 1. (4.111)
k=0
m+p m+p
X X
|Y k ⟩ := L−1 |bk ⟩ = γlk |l⟩ = |Ylk ⟩, (4.112)
l=0 l=0
where |Ylk ⟩ := γlk |l⟩. We first upper bound ∥|Y k ⟩∥ = ∥L−1 |bk ⟩∥, and then use this to
We consider two cases. First, for fixed k ∈ [m + 1]0 , we directly calculate |Ylk ⟩ for
0, if l ∈ [k]0 ;
|bk ⟩, if l = k;
|Ylk ⟩ = (4.113)
l−1
Πj=k [I + A(jh)h]|bk ⟩, if l ∈ [m + 1]0 \ [k + 1]0 ;
Πm−1 k
j=k [I + A(jh)h]|b ⟩, if l ∈ [m + p + 1]0 \ [m + 1]0 .
156
Since maxt∈[0,T ] ∥I + A(t)h∥| ≤ 1 by (4.85), (4.112) gives
m+p m m+p
X X X
k 2 k 2 k 2
∥|Y ⟩∥ = ∥|Yl ⟩∥ ≤ ∥|b ⟩∥ + ∥|bk ⟩∥2
l=0 l=k l=m+1 (4.114)
≤ (m + p + 1 − k)∥|bk ⟩∥2 ≤ (m + p + 1)∥|bk ⟩∥2 .
0, if l ∈ [k]0 ;
k
|Yl ⟩ = (4.115)
|bk ⟩, if l ∈ [m + p + 1]0 \ [k]0 .
m+p m+p
X X
k 2 k 2
∥|Y ⟩∥ = ∥|Yl ⟩∥ = ∥|bk ⟩∥2 = (m + p + 1 − k)∥|bk ⟩∥2 ≤ (m + p + 1)∥|bk ⟩∥2 .
l=0 l=k
(4.116)
∥L−1 ∥2 = sup ∥L−1 |B⟩∥2 ≤ (m+p+1) sup ∥L−1 |bk ⟩∥2 ≤ (m+p+1)2 , (4.118)
∥|B⟩∥≤1 ∥|bk ⟩∥≤1
157
and therefore
as claimed.
We now describe a procedure for preparing the right-hand side |B⟩ of the linear
system (4.12), whose entries are composed of |yin ⟩ and |b((k − 1)h)⟩ for k ∈ [m].
The initial vector yin is a direct sum over spaces of different dimensions, which is
tensor product structure. Specifically, we embed yin into a slightly larger space and
where v0 is some standard vector (for simplicity, we take v0 = |0⟩). If uin lives in a vector
space of dimension n, then zin lives in a space of dimension N nN while yin lives in a
standard techniques, all the operations we would otherwise apply to yin can be applied
158
Lemma 4.5. Assume we are given the value ∥uin ∥, and let Ox be an oracle that maps
are also given the values ∥F0 (t)∥ for each t ∈ [0, T ], and let OF0 be an oracle that
provides the locations and values of the nonzero entries of F0 (t) for any specified t.
Then the quantum state |B⟩ defined in (4.14) (with yin replaced by zin ) can be prepared
using O(N ) queries to Ox and O(m) queries to OF0 , with gate complexity larger by a
N
1 X
|zin ⟩ = √ ∥uin ∥j |j⟩|uin ⟩⊗j |0⟩⊗N −j , (4.122)
V j=1
where
N
X
V := ∥uin ∥2j . (4.123)
j=1
This state can be prepared using N queries to the initial state oracle Ox applied in superposition
N
1 X
|ψint ⟩ := √ ∥uin ∥j |j⟩ ⊗ |0⟩⊗N . (4.124)
V j=1
1 k−1
∥uin ∥ X Y ℓ
|ψint ⟩ = √ ∥uin ∥jℓ 2 |j0 j1 . . . jk−1 ⟩ ⊗ |0⟩⊗N , (4.125)
V j0 ,j1 ,...,jk−1 =0 ℓ=0
159
the k-bit binary expansion of j − 1. Observe that
k−1 1
O 1 X jℓ 2ℓ
|ψint ⟩ = √ ∥uin ∥ |jℓ ⟩ ⊗ |0⟩⊗N (4.126)
ℓ=0
Vℓ j =0
ℓ
where
ℓ+1
Vℓ := 1 + ∥uin ∥2 . (4.127)
Qk−1
(Notice that ℓ=0 Vℓ = V /∥uin ∥2 .) Each tensor factor in (4.126) is a qubit that can be
produced in constant time. Overall, we prepare these k = log2 N qubit states and then
apply Ox N times.
m
1 X
|B⟩ = √ ∥zin ∥|0⟩ ⊗ |zin ⟩ + ∥b((k − 1)h)∥|k⟩ ⊗ |b((k − 1)h)⟩ , (4.128)
Bm k=1
m
X
2
Bm := ∥zin ∥ + ∥b((k − 1)h)∥2 . (4.129)
k=1
This state can be prepared using the above procedure for |0⟩ 7→ |zin ⟩ and m queries to
OF0 with t = (k − 1)h that implement |0⟩ 7→ |b((k − 1)h)⟩ for k ∈ {1, . . . , m}, applied
m
1 X
|ϕint ⟩ = √ ∥zin ∥|0⟩ ⊗ |0⟩ + ∥b((k − 1)h)∥|k⟩ ⊗ |0⟩ . (4.130)
Bm k=1
160
Here the queries are applied conditionally upon the value in the first register: we prepare
|zin ⟩ if the first register is |0⟩ and |b((k−1)h)⟩ if the first register is |k⟩ for k ∈ {1, . . . , m}.
We can prepare |ϕint ⟩ (i.e., perform a unitary transform mapping |0⟩|0⟩ 7→ |ϕint ⟩) in time
complexity O(m) [92] using the known values of ∥uin ∥ and ∥b((k − 1)h)∥.
Overall, we use O(N ) queries to Ox and O(m) queries to OF0 to prepare |B⟩. The
state of the desired form. We now consider the probability of this measurement succeeding.
Lemma 4.6. Consider an instance of the quantum quadratic ODE problem defined in
Problem 4.1, with the QLSA applied to the linear system (4.12) using the forward Euler
method (4.11) with time step (4.65). Suppose the error from Carleman linearization
g
satisfies ∥η(t)∥ ≤ 4
as in (4.67), and the global error from the forward Euler method
g
∥ŷ(T ) − y m ∥ ≤ , (4.131)
4
where g is defined in (4.17). Then the probability of measuring a state |y1k ⟩ for k =
p+1
Pmeasure ≥ , (4.132)
9(m + p + 1)N q 2
161
where q is also defined in (4.17).
Proof. The idealized quantum state produced by the QLSA applied to (4.12) has the form
m+p m+p N
X XX
k
|Y ⟩ = |y ⟩|k⟩ = |yjk ⟩|j⟩|k⟩ (4.133)
k=0 k=0 j=1
where the states |y k ⟩ and |yjk ⟩ for k ∈ [m + p + 1]0 and j ∈ [N ] are subnormalized to
ensure ∥|Y ⟩∥ = 1.
where
m−1 N m+p N
XX XX
|Ybad ⟩ := |yjk ⟩|j⟩|k⟩ + |yjk ⟩|j⟩|k⟩,
k=0 j=1 k=m j=2
m+p
(4.135)
X
|Ygood ⟩ := |y1k ⟩|1⟩|k⟩.
k=m
Note that |y1k ⟩ = |y1m ⟩ for all k ∈ {m, m + 1, . . . , m + p}. We lower bound
First, according to (4.67) and (4.131), the exact solution u(T ) and the approximate
162
solution y1m defined in (4.12) satisfy
g
∥u(T )−y1m ∥ ≤ ∥u(T )− ŷ1 (T )∥+∥ŷ1 (T )−y1m ∥ ≤ ∥η(t)∥+∥ŷ(T )−y m ∥ ≤ . (4.138)
2
Using ∥ŷ(t)∥2 < 4N ∥uin ∥2 by (4.99), and ∥y(kh)−y k ∥ ≤ ∥ŷ(T )−y m ∥ ≤ g/4 < ∥uin ∥/4
∥uin ∥2
k 2 2
∥y ∥ ≤ 2 4N ∥uin ∥ + < 9N ∥uin ∥2 . (4.141)
16
Therefore
p+1
Pmeasure ≥ (4.143)
9(m + p + 1)N q 2
163
as claimed.
Proof. We first present the quantum Carleman linearization (QCL) algorithm and then
The QCL algorithm. We start by rescaling the system to satisfy (4.3) and (4.4). Given a
quadratic ODE (4.1) satisfying R < 1 (where R is defined in (4.2)), we define a scaling
u := γu. (4.144)
du 1
= F2 u⊗2 + F1 u + γF0 (t),
dt γ
(4.145)
u(0) = uin := γuin .
du
= F 2 u⊗2 + F 1 u + F 0 (t),
dt (4.146)
u(0) = uin .
Note that R is invariant under this rescaling, so R < 1 still holds for the rescaled equation.
164
Concretely, we take3
1
γ=p . (4.147)
∥uin ∥r+
After rescaling, the new quadratic ODE satisfies ∥uin ∥r+ = γ 2 ∥uin ∥r+ = 1. Since
∥uin ∥ < r+ by Lemma 4.1, we have r− < ∥uin ∥ < 1 < r+ , so (4.4) holds. Furthermore, 1
is located between the two roots r− and r+ , which implies ∥F 2 ∥·12 +|Re (λ1 )|·1+∥F 0 ∥ <
Having performed this rescaling, we henceforth assume that (4.3) and (4.4) are
satisfied. We then introduce the choice of parameters as follows. Given g and an error
bound ϵ ≤ 1, we define
gϵ g
δ := ≤ . (4.148)
1+ϵ 2
log(2T ∥F2 ∥/δ) log(2T ∥F2 ∥/δ)
N= = . (4.149)
log(1/∥uin ∥) log(r+ )
Since ∥uin ∥/δ > 1 by (4.148) and g < ∥uin ∥, Lemma 4.2 gives
1 N +1 δ
∥u(T ) − ŷ1 (T )∥ ≤ ∥η(T )∥ ≤ T N ∥F2 ∥∥uin ∥N +1 = T N ∥F2 ∥( ) ≤ . (4.150)
r+ 2
Now we discuss the choice of h. On the one hand, h must satisfy (4.65) to satisfy
the conditions of Lemma 4.3 and Lemma 4.4. On the other hand, Lemma 4.3 gives the
1 1
3
In fact, one can show that any γ ∈ r+ , ∥uin ∥ suffices to satisfy (4.3) and (4.4).
165
upper bound
gϵ gϵ δ
∥ŷ1 (T ) − y1m ∥ ≤ 3N 2.5 T h[(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥] ≤ ≤ = .
4 2(1 + ϵ) 2
(4.151)
This also ensures that (4.131) holds since δ ≤ g/2. Thus, we choose
gϵ 1
h ≤ min ′
, ,
12N 2.5 T [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0 ∥] N ∥F1 ∥
(4.152)
2(| Re(λ1 )| − ∥F2 ∥ − ∥F0 ∥)
N (| Re(λ1 )|2 − (∥F2 ∥ + ∥F0 ∥)2 + ∥F1 ∥2 )
u(T ) y1m δ
− m ≤ = ϵ, (4.155)
∥u(T )∥ ∥y1 ∥ g−δ
u(T )
i.e., the normalized output state is ϵ-close to ∥u(T )∥
.
We follow the procedure in Lemma 4.5 to prepare the initial state |ŷin ⟩. We apply
166
the QLSA [46] to the linear system (4.12) with m = p = ⌈T /h⌉, giving a solution
[m + p + 1]0 and j ∈ [N ]. By Lemma 4.6, the probability of obtaining a state |y1k ⟩ for
p+1 1
Pmeasure ≥ 2
≥ . (4.156)
9(m + p + 1)N q 18N q 2
√
By amplitude amplification, we can achieve success probability Ω(1) with O( N q)
Analysis of the complexity. By Lemma 4.5, the right-hand side |B⟩ in (4.12) can be
prepared with O(N ) queries to Ox and O(m) queries to OF0 , with gate complexity larger
matrix with O(N s) nonzero entries in any row or column. By Lemma 4.4 and our choice
3(m + p + 1)
N T [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥]
2.5 2
=O + N T ∥F1 ∥
δ
N T [| Re(λ1 )|2 − (∥F2 ∥ + ∥F0 ∥)2 + ∥F1 ∥2 ]
+
2(| Re(λ1 )| − ∥F2 ∥ − ∥F0 ∥)
N T [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥]
2.5 2
N T ∥F1 ∥2
1
=O + ·
gϵ (1 − ∥uin ∥)2 ∥F2 ∥ + ∥F0 ∥
N T [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥]
2.5 2
=O .
(1 − ∥uin ∥)2 (∥F2 ∥ + ∥F0 ∥)gϵ
(4.157)
167
Here we use ∥F2 ∥ + ∥F0 ∥ < | Re(λ1 )| ≤ ∥F1 ∥ and
1
2(| Re(λ1 )| − ∥F2 ∥ − ∥F0 ∥) > (∥uin ∥ + − 2)(∥F2 ∥ + ∥F0 ∥)
∥uin ∥
(4.158)
1
= (1 − ∥uin ∥)2 (∥F2 ∥ + ∥F0 ∥) > (1 − ∥uin ∥)2 (∥F2 ∥ + ∥F0 ∥).
∥uin ∥
The first inequality above is from the sum of | Re(λ1 )| > ∥F2 ∥∥uin ∥ + ∥F0 ∥/∥uin ∥ and
N 3.5 sT 2 [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥] N sT ∥F2 ∥∥F1 ∥∥F0 ∥∥F0′ ∥
poly log
(1 − ∥uin ∥)2 (∥F2 ∥ + ∥F0 ∥)gϵ (1 − ∥uin ∥)gϵ
sT 2 [(∥F2 ∥ + ∥F1 ∥ + ∥F0 ∥)2 + ∥F0′ ∥] sT ∥F2 ∥∥F1 ∥∥F0 ∥∥F0′ ∥
= poly log / log(1/∥uin ∥)
(1 − ∥uin ∥)2 (∥F2 ∥ + ∥F0 ∥)gϵ (1 − ∥uin ∥)gϵ
(4.159)
√
queries to the oracles OF2 , OF1 , OF0 , and Ox . Using O( N q) steps of amplitude amplification
to achieve success probability Ω(1), the overall query complexity of our algorithm is
and the gate complexity exceeds this by a factor of poly log(nsT ∥F2 ∥∥F1 ∥∥F0 ∥∥F0′ ∥/(1−
R)gϵ)/ log(1/∥uin ∥) .
168
most
queries to the oracles OF2 , OF1 , OF0 , and Ox . Using amplitude amplification to achieve
and the gate complexity is larger by poly log(nsT ∥F2 ∥∥F1 ∥∥F0 ∥∥F0′ ∥/gϵ)/ log(1/∥uin ∥)
as claimed.
solve the quadratic ODE problem when the nonlinearity is sufficiently strong. We quantify
the strength of the nonlinearity in terms of the quantity R defined in (4.2). Whereas there
is an efficient quantum algorithm for R < 1 (as shown in Theorem 4.1), we show here
√
that the problem is intractable for R ≥ 2.
169
√
Theorem 4.2. Assume R ≥ 2. Then there is an instance of the quantum quadratic
ODE problem defined in Problem 4.1 such that any quantum algorithm for producing
a quantum state approximating u(T )/∥u(T )∥ with bounded error must have worst-case
We establish this result by showing how the nonlinear dynamics can be used to
distinguish nonorthogonal quantum states, a task that requires many copies of the given
state. Note that since our algorithm only approximates the quantum state corresponding
to the solution, we must lower bound the query complexity of approximating the solution
of a quadratic ODE.
that the ability to distinguish non-orthogonal states can be applied to solve unstructured
search (and other hard computational problems) [78, 79, 80]. Here we show a similar
Lemma 4.7. Let |ψ⟩, |ϕ⟩ be states of a qubit with |⟨ψ|ϕ⟩| = 1 − ϵ. Suppose we are
either given a black box that prepares |ψ⟩ or a black box that prepares |ϕ⟩. Then any
bounded-error protocol for determining whether the state is |ψ⟩ or |ϕ⟩ must use Ω(1/ϵ)
queries.
Proof. Using the black box k times, we prepare states with overlap (1 − ϵ)k . By the well-
known relationship between fidelity and trace distance, these states have trace distance at
p √
most 1 − (1 − ϵ)2k ≤ 2kϵ. Therefore, by the Helstrom bound (which states that the
170
advantage over random guessing for the best measurement to distinguish two quantum
states is given by their trace distance [121]), we need k = Ω(1/ϵ) to distinguish the states
Lemma 4.7 can be used to establish limitations on the ability of quantum computers
dynamics (such as the Weinberg model [79, 80] and the Gross-Pitaevskii equation [78]),
here we aim to show the difficulty of efficiently simulating more general nonlinear ODEs—
Lemma 4.8. There exists an instance of the quantum quadratic ODE problem as defined
√
in Problem 4.1 with R ≥ 2, and two states of a qubit with overlap 1 − ϵ (for 0 < ϵ <
√
1 − 3/ 10) as possible initial conditions, such that the two final states after evolution
√
time T = O(log(1/ϵ)) have an overlap no larger than 3/ 10.
du1
= −u1 + ru21 ,
dt (4.164)
du2
= −u2 + ru22 ,
dt
for some r > 0, with an initial condition u(0) = [u1 (0); u2 (0)] = uin satisfying ∥uin ∥ = 1.
171
parameter as R. The analytic solution of (4.164) is
1
u1 (t) = ,
R −et (R −1/u1 (0))
(4.165)
1
u2 (t) = .
R −et (R −1/u2 (0))
∗ R
0 ≤ t < t := log ; (4.166)
R −1/u2 (0)
when u2 (0) = 1/ R, we have u2 (t) = 1/ R for all t; and when u2 (0) < 1/ R, u2 (t) goes
Without loss of generality, we assume u1 (0) ≤ u2 (0). For u2 (0) ≥ u1 (0) > 1/ R,
both u1 (t) and u2 (t) are finite within the domain (4.166), which we consider as the domain
of u(t).
Now we consider 1-qubit states that provide inputs to (4.164). Given a sufficiently
θ
2 sin2 = ϵ. (4.167)
2
1
|ϕ(0)⟩ = √ (|0⟩ + |1⟩) (4.168)
2
172
and
π π
|ψ(0)⟩ = cos θ + |0⟩ + sin θ + |1⟩. (4.169)
4 4
√
The initial overlap (4.170) is larger than the target overlap 3/ 10 in Lemma 4.8 provided
√
ϵ < 1 − 3/ 10. For simplicity, we denote
π
v0 := cos θ + ,
4 (4.171)
π
w0 := sin θ + ,
4
and let v(t) and w(t) denote solutions of (4.164) with initial conditions v(0) = v0 and
w(0) = w0 , respectively. Since w0 > 1/ R, we see that w(t) increases with t, satisfying
1 1
≤ √ < w0 < w(t), (4.172)
R 2
and
We now study the outputs of our problem. For the state |ϕ(0)⟩, the initial condition
173
√ √
for (4.164) is [1/ 2; 1/ 2]. Thus, the output for any t ≥ 0 is
1
|ϕ(t)⟩ = √ (|0⟩ + |1⟩). (4.174)
2
For the state |ψ(0)⟩, the initial condition for (4.164) is [v0 ; w0 ]. We now discuss
how to select a terminal time T to give a useful output state |ψ(T )⟩. For simplicity, we
w(t)
K(t) := . (4.175)
v(t)
Noticing that w(t) goes to infinity as t approaches t∗ , while v(t) remains finite within
K(T ) ≥ 2. (4.176)
1
|ψ(T )⟩ = p (|0⟩ + K(T )|1⟩). (4.177)
K(T )2 + 1
K(T ) + 1 3
⟨ϕ(T )|ψ(T )⟩ = p ≤√ (4.178)
2K(T )2 + 2 10
4
More concretely, we take vmax = max{v0 , v(t∗ )} that upper bounds v(t) on the domain [0, t∗ ), in
which v(t∗ ) is a finite value since v0 < w0 . Then there exists a terminal time T such that w(T ) = 2vmax ,
and hence K(T ) = w(T )/v(T ) ≥ 2.
174
using (4.176).
√
∗ R 2
T < t = log < log √ (4.179)
R − w10 2 − w10
since the function log(x/(x − c)) decreases monotonically with x for x > c > 0. Using
√ !
∗ 2 1
T < t < log √ 1
= log 1 + √ , (4.180)
2− sin(θ+ π4 )
2ϵ − ϵ2 − ϵ
Proof. As introduced in the proof of Lemma 4.8, consider the quadratic ODE (4.164); the
two initial states of a qubit |ϕ(0)⟩ and |ψ(0)⟩ defined in (4.168) and (4.169), respectively;
Suppose we have a quantum algorithm that, given a black box to prepare a state
that is either |ϕ(0)⟩ or |ψ(0)⟩, can produce quantum states |ϕ′ (T )⟩ or |ψ ′ (T )⟩ that are
within distance δ of |ϕ(T )⟩ and |ψ(T )⟩, respectively. Since by Lemma 4.8, |ϕ(T )⟩ and
|ψ(T )⟩ have constant overlap, the overlap between |ϕ′ (T )⟩ and |ψ ′ (T )⟩ is also constant
175
for sufficiently small δ. More precisely, we have
3
⟨ϕ(T )|ψ(T )⟩ ≤ √ (4.181)
10
s
3
∥|ϕ(T )⟩ − |ψ(T )⟩ ∥ ≥ 2 1 − √ > 0.32. (4.182)
10
We also have
∥ |ϕ′ (T )⟩ − |ψ ′ (T )⟩ ∥ = ∥(|ϕ(T )⟩ − |ψ(T ))⟩ − (|ϕ(T )⟩ − |ϕ′ (T ))⟩ − (|ψ ′ (T )⟩ − |ψ(T ))⟩ ∥
Lemma 4.7 therefore shows that preparing the states |ϕ′ (T )⟩ and |ψ ′ (T )⟩ requires
time Ω(1/ϵ), as these states can be used to distinguish the two possibilities with bounded
error. By Lemma 4.8, this time is 2Ω(T ) . This shows that we need at least exponential
simulation time to approximate the solution of arbitrary quadratic ODEs to within sufficiently
√
small bounded error when R ≥ 2.
Note that exponential time is achievable since our QCL algorithm can solve the
176
problem by taking N to be exponential in T , where N is the truncation level of Carleman
linearization. (The algorithm of Leyton and Osborne also solves quadratic differential
equations with complexity exponential in T , but requires the additional assumptions that
4.6 Applications
Due to the assumptions of our analysis, our quantum Carleman linearization algorithm
can only be applied to problems with certain properties. First, there are two requirements
have linear dissipation, manifested by Re(λ1 ) < 0; and the dissipation must be sufficiently
stronger than both the nonlinear and the forcing terms, so that R < 1. Dissipation
typically leads to an exponentially decaying solution, but for the dependency on g and
zero.
However, this issue does not arise if the forcing term F0 resists the exponential
decay towards zero, instead causing the solution to decay towards some non-zero (possibly
time-dependent) state. The norm of the state that is exponentially approached can possibly
decay towards zero, but this decay itself must happen slower than exponentially for the
algorithm to be efficient.5
177
possible applications in partial differential equations.
include models of interacting populations of predators and prey [122], dynamics of chemical
reactions [123, 124], and the spread of an epidemic [125]. We now give an example of
the latter, based on the epidemiological model used in [126] to describe the early spread
The so-called SEIR model divides a population of P individuals into four components:
susceptible (PS ), exposed (PE ), infected (PI ), and recovered (PR ). We denote the rate
of transmission from an infected to a susceptible person by rtra , the typical time until
an exposed person becomes infectious by the latent time Tlat , and the typical time an
infectious person can infect others by the infectious time Tinf . Furthermore we assume that
there is a flux Λ of individuals constantly refreshing the population. This flux corresponds
to susceptible individuals moving into, and individuals of all components moving out of,
the population, in such a way that the total population remains constant.
To ensure that there is sufficiently strong linear decay to guarantee Carleman convergence,
that of [127], but denote the vaccination rate, which is approximately equal to the fraction
178
of susceptible individuals vaccinated each day, by rvac . The model is then
dPS PS PI
= −Λ − rvac PS + Λ − rtra PS (4.185)
dt P P
dPE PE PE PI
= −Λ − + rtra PS (4.186)
dt P Tlat P
dPI PI PE PI
= −Λ + − (4.187)
dt P Tlat Tinf
dPR PR PI
= −Λ + rvac PS + . (4.188)
dt P Tinf
constant. Hence we do not need to include the equation for PR in our analysis, which
is crucial since the PR component would have introduced positive eigenvalues. The
Λ
Λ − P − rvac 0 0
0 ,
F0 = F1 = 0 − PΛ − 1
0 , (4.189)
Tlat
1
0 0 Tlat
− PΛ − 1
Tinf
r
0 0 − Ptra 0 0 0 0 0 0
F2 = 0 0 rtra 0 0 0 0 0 0
. (4.190)
P
0 0 0 0 0 0 0 0 0
179
√
∥uin ∥ ≤ P , ∥F0 ∥ = Λ, and ∥F2 ∥ = 2rtra /P , so
√ √
2rtra + 3Λ/P
R≤ . (4.191)
min {rvac , 1/Tlat , 1/Tinf } + Λ/P
√
We see that the condition for guaranteed convergence of Carleman linearization is 2rtra <
√
min {rvac , 1/Tlat , 1/Tinf } − ( 3 − 1). Essentially, the Carleman method only converges if
To assess how restrictive this assumption is, we consider the SEIR parameters
used in [126]. Note that they also included separate components for asymptomatic and
hospitalized persons, but to simplify the analysis we include both of these components in
the PI component. In their work, they considered a city with approximately P = 107
per day, latent time Tlat = 5.2 days, infectious time Tinf = 2.3 days, and transmission
rate rtra ≈ 0.13 days−1 . We let the initial condition be dominated by the susceptible
component so that ∥uin ∥ ≈ P , and we assume7 that rvac > 1/Tlat ≈ 0.19 days−1 . With
the stated parameters, a direct calculation gives R = 0.956, showing that the assumptions
of our algorithm can correspond to some real-world problems that are only moderately
nonlinear.
While the example discussed above has only a constant number of variables, this
example can be generalized to a high-dimensional system of ODEs that models the early
6
Since we require that u(t) does not approach 0 exponentially, we can assume that the travel flux is
some non-zero, but small, value, e.g., Λ = 1 individual per day.
7
This example arguably corresponds to quite rapid vaccination, and is chosen here such that R remains
smaller than one, as required to formally guarantee convergence of the Carleman method. However,
as shown in the upcoming example of the Burgers equation, larger values of R might still allow for
convergence in practice, suggesting that our algorithm might handle lower values of the vaccination rate.
180
spread over a large number of cities with interaction, similar to what is done in [128] and
[129].
with the forcing term f being a function of both space and time. This equation can be cast
in the form of (4.1) by using standard discretizations of space and time. Equations of the
form (4.192) can represent Navier–Stokes-type equations, which are ubiquitous in fluid
mechanics [118], and related models such as those studied in [130, 131, 132] to describe
the formation of large-scale structure in the universe. Similar equations also appear in
models of magnetohydrodynamics (e.g., [133]), or the motion of free particles that stick
to each other upon collision [134]. In the inviscid case, ν = 0, the resulting Euler-type
equations with linear damping are also of interest, both for modeling micromechanical
devices [135] and for their intimate connection with viscous models [136].
often used as a simple model of convective flow [117]. For concreteness, let the initial
condition be u(x, 0) = U0 sin(2πx/L0 ) on the domain x ∈ [−L0 /2, L0 /2], and use
181
Figure 4.1: Integration of the forced viscous Burgers equation using Carleman linearization
on a classical computer (source code available at https://github.com/
hermankolden/CarlemanBurgers). The viscosity is set so that the Reynolds
number Re = U0 L0 /ν = 20. The parameters nx = 16 and nt = 4000 are the number
of spatial and temporal discretization intervals, respectively. The corresponding
Carleman convergence parameter is R = 43.59. Top: Initial condition and solution
L0
plotted at a third of the nonlinear time 31 Tnl = 3U 0
. Bottom: l2 norm of the absolute
error between the Carleman solutions at various truncation levels N (left), and the
convergence of the corresponding time-maximum error (right).
182
Dirichlet boundary conditions u(−L0 /2, 0) = u(L0 /2, 0) = 0. We force this equation
using a localized off-center Gaussian with a sinusoidal time dependence,8 given by f (x, t) =
2
U0 exp − (x−L 0 /4)
2(L0 /32)2
cos(2πt). To solve this equation using the Carleman method, we
discretize the spatial domain into nx points and use central differences for the derivatives
to get
ui+1 − 2ui + ui−1 u2i+1 − u2i−1
∂t ui = ν − + fi (4.194)
∆x2 4∆x
with ∆x = L0 /(nx − 1). This equation is of the form (4.1) and can thus generate the
Carleman system (4.6). The resulting linear ODE can then be integrated using the forward
Euler method, as shown in Figure 4.1. In this example, the viscosity ν is defined such that
were sufficient to resolve the solution. The figure compares the Carleman solution with
the solution obtained via direct integration of (4.194) with the forward Euler method (i.e.,
to equation (4.194) into the definition of R (4.2), we find that Re(λ1 ) is indeed negative
as required, given Dirichlet boundary conditions, but the parameters used in this example
result in R ≈ 44. Even though this does not satisfy the requirement R < 1 of the QCL
algorithm, we see from the absolute error plot in Figure 4.1 that the maximum absolute
error over time decreases exponentially as the truncation level N is incremented (in this
this suggests that in this example, the error of the classical Carleman method converges
183
4.7 Discussion
for a class of quadratic nonlinear differential equations. Compared to the previous approach
nearly quadratic dependence, under the condition R < 1 as defined in (4.2). Qualitatively,
this means that the system must be dissipative and that the nonlinear and inhomogeneous
effects must be small relative to the linear effects. We have also provided numerical results
suggesting the classical Carleman method may work on certain PDEs that do not strictly
satisfy the assumption R < 1. Furthermore, we established a lower bound showing that
√
for general quadratic differential equations with R ≥ 2, quantum algorithms must have
It is natural to ask whether the result of Theorem 4.1 can be achieved with a
classical algorithm, i.e., whether the assumption R < 1 makes differential equations
classically tractable. Clearly a naive integration of the truncated Carleman system (4.6) is
not efficient on a classical computer since the system size is Θ(nN ). But furthermore, it
is unlikely that any classical algorithm for this problem can run in time polylogarithmic
in n. If we consider Problem 4.1 with dissipation that is small compared to the total
evolution time, but let the nonlinearity and forcing be even smaller such that R < 1, then
in the asymptotic limit we have a linear differential equation with no dissipation. Hence
any classical algorithm that could solve Problem 4.1 could also solve non-dissipative
linear differential equations, which is a BQP-hard problem even when the dynamics are
184
unitary [137]. In other words, an efficient classical algorithm for this problem would
imply efficient classical algorithms for any problem that can be solved efficiently by a
do not know the complexity of the quantum quadratic ODE problem. We hope that future
work will close this gap and determine for which R the problem can be solved efficiently
The complexity of our algorithm depends on the parameter q defined in Theorem 4.1,
which characterizes the decay of the final solution relative to the initial condition. This
restricts the utility of our result, since we must have a suitable initial condition and
terminal time such that the final state is not exponentially smaller than the initial state.
renormalization of the state can be used to implement postselection, which would imply
the unlikely consequence BQP = PP (see Section 8 of [53] for further discussion).
However, for inhomogeneous equations the final state need not be exponentially smaller
185
than the initial state even in a long-time simulation, suggesting that our algorithm could
the accuracy of the result. For instance, instead of using just tensor powers of u as
auxiliary variables, one could use other nonlinear functions. Several previous papers on
[73, 138]. They also discuss approximating higher-order terms with lower-order ones in
(4.6) instead of simply dropping them, possibly improving accuracy. Such changes would
however alter the structure of the resulting linear ODE, which could affect the quantum
implementation.
The quantum part of the algorithm might also be improved. In this chapter we limit
ourselves to the first-order Euler method to discretize the linearized ODEs in time. This
is crucial for the analysis in Lemma 4.3, which states the global error increases at most
linearly with T . To establish this result for the Euler method, it suffices to choose the
time step (4.65) to ensure ∥I + Ah∥ ≤ 1, and then estimate the growth of global error
by (4.92). However, it is unclear how to give a similar bound for higher-order numerical
schemes. If this obstacle could be overcome, the error dependence of the complexity
might be improved.
It is also natural to ask whether our approach can be improved by taking features
of particular systems into account. Since the Carleman method has only received limited
attention and has generally been used for purposes other than numerical integration, it
seems likely that such improvements are possible. In fact, the numerical results discussed
in Section 4.6 (see in particular Figure 4.1) suggest that the condition R < 1 is not a
186
strict requirement for the viscous Burgers equation, since we observe convergence even
though R ≈ 44. This suggests that some property of equation (4.193) makes it more
amenable to Carleman linearization than our current analysis predicts. We leave a detailed
A related question is whether our algorithm can efficiently simulate systems exhibiting
dynamical chaos. The condition R < 1 might preclude chaos, but we do not have a proof
of this. More generally, the presence or absence of chaos might provide a more fine-
a state vector that encodes the solution without specifying how information is to be
extracted from that state. Simply producing a state vector is not enough for an end-to-end
application since the full quantum state cannot be read out efficiently. In some cases in
other cases, more sophisticated postprocessing may be required to infer a desired property
of the solution. Our method does not address this issue, but can be considered as a
subroutine whose output will be parsed by subsequent quantum algorithms. We hope that
future work will address this issue and develop end-to-end applications of these methods.
Finally, the algorithm presented in this chapter might be extended to solve related
problems with time-dependent coefficients and boundary value problems. Carleman methods
for such problems are explored in [83], but it is not obvious how to implement those
187
Chapter 5: Conclusion and future work
previous work [52], our algorithm improves the complexity of solving time-dependent
for Poisson’s equation, and generalized the quantum spectral method for general second-
order elliptic equations. Whereas previous algorithms scaled as poly(d, 1/ϵ), our algorithms
In Chapter 4, we have presented the first efficient quantum algorithm for a class of
time over previous quantum algorithms. The key ingredient we adopted and modified is
the Carleman linearization, which provides a long-time stable linear approximation to the
188
nonlinear system. We developed quantum Carleman linearization (QCL) for dissipative
nonlinear differential equations provided the condition R < 1. Here the quantity R is used
to quantify the ℓ2 relative strength of the nonlinearity and forcing to the linear dissipation.
The complexity of this approach is e(T 2 q/ϵ), where T is the evolution time, ϵ is the
allowed error, and q measures the decay of the solution. Moreover, We also provided
a quantum lower bound for the worst-case exponential time complexity of simulating
√
general nonlinear dynamics given R ≥ 2. Beyond the worst-case complexity analysis,
the numerical experiments for the Burgers equation revealed that Carleman linearization
Our work has raised enthusiasm from the communities in quantum physics, computer
science, applied mathematics, fluid and plasma dynamics, and other fields. Detailed open
questions and future directions followed by each topic can be found in the discussion
From the aspect of theories, can we offer exponential quantum speedups for more
general differential equations with provable guarantees? It is possible that the assumptions
on the models can be relaxed. For linear ODEs and PDEs, we wonder if we can develop
relax the elliptic condition for second-order linear PDEs. And for nonlinear differential
equations, we wonder whether the weak nonlinearity condition R < 1 can be relaxed.
Numerical results in Chapter 4 suggest that R < 1 is not a strict requirement for the
viscous Burgers equation, since we observe convergence even though R ≈ 44. This
suggests that some properties of nonlinear systems make them more amenable to Carleman
189
linearization than our current analysis predicts. We also wonder whether the complexity
of quantum algorithms for quadratic ODEs e(T 2 q/ϵ) can be improved. It is implausible to
significantly reduce the dependence on q, since renormalization of the state can implement
postselection, which could imply the unlikely consequence BQP = PP. The error
dependence might be improved if the quantum adaptive FDM or the quantum spectral
method could be applied to the nonlinear case, as a rigorous bound on the condition
number of the resulting quantum linear system is required. It is unclear whether a linear
and boundary value problems of linear and nonlinear differential equations. Despite
that previous studies focus on the performance in terms of the worst-case error, we
may find further improvements when considering the average-case error with random
initial inputs of typical differential equations. Finally, it is also likely that the techniques
presented in this dissertation could inspire new quantum algorithms for related problems
for differential equations and related real-world problems? For the preprocessing procedure,
we are concerned with how to efficiently prepare the initial states by loading classical
data. The ability to prepare such quantum states requires the implementation of quantum
random access memory (qRAM), which is widely thought to be impractical. But for
generate the initial states by a certain efficient routine instead of involving qRAM. For
190
observables of practical interest. Typical instances include the scattering cross section
from a wave propagation, and the free energy from a physical or biological system. Our
algorithms only produce a state vector that encodes the solution, which is not enough for
an end-to-end application since the full quantum state cannot be read out efficiently. In
to infer the desired property of the solution. For producing such classical readouts by
quantum computers, we can investigate the complexity and figure out whether an exponential
quantum speedup is possible. In addition, we are interested in the smallest system that
great interest to estimate the implementation cost (e.g. the one- or two-qubit gate counts)
for early fault-tolerant quantum computers. We hope that future work will address these
191
Bibliography
[1] Michael A. Nielsen and Isaac Chuang. Quantum computation and quantum
information, 2002.
[2] John Watrous. The theory of quantum information. Cambridge university press,
2018.
[4] Seth Lloyd. Universal quantum simulators. Science, pages 1073–1078, 1996.
[5] Ivan Kassal, Stephen P. Jordan, Peter J. Love, Masoud Mohseni, and Alán
Aspuru-Guzik. Polynomial-time quantum algorithm for the simulation of chemical
dynamics. Proceedings of the National Academy of Sciences, 105(48):18681–
18686, 2008. arXiv:0801.2986.
[6] Ian D. Kivlichan, Nathan Wiebe, Ryan Babbush, and Alán Aspuru-Guzik.
Bounding the costs of quantum simulation of many-body physics in real space.
Journal of Physics A: Mathematical and Theoretical, 50(30):305301, 2017.
[7] Borzu Toloui and Peter J. Love. Quantum algorithms for quantum chemistry based
on the sparsity of the CI-matrix, 2013. arXiv:1312.2579.
[8] Ryan Babbush, Dominic W. Berry, Yuval R. Sanders, Ian D. Kivlichan, Artur
Scherer, Annie Y. Wei, Peter J. Love, and Alán Aspuru-Guzik. Exponentially
more precise quantum simulation of fermions in the configuration interaction
representation. Quantum Science and Technology, 3(1):015006, 2017.
[9] Ryan Babbush, Dominic W. Berry, Jarrod R. McClean, and Hartmut Neven.
Quantum simulation of chemistry with sublinear scaling in basis size. Npj
Quantum Information, 5(1):1–7, 2019.
[10] Yuan Su, Dominic W Berry, Nathan Wiebe, Nicholas Rubin, and Ryan Babbush.
Fault-tolerant quantum simulations of chemistry in first quantization. PRX
Quantum, 2(4):040332, 2021. arXiv:2105.12767.
192
[11] Alán Aspuru-Guzik, Anthony D. Dutoi, Peter J. Love, and Martin Head-Gordon.
Simulated quantum computation of molecular energies. Science, 309(5741):1704–
1707, 2005. arXiv:quant-ph/0604193.
[13] Jacob T. Seeley, Martin J. Richard, and Peter J. Love. The Bravyi-Kitaev
transformation for quantum computation of electronic structure. The Journal of
Chemical Physics, 137(22):224109, 2012. arXiv:1208.5986.
[14] Dave Wecker, Bela Bauer, Bryan K. Clark, Matthew B. Hastings, and Matthias
Troyer. Gate-count estimates for performing quantum chemistry on small quantum
computers. Physical Review A, 90(2):022305, 2014. arXiv:1312.1695.
[15] Matthew B. Hastings, Dave Wecker, Bela Bauer, and Matthias Troyer.
Improving quantum algorithms for quantum chemistry. Quantum Information &
Computation, 15(1-2):1–21, 2015. arXiv:1403.1539.
[16] David Poulin, Matthew B. Hastings, David Wecker, Nathan Wiebe, Andrew C.
Doberty, and Matthias Troyer. The Trotter step size required for accurate quantum
simulation of quantum chemistry. Quantum Information & Computation, 15(5-
6):361–384, 2015. arXiv:1406.4920.
[17] Jarrod R. McClean, Ryan Babbush, Peter J. Love, and Alán Aspuru-Guzik.
Exploiting locality in quantum computation for quantum chemistry. The Journal
of Physical Chemistry Letters, 5(24):4368–4380, 2014.
[18] Ryan Babbush, Jarrod McClean, Dave Wecker, Alán Aspuru-Guzik, and Nathan
Wiebe. Chemical basis of Trotter-Suzuki errors in quantum chemistry simulation.
Physical Review A, 91(2):022311, 2015. arXiv:1410.8159.
[19] Markus Reiher, Nathan Wiebe, Krysta M. Svore, Dave Wecker, and Matthias
Troyer. Elucidating reaction mechanisms on quantum computers. Proceedings of
the National Academy of Sciences, 114(29):7555–7560, 2017. arXiv:1605.03590.
[20] Ryan Babbush, Dominic W. Berry, Ian D. Kivlichan, Annie Y. Wei, Peter J.
Love, and Alán Aspuru-Guzik. Exponentially more precise quantum simulation
of fermions in second quantization. New Journal of Physics, 18(3):033032, 2016.
[21] Mario Motta, Erika Ye, Jarrod R. McClean, Zhendong Li, Austin J. Minnich,
Ryan Babbush, and Garnet Kin-Lic Chan. Low rank representations for quantum
simulation of electronic structure. npj Quantum Information, 7(1):1–7, 2021.
[22] Earl Campbell. Random compiler for fast Hamiltonian simulation. Physical
Review Letters, 123(7):070503, 2019. arXiv:1811.08017.
193
[23] Dominic W. Berry, Craig Gidney, Mario Motta, Jarrod R. McClean, and Ryan
Babbush. Qubitization of arbitrary basis quantum chemistry leveraging sparsity
and low rank factorization. Quantum, 3:208, 2019. arXiv:1902.02134.
[24] Andrew M. Childs, Yuan Su, Minh C. Tran, Nathan Wiebe, and Shuchen
Zhu. Theory of Trotter error with commutator scaling. Physical Review X,
11(1):011020, 2021.
[25] Vera von Burg, Guang Hao Low, Thomas Häner, Damian S. Steiger, Markus
Reiher, Martin Roetteler, and Matthias Troyer. Quantum computing enhanced
computational catalysis. Physical Review Research, 3(3):033055, 2021.
[26] Joonho Lee, Dominic Berry, Craig Gidney, William J. Huggins, Jarrod R.
McClean, Nathan Wiebe, and Ryan Babbush. Even more efficient quantum
computations of chemistry through tensor hypercontraction. PRX Quantum,
2(3):030305, 2021. arXiv:2011.03494.
[27] Yuan Su, Hsin-Yuan Huang, and Earl T. Campbell. Nearly tight Trotterization of
interacting electrons. Quantum, 5:495, 2021. arXiv:2012.09194.
[28] Stephen P. Jordan, Keith S.M. Lee, and John Preskill. Quantum algorithms for
quantum field theories. Science, 336(6085):1130–1133, 2012.
[29] John Preskill. Simulating quantum field theory with a quantum computer. In The
36th Annual International Symposium on Lattice Field Theory, volume 334, page
024. SISSA Medialab, 2019. arXiv:1811.10085.
[30] Bela Bauer, Sergey Bravyi, Mario Motta, and Garnet Kin-Lic Chan. Quantum
algorithms for quantum chemistry and quantum materials science. Chemical
Reviews, 120(22):12685–12717, 2020. arXiv:2001.03685.
[31] Yudong Cao, Jonathan Romero, Jonathan P. Olson, Matthias Degroote, Peter D.
Johnson, Mária Kieferová, Ian D. Kivlichan, Tim Menke, Borja Peropadre, Nicolas
P. D. Sawaya, et al. Quantum chemistry in the age of quantum computing.
Chemical Reviews, 119(19):10856–10915, 2019. arXiv:1812.09976.
[32] Ryan Babbush, Nathan Wiebe, Jarrod McClean, James McClain, Hartmut Neven,
and Garnet Kin-Lic Chan. Low-depth quantum simulation of materials. Physical
Review X, 8(1):011044, 2018.
[33] Dominic W. Berry, Graeme Ahokas, Richard Cleve, and Barry C. Sanders.
Efficient quantum algorithms for simulating sparse Hamiltonians. Communications
in Mathematical Physics, 270(2):359–371, 2007. arXiv:quant-ph/0508139.
[34] David Poulin, Angie Qarry, Rolando D. Somma, and Frank Verstraete. Quantum
simulation of time-dependent Hamiltonians and the convenient illusion of Hilbert
space. Physical Review Letters, 106(17):170501, 2011. arXiv:1102.1360.
194
[35] Dominic W. Berry, Andrew M. Childs, Richard Cleve, Robin Kothari, and
Rolando D. Somma. Exponential improvement in precision for simulating sparse
Hamiltonians. Forum of Mathematics, Sigma, 5:e8, 2017. arXiv:1312.1414.
[36] Dominic W. Berry, Andrew M. Childs, Richard Cleve, Robin Kothari, and
Rolando D. Somma. Simulating Hamiltonian dynamics with a truncated taylor
series. Physical Review Letters, 114(9):090502, 2015. arXiv:1412.4687.
[37] Dominic W. Berry, Andrew M. Childs, and Robin Kothari. Hamiltonian simulation
with nearly optimal dependence on all parameters. In IEEE 56th Annual
Symposium on Foundations of Computer Science, pages 792–809. IEEE, 2015.
arXiv:1501.01715.
[38] Gilles Brassard, Peter Hoyer, Michele Mosca, and Alain Tapp. Quantum amplitude
amplification and estimation. Contemporary Mathematics, 305:53–74, 2002.
arXiv:quant-ph/0005055.
[39] Dominic W. Berry and Leonardo Novo. Corrected quantum walk for optimal
hamiltonian simulation. arXiv:1606.03443, 2016.
[40] Leonardo Novo and Dominic W Berry. Improved hamiltonian simulation via a
truncated taylor series and corrections. arXiv:1611.10033, 2016.
[41] Guang Hao Low and Isaac L Chuang. Hamiltonian simulation by qubitization.
Quantum, 3:163, 2019. arXiv:1610.06546.
[42] Guang Hao Low and Isaac L. Chuang. Optimal hamiltonian simulation by
quantum signal processing. Physical Review Letters, 118(1):010501, 2017.
arXiv:1606.02685.
[43] Aram W. Harrow, Avinatan Hassidim, and Seth Lloyd. Quantum algorithm for
linear systems of equations. Physical Review Letters, 103(15):150502, 2009.
arXiv:0811.3171.
[44] Andris Ambainis. Variable time amplitude amplification and quantum algorithms
for linear algebra problems. In 29th Symposium on Theoretical Aspects of
Computer Science, volume 14, pages 636–647. LIPIcs, 2012. arXiv:1010.4458.
[45] Dong An and Lin Lin. Quantum linear system solver based on time-optimal
adiabatic quantum computing and quantum approximate optimization algorithm,
2019. arXiv:1909.05500.
195
[47] András Gilyén, Yuan Su, Guang Hao Low, and Nathan Wiebe. Quantum singular
value transformation and beyond: exponential improvements for quantum matrix
arithmetics. In Proceedings of the 51st Annual ACM SIGACT Symposium on
Theory of Computing, pages 193–204, 2019. arXiv:1806.01838.
[48] Lin Lin and Yu Tong. Optimal quantum eigenstate filtering with application to
solving quantum linear systems. Quantum, 4:361, 2020. arXiv:1910.14596.
[49] Yiğit Subaşı, Rolando D. Somma, and Davide Orsucci. Quantum algorithms for
systems of linear equations inspired by adiabatic quantum computing. Physical
Review Letters, 122(6):060504, 2019. arXiv:1805.10549.
[50] Yu Tong, Dong An, Nathan Wiebe, and Lin Lin. Fast inversion, preconditioned
quantum linear system solvers, and fast evaluation of matrix functions, 2020.
arXiv:2008.13295.
[51] Pedro Costa, Dong An, Yuval R. Sanders, Yuan Su, Ryan Babbush, and
Dominic W. Berry. Optimal scaling quantum linear systems solver via discrete
adiabatic theorem, 2021. arXiv:2111.08152.
[52] Dominic W. Berry. High-order quantum algorithm for solving linear differential
equations. Journal of Physics A: Mathematical and Theoretical, 47(10):105301,
2014. arXiv:1010.2745.
[54] Andrew M. Childs and Jin-Peng Liu. Quantum spectral methods for differential
equations. arXiv:1901.00961, 2019.
[56] Yudong Cao, Anargyros Papageorgiou, Iasonas Petras, Joseph Traub, and Sabre
Kais. Quantum algorithm and circuit design solving the Poisson equation. New
Journal of Physics, 15(1):013021, 2013. arXiv:1207.2485.
[57] Ashley Montanaro and Sam Pallister. Quantum algorithms and the finite element
method. Physical Review A, 93(3):032324, 2016. arXiv:1512.05903.
[58] Pedro C. S. Costa, Stephen Jordan, and Aaron Ostrander. Quantum algorithm
for simulating the wave equation. Physical Review A, 99(1):012323, 2019.
arXiv:1711.05394.
[59] Andrew M. Childs, Jin-Peng Liu, and Aaron Ostrander. High-precision quantum
algorithms for partial differential equations, 2020. arXiv:2002.07868.
196
[60] Alexander Engel, Graeme Smith, and Scott E. Parker. Quantum algorithm for the
Vlasov equation. Physical Review A, 100(6):062315, 2019. arXiv:1907.09418.
[61] Noah Linden, Ashley Montanaro, and Changpeng Shao. Quantum vs. classical
algorithms for solving the heat equation. arXiv:2004.06516.
[62] Alexei Y. Kitaev. Quantum measurements and the abelian stabilizer problem, 1995.
arXiv:quant-ph/9511026.
[63] David Kahaner, Cleve Moler, and Stephen Nash. Numerical methods and software.
Prentice-Hall, Inc., 1989.
[65] Esmail Babolian and Mohammad Mahdi Hosseini. A modified spectral method for
numerical solution of ordinary differential equations with non-analytic solution.
Applied Mathematics and Computation, 132(2-3):341–351, 2002.
[67] Călin Ioan Gheorghiu. Spectral methods for differential problems. Casa Cărţii de
Ştiinţă Cluj-Napoca, 2007.
[68] Jie Shen, Tao Tang, and Li-Lian Wang. Spectral methods: algorithms, analysis
and applications, volume 41. Springer Science & Business Media, 2011.
[69] Yudong Cao, Anargyros Papageorgiou, Iasonas Petras, Joseph Traub, and Sabre
Kais. Quantum algorithm and circuit design solving the Poisson equation. New
Journal of Physics, 15(1):013021, 2013. arXiv:1207.2485.
[70] Pedro C. S. Costa, Stephen Jordan, and Aaron Ostrander. Quantum algorithm
for simulating the wave equation. Physical Review A, 99(1):012323, 2019.
arXiv:1711.05394.
[71] Ivo Babuška and Manil Suri. The h − p version of the finite element method
with quasiuniform meshes. Mathematical Modelling and Numerical Analysis,
21(2):199–238, 1987.
[72] Edward H. Kerner. Universal formats for nonlinear ordinary differential systems.
Journal of Mathematical Physics, 22(7):1366–1371, 1981.
[73] Marcelo Forets and Amaury Pouly. Explicit error bounds for Carleman
linearization, 2017. arXiv:1711.02552.
[74] Sarah K. Leyton and Tobias J. Osborne. A quantum algorithm to solve nonlinear
differential equations. arXiv:0812.4423, 2008.
197
[75] Ilon Joseph. Koopman-von Neumann approach to quantum simulation of nonlinear
classical dynamics, 2020. arXiv:2003.09980.
[77] Seth Lloyd, Giacomo De Palma, Can Gokler, Bobak Kiani, Zi-Wen Liu, Milad
Marvian, Felix Tennie, and Tim Palmer. Quantum algorithm for nonlinear
differential equations, 2020. arXiv:2011.06571.
[78] Andrew M. Childs and Joshua Young. Optimal state discrimination and
unstructured search in nonlinear quantum mechanics. Physical Review A,
93(2):022314, 2016. arXiv:1507.06334.
[79] Daniel S. Abrams and Seth Lloyd. Nonlinear quantum mechanics implies
polynomial-time solution for NP-complete and #P problems. Physical Review
Letters, 81(18):3992, 1998. arXiv:quant-ph/9801041.
[80] Scott Aaronson. NP-complete problems and physical reality. ACM SIGACT News,
36(1):30–52, 2005. arXiv:quant-ph/0502072.
[81] Jin-Peng Liu, Herman Øie Kolden, Hari K. Krovi, Nuno F. Loureiro, Konstantina
Trivisa, and Andrew M. Childs. Efficient quantum algorithm for dissipative
nonlinear differential equations, 2020. arXiv:2011.03185.
[82] Torsten Carleman. Application de la théorie des équations intégrales linéaires aux
systèmes d’équations différentielles non linéaires. Acta Mathematica, 59(1):63–
87, 1932.
[83] Krzysztof Kowalski and Willi-Hans Steeb. Nonlinear Dynamical Systems and
Carleman Linearization. World Scientific, 1991.
[84] Aram W. Harrow, Avinatan Hassidim, and Seth Lloyd. Quantum algorithm for
linear systems of equations. Physical Review Letters, 103(15):150502, 2009.
arXiv:0811.3171.
[85] Tom M. Apostol. Calculus, volume ii: multi-variable calculus and linear algebra,
with applications to differential equations and probability, 1969.
[86] Sangtae Kim and Seppo J. Karrila. Microhydrodynamics: principles and selected
applications. Courier Corporation, 2013.
[88] Stewart Harris. An introduction to the theory of the Boltzmann equation. Courier
Corporation, 2004.
198
[90] Bo Thidé. Electromagnetic field theory. Upsilon books Uppsala, 2004.
[91] R. Byron Bird. Transport phenomena. Appl. Mech. Rev., 55(1):R1–R4, 2002.
[92] Vivek V. Shende, Stephen S. Bullock, and Igor L. Markov. Synthesis of quantum-
logic circuits. IEEE Transactions on Computer-Aided Design of Integrated Circuits
and Systems, 25(6):1000–1010, 2006. arXiv:quant-ph/0406176.
[95] Hans-Joachim Bungartz and Michael Griebel. Sparse grids. Acta Numerica,
13:147–269, 2004.
[97] Jie Shen and Haijun Yu. Efficient spectral sparse grid methods and applications
to high-dimensional elliptic problems. SIAM Journal on Scientific Computing,
32(6):3228–3250, 2010.
[98] Jie Shen and Haijun Yu. Efficient spectral sparse grid methods and applications
to high-dimensional elliptic equations II. Unbounded domains. SIAM Journal on
Scientific Computing, 34(2):A1141–A1164, 2012.
[100] Jianping Li. General explicit difference formulas for numerical differentiation.
Journal of Computational and Applied Mathematics, 183(1):29–52, 2005.
[101] Roger A. Horn and Charles R. Johnson. Matrix analysis. Cambridge University
Press, 2012.
[102] Norbert Schuch and Jens Siewert. Programmable networks for quantum
algorithms. Physical Review Letters, 91(2):027902, 2003. arXiv:quant-
ph/0303063.
[103] Jonathan Welch, Daniel Greenbaum, Sarah Mostame, and Alán Aspuru-Guzik.
Efficient quantum circuits for diagonal unitaries without ancillas. New Journal
of Physics, 16(3):033040, 2014. arXiv:1306.3991.
[104] Dominic W. Berry, Andrew M. Childs, Richard Cleve, Robin Kothari, and
Rolando D. Somma. Simulating Hamiltonian dynamics with a truncated Taylor
series. Physical Review Letters, 114(9):090502, 2015. arXiv:1412.4687.
199
[105] Daniel T. Colbert and William H. Miller. A novel discrete variable representation
for quantum mechanical reactive scattering via the S-matrix Kohn method. The
Journal of Chemical Physics, 96(3):1982–1991, 1992.
[106] Daniel Spielman. Rings, paths, and Cayley graphs (course notes), 2014. http:
//www.cs.yale.edu/homes/spielman/561/lect05-15.pdf.
[107] Tao Tang. Spectral and high-order methods with applications. Science Press
Beijing, 2006.
[108] Andreas Klappenecker and Martin Rotteler. Discrete cosine transforms on quantum
computers. In Proceedings of the 2nd International Symposium on Image and
Signal Processing and Analysis, pages 464–468, 2001. arXiv:quant-ph/0111038.
[109] Martin Rötteler, Markus Püschel, and Thomas Beth. Fast signal transforms for
quantum computers. In Proceedings of the Workshop on Physics and Computer
Science, pages 31–43, 1999.
[110] Markus Püschel, Martin Rötteler, and Thomas Beth. Fast quantum Fourier
transforms for a class of non-abelian groups. In International Symposium on
Applied Algebra, Algebraic Algorithms, and Error-Correcting Codes, pages 148–
159, 1999. arXiv:quant-ph/9807064.
[113] Gerasimos Lyberatos and Christos A. Tsiligiannis. A linear algebraic method for
analysing Hopf bifurcation of chemical reaction systems. Chemical Engineering
Science, 42(5):1242–1244, 1987.
[114] Roger Brockett. The early days of geometric nonlinear control. Automatica,
50(9):2203–2224, 2014.
[115] Andreas Rauh, Johanna Minisini, and Harald Aschemann. Carleman linearization
for control and for state and disturbance estimation of nonlinear dynamical
processes. IFAC Proceedings Volumes, 42(13):455–460, 2009.
[116] Alfredo Germani, Costanzo Manes, and Pasquale Palumbo. Filtering of differential
nonlinear systems via a Carleman approximation approach. In Proceedings of the
44th IEEE Conference on Decision and Control, pages 5917–5922, 2005.
200
[118] Pierre Gilles Lemarié-Rieusset. The Navier-Stokes problem in the 21st century.
CRC Press, 2018.
[119] Germund G. Dahlquist. A special stability problem for linear multistep methods.
BIT Numerical Mathematics, 3(1):27–43, 1963.
[120] Kendall E. Atkinson. An introduction to numerical analysis. John Wiley & Sons,
2008.
[121] Carl W. Helstrom. Quantum detection and estimation theory. Journal of Statistical
Physics, 1:231–252, 1969.
[122] Paul Waltman. Competition models in population biology. SIAM, 1983.
[123] Elliott W. Montroll. On coupled rate equations with quadratic nonlinearities.
Proceedings of the National Academy of Sciences, 69(9):2532–2536, 1972.
[124] Alessandro Ceccato, Paolo Nicolini, and Diego Frezzato. Recasting the mass-
action rate equations of open chemical reaction networks into a universal quadratic
format. Journal of Mathematical Chemistry, 57:1001–1018, 2019.
[125] Fred Brauer and Carlos Castillo-Chavez. Mathematical models in population
biology and epidemiology, volume 2. Springer, 2012.
[126] Chaolong Wang, Li Liu, Xingjie Hao, Huan Guo, Qi Wang, Jiao Huang, Na He,
Hongjie Yu, Xihong Lin, An Pan, Sheng Wei, and Tangchun Wu. Evolving
epidemiology and impact of non-pharmaceutical interventions on the outbreak of
coronavirus disease 2019 in Wuhan, China. Journal of the American Medical
Association, 323(19):1915–1923, 2020.
[127] Gul Zaman, Yong Han Kang, and Il Hyo Jung. Stability analysis and optimal
vaccination of an SIR epidemic model. Biosystems, 93(3):240–249, 2008.
[128] Derdei Bichara, Yun Kang, Carlos Castillo-Chávez, Richard Horan, and Charles
Perrings. SIS and SIR epidemic models under virtual dispersal, 03 2015.
[129] Rany Qurratu Aini, Deden Aldila, and Kiki Sugeng. Basic reproduction number of
a multi-patch SVI model represented as a star graph topology, 10 2018.
[130] Lev Kofman, Dmitri Pogosian, and Sergei Shandarin. Structure of the universe
in the two-dimensional model of adhesion. Monthly Notices of the Royal
Astronomical Society, 242(2):200–208, 1990.
[131] Sergei F. Shandarin and Yakov B. Zeldovich. The large-scale structure of
the universe: turbulence, intermittency, structures in a self-gravitating medium.
Reviews of Modern Physics, 61(2):185–220, 1989.
[132] Massimo Vergassola, Bérengère Dubrulle, Uriel Frisch, and Alain Noullez.
Burgers’ equation, devil’s staircases and the mass distribution for large-scale
structures. Astronomy and Astrophysics, 289:325–356, 1994.
201
[133] P. A. Davidson. An Introduction to Magnetohydrodynamics. Cambridge Texts in
Applied Mathematics. Cambridge University Press, 2001.
[134] Yann Brenier and Emmanuel Grenier. Sticky particles and scalar conservation
laws. SIAM Journal on Numerical Analysis, 35(6):2317–2328, 1998.
[138] Richard Bellman and John M. Richardson. On some questions arising in the
approximate solution of nonlinear differential equations. Quarterly of Applied
Mathematics, 20:333–339, 1963.
202