Schrödingerisation based computationally stable algorithms for ill-posed problems in partial differential equations
Abstract
We introduce a simple and stable computational method for ill-posed partial differential equation (PDE) problems. The method is based on Schrödingerisation, introduced in [S. Jin, N. Liu and Y. Yu, arXiv:2212.13969][S. Jin, N. Liu and Y. Yu, Phys. Rev. A, 108 (2023), 032603], which maps all linear PDEs into Schrödinger-type equations in one higher dimension, for quantum simulations of these PDEs. Although the original problem is ill-posed, the Schrödingerised equations are Hamiltonian systems and time-reversible, allowing stable computation both forward and backward in time. The original variable can be recovered by data from suitably chosen domain in the extended dimension. We will use the (constant and variable coefficient) backward heat equation and the linear convection equation with imaginary wave speed as examples. Error analysis of these algorithms are conducted and verified numerically. The methods are applicable to both classical and quantum computers, and we also lay out quantum algorithms for these methods. Moreover, we introduce a smooth initialisation for the Schrödingerised equation which will lead to essentially spectral accuracy for the approximation in the extended space, if a spectral method is used. Consequently, the extra qubits needed due to the extra dimension, if a qubit based quantum algorithm is used, for both well-posed and ill-posed problems, becomes almost where is the desired precision. This optimizes the complexity of the Schrödingerisation based quantum algorithms for any non-unitary dynamical system introduced in [S. Jin, N. Liu and Y. Yu, arXiv:2212.13969][S. Jin, N. Liu and Y. Yu, Phys. Rev. A, 108 (2023), 032603].
Keywords: ill-posed problems, backward heat equation, Schrödingerisation, quantum algorithms MSCcodes: 65J20, 65M70, 81-08
1 Introduction
In this paper, we are interested in numerically computing ill-posed problems that follow the evolution of a general dynamical system
(1.1) |
whose input data given by . For simplicity, we first consider being Hermitian with real eigenvalues, ordered as
(1.2) |
We assume that , for some , which implies that the system (1.1) contains unstable modes, thus the initial value problem is ill-posed since its solution will grow exponentially in time. This causes significant computational challenges. Normally the numerical errors will grow exponentially unless special care is taken.
Ill-posed or unstable problems appear in many physical applications, for example, fluid dynamics instabilities such as Rayleigh-Taylor and Kevin-Helmholtz instabilities [21, 10], plasma instability [12], and Maxwell equations with negative index of refraction [30, 28]. It also appears in inverse problems [34, 4]. One of the most classical ill-posed problems is the backward heat equation which suffers from the catastrophic Hadamard instability. Usually, some regularization technique is used to make the problems well-posed [8].
In this paper, we study a generic yet simple computational strategy to numerically compute ill-posed problems based on Schrödingerisation, introduced for quantum simulation for dynamical systems whose evolution operator is non-unitary [16, 17]. The idea is to map it to one higher dimension, into Schrödinger-type equations, that obeys unitary dynamics and thereby naturally fitting for quantum simulation. Since the Schrödinger-type equations are Hamiltonian systems that are time-reversible, they can be solved both forwards and backwards in time in a computationally stable way. This makes it suitable for solving unstable problems, as was proposed in our previous work [15]. In this article, we study this method for the classical backward heat equation, and a linear convection equation with imaginary wave speed (or negative index of refraction).
The initial-value problem to the backward heat equation is ill-posed in all three ways: (i) the solution does not necessarily exist; (ii) if the solution exists, it is not necessarily unique; (iii) there is no continuous dependence of the solution on arbitrary input data [27, 18, 13, 23]. This problem is well-posed for final data whose Fourier spectrum has compact support [26]. However, even when the solution exists and is unique, computing the solution is difficult since unstable physical systems usually lead to unstable numerical methods. There have been various treatments of this difficult problem, for example quasi-reversibility methods [35, 32, 22, 3, 25], regularization methods [11, 19, 32, 6], Fourier truncation methods [29], etc. Our approach computes solution consistent with the Fourier truncation method.
The actual implementation of our backward heat equation solver is as follows. First we start with the Fourier transform of the input data denoted by and truncate the Fourier mode in finite domain. This is either achieved automatically if the Fourier mode of has compact support, or we choose a sufficiently large domain in the Fourier space such that outside it the Fourier coefficient of is sufficiently small. The latter can usually be done since the forward heat equation gives a solution that is smooth for so its Fourier coefficient decays rapidly. The Schrödingerisation technique lifts the backward heat equation to a Schrödinger equation in one higher dimension that is time reversible, which can be solved backward in time by any reasonable stable numerical approximation. We then recover for by integrating or pointwise evaluation over suitably chosen domain in the extended variable. Since the time evolution is based on solving the Schrödinger equation, the computational method is stable.
We point out that the truncation in the Fourier space regularizes the original ill-posed problem. Although the initial-value problem to the Schrödingerised equation is well-posed even without this Fourier truncation, to recover one needs finite Fourier mode. We also show that the strategy also applies to variable coefficient heat equation.
We will also apply the same strategy to solve the unstable linear convection equation that has imaginary wave speed.
We will give error estimates on these methods and conduct numerical methods that verify the results of the error analysis. The methods work for both classical and quantum computers. Since the original Schrödingerisation method was introduced for quantum simulation of general PDEs, we will also give the quantum implementation of the computational method.
Moreover, we introduce a smooth initialization for the Schrödingerised equation, so the initial data will be in in the extended space for any integer (see section 4.3). This will lead to essentially the spectral accuracy for the approximation in the extended space, if a spectral method is used. Consequently, the extra qubits needed due to the extra dimension, if a qubit based quantum algorithm is used, for both well-posed and ill-posed problems, becomes almost where is the desired precision. This reduces the complexity of Schrödingerisation based quantum algorithms introduced in [16, 17] for any non-unitary dynamical system.
The rest of the paper is organized as follows. In section 2, we give a brief review of the Schrödingerisation approach for general linear ODEs. In section 3, we study the backward heat equation, for both constant and variable coefficient cases, and show the approximate solution based on the framework of Schrödingerisation. In section 4, the numerical method and error analysis of the backward heat equation are presented. In addition, a smooth initial data with respect to the extended variable is constructed to improve the convergence rates in the extended space. In section 5, we apply this technique to the convection equations with purely imaginary wave speed. In section 6, we show the numerical tests to verify our theories. Section 7 shows the quantum algorithms and the corresponding complexity.
Throughout the paper, we restrict the simulation to a finite time interval . The notation stands for where is independent of the mesh size and time step. Scalar-valued quantities and vector-valued quantities are denoted by normal symbols and boldface symbols, respectively. Moreover, we use a 0-based indexing, i.e. , or , and , denotes a vector with the -th component being and others . We shall denote the identity matrix and null matrix by and 0, respectively, and the dimensions of these matrices should be clear from the context, otherwise, the notation stands for the -dimensional identity matrix.
2 The general framework
We start with the basic framework set up in [15] for forward problems which we first briefly review here.
Using the warped phase transformation for and symmetrically extending the initial data to , one has the following system of linear convection equations [16, 17]:
(2.1) |
This is clearly a hyperbolic system. When the eigenvalues of are all negative, the convection term of (2.1) corresponds to waves moving from the right to the left. One does, however, need a boundary condition on the right-hand side. Since decays exponentially in , one just needs to select large enough, so at this point is essentially zero and a zero incoming boundary condition can be used at . Then can be numerically recovered via
(2.2) |
or
(2.3) |
However, when for some , some of the waves evolving through (2.1) will instead propagate from left to right. If we were solving the problem in domain , we would need a boundary condition at , which is not possible. Instead, we extend the computational domain to also include with a zero incoming boundary at for sufficiently large (so ). In summary, we use zero boundary condition for (2.1) restricted on the finite domain . It is worth noting that extending to the domain is done for the convenience of employing the Fourier spectral method, which requires periodic boundary conditions, and our zero boundary conditions can be regarded, approximately, as periodic boundary condition. It is important to note that any waves–corresponding to positive eigenvalues of – that start from and travel to the right will induce spurious solutions to the domain , therefore, when recovering , one needs to select the correct domain to avoid using these spurious waves. Therefore, the following theorem [15] must be utilized.
Theorem 2.1.
Since (2.1) is a hyperbolic system, thus the initial value problem is well-posed and one can solve it numerically in a stable way (as long as suitable numerical stability condition–the CFL condition–is satisfied). Thus although the original system (1.1) is ill-posed, we can still solve the well-posed problem (2.1), and then recover using Theorem 2.1, as long as !
We also point out here that the discretization of the -derivative in Eq. (2.1) can be achieved not only by using Fourier spectral methods but also through other numerical schemes, such as finite difference or finite element methods. In such cases one can just solve the problem in and, if for some , one needs to supply boundary condition at for the characteristic fields corresponding to positive eigenvalues of . This boundary data can be set arbitrarily, since the induced spurious wave propagating with speed will not be used when we recover using data from , as laid out by Theorem 2.1.
3 The backward heat equation
In this section, we consider the following one-dimensional backward heat equation in the infinite domain,
(3.1) | ||||
where . We want to determine for from the data . The change of variable leads to the following formulation of (3.1),
(3.2) | ||||||
It is easy to see that the eigenvalues of in (1.1) are positive after spatial discretization. Let denote the Fourier transform of and define it by
Let denote the Sobolev norm defined as
When , denotes the -norm, and can be noninteger [9]. For the above solution to be well-defined, we make the following assumptions of for analysis:
-
(H1)
Assume , namely with compact support. This yields a well-posed problem in a subspace of in the sense of Hadamard [26].
-
(H2)
For more general , we will truncate the domain in and begin with the truncated–thus compactly supported–final data.
We remark here that truncation is inherently achieved in our methodology (see (S1) and (S2)), thus providing a form of regularization to the original ill-posed problem.
If the solution of (3.1) in this Sobolev space exists, then it must be unique [9]. We assume is the unique solution of (3.1). Applying the Fourier transform, one gets the exact solution of problem (3.1):
(3.3) |
Denoting , then it is easy to see from (3.3) that
(3.4) |
The Schrödingerisation for (3.1) gives
(3.5) | ||||
where . Again using the Fourier transform technique to (3.5) with respect to the variables and , one gets the Fourier transform of the exact solution of problem (3.5) satisfying
(3.6) |
The solution of is then found to be
(3.7) |
Note the initial-value problem (3.5) is well-posed, even for not being compactly supported in the Fourier space, as can be easily seen from (3.6), which is an oscillatory ODE with purely imaginary spectra. Following the proof in [9, Theorem 5, section 7.3], one has the regularity of under the assumption in (H1) or (H2).
Using the Fourier transform only on gives
(3.8) |
where is the Fourier transform of on the -variable, i.e.
(3.9) |
Here is a linear wave moving to the right, if one starts from and going backward in time (one can also use the change of variable to make the problem (3.5) forward in time for notation comfort), and the solution is computed by the inverse Fourier transformation
(3.10) |
This corresponds to the case of (2.1) in which all eigenvalues of are positive. Therefore, to recover from Eq.(3.10), one needs to choose for time to obtain
This is the basis for the introduction of our stable computational method: we solve the -equation (3.5) with a suitable–stable– computational method, and then recover by either integration or pointwise evaluation using Theorem 2.1.
However, in the continuous space maybe unbounded, thus the condition may not be satisfied. We consider two scenarios here:
-
(S1)
The input data has compact support, namely for for some . While this is not generally true, when is smooth in , its Fourier mode decays rapidly, so one can truncate for with desired accuracy.
-
(S2)
One discretizes equation (3.5) numerically. This usually requires first to truncate the -domain so it is finite, followed by some numerical discretization of the derivative, for example the finite difference or spectral method. Then where is the spatial mesh size in .
3.1 Error estimates
We start by choosing an and denote
(3.11) |
to be a small quantity, which will be chosen with other error terms to meet the numerical tolerance requirement (see Remark 3.2). We define the “approximate” solution of (3.1) as follows:
Definition 3.1.
First, define
(3.12) |
where when , otherwise . Then one obtains the approximate solution by
(3.13) |
with .
With defined as above, we have the following theorem which gives an error estimate to the above approximation.
Theorem 3.1.
Proof.
Remark 3.1.
Remark 3.2.
In order to bound the error by at , the shape estimate of from (3.15) is that is chosen such that
(3.16) |
If , one could choose satisfying
(3.17) |
Considering , one may have with , and then is small. Besides, if is smooth enough, is not necessarily large.
Specifically, if the Fourier transform of has compact support such that when , it yields from (3.15)-(3.16) that there exits no error of recovery by choosing . Take an example, (or ) is a periodic function in . We extend it periodically over the whole field of real domain . The Fourier transform of is , where gives an impulse that is shifted to the left by , likewise the function yields an impulse shifted to the right by . In this case, when . Therefore, the best choice of for (or ) is
(3.18) |
At this point, , one has , for .
In practice, the data at is usually imprecise since it depends on the reading of physical measurements. Consider a perturbed data , which is a small disturb of . The Fourier transform of may not decay as , which leads to the severely ill-posed problems of (3.1). However, after Schrödingerisation, a disturb of the Fourier transform of defined by does not affect the well-posedness of even if is merely in . Assume the measured data satisfies
(3.19) |
Define the approximate solution at time from by
(3.20) | ||||
Now we have the approximate solution .
Theorem 3.2.
3.2 Variable coefficient problems
When the system has a spatially varying coefficient, the Fourier transform cannot be applied, and the analysis in Theorem 3.1 no longer holds. However, the Schrödingerisation still works.
We consider variable coefficient equation
(3.23) |
with Dirichlet’s boundary condition of bounded domain in straightforward way, where and .
Let be the -th eigenfunction corresponding to of the operator such that
(3.24) |
It is easy to see that from the regularity of the second-order elliptic equations [9, Theorem 3, section 6.3]. They form an orthonormal basis set of . According to standard theory on the eigenvalues of symmetric elliptic operators [9, Theorem 1, section 6.5], one has
(3.25) |
We search for solution to (3.23) at time from the data in the form
(3.26) |
where .
Define the equivalent norm of , , again denoted by :
(3.27) |
where . Here denotes norm when , and can also be a noninteger for [20, Proposition 2.1]. For the solution to (3.23) to be well-defined, we assume the expansion of to the orthonormal basis has only finite number of terms, leading to for all , i.e.
(3.28) | ||||
For more general cases, we truncate to the finite terms of expansions of the input data. This regularizes the originally ill-posed problem. Next, assume is in the form of
We correspondingly project the input data on the same basis set and Eq. (3.5) is equivalent to
Note is well-defined even if the series is not truncated. We then define the truncated approximate solution by
(3.29) |
Let
(3.30) |
we choose such that
(3.31) |
for the desired precision . We remark here exists from (3.28) if .
Theorem 3.3.
4 Discretization of the Schrödingerised equation
In this section, we consider the discretization of (3.5) and the corresponding error estimates. For simplicity, we consider to be a periodic function defined in .
4.1 The numerical discretization
We discretize the domain uniformly with the mesh size , where is a positive even integer and the grid points are denoted by . The -D basis functions for the Fourier spectral method are usually chosen as
(4.1) |
Considering the Fourier spectral discretization on , one easily gets
(4.2) |
where and . The matrix is obtained by . By applying matrix exponentials, one has
(4.3) |
However, it is difficult to obtain the numerical solution in the classical computer from (4.3) since the problem is unstable.
We now introduce the discretization of domain as in [15]. First truncating the -region to , where is large enough such that
(4.4) |
with error bound. We point out that the constant is just for numerical analysis which is quite mild for numerical tests, and it is sufficient to take it as approximately .
Then, we can treat at the boundary as . Using the spectral method, one gets the transformation and difference matrix
where with . Applying the discrete Fourier spectral discretization on and , i.e. , it yields
(4.5) |
where . The matrices () or () can be implemented by (inverse) fast Fourier transforms (FFT). Since is a diagonal matrix, the solution at time is computed by
(4.6) |
This approach ensures that no error is introduced in the time discretization. The numerical solution to the system (3.1) is
(4.7) |
where from Theorem 3.1, or can be recovered by second order numerical integration
(4.8) |
with .
4.2 Error analysis for the spatial discretization
Following the general error estimates of Schrödingerisation in the extended domain in [15], we derive the specific estimates for the above approximation to the backward heat equation under the assumption that satisfies (H1) or (H2).
Define the complex and -dimensional space with respect to and , respectively
(4.9) |
The approximation of in the finite space from the numerical solution in Eq. (4.5) is
(4.10) |
where , . Correspondingly, the approximation of is defined by
(4.11) |
where .
It is apparent that the discretization serves as an approximation for the following system with periodic boundary conditions:
(4.12) |
where , and is periodic with a period of , such that
(4.13) |
Since is also periodic with respective to , the Fourier transform on variable of is defined by
(4.14) |
According to the fundamental results on regularity of transport equations, it is well known that the initial value problem (4.12) is well posed, and satisfies [9, Theorem 5, section 7.3]
(4.15) |
In addition, the standard estimate of spectral methods shows
(4.16) |
Lemma 4.1.
Suppose satisfies (3.17), , and satisfies (H1) or (H2). Let . It follows that
Proof.
It is easy to obtain from the system in Eq.(3.7) that
(4.17) |
Similarly, one has
(4.18) |
Using the semi norm in phase space, it yields
According to (4.17) and (4.18), changing the order of integration and letting , it yields
(4.19) |
where and are computed by
Since is periodic with a period of defined in (4.13), , and satisfies Eq. (3.17), one has
(4.20) | ||||
(4.21) |
The proof is completed by inserting (4.20) and (4.21) into (4.19). ∎
Lemma 4.2.
Proof.
Theorem 4.1.
Proof.
As can be seen from Theorem 3.1, the error increases as the computation time progresses from to , which is also reflected in the numerical calculations. The error we provided here corresponds to the error at the moment , which is the worst-case scenario. This aligns with the numerical test in section 6.2, where the error increases as the time approaches zero. Consequently, the error estimation in Theorem 4.1 can be considered as taking the norm in the time direction.
We point out that our algorithm offers three significant advantages. Firstly, the Hamiltonian nature of the semi-discrete system (4.5) allows for the development of a computationally stable scheme, provided the CFL condition is met. Secondly, Theorem 3.1 ensures a straightforward recovery of the original variable from . Although a time-dependent exponential growth factor is introduced at this step, we highlight that the value of is relatively small, as noted in Remark 3.2. Furthermore, by choosing an appropriate interval in the extended space to recover the original variables, essentially truncating the Fourier modes of the input data, the originally ill-posed problem becomes well-posed. Although the initial-value problem to the Schrödingerised equation is well-posed even without this Fourier truncation, to recover one needs finite Fourier mode. Thirdly, our algorithm provides a quantum algorithm for solving ill-posed problems that can be implemented on a quantum computer (see section 7).
4.3 A higher order improvement in
The first order convergence in was due to the lack of regularity in , which is continuous but not in . The convergence rate can be improved by choosing a smoother initial data for in the extended space. For example, consider the following setup:
(4.28) |
where defined in satisfies
(4.29) |
with . It is easy to check the Fourier transform of denoted by satisfies
where , and . In addition, the results in section 2, 3 still hold, since we do not care the solution when .
From Eq.(4.16) and Theorem 4.1, the limitation of the convergence order mainly comes from the non-smoothness of . In order to improve the whole convergence rates, we could smooth by using higher order polynomials, for example
(4.30) |
where is a Hermite interpolation [33, section 2.1.5] and satisfies
Here is an integer. Therefore, and one has the estimate of as
The choice of can be chosen such that . We use in our numerical experiments in section 6.
The linear combination of Hamiltonian simulation (LCHS) method introduced in [2] employs the continuous Fourier transform in , and then applying numerical integration to complete the inverse of Fourier transform, choosing to recover the target variables . Initially, it is a first-order method. Subsequent enhancements in [1] improve the accuracy of the original LCHS algorithms by introducing new kernel functions with faster decay and truncating the Fourier transform integral to a finite interval , where () truncates the continuous Fourier variable . We emphasize that the fundamental principle remains consistent with our approach to smooth extension. Indeed, the Fourier transforms of the kernel functions are for , whereas they exhibit significant differences on the negative real axis. Therefore, the transformed kernel function can be interpreted as a smooth extension in the space.
5 Application to the convection equations with purely imaginary wave speed
In this section, we concentrate on the linear convection equation with an imaginary convection speed which is unstable thus hard to simulate numerically. This is a simple model for more complicated, physically interesting problems such as Maxwell’s equation with negative index of refraction [30, 28] that has applications in meta materials. We consider the one-dimensional model with periodic boundary condition
(5.1) |
where . This problem is ill-posed since, by taking a Fourier transform on , one gets
where is the Fourier transform of in . Assume the solution of (5.1) in Sobolev space exits, the exact solution is
(5.2) |
Clearly, the solution contains exponential growing modes corresponding to . Suppose , then . The Schrödingerisation gives
(5.3) |
where . The Fourier transform shows
(5.4) |
which has a bounded solution for all time, with the exact solution given by
Obviously, the -equation (5.3) and the Hamiltonian system (5.4) can be simulated by a stable scheme in both quantum computers and classical ones. We define an approximate solution to the truncation as follows.
Definition 5.1.
Define
(5.5) |
where is a characteristic function of . The truncated solution is
(5.6) |
with .
Following Theorem 3.1, one gets the estimate for recovery. The proof is omitted here.
Theorem 5.1.
Similar to Remark 3.2, it follows from that is a small number. In order to keep the error within , one could choose . For example, if , then one gets for .
5.1 Discretization for Schrödingerisation of the convection equations with imaginary convective terms
In this section, we give the discretization of Schrödingerisation for the convection equations with purely imaginary wave speed. If we use the spectral method, the algorithm is quite similar to section 4.1. It is obtained by letting in (4.5). It is obvious to see that is Hermitian and has both positive and negative eigenvalues. It is hard to construct a stable scheme if one solves the original problem. However, after Schrödingerisation, it becomes a Hamiltonian system which is quite easy to approximate in both classical and quantum computers. Applying Theorem 5.1, one gets the desired variables.
6 Numerical tests
In this section, we perform several numerical simulations for the backward heat equation and imaginary convection equation by Schrödingeriza-tion in order to check the performance (in recovery and convergence rates) of our methods. For all numerical simulations, we perform the tests in the classical computers by using the Crank-Nicolson method for temporal discretization. In addition, we apply the trapezoidal rule for numerical integration of all of the tests. In order to get higher-order convergence rates of discretization (4.5), we smooth the initial data in (3.5) by choosing
(6.1) |
Therefore, .
6.1 Approximation of smooth solutions
These tests are used to evaluate the recovery and the order of accuracy of our method. More precisely, we want to show the choice of in (3.18) and convergence rates in Theorems 4.1.
In the first test, we consider the backward heat equation in with Dirichlet boundary condition and . We take a smooth solution
(6.2) |
and the initial data is . We use the finite difference method for spatial discretization with defined by
(6.3) |
where is the mesh size. The computation stops at . According to (3.18), one has and . The numerical results are shown in Fig. 1 and Tab. 1. From the plot on the left in Fig. 1, it can be seen that the error between and drops precipitously at , and numerical solutions of Schrödingerisation recovered by choosing one point or numerical integration are close to the exact solution. According to Tab. 1, second-order convergence rates of and are obtained, respectively, where .


order | order | order | ||||
1.64e-03 | - | 3.25e-04 | 2.33 | 8.05e-05 | 2.01 | |
7.48e-04 | - | 1.86e-04 | 2.01 | 4.56e-05 | 2.00 |
Next, we consider another exact solution set by
(6.4) |
with periodic boundary conditions and the initial data is . We apply the spectral method to discrete the -domain. From Theorem 3.1 and Remark 3.2, we deduce that . From Fig. 2, it can be seen that when , and the numerical solutions are very close to the exact solution. In Tab. 2, the errors of the method (4.5)-(4.11) are presented. They show the optimal order , are obtained when , where , and , .


6.2 Approximation of a piece-wise smooth solution
Now, we consider a piece-wise smooth solution in
(6.5) |
The tests are done in the interval and , . We use the finite difference method to simulate the heat equation with the initial condition given by (6.5). Thus, we get the approximate solution to at . According to Theorem 3.1, we choose , and , and for recovering the target variables , and , respectively. Thus, it yields , and . The results are shown in Fig. 3 with computed by (4.5)-(4.11) and , . From the plot of Fig. 3, we find that the error is larger when is smaller, which is consistent with the estimate in Theorem 3.1.



6.3 Input data with noises
In this test, we investigate the cases with the input data containing noises. The exact solution in is set by
(6.6) |
The input data is . The magnitude indicates the noise level of the measurement data, and is a random number such that . The tests are truncated to the interval and . The noise levels are , and . We apply (4.5)-(4.11) to discretize the Schrödingerisation with and . We use the finite difference method to discretize the -domain with , . The results are shown in Fig. 4. According to Theorem 3.2, it is hard to find to satisfy for any fixed , and . However one could choose to get as tends to zero under the assumption , . By investigating the cases of different , we find the numerical solution of Schrödingerisation approximates and tends to the exact solution as goes to zero. The oscillation of the numerical solution comes from two aspects, one is the truncation of the calculation area of , and the other is the finite difference method of calculating the discontinuous input data.






6.4 The convection equation with imaginary wave speed
Now we consider the convection equations with imaginary wave speed to test the accuracy of the recovery. In this test, the exact solution is
(6.7) |
and the simulation stops at . Therefore the source term is obtained by with defined in (6.7). We use the spectral method on and , then get the linear system:
(6.8) |
where , with . In order to obtain unitary dynamical systems to facilitate operation on quantum computers, one needs to use the technique of dealing with source terms in [15]. In Fig. 5, one can find which confirms Theorem 5.1, and the numerical solutions are close to the exact one. Since the solution is not smooth enough when using the spectral method, the oscillation appears in the error between and in Fig. 5.


7 Quantum algorithms for ill-posed problems
The above numerical sch- eme is a new scheme that can be used on both classical and quantum devices based on qubits or qumodes (continuous variables)[14].
7.1 Qubit-based quantum computing
Since Eq.(4.5) is a Hamiltonian system, a quantum simulation can be carried out such as quantum signal processing (QSP)[24], linear combination of unitaries (LCU) [7], etc. The complete circuit for implementing the quantum simulation of is illustrated in Fig. 6, where is a quantum state.
\Qcircuit@C=1em @R=2em
\lstick
& \qw \gateQFT
\qw \multigate1U
\qw \gateIQFT
\qw \qw\meterB— k ⟩
\lstick
\qw \gateQFT
\qw \ghostU \qw \gateIQFT
\qw \qw
For the specific ill-posed problems, the difference from well-posed problems lie in (i) the cost of unitarily evolving the system in Eq. (4.5), and (ii) the cost of final measurement to retrieve , , depending on the specification of possible for different problems.
Since the evolution is unitary, the cost analysis is similar whether one is running forward or backward in time. The only difference is that the initial state used in the forward equation is plus the ancilla state, whereas in the backward equation it is plus the same ancilla state (see Fig. 6). In quantum algorithms, Hamiltonian simulation with nearly optimal dependence on all parameters can be found in [5], with complexity given by the next lemma.
Lemma 7.1.
For a Hamiltonian system with a Hermitian matrix acting on qubits can be simulated within error with queries and additional 2-qubits gates, where , is the sparsity of (maximum number of nonzero entries in each row), is the max-norm (value of largest entry in absolute value), is the evolution time, and
In the case of Eq.(4.5), since the spatial discretization is a second-order scheme , we need qubits for variable where is the dimension. It is crucial to note that if one employs the smooth initialization method for the Schrödingerized equation, as introduced in section 4.3, and discretizes this smooth function in terms of instead, both the initial data and the solution in will belong to the class for any integer . Consequently, if a spectral method is utilized, this approach will yield essentially spectral accuracy in the approximations within the extended space. The extra qubits used in the extension domain is almost . Therefore, the required number of qubits is
the Hermitian matrix satisfies and the sparsity . By applying Lemma 7.1, one can directly obtain the complexity of Schrödingerisation as follows.
Lemma 7.2.
Given sparse-access to the Hermitian matrix , and the initial quantum state has been obtained. The quantum state can be prepared to precision with queries and additional 2-qubits gates, where denotes the case where all logarithmic factors are suppressed.
This reduces the complexity of the Schrödingerisation based quantum algorithms for any dynamical system introduced in [16, 17], wherein the number of 2-qubit gates required for the heat equation is . Furthermore, this quantum approach demonstrates polynomial-level acceleration compared to classical algorithms for solving the forward heat equation.
Different values of affect the probability of retrieving the correct final quantum state for . The final step involves measurement of . There are in principle two schemes from Theorem 2.1, where both schemes give a success probability proportional to , thus measurements are needed, where is a factor depending on , for instance see [16, 14]. The first scheme is the recovery by measurement of one point in Eq. (2.4). This corresponds to accepting the resulting state so long as the measured . In the qubit context for example, the largest value , if . The second scheme is the recovery through the integration method in Eq. (2.5). For a bounded domain, the retrieval process then involves the projection of onto the domain. In this case, . Thus, for large domains where , the first and second schemes are comparable. However, for specific cases, such as in Remark 3.2, the probability to retrieving the state is about
which is close to if we choose . This condition aligns with the requirement for choosing to successfully recover , as analyzed in Remark 3.2.
7.2 Qumode-based analog quantum computing
Since Eq. (2.1) is already of Schrödinger’s form and can be rewritten with obeying
(7.1) |
in the continuous-variable formalism and is the momentum operator conjugate to the position operator which has as eigenvalues and . Since is clearly Hermitian then evolves by unitary evolution generated by the Hamiltonian . We note that here, since we actually want to run the original PDE backwards in time, the state above corresponds in fact to the state . If we want the state at some , then the time to run Eq. (7.1) is over .
In the continuous-variable formulation, how arises is more subtle. The finite can come instead from a kind of energy constraint on the continuous-variable ancillary quantum mode, where is acting like the maximum momentum for the mode (conjugate momentum to ). At the measurement step, here is actually a maximum value of measurable by the physical measurement apparatus. The detailed analysis here is more subtle and depends on the physical system studied, so we leave this to future work.
8 Conclusions
In this paper we present computationally stable numerical methods for two ill-posed problems: the (constant and variable coefficient) backward heat equation and the linear convection equation with imaginary wave speed. The main idea is to use Schrödingerisation, which turns typical ill-posed problems into unitary dynamical systems in one higher dimension. The Schrödingerised system is a Hamiltonian system valid both forward and backward in time, hence suitable for the design of a computationally stable numerical scheme. A key idea is given to recover the solution to the original problem from the Schrödingerised system using data from appropriate domain in the extended space. Some sharp error estimates between the approximate solution and exact solution are provided and also verified numerically. Moreover, a smooth initialization is introduced for the Schrödingerised system which leads to requiring an exponentially lower number of ancilla qubits for qubit-based quantum algorithms for any non-unitary dynamical system.
In the future this technique will be generalized to more realistic and physically interesting ill-posed problems.
Acknowledgement
SJ and NL are supported by NSFC grant No. 12341104, the Shanghai Jiao Tong University 2030 Initiative and the Fundamental Research Funds for the Central Universities. SJ was also partially supported by the NSFC grants No. 12031013. NL also acknowledges funding from the Science and Technology Program of Shanghai, China (21JC1402900). CM was partially supported by China Postdoctoral Science Foundation (No. 2023M732248) and Postdoctoral Innovative Talents Support Program (No. BX20230219).
References
- [1] D. An, A. W. Childs, and L. Lin, Quantum algorithm for linear non-unitary dynamics with near-optimal dependence on all parameters, arXiv:2312.03916, 2023.
- [2] D. An, J. Liu, and L. Lin, Linear combination of Hamiltonian simulation for non-unitary dynamics with optimal state preparation cost , Phys. Rev. Lett., 131 (2023), 150603.
- [3] K.A. Ames, W.C. Gordon, J.F. Epperson, and S.F. Oppenheimer, A comparison of regularizations for an ill-posed problem, Math. Comp., 67 (1998), pp.1451–1471.
- [4] M. Bertero, P. Boccacci, and C. De Mol, Introduction to inverse problems in imaging, CRC press, 2021.
- [5] D.W. Berry, A.M. Childs, R. Kothari, Hamiltonian simulation with nearly optimal dependence on all parameters, IEEE 56th annual symposium of foundation of computer science, 2015.
- [6] U. Biccari, Y. Song, X. Yuan, and E. Zuazua, A two-stage numerical approach for the sparse initial source identification of a diffusion-advection equation, Inverse Problems, 39 (2023), 095003.
- [7] A. M. Childs, and N. Wiebe, Hamiltonian simulation using linear combinations of unitary operations, Quantum Inf. Comput., 12 (2012), pp. 901-924.
- [8] Heinz W. Engl, K. Kunisch, and A. Neubauer, Convergence rates for Tikhonov regularisation of non-linear ill-posed problems, Inverse problems, 5 (1989), 523.
- [9] L. C. Evans, Partial differential equations, American Mathematical Society, 2016.
- [10] T. Funada and D. Joseph, Viscous potential flow analysis of Kelvin–Helmholtz instability in a channel, J. Fluid Mech., 445 (2001), pp.263-283.
- [11] D. N. Háo, A mollification method for ill-posed problems, Numer. Math., 68 (1994) pp.469-506.
- [12] A. Hasegawa, Plasma instabilities and nonlinear effects, volume 8, Springer Science & Business Media, 2012.
- [13] K. Höllig, Existence of infinitely many solutions for a forward backward heat equation, Trans. Amer. Math. Soc., 278 (1983) pp.299-316.
- [14] S. Jin and N. Liu, Analog quantum simulation of partial differential equations, Quantum Sci. Technol., 9 (2024), 035047.
- [15] S. Jin, N. Liu, and C. Ma, On schrödingerization based quantum algorithms for linear dynamical systems with inhomogeneous terms, arXiv:2402.14696v2, 2024.
- [16] S. Jin, N. Liu, and Y. Yu, Quantum simulation of partial differential equations via schrödingerisation, arXiv preprint arXiv:2212.13969., 2022.
- [17] S. Jin, N. Liu, and Y. Yu, Quantum simulation of partial differential equations: Applications and detailed analysis, Phys. Rev. A, 108(3):032603, 2023.
- [18] F. John, Numerical solution of the equation of heat conduction for preceding times, Ann. Mat. Pura Appl., 40 (1955), pp.129-142.
- [19] M. Jourhmane and N. S. Mera, An iterative algorithm for the backward heat conduction problem based on variable relaxation factors, Inverse Problems in Engineering, 10 (2002), pp.293-308.
- [20] H. C. Kaiser and J. Rehberg On stationary Schrödinger–Poisson equations modelling an electron gas with reduced dimension, Math. Methods Appl. Sci., 20 (1997), pp. 1283-1312.
- [21] H. J. Kull, Theory of the Rayleigh-Taylor instability, Phys. Rep., 206 (1991), pp.197-325.
- [22] R. Lattès and J. L. Lions, Méthode de quasi-réversibilité et applications, English translation: R. Bellman, Elsevier, New York, 1969.
- [23] F. Lin, Remarks on a backward parabolic problem, Methods Appl. Anal., 10(2003), pp.245-252.
- [24] G. H. Low and I.L. Chuang, Optimal Hamiltonian simulation by quantum signal processing, Phys. Rev. Lett., 118 (2017), 010501.
- [25] K. Miller, Stabilized quasi-reversibilite and other nearly best possible methods for non-well-posed problems, In: Symposium on Non-Well-Posed Problems and Logarithmic Convexity, in: Lecture Notes in Math., 316 (1973), pp.161-176.
- [26] W. Miranker, A well posed problem for the backward heat equation, Proc. Amer. Math. Soc., 12 (1961), pp.243–247.
- [27] J. Nash, Continuity of solutions of parabolic and elliptic equations, Amer. J. Math., 80(1958), pp.931–954.
- [28] C. G. Parazzoli, R.B. Greegor, K. Li, B. E. C. Koltenbah, and M. Tanielian, Experimental verification and simulation of negative index of refraction using Snell’s law, Phys. Rev. Lett., 90(2003), 107401.
- [29] Z. Qian, C. L. Fu, and R. Shi, A modified method for a backward heat conduction problem, Appl. Math. Comput., 185(2007), pp.564-573.
- [30] R. A. Shelby, D. R. Smith, and S. Schultz, Experimental verification of a negative index of refraction, Science, 292 (2001), pp.77-79.
- [31] J. Shen, T. Tang, and L. Wang, Spectral methods, Springer-Verlag Berliln Heidelberg, 2011.
- [32] R.E. Showalter, The final value problem for evolution equations, J. Math. Anal. Appl., 47(1974), pp.563-572.
- [33] J. Stoer, R. Bulirsch, R. Bartels, W. Gautschi, and C. Witzgall, Introduction to numerical analysis, New York: Springer-Verlag, 1980.
- [34] G. Uhlmann, Inverse problems: seeing the unseen, Bull. Math. Sci., 4(2014), pp.209-279.
- [35] X. T. Xiong, C. L. Fu, and Z. Qian, Two numerical methods for solving a backward heat conduction problem, Appl. Math. Comput., 179(2006), pp.370-377.