0% found this document useful (0 votes)
15 views22 pages

Diffusion PDE

Uploaded by

fakeherolimit
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
15 views22 pages

Diffusion PDE

Uploaded by

fakeherolimit
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 22

DiffusionPDE: Generative PDE-Solving

Under Partial Observation

Jiahe Huang1 Guandao Yang2 Zichen Wang1 Jeong Joon Park1


1
University of Michigan
2
Stanford University
{chloehjh, zzzichen, jjparkcv}@umich.edu
arXiv:2406.17763v1 [cs.LG] 25 Jun 2024

guandao@stanford.edu

Abstract
We introduce a general framework for solving partial differential equations (PDEs)
using generative diffusion models. In particular, we focus on the scenarios where
we do not have the full knowledge of the scene necessary to apply classical solvers.
Most existing forward or inverse PDE approaches perform poorly when the ob-
servations on the data or the underlying coefficients are incomplete, which is a
common assumption for real-world measurements. In this work, we propose Diffu-
sionPDE that can simultaneously fill in the missing information and solve a PDE by
modeling the joint distribution of the solution and coefficient spaces. We show that
the learned generative priors lead to a versatile framework for accurately solving
a wide range of PDEs under partial observation, significantly outperforming the
state-of-the-art methods for both forward and inverse directions. See our project
page for results: jhhuangchloe.github.io/Diffusion-PDE/.

1 Introduction
Partial differential equations (PDEs) are a cornerstone of modern science, underpinning many
contemporary physical theories that explain natural phenomena. The ability to solve PDEs grants
us the power to predict future states of a system (forward process) and estimate underlying physical
properties from state measurements (inverse process).
To date, numerous methods [1, 2] have been proposed to numerically solve PDEs for both the
forward and inverse directions. However, the classical methods can be prohibitively slow, prompting
the development of data-driven, learning-based solvers that are significantly faster and capable of
handling a family of PDEs. These learning-based approaches [3–6] typically learn a deterministic
mapping between input coefficients and their solutions using deep neural networks.
Despite the progress, existing learning-based approaches, much like classical solvers, rely on complete
observations of the coefficients to map solutions. However, complete information on the underlying
physical properties or the state of a system is rarely accessible; in reality, most measurements are
sparse in space and time. Both classical solvers and the state-of-the-art data-driven models often
overlook these scenarios and consequently fail when confronted with partial observations. This
limitation confines their use primarily to synthetic simulations, where full scene configurations are
available by design, making their application to real-world cases challenging.
We present a comprehensive framework, DiffusionPDE, for solving PDEs in both forward and
inverse directions under conditions of highly partial observations—typically just 1~3% of the total
information. This task is particularly challenging due to the numerous possible ways to complete
missing data and find subsequent solutions. Our approach uses a generative model to formulate the
joint distribution of the coefficient and solution spaces, effectively managing the uncertainty and
simultaneously reconstructing both spaces. During inference, we sample random noise and iteratively

Preprint. Under review.


Training
Joint Distribution Diffusion Model Predictions

Coefficient/Initial State: a Solution/Final State: u (on both a and u)


Generated a
...

... ...

Observations
Inference
(on either side or both)
Generated u
Sparse Ground Truth a Sparse Ground Truth u Diffusion Model

Noise PDE Guidance

Observation Guidance

Figure 1: We propose DiffusionPDE, a generative PDE solver under partial observations. Given
a family of PDE with coefficient (initial state) a and solution (final state) u, we train the diffusion
model on the joint distribution of a and u. During inference, we gradually denoise a Gaussian noise,
guided by sparse observation and known PDE function, to recover the full prediction of both a and u
that align well with the sparse observations and the given equation.

denoise it following standard diffusion models [7]. However, we uniquely guide this denoising
process with sparse observations and relevant PDE constraints, generating plausible outputs that
adhere to the imposed constraints. Notably, DiffusionPDE can handle observations with arbitrary
density and patterns with a single pre-trained generative network.
We conduct extensive experiments to show the versatility of DiffusionPDE as a general PDE-solving
framework. We evaluate it on a diverse set of static and temporal PDEs, including Darcy Flow,
Poisson, Helmholtz, Burger’s, and Navier-Stokes equations. DiffusionPDE significantly outperforms
existing state-of-the-art learning-based methods for solving PDEs [3–6, 8] in both forward and inverse
directions with sparse measurements, while achieving comparable results with full observations.
Highlighting the effectiveness of our model, DiffusionPDE accurately reconstructs the complete state
of Burgers’ equation using time-series data from just five sensors (Fig. 4), suggesting the potential of
generative models to revolutionize physical modeling in real-world applications.

2 Related Works
Our work builds on the extensive literature of three areas: forward PDE solvers, inverse PDE solvers,
and diffusion models. Please see relevant surveys for more information [9–13].
Forward PDE Solvers. PDE solvers take the specification of a physics system and predict its state
in unseen space and time by solving an equation involving partial derivatives. Since Most PDEs
are very challenging to solve analytically, people resolve to numerical techniques, such as Finite
Element Method [14, 2] and Boundary Element Method [1, 15]. While these techniques show strong
performance and versatility in some problems, they can be computationally expensive or difficult
to set up for complex physics systems. Recently, advancements in deep-learning methods have
inspired a new set of PDE solvers. Raissi et al. [16, 6] introduce Physics-Informed Neural Networks
(PINNs), which optimize a neural network using PDE constraints as self-supervised losses to output
the PDE solutions. PINNs have been extended to solving specific fluid [17, 18], Reynolds-averaged
Navier–Stokes equations [19], heat equations [20], and dynamic power systems [21]. While PINNs
can tackle a wide range of complex PDE problems, they are difficult to scale due to the need for
network optimization. An alternative approach, neural operators [3, 5], directly learn the mapping
from PDE parameters (e.g.initial and boundary condition) to the solution function. Once trained, this
method avoids expensive network optimization and can instantly output the solution result. This idea

2
has been extended to solve PDE in 3D [22, 23] , multiphase flow [24], seismic wave [25, 26], 3D
turbulence [27, 28], and spherical dynamics [29]. People have also explored using neural networks as
part of the PDE solver, such as compressing the physics state [30–33]. These solvers usually assume
known PDE parameters, and applying them to solve the inverse problem can be challenging.
PDE inverse problem. The inverse problem refers to finding the coefficients of a PDE that can
induce certain observations, mapping from the solution of a PDE solver to its input parameters.
People have tried to extend traditional numerical methods to this inverse problem [34–38], but these
extensions are non-trivial to implement efficiently. There are similar attempts to inverse deep-learning
PDE solvers. For example, one can inverse PINNs by optimizing the network parameters such
that their outputs satisfy both the observed data and the governing equations. iFNO [39] and NIO
[40] tries to extend FNO [3]. Other methods [41, 42] directly learn the operator functions for the
inverse problem. PINO [4] further combines neural operators with physics constraints to improve
the performance of both forward and inverse problems. These methods assume full observations are
available. To address the inverse problem with partial observations, people have tried to leverage
generative priors with Graph neural networks [43, 8]. These works have not demonstrated the ability
to solve high-resolution PDEs, possibly limited by the power of generative prior. We want to leverage
the state-of-the-art generative model, diffusion models, to develop a better inverse PDE solver.
Diffusion models. Diffusion models have shown great promise in learning the prior with higher
resolutions by progressively estimating and removing noise. Models like DDIM [44], DDPM [7], and
EDM [45] offer expressive generative capabilities but face challenges when sampling with specific
constraints. Guided diffusion models [46–49] enhance generation processes with constraints such
as image inpainting, providing more stable and accurate solutions. Prior works on diffusion models
for PDEs highlight the potential of diffusion approaches by generating PDE datasets such as 3D
turbulence [50, 51] and Navier-Stokes equations [52] with diffusion models. Diffusion models can
also be used to model frequency spectrum and denoise the solution space [53], and conditional
diffusion models are applied to solve 2D flows with sparse observation [54]. However, the application
of diffusion models to solve inverse problems under partial observation remains underexplored. In
this work, we aim to take the initial steps towards addressing this gap.

3 Methods
3.1 Overview

To solve physics-informed forward and inverse problems under uncertainty, we start by pre-training a
diffusion generative model on a family of partial differential equations (PDEs). This model is designed
to learn the joint distribution of the PDE coefficients (or the initial state) and its corresponding
solutions (or the final state). Our approach involves recovering full data in both spaces using
sparse observations from either or both sides. We achieve this through the iterative denoising of
random Gaussian noise as in regular diffusion models but with additional guidance from the sparse
observations and the PDE function enforced during denoising. The schematic description of our
approach is shown in Fig. 1.

3.2 Prelimary: Diffusion Models and Guided Diffusion

Diffusion models involve a predefined forward process that gradually adds Gaussian noise to the
data and a learned reverse process that denoises the data to reconstruct the original distribution.
Specifically, Song et al. [55] propose a deterministic diffusion model that learns an N -step denoising
process that eventually outputs a denoised data xN and satisfies the following ordinary differential
equations (ODE) at each timestep ti where i ∈ {0, 1, ..., N − 1}

dx = −σ̇(t)σ(t)∇x log p x; σ(t) dt. (1)

Here ∇x log p x; σ(t) is the score function [56] that helps to transform samples from a normal
distribution N (0, σ(t0 )2 I) to a target probability distribution p(x; σ(t)). To estimate the score
function, Karras et al. [45] propose to learn a denoiser function D(x; σ) such that
∇x log p x; σ(t) = (D(x; σ(t)) − x)/σ(t)2

(2)

To enable control over the generated data, guided diffusion methods [48] add guidance gradients to
the score function during the denoising process. Recently, diffusion posterior sampling (DPS) [46]

3
Algorithm 1 Sparse Observation and PDE Guided Diffusion Sampling Algorithm.
1: input DeterministicSampler Dθ (x; σ), σ(ti∈{0,...,N } ), TotalPointCount m, ObservedPointCount
n, Observation y, PDEFunction
 f , Weights ζobs , ζpde
2: sample x0 ∼ N 0, σ(t0 )2 I ▷ Generate initial sampling noise
3: for i ∈ {0, . . . , N − 1} do
4: x̂iN ← Dθ (xi ; σ(t i )) ▷ Estimate the denoised data at step ti
5: di ← xi − x̂iN /σ(ti ) ▷ Evaluate dx/dσ(t) at step ti
6: xi+1 ← xi + (σ(ti+1 ) − σ(ti ))di ▷ Take an Euler step from σ(ti ) to σ(ti+1 )
7: if σ(ti+1 ) ̸= 0 then
8: x̂iN ← Dθ (xi+1 ; σ(t i+1 )) ▷ Apply 2nd order correlation unless σ = 0
9: d′i ← xi+1 − x̂iN /σ(ti+1 ) ▷ Evaluate dx/dσ(t) at step ti+1
1 1 ′

10: xi+1 ← xi + (σ(ti+1 ) − σ(ti )) d
2 i + d
2 i ▷ Apply the trapezoidal rule at step ti+1
11: end if
12: Lobs ← n1 ∥y − x̂iN ∥22 ▷ Evaluate the observation loss of x̂iN
13: Lpde ← 1
m ∥0 − f (x̂iN )∥22 ▷ Evaluate the PDE loss of x̂iN
14: xi+1 ← xi+1 − ζobs ∇xi Lobs − ζpde ∇xi Lpde ▷ Guide the sampling with Lobs and Lpde
15: end for
16: return xN ▷ Return the denoised data

made notable progress in guided diffusion for tackling various inverse problems. DPS uses corrupted
measurements y derived from x to guide the diffusion model in outputting the posterior distribution
p(x|y). A prime application of DPS is the inpainting problem, which involves recovering a complete
image from sparsely observed pixels, which suits well with our task. This approach modifies Eq. 1 to
 
dx = −σ̇(t)σ(t) ∇x log p x; σ(t) + ∇x log p y|x; σ(t) dt. (3)

DPS [46] showed that under Gaussian noise assumption of the sparse measurement operator M(·),
i.e., y|x ∼ N (M(x), δ 2 I) with some S.D. δ, the log-likelihood function can be approximated with:
1
∇x log p y|xi ; σ(ti ) ≈ ∇xi log p y|x̂iN ; σ(ti ) ≈ − 2 ∇xi ∥y − M(x̂iN (xi ; σ(ti ))∥22 , (4)
 
δ
where x̂iN := D(xi ; σ(ti )) denotes the estimation of the final denoised data at each denoising step i.
Applying the Baye’s rule, the gradient direction of the guided diffusion is therefore:
∇xi log p(xi |y) ≈ s(xi ) − ζ∇xi ∥y − M(x̂iN )∥22 , (5)

where s(x) = ∇x log p x is the original score function, and ζ = 1/δ 2 .

3.3 Solving PDEs with Guided Diffusion

Our work focuses on two classes of PDEs: static PDEs and dynamic time-dependent PDEs. Static
systems (e.g., Darcy Flow or Poisson equations) are defined by a time-independent function f :
f (c; a, u) = 0 in Ω ⊂ Rd , u(c) = g(c) in ∂Ω, (6)
where Ω is a bounded domain, c ∈ Ω is a spatial coordinate, a ∈ A is the PDE coefficient field,
and u ∈ U is the solution field. ∂Ω is the boundary of the domain Ω and u|∂Ω = g is the boundary
constraint. We aim to recover both a and u from sparse observations on either a or u or both.
Similarly, we consider the dynamic systems (e.g., Navier-Stokes):
f (c, τ ; a, u) = 0, in Ω × (0, ∞)
u(c, τ ) = g(c, τ ), in ∂Ω × (0, ∞) (7)
u(c, τ ) = a(c, τ ), in Ω̄ × {0}
where τ is a temporal coordinate, a = u0 ∈ A is the initial condition, u is the solution field, and
u|Ω×(0,∞) = g is the boundary constraint. We aim to simultaneously recover both a and the solution
uT := u(·, T ) at a specific time T from sparse observations on either a, uT , or both.
Finally, we explore the recovery of the states across all timesteps u0:T in 1D dynamic systems
governed by Burger’s equation. Our network Dθ models the distribution of all 1D states, including
the initial condition u0 and solutions u1:T stacked in the temporal dimension, forming a 2D dataset.

4
Sparse observations on Sparse Observations on 
 Sparse Observations on Both
coefficients Solutions Coefficients and Solutions Ground Truth Partial Observations

Recovered Coefficients

Error Rate: 2.8% Error Rate: 3.2% Error Rate: 2.6%


Recovered Solutions

Relative Error: 2.5% Relative Error: 1.0% Relative Error: 0.9%

Figure 2: Different from forward and inverse PDE solvers, DiffusionPDE can take sparse observations
on either the coefficient a or the solution u to recover both of them, using one trained network. Here,
we show the recovered a and u of the Darcy’s eqaution given sparse observations on a, u, or both.
Compared with the ground truth, we see that our method successfully recovers the PDE in all cases.

Guided Diffusion Algorithm In the data-driven PDE literature, the above tasks can be achieved by
learning directional mappings between a and u (or uT for dynamic systems). Thus, existing methods
typically train separate neural networks for the forward solution operator F : A → U and the inverse
solution operator I : U → A.
Our method unifies the forward and inverse operators with a single network and an algorithm using
the guided diffusion framework. DiffusionPDE can handle arbitrary sparsity patterns with one
pre-trained diffusion model Dθ that learns the joint distribution of A and U, concatenated on the
channel dimension, denoted X . Thus, our data x ∈ X , where X := A × U. We follow the typical
diffusion model procedures [45] to train our model on a family of PDEs.
Once we train the diffusion model Dθ , we employ our physics-informed DPS [46] formulation during
inference to guide the sampling of x ∈ X that satisfies the sparse observations and the given PDE, as
detailed in Algorithm 1. We follow Eq. 5 to modify the score function using the two guidance terms:

∇xi log p(xi |y obs , f ) ≈ ∇xi log p xi ) − ζobs ∇xi Lobs − ζpde ∇xi Lpde , (8)

where xi is the noisy data at denoising step i, y obs are the observed values, and f (·) = 0 is
the underlying PDE condition. Lobs and Lpde respectively represent the MSE loss of the sparse
observations and the PDE equation residuals:
n
1 1X
Lobs (xi , y obs ; Dθ ) = ∥y obs − x̂iN ∥22 = (y (oj ) − x̂iN (oj ))2 ,
n n j=1 obs
(9)
1 1 XX
Lpde (xi ; Dθ , f ) = ∥0 − f (x̂iN )∥22 = f (cj , τk ; ûj , âj )2 ,
m m j
k

where x̂iN = Dθ (xi ) is the clean image estimate at denoising timestep i, which can be split into
coefficient ûi and solution âi . Here, m is the total number of grid points (i.e., pixels), n is the number
of sparse observation points. oj represents the spatio-temporal coordinate of jth observation. Note
that, without loss of generality, Lpde can be accumulated for all applicable PDE function f in the
system, and the time component τk is ignored for static systems.

4 Experiments

4.1 PDE Problem Settings

We show the usefulness of DiffusionPDE across various PDEs for inverse and forward problems and
compare it against recent learning-based techniques. We test on the following families of PDEs.

5
Guided with
 Guided with PDE Loss
 Guided with
 Guided with PDE Loss

Ground Truth Observation Loss and Observation Loss Ground Truth Observation Loss and Observation Loss

Coefficient Absolute Error

Solution Absolute Error


Coefficient

Solution
Figure 3: Usefulness of PDE loss. We visualize the absolute errors of the recovered coefficient and
solution of the Helmholtz equation with and w/o PDE loss. We compare having only the observation
loss with applying the additional PDE loss. The errors drop significantly when using PDE loss.

Darcy Flow. Darcy flow describes the movement of fluid through a porous medium. In our
experiment, we consider the static Darcy Flow with a no-slip boundary ∂Ω
−∇ · (a(c)∇u(c)) = q(c), c∈Ω
(10)
u(c) = 0, c ∈ ∂Ω
Here the coefficient a has binary values. We set q(c) = 1 for constant force. The PDE guidance
function is thus f = ∇ · (a(c)∇u(c)) + q(c).
Inhomogeneous Helmholtz Equation. We consider the static inhomogeneous Helmholtz Equation
with a no-slip boundary on ∂Ω, which describes wave propagation:
∇2 u(c) + k 2 u(c) = a(c), c∈Ω
(11)
u(c) = 0, c ∈ ∂Ω
The coefficient a is a piecewise constant function and k is a constant. Note 11 is the Poisson
equation when k = 0. Setting k = 1 for Helmholtz equations, the PDE guidance function is
f = ∇2 u(c) + k 2 u(c) − a(c).
Non-bounded Navier-Stokes Equation. We study the non-bounded incompressive Navier-Stokes
equation regarding the vorticity.
∂t w(c, τ ) + v(c, τ ) · ∇w(c, τ )= ν∆w(c, τ ) + q(c), c ∈ Ω, τ ∈ (0, T ]
(12)
∇ · v(c, τ ) = 0, c ∈ Ω, τ ∈ [0, T ]
Here w = ∇ × v is the vorticity, v(c, τ ) is the velocity at c at time τ , and q(c) is a force field. We
set the viscosity coefficient ν = 10−3 and correspondingly the Reynolds number Re = ν1 = 1000.
DiffusionPDE learns the joint distribution of w0 and wT and we take T = 10 which simulates
1 second. Since T ≫ 0, we cannot accurately compute the PDE loss from our model outputs.
Therefore, given that ∇ · w(c, τ ) = ∇ · (∇ × v) = 0, we use simplified f = ∇ · w(c, τ ).
Bounded Navier-Stokes Equation. We study the bounded 2D imcompressive Navier Stokes
regarding the velocity v and pressure p.
1
∂t v(c, τ ) + v(c, τ ) · ∇v(c, τ ) + ∇p = ν∇2 v(c, τ ), c ∈ Ω, τ ∈ (0, T ]
ρ (13)
∇ · v(c, τ ) = 0, c ∈ Ω, τ ∈ (0, T ].
We set the viscosity coefficient ν = 0.001 and the fluid density ρ = 1.0. We generate 2D cylinders
of random radius at random positions inside the grid. Random turbulence flows in from the top of the
grid, with the velocity field satisfying no-slip boundary conditions at the left and right edges, as well
as around the cylinder ∂Ωlef t,right,cylinder . DiffusionPDE learns the joint distribution of v0 and vT
at T = 4, which simulates 0.4 seconds. Therefore, we similarly use f = ∇ · v(c, τ ) as before.
Burgers’ Equation. We study the Burgers’ equation with periodic boundary conditions on a 1D
spatial domain of unit length Ω = (0, 1). We set the viscosity to ν = 0.01. In our experiment, the
initial condition u0 has a shape of 128 × 1, and we take 127 more time steps after the initial state to
form a 2D u0:T of size 128 × 128.
∂t u(c, τ ) + ∂c (u2 (c, τ )/2) = ν∂cc u(c, τ ), c ∈ Ω, τ ∈ (0, T ]
(14)
u(c, 0) = u0 (c), c∈Ω
We can reliably compute f = ∂t u(c, τ ) + ∂c (u2 (c, τ )/2) − ν∂cc u(c, τ ) with finite difference since
we model densely on the time dimension.

6
Table 1: Relative errors of solutions (or final states) and coefficients (or initial states) when solving
forward and inverse problems respectively with sparse observations. Error rates are used for the
inverse problem of Darcy Flow.
DiffusionPDE PINO DeepONet PINNs FNO
Forward 2.5% 35.2% 38.3% 48.8% 28.2%
Darcy Flow
Inverse 3.2% 49.2% 41.1% 59.7% 49.3%
Forward 4.5% 107.1% 155.5% 128.1% 100.9%
Poisson
Inverse 20.0% 231.9% 105.8% 130.0% 232.7%
Forward 8.8% 106.5% 123.1% 142.3% 98.2%
Helmholtz
Inverse 22.6% 216.9% 132.8% 160.0% 218.2%
Non-bounded Forward 6.9% 101.4% 103.2% 142.7% 101.4%
Navier-Stokes Inverse 10.4% 96.0% 97.2% 146.8% 96.0%
Bounded Forward 3.9% 81.1% 97.7% 100.1% 82.8%
Navier-Stokes Inverse 2.7% 69.5% 91.9% 105.5% 69.6%

4.2 Dataset Preparation and Training

We first test DiffusionPDE on jointly learning the forward mapping F : A → U and the inverse
mapping I : U → A given sparse observations. In our experiments, we define our PDE over the unit
square Ω = (0, 1)2 , which we represent as a 128 × 128 grid. We utilize Finite Element Methods
(FEM) to generate our training data. Specifically, we run FNO’s [3] released scripts to generate
Darcy Flows and the vorticities of the Navier-Stokes equation. Similarly, we generate the dataset of
Poisson and Helmholtz using second-order finite difference schemes. To add more complex boundary
conditions, we use Difftaichi [57] to generate the velocities of the bounded Navier-Stokes equation.
We train the joint diffusion model for each PDE on three A40 GPUs for approximately 4 hours, using
50,000 data pairs. For Burgers’ equation, we train the diffusion model on a dataset of 50,000 samples
produced as outlined in FNO [3]. We randomly select 5 out of 128 spatial points on Ω to simulate
sensors that provide measurements across time.

4.3 Baseline Methods


We compare DiffusionPDE with state-of-the-art learning-based methods, including PINO [4], Deep-
ONet [5], PINNs [6], and FNO [3]. However, note that none of these methods show operation on
partial observations. These methods can learn mappings between a and u or u0 and u1:T with full
observations, allowing them to also solve the mapping between u0 and uT . PINNs map input a
to output u by optimizing a combined loss function that incorporates both the solution u and the
PDE residuals. DeepONet employs a branch network to encode input function values sampled at
discrete points and a trunk network to handle the coordinates of the evaluated outputs. FNO maps
from the parametric space to the solution space using Fourier transforms. PINO enhances FNO by
integrating PDE loss during training and refining the model with PDE loss finetuning. We train all
four baseline methods on both forward and inverse mappings using full observation of a or u for
both static and dynamic PDEs. We tried training the baseline models on partial observations, but we
noticed degenerate training outcomes (see supplementary for details). Overall, they are intended for
full observations and may not be suitable for sparse measurements.
More closely related to our method, GraphPDE [8] demonstrates the ability to recover the initial
state using sparse observations on the final state, a task that other baselines struggle with. Therefore,
we compare against GraphPDE for the inverse problem of bounded Navier-Stokes (NS) equation,
which is the setup used in their report. GraphPDE uses a trained latent space model and a bounded
forward GNN model to solve the inverse problem with sparse sensors and thus is incompatible with
unbounded Navier-Stokes. We create bounded meshes using our bounded grids to train the GNN
model and train the latent prior with v0:T for GraphPDE.
4.4 Main Evaluation Results
We respectively address the forward problem and the inverse problem with sparse observations of a
or u. For the forward problem, we randomly select coefficients (initial states) as sparse observations

7
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs

Non-bounded Navier-Stokes Equation


Initial State
Forward

Final State
y

x 500 Random Points Relative Error: 6.9% Relative Error: 103.2% Relative Error: 101.4%Relative Error: 101.4% Relative Error: 142.7%

Inverse

Initial State
Final State

x Relative Error: 10.4% Relative Error: 97.2% Relative Error: 96.0% Relative Error: 96.0% Relative Error: 146.8%
5 Random

Burgers’ Equation

Sensors
Time

Time

x Relative Error: 2.68% Relative Error: 92.0% Relative Error: 85.7% Relative Error: 90.0% Relative Error: 92.4%
Figure 4: We compare DiffusionPDE with state-of-the-art neural PDE solvers [3–6]. In the forward
Navier-Stokes problem, we give 500 sparse observations of the initial state to solve for the final
state. In the inverse set-up, we take observations of the final state and solve for the initial. For the
Burgers’ equation, we use 5 sensors throughout all time steps and want to recover the solution at
all time steps. Note that we train on neighboring snapshot pairs for the baselines in order to add
continuous observations of the Burgers’ equation. Results show that existing methods do not support
PDE solving under sparse observations, and we believe they are not easily extendable to do so. We
refer readers to the supplementary for a complete set of visual results.

and then compare the predicted solutions (final states) with the ground truth. Specifically, we select
500 out of 128 × 128 points, approximately 3%, on the coefficients of Darcy Flow, Poisson equation,
Helmholtz equation, and the initial state of the non-bounded Navier-Stokes equation. For the bounded
Navier-Stokes equation, we use 1% observed points beside the boundary of the cylinder in 2D.
Similarly, for the inverse problem, we randomly sample points on solutions (final states) as sparse
observations, using the same number of observed points as in the forward model for each PDE.
We show the relative errors of all methods regarding both forward and inverse problems in Table
1. Since the coefficients of Darcy Flow are binary, we evaluate the error rates of our prediction.
Non-binary data is evaluated using mean pixel-wise relative error. We report error numbers aver-
aged across 1,000 random scenes and observations for each PDE. DiffusionPDE outperforms other
methods including PINO [4], DeepONet [5], PINNs [6], and FNO [3] for both directions with sparse
observations, demonstrating the novelty and uniqueness of our approach. For the inverse problems of
the Poisson and Helmholtz equations, DiffusionPDE exhibits higher error rates due to the insufficient
constraints within the coefficient space, produced from random fields. In Fig. 4, we visualize the
results for solving both the forward and inverse problem of the non-bounded Navier-Stokes. We refer
to the supplementary for additional visual results. While other methods may produce partially correct
results, DiffusionPDE outperforms them and can recover results very close to the ground truth.
For the inverse problem of the bounded Navier-Stokes equation, we further compare DiffusionPDE
with GraphPDE, as illustrated in Fig. 5. Our findings reveal that DiffusionPDE surpasses GraphPDE
[8] in accuracy, reducing the relative error from 12.0% to 2.7% with only 1% observed points.
We further show whether DiffusionPDE can jointly recover both a and u by analyzing the retrieved
a and u with sparse observations on different sides as well as on both sides. In Fig. 2, we recover the
coefficients and solutions of Darcy Flow by randomly observing 500 points on only coefficient space,
only space solution space, and both. Both coefficients and solutions can be recovered with low errors
for each situation. We therefore conclude that DiffusionPDE can solve the forward problem and the
inverse problem simultaneously with sparse observations at any side without retraining our network.

4.5 Recovering Solutions Throughout a Time Interval

We demonstrate that DiffusionPDE is capable of retrieving all time steps throughout the time
interval [0, T ] from continuous observations on sparse sensors. To evaluate its ability to recover
u0:T with sparse sensors, we study the 1D dynamic Burgers’ equation, where DiffusionPDE learns

8
Ground Truth Observed Final State Cylinder Boundary
GraphPDE DiffusionPDE Ground Truth

Initail State
Inverse

Final State

1% Random Points Relative Error: 12.0% Relative Error: 2.7%

Figure 5: We compare GraphPDE [8] and our method for solving the inverse bounded Navier-Stokes
equation. Given the boundary conditions and 1% observations of the final vorticity field, we solve the
initial vorticity field. We set the fliuds to flow in from the top, with boundary conditions at the edges
and a middle cylinder. While GraphPDE can recover the overall pattern of the initial state, it suffers
from noise when the fluid passes the cylinder and misses the high vorticities at the bottom.

the distribution of u0:T using a 2D diffusion model. To apply continuous observation on PINO,
DeepONet, FNO, and PINNs, we train them on neighboring snapshot pairs. Our experiment results
in a test relative error of 2.68%, depicted in Fig. 4, which is significantly lower than other methods.

4.6 Additional Analysis

We examine the effects of different components of our algorithm such as PDE loss and observation
samplings. We strongly encourage readers to view the supplementary for more details of these
analyses as well as additional experiments.
PDE Loss. To verify the role of the PDE guidance loss of Eq. 8 during the denoising process, we
visualize the errors of recovered a and u of Helmholtz equation with or without PDE loss. Here, we
run our DPS algorithm with 500 sparse observed points on both the coefficient a and solution u and
study the effect of the additional PDE loss guidance. The relative error of u reduces from 9.3% to
0.6%, and the relative error of a reduces from 13.2% to 9.4%. Therefore, we conclude that PDE
guidance helps smooth the prediction and improve the accuracy.
Number of Observations. We examine the results of DiffusionPDE in solving forward and inverse
problems when there are 100, 300, 500, and 1000 random observations on a, u, or both a and u. The
error of DiffusionPDE decreases as the number of sparse observations increases. DiffusionPDE is
capable of recovering both a and u with errors 1% ∼ 10% with approximately 6% observation points
at any side for most PDE families. DiffusionPDE becomes insensitive to the number of observations
and can solve the problems well once more than 3% of the points are observed.
Observation Sampling Pattern We show that DiffusionPDE is robust to different sampling patterns
of the sparse observations, including grid and non-uniformly concentrated patterns. Note that even
when conditioned on the full observations, our approach performs on par with the current best
methods, likely due to the inherent resilience of our guided diffusion algorithm.

5 Conclusion and Future Work


In this work, we develop DiffusionPDE, a diffusion-based PDE solver that addresses the challenge
of solving PDEs from partial observations by filling in missing information using generative priors.
We formulate a diffusion model that learns the joint distribution of the coefficient (or initial state)
space and the solution (or final state) space. During the sampling process, DiffusionPDE can flexibly
generate plausible data by guiding its denoising with sparse measurements and PDE constraints. Our
new approach leads to significant improvements over existing state-of-the-art methods, advancing
toward a general PDE-solving framework that leverages the power of generative models.
Several promising directions for future research have emerged from this work. Currently, Diffusion-
PDE is limited to solving slices of 2D dynamic PDEs; extending its capabilities to cover full time
intervals of these equations presents a significant opportunity. Moreover, the model’s struggle with
accuracy in spaces that lack constraints is another critical area for exploration. DiffusionPDE also
suffers from a slow sampling procedure, and a faster solution might be desired.

9
References
[1] Ferri MH Aliabadi. Boundary element methods. In Encyclopedia of continuum mechanics,
pages 182–193. Springer, 2020.

[2] Pavel Ŝolín. Partial differential equations and the finite element method. John Wiley & Sons,
2005.
[3] Zongyi Li, Nikola Kovachki, Kamyar Azizzadenesheli, Burigede Liu, Kaushik Bhattacharya,
Andrew Stuart, and Anima Anandkumar. Fourier neural operator for parametric partial differen-
tial equations. arXiv preprint arXiv:2010.08895, 2020.
[4] Zongyi Li, Hongkai Zheng, Nikola Kovachki, David Jin, Haoxuan Chen, Burigede Liu, Kamyar
Azizzadenesheli, and Anima Anandkumar. Physics-informed neural operator for learning partial
differential equations. ACM/JMS Journal of Data Science, 2021.
[5] Lu Lu, Pengzhan Jin, Guofei Pang, Zhongqiang Zhang, and George Em Karniadakis. Learning
nonlinear operators via deeponet based on the universal approximation theorem of operators.
Nature machine intelligence, 3(3):218–229, 2021.
[6] Maziar Raissi, Paris Perdikaris, and George E Karniadakis. Physics-informed neural networks:
A deep learning framework for solving forward and inverse problems involving nonlinear partial
differential equations. Journal of Computational physics, 378:686–707, 2019.
[7] Jonathan Ho, Ajay Jain, and Pieter Abbeel. Denoising diffusion probabilistic models. Advances
in neural information processing systems, 33:6840–6851, 2020.
[8] Qingqing Zhao, David B Lindell, and Gordon Wetzstein. Learning to solve pde-constrained
inverse problems with graph networks. arXiv preprint arXiv:2206.00711, 2022.
[9] Lawrence C Evans. Partial differential equations, volume 19. American Mathematical Society,
2022.
[10] Kenji Omori and Jun Kotera. Overview of pdes and their regulation. Circulation research,
100(3):309–327, 2007.
[11] Ryan Po, Wang Yifan, Vladislav Golyanik, Kfir Aberman, Jonathan T Barron, Amit H Bermano,
Eric Ryan Chan, Tali Dekel, Aleksander Holynski, Angjoo Kanazawa, et al. State of the art on
diffusion models for visual computing. arXiv preprint arXiv:2310.07204, 2023.
[12] Walter A Strauss. Partial differential equations: An introduction. John Wiley & Sons, 2007.
[13] Ling Yang, Zhilong Zhang, Yang Song, Shenda Hong, Runsheng Xu, Yue Zhao, Wentao Zhang,
Bin Cui, and Ming-Hsuan Yang. Diffusion models: A comprehensive survey of methods and
applications. ACM Computing Surveys, 56(4):1–39, 2023.
[14] Alfio Quarteroni and Alberto Valli. Numerical approximation of partial differential equations,
volume 23. Springer Science & Business Media, 2008.
[15] Sergio R Idelsohn, Eugenio Onate, Nestor Calvo, and Facundo Del Pin. The meshless finite
element method. International Journal for Numerical Methods in Engineering, 58(6):893–912,
2003.
[16] Maziar Raissi, Paris Perdikaris, and George Em Karniadakis. Physics informed deep learn-
ing (part i): Data-driven solutions of nonlinear partial differential equations. arXiv preprint
arXiv:1711.10561, 2017.
[17] Shengze Cai, Zhiping Mao, Zhicheng Wang, Minglang Yin, and George Em Karniadakis.
Physics-informed neural networks (pinns) for fluid mechanics: A review. Acta Mechanica
Sinica, 37(12):1727–1738, 2021.
[18] Zhiping Mao, Ameya D Jagtap, and George Em Karniadakis. Physics-informed neural networks
for high-speed flows. Computer Methods in Applied Mechanics and Engineering, 360:112789,
2020.

10
[19] Hamidreza Eivazi, Mojtaba Tahani, Philipp Schlatter, and Ricardo Vinuesa. Physics-informed
neural networks for solving reynolds-averaged navier–stokes equations. Physics of Fluids,
34(7), 2022.
[20] Shengze Cai, Zhicheng Wang, Sifan Wang, Paris Perdikaris, and George Em Karniadakis.
Physics-informed neural networks for heat transfer problems. Journal of Heat Transfer,
143(6):060801, 2021.
[21] George S Misyris, Andreas Venzke, and Spyros Chatzivasileiadis. Physics-informed neural
networks for power systems. In 2020 IEEE power & energy society general meeting (PESGM),
pages 1–5. IEEE, 2020.
[22] Zongyi Li, Nikola Kovachki, Chris Choy, Boyi Li, Jean Kossaifi, Shourya Otta, Moham-
mad Amin Nabian, Maximilian Stadler, Christian Hundt, Kamyar Azizzadenesheli, et al.
Geometry-informed neural operator for large-scale 3d pdes. Advances in Neural Information
Processing Systems, 36, 2024.
[23] Louis Serrano, Lise Le Boudec, Armand Kassaï Koupaï, Thomas X Wang, Yuan Yin, Jean-Noël
Vittaut, and Patrick Gallinari. Operator learning with neural fields: Tackling pdes on general
geometries. Advances in Neural Information Processing Systems, 36, 2024.
[24] Gege Wen, Zongyi Li, Kamyar Azizzadenesheli, Anima Anandkumar, and Sally M Benson.
U-fno—an enhanced fourier neural operator-based deep-learning model for multiphase flow.
Advances in Water Resources, 163:104180, 2022.
[25] Fanny Lehmann, Filippo Gatti, Michaël Bertin, and Didier Clouteau. Fourier neural operator
surrogate model to predict 3d seismic waves propagation. arXiv preprint arXiv:2304.10242,
2023.
[26] Bian Li, Hanchen Wang, Shihang Feng, Xiu Yang, and Youzuo Lin. Solving seismic wave
equations on variable velocity models with fourier neural operator. IEEE Transactions on
Geoscience and Remote Sensing, 61:1–18, 2023.
[27] Zhijie Li, Wenhui Peng, Zelong Yuan, and Jianchun Wang. Fourier neural operator approach
to large eddy simulation of three-dimensional turbulence. Theoretical and Applied Mechanics
Letters, 12(6):100389, 2022.
[28] Wenhui Peng, Zelong Yuan, Zhijie Li, and Jianchun Wang. Linear attention coupled fourier
neural operator for simulation of three-dimensional turbulence. Physics of Fluids, 35(1), 2023.
[29] Boris Bonev, Thorsten Kurth, Christian Hundt, Jaideep Pathak, Maximilian Baust, Karthik
Kashinath, and Anima Anandkumar. Spherical fourier neural operators: Learning stable
dynamics on the sphere. In International conference on machine learning, pages 2806–2823.
PMLR, 2023.
[30] Peter Yichen Chen, Jinxu Xiang, Dong Heon Cho, Yue Chang, GA Pershing, Henrique Teles
Maia, Maurizio M Chiaramonte, Kevin Carlberg, and Eitan Grinspun. Crom: Continu-
ous reduced-order modeling of pdes using implicit neural representations. arXiv preprint
arXiv:2206.02607, 2022.
[31] Zilu Li, Guandao Yang, Xi Deng, Christopher De Sa, Bharath Hariharan, and Steve Marschner.
Neural caches for monte carlo partial differential equation solvers. In SIGGRAPH Asia 2023
Conference Papers, pages 1–10, 2023.
[32] Thomas Müller, Fabrice Rousselle, Jan Novák, and Alexander Keller. Real-time neural radiance
caching for path tracing. arXiv preprint arXiv:2106.12372, 2021.
[33] Hong Chul Nam, Julius Berner, and Anima Anandkumar. Solving poisson equations using
neural walk-on-spheres. In ICLR 2024 Workshop on AI4DifferentialEquations In Science, 2024.
[34] M Cho, B Jadamba, R Kahler, AA Khan, and M Sama. First-order and second-order adjoint
methods for the inverse problem of identifying non-linear parameters in pdes. Industrial
Mathematics and Complex Systems: Emerging Mathematical Models, Methods and Algorithms,
pages 147–163, 2017.

11
[35] Colin Fox and Geoff Nicholls. Statistical estimation of the parameters of a pde. Can. appl.
Math. Quater, 10:277–810, 2001.
[36] Bastian Harrach. An introduction to finite element methods for inverse coefficient problems in
elliptic pdes. Jahresbericht der Deutschen Mathematiker-Vereinigung, 123(3):183–210, 2021.
[37] Krishna Kumar and Yonjin Choi. Accelerating particle and fluid simulations with differentiable
graph networks for solving forward and inverse problems. In Proceedings of the SC’23
Workshops of The International Conference on High Performance Computing, Network, Storage,
and Analysis, pages 60–65, 2023.
[38] Tristan van Leeuwen and Felix J Herrmann. A penalty method for pde-constrained optimization
in inverse problems. Inverse Problems, 32(1):015007, 2015.
[39] Da Long and Shandian Zhe. Invertible fourier neural operators for tackling both forward and
inverse problems. arXiv preprint arXiv:2402.11722, 2024.
[40] Roberto Molinaro, Yunan Yang, Björn Engquist, and Siddhartha Mishra. Neural inverse
operators for solving pde inverse problems. arXiv preprint arXiv:2301.11167, 2023.
[41] Maarten V de Hoop, Matti Lassas, and Christopher A Wong. Deep learning architectures
for nonlinear operator functions and nonlinear inverse problems. Mathematical Statistics and
Learning, 4(1):1–86, 2022.
[42] Samira Pakravan, Pouria A Mistani, Miguel A Aragon-Calvo, and Frederic Gibou. Solving
inverse-pde problems with physics-aware neural networks. Journal of Computational Physics,
440:110414, 2021.
[43] Valerii Iakovlev, Markus Heinonen, and Harri Lähdesmäki. Learning continuous-time pdes
from sparse data with graph neural networks. arXiv preprint arXiv:2006.08956, 2020.
[44] Jiaming Song, Chenlin Meng, and Stefano Ermon. Denoising diffusion implicit models. arXiv
preprint arXiv:2010.02502, 2020.
[45] Tero Karras, Miika Aittala, Timo Aila, and Samuli Laine. Elucidating the design space of
diffusion-based generative models. In Proc. NeurIPS, 2022.
[46] Hyungjin Chung, Jeongsol Kim, Michael T Mccann, Marc L Klasky, and Jong Chul Ye. Diffu-
sion posterior sampling for general noisy inverse problems. arXiv preprint arXiv:2209.14687,
2022.
[47] Hyungjin Chung, Byeongsu Sim, Dohoon Ryu, and Jong Chul Ye. Improving diffusion models
for inverse problems using manifold constraints. Advances in Neural Information Processing
Systems, 35:25683–25696, 2022.
[48] Prafulla Dhariwal and Alexander Nichol. Diffusion models beat gans on image synthesis.
Advances in neural information processing systems, 34:8780–8794, 2021.
[49] Sihan Xu, Yidong Huang, Jiayi Pan, Ziqiao Ma, and Joyce Chai. Inversion-free image editing
with natural language. arXiv preprint arXiv:2312.04965, 2023.
[50] Christian Jacobsen, Yilin Zhuang, and Karthik Duraisamy. Cocogen: Physically-consistent and
conditioned score-based generative models for forward and inverse problems. arXiv preprint
arXiv:2312.10527, 2023.
[51] Marten Lienen, David Lüdke, Jan Hansen-Palmus, and Stephan Günnemann. From zero
to turbulence: Generative modeling for 3d flow simulation. In The Twelfth International
Conference on Learning Representations, 2023.
[52] Gefan Yang and Stefan Sommer. A denoising diffusion model for fluid field prediction. arXiv
preprint arXiv:2301.11661, 2023.
[53] Phillip Lippe, Bas Veeling, Paris Perdikaris, Richard Turner, and Johannes Brandstetter. Pde-
refiner: Achieving accurate long rollouts with neural pde solvers. Advances in Neural Informa-
tion Processing Systems, 36, 2024.

12
[54] Dule Shu, Zijie Li, and Amir Barati Farimani. A physics-informed diffusion model for high-
fidelity flow field reconstruction. Journal of Computational Physics, 478:111972, 2023.
[55] Yang Song, Jascha Sohl-Dickstein, Diederik P Kingma, Abhishek Kumar, Stefano Ermon, and
Ben Poole. Score-based generative modeling through stochastic differential equations. In
International Conference on Learning Representations, 2021.
[56] Aapo Hyvärinen and Peter Dayan. Estimation of non-normalized statistical models by score
matching. Journal of Machine Learning Research, 6(4), 2005.
[57] Yuanming Hu, Luke Anderson, Tzu-Mao Li, Qi Sun, Nathan Carr, Jonathan Ragan-Kelley, and
Frédo Durand. Difftaichi: Differentiable programming for physical simulation. arXiv preprint
arXiv:1910.00935, 2019.

13
Appendix
A Overview
In this supplementary material, we provide detailed additional content that complements the main
paper. Section B elaborates on the implementation of our data generation process. Section C includes
the detailed implementation of our sampling progress, while Section D delves into the reductions in
error achieved by integrating PDE loss. Section E includes comprehensive visual results for both
forward and inverse computations using sparse observations, not featured in the main text. In Section
F, we discuss results from full observation scenarios for all methods. Section G justifies our choice
of training the baselines on complete observation data. To evaluate the stability of DiffusionPDE,
Section H includes the standard deviation of DiffusionPDE, and Section I examines the robustness of
our DiffusionPDE model against random noise or different observation locations. Section J explores
the effects of varying observation numbers on result accuracy, providing a deeper understanding of
our model’s performance under different conditions. Finally, Section K studies the performance of
DiffusionPDE across different resolutions.

B Data Generation Details


We generate 50,000 samples for each PDE and all diffusion models are trained on Nvidia A40 GPUs.

B.1 Static PDEs

We derived the methods of data generation for static PDEs from [3]. We first generate Gaussian
random fields on (0, 1)2 so that µ ∼ N (0, (−∆ + 9I)−2 ). For Darcy Flow, we let a = f (µ) so that:
if µ(x) ≥ 0

a(x) = 12,
a(x) = 3, if µ(x) < 0
For the Poisson equation and Helmholtz equation, we let a = µ as the coefficients. We then use
second-order finite difference schemes to solve the solution u and enforce the no-slip boundary
condition for solutions by multiplying a mollifier sin(πx1 ) sin(πx2 ) for point x = (x1 , x2 ) ∈ (0, 1)2 .
Both a and u have resolutions of 128 × 128.

B.2 Non-bounded Navier-Stokes Equation


We derived the method to generate non-bounded Navier-Stokes equation from [3]. The initial
condition w0 is generated by Gaussian random field N (0, 71.5 (−∆+49I)−2.5 ). The forcing function
follows the fixed pattern for point (x1 , x2 ):
1
q(x) = (sin(2π(x1 + x2 )) + cos(2π(x1 + x2 )))
10
We then use the pseudo-spectral method to solve the Navier-Stokes equations in the stream-function
formulation. We transform the equations into the spectral domain using Fourier transforms, solving
the vorticity equation in the spectral domain, and then using inverse Fourier transforms to compute
nonlinear terms in physical space. We simulate for 1 second with 10 timesteps, and wt has a resolution
of 128 × 128.

B.3 Bounded Navier-Stokes Equation

We use Difftaichi [57] to generate data for the bounded Navier-Stokes equation. Specifically,
we apply the Marker-and-Cell (MAC) method by solving a pressure-Poisson equation to enforce
incompressibility and iterating through predictor and corrector steps to update the velocity and
pressure fields. The grid is of the resolution 128 × 128 and the center of the cylinder is at a random
location in [30, 60] × [30, 90] with a random radius in [5, 20]. The fluid flows into the grid from the
upper boundary with a random initial vertical velocity in [0.5, 3]. We simulate for 1 second with 10
timesteps and study steps 4 to 8 when the turbulence is passing the cylinder.

B.4 Burgers’ Equation

We derived the method to generate Burgers’ equation from [3]. The initial condition u0 is generated
by Gaussian random field N (0, 625(−∆ + 25I)−2 ). We solve the PDE with a spectral method and
simulate 1 second with 127 additional timesteps. The final u0:T space has a resolution of 128 × 128.

14
C Guided Sampling Details
For experiments with sparse observations or sensors, we find that DiffusionPDE performs the best
when weights ζ are selected as shown in Table 2. During the initial 80% of iterations in the sampling
process, guidance is exclusively provided by the observation loss Lobs . Subsequently, after 80% of
the iterations have been completed, we introduce the PDE loss Lpde , and reduce the weighting factor
ζobs for the observation loss, by a factor of 10. This adjustment shifts the primary guiding influence
to the PDE loss, thereby aligning the diffusion model more closely with the dynamics governed by
the partial differential equations.

Table 2: The weights assigned to the PDE loss and the observation loss vary depending on whether
the observations pertain to the coefficients (or initial states) a or to the solutions (or final states) u.
Darcy Flow Poisson Helmholtz Non-bounded Bounded Burgers’
Navier-Stokes Navier-Stokes equation
a 2.5 × 103 4 × 102 2 × 102 5 × 102 2.5 × 102 3.2 × 102
ζobs
u 106 2 × 104 3 × 104 5 × 102 2.5 × 102 -
ζpde 103 102 102 102 102 102

D Improvement in Prediction through PDE Loss Term


DiffusionPDE performs better when we apply the PDE loss term Lpde in addition to the observation
loss term Lobs as guidance, as shown in Table 3. The errors in both the coefficients ( initial states) a
and the solutions (final states) u significantly decrease. We also visualize the recovered a and u and
corresponding absolute errors of Darcy Flow, Poisson equation, and Helmholtz equation in Fig. 6. It
is demonstrated that the prediction becomes more accurate with the combined guidance of PDE loss
and observation loss than with only observation loss.

Table 3: DiffusionPDE’ prediction errors of coefficients (initial states) a and solutions (final states) u
with sparse observation on both a and u, guided by different loss functions.
Loss Function Side Darcy Flow Poisson Helmholtz Non-bounded Bounded
Navier-Stokes Navier-Stokes
a 4.6% 12.1% 13.2% 8.2% 6.4%
Lobs
u 4.8% 6.5% 9.3% 7.6% 3.3%
a 3.4% 10.3% 9.4% 4.9% 1.7%
Lobs + Lpde
u 1.7% 0.3% 0.6% 0.6% 1.4%

Guided with
 Guided with PDE Loss
 Guided with
 Guided with PDE Loss

Observation Loss and Observation Loss Ground Truth Observation Loss and Observation Loss
Coefficient Absolute Error
Recovered Coefficients

Error Rate: 3.3% Error Rate: 2.6%


Solution Absolute Error
Recovered Solutions

Relative Error: 5.1% Relative Error: 0.9%

(a) Recovered coefficients, solutions, and corresponding absolute errors of Darcy Flow.

15
Guided with
 Guided with PDE Loss
 Guided with
 Guided with PDE Loss

Observation Loss and Observation Loss Ground Truth Observation Loss and Observation Loss

Coefficient Absolute Error


Recovered Coefficients

Relative Error: 12.1% Relative Error: 10.3%

Solution Absolute Error


Recovered Solutions

Relative Error: 6.5% Relative Error: 0.3%

(b) Recovered coefficients, solutions, and corresponding absolute errors of Poisson equation.
Guided with
 Guided with PDE Loss
 Guided with
 Guided with PDE Loss

Observation Loss and Observation Loss Ground Truth Observation Loss and Observation Loss

Coefficient Absolute Error


Recovered Coefficients

Relative Error: 13.2% Relative Error: 9.4%


Solution Absolute Error
Recovered Solutions

Relative Error: 9.3% Relative Error: 0.6%

(c) Recovered coefficients, solutions, and corresponding absolute errors of Helmholtz equation.
Figure 6: Recovered coefficients, solutions, and their corresponding visualized absolute errors for
various PDE families.

E Additional Results on All PDEs with Sparse Observation


We present the recovered results of another Burgers’ equation in Fig. 7. DiffusionPDE outperforms
all other methods with 5 sensors for continuous observation. We also present the recovered results
for both the forward and inverse problems of all other PDEs with sparse observations, as shown in
Fig. 8. Specifically, we solve the forward and inverse problems for the Darcy Flow, Poisson equation,
Helmholtz equation, and non-bounded Navier-Stokes equation using 500 random points observed
in either the solution space or the coefficient space. Additionally, for the bounded Navier-Stokes
equation, we observe 1% of the points in the velocity field. Our findings indicate that DiffusionPDE
outperforms all other methods, providing the most accurate solutions.

Additional Data Setting for Darcy Flow To further demonstrate the generalization capability of
our model, we conducted additional tests on different data settings for Darcy Flow. In Fig. 9, we
solve the forward and inverse problems of Darcy Flow with 500 observation points, adjusting the
binary values of a to 20 and 16 instead of the original 12 and 3 in Section B, i.e.,
if µ(x) ≥ 0

a(x) = 20,
a(x) = 16, if µ(x) < 0
Our results indicate that DiffusionPDE performs equally well under these varied data settings,
showcasing its robustness and adaptability.

16
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs

5 Random Sensors

Time

x Relative Error: 5.1% Relative Error: 78.2% Relative Error: 96.2% Relative Error: 96.3% Relative Error: 81.0%
Figure 7: Results of another Burgers’ equation recovered by 5 sensors throughout the time interval.
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs
Coefficients

Forward
y Solutions

x 500 Random Points Relative Error: 2.5% Relative Error: 38.3% Relative Error: 35.2% Relative Error: 28.2% Relative Error: 48.8%
Coefficients

Inverse
Solutions

x Error Rate: 3.2% Error Rate: 41.1% Error Rate: 49.2% Error Rate: 49.3% Error Rate: 59.7%
(a) Forward and inverse results of Darcy Flow recovered by 500 observation points.
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs
Coefficients

Forward
Solutions

x 500 Random Points Relative Error: 4.5% Relative Error: 155.5% Relative Error: 107.1% Relative Error: 100.9% Relative Error: 128.1%
Coefficients

Inverse
Solutions

x Relative Error: 20.0% Relative Error: 105.8%Relative Error: 231.9% Relative Error: 232.7% Relative Error: 130.0%
(b) Forward and inverse results of Poisson equation recovered by 500 observation points.
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs
Coefficients

Forward
Solutions

x 500 Random Points Relative Error: 8.8% Relative Error: 123.1% Relative Error: 106.5% Relative Error: 98.2% Relative Error: 142.3%
Coefficients

Inverse
Solutions

x Relative Error: 22.6% Relative Error: 132.8% Relative Error: 216.9% Relative Error: 218.2% Relative Error: 160.0%
(c) Forward and inverse results of Helmholtz equation recovered by 500 observation points.
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs
Initial State

Forward
Final State

x 500 Random Points Relative Error: 5.7% Relative Error: 100.1% Relative Error: 99.5% Relative Error: 99.1% Relative Error: 141.1%

Inverse
Initial State
Final State

x Relative Error: 9.5% Relative Error: 96.9% Relative Error: 98.2% Relative Error: 98.1% Relative Error: 150.7%
(d) Forward and inverse results of another non-bounded Navier-Stokes equation recovered by 500 observation
points.

17
Observations Ground Truth DiffusionPDE DeepONet PINO FNO PINNs

Initial State
Forward

Final State
y

x 1% Random Points Relative Error: 3.9% Relative Error: 97.7% Relative Error: 81.1% Relative Error: 82.8% Relative Error: 100.1%

Inverse

Initial State
Final State

x Relative Error: 2.7% Relative Error: 91.9% Relative Error: 69.5% Relative Error: 69.6% Relative Error: 105.5%
(e) Forward and inverse results of bounded Navier-Stokes equation recovered by 1% observation points.
Ground Truth Observed Final State Cylinder Boundary
GraphPDE DiffusionPDE Ground Truth

Initail State
Inverse
Final State

1% Random Points Relative Error: 10.5% Relative Error: 5.6%

(f) Inverse results of DiffusionPDE and GraphPDE of another bounded Navier-Stokes equation recovered by 1%
observation points and the known boundary of the cylinder.
Figure 8: Results of forward and inverse problems for different PDE families with sparse observation.

Observed Coefficients Recovered Solutions Ground Truth Observed Solutions Recovered Solutions Ground Truth

Forward Inverse

500 Random Points Relative Error: 2.0% 500 Random Points Error Rate: 4.9%
Figure 9: Forward and inverse results of Darcy Flow recovered by 500 observation points under a
different data setting.

F Solving Forward and Inverse Problems with Full Observation


We have also included the errors of all methods when solving both the forward and inverse problems
with full observation, as displayed in Table 4.
Table 4: Relative errors of solutions (or final states) and coefficients (or initial states) when solving
forward and inverse problems with full observations. Error rates are used for the inverse problem of
Darcy Flow.
DiffusionPDE PINO DeepONet PINNs FNO
Forward 2.2% 4.0% 12.3% 15.4% 5.3%
Darcy Flow
Inverse 2.0% 2.1% 8.4% 10.1% 5.6%
Forward 2.7% 3.7% 14.3% 16.1% 8.2%
Poisson
Inverse 9.8% 10.2% 29.0% 33.1% 13.6%
Forward 2.3% 4.9% 17.8% 18.1% 11.1%
Helmholtz
Inverse 4.0% 4.9% 28.1% 29.2% 5.0%
Non-bounded Forward 6.1% 1.1% 25.6% 27.3% 2.3%
Navier-Stokes Inverse 8.6% 6.8% 19.6% 27.8% 6.8%
Bounded Forward 1.7% 1.9% 13.3% 18.6% 2.0%
Navier-Stokes Inverse 1.4% 2.9% 6.1% 7.6% 3.0%

18
In general, DiffusionPDE and PINO outperform all other methods, and DiffusionPDE performs the
best for all static PDEs. DiffusionPDE is capable of solving both forward and inverse problems with
errors of less than 10% for all classes of discussed PDEs and is comparable to the state-of-the-art.

G Training Baselines Methods on Partial Inputs


For our main experiments, we opt to train the baseline models (PINO, DeepONet, PINNs, FNO) on
full observations for several compelling reasons: First, physics-informed models such as PINNs and
PINO are unable to effectively compute the PDE loss when only sparse observations are available.
Second, other models like DeepONet and FNO perform poorly with sparse observations. For instance,
training the DeepONet model on 500 uniformly random points for each training sample in the context
of the forward problem of Darcy Flow leads to testing outcomes that are consistently similar, as
illustrated in Fig. 10, regardless of the testing input. This pattern suggests that the model tends to
generate a generalized solution that minimizes the average error across all potential solutions rather
than converging based on specific samples. Furthermore, the partial-input-trained model exhibits
poor generalization when faced with a different distribution of observations from training, indicating
that it lacks flexibility—a critical attribute of our DiffusionPDE.
500 Observed Points 1000 Observed Points Full Observation

Figure 10: Predicted solutions obtained using the DeepONet model trained with 500 observation
points across different numbers of observation points.

H Standard Deviation of DiffusionPDE Experiment Results


We further assess the statistical significance of our DiffusionPDE by analyzing the standard devia-
tions for forward and inverse problems under conditions of 500 sparse observation points and full
observation, respectively, as detailed in Table 5. We evaluate our model using test sets comprising
1,000 samples for each PDE. Our findings confirm that full observation enhances the stability of
the results, a predictable outcome as variability diminishes with an increase in observation points.
The standard deviations are notably higher for more complex PDEs, such as the inverse problems
of the Poisson and Helmholtz equations, reflecting the inherent challenges associated with these
computations. Overall, DiffusionPDE demonstrates considerable stability, evidenced by relatively
low standard deviations across various tests.
Table 5: Standard deviation of DiffusionPDE when solving forward and inverse problems with sparse
or full observations.
Sparse Observations Full Observations
Forward 2.5 ± 0.7% 2.2 ± 0.1%
Darcy Flow
Inverse 3.2 ± 0.9% 2.0 ± 0.1%
Forward 4.5 ± 0.9% 2.7 ± 0.1%
Poisson
Inverse 20.0 ± 1.8% 9.8 ± 0.7%
Forward 8.8 ± 1.0% 2.3 ± 0.1%
Helmholtz
Inverse 22.6 ± 1.7% 4.0 ± 0.6%
Non-bounded Forward 6.9 ± 0.9% 6.1 ± 0.2%
Navier-Stokes Inverse 10.4 ± 1.0% 8.6 ± 0.3%
Bounded Forward 3.9 ± 0.2% 1.7 ± 0.1%
Navier-Stokes Inverse 2.7 ± 0.2% 1.4 ± 0.1%

19
Figure 11: Relative errors of recovered Darcy Flow solutions with sparse noisy observation.

I Robustness of DiffusionPDE
We find that DiffusionPDE is robust against sparse noisy observation. In Fig. 11, we add Gaussian
noise to the 500 observed points of Darcy Flow coefficients. Our DiffusionPDE can maintain a relative
error of around 10% with a 15% noise level concerning the forward problem, and the recovered
solutions are shown in Fig. 12. Baseline methods such as PINO also exhibit robustness against
random noise under sparse observation conditions; this is attributed to their limited applicability to
sparse observation problems, leading them to address the problem in a more randomized manner.

Clean Observations No Noise 1% Noise 3% Noise 5% Noise 10% Noise 15% Noise
Recovered Solutions
Coefficients

Forward

Add Noise

Relative Error: 2.5% Relative Error: 6.3% Relative Error: 7.8% Relative Error: 7.4% Relative Error: 7.6% Relative Error: 10.1%

Figure 12: Recovered solutions for Darcy Flow with noisy observations.

Robustness on Sampling Patterns Moreover, as mentioned in the main document, we investigate


the robustness of DiffusionPDE on different sampling patterns of the observation points. Here,
we address the forward problem of Darcy Flow using 500 observed coefficient points, which are
non-uniformly concentrated on the left and right sides or are regularly distributed across the grid, as
depicted in Fig. 13. Our results demonstrate that DiffusionPDE flexibly solves problems with arbitrary
sparse observation locations within the spatial domain, without re-training the neural network model.

Observed Coefficients Recovered Solutions Observed Coefficients Recovered Solutions Observed Coefficients Recovered Solutions

Forward Forward Forward

Relative Error: 6.2% Relative Error: 8.3% Relative Error: 2.3%


Figure 13: Recovered solutions for Darcy Flow with observations sampled using non-uniform
distributions.

20
J Solving Forward and Inverse Problems with Different Numbers of
Observations

We also investigate how our DiffusionPDE handles varying degrees of sparse observation. Experi-
ments are conducted on the Darcy Flow, Poisson equation, Helmholtz equation, and non-bounded
Navier-Stokes equation. We examine the results of DiffusionPDE in solving forward and inverse
problems when there are 100, 300, 500, and 1000 random observations on a, u, or both a and u,
as shown in Fig. 14. We have observed that the error of DiffusionPDE decreases as the number
of sparse observations increases. Overall, we recover u better than a. DiffusionPDE can recover u
with approximately 2% observation points at any side pretty well. DiffusionPDE is also capable of
recovering both a and u with errors 1% ∼ 10% with approximately 6% observation points at any
side for most PDE families. We also conclude that our DiffusionPDE becomes insensitive to the
number of observations once more than 3% of the points are observed.

(a) Error rates for Darcy Flow and relative errors for other PDEs of recovered coefficients or initial states a.

(b) Relative errors of recovered solutions or final states u.


Figure 14: Error rate or relative error of both coefficients (or initial states) a and solutions (or final
states) u with different numbers of observations.

21
K Solving Forward and Inverse Problems across Varied Resolutions
To evaluate the generalizability of DiffusionPDE, we implemented the model on various resolutions,
including 64 × 64 and 256 × 256, while maintaining the same percentage of observed points. For
resolutions of 64 × 64, 128 × 128, and 256 × 256, we observe 125, 500, and 2000 points on a or u
respectively, which are approximately 3% for each resolution. Overall, DiffusionPDE is capable of
handling different resolutions effectively. For instance, Table 6 presents the forward relative errors
of the solution u and inverse error rates of the coefficient a for the Darcy Flow, demonstrating that
DiffusionPDE performs consistently well with similar error rates across various resolutions.

Table 6: Forward relative errors and inverse error rates of Darcy Flow across different resolutions.
Resolution Forward Relative Error Inverse Error Rate
64 × 64 2.9% 4.3%
128 × 128 2.5% 3.2%
256 × 256 3.1% 4.1%

22

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy