Conference PaperPDF Available

Projection-based multifidelity linear regression for data-poor applications

Authors:
Projection-based multifidelity linear regression for data-poor
applications
Vignesh Sella, Julie Pham, Anirban Chaudhuri, Karen Willcox §
University of Texas at Austin, TX 78712, USA
Surrogate modeling for systems with high-dimensional quantities of interest is important
for many applications in science and engineering, but remains a challenge in situations where
training data are expensive to acquire. This work develops multifidelity (MF) approaches for
multivariate multi-output linear regression for data-poor applications with high-dimensional
outputs. The MF approach combines information from many low-cost, low-fidelity (LF) model
evaluations with limited expensive, high-fidelity (HF) model evaluations. We implement and
contrast three MF linear regression methods with projections to a lower-dimensional space
through proper orthogonal decomposition basis vectors. The three MF linear regression
approaches developed in this work are: (i) an additive method based on the Kennedy-O’Hagan
framework, (ii) a direct data augmentation using LF data, and (iii) a data augmentation method
using explicit linear regression mapping between LF and HF data. The data augmentation
technique combines the HF and the LF data into one training data set and the linear regression
model is trained using weighted least squares with different weights for the different fidelity
models. We apply the projection-enabled MF linear regression methods to approximate the
surface pressure field on a hypersonic vehicle in flight. The MF linear regression outperforms the
single-fidelity linear regression in the low data regime of 3
10 HF samples with an improvement
in the range of approximately 312% in median accuracy for similar computational cost.
I. Introduction
An important challenge in scientific machine learning research is to develop methods that can exploit and maximize
the amount of learning possible from scarce data [
1
]. The need for such research arises often in science and engineering,
and especially in the case of computational fluid dynamics (CFD), since expensive-to-evaluate high-fidelity (HF)
models make many-query problems such as uncertainty quantification, risk analysis, optimization, and optimization
under uncertainty computationally prohibitive [
2
]. Thus, there is a need for surrogate models that can approximate the
solutions to HF models to accelerate the design and analysis process. However, lack of sufficient HF data adversely
affects the accuracy of surrogate models. We propose multifidelity linear regression methods that can use multiple
cheaper lower-fidelity (LF) information sources along with the limited HF data for training linear regression models.
Linear regression [
3
6
], including polynomial regression or the response surface methodology (RSM) [
7
], is a
regression analysis technique that has been extensively used for surrogate modelling and prediction in aerospace
applications. Madsen et al. [
8
] utilized RSM using polynomials for diffuser shape optimization with up to five input
design variables. Nakamura et al. [
9
] applied linear regression methods to predict fluid-flow over two-dimensional
cylinders and the state-estimation from wall characteristics in turbulent flow using training samples on the order of
𝑂(
10
3)
. Another line of work seeks to use low-order polynomial RSM approximations as a tool within multidisciplinary
design optimization frameworks in order to deal with data plagued by high computational cost and noise in aerospace
design applications [
10
,
11
]. Recently, there has also been a growing wealth of literature on using even more data to
train and create surrogate models. Ladický et al. [
12
] investigated the feasibility of random forests to approximate fluid
particle behaviour in time and compared it to the state-of-the-art position based fluid approach using
𝑂(
10
9)
training
samples. In many of these studies, the amount of training data procured to train the respective surrogate models was
a computationally intensive effort. While these levels of data procuration are possible, they are not always plausible
without access to significant computing resources. In this work, we propose multifidelity versions of linear regression
that can alleviate the issue of a lack of data.
Graduate Research Assistant, Oden Institute of Computational Engineering and Sciences, AIAA Student Member
Graduate Research Assistant, Deptartment of Aerospace Engineering and Engineering Mechanics, AIAA Student Member
Research Associate, Oden Institute of Computational Engineering and Sciences, AIAA Member
§Director, Oden Institute of Computational Engineering and Sciences, AIAA Fellow
1
Multifidelity (MF) linear regression methods that can fuse information from cheaper LF data with the sparse HF
data have been explored to address the prohibitive data requirement for regression in many applications. A MF linear
regression technique was proposed by Balabanov et al. [
13
] that used a quadratic RSM created using many coarse finite
element structural model simulations and a correction RSM using a small number of finer mesh finite element model
simulations for civilian transport wing design. Along a similar vein, the work in [
14
17
] on surrogate modelling in
computational fluid dynamics and structural mechanics for low-dimensional problems found MF linear regression to to
be more efficient and have a higher accuracy than a single-fidelity approach given the same computational cost. There
has also been development on MF surrogate modelling using artificial neural networks [
18
,
19
] and using Gaussian
process regression [
20
,
21
]. The focus of this paper is on linear regression that can tackle problems in the ultra low-data
regime and with high-dimensional outputs.
In this work, we develop and compare three MF linear regression methods: (i) additive structure using the
Kennedy-O’Hagan [
22
] approach, (ii) data augmentation of LF data directly, and (iii) data augmentation via an explicit
mapping between LF outputs and HF outputs. The MF regression algorithms can be generalized to any other choice
of underlying regression technique. However, we implement the MF regression algorithms using multi-output linear
(or polynomial) regression since we are dealing with a limited number of HF samples. For problems involving
high-dimensional quantities of interest, we show how the MF linear regression methods combine with data-driven
projection techniques to improve the performance and accuracy. We apply the projection-enabled MF linear regression
techniques to predict the pressure loads over the surface of a hypersonic vehicle.
The remainder of this paper is organized as follows. Section II describes the MF regression problem setup. Section
III describes the different MF linear regression methods developed in this work. Section IV describes the hypersonic
vehicle application and provides a detailed analysis of the performance of the algorithms. Finally, Section V concludes
the paper with remarks on the efficacy and performance of the proposed methods.
II. Problem Setup
This paper considers linear regression problems wherein the
𝑑
inputs to a system are
𝒙 X R𝑑
defined on the
input space
X
, and the output quantity of interest is
𝑚
-dimensional
𝒚 Y R𝑚
defined on the output space
Y
. In
our target applications,
𝒚
is a high-dimensional quantity with
𝑚
in thousands. For a single-fidelity case, a set of
𝑁
input-output samples
(𝑿,𝒀)
, where
𝑿=[𝒙1, ..., 𝒙𝑁] R𝑑×𝑁
and
𝒀=[𝒚1, ..., 𝒚𝑁] R𝑚×𝑁
, are available for training
the surrogate model through linear regression using ordinary least squares (OLS). In many aerospace applications,
only sparse data is available for training these surrogates due to the high computational cost associated with the HF
model evaluations. The goal of this work is to develop a multifidelity linear regression method that has the potential to
build accurate surrogates in a data-poor regime using linear regression as the underlying regression method. Given
input-output data of varying fidelity levels, multifidelity linear regression can efficiently approximate the HF quantity of
interest by incorporating information from the cheaper LF models. The training data consists of
𝑁HF
samples from the
HF model
(𝑿HF,𝒀HF )
and
𝑁LF
samples from the LF model
(𝑿LF,𝒀LF )
. In this work, we will be considering bifidelity
problems, wherein there are only two models with different levels of fidelity, but the general idea can be extended to
more than two fidelity levels.
III. Multifidelity Linear Regression
In this section, we first discuss how we implement dimensionality reduction to efficiently approximate output
quantities III.A. We then describe two overarching multifidelity linear regression methods, both of which incorporate
dimensionality reduction for high-dimensional outputs: an additive method in Section III.B and data augmentation via
synthetic data generation in Section III.C. Furthermore, all the methods presented here work for the general case of
non-collocated samples, i.e. the 𝒙LF and 𝒙HF in each input training set do not need to be at the same locations.
A. Output Dimensionality Reduction Using Proper Orthogonal Decomposition
This work targets applications with high-dimensional output quantity of interest
𝒚
and with limited training data
for the surrogate. We use dimensionality reduction techniques such as proper orthogonal decomposition (POD) to
reduce output dimensions and train a compact surrogate model with limited data. Similar projection-enabled techniques
have been used for parameteric reduced order models [
23
25
] and for neural networks [
26
]. In this work, the POD
basis vectors are obtained by taking the singular value decomposition (SVD) of the training data matrix consisting of
𝑁
samples
𝒀R𝑚×𝑁
centered by the sample average mean of the training data
𝒀
. In our target applications,
𝑚
is in
2
thousands and the number of samples 𝑁𝑚. The thin SVD of the centered training data matrix is
𝒀𝒀=𝑼𝚺𝑽,(1)
where
𝑼R𝑚×𝑁
and
𝑽R𝑁×𝑁
are orthogonal matrices, and
𝚺R𝑁×𝑁
is a square diagonal matrix consisting of
the singular values,
𝜎
, of the centered training data matrix. The reduced basis for projection to a lower-dimensional
subspace of size
𝑘𝑚
(and
𝑘𝑁
)
𝑼𝑘R𝑚×𝑘
consists of the first
𝑘
columns of the left singular vectors
𝑼
. The
projection of an output sample
𝒚
on the low-dimensional subspace is given by the reduced state
𝒄=𝑼
𝑘𝒚
. The dimension
𝑘
is chosen such that the cumulative energy captured by the first
𝑘
POD modes is larger than a specified tolerance of
𝜖
as given by
Í𝑘
𝑖=1𝜎2
𝑖
Í𝑁
𝑖=1𝜎2
𝑖
> 𝜖, (2)
where 𝜎𝑖is the 𝑖-th singular value.
B. Multifidelity Linear Regression Using an Additive Structure
In this method, we use the Kennedy-O’Hagan approach [
22
] to build the additive multifidelity linear regression.
This method assumes that the relationship between the LF and the HF data can be well modeled linearly. We begin
by training the LF surrogate model,
𝒇LF
, that fits the reduced state of the LF outputs
(𝑼LF
𝑘)(𝒀LF 𝒀LF)
using OLS,
where
𝑼LF
𝑘
is the LF reduced basis and
𝒀LF
is the mean of the LF training data (see Eqn.
(1)
). Then, we use the LF
surrogate to predict data at the same input locations as the HF data. We reconstruct the LF surrogate predictions in
the full-dimensional space as
(𝑼LF
𝑘𝒇LF (𝑿HF) + 𝒀LF)
. We use the predicted data at the same locations to estimate the
discrepancy as
𝜹(𝑿HF)=𝒀HF (𝑼LF
𝑘𝒇LF (𝑿HF) + 𝒀LF ).(3)
We fit a linear regression for discrepancy,
𝒇𝛿
, to the projected states of the discrepancy data via similar steps as for
𝒇LF
.
We can then use the discrepancy surrogate along with the LF surrogate to construct our MF linear regression surrogate as
𝒇MF (𝒙)=𝑼𝛿
𝑘𝒇𝛿(𝒙) + 𝜹+𝑼LF
𝑘𝒇LF (𝒙) + 𝒀LF (4)
where
𝑼𝛿
𝑘
is reduced basis obtained from the SVD on the discrepancy data and
𝜹
is the mean of the discrepancy data.
We summarize the additive multifidelity linear regression approach in Alg. 1.
Algorithm 1 Multifidelity linear regression via additive method (with output dimensionality reduction)
Input: HF and LF training data (𝑿LF,𝒀LF )and (𝑿HF ,𝒀HF), new input locations for prediction 𝑿
Output: Output predictions b
𝒀MF at inputs 𝑿from MF surrogate
1: Use 𝑼LF
𝑘from the SVD of 𝒀LF (Eqn. 1) to obtain the reduced states 𝑪LF R𝑘×𝑁LF :
𝑪LF =𝑼LF
𝑘𝒀LF 𝒀LF
2: Train LF linear regression surrogate model 𝒇LF on (𝑿LF,𝑪LF )using OLS linear regression
3: Predict and reconstruct LF outputs at the HF input locations (𝑼LF
𝑘𝒇LF (𝑿HF) + 𝒀LF )
4: Estimate discrepancy data 𝜹using Eqn. (3)
5: Use 𝑼𝛿
𝑘from the SVD of 𝜹to project the discrepancy to the reduced state: 𝑪𝛿=(𝑼𝛿
𝑘)(𝜹𝜹)
6: Train discrepancy linear regression 𝒇𝛿on (𝑿HF,𝑪𝛿) using OLS linear regression
7: Set 𝒇MF (𝑿HF)as the linear combination of 𝒇𝛿(𝑿HF),𝒇LF (𝑿HF ), and the known bias terms using Eqn. (4)
8: Predict outputs b
𝒀MF at new input locations 𝑿:
b
𝒀MF =𝒇MF (𝑿)
3
C. Multifidelity Linear Regression Through Data Augmentation
In the data augmentation method, we generate synthetic data from LF model evaluations and augment the HF training
set with the synthetic data. Synthetic data in this context refers to data which comes from a different source or fidelity
level than the HF model. We describe two approaches for obtaining synthetic data in Section III.C.1 and the general
data augmentation method for multifidelity linear regression using the generated synthetic data in Section III.C.2.
1. Synthetic data for data augmentation
We explore two ways of using the LF training data to create the synthetic data: (i) augmenting the LF training data
directly, and (ii) using an explicit linear regression mapping of LF data to HF data. In the first approach, the LF training
data is directly fed in as synthetic data to be augmented to the HF training data as
(𝑿syn
LF ,𝒀syn
LF )=(𝑿LF,𝒀LF )
. There is
no need to build a LF linear regression to implement the direct data augmentation. In the second approach, an explicit
map between the LF and HF data is created. To create the explicit map, we need HF and LF data at the same input
locations. However, in the general case, we do not assume that LF and HF samples are available at the same locations.
Thus, we construct the LF surrogate model
𝒇LF
and use this to predict LF surrogate outputs at the HF input locations
similar to the process shown for the additive approach in Steps 1-3 in Alg. 1. Note that if LF and HF data were available
at the same locations one can skip the initial step to create a LF surrogate model and directly train the explicit map on
the collocated data. Then, we train the explicit linear regression map
𝒇LF↦→HF
from reduced LF outputs to reduced HF
outputs using data at the same location
(b
𝑪LF,𝑪HF )
as detailed in Alg. 2. After constructing the explicit map, we can
generate the synthetic data at all the LF input locations as
𝒀syn
LF =𝑼HF
𝑘𝒇LF↦→HF (𝑪LF) + 𝒀HF
to obtain the training set
(𝑿syn
LF =𝑿LF,𝒀syn
LF )
. We summarize this process in Alg. 2. Once the HF training data is augmented with the synthetic
data, the multifidelity linear regression is trained through the process described in the next section.
Algorithm 2 Synthetic data generation via an explicit mapping using linear regression
Input: HF and LF training data (𝑿LF,𝒀LF )and (𝑿HF ,𝒀HF)
Output: Synthetic data 𝒀syn
LF at inputs 𝑿LF from the LF to HF surrogate map
1: Reduce 𝒀HF using 𝑼HF
𝑘from the SVD of 𝒀HF to obtain the reduced states 𝑪HF R𝑘×𝑁HF :
𝑪HF =𝑼HF
𝑘𝒀HF 𝒀HF
2:
Obtain
(𝑼LF
𝑘𝒇LF (𝑿HF) + 𝒀LF)
from Alg. 1 (steps 1-3) and project it to obtain the reduced states for explicit map
training:
b
𝑪LF =𝑼HF
𝑘𝑼LF
𝑘𝒇LF (𝑿HF) + 𝒀LF 𝒀HF
3: Train LF ↦→ HF linear regression model 𝒇LF↦→HF on (b
𝑪LF,𝑪HF )using OLS linear regression
4: Use 𝑼LF
𝑘from the SVD of 𝒀LF to project all the LF data to the reduced state 𝑪LF
5: Generate synthetic data 𝒀syn
LF at 𝑿LF locations and reconstruct the output in the full-dimensional space:
𝒀syn
LF =𝑼HF
𝑘𝒇LF↦→HF (𝑪LF)+𝒀HF
2. Data augmentation with weighted least squares
The data augmentation method augments the HF training dataset with the synthetic data generated from the LF
training data and performs weighted least squares (WLS) linear regression to train the MF surrogate model. In this work,
we use the two approaches described in Section III.C.1 for generating the synthetic data,
(𝑿syn
LF ,𝒀syn
LF )
. We then augment
our HF training dataset with the transformed LF synthetic data to get the training data of size
𝑁HF +𝑁LF
for the MF linear
regression as
([ 𝑿HF ,𝑿syn
LF ],[𝒀HF,𝒀syn
LF ])
where typically,
𝑁LF > 𝑁HF
. While we can train our MF surrogate model
directly with this augmented dataset, we know that the synthetic data is lower-fidelity. OLS has an underlying assumption
of homoscedasticity, or constant variance in the errors. Since we know that
𝒀syn
LF
is inherently a LF approximation to the
true HF quantity, we use different sample weights to account for the expected heteroscedasticity when training on data
from different sources. The sample weight matrix
𝑾=diag(𝑤1, . . . , 𝑤 𝑁HF+𝑁LF )
is used in the MF linear regression
training through WLS [
5
]. For all HF samples we use the sample weight
𝑤𝑖=
1
, 𝑖 =
1
, . . . , 𝑁HF
and for all synthetic
4
data samples generated from LF data we use equal sample weights of
𝑤𝑖=𝑤syn <
1
, 𝑖 =𝑁HF +
1
, . . . , 𝑁HF +𝑁LF
.
Furthermore, in the results section we note the impact of various sample weighting schemes on the numerical examples
we have applied MF regression to. We summarize the data augmentation method for MF linear regression in Alg. 3. As
we show in the results, using the direct LF data approach is more sensitive to the choice of weight parameter
𝑤syn
in
WLS as compared to the explicit linear regression map approach. One way to further improve the performance for both
the approaches would be select an optimal hyperparameter
𝑤syn
by minimizing the leave-one-out cross-validation error.
Algorithm 3 Multifidelity linear regression via data augmentation
Input: HF and LF training data
(𝑿LF,𝒀LF )
and
(𝑿HF,𝒀HF )
, synthetic sample weight
𝑤syn
, new input locations
for prediction 𝑿
Output: Output predictions b
𝒀MF at inputs 𝑿from MF surrogate
1: Generate synthetic data by transforming the LF data: (𝑿LF,𝒀LF) ↦→ ( 𝑿syn
LF ,𝒀syn
LF )using Alg. 2
2: Augment the training dataset to contain 𝑁HF +𝑁LF samples: ([ 𝑿HF ,𝑿syn
LF ],[𝒀HF,𝒀syn
LF ])
3:
Use
𝑼HF
𝑘
from the SVD of
𝒀HF
(Eqn. 1) to obtain the reduced states of MF training data outputs
𝑪MF R𝑘×( 𝑁HF +𝑁LF)
:
𝑪MF =𝑼HF
𝑘[𝒀HF,𝒀syn
LF ] 𝒀HF
4: Set up sample weight matrix 𝑾using 𝑤𝑖=1, 𝑖 =1, . . . , 𝑁HF and 𝑤𝑗=𝑤syn, 𝑗 =𝑁HF +1, . . . , 𝑁HF +𝑁LF
5:
Train MF linear regression surrogate model
𝒇MF
on
([ 𝑿HF ,𝑿syn
LF ],𝑪MF)
with sample weights
𝑾
using WLS linear
regression
6: Predict b
𝒀MF at new input locations 𝑿by reconstructing the output of 𝒇MF (𝑿)in the full-dimensional space:
b
𝒀MF =𝑼HF
𝑘𝒇MF (𝑿) + 𝒀HF
IV. Multifidelity Linear Regression for Adapted IC3X Testbed Problem
In the following sections, we present the results for a testbed problem in the computational fluid dynamics domain
described in Section IV.A. The HF and the LF model used for the MF linear regression are described in Section IV.B.
Then, we present results for the projection-enabled MF linear regression techniques proposed in this work in Section IV.C.
A. IC3X Problem Description
The hypersonic vehicle considered in this work is the Initial Concept 3.X (IC3X). The IC3X was initially proposed
by Pasiliao et al. [
27
], and a finite element model for the vehicle was developed by Witeof et al. [
28
]. The distributed
pressure load over the surface of a hypersonic vehicle, which is a primary quantity of interest, varies as a function of
Mach number, angle of attack, and sideslip angle. Based on a nominal mission trajectory for this geometry, we consider
the interval of Mach numbers
𝑀 [
5
,
7
]
, angles of attack
𝛼 [
0
,
8
]
, and sideslip angles
𝛽 [
0
,
8
]
. The pressure
field is simulated by solving the inviscid Euler equations via a Cartesian volume mesh using the flow solver package
Cart3D [
29
31
]. The pressure field solution computed by Cart3D is a vector of dimension
𝑚=
55966. A pressure field
solution at a given operating condition of 𝑀=7,𝛼=4, and 𝛽=0is shown in Figure 1.
In order to gain design insights for performance, stability, and reliability of a hypersonic vehicle, CFD simulations
are required over a range of flight conditions. For example, stability analyses for a hypersonic vehicle requires an
understanding of the pressure field surrounding the vehicle over the operating range of Mach number, angle of attack,
and sideslip angle of the vehicle. However, high-fidelity CFD solutions are computationally intensive due to the fine
mesh size required to adequately capture the physics of hypersonic flight. In this work, we address the prohibitive
computational cost through constructing cheaper approximations using MF linear regression techniques that reduce the
number of HF model evaluations required to make accurate predictions of the pressure fields by combining with data
from cheaper LF models.
B. Model Specifications and Data Generation
We can construct different levels of fidelity for the pressure field solution by leveraging Cart3D’s inherent mesh
adaptation over multiple adaptation cycles. We control the number of adaptations and the error tolerance. In this work,
5
Fig. 1 Cart3D pressure solution at 𝑀=7,𝛼=4, and 𝛽=0
we define two levels of fidelity for simulating the pressure field: (i) the HF model with a finer volume mesh after more
mesh adaptations and (ii) the LF model with a coarser volume mesh after fewer mesh adaptations and with a lower
error tolerance. We control the maximum number of mesh refinements ("Max Refinement"), the maximum number
of adaptation processes ("Max Adaptations"), error tolerance, and the number of iterations per adaptation process
("Cycles/Adaptation") to generate the different fidelity levels. The specifications for the HF model and the LF model
used in this work are described in Table 1. We also provide the relative computational cost in terms of one HF model
evaluation.
While the choice of HF and LF sample sizes is problem- and resource-dependent, in this case we use a very limited
number of HF samples,
𝑁HF {
3
,
5
,
10
}
, a LF training sample size of
𝑁LF =
80, and a HF testing sample size of
𝑁test
HF =
50 to analyze the algorithms’ effectiveness in the ultra low-data regime. We present the results while accounting
for the computational cost of using the additional 80 LF samples given by 80
/
127
=
0
.
63 equivalent HF samples. In all
of our results, we bootstrap the available dataset to provide a measure of robustness of each method over 50 repetitions
of the training and the testing samples.
Table 1 Model specifications
Model Type Max Refinement Max Adaptations Error Tolerance Cycles/Adaptation Cost (# HF)
HF 7 12 1e-3 175 1
LF 5 2 5e-3 50 1/127
C. Results for Projection-Enabled Multifidelity Linear Regression
We first analyze the dimensionality reduction on our training datasets of
𝑁HF =
10 and
𝑁LF =
80 to select an
appropriate lower-dimensional subspace size. Figures 2 and 3 show the singular value decay and the cumulative energy
plots for the LF and HF data, respectively. We show the median of 50 repetitions of SVD computations and the 25th and
75th percentile shaded around the median curve. There is not much variability in the singular values as seen from the
plots. We use a tolerance of
𝜖=
0
.
995 for the cumulative energy to decide the size of the low-dimensional subspace
using Eqn.
(2)
, which leads to
𝑘=
7for most LF training datasets and
𝑘=
4for most HF training data sets. This
facilitates the use of lower dimensional representations of the data for the surrogate models to be trained on, without
significant loss of information.
We apply the three MF linear regression methods described in Section III to the prediction of the pressure field on
the IC3X testbed hypersonic vehicle. We evaluate the performance of a surrogate model through the normalized L2
6
(a) Singular Values (b) Cumulative energy
Fig. 2 SVD on 80 LF training data
(a) Singular Values (b) Cumulative energy
Fig. 3 SVD on 10 HF training data
accuracy metric given by (1𝜖L2 ), where the normalized L2 error 𝜖L2 is defined by
𝜖L2
=
1
𝑁test
HF
𝑁test
HF
𝑖=1
𝒚𝑖ˆ
𝒚𝑖2
𝒚𝑖2
,(5)
where
.2
is the L2 vector norm,
𝒚𝑖
is the HF model solution at
𝑖th
test sample, and
ˆ
𝒚𝑖
is the surrogate prediction at
𝑖th
test sample. Note that the results for the single-fidelity (SF) surrogate model refer to the linear regression which was
trained on the HF pressure field data only. Since the surrogate models were trained on 50 repetitions of the training and
test dataset, we present the median, 25th, and 75th percentiles of the test accuracies. For the SF model, the order of the
polynomial was limited by the number of samples available limiting the choice to a linear equation in all cases. The
MF linear regression with the additive structure also used a linear polynomial since it is trained on the same amount of
HF data albeit with the discrepancy added. Lastly, both the MF surrogate models using the data augmentation methods
were able to be trained using a polynomial of order two since the number of samples available to train was larger by the
nature of the algorithms.
We first study the impact of different sample weighting schemes on the results of the two data augmentation
methods as shown in Figure 4. To illustrate the effect of sample weighting schemes, a weighting scheme of
𝑤HF =
1
, 𝑤syn ={
0
.
9
,
0
.
5
,
0
.
1
,
0
.
01
}
is tested. We find that the direct data augmentation method is sensitive to the
choice of
𝑤syn
, with a variation of up to 10% in median accuracy. On the contrary, the explicit map data augmentation
method is less sensitive to changes in the sample weight, with a variation of up to 2% in median accuracy. In both cases,
optimal selection of the synthetic sample weight could improve the accuracy of the MF linear regression using data
augmentation.
7
(a) Direct Augmentation (b) Explicit Map
Fig. 4 Comparison of sample weighting schemes for MF linear regression using data augmentation
Figure 5 shows the comparison of the three different MF linear regression methods proposed in this work with the
SF surrogate model. The additive MF method performs similar to the SF linear regression and does not offer significant
increase in accuracy for this application. The additive approach shows modest improvements for
𝑁HF
5samples. In
contrast, both the data augmentation techniques (using
𝑤syn =
0
.
01) perform better than the additive approach and show
significant improvement in accuracy over the SF linear regression for equivalent computational cost. Furthermore, the
robustness of both the MF linear regression models with data augmentation is markedly better than the SF surrogate
model. This is to be expected as the MF surrogate model sees more data when training the linear regression model.
The extra LF samples in data augmentation methods are of course not fully representative of the HF model, so we use
sample weights of 1 for the HF samples and
𝑤syn =
0
.
01 for the synthetic data generated from the LF samples. The MF
method with explicit map for data augmentation performs the best of the three presented here, specifically in the fewer
data regime of
𝑁HF [
3
,
5
]
HF samples. Table 2 provides the median accuracies of each regression method for
𝑁HF =
3, 5, and 10 HF samples. We can see that the data augmentation technique using explicit map leads to an improvement
of approximately 12% compared to the SF model for
𝑁HF =
3HF samples and 3
.
6% compared to the SF model for
𝑁HF =10 HF samples.
Table 2 Multifidelity linear regression results
Model Type # LF Samples # HF Samples Median Normalized L2 Test Accuracy
SF
- 3 0.769
- 5 0.870
- 10 0.898
MF - Additive 80
3 0.765
5 0.892
10 0.919
MF - Direct data augmentation 80
3 0.854
5 0.903
10 0.935
MF - Explicit map data augmentation 80
3 0.890
5 0.919
10 0.934
8
(a) Comparison of all MF methods (b) Additive Method
(c) Data augmentation: direct augmentation (d) Data augmentation: explicit mapping
Fig. 5 Comparison of MF linear regression methods to baseline SF linear regression (DA denotes data
augmentation in the legends)
Finally, we look at a comparison of the absolute errors in pressure prediction between the SF surrogate and the
MF surrogate using the direct data augmentation method with explicit map. For a random test sample, we predict the
pressure field using the surrogates and show the absolute error compared to the HF model simulation. We show a
contour plot of the errors on the vehicle body in Figure 6, providing some context for the gains the MF surrogate model
nets.
V. Conclusion
This work presents and contrasts multifidelity linear regression methods for problems in the ultra low-data regime
with two overarching ideas: (i) using discrepancy/additive structure and (ii) using data augmentation. In the additive
structure for MF linear regression, we use a linear regression model to calibrate the LF data and better align it to the HF
data based on the Kennedy O’Hagan approach. In the MF linear regression using data augmentation, we transform
the LF data in two different ways and augment the transformed data to the HF dataset to perform a weighted least
squares linear regression. In all these methods we embed dimensionality reduction through the proper orthogonal
decomposition to tackle high-dimensional outputs. A numerical example on the prediction of the pressure load upon a
hypersonic vehicle in-flight is used to compare and contrast the various MF approaches. For this application and HF
samples in the range of 3 to 10, we find that the additive approach does not substantially improve the accuracy compared
to the baseline SF surrogate model. The data augmentation techniques produce robust and accurate surrogate models,
9
(a) SF Surrogate Model Errors (b) MF Surrogate Model Errors
Fig. 6 Comparison of errors in pressure field prediction
with up to approximately 3
12% in median accuracy gain in the low-data regime as compared to the SF surrogate. The
direct data augmentation method had comparable accuracy to the explicit mapping method, but showed more sensitivity
to the selection of the synthetic data weight in the weighted least squares regression. One possible future direction is
to explore improvements in the performance of the data augmentation methods through an optimal hyperparameter
selection strategy for the weight associated with the synthetic data in the weighted least squares regression.
Acknowledgments
This work has been supported in part by the ARPA-E Differentiate award number DE-AR0001208 and AFOSR
grant FA9550-21-1-0089 under the NASA University Leadership Initiative (ULI).
References
[1]
Baker, N., Alexander, F., Bremer, T., Hagberg, A., Kevrekidis, Y., Najm, H., Parashar, M., Patra, A., Sethian, J., Wild, S.,
Willcox, K., and Lee, S., “Workshop Report on Basic Research Needs for Scientific Machine Learning: Core Technologies for
Artificial Intelligence,” 2019. https://doi.org/10.2172/1478744.
[2]
Peherstorfer, B., Willcox, K., and Gunzburger, M., “Survey of Multifidelity Methods in Uncertainty Propagation, Inference, and
Optimization,” SIAM Review, Vol. 60, No. 3, 2018, pp. 550–591. https://doi.org/10.1137/16M1082469.
[3]
Myers, R., and Montgomery, D., Response Surface Methodology: Process and Product Optimization Using Designed Experi-
ments, Wiley Series in Probability and Statistics, Wiley, 1995. URL https://books.google.com/books?id=7xvvAAAAMAAJ.
[4] Seber, G. A., and Lee, A. J., Linear regression analysis, John Wiley & Sons, 2012.
[5]
Montgomery, D., Peck, E., and Vining, G., Introduction to Linear Regression Analysis, Wiley Series in Probability and Statistics,
Wiley, 2021. URL https://books.google.com/books?id=tCIgEAAAQBAJ.
[6]
Simpson, T., Poplinski, J., Koch, P. N., and Allen, J., “Metamodels for Computer-based Engineering Design: Survey and
recommendations,” Engineering with Computers, Vol. 17, No. 2, 2001, pp. 129–150. https://doi.org/10.1007/pl00007198.
[7]
Box, G. E. P., and Wilson, K. B., “On the Experimental Attainment of Optimum Conditions,” Journal of the Royal
Statistical Society: Series B (Methodological), Vol. 13, No. 1, 1951, pp. 1–38. https://doi.org/https://doi.org/10.1111/j.2517-
6161.1951.tb00067.x.
[8]
Madsen, J. I., Shyy, W., and Haftka, R. T., “Response Surface Techniques for Diffuser Shape Optimization, AIAA Journal,
Vol. 38, No. 9, 2000, pp. 1512–1518. https://doi.org/10.2514/2.1160.
10
[9]
Nakamura, T., Fukami, K., and Fukagata, K., “Identifying key differences between linear stochastic estimation and neural
networks for fluid flow regressions,” Scientific Reports, Vol. 12, No. 1, 2022. https://doi.org/10.1038/s41598- 022-07515-7.
[10]
Hosder, S., Watson, L. T., Grossman, B., Mason, W. H., Kim, H., Haftka, R. T., and Cox, S. E., “Polynomial Response Surface
Approximations for the Multidisciplinary Design Optimization of a High Speed Civil Transport,” Optimization and Engineering,
Vol. 2, No. 4, 2001, pp. 431–452. https://doi.org/10.1023/a:1016094522761.
[11]
Mack, Y., Goel, T., Shyy, W., and Haftka, R., “Surrogate Model-Based Optimization Framework: A Case Study in Aerospace
Design,” Studies in Computational Intelligence, Springer Berlin Heidelberg, 2007, pp. 323–342. https://doi.org/10.1007/978- 3-
540-49774-5_14.
[12]
Ladický, L., Jeong, S., Solenthaler, B., Pollefeys, M., and Gross, M., “Data-Driven Fluid Simulations Using Regression Forests,”
ACM Trans. Graph., Vol. 34, No. 6, 2015. https://doi.org/10.1145/2816795.2818129.
[13]
Balabanov, V., Grossman, B., Watson, L., Mason, W., and Haftka, R., “Multifidelity response surface model for HSCT wing
bending material weight,” 7th AIAA/USAF/NASA/ISSMO Symposium on Multidisciplinary Analysis and Optimization, 1998, p.
4804.
[14]
Madsen, J. I., and Langthjem, M., “Multifidelity Response Surface Approximations for the Optimum Design of Diffuser Flows,
Optimization and Engineering, Vol. 2, No. 4, 2001, pp. 453–468. https://doi.org/10.1023/a:1016046606831.
[15]
Vitali, R., Haftka, R. T., and Sankar, B. V., “Multi-fidelity design of stiffened composite panel with a crack, Structural and
Multidisciplinary Optimization, Vol. 23, No. 5, 2002, pp. 347–356.
[16]
Zhang, Y., Kim, N. H., Park, C., and Haftka, R. T., “Multifidelity Surrogate Based on Single Linear Regression,” AIAA Journal,
Vol. 56, No. 12, 2018, pp. 4944–4952. https://doi.org/10.2514/1.j057299.
[17]
Park, C., Haftka, R. T., and Kim, N. H., “Remarks on multi-fidelity surrogates,” Structural and Multidisciplinary Optimization,
Vol. 55, No. 3, 2016, pp. 1029–1050. https://doi.org/10.1007/s00158-016-1550- y.
[18]
Meng, X., and Karniadakis, G. E., “A composite neural network that learns from multi-fidelity data: Application to
function approximation and inverse PDE problems,” Journal of Computational Physics, Vol. 401, 2020, p. 109020.
https://doi.org/https://doi.org/10.1016/j.jcp.2019.109020.
[19]
Guo, M., Manzoni, A., Amendt, M., Conti, P., and Hesthaven, J. S., “Multi-fidelity regression using artificial neural networks:
Efficient approximation of parameter-dependent output quantities,” Computer Methods in Applied Mechanics and Engineering,
Vol. 389, 2022, p. 114378. https://doi.org/10.1016/j.cma.2021.114378.
[20]
Forrester, A. I., Sóbester, A., and Keane, A. J., “Multi-fidelity optimization via surrogate modelling,” Proceedings of the royal
society a: mathematical, physical and engineering sciences, Vol. 463, No. 2088, 2007, pp. 3251–3269.
[21]
Le Gratiet, L., and Garnier, J., “Recursive co-kriging model for design of computer experiments with multiple levels of fidelity,”
International Journal for Uncertainty Quantification, Vol. 4, No. 5, 2014.
[22]
Kennedy, M. C., and O’Hagan, A., “Predicting the Output from a Complex Computer Code When Fast Approximations Are
Available,” Biometrika, Vol. 87, No. 1, 2000, pp. 1–13. URL http://www.jstor.org/stable/2673557.
[23]
Benner, P., Gugercin, S., and Willcox, K., “A survey of projection-based model reduction methods for parametric dynamical
systems, SIAM Review, Vol. 57, No. 4, 2015, pp. 483–531.
[24]
Swischuk, R., Mainini, L., Peherstorfer, B., and Willcox, K., “Projection-based model reduction: Formulations for physics-based
machine learning,” Computers & Fluids, Vol. 179, 2019, pp. 704–717. https://doi.org/https://doi.org/10.1016/j.compfluid.2018.
07.021.
[25]
Guo, M., and Hesthaven, J. S., “Data-driven reduced order modeling for time-dependent problems,” Computer Methods in
Applied Mechanics and Engineering, Vol. 345, 2019, pp. 75–99.
[26]
O’Leary-Roseberry, T., Villa, U., Chen, P., and Ghattas, O., “Derivative-informed projected neural networks for high-dimensional
parametric maps governed by PDEs, Computer Methods in Applied Mechanics and Engineering, Vol. 388, 2022, p. 114199.
https://doi.org/10.1016/j.cma.2021.114199.
[27]
Pasiliao, C. L., Sytsma, M. J., Neergaard, L., Witeof, Z., and Trolier, J. W., “Preliminary Aero-thermal Structural Simulation,”
14th AIAA Aviation Technology, Integration, and Operations Conference, American Institute of Aeronautics and Astronautics,
2014. https://doi.org/10.2514/6.2014-2292.
11
[28]
Witeof, Z., and Neergaard, L., “Initial concept 3.0 finite element model definition, Eglin Air Force Base, Air Force Research
Laboratory, AFRL-RWWV-TN-2014-0013, FL, 2014.
[29]
NASA Ames Research Center, “Automated Triangle Geometry Processing for Surface Modeling and Cartesian Grid Generation
(Cart3D),” , Accessed 2022. URL https://software.nasa.gov/software/ARC-14275-1.
[30]
Aftosmis, M., Berger, M., and Adomavicius, G., “A parallel multilevel method for adaptively refined Cartesian grids with
embedded boundaries,” 38th AIAA Aerospace Sciences Meeting and Exhibit, 2000, pp. 2000–0808. https://doi.org/10.2514/6.
2000-808.
[31]
Aftosmis, M. J., Berger, M. J., and Melton, J. E., “Robust and Efficient Cartesian Mesh Generation for Component-Based
Geometry, AIAA Journal, Vol. 36, No. 6, 1998, pp. 952–960. https://doi.org/10.2514/2.464.
12
... Parida et al. 2023 also augmented earthquake signals using a discrete wavelet transform method, creating new samples by perturbing the wavelet coefficients of original signals. Additionally, many surrogate models trained on image-based snapshots, e.g., [Dong et al. 2022]; thus employed random rotations to ensure orientation invariance, and in some multi-fidelity studies, e.g., [Sella et al. 2023], LF training samples directly augment datasets. Our research particularly augments datasets for multiscale surrogate modeling without any large-scale data, a novel approach in this context. ...
Thesis
Articular cartilage (AC), a crucial soft tissue for pain-free movement, undergoes significant biomechanical stress, warranting research through non-invasive computer simulations. Such simulations commonly use time-intensive numerical methods, such as finite element (FE) analysis. Artificial intelligence (AI), with machine learning (ML) as a surrogate model, presents a faster alternative, imitating numerical analysis with numerical data samples. Despite its potential, this method often requires extensive training and large datasets, which is inefficient. This study first introduces an efficient biomechanical model incorporating multi-physics modeling and a novel pre-stressing algorithm (PSA), generating few training samples across different fidelities and scales. We then propose efficient, multi-fidelity, generalizable surrogate models, enhanced by innovative preprocessing and training algorithms. Finally, the research code, including our developed Fortran subroutines and Python scripts, is open-sourced. Empirical results emphasize the significance of pre-stressing in multiphasic modeling and the value of efficient surrogate modeling in reducing computational time. The scalability and generalizability of the hybrid ML (HML) framework are confirmed through multiscale simulations, demonstrating the impact of our physics-constrained data augmentation (DA) and graph neural network (GNN) implementation. Therefore, while this study potentially advances the field with the developed PSA implementation, it also presents a straightforward approach to address computational challenges and data scarcity in ML-based surrogate modeling of cartilage biomechanics. Such research, though tailored to specific cartilage modeling, potentially paves the way for broader applications of our synergistic methodology through its open-source availability.
... Moreover, problem-specific algorithms have been developed, e.g., to augment datasets, e.g., by using pairwise feature differences, if applicable [40] or the application of signal decomposition [41]. Additionally, in certain multi-fidelity studies, low-fidelity (LF) training samples were added to the training sets for augmentation [42]. Our research presents a novel DA specifically tailored to VOLUME 12, 2024 multiscale numerical problems where extensive training data is not available. ...
Article
Full-text available
Articular cartilage (AC) is essential for minimizing friction in the human knee, but its healthy function is highly influenced by biomechanical factors such as weight bearing. Non-invasive biomechanical and numerical simulations are widely used to study AC but often require complex and costly numerical approximations. Machine learning (ML) provides a more efficient alternative and uses data from these numerical methods for training. Hybrid ML models (HML) complemented by reduced-order numerical models can achieve similar outcomes with minimal data input but may have problems with generalizability across different scales. In this study, we present an extended HML framework (EHML) for developing a multiscale surrogate model specifically tailored for knee cartilage simulations. Our approach is based on integrating hybrid graph neural networks (GNNs) with tissue-scale data and aims to achieve remarkable few-shot learning and potential zero-shot generalizability for large-scale analysis. The main proposed idea is a physics-constrained data augmentation (DA) technique coupled with a set of pre-processing and customization algorithms to bridge the scales. Specifically, we integrate feature transformations, resampling, and cost-sensitive functions to manage the observed data imbalances, all within a customized, memory-efficient training framework. Our rigorous testing using an advanced multi-physics cartilage model demonstrates the viability of our approach. Comparative analyses underscore the significant role of pre-processing and DA methods in enhancing generalizability and efficiency. They helped reduce the normalized mean squared errors to 0.1 or less (compared to the ablated model with its error of 2 or higher). Therefore, this work represents an important step towards addressing the challenges of limited generalizability and efficiency of existing ML-based surrogate models and opens new possibilities for their application in more complex simulations.
Article
Full-text available
Neural networks (NNs) and linear stochastic estimation (LSE) have widely been utilized as powerful tools for fluid-flow regressions. We investigate fundamental differences between them considering two canonical fluid-flow problems: (1) the estimation of high-order proper orthogonal decomposition coefficients from low-order their counterparts for a flow around a two-dimensional cylinder, and (2) the state estimation from wall characteristics in a turbulent channel flow. In the first problem, we compare the performance of LSE to that of a multi-layer perceptron (MLP). With the channel flow example, we capitalize on a convolutional neural network (CNN) as a nonlinear model which can handle high-dimensional fluid flows. For both cases, the nonlinear NNs outperform the linear methods thanks to nonlinear activation functions. We also perform error-curve analyses regarding the estimation error and the response of weights inside models. Our analysis visualizes the robustness against noisy perturbation on the error-curve domain while revealing the fundamental difference of the covered tools for fluid-flow regressions.
Article
Full-text available
Highly accurate numerical or physical experiments are often very time-consuming or expensive to obtain. When time or budget restrictions prohibit the generation of additional data, the amount of available samples may be too limited to provide satisfactory model results. Multi-fidelity methods deal with such problems by incorporating information from other sources, which are ideally well-correlated with the high-fidelity data, but can be obtained at a lower cost. By leveraging correlations between di↵erent data sets, multi-fidelity methods often yield superior generalization when compared to models based solely on a small amount of high-fidelity data. In the current work, we present the use of artificial neural networks applied to multi-fidelity regression problems. By elaborating a few existing approaches, we propose new neural network architectures for multi-fidelity regression. The introduced models are compared against a traditional multi-fidelity regression scheme-co-kriging. A collection of artificial benchmarks are presented to measure the performance of the analyzed models. The results show that cross-validation in combination with Bayesian optimization leads to neural network models that outperform the co-kriging scheme. Additionally, we show an application of multi-fidelity regression to an engineering problem. The propagation of a pressure wave into an acoustic horn with parametrized shape and frequency is considered, and the index of reflection intensity is approximated using the proposed multi-fidelity models. A finite element, full-order model and a reduced-order model built through the reduced basis method are adopted as the high-and low-fidelity, respectively. It is shown that the multi-fidelity neural networks return outputs that achieve a comparable accuracy to those from the expensive, full-order model, using only very few full-order evaluations combined with a larger amount of inaccurate but cheap evaluations of the reduced order model.
Article
Full-text available
Many-query problems – arising from, e.g., uncertainty quantification, Bayesian inversion, Bayesian optimal experimental design, and optimization under uncertainty – require numerous evaluations of a parameter-to-output map. These evaluations become prohibitive if this parametric map is high-dimensional and involves expensive solution of partial differential equations (PDEs). To tackle this challenge, we propose to construct surrogates for high-dimensional PDE-governed parametric maps in the form of derivative-informed projected neural networks (DIPNets) that parsimoniously capture the geometry and intrinsic low-dimensionality of these maps. Specifically, we compute Jacobians of these PDE-based maps, and project the high-dimensional parameters onto a low-dimensional derivative-informed active subspace; we also project the possibly high-dimensional outputs onto their principal subspace. This exploits the fact that many high-dimensional PDE-governed parametric maps can be well-approximated in low-dimensional parameter and output subspaces. We use the projection basis vectors in the active subspace as well as the principal output subspace to construct the weights for the first and last layers of the neural network, respectively. This frees us to train the weights in only the low-dimensional layers of the neural network. The architecture of the resulting neural network then captures, to first order, the low-dimensional structure and geometry of the parametric map. We demonstrate that the proposed projected neural network achieves greater generalization accuracy than a full neural network, especially in the limited training data regime afforded by expensive PDE-based parametric maps. Moreover, we show that the number of degrees of freedom of the inner layers of the projected network is independent of the parameter and output dimensions, and high accuracy can be achieved with weight dimension independent of the discretization dimension.
Article
Full-text available
A data-driven reduced basis (RB) method for parametrized time-dependent problems is proposed. This method requires the offline preparation of a database comprising the time history of the full-order solutions at parameter locations. Based on the full-order data, a reduced basis is constructed by the proper orthogonal decomposition (POD), and the maps between the time/parameter values and the projection coefficients onto the RB are approximated as a regression model. With a natural tensor grid between the time and the parameters in the database, a singular-value decomposition (SVD) is used to extract the principal components in the data of projection coefficients. The regression functions are represented as the linear combinations of several tensor products of two Gaussian processes, one of time and the other of parameters. During the online stage, the solutions at new time/parameter locations in the domain of interest can be recovered rapidly as outputs from the regression models. Featuring a non-intrusive nature and the complete decoupling of the offline and online stages, the proposed approach provides a reliable and efficient tool for approximating parametrized time-dependent problems, and its effectiveness is illustrated by non-trivial numerical examples.
Article
Full-text available
Multifidelity surrogates (MFS) combine low-fidelity models with few high-fidelity samples to infer the response of the high-fidelity model for design optimization or uncertainty quantification. Most publications in MFS focus on Bayesian frameworks based on Gaussian process. Other types of surrogates might be preferred for some applications. In this paper, a simple and yet powerful MFS based on single linear regression is proposed, termed as linear regression multifidelity surrogate (LR-MFS), especially for fitting high-fidelity data with noise. The LR-MFS considers the low-fidelity model as a basis function and identifies unknown coefficients of both the low-fidelity model and the discrepancy function using a single linear regression. Because the proposed LR-MFS is obtained from standard linear regression, it can take advantage of established regression techniques such as prediction variance, D-optimal design, and inference. The LR-MFS is first compared with three Bayesian frameworks using a benchmark dataset from the simulations of a fluidized-bed process. The LR-MFS showed a comparable accuracy with the best Bayesian frameworks. The effect of combining multiple low-fidelity models was also discussed. Then the LR-MFS is evaluated using an algebraic function with different sampling plans. The LR-MFS bested co-kriging for 55∼63% cases with an increasing number of high-fidelity (HF) samples. The sources of uncertainty with an increasing number of samples were also discussed. For both examples, the LR-MFS proved to be better than fitting only HF samples and robust with noisy data.
Article
Currently the training of neural networks relies on data of comparable accuracy but in real applications only a very small set of high-fidelity data is available while inexpensive lower fidelity data may be plentiful. We propose a new composite neural network (NN) that can be trained based on multi-fidelity data. It is comprised of three NNs, with the first NN trained using the low-fidelity data and coupled to two high-fidelity NNs, one with activation functions and another one without, in order to discover and exploit nonlinear and linear correlations, respectively, between the low-fidelity and the high-fidelity data. We first demonstrate the accuracy of the new multi-fidelity NN for approximating some standard benchmark functions but also a 20-dimensional function that is not easy to approximate with other methods, e.g. Gaussian process regression. Subsequently, we extend the recently developed physics-informed neural networks (PINNs) to be trained with multi-fidelity data sets (MPINNs). MPINNs contain four fully-connected neural networks, where the first one approximates the low-fidelity data, while the second and third construct the correlation between the low- and high-fidelity data and produce the multi-fidelity approximation, which is then used in the last NN that encodes the partial differential equations (PDEs). Specifically, by decomposing the correlation into a linear and nonlinear part, the present model is capable of learning both the linear and complex nonlinear correlations between the low- and high-fidelity data adaptively. By training the MPINNs, we can: (1) obtain the correlation between the low- and high-fidelity data, (2) infer the quantities of interest based on a few scattered data, and (3) identify the unknown parameters in the PDEs. In particular, we employ the MPINNs to learn the hydraulic conductivity field for unsaturated flows as well as the reactive models for reactive transport. The results demonstrate that MPINNs can achieve relatively high accuracy based on a very small set of high-fidelity data. Despite the relatively low dimension and limited number of fidelities (two-fidelity levels) for the benchmark problems in the present study, the proposed model can be readily extended to very high-dimensional regression and classification problems involving multi-fidelity data.
Article
In many situations across computational science and engineering, multiple computational models are available that describe a system of interest. These different models have varying evaluation costs and varying fidelities. Typically, a computationally expensive high-fidelity model describes the system with the accuracy required by the current application at hand, while lower-fidelity models are less accurate but computationally cheaper than the high-fidelity model. Outer-loop applications, such as optimization, inference, and uncertainty quantification, require multiple model evaluations at many different inputs, which often leads to computational demands that exceed available resources if only the high-fidelity model is used. This work surveys multifidelity methods that accelerate the solution of outer-loop applications by combining high-fidelity and low-fidelity model evaluations, where the low-fidelity evaluations arise from an explicit low-fidelity model (e.g., a simplified physics approximation, a reduced model, a data-fit surrogate) that approximates the same output quantity as the high-fidelity model. The overall premise of these multifidelity methods is that low-fidelity models are leveraged for speedup while the high-fidelity model is kept in the loop to establish accuracy and/or convergence guarantees. We categorize multifidelity methods according to three classes of strategies: adaptation, fusion, and filtering. The paper reviews multifidelity methods in the outer-loop contexts of uncertainty propagation, inference, and optimization.
Article
This paper considers the creation of parametric surrogate models for applications in science and engineering where the goal is to predict high-dimensional output quantities of interest, such as pressure, temperature and strain fields. The proposed methodology develops a low-dimensional parametrization of these quantities of interest using the proper orthogonal decomposition (POD), and combines this parametrization with machine learning methods to learn the map between the input parameters and the POD expansion coefficients. The use of particular solutions in the POD expansion provides a way to embed physical constraints, such as boundary conditions and other features of the solution that must be preserved. The relative costs and effectiveness of four different machine learning techniques—neural networks, multivariate polynomial regression, k-nearest-neighbors and decision trees—are explored through two engineering examples. The first example considers prediction of the pressure field around an airfoil, while the second considers prediction of the strain field over a damaged composite panel. The case studies demonstrate the importance of embedding physical constraints within learned models, and also highlight the important point that the amount of model training data available in an engineering setting is often much less than it is in other machine learning applications, making it essential to incorporate knowledge from physical models.
Chapter
Simultaneous Interval EstimationConfidence Bands for the Regression SurfacePrediction Intervals and Bands for the ResponseEnlarging the Regression Matrix