Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Received: 30 September 2022 Revised: 28 December 2022 Accepted: 20 January 2023

DOI: 10.1002/eqe.3839

SPECIAL ISSUE ARTICLE

Surrogate modeling of structural seismic response using


probabilistic learning on manifolds

Kuanshi Zhong1 Javier G. Navarro2 Sanjay Govindjee3


Gregory G. Deierlein1

1 Department of Civil and Environmental


Engineering, Stanford University, Abstract
Stanford, California, USA Nonlinear response history analysis (NLRHA) is generally considered to be a reli-
2 Department of Civil, Environmental and able and robust method to assess the seismic performance of buildings under
Architectural Engineering, University of
Colorado Boulder, Boulder, Colorado, strong ground motions. While NLRHA is fairly straightforward to evaluate indi-
USA vidual structures for a select set of ground motions at a specific building site,
3 Department of Civil and Environmental
it becomes less practical for performing large numbers of analyses to evaluate
Engineering, University of California,
either (1) multiple models of alternative design realizations with a site-specific
Berkeley, Berkeley, California, USA
set of ground motions, or (2) individual archetype building models at multi-
Correspondence ple sites with multiple sets of ground motions. In this regard, surrogate models
Sanjay Govindjee.
Email: s_g@berkeley.edu
offer an alternative to running repeated NLRHAs for variable design realiza-
tions or ground motions. In this paper, a recently developed surrogate modeling
Current address technique, called probabilistic learning on manifolds (PLoM), is presented to
Kuanshi Zhong, University of Cincinnati,
Department of Civil and Architecture estimate structural seismic response. Essentially, the PLoM method provides an
Engineering and Construction efficient stochastic model to develop mappings between random variables, which
Management
can then be used to efficiently estimate the structural responses for systems with
Funding information variations in design/modeling parameters or ground motion characteristics. The
Centro de Formación Interdisciplinaria PLoM algorithm is introduced and then used in two case studies of 12-story
Superior; National Science Foundation,
buildings for estimating probability distributions of structural responses. The
Grant/Award Numbers: CMMI-1612843,
2131111 first example focuses on the mapping between variable design parameters of a
multidegree-of-freedom analysis model and its peak story drift and acceleration
responses. The second example applies the PLoM technique to estimate struc-
tural responses for variations in site-specific ground motion characteristics. In
both examples, training data sets are generated for orthogonal input parameter
grids, and test data sets are developed for input parameters with prescribed sta-
tistical distributions. Validation studies are performed to examine the accuracy
and efficiency of the PLoM models. Overall, both examples show good agreement
between the PLoM model estimates and verification data sets. Moreover, in con-
trast to other common surrogate modeling techniques, the PLoM model is able

This is an open access article under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License, which permits use and distribution in any medium,
provided the original work is properly cited, the use is non-commercial and no modifications or adaptations are made.
© 2023 The Authors. Earthquake Engineering & Structural Dynamics published by John Wiley & Sons Ltd.

Earthquake Engng Struct Dyn. 2023;52:2407–2428. wileyonlinelibrary.com/journal/eqe 2407


10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2408 ZHONG et al.

to preserve correlation structure between peak responses. Parametric studies are


conducted to understand the influence of different PLoM tuning parameters on
its prediction accuracy.

KEYWORDS
incremental dynamic analysis, machine learning, multiple stripe analysis, probabilistic
learning on manifolds, seismic response prediction, site-specific

1 INTRODUCTION

Estimating structural response to earthquake ground motions and quantifying its variability is an essential task in
Performance-Based Earthquake Engineering (PBEE).1,2 In engineering practice, nonlinear response history analy-
sis (NLRHA) has become increasingly prevalent in recent years for seismic design and performance assessment of
new and existing buildings. The analysis model is created based on the structural design information and used to
calculate the dynamic response history of the structure under a set of ground motion records. The peak and cumu-
lative response quantities of the response history, referred to as Engineering Demand Parameters (EDPs), can be
then used to estimate damage to structural and nonstructural components. As a result of uncertainties in ground
motions, structural behaviors, numerical models, and construction material properties, the structural responses are not
deterministic. To account for this randomness, two sources of uncertainty are typically considered, namely, ground
motion record-to-record variability and structural modeling uncertainty.3–5 To explicitly quantify variability in the
responses, NLRHAs are typically conducted using sets of multiple input ground motions6–8 and/or structural modeling
parameters.9–12
In more complex applications such as reliability analysis, regional earthquake simulation, and seismic design optimiza-
tion, the above workflow is repeated to evaluate the structural responses for different realizations. Although available
computational power has been growing, the computational cost can still be excessive, especially when high-fidelity mod-
els and/or a large number of ground motions are used to reliably estimate small failure probability. Moreover, selecting
appropriate ground motions that represent the site-specific seismic hazard13–17 requires special care and effort.
As an alternative to running repeated NLRHAs for multiple models and/or multiple ground motion sets, surrogate
modeling (metamodeling) techniques are utilized to train statistical models from a set of simulation data to more effi-
ciently estimate structural responses. From ordinary linear regression to advanced machine learning,18 the general concept
behind surrogate models is to construct a function that maps the structural parameters and/or ground-motion parameters
to the structural responses.
In theory, provided sufficient data, a learned mapping function could be substituted for an explicit NLRHA. In
practice, surrogate modeling is an evolving field with alternative algorithms, statistical models, and computational
design strategies.19–27 The response surface method is one common technique where linear regression,28,29 polynomial
regression,11,30 kernel regression,31,32 and regularized regression methods33 can be implemented to relate structural design
parameters and/or ground motion characteristics to the EDPs or damage indicators. More recently, the Gaussian pro-
cess regression (or Kriging) method has received attention as it naturally characterizes the uncertainties in the predicted
responses, and its application for meta-modeling has been investigated.24,34 Artificial neural networks, along with other
advanced machine learning methods (e.g., random forest regression), are also being explored35–37 to map earthquake char-
acteristics to structural response demands. It is also an emerging field of adapting and utilizing deep learning methods
to model the structural response time history as sequential data (e.g., using recurrent neural networks, long short-term
memory architectures, or more recently tranformers38 ). The network architecture is designed and trained to map the
earthquake record to the structural response history.39,40 It is worth mentioning that the optimal surrogate model is
problem-dependent. Linear response surface models can be sufficient and most efficient in many cases, but they may
not be applicable for complex data structures. Complex models are more capable of capturing highly nonlinear data rela-
tionships, with the tradeoff being that the training of those models can require substantial computational resource. For
example, methods that can estimate the entire response history response obviously provides a more complete descrip-
tion of the structural behavior; however, the application of methods to do this is limited when the ground motion time
history record is not available for the given building site, which is common in regional seismic risk analysis problems.
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2409

Moreover, even when complete response histories are available, the results are often distilled down to simpler scalar
response metrics.
This study focuses on the surrogate modeling of peak (or cumulative) response measures of structures under earthquake
loads, that is, scalar vectors of EDPs, wherein structural design parameters of interest and accessible seismic intensity
measures are central elements of the surrogatating process. In this regard, there are important challenges to be recognized.
First, most studies prescribe the probability distribution type for the responses vector, such as a Gaussian distribution,
which may not represent the actual response (e.g., where the pre- and post-yield displacement responses may follow
two different distributions). Second, few studies20,22 have captured the correlation structure between different response
quantities (e.g., the interdependence between peak story drift ratios along the building height at different drift levels),
which can be important for accurate damage and loss assessment. Also, some methods are less efficient or effective for
engineering problems with high-dimensional uncertainty and limited sample sizes due to the expensive computational
cost for generating training data. Moreover, as specific to earthquake engineering applications, few studies have been
conducted on models that are generally applicable for a broad set of alternative structural designs23 or alternative site-
specific ground motion characteristics.32
To address these challenges, this paper explores the use of Soize and Ghanem’s recently developed machine learning
methodology, probabilistic learning on manifolds (PLoM).41–45 This methodology is just coming to the attention of engi-
neering researchers, with only a handful of papers utilizing the method outside of the research groups of the originators of
the PLoM methodology.46–50 In this paper, we show the utility of PLoM for estimating the joint probability distribution of
EDPs under earthquake ground motions. In particular, PLoM is proposed for generating realizations of a random vector
with values in a finite-dimensional Euclidean space that preserve the correlation structure of the original data and can
comply with user-defined statistical constraints. One promising feature of PLoM is its capability of achieving this in high-
dimensional random spaces using relatively small sample sizes. Our implementation of PLoM with constraints, which
was originally proposed by Soize and Ghanem,44 is to our knowledge the first independent implementation and test of
this important variant of the PLoM methodology.
The paper is organized as follows. Section 2 first discusses the general idea of applying surrogate modeling methods to
estimate earthquake structural responses and the distinct features of the proposed method, including the problem formu-
lation and an overview of the PLoM algorithm. Sections 3 and 4 present two case studies to demonstrate the use of PLoM
for predicting seismic responses for (1) alternative structural designs under a fixed set of ground motions and (2) a single
structural model with alternative ground motion characteristics for different geologic sites. Structural response statistics
from the PLoM prediction and direct NLRHAs are compared to examine the accuracy and efficiency of the proposed
method. Section 5 summarizes the case studies and aspects related to the PLoM modeling parameters and future studies.

2 PROBLEM STATEMENT

2.1 Problem overview

NLRHA has become an essential component in earthquake engineering applications, for example, examining structural
safety under critical earthquake scenarios, optimizing the structural design, and estimating the probabilities of various
damage stages and economic loss levels.
The inputs of the NLRHA consist of: (1) structural parameters 𝐱𝑆 and (2) ground motion parameters 𝐱𝐺𝑀 ; and the out-
puts are structural responses 𝐱𝑅 . The structural parameters can be characterized through building inventory data (e.g.,
year constructed, number of stories), structure design properties (e.g., construction material properties, framing dimen-
sions, member sizes), and structural modeling parameters (e.g., story stiffness, hinge yield moment). The ground motion
parameters can include ground motion intensity measures (e.g., response spectral acceleration at single or multiple periods
𝑆𝑎(𝑇𝑖 )) and other features (e.g., duration, spectral shape, pulse indices) abstracted from a seismogram; and the structural
responses can include EDPs, structural collapse indicators, and damage and loss indices.
Uncertainty in the model inputs is typically considered by evaluating multiple numerical models, which will need to be
built and analyzed with variable combinations of input parameters. The sampling and analyses of multiple realizations
can be computationally expensive, especially if high-fidelity models are employed; this can be impractical for analyses of
large building or bridge inventories for optimization studies, regional earthquake simulations, and other applications. In
such cases, surrogate models are intended to replace the explicit NLRHAs with a mapping function that is trained using
data generated by detailed NLRHAs and evaluating the response outputs more efficiently.
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2410 ZHONG et al.

The proposed surrogate modeling method using PLoM has three features that distinguish it from other prevailing
methods:

1. Instead of developing a functional (parametric) mapping from the vector-valued input parameters to output responses,
it learns a probabilistic mapping between the joint distributions of input parameters and output responses.
2. No specific distribution type is prescribed to the joint distribution of input parameters and output responses. Combined
with the first feature, the proposed method is naturally more flexible and able to capture local geometric structure of
data (e.g., nonstationary correlation structures, multimodal responses).
3. It couples a nonlinear dimension reduction (NDR) method called diffusion maps with Markov chain Monte Carlo
(MCMC) methods to effectively discover localized data clusters and efficiently generate new samples from the learned
distribution for high-dimensional problems.

To further elaborate on these features and the basis for investigating the effectiveness and efficiency of the proposed
surrogate modeling method, the following sections will abstract the problem formulation, summarize the key steps of the
implemented PLoM algorithm, and discuss the use cases focused upon in this study.

2.2 Abstract problem formulation

At an abstract level, the key problem to be addressed in this study is formulated as follows:

1. A mathematical model 𝐟 ∶ ℝ𝑛𝑆 +𝑛𝐺𝑀 → ℝ𝑛𝑅 is assumed between the structural and ground-motion parameter vectors
and the structural response, that is, 𝐱𝑅 = 𝐟 (𝐱𝑆 , 𝐱𝐺𝑀 ), where 𝐱𝑅 ∈ ℝ𝑛𝑅 is a vector of selected EDPs, 𝐱𝑆 ∈ ℝ𝑛𝑆 is a vec-
tor of structural parameters, and 𝐱𝐺𝑀 ∈ ℝ𝑛𝐺𝑀 is a vector of ground motion intensity measures. In this study, 𝐟 is the
model of the structural system, evaluated by NLRHA. It maps realizations of the underlying random vectors of struc-
tural parameters, 𝐗𝑆 , and ground-motion characteristics, 𝐗𝐺𝑀 , to a realization of the random vector of EDPs, 𝐗𝑅 . The
primary goal is to efficiently sample this later random vector. To that end, we slightly reformulate the set-up.
2. We denote realizations of an aggregate random vector, 𝐗 = (𝐗𝑆 , 𝐗𝐺𝑀 , 𝐗𝑅 ), as 𝐱 = (𝐱𝑆 , 𝐱𝐺𝑀 , 𝐱𝑅 ) in ℝ𝑛𝑆 +𝑛𝐺𝑀 +𝑛𝑅 . Such
realizations define the manifold for the response of the structure or structures of interest together with their seis-
mic excitation. It will also be convenient to define the matrix-valued random variable [𝔛] whose 𝑁 columns are
independent copies of 𝐗. A realization [𝔵] = [𝐱(1), 𝐱(2), … , 𝐱(𝑁)] of [𝔛] is an (𝑛𝑆 + 𝑛𝐺𝑀 + 𝑛𝑅 )-by-𝑁 matrix.
3. An initial matrix-valued sample, [𝔵0 ], is generated of the random matrix [𝔛]. The sample is generated so as to cover a
wide range of structural design parameters and ground motion parameters. In machine-learning parlance, this realiza-
tion will constitute our training data—representing 𝑁 points in the (𝑛𝑆 + 𝑛𝐺𝑀 + 𝑛𝑅 )-dimensional random space that
represents the system of interest. The expectation is that the data points are concentrated on a manifold of dimension
much less that (𝑛𝑆 + 𝑛𝐺𝑀 + 𝑛𝑅 ).
4. From the initial realization [𝔵0 ], new matrix realizations [𝔵new ] of [𝔛] are now generated. Each such realization con-
tains 𝑁 new realizations of the aggregate random vector 𝐗. The generation of the new samples is done so as to be
statistically faithful to the response manifold of the system being studied—the generation is additionally devised so as
to be efficient even when the dimensionality of the random space is large. We do this using the PLoM method.41 In the
variant of the PLoM algorithm employed in this paper, we generate these realizations such that they can additionally
follow prescribed statistical constraints (e.g., first and second moments) to represent specific categories of structural
designs or site hazards.44 In particular, we consider two separate problem classes that can take advantage of such an
algorithm—one where the structural parameters are varied and the other where the ground motion parameters are
varied.

2.3 PLoM conceptualization

As the PLoM methodology is relatively new, key steps of the method are described through an illustrative example in this
section (see Figure 1), and further details of the mathematical implementation of the PLoM algorithm are provided in
Appendix A. Readers are also referred to the original papers41–45 for further details.
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2411

F I G U R E 1 Flowchart of the implemented PLoM method. (A) Key steps of the PLoM algorithm. (B) Example initial data sample. (C)
Transformed data matrix [𝐇]. (D) Estimated nonparametric probability density of [𝐇]. (E) Reduced-order representation by diffusion maps
[𝐙][𝐠]𝑇 and new realizations. (F) New realizations in the original coordinate.

As noted in Section 2.2, the initial input to the PLoM model is an initial sample of data from the random matrix [𝔛]
whose realizations are 𝑛-by-𝑁 matrices, where 𝑛 is the dimension of the random data to be modeled and 𝑁 is the number
of available samples. In this example, the input data, shown in Figure 1B, are from incremental dynamic analyses (IDA) of
a 12-story building using a set of 49 earthquake ground motions. The data in Figure 1B, which are used as the initial sample
[𝐗raw ]𝑛×𝑁 , consist of 623 pairs of maximum story drift ratio (𝑆𝐷𝑅𝑚𝑎𝑥 ) and response acceleration at the first period of the
frame (𝑆𝑎) of various ground motions (𝑛 = 2 and 𝑁 = 623). In this example, we are interested to use the PLoM model to
learn the data structure and generate new estimates of response to represent ground shaking with a different mean value
from the set of input data. This exercise mimics the use of the IDA-based surrogate model to predict structural responses
to site-specific earthquake hazard characteristics for a specified shaking intensity or return period.
Referring to the PLoM algorithm flowchart in Figure 1A, the initial sample [𝐗raw ]𝑛×𝑁 is first scaled and normalized
by row to regularize the range of each individual dimension. Then principal component analysis (PCA) is performed
to project the data matrix [𝐗]𝑛×𝑁 to a set of orthogonal bases (PCA bases), see Figure 1C, where each data point in the
resulting matrix [𝐇]𝜈×𝑁 (𝜈 ≤ 𝑛) corresponds to one realization in [𝐗]𝑛×𝑁 . The selection of 𝜈 is dependent on a user-defined
parameter 𝜖PCA (it measures the error between the original data sample and the sample reconstructed using the 𝜈 PCA
bases). In this simple example 𝜈 = 2; thus, this step only decouples the representation, though in general one expects
𝜈 < 𝑛. In fact, in the PLoM algorithm, the use of PCA is primarily used for the purpose of decoupling the data to facilitate
the next algorithmic step of kernel density estimation (KDE).
The [𝐇]𝜈×𝑁 is viewed as a collection of 𝑁 realizations for 𝜈 linearly independent variables. As shown in Figure 1D, the
joint probability distribution 𝑝𝐇 is constructed by KDE using a Gaussian kernel with the optimal Silverman bandwidth.
The ultimate goal is to conditionally sample new data from this probability density. Before performing the conditional
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2412 ZHONG et al.

sampling, which will be introduced later, one more matrix transformation is performed to preserve local data concen-
tration. Referring to Figure 1C, although the data are linearly uncorrelated between two dimensions, note the nonlinear
relationship between the data along with multimodal structures. This local concentration of the data is preserved by
applying the concept of diffusion maps51 to further transform [𝐇]𝜈×𝑁 to a random matrix [𝐙]𝜈×𝑚 , that is, [𝐇] = [𝐙][𝐠]𝑇 ,
[𝐙] = [𝐇][𝐠]([𝐠]𝑇 [𝐠])−1 , where 𝑚 is the number of data clusters that are discovered and characterized by the diffusion-
map basis [𝐠]𝑁×𝑚 (𝑚 ≤ 𝑁). The diffusion-map basis is dependent on a user-defined kernel parameter 𝜖k , which controls
the resolution for clustering localized data. Typically, a larger 𝜖k guides the diffusion-maps basis to view the data more
collectively, and vice versa. The main function of this step is to apply the NDR using the diffusion-map basis. Similar
application of NDR can be seen in previous studies, for example, using the kernel principal component analysis (KPCA)
in surrogate models for high-dimensional data.52
A larger value of 𝑚 essentially reproduces finer local geometric structures (e.g., 𝑚 = 𝑁 would fully recover [𝐇]), while
an optimal 𝑚 is selected for the reduced-order representation [𝐙]𝜈×𝑚 by a special convergence criterion (please see more
details in Appendix A). For instance, Figure 1E plots the reduced-order representation by [𝐙] and [𝐠]𝑇 (with the first
153 clusters) as red dots along with the original [𝐇] data as black dots. The diffusion map basis is seen to very faithfully
represent the data even with a four-fold reduction in the data space.
The final step in the PLoM procedure is to generate new samples of [𝐙]𝜈×𝑚 and then transform them back to the original
coordinate system to obtain the new sample [𝐗new ]. This is performed by first formulating and solving a Galerkin projec-
tion of an Itô stochastic differential equation (ISDE), which views the logarithmic nonparametric probability distribution
as the potential of a stochastic dissipative Hamiltonian system (please see more details in Appendix A). The ISDE contains
a free parameter 𝑓0 (usually taken as 1.5) to damp the transition response in the system. For the earthquake engineering
applications considered in this paper, we also wish to obtain new samples from the surrogate, which satisfy additional,
say, design constraints. This is achieved by imposing a vector-value constraint (e.g., the mean values of multiple individual
input parameters) in the PLoM algorithm. In this case, the ISDE is numerically solved to efficiently produce new samples
([𝐙new ]𝜈×𝑚 ) using a potential whose probability density satisfies the constraint. From the space of such probability den-
sities, we chose the one that minimizes the Kullback–Leibler (KL) divergence with respect to 𝑝𝐇 within a user-defined
tolerance (𝜖KL ). For instance, in this example, suppose we are interested in sampling response data for ground motions
with a mean intensity of 𝑆𝑎 equal to 0.5 g, as opposed to the mean intensity of 𝑆𝑎 equal to 0.22 g in the original set of IDA
data. Figure 1E plots the new realizations for [𝐇] (green points), selected utilizing this constraint, against the original [𝐇].
It is seen that we properly generate samples faithful to the original data conditioned on the constraint.
Finally, the new samples [𝐙new ]𝜈×𝑚 are lifted to the original representation [𝐗new ]𝑛×𝑁 . The green data points in
Figure 1F are the new data samples generated when we impose the mean of 𝑆𝑎 to be 0.5 g. Shown in black in the figure is
the training data of the data set, which has a mean of 0.22 g. The relationship between the output response 𝑆𝐷𝑅𝑚𝑎𝑥 and
the input 𝑆𝑎 is learned and translated into the resulting marginal distribution of 𝑆𝐷𝑅𝑚𝑎𝑥 in the new data sample. The
mean of the log 𝑆𝐷𝑅𝑚𝑎𝑥 of the new sample is 0.029, versus 0.013 for the mean of the log 𝑆𝐷𝑅𝑚𝑎𝑥 for the entire training
data set. The example in this section is meant to illustrate in a very simple setting the essence of the PLoM surrogate
modeling process, which will be further illustrated for multiple output response quantities and more input variables, and
validated against “ground truth” test data sets in Sections 3 and 4.

2.4 Use cases

A key application of the proposed PLoM surrogate modeling method is estimating a set of structural response realiza-
tions based on a specified target distribution of input parameters. Hence, it is aligned closely with seismic risk analysis
and response estimation with design/modeling variability where the uncertainty of input parameters (𝐱𝑆 and 𝐱𝐺𝑀 ) are
explicitly considered and propagated to the response parameter (𝐱𝑅 ).
Two specific applications are considered in this study. The first involves simulation of structural responses with variable
model parameters and known (fixed) ground motions. Thus, 𝐱𝐺𝑀 is treated as fixed and variable structural model param-
eters 𝐱𝑆 will be considered in estimating seismic response 𝐱𝑅 to investigate alternative structural design and modeling
characteristics. The second involves the seismic response 𝐱𝑅 of a single structural model 𝐱𝑆 to ground motions 𝐱𝐺𝑀 with
variable spectral intensities, spectral shapes, and duration. This case is relevant to evaluating the response of a structure,
which is located at different sites. These two problem classes encompass common use cases and thus provide important
test cases for the proposed surrogate modeling framework.
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2413

F I G U R E 2 Use Case I (A) MDOF model and modeling parameters, (B) distributions of training data and three building variants, and
(C–E) comparisons of 500 training data realizations (blue) and 1000 test realizations (red, green, orange) of the seven design parameters for
each of three Case I buildings (red, green, orange).

TA B L E 1 Multiple-degree-of-freedom (MDOF) modeling parameters of training samples and three test building archetypes.
Modeling parameter Training Building 1 Building 2 Building 3
Distribution type Uniform Lognormal Lognormal Lognormal
Floor weight w (kips) [20, 300] 𝑚 = 100, 𝜎 = 25 𝑚 = 50, 𝜎 = 5 𝑚 = 150, 𝜎 = 15
Story stiffness 𝐾𝑥 (kip/in) [100, 1000] 𝑚 = 600, 𝜎 = 150 𝑚 = 500, 𝜎 = 50 𝑚 = 500, 𝜎 = 50
Story stiffness 𝐾𝑦 (kip/in) [100, 1000] 𝑚 = 500, 𝜎 = 100 𝑚 = 500, 𝜎 = 50 𝑚 = 500, 𝜎 = 50
Yield force 𝐹𝑦𝑥 (kips) [50, 500] 𝑚 = 300, 𝜎 = 30 𝑚 = 300, 𝜎 = 30 𝑚 = 200, 𝜎 = 25
Yield force 𝐹𝑦𝑦 (kips) [50, 500] 𝑚 = 200, 𝜎 = 40 𝑚 = 300, 𝜎 = 30 𝑚 = 200, 𝜎 = 25
Strength hardening ratio 𝛼𝑥 [0.02, 0.09] 𝑚 = 0.4, 𝜎 = 0.1 𝑚 = 0.3, 𝜎 = 0.05 𝑚 = 0.1, 𝜎 = 0.02
Strength hardening ratio 𝛼𝑦 [0.02, 0.09] 𝑚 = 0.4, 𝜎 = 0.1 𝑚 = 0.3, 𝜎 = 0.05 𝑚 = 0.1, 𝜎 = 0.02
Number of realizations 500 1000 1000 1000
Note: 𝑚 and 𝜎 stand for the median and standard deviation of a distribution, respectively.

3 ESTIMATING SEISMIC RESPONSE WITH DESIGN AND MODELING VARIABILITY

For the first use case with structural design and modeling variability, consider a multiple-degree-of-freedom (MDOF)
model representing a 12-story building (Figure 2), whose story force–displacement relationships are characterized by bi-
linear springs in two horizontal directions. Assuming that the spring properties and floor (masses) weights are constant
up the building height, the model has seven modeling parameters, namely, floor weight (𝑤), along with story stiffness (𝐾𝑥
and 𝐾𝑦 ), yield forces (𝐹𝑦𝑥 and 𝐹𝑦𝑦 ), and strain hardening ratios (𝛼𝑥 and 𝛼𝑦 ) of the nonlinear springs in the two horizontal
directions. Each individual combination of the seven parameters corresponds to one sample point in the design domain.
The seven modeling parameters of the building are assumed to follow a joint Gaussian distribution in logarithmic space,
which can be defined by its mean vector 𝐦7×1 and covariance matrix 𝚺7×7 .
The goal is to build a surrogate model that is capable of estimating the seismic response of different building archetypes
within the design domain. To construct this surrogate model, we first sample 500 realizations from uniform distributions of
the modeling parameters (Table 1) and then analyze these sampled building realizations under a suite of ground motions.
The analysis is performed using the NHERI-SimCenter EE-UQ application,53 which invokes OpenSees54 to perform the
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2414 ZHONG et al.

F I G U R E 3 Response analysis result for the multiple-degree-of-freedom (MDOF) models. (A) Response spectra of selected 100 ground
motions. (B) MDOF numerical model. (C) Maximum story drift ratio (𝑆𝐷𝑅𝑚𝑎𝑥 ) in the horizontal X direction for the first story.

NLRHA of the MDOF model. The input modeling parameters and output response quantities constitute the training data
set to develop the PLoM surrogate model.
To test the surrogate model, we construct three unique building designs whose mean and standard deviation of modeling
parameters are summarized in Table 1. The modeling parameters of building design 1 are assumed to related to one another
following the correlation coefficient matrix presented in Equation 1, whereas the modeling parameters for designs 2 and
3 are assumed to be statistically independent from each other. For each building, 1000 realizations are drawn from the
corresponding log-normal (joint) distributions (Figure 2), and the resulting structural models are analyzed using the same
suite of ground motions as the training set. The structural response statistics estimated from the NLRHA results (i.e., direct
simulation) of each of the three building models are used as “ground truth” against which the estimates determined using
the PLoM surrogate model are compared. It is important to observe that the surrogate is trained on the 500 input vectors
represented as blue points in Figure 2C–E (together with the associated NLRHA output), not on the 3 × 1000 = 3000 input
realizations (plus associated output) represented by the green, red, and orange contour lines in the figure.

𝑤 𝐾𝑥 𝐾𝑦 𝐹𝑦𝑥 𝐹𝑦𝑦 𝛼𝑥 𝛼𝑦
⎡1.0 0.0 0.0 0.0 0.0 0.0 0.0⎤ 𝑤
⎢ ⎥
⎢ 1.0 0.6 0.3 0.0 0.1 0.0⎥ 𝐾𝑥
⎢ 1.0 0.0 0.3 0.0 0.1⎥⎥ 𝐾𝑦

= ⎢ 0.2 0.0⎥
Bldg1
Σ7×7 1.0 0.2 𝐹𝑦𝑥 (1)
⎢ ⎥
⎢ sym. 1.0 0.0 0.2⎥ 𝐹𝑦𝑦
⎢ ⎥
⎢ 1.0 0.5⎥ 𝛼𝑥
⎢ ⎥
⎣ 1.0⎦ 𝛼𝑦

3.1 Benchmark direct simulation

The three test building archetypes are assumed to be located in Los Angeles, and we are interested in their structural
responses under an earthquake ground motion with a return period of 2475 years. Ground motion selection is conducted
to select and scale 100 pairs of records from the PEER NGA West 2 database55 whose average response spectrum matches
closely to the target 2475-year uniform hazard spectrum (UHS) at the site (Figure 3A). The selected ground motion records
are used to analyze the dynamic response of the MDOF model where the maximum story drift ratios (𝑆𝐷𝑅𝑚𝑎𝑥 ) and peak
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2415

F I G U R E 4 Prediction under PLoM constraints. (A) Comparison of probability density of floor weight in the training, test, and PLoM
realizations. (B) Error evolution of the Kullback–Leibler iteration on a log scale with 𝜖KL = 0.005.

floor accelerations (𝑃𝐹𝐴) in both horizontal directions constitute the 48 EDPs focused upon in this study (Figure 3B,
12 stories with four EDPs for each story). The 1st story 𝑆𝐷𝑅𝑚𝑎𝑥 in the horizontal X direction is plotted against different
design/modeling parameters in Figure 3C (including raw data for the training data set, and joint density contours for the
1000 realizations associated with each test building). The response data of the training building set (i.e., the blue dots),
because of its wide design domain, mostly covers the three test buildings. Note that the training data input values (plotted
on the ordinate) were uniformly sampled, and the data are plotted on log–log axes.

3.2 Training and prediction

The training data has a dimension of 𝑛 = 55 variables (including 𝑛𝑆 = 7 design parameters and 𝑛𝑅 = 48 EDPs) and 𝑁 =
500 sample points (Table 1). Thus, the training sample [𝔵0 ] is a matrix of size 55-by-500. The training dataset is used as the
initial data in the PLoM algorithm to train a surrogate model with the following tuning parameters: (1) preselected PCA
error threshold 𝜖PCA = 1 × 10−6 , (2) diffusion kernel parameter 𝜖k = 10, (3) diffusion basis error threshold 𝜖db = 1 × 10−6 ,
(4) tolerance for KL divergence 𝜖KL = 0.005, and (5) dissipative coefficient of the Wiener term 𝑓0 = 1.5. Note the PCA
error is selected so that 𝜈 = 𝑛, which is based on two considerations. First, the main goal of PCA is to normalize the data
and, thereby improve the numerical behavior when using multidimension Gaussian KDE, as opposed to trying to reduce
the problem dimension. Second, setting 𝜈 = 𝑛 can minimize the modification of local data structure from PCA when new
realizations are transformed back to the original space. Note, as discussed later, for the same reasons, the 𝜖PCA used in the
second example is also selected to retain all components. The diffusion basis error threshold resulted in a diffusion basis
dimension of m = 52, a 10-fold reduction from the original data sample. See Appendix A for further details.
The PLoM model, which is trained on uniform samples, for example, see floor weight training distribution in Figure 2B,
is used to generate new realizations for each test building by introducing constraints on the first moments of the seven
modeling parameters (as summarized in Table 1). The constraints are applied by solving the optimization problem whose
Lagrangian consists of the KL divergence between the pdf of the initial data and the space of pdfs that satisfy the first
moment constraints44 ; for example, for test building 2, Figure 4A contrasts the probability density functions of the floor
weight in the training (uniform distribution), the test set, and the PLoM generated dataset. The corresponding error
index history in the optimization problem, which must be solved to identify the Lagrange multipliers to properly define
the constrained nonparametric density, 𝑝𝐇𝑐 (𝐡), is plotted in Figure 4B. See Appendix A for additional details on the
error computation.

3.3 Results and comparisons

Figure 5A contrasts three (marginal) probability densities of the first-story X-direction 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 estimated from
the (1) training dataset (blue), (2) “ground truth” test building 1 response data (orange), and (3) PLoM prediction (black).
The joint distribution predicted by PLoM has very close marginal distributions to the “ground truth” and also well captures
the localized data concentration. Figure 5B–D summarizes the median and 25%–75% statistics of the PLoM predicted X
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2416 ZHONG et al.

F I G U R E 5 Comparison of structural responses from test building simulation and PLoM prediction. (A) First story 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴
distribution. (B) Median and 25%–75% box plots for 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 along the building height (Test Building 1). (C) 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 box
plots (Test Building 2). (D) 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 box plots (Test Building 3).

FIGURE 6 Runtime comparison: direct analysis versus training simulation and PLoM prediction.

and Y story responses (gray) along the building heights against the “ground truth” results for the three building archetypes
(orange, green, and red). As with the shown marginal for building archetype 1, the results are seen to be in good agreement.
Figure 6 compares on the same computational resources the run time of the two approaches: (1) response history analy-
sis (direct simulation) and (2) training set simulation plus PLoM prediction. The total run time of the direct simulation of
1000 test cases for each of the three test buildings is 20,208 CPU-sec on the hardware used. This is 4.6 times larger than the
entire PLoM prediction process of (4434 CPU-sec), which is dominated by the overhead cost of simulating the 500 training
points (3897 CPU-sec). Note, the difference in run time (i.e., the advantage of the PLoM prediction approach regarding the
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2417

TA B L E 2 Site-specific ground motion intensity measures (logarithmic mean).


San Francisco Los Angeles
Structure Site and return period (yr) 224 475 975 2475 4975 9950 224 475 975 2475 4975 9950
12-story 𝑇1 = 2.00 s 𝑆𝑎(𝑇1 ) (g) 0.10 0.14 0.20 0.28 0.35 0.43 0.17 0.26 0.37 0.55 0.71 0.90
𝑆𝑎𝑅𝑎𝑡𝑖𝑜 (0.04 s, 2 s, 2 s) 0.61 0.66 0.71 0.77 0.81 0.86 0.61 0.66 0.70 0.76 0.81 0.84
𝑆𝑎𝑅𝑎𝑡𝑖𝑜 (0.04 s, 2 s, 6 s) 1.72 1.78 1.85 1.94 2.02 2.09 1.64 1.71 1.77 1.85 1.92 1.98
𝐷𝑆5−75 (s) 12.7 13.9 15.0 16.1 17.2 18.5 8.9 8.6 8.4 8.2 8.3 8.4
62 63
Note: 𝑆𝑎𝑅𝑎𝑡𝑖𝑜 is computed using the ground motion model by Boore et al. 𝐷𝑆5−75 is computed using the ground motion model by Afshari and Stewart.

computational efficiency) will be even more pronounced when the PLoM surrogate model is applied to investigate large
inventories of buildings in regional earthquake risk analyses, for structural design optimization, or other applications that
require repeated evaluations of building models.

4 ESTIMATING SEISMIC RESPONSE WITH GROUND MOTION UNCERTAINTY

For considering ground motion as uncertain, we focus on a 12-story concrete moment frame, which was idealized and
modeled as a two-dimensional model in OpenSees.54 The building, originally developed in the FEMA P695 project,3 was
assumed to be designed per ASCE 7-05 for the Seismic Design Category D maximum (i.e., 𝑆𝑀1 = 0.9 g). The moment
frame system is modeled using concentrated plastic hinges in beams and columns, nonlinear beam–column joint panel
zones, and a leaning column to capture the 𝑃-Δ effects from the gravity system. The lumped plastic hinges employ the phe-
nomenological Ibarra–Medina–Krawinkler model,56 where the cyclic degradation parameters were calibrated to capture
the in-cycle and cyclic deterioration.57 The fundamental natural period of the frame model is 2.0 s.
Similar to the first problem case, comparisons are made between the “ground truth” result from direct response history
analysis and the prediction from a PLoM surrogate model trained on separate input parameters. In this case, ground
motion intensity measures, that is, spectral acceleration at the fundamental period 𝑆𝑎(𝑇1 ), significant duration 𝐷𝑆5−75 ,58
and response spectral shape measure 𝑆𝑎𝑅𝑎𝑡𝑖𝑜(𝑇𝐴 , 𝑇1 , 𝑇𝐵 ),59 are treated as input parameters to predict structural responses
(Section 4). Here, we train PLoM models on incremental dynamic analysis (IDA) data for predicting EDPs including
peak story drift ratios and peak floor accelerations at an arbitrary site. The training data are created by a set of IDA60
with a total of 1059 data points (49 ground motions applied at about 21 increments). Here, four ground motion intensity
measures (namely: spectral intensity, 𝑆𝑎; significant duration, 𝐷𝑆5−75 ; and spectral shape, 𝑆𝑎𝑅𝑎𝑡𝑖𝑜(0.02𝑇1 , 𝑇1 , 3𝑇1 ), and
𝑆𝑎𝑅𝑎𝑡𝑖𝑜(0.02𝑇1 , 𝑇1 , 𝑇1 )) are used to select IDA ground motions uniformly covering the expected range of the random
inputs. The set of 49 ground motions is selected based on a 7 × 7 grid to cover the range of significant duration and
spectral shape.32
To validate the effectiveness of the PLoM model, multiple stripe analysis (MSA) is conducted for two test site locations,
namely, Los Angeles (−118.26 W, 34.05 N, 𝑉𝑆,30 = 360 m/s) and San Francisco (−122.39 W, 37.79 N, 𝑉𝑆,30 = 873 m/s).
Table 2 summarizes the logarithmic mean values of the target ground motion intensity measures computed for the two
sites. Next, we perform 1200 runs of time history analysis, which are conducted following the multistripe analysis (MSA)61
method for the two sites (i.e., six intensity levels per site and 100 ground motions per intensity level). The mean, standard
deviation, and correlation coefficient of the EDPs from the MSA are then taken as the ground truth and compared with
the prediction by the PLoM model applied to these two sites using constrained sampling from the overall surrogate. For
this example, the random dimension is 𝑛 = 𝑛GM + 𝑛R = 28, where 𝑛GM = 4 and 𝑛R = 24. The number of nonsite-specific
IDA training points is 𝑁 = 1059 and the number of site-specific MSA generated test points is 996. The IDA data and the
MSA data are discussed in fuller detail in the next two sections.

4.1 Benchmark MSA (Ground Truth) results

For both test sites, seismic hazard disaggregation results are obtained from the USGS Unified Hazard Tool at six differ-
ent intensity levels with return periods from 224 to 9950 years. Based on the mean earthquake magnitude, distance, and
response spectral shape factor (𝜖), the conditional mean spectrum (CMS)64 and conditional significant duration65 are com-
puted as site-specific target intensity measures. At each intensity level, 100 ground motion records are selected15 to fit the
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2418 ZHONG et al.

F I G U R E 7 Multiple stripe analysis (MSA) site-specific ground motion selection. (A) Response spectra: 100 ground motions matching
the target conditional mean spectrum 𝑆𝑎𝑇𝑎𝑟𝑔𝑒𝑡 of each return periods (224, 475, 975, 2475, 4950, 9900 years)—the figure shows individual 100
records for the return period of 2475 years. (B) Significant duration (𝐷𝑆5−75 ): selected 100 ground motions versus site-specific target
distribution at 2475 years (LA: Los Angeles, SF: San Francisco).

F I G U R E 8 Key intensity measures 𝑆𝑎(2 s), 𝑆𝑎𝑅𝑎𝑡𝑖𝑜(0.04 s, 2 s, 6 s), 𝑆𝑎𝑅𝑎𝑡𝑖𝑜(0.04 s, 2 s, 2 s), and 𝐷𝑆5−75 of ground motion records used
in MSA (red, green, and orange) and IDA (blue). Note that the axes are log–log.

site-specific target intensity measures (see Figure 7). For the selected ground motion records, the 𝑆𝑎(2 s), 𝑆𝑎𝑅𝑎𝑡𝑖𝑜(0.04 s,
2 s, 6 s), 𝑆𝑎𝑅𝑎𝑡𝑖𝑜(0.04 s, 2 s, 2 s), and 𝐷𝑆5−75 are computed as the key intensity measures. Figure 8 presents the sample
calculations for the test sites at the return periods of 475 and 2475 years. Next nonlinear times history analyses are con-
ducted using the selected ground motion records at each intensity level. Overall 1200 simulations (two sites, six intensity
levels for each site, and 100 ground motion per intensity level) are completed in which the frame structure collapses in
204 simulations (i.e., infinite story drift ratio due to lateral instability). In the remaining 996 simulations, the maximum
story drift ratios and peak floor accelerations (24 EDPs) are recorded. Figure 9 plots sample maximum story drift ratio
(𝑆𝐷𝑅𝑚𝑎𝑥 ) of Story 1 and peak floor acceleration (𝑃𝐹𝐴) of Story 12 against two intensity measures of the corresponding
input ground motions (𝑆𝑎(2𝑠) and 𝑆𝑎𝑅𝑎𝑡𝑖𝑜).
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2419

FIGURE 9 MSA and IDA results: EDPs versus ground motion intensity measures.

4.2 Training IDA results

To generate training data, we consider the use of IDA. IDA is widely used for estimating probabilistic distributions of
structural responses under seismic excitation. The major difference between MSA and IDA is that instead of using unique
ground motion sets selected for individual earthquake intensity levels, the structure is analyzed using a generic set of
ground motion records in IDA. This difference with IDA renders the results usually problematic to interpret for a specific
site. In this study, IDA will be used to generate a training data set for a PLoM model, where the 12-story frame model
is analyzed using 49 ground motion records32 with gradually increasing 𝑆𝑎(2𝑠) intensity until the structure collapses in
the analysis. The other key intensity measures of these ground motions are also computed and plotted in Figure 8 (in
blue). These analyses result in 1059 data points in ℝ28 . In comparison with the MSA results, the EDP data from IDA have
wider distributions, as shown in Figure 9. This is because the input ground motions used in IDA do not represent the
seismic hazard at any specific site and their intensity measures (e.g., 𝑆𝑎𝑅𝑎𝑡𝑖𝑜) have larger ranges compared to the MSA
ground motions. The IDA results are used as the input training data for the PLoM surrogate to predict the site-specific
MSA results.
The tuning parameters of the PLoM algorithm are taken in a similar manner as the first example: (1) preselected PCA
error threshold 𝜖PCA = 1 × 10−6 , (2) diffusion kernel parameter 𝜖k = 50, (3) diffusion basis error threshold 𝜖db = 1 × 10−6 ,
(4) tolerance for KL divergence 𝜖KL = 0.005, and (5) dissipative coefficient of the Wiener term 𝑓0 = 1.5. The PCA tolerance
leaves 𝜈 = 𝑛. The diffusion basis tolerance results in 𝑚 = 31 for a 33-fold reduction in the size of the data space.

4.3 Results and discussion

The statistics of maximum story drift ratio (𝑆𝐷𝑅𝑚𝑎𝑥 ) and peak floor acceleration (𝑃𝐹𝐴) are computed and contrasted
between the training data set (raw IDA results), test data set (MSA results), and PLoM prediction. Figure 10A compares
the box-plots for 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 statistics for the Los Angeles site for a 2475-year return period event. The median EDPs
from the training data set (blue) are seen to be quite different from the MSA estimates (green), and the variation in the
training data set is much larger from the MSA data. The estimated median and variation of EDPs based on the PLoM
model (gray) are seen to be in good agreement with the ground truth MSA results. Shown in Figure 11 is an expanded set
of response statistics for the same 12-story frame subjected to ground motions representative of the 475-year and 2475-year
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2420 ZHONG et al.

F I G U R E 1 0 MSA result and PLoM prediction for Los Angeles site at 2475-year return period. (A) Statistics of 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴:
training (blue), MSA (green), and PLoM (gray). (B) Correlation coefficient matrix for 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴: MSA (ground truth) above the
diagonal, PLoM prediction below the diagonal.

FIGURE 11 Statistics of 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 at 475- and 2475-year return periods (MSA vs. PLoM). (A) Los Angeles site. (B) San Francisco
site.

hazards at building sites in Los Angeles and San Francisco. In these analyses, the agreement is likewise good between the
median values and dispersion from the MSA and PLoM analyses.
The joint distribution of EDPs, which is estimated from numerical simulations, is important for assessing the damage
and loss of the structure. The correlation coefficients between EDPs directly estimated from the MSA results are plotted
in the top-left triangle of Figure 10B against the correlation coefficients estimated from the PLoM prediction (bottom right
triangle). Good agreement is seen in the estimated correlation coefficients between 𝑆𝐷𝑅𝑚𝑎𝑥 s and correlations between
𝑃𝐹𝐴s at different stories, whereas the cross-correlation coefficients between 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 estimated from the PLoM
prediction can be relatively larger in some cases (e.g., the correlation between 𝑆𝐷𝑅1 and 𝑃𝐹𝐴12 ). The relative errors
between predicted and test EDP metrics (i.e., median vector, dispersion vector, and correlation coefficient matrix) are
computed by ‖(⋅)PLoM − (⋅)MSA ‖∕‖(⋅)MSA ‖, where (⋅) is to be replaced by median vector, dispersion vector, or correlation
coefficient matrix, resulting in the values of 0.05, 0.23, and 0.20, respectively. Overall, the median vector error is very good,
and the error in dispersion and correlations are reasonable, given the inherent variability in nonlinear response analyses.
To consider the combined uncertainties in the calculated EDPs (e.g., building drift) and the earthquake hazard, the risk
can be quantified by the mean annual frequency of exceeding a certain EDP level, 𝜆𝐸𝐷𝑃 (𝑥). The earthquake hazard can be
characterized by the seismic hazard curve that describes the mean annual frequency for response spectral acceleration (𝑆𝑎)
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2421

F I G U R E 1 2 Mean annual frequency of exceedance (A) 𝑆𝑎(𝑇 = 2.0 s). (B) Maximum story drift ratio in the building. (C) Maximum peak
floor acceleration in the building.

FIGURE 13 Influence of 𝜖𝑘 on relative error for predicting median, dispersion, and correlation coefficient.

to exceed different 𝑆𝑎 levels (𝜆𝑆𝑎 ). For instance, Figure 12A plots the seismic hazard curve for 𝑆𝑎(2𝑠) at the two case study
sites (San Francisco and Los Angeles). The 𝜆𝐸𝐷𝑃 (𝑥) can be evaluated by Equation 2 where the probability of exceeding a
certain 𝐸𝐷𝑃 level at a given 𝑆𝑎, 𝑃(𝐸𝐷𝑃 ≥ 𝑥|𝑆𝑎), considers both the noncollapse (NC) and collapse (C) cases (Equation 3).
In this example, the collapse probability 𝑃(C|𝑆𝑎) and noncollapse probability 𝑃(NC|𝑆𝑎) are estimated from the MSA, and
the exceedance probability conditional on the noncollapse case, 𝑃(𝐸𝐷𝑃 ≥ 𝑥|NC, 𝑆𝑎), can be evaluated from the MSA as
well as the PLoM prediction. Figure 12B,C contrasts the annual frequency of exceedance curves computed based on the
MSA ground truth result and PLoM prediction. Overall, the story drift and floor acceleration exceedance curves compare
well between the MSA and PLoM analyses.

| 𝑑𝜆𝑆𝑎 (𝑆𝑎) |
𝜆𝐸𝐷𝑃 (𝑥) = [𝑃(𝐸𝐷𝑃 ≥ 𝑥|𝑆𝑎)]|| |𝑑(𝑆𝑎)
| (2)
∫𝑆𝑎 | 𝑑(𝑆𝑎) |
𝑃(𝐸𝐷𝑃 ≥ 𝑥|𝑆𝑎) = 𝑃(𝐸𝐷𝑃 ≥ 𝑥|NC, 𝑆𝑎) × 𝑃(NC|𝑆𝑎) + 𝑃(C|𝑆𝑎) (3)

We also explore the influence of different tuning parameters on the prediction accuracy via parametric studies. For each
tuning parameter, the aforementioned three relative error measures are computed from the corresponding prediction as
the parameter is varied. Two tuning parameters along with a data filter are found to be very relevant to the prediction
accuracy. The two tuning parameters are the diffusion kernel length scale 𝜖k and the tolerance for KL divergence 𝜖KL .
Figure 13 plots the relative error curves versus 𝜖k where two major findings are: (1) the mean prediction is stable and
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2422 ZHONG et al.

FIGURE 14 Influence of 𝜖KL on relative error for predicting median EDPs.

FIGURE 15 Influence of data filter (number of dispersion around the target center) on relative prediction errors.

not sensitive to the kernel parameter; (2) the error of predicting dispersion and the correlation coefficient decreases as 𝜖k
increases and converges when 𝜖k ≥ 20. Typically, models with smaller 𝜖k values tend to view training data more separate,
which would overlook the global data trend, while further increasing 𝜖𝑘 to large numbers (e.g., greater than 1000 in this
case study example) does not help improve the accuracy and would increase the error if there are localized geometric
structures in the data as larger 𝜖𝑘 values guide the model to view the data more collectively.
In contrast, we find that the accuracy of the median prediction is quite dependent on the KL divergence tolerance.
Figure 14 plots the relative median prediction error versus 𝜖KL where the error tends to converge when 𝜖KL ≤ 0.02. We
also notice that the dispersion and correlation coefficient prediction is not very sensitive to 𝜖KL . In addition, when the
goal is to only predict a localized data set, applying a filter by only including the data falling within ±𝑘𝜎 around the target
prediction center could significantly improve the accuracy of predicting the dispersion and correlation coefficient. For
instance, Figure 15 plots the relative prediction errors versus the data filter, namely the number of standard deviations 𝑘.
It indicates (1) either a too wide or too narrow data filter would increase the prediction error, and (2) an optimal 𝑘 value
ranges from 1.5 to 3 for different combinations of site and return period.

5 SUMMARY AND CONCLUSIONS

In this paper, PLoM is used as an alternative method of predicting seismic response of building structures. The PLoM algo-
rithm proposed by Soize and Ghanem41,44 is implemented into an open-source Python package, PLoM66 and SimCenter
tools,53 thus freely available to the research community and built-in to the NHERI SimCenter tool chain.67 The algorithm
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2423

consists of five major components: (1) PCA to decorrelate the random variables; (2) KDE for probability density approxi-
mation; (3) Itô stochastic differential equation for sample modeling; (4) Diffusion-map projection for response manifold
structure preservation; and (5) MCMC simulation for sample generation.
Two use cases are provided to validate and demonstrate the power of the PLoM software package applied to seismic
response estimation. The first use case investigates the use of PLoM for predicting responses of buildings with different
design parameters. The 12-story archetype building is idealized and modeled as a MDOF system with seven design param-
eters, and its story EDPs (𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴) are of engineering interest. The training data set is designed to cover a wide
range of each design parameter (Table 1). The PLoM algorithm learns the training data (500 data points, 55 variables) and
predicts seismic responses of three test buildings with differently constrained design/modeling parameters. The joint dis-
tribution of EDPs predicted by the PLoM model is found to be in good agreement with the “ground truth” result estimated
from direct simulation of the building responses (1000 data points for each test building). For this example, the PLoM cal-
culations are about five times faster than the direct NLRHA simulation approach (including training of the PLoM model),
and applying the PLoM surrogate model to large building inventories or repeated design optimization problems would be
orders of magnitude faster than direct simulation.
The second use case investigates the use of a PLoM model for predicting building responses under site-specific seis-
mic hazards. A 12-story reinforced concrete moment frame is used as the archetype structure. MSA is conducted for the
archetype structure at two sites, Los Angeles and San Francisco, at six intensity levels under 100 ground motions for each.
In this use case, the PLoM models are trained using noncollapse IDA data (1059 data points, 28 variables) for EDPs includ-
ing maximum story drift ratio (𝑆𝐷𝑅𝑚𝑎𝑥 ) and peak floor acceleration (𝑃𝐹𝐴). The trained model is then used to generate
new realizations with the site-specific constraints on the major ground motion intensity measures (𝑆𝑎, 𝑆𝑎𝑅𝑎𝑡𝑖𝑜, and
𝐷𝑆5−75 ). The predicted median, dispersion, and correlation coefficients of EDPs are found in good agreement with the
estimates based on MSA data (996 data points). Similarly, good agreement is seen in the mean annual exceedance curves
of 𝑆𝐷𝑅𝑚𝑎𝑥 and 𝑃𝐹𝐴 between the MSA and PLoM prediction.
Along with the observed agreement on EDP distributions between the PLoM surrogate model and direct simula-
tion by NLRHA, as well as the promising computational efficiency of the PLoM approach, there are several topics
that could be further investigated and explored. First, the design/modeling variability use case examined in this study
has seven structural modeling parameters, which can be expanded to include more detailed parameters, that is, the
application using the PLoM method can be scaled to develop surrogate models for more complex structural systems
and examine their prediction accuracy and efficiency (the computational cost of those more complex models can be
longer and the efficiency of the PLoM approach can be even more valuable). Second, based on the parametric study
on the PLoM tuning parameters (𝜖k , 𝜖KL , and data filter size), we recommend to use (1) a relatively large 𝜖k (not less
than 20), (2) a smaller 𝜖KL (not larger than 0.02), and (3) 1.5 to 3.0 number of dispersion for the data filter size to
reduce the prediction error; however, these tuning parameters may vary for different structural systems with unique
response characteristics. Hence, future investigations can systematically explore and optimize them for common struc-
tural systems. Moreover, the strength of the PLoM algorithm in capturing the local data concentration and correlation
between structural responses can be very useful to improve the accuracy and efficiency of damage and loss assess-
ment. Hence, the application of PLoM to predict damage and loss under earthquakes can be explored and validated.
Finally, future studies are suggested to extend this study by investigating PLoM surrogate models that simultaneously
consider both the design variables (including discrete design parameters, e.g., number of stories) and ground motion
characteristics as input predictors for EDPs. This is related to the usage of PLoM surrogate models in regional earth-
quake simulations which are known as computational costly problems and invoke both ground motion and structural
design/modeling uncertainty.

AC K N OW L E D G M E N T S
This material is based upon work supported by the National Science Foundation under Grant (CMMI-1612843 and 2131111).
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the authors and do not
necessarily reflect the views of the National Science Foundation. The CFIS (Centro de Formación Interdisciplinaria Supe-
rior) is also acknowledged for funding support. The authors would also like to thank Dr. Roger Ghanem for discussions
on the PLoM algorithm.

D A T A AVA I L A B I L I T Y S T A T E M E N T
The numerical models and NLRHA results are published and available in the DesignSafe Data Depot as project PRJ-
3670 https://www.designsafe-ci.org/data/browser/public/designsafe.storage.published/PRJ-3670. The source code of the
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2424 ZHONG et al.

PLoM algorithm is available via the github page (https://github.com/sanjayg0/PLoM), and the integration of the PLoM
module is also available in NHERI-SimCenter applications EE-UQ (https://simcenter.designsafe-ci.org/research-tools/
ee-uq-application/) and quoFEM (https://simcenter.designsafe-ci.org/research-tools/quofem-application/).

ORCID
Kuanshi Zhong https://orcid.org/0000-0003-2418-9744
Sanjay Govindjee https://orcid.org/0000-0003-0711-3633

REFERENCES
1. Cornell CA, Krawinkler H. Progress and challenges in seismic performance assessment. PEER Newsletter. 2000.
2. Moehle J, Deierlein GG. A framework methodology for performance-based earthquake engineering. In: 13th World Conference on
Earthquake Engineering, 13 WCEE, Vancouver, Canada. 2004.
3. Applied Technology Council. Quantification of Building Seismic Performance Factors. Technical Report FEMA 695. Federal Emergency
Management Agency, Washington, D.C.; 2009.
4. Applied Technology Council. Seismic Performance Assessment of Buildings Volume 1 - Methodology. Technical Report FEMA P58-1. Federal
Emergency Management Agency, Washington, D.C.; 2012.
5. Applied Technology Council. Seismic Performance Assessment of Buildings Volume 2 - Implementation Guide. Technical Report FEMA
P58-2. Federal Emergency Management Agency, Washington, D.C.; 2012.
6. Goulet CA, Haselton CB, Mitrani-Reiser J, et al. Evaluation of the seismic performance of a code-conforming reinforced-concrete frame
building—from seismic hazard to collapse safety and economic losses. Earthq Eng Struct Dyn. 2007;36(13):1973-1997.
7. Kramer SL. Geotechnical Earthquake Engineering. Pearson Education; 1996.
8. Yang T, Moehle J, Stojadinovic B, Der Kiureghian A. Seismic performance evaluation of facilities: methodology and implementation. J
Struct Eng. 2009;135(10):1146-1154.
9. Celik OC, Ellingwood BR. Seismic fragilities for non-ductile reinforced concrete frames—role of aleatoric and epistemic uncertainties.
Struct Saf. 2010;32(1):1-12.
10. Gokkaya BU, Baker JW, Deierlein GG. Quantifying the impacts of modeling uncertainties on the seismic drift demands and collapse risk
of buildings with implications on seismic design checks. Earthq Eng Struct Dyn. 2016;45(10):1661-1683.
11. Liel AB, Haselton CB, Deierlein GG, Baker JW. Incorporating modeling uncertainties in the assessment of seismic collapse risk of buildings.
Struct Saf. 2009;31(2):197-211.
12. Padgett JE, DesRoches R. Sensitivity of seismic response and fragility to parameter uncertainty. J Struct Eng. 2007;133(12):1710-1718.
13. Baker JW, Cornell AC. Spectral shape, epsilon and record selection. Earthq Eng Struct Dyn. 2006;35(9):1077-1095.
14. Baker JW, Lee C. An improved algorithm for selecting ground motions to match a conditional spectrum. J Earthq Eng. 2018;22(4):708-723.
15. Jayaram N, Lin T, Baker JW. A computationally efficient ground-motion selection algorithm for matching a target response spectrum
mean and variance. Earthq Spectra. 2011;27(3):797-815.
16. Kohrangi M, Vamvatsikos D, Bazzurro P. Site dependence and record selection schemes for building fragility and regional loss assessment.
Earthq Eng Struct Dyn. 2017;46(10):1625-1643.
17. Kwong NS, Chopra AK, McGuire RK. A framework for the evaluation of ground motion selection and modification procedures. Earthq
Eng Struct Dyn. 2015;44(5):795-815.
18. Xie Y, Ebad Sichani M, Padgett JE, DesRoches R. The promise of implementing machine learning in earthquake engineering: a state-of-
the-art review. Earthq Spectra. 2020;36(4):1769-1801.
19. Burton H, Xu H, Yi Z. Design of computer experiments for developing seismic surrogate models. Earthq Spectra. 2021;38:87552930211033309.
20. Du A, Padgett JE. Investigation of multivariate seismic surrogate demand modeling for multi-response structural systems. Eng Struct.
2020;207:110210.
21. Ghosh J, Padgett JE, Dueñas-Osorio L. Surrogate modeling and failure surface visualization for efficient seismic vulnerability assessment
of highway bridges. Probab Eng Mech. 2013;34:189-199.
22. Goda K, Tesfamariam S. Multi-variate seismic demand modelling using copulas: application to non-ductile reinforced concrete frame in
Victoria, Canada. Struct Saf. 2015;56:39-51.
23. Gudipati VK, Cha EJ. Surrogate modeling for structural response prediction of a building class. Struct Saf. 2021;89:102041.
24. Kyprioti AP, Taflanidis AA. Kriging metamodeling for seismic response distribution estimation. Earthq Eng Struct Dyn. 2021;50(13):3550-
3576.
25. Mangalathu S, Heo G, Jeon JS. Artificial neural network based multi-dimensional fragility development of skewed concrete bridge classes.
Eng Struct. 2018;162:166-176.
26. Yi Sr, Wang Z, Song J. Bivariate Gaussian mixture-based equivalent linearization method for stochastic seismic analysis of nonlinear
structures. Earthq Eng Struct Dyn. 2018;47(3):678-696.
27. Zou J, Welch D, Zsarnoczay A, Taflanidis A, Deierlein G. Surrogate modeling for the seismic response estimation of residential wood frame
structures. In: Proceedings of the 17th World Conference on Earthquake Engineering, 17 WCEE, Sendai, Japan. 2020.
28. Mackie K, Stojadinović B. Probabilistic seismic demand model for California highway bridges. J Bridge Eng. 2001;6(6):468-481.
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2425

29. Jeon JS, Mangalathu S, Song J, Desroches R. Parameterized seismic fragility curves for curved multi-frame concrete box-girder bridges
using Bayesian parameter estimation. J Earthq Eng. 2019;23(6):954-979.
30. Pan Y, Agrawal AK, Ghosn M. Seismic fragility of continuous steel highway bridges in New York State. J Bridge Eng. 2007;12(6):689-699.
31. Huang H, Burton HV. Classification of in-plane failure modes for reinforced concrete frames with infills using machine learning. J Build
Eng. 2019;25:100767.
32. Zhong K, Chandramohan R, Baker JW, Deierlein GG. Site-specific adjustment framework for incremental dynamic analysis (SAF-IDA).
Earthq Spectra. 2022;38:1893–1917.
33. Mangalathu S, Jeon JS, DesRoches R. Critical uncertainty parameters influencing seismic performance of bridges using Lasso regression.
Earthq Eng Struct Dyn. 2018;47(3):784-801.
34. Gidaris I, Taflanidis AA, Mavroeidis GP. Kriging metamodeling in seismic risk assessment based on stochastic ground motion models.
Earthq Eng Struct Dyn. 2015;44(14):2377-2399.
35. Mitropoulou CC, Papadrakakis M. Developing fragility curves based on neural network IDA predictions. Eng Struct. 2011;33(12):3409-3421.
36. Liu Z, Zhang Z. Artificial neural network based method for seismic fragility analysis of steel frames. KSCE J Civ Eng. 2018;22(2):708-717.
37. Mangalathu S, Jeon JS. Stripe-based fragility analysis of multispan concrete bridge classes using machine learning techniques. Earthq Eng
Struct Dyn. 2019;48(11):1238-1255.
38. Vaswani A, Shazeer N, Parmar N, et al. Attention is all you need. In: Guyon I, Luxburg UV, Bengio S, et al., eds. Advances in Neural
Information Processing Systems. Vol. 30. Curran Associates, Inc.; 2017.
39. Ni P, Sun L, Yang J, Li Y. Multi-end physics-informed deep learning for seismic response estimation. Sensors. 2022;22(10):3697.
40. Zhang R, Chen Z, Chen S, Zheng J, Büyüköztürk O, Sun H. Deep long short-term memory networks for nonlinear structural seismic
response prediction. Comput Struct. 2019;220:55-68.
41. Soize C, Ghanem R. Data-driven probability concentration and sampling on manifold. J Comput Phys. 2016;321:242-258.
42. Soize C, Ghanem R, Safta C, et al. Entropy-based closure for probabilistic learning on manifolds. J Comput Phys. 2019;388:518-533.
43. Soize C, Ghanem R. Polynomial chaos representation of databases on manifolds. J Comput Phys. 2017;335:201-221.
44. Soize C, Ghanem R. Physics-constrained non-Gaussian probabilistic learning on manifolds. Int J Numer Methods Eng. 2020;121(1):110-145.
45. Soize C, Ghanem RG, Desceliers C. Sampling of Bayesian posteriors with a non-Gaussian probabilistic learning on manifolds from a small
dataset. Stat Comput. 2020;30:1433-1457.
46. Wan Z, Chen J, Beer M. Pathways for uncertainty quantification through stochastic damage constitutive models of concrete. In: 13th
International Conference on Application of Statistics and Probability in Civil Engineering, ICASP13, Seoul, South Korea. 2019.
47. Wu L, Nguyen VD, Adam L, Noels L. An inverse micro-mechanical analysis toward the stochastic homogenization of nonlinear random
composites. Comput Methods Appl Mech Eng. 2019;348:97-138.
48. Guilleminot J, Dolbow JE. Data-driven enhancement of fracture paths in random composites. Mech Res Commun. 2020;103:103443.
49. Almeida JO, Rochinha FA. A probabilistic learning approach applied to the optimization of wake steering in wind farms. J Comput Inf Sci
Eng. 2022;23(1):011003.
50. Upadhyay K, Giovanis DG, Alshareef A, et al. Data-driven uncertainty quantification in computational human head models. Comput
Methods Appl Mech Eng. 2022;398:115108.
51. Coifman RR, Lafon S, Lee AB, et al. Geometric diffusions as a tool for harmonic analysis and structure definition of data: diffusion maps.
Proc Natl Acad Sci. 2005;102(21):7426-7431.
52. Peng Y, Zhou T, Li J. Surrogate modeling immersed probability density evolution method for structural reliability analysis in high
dimensions. Mech Syst Sig Process. 2021;152:107366.
53. McKenna F, Zhong K, Gardner M, Zsarnoczay A, Wang C, Elhaddad W. NHERI-SimCenter/EE-UQ: version 3.0.0 (v3.0.0). Zenodo; 2022.
54. Mazzoni S, McKenna F, Scott MH, et al. OpenSees Command Language Manual. Pacific Earthquake Engineering Research (PEER) Center;
2006:264.
55. Ancheta TD, Darragh RB, Stewart JP, et al. NGA-West2 database. Earthq Spectra. 2014;30(3):989-1005.
56. Ibarra LF, Medina RA, Krawinkler H. Hysteretic models that incorporate strength and stiffness deterioration. Earthq Eng Struct Dyn.
2005;34(12):1489-1511.
57. Haselton CB, Liel AB, Lange ST, Deierlein GG. Beam-Column Element Model Calibrated for Predicting Flexural Response Leading to Global
Collapse of RC Frame Buildings. Pacific Earthquake Engineering Research Center; 2008.
58. Bommer JJ, Martinez-Pereira A. The effective duration of earthquake strong motion. J Earthq Eng. 1999;3(02):127-172.
59. Eads L, Miranda E, Lignos DG. Spectral shape metrics and structural collapse potential. Earthq Eng Struct Dyn. 2016;45(10):1643-1659.
60. Vamvatsikos D, Cornell CA. Incremental dynamic analysis. Earthq Eng Struct Dyn. 2002;31(3):491-514.
61. Jalayer F. Direct Probabilistic Seismic Analysis: Implementing Non-linear Dynamic Assessments. PhD thesis. Stanford University, Stanford,
CA; 2003.
62. Boore DM, Stewart JP, Seyhan E, Atkinson GM. NGA-West2 equations for predicting PGA, PGV, and 5% damped PSA for shallow crustal
earthquakes. Earthq Spectra. 2014;30(3):1057-1085.
63. Afshari K, Stewart JP. Physically parameterized prediction equations for significant duration in active crustal regions. Earthq Spectra.
2016;32(4):2057-2081.
64. Lin T, Baker JW. Introducing adaptive incremental dynamic analysis: a new tool for linking ground motion selection and structural
response assessment. In: Proceedings of 11th International Conference on Structural Safety & Reliability (ICOSSAR 2013), New York,
NY. The International Association for Structural Safety and Reliability. 2013.
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2426 ZHONG et al.

65. Chandramohan R, Baker JW, Deierlein GG. Quantifying the influence of ground motion duration on structural collapse capacity using
spectrally equivalent records. Earthq Spectra. 2016;32(2):927-950. doi: https://doi.org/10.1193/122813eqs298mr2
66. Zhong K, Navarro JG, Govindjee S. PLoM python package v1.0. 2021. Accessed January 31, 2023. https://github.com/sanjayg0/PLoM
67. Deierlein GG, McKenna F, Zsarnóczay A, et al. A cloud-enabled application framework for simulating regional-scale impacts of natural
hazards on the built environment. Frontiers Built Environ. 2020;6:558706.
68. Ortiz JB, Kelli M, Grant MJ, et al. Trajectory Optimization via Unsupervised Probabilistic Learning on Manifolds. Technical Report
SAND2021-11867. Sandia National Laboratories (SNL-NM), Albuquerque, NM, United States; Sandia National Laboratories, SNL
California; Livermore, CA; 2021.

How to cite this article: Zhong K, Navarro JG, Govindjee S, Deierlein GG. Surrogate modeling of structural
seismic response using probabilistic learning on manifolds. Earthquake Engng Struct Dyn. 2023;52:2407–2428.
https://doi.org/10.1002/eqe.3839

APPENDIX A: PLOM METHODOLOGY


This section provides more specificity of the basic PLoM methodology, somewhat following Safta et al.68 We then add a
description for how one can enforce constraints on the Markov Chain Monte Carlo (MCMC) Itô sampler that is embedded
in the PLoM algorithm:

1. First, given the possible general heterogeneity of data, we assume throughout that the realization [𝔵0 ] has already been
scaled such that each row takes on nondimensional values between 0 and 1. This operation is linear and invertible,
allowing for reconstruction of dimensional values as a post-processing step at the end of the PLoM algorithm (not
explicitly discussed in what follows).
2. The next step in the PLoM methodology is to apply PCA in order to help decorrelate the data, a step which improves the
performance of later steps in the PLoM procedure. For PCA, one first computes the 𝑛-by-𝑛 correlation matrix of the data
1 ∑𝑁 0 1 ∑𝑁
[𝐂] = ̄ 0 (𝑙) − 𝐱)
𝑙=1 (𝐱 (𝑙) − 𝐱)(𝐱 ̄ 𝑇 , where 𝐱̄ = 0 0
𝑙=1 𝐱 (𝑙) and 𝐱 (𝑙) is realization 𝑙 of the random vector 𝐗, that
𝑁−1 𝑁
is, the 𝑙 th data point. The dominant 𝜈 eigenvalues of the correlation matrix are next arranged in a diagonal matrix [𝛍]
and the corresponding eigenvectors placed in an 𝑛-by-𝜈 matrix [𝚽]. If one now defines the 𝑛-by-𝑁 matrix [̄𝔵], whose
identical columns are given by 𝐱, ̄ then one can define a 𝜈-by-𝑁 random matrix [ℌ] = [𝛍]−1∕2 [𝚽]𝑇 ([𝔛] − [̄𝔵]). The
columns of [ℌ] are formed by identical copies of a random vector that we denote as 𝐇 with realizations 𝐡. The matrix-
valued realizations of [ℌ] are denoted as [𝔥]. Using the realization [𝔥0 ] = [𝛍]−1∕2 [𝚽]𝑇 ([𝔵0 ] − [̄𝔵]), the selection of 𝜈 is
made such that the recovery of the data in the original space, namely, [𝔵0 ]rec = [̄𝔵] + [𝚽][𝛍]1∕2 [𝔥0 ], has an expectation
error below a pre-selected threshold 𝜖PCA . Note that the empirical mean of the columns of [𝔥0 ] is the zero vector and
that the empirical correlation matrix of these same columns is the 𝜈-by-𝜈 identity matrix.
3. The next step in the PLoM algorithm is to construct a nonparametric probability density for the random matrix [ℌ]
using the initial data. This is accomplished by first creating a kernel density estimate (KDE) for the random vector 𝐇.
Omitting the details, this results in an explicit expression 𝑝𝐇 (𝐡) in terms of sums of exponentials depending on the 𝜈-by-
1 vector-valued realizations 𝐡0 (1), 𝐡0 (2), … , 𝐡0 (𝑁). The joint probability distribution 𝑝[ℌ] ([𝔥]) for [ℌ] is then formed via
a product of 𝑁 copies of 𝑝𝐇 (𝐡). In our implementation, we utilize a Gaussian kernel involving two scalar parameters,
one sets the bandwidth to the optimal Silverman value and the other is set to ensure compatibility with the empirical
correlation structure of the reduced and transformed data sample [𝔥0 ].44 Thus there are no user adjustable parameters
in this step. The use of Gaussian kernels is helpful for the later use of the KDE approximation when generating new
samples.
4. An important property of the data sample is how the vector-valued realizations 𝐡0 (1), 𝐡0 (2), … , 𝐡0 (𝑁) are concentrated
in ℝ𝜈 . In how PLoM generates new samples of [ℌ] and hence [𝔛], it attempts to preserve this concentration. It does
so using the concept of diffusion maps and diffusion basis vectors. In particular, we represent the random 𝜈-by-𝑁
matrix [ℌ] by a second random matrix [ℨ] of dimension 𝜈-by-𝑚, where 𝑚 < 𝑁 (often much less so). The two random
matrices are related to each other by the relation [ℌ] = [ℨ][𝐠]𝑇 , where [𝐠] is a deterministic matrix, whose columns are
called the diffusion basis vectors and are computed from the data sample [𝔥0 ]. Realizations of the random matrix [ℨ]
generate realizations of [ℌ] that preserve any clustering present in the 𝐇 representation of the data sample. Assuming
invertability of the 𝑚-by-𝑚 matrix [𝐠]𝑇 [𝐠], we can also write [ℌ][𝐚] = [ℨ], where [𝐚] = [𝐠]([𝐠]𝑇 [𝐠])−1 can be thought
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ZHONG et al. 2427

of as a Moore–Penrose inverse, though in the PLoM methodology it is more properly understood to arise out of the use
of a Galerkin projection as part of Step 5.
The idea that is exploited to construct [𝐠] is to consider the data points in ℝ𝜈 as the 𝑁 vertices of a fully connected
graph. From this graph, one builds a 𝑁-by-𝑁 diffusion matrix [𝐊], such that the (𝑙, 𝑚) entry is given by 𝑘(𝐡0 (𝑙), 𝐡0 (𝑚)),
where 𝑘(⋅, ⋅) is a symmetric diffusion kernel that defines a distance metric between nodes on the graph. In our imple-
1
mentation, 𝑘(𝐡0 (𝑙), 𝐡0 (𝑚)) = exp(− ‖𝐡0 (𝑙) − 𝐡0 (𝑚)‖2 ), where 𝜖k is a user-selected parameter that controls the scale
4𝜖k
of the metric. If the rows of [𝐊] are rescaled to each have a row-sum of 1, then we arrive at a matrix, which we denote
as [𝐏]. This matrix functions as a transition probability matrix on the graph and is related to the graph Laplacian. The
right-eigenvectors and eigenvalues of [𝐏] serve to cluster the data. By this we mean that for points near each other
on the graph (as measured by the diffusion kernel), they will have entries in the eigenvectors that are close in value
when viewed at a fixed resolution. The largest eigenvalue Λ1 = 1 and the corresponding vector is a vector of constant
values; this is the coarsest resolution of the graph and all points are considered to be near each other. The remaining
vectors, with progressively decreasing eigenvalues Λ2 > Λ3 > ⋯ > Λ𝑁 , separate points on the graph in progressively
finer clusters. The columns of the 𝑁-by-𝑚 matrix [𝐠] are the first 𝑚 eigenvectors of [𝐏], where we start counting from
the second eigenvector. 𝑚 is properly selected so that the normalized error between the correlation matrix [𝐂] of the
original data sample and the correlation matrix generated from the recovered data sample after both PCA projection
and the diffusion basis projection, [̄𝔵] + [𝚽][𝛍]1∕2 [𝔥0 ][𝐚][𝐠]𝑇 , is below a specified error tolerance, 𝜖db . As a simple rule,
one can select 𝑚 ≥ 3 such that Λ𝑚 ∕Λ2 < 0.1.
5. The final step in the basic implementation of PLoM is to generate samples of the 𝜈-by-𝑚 random matrix [ℨ]. Each
such sample [𝔷new ] can then be used to construct a sample of [𝔛], namely, [𝔵new ] = [̄𝔵] + [𝚽][𝛍]1∕2 [𝔷new ][𝐠]𝑇 . Noting
that 𝑝[ℌ] ([𝔥]) is known, it is possible to construct a stochastic dissipative Hamiltonian dynamical system whose tra-
jectories properly sample 𝑝[ℌ] ([𝔥]). To preserve, however, the concentration of measure seen in [𝔵0 ], these stochastic
differential equations are projected onto the diffusion basis [𝐠]; that is, a Galerkin projection is utilized. The resulting
Itô equations that are numerically solved using a Störmer–Verlet scheme are

𝑑[(𝑠)] = 𝑑[(𝑠)]𝑑𝑠 (A1)

1 √
𝑑[(𝑠)] = []([(𝑠)])𝑑𝑠 − 𝑓0 [(𝑠)]𝑑𝑠 + 𝑓0 [𝑑(𝑠)] , (A2)
2

where the initial conditions are [(0)] = [𝔥0 ][𝐚] and [(0)] = [ ][𝐚] where [ ] is a 𝜈-by-𝑁 random matrix whose
columns are standard normal vectors in ℝ𝜈 . The parameter 𝑓0 is a user selected parameter to control the dissipative
nature of the dynamics and [𝑑(𝑠)] = [𝑑𝔚(𝑠)][𝐚] is a matrix-valued projection of [𝑑𝔚(𝑠)] whose 𝑁 columns are
independent increments of a Wiener process in ℝ𝜈 . The matrix-valued driver of the velocities is the projection onto
the diffusion basis of the unconstrained system’s potential gradient, namely, []([(𝑠)]) = ∇ ln 𝑝[ℌ] ([(𝑠)][𝐠]𝑇 )[𝐚].
Note the gradient is with respect to the entire argument of the probability distribution. The numerical solution of these
equations produces a Markov Chain with the desired distribution in an efficient manner strictly based of the initial 𝑁
samples in ℝ𝑛 . Samples from this chain are now available to compute statistics of the system of interest.
6. In the case that we wish to sample the response manifold while simultaneously imposing constraints on the samples so
that we are trying to sample a random variable 𝐗𝑐 instead of 𝐗, then an iterative but efficient modification of the above
five steps can be employed. The technical details of the process as outlined by Soize and Ghanem44 are somewhat
involved but the general concept is modestly straightforward. If we wish instead to generate samples, for the same
system for which we have an initial sample, but under a vector-valued constraint of the form ∫ℝ𝜈 𝐜(𝐡)𝑝𝐇𝑐 (𝐡) 𝑑𝐡 = 𝛄,
where 𝛄 ∈ ℝ𝑐 is given, then we can mostly utilize the procedure outline above. The primary change however is that the
underlying KDE density 𝑝𝐇 (𝐡) needs to be modified to account for the constraint. We will denote the “constrained”
density as 𝑝𝐇𝑐 (𝐡) and it will be determined by minimizing the Kullback–Leibler (KL) divergence with respect to the
KDE density. Thus, 𝑝𝐇𝑐 (𝐡) is taken as

𝑝𝐇𝑐 (𝐡) = arg min 𝑞(𝐡) ln (𝑞(𝐡)∕𝑝𝐇 (𝐡)) 𝑑𝐡 , (A3)


𝑞(𝐡) ∫ℝ𝜈

where the minimization is carried out over the set of normalized densities that satisfy the given vector-valued
constraint; throughout, we assume that the underlying scalar-valued constraints are independent. The constrained
minimization is effected by the use of Lagrange multipliers (𝜆, 𝛌) ∈ ℝ1+𝑐 to impose the normalization condition and the
10969845, 2023, 8, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3839, Wiley Online Library on [13/06/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2428 ZHONG et al.

vector constraint. The KL divergence functions to some degree as a distance metric on the space of probability measures,
but it should be recognized that it is neither symmetric, nor does it satisfy the triangle inequality. Notwithstanding,
minimizing the KL divergence determines a density close to the original density but satisfying the desired constraints.
Taking the variation of Equation A3 with respect to 𝑞(𝐡) yields an expression for the desired density as

𝑝𝐇𝑐 (𝐡) = 𝑝𝐇 (𝐡) exp[−𝜆 − 𝛌 ⋅ 𝐜(𝐡)] , (A4)

where the multipliers need to be determined from the normalization condition ∫ℝ𝜈 𝑝𝐇𝑐 (𝐡) 𝑑𝐡 = 1 and the constraint
∫ℝ𝜈 𝐜(𝐡)𝑝𝐇𝑐 (𝐡) 𝑑𝐡 = 𝛄. The determination of the multipliers is performed in an iterative manner using Newton’s
method to a relative tolerance 𝜖KL . The key point to notice is that to evaluate the Newton residual one needs to evaluate
the normalization and constraint integrals. These integrals can be computed using Monte Carlo samples, which we
can conveniently generate using the Itô stochastic differential equation set up in the basic PLoM algorithm. Once the
Lagrange multipliers converge, the density 𝑝𝐇𝑐 (𝐡) can be used in place of 𝑝𝐇 (𝐡) in the basic PLoM setup, thus yielding
the desired samples. It is noted that the Newton iterations in our experience converge quickly and the process remains
efficient.

To summarize, the PLoM methodology utilizes PCA to normalize and linearly de-correlate feature space; it estimates
the PDF of the data using a KDE approximation; it then uses a diffusion basis to reduce the size of the graph of the data
and more importantly preserve the concentration of measure on the response manifold; and last, it utilizes a Hamiltonian
MCMC sampler built upon an Itô stochastic differential equation. In this way, the linear and nonlinear structure of the
data is preserved in the new samples, which are produced efficiently and can then be used to compute statistical estimates
associated with the system of interest. At its essence, the PLoM algorithm treats the initial data sample as a realization of
a random graph and then provides a logic for constructing an efficient basis for the graph and a generator for statistically
similar graphs. In the case that one wishes to sample the response manifold in the presence of additional constraints, the
KL divergence can be minimized to produce a modified-KDE for sampling with the Hamiltonian MCMC sampler. For
the earthquake engineering questions raised in the introduction, these features lead to a substantial numerical efficiency
with little loss in accuracy.

You might also like