Advanced Reduced Order Methods and Applications in Computational Fluid Dynamics
Advanced Reduced Order Methods and Applications in Computational Fluid Dynamics
org/terms-privacy
Advanced
Methods and
Computational
Applications in
Fluid Dynamics
Reduced Order
Computational Science & Engineering
The SIAM series on Computational Science and Engineering publishes research monographs, advanced
undergraduate- or graduate-level textbooks, and other volumes of interest to an interdisciplinary CS&E community of
computational mathematicians, computer scientists, scientists, and engineers. The series includes introductory volumes
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
aimed at a broad audience of mathematically motivated readers interested in understanding methods and applications
within computational science and engineering, monographs reporting on the most recent developments in the field,
and volumes addressed to specific groups of professionals whose work relies extensively on computational science and
engineering.
SIAM created the CS&E series to support access to the rapid and far-ranging advances in computer modeling and
simulation of complex problems in science and engineering, to promote the interdisciplinary culture required to meet
these large-scale challenges, and to provide the means to the next generation of computational scientists and engineers.
Editor-in-Chief
Donald Estep, Simon Fraser University
Editorial Board
Ben Adcock Serkan Gugercin David Keyes
Simon Fraser University Virginia Tech Columbia University
Daniela Calvetti Jan S. Hesthaven Ralph C. Smith
Case Western Reserve University Ecole Polytechnique Fédérale North Carolina State University
de Lausanne
Omar Ghattas Karen Willcox
University of Texas at Austin Johan Hoffman University of Texas at Austin
KTH Royal Institute of Technology
Chen Greif
University of British Columbia
Series Volumes
Rozza, Gianluigi, Stabile, Giovanni, and Ballarin, Rostamian, Rouben, Programming Projects in C for
Francesco, Advanced Reduced Order Methods and Students of Engineering, Science, and Mathematics
Applications in Computational Fluid Dynamics Smith, Ralph C., Uncertainty Quantification: Theory,
Gatto, Paolo, Mathematical Foundations of Finite Implementation, and Applications
Elements and Iterative Solvers Dankowicz, Harry and Schilder, Frank, Recipes for
Adcock, Ben, Brugiapaglia, Simone, and Continuation
Webster, Clayton G., Sparse Polynomial Approximation Mueller, Jennifer L. and Siltanen, Samuli, Linear and
of High-Dimensional Functions Nonlinear Inverse Problems with Practical Applications
Hoffman, Johan, Methods in Computational Science Shapira, Yair, Solving PDEs in C++: Numerical Methods
Da Veiga, Sébastien, Gamboa, Fabrice, Iooss, Bertrand, in a Unified Object-Oriented Approach, Second Edition
and Prieur, Clémentine, Basics and Trends in Sensitivity Borzì, Alfio and Schulz, Volker, Computational
Analysis: Theory and Practice in R Optimization of Systems Governed by Partial
Vidyasagar, M., An Introduction to Compressed Sensing Differential Equations
Antoulas, A. C., Beattie, C. A., and Güğercin, S., Ascher, Uri M. and Greif, Chen, A First Course in
Interpolatory Methods for Model Reduction Numerical Methods
Sipahi, Rifat, Mastering Frequency Domain Techniques Layton, William, Introduction to the Numerical Analysis of
for the Stability Analysis of LTI Time Delay Systems Incompressible Viscous Flows
Bardsley, Johnathan M., Computational Uncertainty Ascher, Uri M., Numerical Methods for Evolutionary
Quantification for Inverse Problems Differential Equations
Hesthaven, Jan S., Numerical Methods for Conservation Zohdi, T. I., An Introduction to Modeling and Simulation
Laws: From Analysis to Algorithms of Particulate Flows
Sidi, Avram, Vector Extrapolation Methods with Biegler, Lorenz T., Ghattas, Omar, Heinkenschloss,
Applications Matthias, Keyes, David, and van Bloemen Waanders,
Borzì, A., Ciaramella, G., and Sprengel, M., Formulation Bart, editors, Real-Time PDE-Constrained Optimization
and Numerical Solution of Quantum Control Problems Chen, Zhangxin, Huan, Guanren, and Ma, Yuanle,
Benner, Peter, Cohen, Albert, Ohlberger, Mario, Computational Methods for Multiphase Flows in Porous
and Willcox, Karen, editors, Model Reduction and Media
Approximation: Theory and Algorithms Shapira, Yair, Solving PDEs in C++: Numerical Methods
Kuzmin, Dmitri and Hämäläinen, Jari, Finite Element in a Unified Object-Oriented Approach
Methods for Computational Fluid Dynamics: A Practical
Guide
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Advanced
Reduced Order
Methods and
Applications in
Computational
Fluid Dynamics
GIANLUIGI ROZZA
SISSA, International School for Advanced Studies
Trieste, Italy
GIOVANNI STABILE
SISSA, International School for Advanced Studies
Trieste, Italy
FRANCESCO BALLARIN
Catholic University of the Sacred Heart
Brescia, Italy
10 9 8 7 6 5 4 3 2 1
All rights reserved. Printed in the United States of America. No part of this book may be reproduced, stored, or transmitted in any manner
without the written permission of the publisher. For information, write to the Society for Industrial and Applied Mathematics, 3600 Market
Street, 6th Floor, Philadelphia, PA 19104-2688 USA.
No warranties, express or implied, are made by the publisher, authors, and their employers that the programs contained in this volume
are free of error. They should not be relied on as the sole basis to solve a problem whose incorrect solution could result in injury to person
or property. If the programs are employed in such a manner, it is at the user’s own risk and the publisher, authors, and their employers
disclaim all liability for such misuse.
Trademarked names may be used in this book without the inclusion of a trademark symbol. These names are used in an editorial context
only; no infringement of trademark is intended.
is a registered trademark.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
To our families, to our younger collaborators,
to little Petra and Clara representing the future,
and to all people who contributed to and inspired this book
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Contents
Preface xxxvii
vii
viii Contents
2.3 ROMs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.1 RB Approximation . . . . . . . . . . . . . . . . . . . . . . . . 26
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
III Advances in Reduced Order Models for Computational Fluid Dynamics 249
Bibliography 415
Index 461
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
List of Contributors
Francesco Ballarin
Department of Mathematics and Physics, Catholic University of the Sacred Heart,
Brescia, Italy
Nicola Demo
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Michele Girfoglio
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Martin W. Hess
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Saddam Hijazi
Institute of Mathematics, University of Potsdam, Germany
Efthymios N. Karatzas
Department of Mathematics, Aristotle University of Thessaloniki, Greece
Moaad Khamlich
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Andrea Lario
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Laura Meneghetti
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
xvii
xviii List of Contributors
Andrea Mola
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Umberto Morelli
Departamento de Matemática Aplicada, Universidade de Santiago de Compostela,
Spain
Monica Nonino
Department of Mathematics, University of Vienna, Austria
Federico Pichi
Chair of Computational Mathematics and Simulation Science, École Polytechnique
Fédérale de Lausanne, Switzerland
Francesco Romor
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Gianluigi Rozza
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Nirav Shah
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Pierfrancesco Siena
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Giovanni Stabile
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Maria Strazzullo
Department of Mathematical Sciences “G.L. Lagrange”, Politecnico di Torino, Italy
Marco Tezzele
Oden Institute for Computational Engineering and Sciences, University of Texas at
Austin, USA
Davide Torlo
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Zakia Zainib
Technische Universität Dortmund, Germany
Matteo Zancanaro
mathLab, Mathematics Area, SISSA, Scuola Internazionale Superiore di Studi
Avanzati, Trieste, Italy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
List of Figures
2.1 Left: example of a triangulation over the unit square resulting from Listing
2.1. Right: highlighted edges for the application of boundary conditions. . . . 18
2.2 P1 and P2 element types for two- and three-dimensional cells in (a), (b) and
(c), (d), respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Basis function φj and its support highlighted in pink. . . . . . . . . . . . . . . 19
2.4 Solution of the system (2.10), with F = sin(µ[0]x) cos(µ[1]y), for several
values of µ. Here, x and y indicate the spatial coordinates of the spatial do-
main. Left: µ = (2π, 2π). Center: µ = (2π, π). Right: µ = (π, π). . . . . . 20
2.5 Flow past a cylinder: solution of the system (2.23) with a parabolic inlet on the
left boundary. Velocity and pressure fields from left to right, respectively. No-
slip conditions on the bottom and top walls and on the cylinder and Neumann
boundary conditions on the right outflow. . . . . . . . . . . . . . . . . . . . . 23
2.6 Flow past a cylinder: solution of the system (2.25) with a parabolic inlet on the
left boundary and velocity and pressure fields from left to right, respectively.
No-slip conditions on the bottom and top walls and on the cylinder and Neu-
mann boundary conditions on the right outflow. The solution is represented
for several values of Re. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.7 Solution manifold at the discrete level represented by the snapshots. . . . . . . 27
2.8 Two-dimensional domain, composed of two subdomains, for the steady heat
conduction problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.9 Snapshots and basis for the thermal block problem. . . . . . . . . . . . . . . . 33
2.10 FE (369 nonzero elements) and RB (225 nonzero elements) structures of the
assembled systems, respectively left and right. . . . . . . . . . . . . . . . . . 35
2.11 Left: parametric domain Ωo (µ). Right: reference domain Ω. . . . . . . . . . . 36
2.12 Top: Stokes flow in deformable channels (RBniCS tutorial 12). Bottom:
Backward-facing step for moderate Reynolds, Navier–Stokes equations (RB-
niCS tutorial 17). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.13 Velocity and pressure reduced solution for the Navier–Stokes equations in the
backward-facing step channel, varying the number of supremizers Ns . . . . . . 42
2.14 Reduced manifold MN ,K for time-dependent problem. . . . . . . . . . . . . 43
2.15 Left: RB temperature solution. Right: computational mesh for Ω = (−1, 1)2
with EIM interpolation points. . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.16 Convergence of the RB method for the elliptic problem with Gaussian source,
with respect to the number of basis functions, for EIM (left) and RB (right). . . 58
xix
xx List of Figures
problem (3.30). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.6 FE solution, velocity magnitude (top) and temperature (bottom), for µ =
4523, µ = 55732, and µ = 642639 (left to right). . . . . . . . . . . . . . . . . 76
3.7 Error evolution for the EIM for µ ∈ [103 , 105 ] (left) and µ ∈ [105 , 106 ] (right). 76
3.8 Evolution of the a posteriori error bound in the greedy algorithm for µ ∈
[103 , 105 ] (left) and µ ∈ [105 , 106 ] (right). . . . . . . . . . . . . . . . . . . . . 77
3.9 A posteriori error bound for N = Nmax for µ ∈ [103 , 105 ] (left) and µ ∈
[105 , 106 ] (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
3.10 FE snapshots for µg = 0.5 (top left), µg = 1 (bottom left), µg = 1.5 (middle),
and µg = 2 (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.11 Error evolution for the EIM and for the Boussinesq VMS-Smagorinsky model
with µ ∈ [103 , 104 ] × [0.5, 2]. . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.12 Evolution of the a posteriori error bound in the greedy algorithm for only the
geometrical parameter (left) and for both physical and geometrical parameters
(right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
3.13 A posteriori error bound for only the geometrical parameter (left) and for both
physical and geometrical parameters (right), for N = Nmax . . . . . . . . . . . 81
5.6 Left: RB density functions at µ = 0.8 with τ ∈ {0.15, 0.25} for the first three
branches. Right: RB bifurcation diagram for the multiparameter test case with
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
infinity norm of the density as output. The blue dashed lines show the critical
values of µ for both branching phenomena. . . . . . . . . . . . . . . . . . . . 115
5.7 Left: error between the branches of the bifurcation diagram computed with
FE and RB with EIM/DEIM in the µ-NB plane. Right: high-fidelity and RB
real part φ of the solution X for the 1-dark soliton stripe branch, left and right,
respectively, at µ = 0.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.8 Left: RB bifurcation diagrams for the SVK unforced beam. Right: high-
fidelity zeroth and second mode displacement u for the SVK beam with B =
(0, 0) at µ = 0.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.9 Left: RB bifurcation plot for SVK beam with B = (0, −1000) for five random
pairs (E, ν) ∈ [105 , 107 ] × [0.25, 0.42]. Right: RB bifurcation plot for SVK
beam with B = (0, −1000) and l ∈ [0.5, 1]. . . . . . . . . . . . . . . . . . . . 119
5.10 Left: RB bifurcation diagram for tubular unforced geometries with l ∈ [10, 20].
Right: representative solutions of the three-dimensional SVK model with l ∈
{2, 20}. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.11 Domain Ω, which represents a straight channel with a narrow inlet. . . . . . . 120
5.12 Left: RB bifurcation diagram for the NS system. Right: velocity magnitude
of the asymmetric branch varying the viscosity µ. . . . . . . . . . . . . . . . . 122
5.13 Representative solutions for the NS system at µ = 0.5, velocity and pressure
fields, lower and middle branch, top and bottom respectively. . . . . . . . . . . 122
5.14 RB errors with respect to µ ∈ P for the velocity and pressure of the asymmet-
ric branch. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.15 Eigenvalues of the state eigenproblem in the complex plane for the NSE: stable
and unstable solutions, left and right panels, respectively. . . . . . . . . . . . . 123
7.5 The figure shows a general overview of the mesh with dimension and bound-
aries (upper left), a table with the imposed values at the boundaries (bottom),
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
and a zoom of the mesh near the cylinder (upper right). . . . . . . . . . . . . . 160
7.6 Kinetic energy relative error in the cylinder example for ν = 0.005 and
ν = 0.005625. The kinetic energy relative error is plotted over time for the
two values of viscosity and for the two different models: with supremizer sta-
bilization (USUP and PSUP) and PPE stabilization (UPPE and PPPE). The
ROM solutions are obtained with 15 modes for velocity, 10 modes for pres-
sure, and 12 modes for supremizers. The time window in this case is wider
than the one used to generate the RB spaces (∆T = 10 s). . . . . . . . . . . . 161
7.7 First four basis functions for velocity (first row), pressure (second row), and
supremizers (third row). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
7.8 Comparison of the velocity field for high-fidelity (UHF, first row), supremizer
stabilized ROM (USUP, second row), and PPE stabilized ROM (UPPE, third
row). The fields are depicted for different time instants equal to t = 195 s,
200 s, 230 s, and 270 s and increasing from left to right. The ROM solu-
tions are obtained with 15 modes for velocity and 10 modes for pressure, and
for the SUP-ROM only with 12 additional supremizer modes. The velocity
magnitude is shown in the image legends. . . . . . . . . . . . . . . . . . . . . 162
7.9 Comparison of the pressure field for high-fidelity (PHF, first row), supremizer
stabilized ROM (PSUP, second row), and PPE stabilized ROM (PPPE, third
row). The fields are depicted for different time instants equal to t = 195 s,
200 s, 230 s, and 270 s and increasing from left to right. The ROM solu-
tions are obtained with 15 modes for velocity and 10 modes for pressure, and
for the SUP-ROM only with 12 additional supremizer modes. The pressure
magnitude is shown in the image legends. . . . . . . . . . . . . . . . . . . . . 162
7.10 Error analysis for the velocity (left plot) and pressure (right plot) fields in
the cylinder example with ν = 0.005625. The L2 -norm of the relative error
is plotted over time for the two different models: with supremizer stabiliza-
tion (USUP and PSUP, continuous red line) and PPE stabilization (UPPE and
PPPE, dot-dashed blue line). The ROM solutions are obtained with 15 modes
for velocity, 10 modes for pressure, and 12 modes for supremizers. . . . . . . 163
7.11 Error analysis for the velocity (left plot) and pressure (right plot) fields in the
cylinder example with ν = 0.005 and ν = 0.005625. The L2 -norm of the
relative error is plotted over time for the two values of viscosity and for the
two different models: with supremizer stabilization (USUP and PSUP) and
PPE stabilization (UPPE and PPPE). The ROMs are obtained with 15 modes
for velocity, 10 modes for pressure, and 12 modes for supremizers. The time
window is in this case wider than the one used to generate the RB spaces
(∆T = 10 s). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
8.1 The computational domain used in the numerical simulations; all lengths are
described in terms of the characteristic length D, which is equal to 1 m. . . . . 181
8.2 Cumulative ignored eigenvalue decay. In the plot, the solid black line refers to
the velocity eigenvalues, the dashed magenta line indicates the pressure eigen-
values, and the dash-dotted blue line refers to the eddy viscosity eigenvalues. . 182
8.3 k- turbulence model case, velocity fields for the value of the parameter U =
7.0886 m/s: (a) shows the FOM velocity, while in (b) one can see the U-ROM
velocity, and finally in (c) we have the H-SUP-ROM velocity. . . . . . . . . . 183
List of Figures xxiii
8.4 k- turbulence model case, pressure fields for the value of the parameter U =
7.0886 m/s: (a) shows the FOM pressure, while in (b) one can see the U-ROM
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
8.18 The time evolution of the L2 relative errors of the pressure-reduced approx-
imations for both the U-ROM and the H-SUP-ROM. The curves correspond
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
to the case run with the parameter value Uin = 7.75 m/s: (a) shows the er-
ror curve for the U-ROM, while (b) depicts the case of the H-SUP-ROM. The
error values in both graphs are in percentages. . . . . . . . . . . . . . . . . . . 193
8.19 Lift coefficient curves for the cross-validation test done for the parameter value
Uin = 7.75 m/s for the time range [0, 8] s; the figure shows the FOM, the
U-ROM, and the H-SUP-ROM lift coefficient histories: (a) the full range is
shown, and (b) the last 2 s of Cl is shown. . . . . . . . . . . . . . . . . . . . . 194
8.20 The graph of the L2 relative errors for the lift coefficient curve versus the
number of modes used in the online stage in the U-ROM and the H-SUP-ROM.
The curves correspond to the case run with the parameter value Uin = 7.75
m/s. The error is computed between the lift coefficient curve obtained by the
FOM solver and the one reconstructed from both the U-ROM and the H-SUP-
ROM for the time range [0, 8] s: (a) shows the error curve for the U-ROM,
where Nr is the number of modes used in the online stage for all variables (by
construction of the U-ROM, it is not possible to choose a different number of
online modes for the reduced variables). (b) depicts the case of the H-SUP-
ROM, where one can see the error values by varying the number of modes used
for the pure velocity, with different fixed settings for the three other variables
(the pressure, the supremizers, and the eddy viscosity). The error values in
both graphs are in percentages. . . . . . . . . . . . . . . . . . . . . . . . . . . 195
8.21 The graph of the peak relative errors for the lift coefficient curves for varied
values of the number of modes used in the online stage in the U-ROM and the
H-SUP-ROM. The curves correspond to the case run with the parameter value
Uin = 7.75 m/s. The error is computed between the peak values of the lift
coefficient curve obtained by the FOM solver and the ones reconstructed from
both the U-ROM and the H-SUP-ROM for the time range [0, 8] s: (a) shows
the error curve for the U-ROM, where Nr is the number of modes used in the
online stage for all variables (by construction of the U-ROM it is not possible
to choose a different number of online modes for the reduced variables). (b)
depicts the case of the H-SUP-ROM. The error values in both graphs are in
percentages. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
8.22 Lift coefficient curves for the cross-validation test done for the parameter value
Uin = 7.75 m/s for the time range [0, 8] s. The figure shows the FOM, the SU-
PPE-ROM, and the H-PPE-ROM lift coefficient histories: (a) the full range is
shown, and (b) the last 2 s of Cl is shown. . . . . . . . . . . . . . . . . . . . . 197
8.23 The graph of the L2 relative errors for the lift coefficient curve versus number
of modes used in the online stage in the SU-PPE-ROM and the H-PPE-ROM.
The curves correspond to the case run with the parameter value Uin = 7.75
m/s. The error is computed between the lift coefficient curve obtained by the
FOM solver and the one reconstructed from both the U-ROM and the H-PPE-
ROM for the time range [0, 8] s: (a) shows the error curve for the SU-PPE-
ROM, where Nu is the number of modes used in the online stage for both the
velocity and the eddy viscosity, while Np is the number of modes used for the
pressure field. (b) depicts the case of the H-PPE-ROM, where one can see the
error values varying the number of modes used for the velocity (including the
lifting velocity mode) with different fixed settings for the two other variables
(the pressure and the eddy viscosity). The error values in both graphs are in
percentages. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
List of Figures xxv
8.24 The graph of the peak relative errors for the lift coefficient curves for varied
values of the number of modes used in the online stage in the SU-PPE-ROM
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
and the H-PPE-ROM. The curves correspond to the case run with the param-
eter value Uin = 7.75 m/s. The error is computed between the peak values
of the lift coefficient curve obtained by the FOM solver and the ones recon-
structed from both the SU-PPE-ROM and the H-PPE-ROM for the time range
[0, 8] s: (a) shows the error curve for the U-ROM, where Nu is the number
of modes used in the online stage for both the velocity and the eddy viscosity,
while Np is the number of modes used for the pressure field. (b) depicts the
case of the H-PPE-ROM. The error values in both graphs are in percentages. . 199
8.25 The time evolution of the L2 relative errors of the velocity-reduced approxima-
tions for both the SU-PPE-ROM and the H-PPE-ROM. The curves correspond
to the case run with the parameter value Uin = 7.75 m/s: (a) shows the er-
ror curve for the SU-PPE-ROM. (b) depicts the case of the H-PPE-ROM. The
error values in both graphs are in percentages. . . . . . . . . . . . . . . . . . . 199
8.26 The time evolution of the L2 relative errors of the pressure-reduced approxi-
mations for both the SU-PPE-ROM and the H-PPE-ROM. The curves corre-
spond to the case run with the parameter value Uin = 7.75 m/s: (a) shows the
error curve for the SU-PPE-ROM. (b) depicts the case of the H-PPE-ROM.
The error values in both graphs are in percentages. . . . . . . . . . . . . . . . 200
8.27 Lift coefficient curves for the cross-validation test done for the parameter value
Uin = 11.75 m/s for the time range [0, 10] s. The figure shows the FOM and
the H-SUP-ROM lift coefficient histories: (a) the full range is shown; (b) the
last 3-s history of Cl is shown. . . . . . . . . . . . . . . . . . . . . . . . . . . 201
8.28 The lift coefficient curves obtained using both the k- and SST k-ω turbulence
models and the H-SUP-ROM ones. The case considered is nonparametrized,
with Uin = 10 m/s corresponding to Re = 105 . The plot is for the time
range t ∈ [6, 8]; the H-SUP-ROM achieved relative L2 errors (over the range
t ∈ [0, 8]) less than 5% in both cases. . . . . . . . . . . . . . . . . . . . . . . 202
9.1 First three POD modes obtained from the resolution of a parametrized Navier–
Stokes problem, describing the flow past a cylinder. . . . . . . . . . . . . . . . 206
9.2 In the left column are the eigenvalue positions with respect to the unit circle,
and in the right column are the corresponding DMD modes. From top to
bottom we have a stable, a convergent, and a divergent mode, corresponding
to an eigenvalue on the unit circle, inside it, and outside it, respectively. . . . . 209
9.3 On the left is the initial dataset containing features at different time scales.
In the middle is the DMD reconstruction, and on the right is the mrDMD
reconstruction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
9.4 On the left is the initial dataset, in the middle is the DMD reconstruction, and
on the right is the cDMD reconstruction. . . . . . . . . . . . . . . . . . . . . 211
9.5 Computational savings of cDMD with respect to the exact DMD for an in-
creasing dimension of the snapshot matrix. . . . . . . . . . . . . . . . . . . . 211
9.6 Representation of the computational pipeline used for the shape optimization
problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
9.7 Example of bulbous bow deformations using the FFD method. . . . . . . . . . 213
9.8 Comparison of the average relative L2 error of the ROM in a naval shape opti-
mization problem as a function of the number of modes (left) and the number
of snapshots used (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
9.9 Flowchart of the FOM/ROM procedure. . . . . . . . . . . . . . . . . . . . . . 214
xxvi List of Figures
9.10 Sketch of the computational domain used for the FOM. The sphere diameter is
D = 0.01 m. Successive mesh refinement layers (R2, R3, R4) are performed
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
using the cell splitting approach until the finest grid spacing 0.001D is reached
in the region R5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
9.11 Normalized singular values for streamwise velocity and pressure snapshots. . . 216
9.12 Isosurfaces of the DMD modes (left) at a value of −15 × 10−5 m/s versus
the isosurfaces of POD modes (right) at a value of 6.5 m/s for the streamwise
velocity field. Modes: 1, 2, 8, and 36. . . . . . . . . . . . . . . . . . . . . . . 216
9.13 Isosurfaces of the DMD modes (left) at a value of −5 × 10−4 Pa versus the
isosurfaces of POD modes (right) at a value of 300 Pa for the pressure field.
Modes: 1, 2, 8, and 36. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
9.14 Relative error percentage in the Frobenius norm for velocity and pressure (top
and bottom, respectively) fields, where half-sample-rate snapshots are used to
train the reduced model, DMD (left) or PODI (right), for a particular trun-
cation (r) while predicting the intermediate snapshots. Plots show only the
snapshotwise prediction error, while disregarding errors in the training set,
which are almost null. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
9.15 Linear dipole (top) and nonlinear quadrupole (bottom) terms of FWH equa-
tion evaluated from LES data and compared to corresponding DMD (left) and
PODI (right) reduced models at different truncation ranks (r). . . . . . . . . . 218
9.16 Flowchart representing the proposed computational pipeline. . . . . . . . . . . 218
9.17 Sketch of the computational domain used to solve the fluid dynamics problem
in its reference configuration. The left plot is a zoom on the mesh near the
wing. The right picture is a schematic view of the domain with the main
geometrical dimensions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
9.18 The temporal evolution of the lift coefficient from 1 s to 30 s for nine different
parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
9.19 Airfoil shape functions with respect to the profile abscissa. The leading edge
corresponds to x = 0. The bump functions are rescaled by a factor of 0.2 for
illustrative reasons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
9.20 On the left is the sufficiency summary plot for the lift coefficient at times
t = 10.0, 14.0, 18.0 s (top to bottom). On the right are the first eigenvector
components at the corresponding parameters. . . . . . . . . . . . . . . . . . . 221
9.21 The relative error of the approximated outputs at different times. The relative
error is computed on 100 test samples, using the high-fidelity lift coefficient
to train the regression for t ≤ 20 s, while for t > 20 s the DMD forecasted
states are used for the training. . . . . . . . . . . . . . . . . . . . . . . . . . . 222
10.1 Full-order, steady-state solution for ν = 10: velocity in the x direction (top)
and y direction (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
10.2 Full-order, steady-state solution for ν = 0.05: velocity in the x direction (top)
and the y direction (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
10.3 Mean and maximum relative L2 (Ω) error in the velocity with increasing basis
size. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
12.1 Comparison of multivariate quadrature rules with order of accuracy 15. . . . . 256
12.2 Heat transfer problem: error decay with respect to the number of reduced ba-
sis functions, comparison between weighted and not weighted algorithms and
between uniform and Beta(20,10) distributed training samples. Left: POD;
right: greedy algorithm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
12.3 Reference domain D. The reference parameter is µ = (1., 1.5, 1., 1.5, µ4 ). . . 258
12.4 Stokes problem: error decay with respect to the number of RB functions,
comparison between unweighted (left) and weighted (right) algorithms and
between the Smolyak, tensor product, uniform, and Beta(75,75) distributed
training samples. For the weighted algorithm, the Smolyak rule, the tensor
product rule, and Beta(75,75) sampling coincide. . . . . . . . . . . . . . . . . 259
12.5 Graetz problem: error decay with respect to the number of RB functions,
and comparison between weighted and unweighted algorithms and between
uniform and Beta(5,3) distributed training samples. With online stabilization
(left), without online stabilization (right). POD algorithm (top), greedy algo-
rithm (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
12.6 Graetz problem: weighted algorithms and Beta(5,3) distributed training sam-
ples. Error of stabilized and nonstabilized online with respect to the advection
parameter µ. Comparison with selective online strategy. Left: POD; right:
Greedy. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
12.7 Domain D, the Gulf of Trieste. Orange: Miramare reserve Dobs . Pollutant
spill Du . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
12.8 Optimal control problem: error decay with respect to the number of RB func-
tions; comparison between state variable (left) and control variable (right) with
an unweighted POD algorithm between uniform, Beta(20, 5), Beta(5, 5), and
Beta(75,75) distributed training samples. . . . . . . . . . . . . . . . . . . . . 263
13.1 Embedded FEM geometrical tools for (i) CutFEM and (ii) SBM. . . . . . . . 267
13.2 (i) Parametrized geometry. (ii) Some RB components for µ ∈ [−0.5, 0.5]
parametrized geometry. (iii) The full/reduced order solution and the absolute
error (µ = −0.015). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
13.3 Heat exchange problem: (i) eigenvalue decay and (ii) mean relative errors. . . 270
13.4 Stokes with SBM experiment: sketch of the mesh, the embedded domain, and
the parameters considered in the numerical examples. . . . . . . . . . . . . . 272
13.5 Stokes system ROM-SBM POD velocity and pressure components for (i) µ1
geometry and (ii) (µ0 , µ1 ) geometry. . . . . . . . . . . . . . . . . . . . . . . 273
13.6 Stokes SBM: the full and the reduced solution and the absolute error for
velocity and pressure for the (i) one-dimensional and (ii) two-dimensional
parametrization case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
13.7 Stokes (SBM) parametrization: (i) µ1 geometry, with the relative errors for
velocity and pressure in (i)a and the execution times in (i)b ; (ii) (µ0 , µ1 ) ge-
ometry, with the velocity and pressure fields with and without supremizer sta-
bilization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
xxviii List of Figures
13.8 Shape parametrization with large deformations and a zoom into the embedded
cylinder visualizing the extended solution. . . . . . . . . . . . . . . . . . . . . 277
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
13.9 Six classical (i) and six improved (ii) POD modes. . . . . . . . . . . . . . . . 278
13.10 (i) Poisson system (CutFEM): eigenvalue decay and error analysis between
reduced order and high-fidelity approximations with and without transport;
(ii) steady Stokes (CutFEM): relative errors with and without transport. . . . . 279
13.11 Circular embedded geometry (parametrized). . . . . . . . . . . . . . . . . . . 280
13.12 Cahn–Hilliard (CutFEM): ROM basis results, the first six modes. . . . . . . . 281
13.13 Results for the embedded circle geometrical parametrization µ = 0.4261 and
t = [28, 36, 46, 100]dt. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
13.14 Cahn–Hilliard (CutFEM): the FOM mass evolution with respect to time, to-
gether with its RB approximation and the conservation of mass. . . . . . . . . 282
14.12 Pressure snapshots (left column) at time t = 0.005 (top) and final time t =
0.015 (bottom). Reduced order pressure simulation (right column) at time
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
t = 0.005 and time t = 0.015. The reduced simulation was obtained with
N = 4 basis functions for each component of the solution of the FSI problem. 309
14.13 Displacement snapshots (left column) at time t = 0.005 (top) and final time
t = 0.015 (bottom). Reduced order displacement simulation (right column)
at time t = 0.005 and time t = 0.015. The reduced simulation was obtained
with N = 4 basis functions for each component of the solution of the FSI
problem. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
14.14 Analysis of the behavior of the relative error for the fluid pressure approxima-
tion. Dashed lines were obtained by employing the preprocessing procedure,
continuous lines were obtained using a standard model order reduction. . . . . 309
14.15 Behavior of the mean relative error for the fluid pressure, depending on the
number N of basis functions employed, with preprocessing (red) and without
preprocessing (blue). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
16.1 Illustration of the inactive variable sampling strategy. Successive steps are
depicted all at once. We highlight the Chebyshev center, the selection of the
next sample using the hit and run method, and the polytope defined by (16.8). . 329
16.2 Illustration of the design of a response surface using KAS and Gaussian pro-
cess regression (GPR). The factorization through an RKHS is emphasized. . . 331
16.3 Illustration of the global AS direction, highlighted in orange in the left panel.
On the right is the corresponding sufficient summary plot. . . . . . . . . . . . 333
16.4 Comparison between K-means, K-medoids, and hierarchical top-down clus-
tering using the AS-induced distance metric for the quartic example of Fig-
ure 16.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 334
16.5 Illustrative scheme of a multifidelity regression procedure that employs AS
as a low-fidelity model. The function to be approximated is the hyperbolic
paraboloid from (16.16). Starting from 10 high-fidelity data points (depicted
in blue and in white) we construct as a low-fidelity model a response surface
which is constant along the inactive subspace. . . . . . . . . . . . . . . . . . . 336
xxx List of Figures
16.6 NLL application two-dimensional example. On the left we have the target
function against the first transformed coordinate, in the middle the loss func-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
tion decay, and on the right the transformed parameter space. . . . . . . . . . . 337
16.7 On the y-axis: basic reproduction number R0 of the SEIR model for the Ebola
epidemic. Left: GPR on the one-dimensional AS. Middle: GPR on the one-
dimensional KAS. Right: GPR on the one-dimensional active variable of the
NLL method. In the plots are also represented the 200 training points used to
find the reduced parameter space and build the GPRs. . . . . . . . . . . . . . . 339
16.8 Each one of the six plots represents the GPR on the reduced space of one
out of six partitions of the parameter space found with LAS. On the y-axis is
the basic reproduction number R0 of the SEIR model for the Ebola epidemic.
On the x-axis is the reduced one-dimensional AS of the relative cluster. In
the plots are also represented the 200 training points scattered among all six
clusters and used to find the partition of the parameter space with hierarchical
top-down clustering and build the GPRs. . . . . . . . . . . . . . . . . . . . . 339
16.9 ASGA scheme. The main steps of the classical GA are depicted from top to
bottom. Projections onto and from a lower-dimension AS are highlighted with
yellow boxes, which are specific to ASGA. . . . . . . . . . . . . . . . . . . . 341
17.1 Sketch of the three deformation maps which compose the FFD morphing. . . 348
17.2 Interpolations corresponding to six different RBFs. The exact function is de-
picted with a red dashed line. The parameter ε is equal to the inverse of the
average distance between the sampling points. . . . . . . . . . . . . . . . . . 349
17.3 Interpolation error for different types of RBFs. Refer to Figure 17.2 for the
actual interpolations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
17.4 Front view (top) and side view (bottom) of a planing hull featuring a double
chine line and substantially straight walls above the chine line. . . . . . . . . 352
17.5 Vertical section of a planing hull perpendicular to its longitudinal axis. The
bottom vertex of the section is the intersection of the section plane and the keel
line. The intersections of the two chine lines are also visible on the bottom left
side of the section. The first step of the parametrization strategy is a rigid
rotation of the bottom part of the section around the keel intersection. This is
combined with a rigid translation of the chine and wall surface sections needed
to keep the top part of the section in contact with the displaced bottom part. . 353
17.6 Front view (left) and lateral view (right) of a ship propeller. In both images,
one of the five propeller blades’ surface is transparent, and a set of airfoil cylin-
drical sections is indicated in red to show how the blade surface is generated
as the envelope surface of the single sections. . . . . . . . . . . . . . . . . . 354
17.7 The airfoil section pitch as a function of relative radial coordinate. The plot
shows both an initial pitch distribution (red continuous line) and the control
points associated with its B-spline interpolation (red piecewise linear line). A
new, smooth pitch radial distribution (blue continuous line) is then obtained
using a displaced set of pitch control points (blue piecewise linear line). . . . 355
17.8 FFD application to the STL triangulation representing the bulbous bow of a
cruise ship hull. The left panel, which represents the original bow geometry,
also displays the FFD undeformed control point lattice. On the right, the vis-
ible displacement of the control point results in a deformed STL geometry.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
List of Figures xxxi
17.9 Two examples of a triangulated surface of a carotid artery deformed with RBF.
The original geometry is emphasized by the blue dots, while the RBF control
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
points are depicted in red (the deformed ones) and in green (the undeformed
ones). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
17.10 A simplified two-dimensional example of the application of FFD to a NURBS
curve. On the left, the procedure starts by placing an FFD lattice (red dots)
on top of a NURBS curve. On the right, the displacement of the FFD control
points is used to modify the position of the NURBS control points (black dots).
The modified control polygon results in the morphed NURBS curve displayed
on the right. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358
17.11 An application of the FFD algorithm to the NURBS surface specified by an
IGES file. A front view of the DTMB-5415 navy combatant hull featuring a
modified and enlarged bow is displayed on the left; the original hull surface is
displayed on the right for reference. . . . . . . . . . . . . . . . . . . . . . . . 359
17.12 A side view of the IGES geometry of the DTMB-5415 navy combatant hull
shown in Figure 17.11. The top picture depicts the CAD surface of the hull,
modified via FFD. The bottom picture shows the original hull surface. . . . . 359
17.13 Representation of a family of ship propellers generated with bottom-up con-
struction. The different propellers are obtained in this case by scaling the skew
angle curve. In particular, in the image an original propeller blade shape is pre-
sented along with two modifications resulting from scaling the skew curve by
factors of 0.8 and 1.1, respectively. . . . . . . . . . . . . . . . . . . . . . . . 360
17.14 Examples of computational domain deformation using FFD: the first picture
(left) shows the edges of the original mesh, while the others illustrate two
deformed domains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
17.15 The application of RBF shape parametrization to the Ahmed body and the
surrounding volumetric mesh for CFD simulations. In the test illustrated, the
inclination of the diagonal surface at the rear end of the Ahmed body is modi-
fied to investigate its effect on aerodynamic performance. Two configurations,
including a vertical cut of the volumetric mesh, are shown in the top images.
For reference, path lines of mean velocity field resulting from the correspond-
ing unsteady RANS simulations are shown in the bottom pictures. . . . . . . 362
17.16 Example of computational domain deformation applied to a propeller design
problem, seen at the radial plane. The blue edges refer to the original mesh,
whereas the red ones show the deformed grid. . . . . . . . . . . . . . . . . . 363
17.17 A sectional view of a volumetric grid deformation carried out using RBFs in
a naval engineering problem. The blue lines denote the original volumetric
mesh edges, while the red lines indicate the deformed configuration. . . . . . 364
18.1 CABG: Boundary control magnitude mm2 /s2 (left) and eigenvalue reduc-
18.7 The mitral valve between the left atrium and the left ventricle produces the
Coandă effect. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
List of Tables
2.1 Approximation accuracy for the ROM with respect to the number of suprem-
izers exploited. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.1 Computational time for FE solution and RB online phase, with the speedup
and the relative error. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.2 Computational time for FE and RB solutions, with the speedup and the error,
for problem (3.30): Ra ∈ [103 , 105 ] (top) and Ra ∈ [105 , 106 ] (bottom) . . . . 78
3.3 Computational time for FE and RB solutions, with the speedup and the error,
for the Boussinesq VMS-Smagorinsky model with µg ∈ [0.5, 2]. . . . . . . . . 81
3.4 Computational time for FE and RB solutions, with the speedup and the error,
for the Boussinesq VMS-Smagorinsky model with µ ∈ [103 , 104 ] × [0.5, 2]. . 81
6.1 Computational time for FE solution and RB online phase, with the speedup
and the error. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.1 The table contains the cumulative eigenvalues for the lid-driven cavity test.
The first, second, and third columns report the cumulative eigenvalues for the
velocity, pressure, and supremizer fields, respectively. The last column con-
tains the value of the inf-sup constant, in the supremizer stabilization case, for
different numbers of supremizer modes and with a fixed number of velocity
and pressure modes (10 modes for velocity and 10 modes for pressure). . . . . 158
7.2 The table contains the cumulative eigenvalues for the cylinder problem. In
the first, second, and third columns are reported the cumulative eigenvalues
for the velocity, pressure, and supremizer fields, respectively, as a function of
the number of modes. In the last column is reported the value of the inf-sup
constant for the supremizer stabilization case for different numbers of suprem-
izer modes with a fixed number of velocity and pressure modes (15 modes for
velocity and 10 modes for pressure). . . . . . . . . . . . . . . . . . . . . . . . 161
7.3 The table contains the computational time for the supremizer (SUP) and the
PPE stabilization techniques. In the cavity experiment, the SUP-ROM is ob-
tained with 10 modes for velocity, pressure, and supremizers, while the PPE-
ROM is obtained with 10 modes for pressure and supremizers. In the cylinder
experiment, the SUP-ROM is obtained with 15 modes for velocity, 10 for pres-
sure, and 12 for supremizers, while the PPE-ROM is obtained with 15 modes
for velocity and 10 for pressure. . . . . . . . . . . . . . . . . . . . . . . . . . 164
8.1 Offline parameter samples and the corresponding snapshot data. . . . . . . . . 190
8.2 Summary of the accuracy and the efficiency results for the ROMs considered
in the problem of flow around a cylinder. . . . . . . . . . . . . . . . . . . . . 200
xxxiii
xxxiv List of Tables
12.1 Tables for the selective stabilization approach. Given a certain threshold ad-
vection coefficient ε, we obtain the mean error by computing only a percentage
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
14.1 Physical and geometrical constants and parameters for the geometrically and
physically parametrized leaflets test case. . . . . . . . . . . . . . . . . . . . . 296
14.2 Problem data for the test case of a flow in a channel with deformable walls. . . 303
16.1 Parameter ranges for the Ebola model. Data taken from [167]. . . . . . . . . . 339
16.2 Comparison of the R2 scores of the response surfaces for the Ebola model
built with the AS, KAS, LAS, and NLL methods. . . . . . . . . . . . . . . . . 340
18.1 L2 -norm relative errors for pressure p, WSS, and velocity components ux , uy ,
and uz for P F = 3.45 l/min. . . . . . . . . . . . . . . . . . . . . . . . . . . 376
18.2 L2 -norm relative errors for pressure p, WSS, and velocity components ux , uy ,
and uz for P F = 4.35 l/min. . . . . . . . . . . . . . . . . . . . . . . . . . . 376
20.1 Results obtained for the reduced net POD+FNN (7) trained on CIFAR10 with
different structures for the FNN. . . . . . . . . . . . . . . . . . . . . . . . . . 408
20.2 Results obtained with CIFAR10 dataset. . . . . . . . . . . . . . . . . . . . . . 409
20.3 Results obtained with a custom dataset. . . . . . . . . . . . . . . . . . . . . . 409
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
List of Algorithms
xxxv
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Preface
Reduced order modeling is an important and fast-growing research field in computational sci-
ence and engineering, motivated by several reasons, of which we mention just a few: parametric
computing, repetitive computational environments, real-time computing, increasing complexity
of scenarios with uncertainties, and better computational performance in optimization, control,
and inverse problems. This book deals with recent developments in model order reduction in
fluid dynamics, including challenging topics such as turbulence, optimal control, flow stability,
aerodynamics, shape optimization, inverse problems, multiphysics, and uncertainty quantifica-
tion. We present here an integrated, reduced, parametric computational framework characterized
by a wide portability to be embedded in the most modern computational pipelines, including
automatic learning and digital twin developments.
After this preface and an overview and motivation in Chapter 1, the first part is concerned
with finite element–based reduced order modeling with a focus on laminar computational fluid
dynamics (Chapter 2), then with the introduction of a simple turbulent pattern (Chapter 3), and
then with the optimal flow control framework (Chapter 4). In Chapter 5, bifurcation problems
are studied, while in Chapter 6 the focus is on transport-dominated problems.
The second part of the book deals with finite volume and spectral element methods and
discontinuous Galerkin-based reduced order modeling. Chapters 7 and 8 deal with finite volume–
based reduced order modeling in computational fluid dynamics (CFD) from laminar to turbulent
flows. In Chapter 9 we deal with nonintrusive data-driven reduced order models. Chapter 10
introduces spectral element method–based reduced order modeling, while Chapter 11 deals with
discontinuous Galerkin–based reduced order modeling.
The third part deals with advanced reduced order modeling in CFD. Chapter 12 deals with
weighted reduced order modeling for uncertainty quantification, Chapter 13 with model order
reduction for embedded methods and level-set geometries, Chapter 14 with multiphysics prob-
lems (fluid-structure interaction), and Chapter 15 with bifurcations in parametric multiphysics
settings.
The fourth part is concerned with perspectives and applications. In Chapter 16 we deal with
reduction in parameter space, and Chapter 17 deals with geometrical parametrization and appli-
cations. Hemodynamics applications are introduced as examples in Chapter 18. In Chapter 19
we introduce our scientific computing open-source libraries and Python tools. Last, but not least,
Chapter 20 provides perspectives and current preliminary development to improve model reduc-
tion by automatic learning, concluding with the digital twin concept.
We accompany this book with our open-source software collection and worked problems,
available at mathlab.sissa.it/cse-software.
We acknowledge all the chapter contributors (SISSA mathLab current members and past
members) as listed in the frontmatter and at the beginning of each chapter. We acknowledge our
long-lasting national and international research collaborations, without which this work would
not exist.
xxxvii
xxxviii Preface
We also acknowledge our several industrial partners for inspiration, trust, and challenging
ideas, since they motivated these research developments. We acknowledge Fincantieri, Danieli,
Electrolux Professional, MICAD, Monte Carlo Yachts, CETENA, ie-fluids, Engys, Optimad, and
Bormioli.
The book is based on preliminary lecture notes and slides used at the first Summer School on
Advanced Reduced Order Methods and Applications in CFD (AROMA-CFD), held in Trieste at
SISSA in July 2019; see https://indico.sissa.it/event/34/.
This project has been carried out during lockdowns and the pandemic spread of COVID-19,
and it has represented for all of us a good way to remain optimistic about the future. This is the
reason the project’s “secret” name was “Decameron,” based on Boccaccio’s manuscript.
We acknowledge the strong support provided by the European Research Council Executive
Agency through the Consolidator Grant AROMA-CFD (grant 681447), as well as FARE-X-
AROMA-CFD project and PRIN project NA-FROM-PDEs by the Italian Ministry for Univer-
sities, INdAM GNCS national group, H2020 MSCA EID ROMSOC (grant 765374), H2020
MSCA RISE ARIA (grant 872442), and regional research programs supported by the European
Social Fund (FSE FVG) and POR-FESR Friuli Venezia-Giulia (Regional Development European
Funds).
We hope that you take some inspiration from this book and that this work will be known as
the AROMA book.
Trieste, Italy, 20 April 2022 Gianluigi Rozza, Giovanni Stabile, Francesco Ballarin
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 1
1
2 Chapter 1. Overview and Motivation
A(µ)x = b, (1.1)
with matrix A ∈ RN ×N , solution vector x ∈ RN , and load vector b ∈ RN . The system
dimension N is typically so large that it is prohibitive to solve the system for many parameter
values µ ∈ D. The parameter domain D depends on the application. It can denote a range of
physical parameters, different geometry configurations, realizations of statistical quantities, or
any combination, which can be posed in a rigorous mathematical form.
In order to solve for many parameter values, the system is projected onto a low-dimensional
ROM space of dimension N . Given a trial basis V ∈ RN ×N , which serves as a projection
matrix, the Galerkin projection of (1.1) is
V T A(µ)V xN = V T b. (1.2)
The resulting solution will be of dimension N , but it can be reprojected to the full-order
space as x = V xN . A Petrov–Galerkin projection uses a test space basis W ∈ RN ×N that is
distinct from the trial space V :
W T A(µ)V xN = W T b. (1.3)
The following sections and chapters will explain much more clearly how to obtain the basis
spaces and how to apply them in different settings.
scenario would be to run the offline stage on an HPC cluster or overnight on a workstation. The
result of the offline stage is the reduced order model (also abbreviated as ROM). The reduced
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
model can be a projected system or an approximated response surface, or it can have any other
form, depending on which ROM method was used.
The reduced model can be evaluated with low computational effort and typically requires
low storage. It can thus be used by embedded systems, portable systems, or even phones. The
evaluation of the ROM for parameters of interest is called the online stage. The online stage
refers to any evaluation of the ROM, whether it is a parameter sweep directly after completion
of the offline stage or whether it is done at a later stage on an embedded device for particular
parameters of interest.
The offline-online decomposition was crucial for the success of RB methods. Early work
[199, 198, 459, 421, 424, 423, 422, 460, 368, 41] could not achieve large speedup factors due to
a dependence on the large-scale dimension in the computations.
f (v) = Av + b (1.5)
The parameter must not be present explicitly, so Θi = 1 is allowed. See [505] for some explicit
examples.
The affine parameter dependency is often important for the offline-online decomposition.
When the fi in (1.6) can be precomputed approximately in the offline phase and stored as low-
order quantities, then the ROM can solve a reduced system which has the same affine parameter
dependency.
The actual computation of the affine dependency can be intricate, e.g., in the case of ge-
ometric parameter dependency. Some software packages invoke symbolic computations in the
parameter to find a minimal affine expansion. An overview of current RB and more general ROM
software can be found in [258] or on the website www.modelreduction.org.
A(µ)x = b, (1.7)
Q
!
X
A(µ)x = Θi (µ)Ai x = b. (1.8)
i=1
If the parameter dependency is nonaffine, an affine representation can be approximated using the
empirical interpolation method (EIM); see [245, 502].
POD
Collect a few snapshot solutions to (1.7) for a chosen set of parameter vectors. The chosen set
of parameters is often taken to be uniformly distributed or uses a Latin hypercube sampling. The
number of snapshot solutions actually requires an informed decision and needs some understand-
ing of how many snapshots are needed to cover all features of solutions in the parameter domain.
In the case of strongly varying solutions in part of the parameter domain, the snapshot sampling
can also happen nonuniformly.
The snapshot solution vectors {x(µ1 ), . . . , x(µL )} are collected in a snapshot matrix S ∈
R N ×L
by simply arranging them columnwise. The SVD of S computes the matrix factorization
S = U ΣW T , (1.9)
where U is an N × N orthogonal matrix, Σ is a diagonal matrix with nonnegative entries, and
V is an L × L orthogonal matrix. It is assumed that the SVD is computed such that the entries
of Σ are sorted in descending order. The diagonal entries of Σ are the singular values of S.
The columns of U are called the left singular vectors and the columns of W are called the right
singular vectors. The Schmidt–Eckart–Young–Mirsky theorem [178, 529, 400] guarantees that
the N -dimensional, best-approximating subspace to S is given by the first N left singular vectors.
See [556, 236] for a textbook reference. Thus, define a projection matrix V ∈ RN ×N as the first
N columns of U .
Computationally, it is often beneficial to compute the first N left singular vectors with the
method of snapshots. Here, the correlation matrix M = S T S is computed and the eigendecom-
position of M is given with M zi = λi zi for 1 ≤ i ≤ N . The ith left singular vector is then
−1
given by ui = λi 2 Szi .
The reduced dimension N is often chosen as a percentage of the POD energy. The total POD
energy is given by the sum of all singular values. The value of N is then computed as the lowest
number of singular values whose sum reaches a prescribed percentage of the total POD energy.
Given V , the system (1.8) can be projected onto the low-order space as
Q
!
X
V T A(µ)V xr = Θi (µ)V T Ai V xr = V T b, (1.10)
i=1
which solves for a reduced order solution xr . The approximate full-order solution is then given
by x ≈ V xr .
During the offline stage, the quantities Ari = V T Ai V and br = V T b are computed such that
the ROM is given by
Q
!
X
Θi (µ)Ari xr = br . (1.11)
i=1
Greedy Algorithm
The greedy algorithm or greedy sampling was introduced for RB methods in [595]. Like the
POD, it also computes a projection space V , which is applied as shown in (1.10) and (1.11).
1.2. Basic Principles 5
The parameter domain is replaced with a discrete sample set Ξ, which can be much more dense
than the sample set for the POD, since it is not necessary to compute the snapshot solution at
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
every parameter location. Instead, an a posteriori error indicator or error estimator is evaluated
over the discrete sample set and the snapshot solution is only computed where the current error
is assumed to be at its maximum. This snapshot is then added to the projection space, and the
procedure is iterated until the maximum estimator falls below a prescribed threshold or the ratio
of initially estimated error to current error falls below a prescribed threshold.
The most common and easy-to-evaluate a posteriori error indicator is the residual. The resid-
ual r(µ) is given by (dropping the µ-dependency for more clear notation)
r = b − AV xr . (1.12)
Since only the norm of the residual is required in the greedy algorithm, this computation can be
recast into a form which is independent of N as
krk2 = rT r (1.13)
T
= (b − AV xr ) (b − AV xr ) (1.14)
T T
= b b − 2b AV xr + xTr V T AT AV xr . (1.15)
Using the affine parameter dependency, the terms occurring in (1.15) can be largely precomputed,
except for the ROM solution xr ; see [281], [505].
Either the greedy iteration can be initialized with an empty V , in which case the snapshot
solution will be added where b = b(µ) attains its maximum norm, or an initial snapshot location
is prescribed. The next iterate is found by finding
µmax = arg max ∆(µ) = arg max kr(µ)k2 . (1.16)
µ∈Ξ µ∈Ξ
The optimality of the greedy approximation can sometimes be established; see [81, 60].
There are more elaborate error indicators and even error estimators which allow one to rig-
orously bound the error. These methods rely on estimates of the stability constant, i.e., the coer-
civity or inf-sup constant. A common technique is the successive constraint method; see [301],
where the method was introduced, and for an extension [120, 121, 591, 278]; interpolation tech-
niques for the stability constant are investigated in [271, 388, 302]. More general time-dependent
and inf-sup stable problems are treated, e.g., in [594, 164, 631, 387, 107].
POD-Greedy
The POD-Greedy method is the standard approach to treating time-dependent problems [180,
260, 346]. The time trajectory is treated with the POD method, while parameters in parameter
space are identified with the greedy algorithm. Theoretical considerations on the convergence
rates of the method are given in [256].
The POD-greedy algorithm is initialized with a basis VN , e.g., computed from the POD of
the trajectory at an initially predefined parameter location of dimension N . An error estimator
or error indicator is then evaluated over a fine grid of the parameter domain to determine the
currently least well approximated parameter location as defined by the ROM associated with
the projection space VN . This is an analogous procedure to the greedy algorithm for stationary
problems, except that the error estimators are more technical; see, e.g., [260].
The trajectory at the greedy-determined parameter location is computed, and its orthogonal
complement with respect to the current projection space VN is determined. The POD of the com-
plement is computed, and the dominant modes are added to the projection space, thus updating
VN and the dimension N . This procedure is continued until the maximum error in the greedy
procedure falls below some predefined tolerance threshold.
6 Chapter 1. Overview and Motivation
1.2.4 Stabilization
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
There are many different concepts of stability in mathematics. This section considers mainly
solvability at a parameter µ. This means that the system matrix A(µ) is invertible, which means
that it has no zero eigenvalues. In particular, it is important that the projected system matrix be
stable when the large-scale system matrix is.
Coercive Problems
Coercive problems carry the stability of the large-scale model over to the ROM. Examples of
coercive problems are heat conduction and mass convection; see [474] for elliptic and parabolic
parametric PDEs. A typical PDE of interest is the Poisson equation, where the system matrix A
assembles to the discrete variational form of the Laplace operator.
In particular, assume there exists a parameter-dependent bilinear form a(·, ·; µ) and a linear
form f (·; µ) defined over a function space X associated with a domain Ω, such that the solution
u(µ) satisfies
a(u(µ), v; µ) = f (v; µ) ∀v ∈ X(Ω). (1.17)
When the function space X is a discrete space, then this is an instance of the parametrized
equation A(µ)x = b, as introduced in (1.1).
The bilinear form is coercive over X and for all parameters µ if
a(w, w; µ)
∃α0 > 0 : α(µ) = inf ≥ α0 ∀µ. (1.18)
w∈X kwk2X
The property (1.18) translates on the matrix level as a positive definite system matrix A(µ)
whose smallest eigenvalue is bounded away from zero for all µ. In particular, this implies that
A(µ) is invertible and there exists a unique solution.
To cover negative definite models, the definition is sometimes altered as
|a(w, w; µ)|
∃α0 > 0 : α(µ) = inf ≥ α0 ∀µ (1.19)
w∈X kwk2X
since VN ⊂ X. This implies that the coercivity constant of the reduced system is also bounded
away from zero and the ROM is thus considered stable.
a(w, v; µ)
∃β0 > 0 : β(µ) = inf sup ≥ β0 ∀µ, (1.21)
w∈X v∈X kwkX kvkX
1.2. Basic Principles 7
where β(µ) is called the inf-sup stability constant at µ. At the matrix level, inf-sup stability
corresponds to an invertible, indefinite system matrix A(µ) and the value of β(µ) corresponds to
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
It holds that
a(w, v; µ)
T µ w = arg sup , (1.23)
v∈X kvkX
and T µ is called the “supremizing” operator. The inf-sup constant can then be expressed as
kT µ wkX
β(µ) = inf . (1.24)
w∈X kwkX
The restriction of X to VN will then upperbound the large-scale inf-sup constant as
kT µ wkX kT µ wkX
inf ≤ inf . (1.25)
w∈X kwkX w∈VN kwkX
By choosing a Petrov–Galerkin projection with the test space VN and trial space T µ VN , the
ROM is stable, with the inf-sup constant of the reduced system βN (µ) given as
kT µ wkX
βN (µ) = inf . (1.26)
w∈VN kwkX
The stabilizing Petrov–Galerkin projection solves the normal equations known from the
residual minimization approach minx∈VN kb − Axk and linear least squares in the projection
space VN since the linear system being solved is
with X also denoting the inner product matrix associated with the space X for notational conve-
nience.
CFD
A particular issue when computing stable ROMs in fluid dynamics arises from different stability
concepts usually called the Brezzi inf-sup condition and the Babuška inf-sup condition. Both
concepts are introduced in [360], while here we will only discuss the Brezzi inf-sup condition
since it is typically used to define stable ROMs.
Viscous flows are described by the incompressible Navier–Stokes equations, which are typ-
ically written in terms of the velocity u : Ω × [0, T ) → Rd and pressure p : Ω × [0, T ) → R
as
∂u 1
+ (u · ∇) u + ∇p − ∆u = 0, Ω × (0, T ), (1.28)
∂t Re
∇ · u = 0, Ω × (0, T ), (1.29)
u(x, 0) = u0 (x), (1.30)
8 Chapter 1. Overview and Motivation
where Ω ⊂ Rd denotes the physical domain, Re denotes the Reynolds number, and appropriate
boundary conditions are supposed to be provided; see [227].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
with linear forms F (·; µ) and G(·; µ) as well as parametrized bilinear and trilinear forms
∂v ∂w
Z
a(v, w; µ) = νij (µ) dΩ, (1.33)
Ω ∂xi ∂xi
∂wj
Z
b(q, w; µ) = − qχij (µ) dΩ, (1.34)
∂xi
Z Ω
∂wm
c(v, w, z; µ) = vi χij (µ) zm dΩ, (1.35)
Ω ∂xj
where the parametrized tensors ν(µ) and χ(µ) might depend on the parameters as well as spatial
coordinates. Einstein summation over repeated indices is used.
Given discrete spaces Vh and Qh to V and Q, define the full-order (Brezzi) inf-sup constant
b(q, v; µ)
βh (µ) = inf sup . (1.36)
q∈Qh v∈Vh kqkQ kvkV
The condition βh (µ) > 0 ensures stability. Note that this only takes the bilinear form b(·, ·; µ)
into account and not the Navier–Stokes operator. To have a stable ansatz space in the sense that
βh (µ) > 0, Taylor–Hood elements can be used, e.g., [74], but there are also other options [468].
Given reduced order spaces VN ⊂ Vh and QN ⊂ Qh , the reduced order Brezzi inf-sup
constant is
b(q, v; µ)
βN (µ) = inf sup , (1.37)
q∈QN v∈VN kqkQ kvkV
where βN (µ) > 0 is a necessary and sufficient condition for ROM stability.
It does not hold that βh (µ) > 0 implies βN (µ) > 0. Instead, the velocity space is enriched in
a particular way so that stability carries over to the ROM. Define for each pressure basis function
Φj ∈ QN
b(Φj , v; µ)
T µ Φj = arg sup , (1.38)
v∈Vh kvkV
such that
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
b(q, v; µ)
0 < βh (µ) = inf sup (1.41)
q∈Qh v∈Vh kqkQ kvkV
b(q, v; µ)
≤ inf sup (1.42)
q∈QN v∈Vh kqkQ kvkV
b(q, T µ q; µ)
= inf (1.43)
q∈QN kqkQ kT µ qkV
b(q, v; µ)
≤ inf sup (1.44)
q∈QN v∈V ∗ kqkQ kvkV
N
∗
= βN (µ). (1.45)
With this velocity space enrichment, the stability of the large-scale system carries over to the
ROM.
where f : Rd → R and A ∈ Rn×d is a tall matrix, with 1 ≤ d < n. The aim is to estimate the
matrix A using only point queries of f or of its gradient.
For the actual implementation of parameter space reduction methods we recommend the
open-source Python package ATHENA [496].
Active Subspaces
Active subspaces (ASs) [134] are one of the most used techniques for linear reduction in input
spaces. ASs have also been used in a multifidelity setting to enhance the accuracy of Gaussian
process regression [494]. More information can also be found in Chapter 16.
To estimate the matrix A described above, ASs construct the uncentered covariance matrix
Σ of the gradients of the function of interest f with respect to the input parameters x:
Z
Σ := Eρ [∇x f ∇x f T ] = (∇x f )(∇x f )T ρ dL, (1.47)
10 Chapter 1. Overview and Motivation
where Eρ denotes the expected value with respect to the probability density function ρ. By
decomposing the symmetric positive definite matrix Σ as W ΛW T , we can retain the first r
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
eigenvectors and store them in W1 . Then we project the inputs over the AS defined by the span
of W1 and we obtain the reduced inputs: y = W1T x.
Kernel-Based ASs
The kernel-based extension of active subspaces (KAS) [493] is a nonlinear technique which
consists of a ridge approximation of f factorized through a projection onto a reproducing kernel
Hilbert space (RKHS). This is done by introducing a feature map φ which maps the inputs into a
higher-dimensional space. Then we seek an AS in this intermediate space.
The new covariance matrix for this case is
Z
Σ := (∇φ(x) f (φ(x)))(∇φ(x) f (φ(x)))T dµ(φ(x)), (1.48)
φ(x)
where µ := φ ◦ L. As we can see, the choice of the feature map is crucial. A common choice
in the machine learning community is the following, which has the advantage that the gradients
can be computed analytically:
r
2
φ(x) = σf cos(Qx + b), (1.49)
D
where D is the number of features, σf is a hyperparameter to be tuned, Q is a matrix whose
rows are sampled from a given probability distribution, and b is a vector whose components are
sampled independently and uniformly in [0, 2π].
Local ASs
Following a completely different strategy, to achieve good linear dimensionality reduction we can
exploit domain decomposition. Local active subspaces (LASs) [495] do so by using a clustering
technique equipped with a supervised distance metric. Then on each cluster a ridge approxima-
tion is performed. Given a pair of inputs, we define the AS-induced distance as
q
kxi − xj kΛ = (xi − xj )T W Λ2 W T (xi − xj ), (1.50)
where Λ and W are the matrices obtained from the decomposition of the covariance matrix
described above. We can observe that by using this formulation we incorporate information
from the global trend given by the global AS. Effective clustering techniques are K-medoids and
hierarchical top-down. The latter is particularly versatile since it allows the selection of more
sophisticated divisions; e.g., it is possible to automatically select the optimal AS dimension on
each cluster.
FFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
FFD consists of creating a lattice of control points around the object of interest and propagating
the deformation of this lattice to the object. This is done by mapping ψ from the physical domain
Ω to a reference domain Ω̂. The FFD parameters µ ∈ P are the displacements of the control
points along the axes. The propagation map T̂ (·, µ) : Ω̂ → Ω̂(µ) is usually composed by
Bernstein polynomials, but other choices are also possible. The FFD map M is the composition
of the three different geometric deformations, i.e.,
The FFD-induced deformation does not depend on the topology of the object to be morphed,
so it is considered very versatile and nonintrusive, especially in industrial contexts [508, 517,
571, 157, 154].
RBFs
A smooth real-valued function whose value depends only on the distance between the input
and some reference point is called a radial basis function (RBF). This class of functions can be
exploited to induce a deformation by moving some control points placed over the parametrized
object and interpolating them with a new surface.
Let {xCi }N NC
i=1 ∈ Ω be such control points, ϕ be a radial basis function, and {γi }i=1 be the
C
NC
weights associated with the RBF centered in {xCi }i=1 . We define the deformation map M for
the RBF interpolation technique as
NC
X
M(x, µ) := s(x, µ) + γi (µ)ϕ(kx − xCi k) ∀x ∈ Ω, µ ∈ P, (1.52)
i=1
where s(·, µ) is a polynomial term of degree one, and µ is the parameter vector composed of
the displacements of the RBF control points. To compute the weights γi we need to impose the
interpolatory constraints
where yCi are the new locations of the control points after the displacement of xCi . For the
remaining unknowns, we need to impose the conservation of the total force and momentum.
For applications of RBF interpolation we suggest among others [568, 390] for a carotid artery
and [573, 161] for a NACA airfoil profile.
without making assumptions on the underlying model. This is done using two different neural
networks: the shifting net calculates the spatial coordinates of the shifted snapshots, while the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
interpolation net evaluates the reference configuration in the coordinates of the shifted snapshots.
It is worth noting that along with the expansion of ANNs to solving complex problems, the
average architecture size increased as well, in terms of both number of layers and number of
neurons. To overcome this issue it is possible to apply reduced order methods, such as POD and
AS, to compress a trained neural network [394]. This leads to remarkable memory savings while
retaining a high level of accuracy. This method has been successfully applied to convolutional
nets, but its generality allows application to other types of architectures.
Part I
Finite Element–Based
Reduced Basis Method
in Computational Fluid
Dynamics
2.1 Introduction
The goal of this book is to give a comprehensive overview of reduced order models (ROMs)
and their applications in fluid dynamics. The main idea of ROMs is to identify a basis able to
capture the parametrized behavior of the system under investigation [281]. ROMs aim at rapidly
obtaining many solutions, corresponding to several values of parameters that can represent both
physical and geometrical features. This technique is important in several fields, from applied
sciences to industry. This chapter focuses on reduced order modeling (ROM) based on the fi-
nite element method (FEM), a classical projection-based approach to solving partial differential
equations (PDEs) [477]. In Section 2.2 we introduce the FEM for solving elliptic PDEs and
some fluid dynamics problems, dealing with both linear and nonlinear equations. Then, the
main ideas behind ROM techniques are described in Section 2.3, where we propose the most
common algorithms exploited in model reduction, i.e., proper orthogonal decomposition (POD)
[34, 85, 114, 281] and the greedy algorithm [81, 281]. The latter procedure uses error estimators
between the FEM and the ROM solutions and is the topic of Section 2.4. Finally, some advanced
techniques involving more complicated problems are treated in Section 2.5.
15
16 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
Our description starts from a simple Poisson problem setting, which paves the way for more
complex applications. The system we introduce is a milestone in the numerical approximation
of PDEs and it represents a very useful tool for understanding basic concepts of numerical dis-
cretization, such as weak formulation, well-posedness results, discrete polynomial spaces, and
estimates of the approximation error [477].
Let us consider a domain Ω ⊂ Rd with d = 1, 2, 3, and, furthermore, let us assume that Ω
has a sufficiently regular boundary ∂Ω. The aim is to find an unknown solution u ∈ C 2 (Ω) such
that
in Ω,
−∆u = f
(2.1)
u=0 on ∂Ω,
where f = f (x) ∈ C 0 (Ω) is a given forcing term acting on the system at hand. The problem
(2.1) is characterized by homogeneous Dirichlet boundary conditions. However, all the strategies
we will present easily adapt to more complicated settings and several variants of the boundary
conditions (such as nonhomogeneous Dirichlet boundary conditions and homogeneous and non-
homogeneous Neumann boundary conditions) and more general elliptic operators.
To solve this prototypical differential problem, we analyze the weak formulation of the sys-
tem. Namely, we move from a second-order description of the operators to an integral form
of the first order of differentiation. To reach this goal, we multiply the Poisson equation by a
smooth and arbitrary function v. Thus, applying the divergence theorem, it is a matter of simple
computations to show that the problem (2.1) can be recast in the form
Z Z Z
∇u · ∇v dΩ − ∇u · vn dΓ = f vdΩ, (2.2)
Ω ∂Ω Ω
where n is the outward unit normal vector to ∂Ω. Combining this integral version of the system
and the homogeneous boundary condition, we can weaken the regularity assumption and find the
.
solution u ∈ V = H01 (Ω), considering the forcing term f ∈ L2 (Ω). Furthermore, thanks to the
assumption u = 0 ∈ ∂Ω, the boundary term of (2.2) vanishes, and by defining the bilinear form
a : V × V → R and the functional F : V → R as
Z Z
a(u, v) = ∇u · ∇v dΩ, F (v) = f v dΩ,
Ω Ω
the weak formulation (2.2) reads as follows: find the solution u ∈ V such that
The well-posedness of the problem, i.e., the proof of the existence, uniqueness, and stability of
the solution, is related to the following theorem.
M
kukV ≤ ,
α
2.2. The FEM 17
.
where M = kV 0 k is the continuity constant of the functional F and
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
. a(v, v)
α = inf
v∈V kvk2 V
In the next subsection, we will analyze the FEM as a numerical approach to dealing with the
elliptic problem (2.3) at a finite-dimensional level.
23
24 Bottom ().mark( boundaries , 1)
25 Left ().mark( boundaries , 2)
26 Top ().mark( boundaries , 3)
27 Right ().mark( boundaries , 4)
28
29 dx = Measure ("dx")( subdomain_data = subdomains )
30 ds = Measure ("ds")( subdomain_data = boundaries )
Listing 2.1. Mesh generation in FEniCS.
Figure 2.1. Left: example of a triangulation over the unit square resulting from Listing 2.1.
Right: highlighted edges for the application of boundary conditions.
In Figure 2.1, we represent a conformal grid over a unit square. The triangulation is obtained
though the reported FEniCS code; see Listing 2.1. Once the physical domain is discretized, the
next step of the approximation relies on the function space definition. The FE space is
V N (r) = {v N ∈ C 0 (Ω) | v|NK ∈ Pr ∀K ∈ T N }, (2.4)
where Pr is the space of polynomials up to degree r ≥ 1. For the sake of notation, we will
drop the degree dependence from the space dimension and we will write V N instead of V N (r) .
However, it is clear that the higher r is, the higher N will be. Indeed, to represent the solution
v N over each cell of T N , its value must be known on a number of properly chosen nodes, i.e.,
specific points over the cell element. Figure 2.2 shows the required cell points to represent an FE
function with P1 and P2 elements, in both two and three dimensions. We also report the FEniCS
code that creates the related FE function spaces in Listing 2.2. The set of nodes will be denoted
by Nj for j = 1, . . . , N . A basis of the FE space V N can be set as the characteristic Lagrangian
functions φj ∈ V N for j = 1, . . . , N such that
if i 6= j,
0
φj (Ni ) = δij = (2.5)
1 otherwise.
The assumption φj ∈ V N is verified since the bases, locally, are polynomials of degree at most
r. The structure of the bases is represented in Figure 2.3: φj has a local support Sj centered in
Nj , i.e., it is nonzero only in a neighborhood of Nj .
1 V1 = FunctionSpace (mesh , "CG", 1)
2 V2 = FunctionSpace (mesh , "CG", 2)
Listing 2.2. FE function space definitions in FEniCS.
2.2. The FEM 19
P1 P2 P1 P2
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 2.2. P1 and P2 element types for two- and three-dimensional cells in (a), (b) and (c), (d),
respectively.
φj
Nj
Thanks to the bases definition, a generic FE function v N ∈ V N can be expressed as the linear
combination of the {φi }N
i=1 as
XN
vN = vi φi (2.6)
i=1
for some unique, yet unknown, real coefficients {uj }N . Now, imposing that the FE solution
satisfy the weak formulation for every basis function, we get the following linear system of N
equations in N unknowns:
N
X
uj a(φj , φi ) = F (φi ) ∀i = 1, . . . , N . (2.8)
j=1
The equations (2.8) can be recast in an algebraic setting. Indeed, let us define the symmetric
positive definite stiffness matrix A ∈ RN ×N as
Z Z
Aij = a(φj , φi ) = ∇φj · ∇φi dΩ = ∇φj · ∇φi dΩ ∀i, j = 1, . . . , N (2.9)
Ω Sj ∩Si
and the forcing vector f ∈ RN where fi = F (φi ). Furthermore, we indicate by u the column
vector of the unknown coefficients of the solution. Then, the equations (2.8) can be recast as the
following linear system of dimension N :
Au = f . (2.10)
20 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
We underline that the linear system (2.10) is sparse because the bases are nonzero only in the
local support. In FEniCS, the linear system (2.10) can be built and, thus, solved for different
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
parameter values (see Figure 2.4), as is done in Listings 2.3 and 2.4, respectively.
1 # Define a function space
2 V = FunctionSpace (mesh , "CG", 1)
3 u = TrialFunction (V)
4 v = TestFunction (V)
5
6 # Create stiffness matrix
7 a = inner (grad(u), grad(v))*dx
8 A = assemble (a)
9
10 # Create forcing vector
11 f = 5*v*dx
12 F = assemble (f)
Listing 2.3. Linear system definition in FEniCS.
A value which gives information about the solvability of the discrete system is the condition
number, and it is related to the Euclidean norm of the matrix A. In this context, we will always
work with the 2-norm, defined as
v
uN N
uX X
kAk2 = t |aij |2 .
i=1 j=1
In the case of a symmetric positive definite matrix, the condition number can be computed as
λmax (A)
K2 (A) = , (2.11)
λmin (A)
where λmax (A) and λmin (A) are the maximum and the minimum singular values of A, re-
spectively. Furthermore, under regularity assumptions on the mesh T N , the condition number
verifies K2 (A) ≤ Ch−2 , where C is a constant which does not depend on h yet may depend on
the polynomial degree r. We stress that the condition number deteriorates as h goes to zero.
Figure 2.4. Solution of the system (2.10), with F = sin(µ[0]x) cos(µ[1]y), for several values
of µ. Here, x and y indicate the spatial coordinates of the spatial domain. Left: µ = (2π, 2π). Center:
µ = (2π, π). Right: µ = (π, π).
Namely, using the FE method, we are dealing with a finite-dimensional approximation of the
continuous problem solution u. The natural question that arises concerns the error one is com-
mitting with respect to the continuous solution of the problem at hand. How far am I from u?
How is the error related to such features as the regularity of the function and the refinement of
the mesh? Answers can be found in the next theorem [477].
Theorem 2.2. Let u ∈ V be the exact solution of the weak formulation of the Poisson equation
(2.3) and uN ∈ V N its approximated solution using an FE discretization of dimension N and
degree r. If u ∈ C 0 (Ω) ∩ H p+1 (Ω) for p > 0, then the following a priori error estimate holds
for a positive constant C:
ku − uN kV ≤ Chs |u|H s+1 (Ω) , (2.12)
where s = min{r, p}.
The proof of the theorem relies on a combination of several fundamental lemmas in the
numerical analysis of the FEM for elliptic problems. As expected, the result states that the
accuracy increases polynomially as h decreases, up to the maximum regularity allowed by u ∈
C 0 (Ω) ∩ H p+1 (Ω).
In the next section, we will deal with a first example of the FEM applied to fluid dynamics.
Indeed, we will show some results and specific FE techniques to deal with the Stokes equations,
the model that describes the flow of incompressible fluids under high diffusivity action.
in Ω,
−ν∆u + ∇p = f
(2.13)
div(u) = 0 on ∂Ω,
where u is the vector fluid velocity, p is the scalar pressure variable divided by the density ρ,
ν = µ/ρ is the kinematic viscosity, and µ is the dynamic viscosity. Here, f is the forcing term
per unit volume. From these quantities, we can define the Reynolds number as
. u∞ L
Re = , (2.14)
ν
22 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
where L is a representative length of the spatial domain and u∞ is the velocity field. The
Reynolds number typically expresses the relation between inertial and viscous forces. Namely,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Re is large when inertial forces are dominant. This is not the case for Stokes flows, where inertial
forces are negligible with respect to viscous forces, i.e., Re → 0.
The boundary conditions will be of either Dirichlet type, i.e., u = g on ΓD ⊂ ∂Ω, or
∂n − pn = h. Also, for these specific equations, we can write (2.13) in a
Neumann type, i.e., ν ∂u
weak form. To this purpose, we multiply the equation
−ν∆u + ∇p = f
by a smooth and arbitrary vector function v, and, exploiting the divergence theorem, we obtain
Z
∂u
Z Z Z
ν∇u : ∇v dΩ − pdiv(v) dΩ = f · v dΩ + ν − pn · vdΓ. (2.15)
Ω Ω Ω ΓN ∂n
The same strategy can be applied to div(u) = 0 as well. The latter expression is known as
the continuity equation. Thus, exploiting the multiplication for a smooth and arbitrary scalar
function q, the continuity constraint is
Z
qdiv(u)dΩ = 0. (2.16)
Ω
To guarantee that the integral formulation is meaningful, let us consider the velocity fields in
. .
V = HΓ1D (Ω) and the scalar field in Q = L2 (Ω), where V is the space of vector functions of
H 1 (Ω) regularity which vanish over ΓD . With a similar argument with respect to the Poisson
equation, we define the bilinear forms a : V × V → R and b : V × Q → R, where
Z Z
a(u, v) = ν∇u : ∇v dΩ, b(u, q) = − qdiv(u)dΩ, (2.17)
Ω Ω
and the functional F : V → R, where
Z Z
F (v) = f · v dΩ + h · vdΓ. (2.18)
Ω ΓN
Namely, the weak formulation of the Stokes problem is to find the pair (u, p) ∈ V × Q such that
a(u, v) + b(v, p) = F (v) ∀v ∈ V,
(2.19)
b(u, q) = 0 ∀q ∈ Q.
The problem (2.19) is, then, a saddle-point problem, which is well posed thanks to the following
theorem [74].
Theorem 2.3 (Brezzi). The weak formulation of the Stokes problem admits a unique solution if
and only if
1. a : V × V → R is bilinear; coercive on the space
Vdiv = {v ∈ V : b(v, q) = 0 ∀q ∈ Q},
with coercivity constant α; and continuous with continuity constant γ;
2. b : V × Q → R is bilinear and continuous, and there exists a constant β > 0 such that
the inf-sup condition is verified, i.e.,
b(v, q)
inf sup ≥ 0. (2.20)
q∈Q v∈V kvkV kqkQ
Furthermore, the following stability estimates hold:
1 1 γ
kukV ≤ kF kV 0 , kpkQ ≤ 1+ kF kV 0 . (2.21)
α β α
2.2. The FEM 23
However, when the system is discretized with the FEM, the problem may not satisfy the inf-
sup condition (2.21). In this case, there exists a spurious pressure mode η ∈ Q, η 6= 0, such that
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
b(v, η) = 0 for all v ∈ V . This means that, if the pair (u, p) ∈ V × Q is a solution, then the
pair (u, p + τ η) ∈ V × Q would also be a solution for any scaling factor τ ∈ R. In this case, no
stability estimate holds for the pressure field p; indeed, its norm could explode for very large τ .
However, the stability estimate for u is still valid.
To avoid this issue, the velocity and pressure spaces cannot be independently chosen; in
particular, the polynomials’ degrees must verify some features. One of the main FE approaches
used for Stokes equations relies on Taylor–Hood elements. Namely, the approximation consists
of exploiting the polynomial space pair Pr -Pr−1 with r ≥ 2 for V and Q, respectively. From now
on, we will always use this conventional choice for the FE approximations of Stokes systems.
Now we can focus on the algebraic formulation of this fluid dynamics system. Let us denote
Np
by {ϕj }Nj=1 ∈ V
u N
and {ψk }k=1 ∈ QN the Lagrangian basis functions of the spaces V and Q,
.
with N = Nu + Np , representing the global dimension of the space. Thus, the discrete solutions
can be expressed as
Nu Np
X X
N N
u = uj ϕj , p = uk ψk . (2.22)
j=1 k=1
AU + B T P = f ,
(2.23)
BP = 0,
A BT
U f
= . (2.24)
B 0 P 0
In this algebraic context, the equivalent of the inf-sup condition (2.20) is ker(B T ) = {0}; when
it holds, no spurious pressure modes are present. In Figure 2.5 we present the solution of a flow
past a cylinder, a classical benchmark in fluid dynamics. We also show the FEniCS code to solve
it in Listing 2.5. In the next subsection, we will focus our attention on how to approximate a
more complicated model, the nonlinear Navier–Stokes equations.
Figure 2.5. Flow past a cylinder: solution of the system (2.23) with a parabolic inlet on the left
boundary. Velocity and pressure fields from left to right, respectively. No-slip conditions on the bottom and
top walls and on the cylinder and Neumann boundary conditions on the right outflow.
1 # Function spaces
2 V_element = VectorElement (" Lagrange ", mesh. ufl_cell () , 2)
3 Q_element = FiniteElement (" Lagrange ", mesh. ufl_cell () , 1)
4 W_element = MixedElement (V_element , Q_element ) # Taylor -Hood
5 W = FunctionSpace (mesh , W_element )
6
7 # Test and trial functions
24 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
in Ω,
−ν∆u + (u · ∇)u + ∇p = f
(2.25)
div(u) = 0 on ∂Ω,
where (u · ∇)u is a nonlinear term describing the convective action of the flow. The correspond-
ing weak formulation requires the introduction of an additional trilinear form C : V ×V ×V → R
defined as Z
c(u, w, v) = [(u · ∇)w] · v dΩ. (2.26)
Ω
Thus, the weak formulation of the Navier–Stokes equations is to find the pair (u, p) ∈ V × Q
such that
a(u, v) + c(u, u, v) + b(v, p) = F (v) ∀v ∈ V,
(2.27)
b(u, q) = 0 ∀q ∈ Q.
In Listing 2.6, we show how to deal with such a system in FEniCS, from the function space
definition to the solution of the equations. In Figure 2.6 we show some parametric solutions of
the problem at hand, varying the Reynolds number. As expected, the solution for Re = 1 is quite
similar to the Stokes one presented in Figure 2.5.
1 # Test and trial functions
2 (v, q) = TestFunction (W)
3
4 # Solution
5 up = Function (W)
6 (u, p) = split (up)
7
8 # Residual
9 F = (nu * inner (grad(u), grad(v)) * dx
10 + inner (grad(u) * u, v) * dx
11 - div(v) * p * dx
12 + div(u) * q * dx
13 - inner (f, v) * dx)
14
15 # Jacobian
16 J = derivative (F, up)
17
18 # Boundary conditions
2.2. The FEM 25
19 bc = ...
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
20
21 # Solve the problem
22 problem = NonlinearVariationalProblem (F, up , bc , J)
23 solver = NonlinearVariationalSolver ( problem )
24 solver . solve ()
Listing 2.6. Navier–Stokes problem in FEniCS.
The existence and the uniqueness of the solution are guaranteed by the Banach contraction
theorem when
kf kV 0 ≤ C,
where the constant C depends on the continuity constant of the trilinear form and on the size and
shape of the domain Ω, represented by the constant of the Poincaré inequality. This assumption,
i.e., the small data assumption, is verified when either f is controlled in norm or the Reynolds
number Re is not large. However, this is not always the case in practice, and we may be interested
in cases lacking uniqueness [446, 452, 454]. From now on, we will always assume that the
problem at hand is well posed. In order to discretize it, one can use the same Taylor–Hood
spaces as in the Stokes case. Furthermore, to deal with the nonlinearity of the system related
to the convective trilinear term, a Newton method can be used. It results in the solution of the
following iterative scheme: given (un , pn ) ∈ V × Q, find the pair (un+1 , pn+1 ) ∈ V × Q such
that
a(un+1 , v) + c(un , un+1 , v) + c(un+1 , un , v) + b(v, pn+1 ) = F (v) ∀v ∈ V,
b(un+1 , q) = 0 ∀q ∈ Q.
(2.28)
The iterations shall be repeated until a convergence criterion is reached. In this chapter we will
not consider time dependency and time evolution. However, the extension is straightforward
and the simulations can be performed through the standard finite difference method in time,
e.g., backward Euler. In the next section, we will introduce the basic ideas behind ROMs and
the principal algorithms related to such an approximation. The different procedures will be
complemented by some numerical examples.
Figure 2.6. Flow past a cylinder: solution of the system (2.25) with a parabolic inlet on the left
boundary and velocity and pressure fields from left to right, respectively. No-slip conditions on the bottom
and top walls and on the cylinder and Neumann boundary conditions on the right outflow. The solution is
represented for several values of Re.
26 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
2.3 ROMs
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Dealing with the approximation of parametrized problems can be a critical task from the compu-
tational viewpoint. Thus, the idea is to replace the original high-dimensional problem, i.e., the
FE approximation, with a reduced one that is easy to manage. In this section, we provide a short
mathematical and practical introduction to ROMs, and in particular to the reduced basis (RB)
method, by which we can reduce the computational costs of our problems. We will present some
preliminary applications to elliptic PDEs and Stokes system test cases, which were discussed in
the previous section. We will have a look at the building blocks of the RB technique: (i) param-
eter separability and (ii) RB space generation via POD and the greedy algorithm. Moreover, we
will highlight the differences with respect to the FE method case and we will show examples of
implementation within RBniCS [38].
2.3.1 RB Approximation
As we saw in the previous section, the preliminary step for the discretization of a PDE is the pro-
jection of its weak formulation in a finite-dimensional setting, which results in a Galerkin system
with N degrees of freedom. In a parametrized many-query and real-time context, the compu-
tational burden can become unaffordable; for this reason, here we focus on the RB method.
Roughly speaking, this method consists of a projection of the high-fidelity problem onto a sub-
space of smaller dimension spanned by some properly chosen basis functions; thus it shares
many features with the Galerkin-FE method.
More precisely, the mathematical setting and the goals of this approximation are as follows.
Let µ denote parameters which affect the solution of a problem of interest (e.g., due to uncer-
tainty, optimization, or physical or geometrical features). Let P ⊂ Rp be the parameter domain;
given a tolerance > 0, we want to build an efficient approximation of the FE solution for any
given parameter µ ∈ P. This consists of evaluating the map µ ∈ P → uN (µ) ≈ uN (µ),
possibly achieving the desired accuracy
kuN (µ) − uN (µ)kN
V ≤ ,
but more efficiently than the FEM; i.e., the RB evaluation µ → uN (µ) should be much faster
than its FE counterpart, or ∂tN
comp ∂tcomp .
N
• In the many-query context, we deal with optimization problems, or we explore the param-
eter domain, requiring multiple evaluations of the solution
tcomp (µj → uN (µj ), j = 1, . . . , J) = ∂tN
comp J as J → ∞.
We can thus introduce the notion of solution manifold in the discrete context, comprising all
solutions of the parametric problem under variation of the parameters, i.e.,
MN = {u(µ) | µ ∈ P ⊂ V N },
where each uN (µ) ∈ V N corresponds to the solution of the parametric FE problem. The key
assumption for the applicability of any reduced model is that the solution manifold in Figure 2.7
2.3. ROMs 27
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
uN (µntrain )
N
M
uN (µ)
N 1
u (µ )
Figure 2.7. Solution manifold at the discrete level represented by the snapshots.
can be represented by the span of a low number of basis functions, i.e., N N , with a small
error. Thus, we start obtaining some information about the behavior of the solution with respect
to the parameter, i.e., the computation of FE solutions for prescribed parameter values (the so-
called snapshots), and then we extract the RB.
As we can see in the following example, to obtain good approximation properties, a cru-
cial assumption is the smooth parametric dependence of uN (µ) on µ. We refer the reader to
Chapter 5 for complex applications concerning the lack of smoothness related to bifurcating
phenomena.
Example 2.4 (linear superposition as RB method). Consider the second-order PDE on the in-
terval I = (0, 1):
(1 + µ)u00 = 1 in I,
(2.29)
u(0) = u(1) = 1.
For any µ ∈ [0, 1] we want to find its solution u(µ) ∈ C 2 (I).
Within this simple example we can compute the two snapshots given by
(
u0 = u(µ = 0) = 12 x2 − 21 x + 1,
(2.30)
u1 = u(µ = 1) = 14 x2 − 41 x + 1,
and by taking them as an RB, we can represent the solution manifold by means of VN =
span{u0 , u1 }.
Therefore, the reduced solution for any µ ∈ [0, 1] is
2 2
uN (µ) = − 1 u0 − − 2 u1 , (2.31)
1+µ 1+µ
which is actually equal to the exact solution u(µ) in the continuous counterpart of the solution
manifold MN , contained in a two-dimensional subspace of the infinite-dimensional space C 2 (I)
thanks to the linear superposition property.
Figure 2.8. Two-dimensional domain, composed of two subdomains, for the steady heat conduc-
tion problem.
A parametric heat flux µ2 occurs over Γbase , while the left and right boundaries Γside are in-
sulated, and the top boundary Γtop is kept at a reference temperature, which for the sake of
simplicity is taken as zero. The formulation of this test case, called the thermal block problem
and reported as the first tutorial in the RBniCS library, reads as follows: for a given parameter
µ = (µ1 , µ2 ) ∈ P = [0.1, 10] × [−1, 1] find u(µ) such that
−div(κ(µ1 )∇u(µ)) = 0 in Ω,
u(µ) = 0 on Γtop ,
κ(µ1 )∇u(µ) · n = 0 on Γside ,
on Γbase ,
κ(µ1 )∇u(µ) · n = µ2
where n denotes the outer normal to the boundaries Γside and Γbase and the conductivity is
defined as in (2.32).
As we said, in order to approximate numerically the solution of the model under considera-
tion, we need to derive its weak formulation. If we define the function space V as
the problem reads as follows: given a parameter µ ∈ P, find u(µ) ∈ V such that
Qa Qf
θfq (µ)fq (v),
X X
a(u, v; µ) = θaq (µ)aq (u, v), f (v; µ) = (2.34)
q=1 q=1
and Z
f (v; µ) = µ2 v dΓ .
|{z} Γbase
θ1f (µ) | {z }
f1 (v)
Therefore, the standard pipeline for the discretization of a general weak formulation of the form
(2.33), via the Galerkin methodology, for a given parameter µ ∈ P reads as follows:
• Continuous problem: Find u(µ) ∈ V such that
a (u(µ), v; µ) = f (v; µ) ∀v ∈ V.
a uN (µ), v N ; µ = f (v N ; µ) ∀v N ∈ V N .
a uN (µ), v N ; µ = f (v N ; µ) ∀v N ∈ V N .
These considerations bring us to the following questions: How can we construct the RB space
V N ? Is it possible to choose a number of basis functions N N that can approximate the
high-fidelity manifold? In the next subsections we will provide some answers to build the RB
approximation.
1 N i N
CP ij
OD
= u (µtrain ), uN (µjtrain ) , 1 ≤ i, j ≤ ntrain .
ntrain V
Then, we want to search for the eigenpairs (ψ k ∈ Rntrain , λk ∈ R+,0 ) of the eigenproblem
C P OD ψ k = λk ψ k , 1 ≤ k ≤ ntrain .
Arranging the eigenvalues in descending order as λ1 ≥ λ2 ≥ · · · ≥ λntrain ≥ 0, we can now
identify Ψk ∈ V for 1 ≤ k ≤ ntrain as
nX
train
k k
Ψ = ψm u(µm
train ).
m=1
to be used as a criterion to select the dimension of the POD basis by looking at the retained energy
from the discarded snapshots. Thus, the POD-RB spaces are give by XN P OD
= span{Ψn , 1 ≤
n ≤ N }, 1 ≤ N ≤ Nmax , with orthonormalized basis functions, i.e., (Ψn , Ψm )X = δn,m , 1 ≤
n, m ≤ ntrain .
It is clear that one of the disadvantages of this technique is the huge number of FE solutions
that have to be computed to obtain a fairly accurate representation of the high-fidelity mani-
fold. Another one concerns the computation of the eigenproblem of dimension ntrain × ntrain .
Despite this, this methodology is still widely used in many contexts, including time-dependent
problems and applications in CFD.
We conclude this section with an example of the implementation in RBniCS of the offline
phase regarding the POD technique (Listing 2.7).
1 def offline (self):
2 # Loop over training set
3 for (mu_index , mu) in enumerate (self. training_set ):
4 # Choose parameter mu
5 self. truth_problem . set_mu (mu)
6 # Truth solve for mu
7 snapshot = self. truth_problem . solve ()
8 # Update snapshots matrix
9 self. update_snapshots_matrix ( snapshot )
10 # Perform POD
11 self. compute_basis_functions ()
12 # Build reduced operators
13 self. reduced_problem . build_reduced_operators ()
Listing 2.7. rbnics/reduction_methods/base/pod_galerkin_reduction.py.
Thus, the so-called proto version of the greedy algorithm over Ξtrain reads as follows.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
The bottlenecks of this methodology are (again) the need to compute ntrain FE solutions and
the suboptimality in the sense that the greedy strategy aims to be optimal in the maximum norm
over Ξtrain rather than L2 for the POD basis. To overcome the first issue, the “actual” method,
instead of building a huge dataset, requires only one high-fidelity solution per iteration, and a
total of N solutions for a basis of dimension N .
Hence, the key ingredient is the selection of the parameters for which the solutions are com-
puted. This choice is usually guided by an error estimator ∆N (µ), which is used at the generic
N th iteration to find the worst approximated solution by the RB space of dimension N in the
whole parameter space Ξtrain . Thus, this version reads as follows.
In order for this to be an improvement of the prototype greedy algorithm, we need ∆N (µ) to
be a sharp, inexpensive1 a posteriori error bound for the ROM error, i.e., for N = 1, . . . , Nmax
we want
||uN (µ) − uN (µ)||V ≤ ∆N (µ) ∀µ ∈ P.
This way, Algorithm 2.2 only computes at most N snapshots, in contrast to ntrain of POD and
proto-greedy. Given the fundamental importance of the a posteriori error estimator ∆N (µ), let
us now detail further the approximation capabilities of the greedy technique. Indeed, the explicit
expressions for the error estimators are postponed to the next section.
Proposition 2.5 ([81]). Assume that there exists an optimal space VN∗ and an optimal u∗N (µ) ∈
VN∗ such that
||uN (µ) − u∗N (µ)||V ≤ ce−αN for some c, α > 0.
Then the RB VN constructed by the greedy method still guarantees exponential convergence
A natural question that arises is how reasonable the assumption required in the previous
proposition is. As concerns elliptic problems, one can explicitly construct a (possibly subop-
timal) space VN∗∗ which features exponential convergence for the model problem described in
1 Marginal cost (= average asymptotic cost) is independent of N .
32 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
Subsection 2.2.1 [81, 386]. Moreover, in the past decades the scientific community has sup-
ported its numerical evidence by means of a great number of different applications.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
It can still be applied when dealing with fluid dynamics problems if one considers moderate
Re numbers, while it is not necessarily true for convection-dominated problems. Indeed, in the
latter case, i.e., Re 1, linear RB approximation techniques suffer from the slow decay of the
Kolmogorov n-width [133, 393], where the exponential decay assumption is not well suited. For
this reason, there are several ongoing efforts to restore this property in that case.
As we said, by construction, the POD basis is orthonormal, while the same does not hold
for the greedy technique, which in principle only provides the basis as the computed snapshots
VN = span{uN (µn ), 1 ≤ n ≤ N }, which are not (necessarily) orthogonal. Thus, we will apply
the Gram–Schmidt orthonormalization during the greedy strategy,
to obtain a more suitable basis for the RB space. The Gram–Schmidt orthonormalization proce-
dure is as follows:
3: for n = 2, . . . , NmaxP:
n−1
4: z n = uN (µn ) − m=1 (uN (µn ), ζm )X ζm . Orthogonal projection onto Vn−1
5: ζn = z /||z ||V
n n
. Normalization
6: end
As a result of this process we obtain the orthogonality condition (ζn , ζm )X = δn,m for 1 ≤
n, m ≤ Nmax , where δn,m is the Kronecker delta symbol.
Again, let us conclude this subsection with an example of the implementation in RBniCS of
the offline phase concerning the greedy algorithm (Listing 2.8).
1 def offline (self):
2 # Greedy loop with stopping criteria
3 while self. reduced_problem .N < self.Nmax and relative_error_estimator_max
>= self.tol:
4 # Truth solve for mu
5 snapshot = self. truth_problem . solve ()
6 # Update snapshots matrix
7 self. update_snapshots_matrix ( snapshot )
8 # Build reduced operators
9 self. reduced_problem . build_reduced_operators ()
10 # Build error estimation operators
11 self. reduced_problem . build_error_estimation_operators ()
12 # Choose next mu value with Greedy
13 self. greedy ()
Listing 2.8. rbnics/reduction_methods/base/rb_reduction.py.
Figure 2.9. Snapshots and basis for the thermal block problem.
For example, as regards the rate of convergence, for the FE we have the standard polynomial
decay [477] depending on both the approximation degree r and the regularity q:
M
ku − uN kV ≤ Chs |u|H s+1 (Ω) for s = min{r, q}. (2.35)
α
On the contrary, as we observed in the previous subsection, for the RB technique we have an
exponential order, due to the estimate
||uN − uN ||V ≤ Ce−βN for some C, β > 0,
where C and β usually depend very mildly on r, q, and h.
Another fundamental difference is about the meaning of the bases which span the discretized
space. Indeed, the FE basis functions have local support and are not necessarily orthonormal,
while the RB ones have by construction global support and are orthonormal (directly for POD,
or by Gram–Schmidt postprocessing for greedy; see Figure 2.9 for an example regarding the
thermal block problem).
Let us now describe the system assembly of the offline and online procedures. Within the RB
methodology, we can express every element of the space V N in terms of its basis {ζn } ∈ V N as
the expansion
XN
uN (µ) = uN
j (µ)ζj . (2.36)
j=1
Then, if a(u, v; µ) satisfies the affine parameter dependence assumption (and assuming the
external force f (·) to be nonparametric), we can insert equation (2.36) into (2.34), obtaining
N
X
a(ζj , ζi ; µ)uN
j = f (ζi ), 1 ≤ i ≤ N,
j=1
N Qa
!⇓
X X
θaq (µ)aq (ζj , ζi ) uN
j = f (ζi ), 1 ≤ i ≤ N.
j=1 q=1
34 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
In this view, the offline-online decomposition for an efficient evaluation of the RB solution uN
j
corresponding to the parameter value µ, reads as follows for 1 ≤ j ≤ N :
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
j (µ) for 1 ≤ j ≤ N .
– Solve for uN
In the evaluation process of uN j for 1 ≤ j ≤ N , we note that from the jth column of the
matrix Z N = {(ζj )k }j,k ∈ R N ×N
, denoting the coefficients when the jth basis function ζj is
expressed in terms of the basis functions {ϕN
k }k , we can write
N
X
ζj = (ζj )k ϕN
k ,
k=1
and thus we obtain the following expression for the RB block corresponding to the parameter-
independent form aq (·, ·):
P
N PN
[AN,q ]ij = aq (ζj , ζi ) = aq k=1 (ζj )k ϕN
k , l=1 (ζ i ) l ϕN
l
PN PN q N N (2.37)
= k=1 l=1 (ζj )k a ϕk , ϕl (ζi )l
= Z TN AN ,q Z N .
Of course, having considered a new (reduced) system, one can investigate its algebraic sta-
bility, and in particular its connection to the orthonormality assumption on Z N .
Proposition 2.6. Assume that a(·, ·; µ) is symmetric and that Z N is orthonormal. Then the
condition number of
Qa
X
AN (µ) = θaq (µ)AN,q
q=1
is bounded independently on N .
Proof. For the sake of simplicity, let us remove µ from the notation. Let vN = [vjN ]N
j=1 be an
N
eigenvector associated with λmax (A ), and define
N
X
N
v = vjN ζj . (2.38)
j=1
Due to orthonormality,
X N N
X N
X N
X
kv N k2V = viN ζi , vjN ζj = viN vjN (ζi , ζj ) = kvN k2 . (2.39)
i=1 j=1 i,j=1 i=1
2.3. ROMs 35
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 2.10. FE (369 nonzero elements) and RB (225 nonzero elements) structures of the assem-
bled systems, respectively left and right.
Similarly, one obtains that λmin (AN ) ≥ α, where α is the coercivity constant. We can conclude
by taking the ratio K2 ((AN )) = λmax (AN )/λmin (AN ), which is bounded independently by the
stability constants of the system.
Finally, let us summarize what we have understood in this section concerning the properties
of the FE and the RB assembled systems.
• Dimension:
• Conditioning:
We are now ready to show an application of the RB methodology within the context of
geometrical parametrization.
Ωo (µ) = [−2, 2] × [−2, 2] \ [µ0 , µ1 ] × [µ0 , µ1 ], where the geometrical parameter is µ = (µ0 , µ1 )
in the parameter space P = [0.5, 1.5]2 .
From the original parametrized domain Ωo (µ), which can be decomposed as
K[
dom
k
Ωo (µ) = Ωo (µ),
k=1
to ensure a continuous piecewise-affine global mapping T aff (·; µ) from Ω onto Ωo (µ).
Thus, we can now describe the weak formulation on the original parametrized domain. For
v, w ∈ H 1 (Ωo (µ)), we consider
∂w
K
Xdom Z ∂xo1
∂v ∂v k ∂w
ao (v, w; µ) = ∂xo1 ∂xo2 v Koij (µ) ∂xo2
,
Ωk
o (µ)
k=1 w
2.3. ROMs 37
where Kok : P → R3×3 is symmetric and positive definite for 1 ≤ k ≤ Kdom ; note that Kok being
affine in xo is also permissible.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Of course, in order to make the parameters explicit and to be able to run the reduction ma-
chinery, we exploit the change of variables formula, resulting in the following problem. For
v, w ∈ H 1 (Ω) we consider
∂w
KXdom Z ∂x1
∂v ∂v
k ∂w
a(v, w; µ) = ∂x1 ∂x2 v Kij (µ) ∂x ,
k 2
Ω
k=1 w
where
aff,k −1 0
(G )
Kk (µ) = |detG aff,k (µ)|D(µ)Kok (µ)DT (µ) and D(µ) = 0 .
0 0 1
Finally, we can make a few remarks and point out some issues when dealing with problems
characterized by geometrical parametrization.
• One of the requirements of the above methodology is the existence of an analytical map-
ping between subdomains, so that snapshots are defined in a common mesh. For recent
extensions to the removal of this assumption we refer the reader to [324, 325].
• An affine mapping is used in order to ensure the availability of a parameter-separable
formulation, but more general mappings are possible and very useful in CFD [508, 569].
• The pullback procedure (change of variables in integrals) required for the assembly on the
reference domain is usually tedious and error-prone but can be carried out automatically
by symbolic processing tools once the expression of the mapping is provided. For this
reason, RBniCS offers automatic pullback capabilities.
T
AN B N UN FN
= , (2.44)
BN 0 PN 0
where we have introduced the reduced quantities, by means of the RB matrix, as follows:
uT u
AN = Z N AZN ,
pT u
BN = ZN BZN ,
uT
FN = ZN F,
p
and ZNu
, ZN are built using either a greedy or a POD training.
As an example, we present the space generation algorithms corresponding to Listings 2.7 and
2.8 for the Stokes problem.
We remark that, in order to deal with a well-posed problem, the assumptions of Theorem
2.3 have to be satisfied. In particular, the algebraic equivalent of the inf-sup condition (2.20) for
the FE discretization, which reads as ker(B T ) = {0}, must be met. Otherwise, all elements in
ker(B T ) are spurious pressure modes. Even if we assume that ker(B T ) = {0} holds at the FE
level, it cannot be the case at the reduced level. Let us call ϕm N an element of the basis of VN
(i.e., represented by a column of ZN u
) and ψN
k
an element of the basis of QN (i.e., represented
p
by a column of ZN ). Then, we obtain BN = [b(ϕm N , ψN )] = 0, because ϕN is obtained as a
k m
linear combination (POD or Gram–Schmidt) of weakly divergence-free snapshots {uN (µN )}.
Thus, ker(BN T
) is definitely larger than {0} and is actually the whole reduced pressure space
QN . Unfortunately, this means that every pressure basis function is a spurious pressure mode. In
the next subsection, we will describe how to handle the inf-sup issue at the reduced level, relying
on a space enrichment approach.
(q, B(µ)v)
FE inf-sup constant: β N (µ) = inf sup ; (2.45)
q6=0 v6=0 kvkV kqkQ
(qN , BN (µ)vN )
RB inf-sup constant: βN (µ) = inf sup . (2.46)
q 6=0 v 6=0
N N
kvN kVN kqN kQN
As we have seen, we currently have βN (µ) = 0 due to the spurious pressure modes. Therefore,
we aim at changing the reduced space VN so that βN (µ) ≥ β N (µ).
Starting from β N (µ), and thanks to the fact that QN ⊂ QN , we can restrict the inf to a
smaller subspace, possibly increasing the infimum value, obtaining
We would like to do the same for the sup argument, owing to the fact that the inclusion still holds
for the velocity spaces VN ⊂ V N . Unfortunately, in general, restricting a sup to a smaller space
might decrease the supremum, unless we were sure that the supremizer sµ (q) belongs the space
VN for every q ∈ QN , i.e.,
and enrich VN as
k
VeN = VN ⊕ span{sµ (ψN ), k = 1, . . . , N }. (2.51)
This allows us to continue the argument discussed earlier, reaching the desired inequality among
the inf-sup constants [511]:
(i) (Zp qN , B(µ)v) (ii) (Zp qN , B(µ)sµ (Zp qN ))
β N (µ) ≤ inf sup = inf
q 6=0 v6=0
N
kvkV kZp qN kQ q 6=0
N
ksµ (Zp qN )kV kZp qN kQ
(iii) (B T (µ)Zp qN , Zu vN )
≤ inf sup
q 6=0 v 6=0
N N
kZu vN kV kZp qN kQ
(iv) (ZuT B T (µ)Zp qN , vN ) (qN , BN (µ)vN )
= inf sup = inf sup = βN (µ),
q 6=0 v 6=0
N N
kZu vN kV kZp qN kQ q 6=0 v 6=0
N N
kvN kVN kqN kQN
40 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
where we have exploited the following facts: (i) QN ⊂ QN ; (ii) the supremizer’s definition; (iii)
the exact enrichment of the velocity space by supremizer solutions, resulting in the space VeN ;
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Examples of CFD applications are the flow in deformable channels governed by the Stokes
system or the behavior of the Navier–Stokes nonlinear model in a backward-facing step channel,
as in the ones shown in Figure 2.12. Let us focus on the latter model, where we set the Reynolds
number to Re = 150, and we want to study the stability of the online problem with respect to
the number of supremizers Ns considered [34].
We report in Table 2.1 the approximation property of the reduced model, considering both
the inf-sup constant and the norm of the reduced pressure.
Table 2.1. Approximation accuracy for the ROM with respect to the number of supremizers exploited.
Figure 2.12. Top: Stokes flow in deformable channels (RBniCS tutorial 12). Bottom: Backward-
facing step for moderate Reynolds, Navier–Stokes equations (RBniCS tutorial 17).
As we can see from the first row in Figure 2.13, if we do not enrich the RB space with the
supremizers, we cannot avoid spurious pressure modes. Numerically, this is also confirmed by
the large value of kpN k2 and the incorrect approximation of the velocity.
If too few supremizers are considered, say 0 < Ns < Np /3, the velocity is correctly approx-
imated, yet the pressure is not recovered accurately (second row in Figure 2.13).
However, if enough supremizers are considered (say, Ns > Np /3), the ROM allows us to get
a better qualitative agreement (both for velocity and pressure) and is inf-sup stable (βN (µ) >
0) (third row in Figure 2.13). It is therefore, in general, not necessary to assume βN (µ) ≥
βh (µ) to obtain good results. As in the previous proof, even considering POD-approximate
supremizer enrichment, the inequality βN (µ) ≥ βh (µ) > 0 holds as long as Ns ≥ Np (last row
in Figure 2.13).
Ns = 0 Ns = 1
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Ns = 5 Ns = 7
Ns = 10 Ns = 15
Ns = 30
Figure 2.13. Velocity and pressure reduced solution for the Navier–Stokes equations in the
backward-facing step channel, varying the number of supremizers Ns .
.
We apply a semidiscretization approach, obtaining the following formulation: ∀k ∈ K =
{1, . . . , K}, find u (µ) ∈ V such that
k
k
u (µ) − uk−1 (µ)
m , v; µ + a(uk (µ), v; µ) = g(tk )f (v, µ) ∀v ∈ V,
∆t
where T has been divided in K subintervals of length ∆t. Then, the discrete problem is recovered
by exploiting a finite difference approximation of the temporal derivative and a space discretiza-
tion via FE. So the finite-dimensional problem reads as follows: ∀k ∈ K, find uN ,k (µ) ∈ V N
such that
N ,k
u (µ) − uN ,k−1 (µ)
m , v; µ + a(uN ,k (µ), v; µ) = g(tk )f (v, µ) ∀v ∈ V N .
∆t
Finally, we can apply model order reduction to the parabolic system, and the reduced problem
reads as follows: ∀k ∈ K, find uN,k (µ) ∈ V N such that
N,k
u (µ) − uN,k−1 (µ)
m , v; µ + a(uN,k (µ), v; µ) = g(tk )f (v, µ) ∀v ∈ V N .
∆t
Let us remark that when dealing with parabolic problems, if we assume that the form a(·, ·)
is stable and that the θm,a
q
(µ), 1 ≤ q ≤ Qm,a , are smooth, then the reduced manifold MN ,K =
{u (µ)| ∀k ∈ K, ∀µ ∈ P} lies on a smooth (P + 1)-dimensional manifold, where P is the
N ,k
5: N ←− N + M2 . Update dimension RB
6: . Perform POD on Z
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
{ζn , 1 ≤ n ≤ N } = P OD(Z, N )
7: XN = span{ζn , 1 ≤ n ≤ N } . Build reduced space
8: µ∗ = arg maxµ∈Ξtrain ∆KN (µ) . Error estimator
9: S∗ ←− S∗ ∪ µ∗ . Chosen parameters
10: end
uN ,k (µntrain )
N ,k
M
uN ,k (µ)
k=0
uN ,k (µ1 )
k=K
P
Note how time t and parameters µ are treated in a different way in the sampling procedure.
Other model reduction techniques may treat them as one “combined” parameter µ e = (µ, t), as
in the proper generalized decomposition (PGD) approach [123].
Let us conclude by highlighting the main differences between the greedy and the POD-greedy
procedure, as implemented in RBniCS (Listings 2.9 and 2.10).
1 def update_basis_matrix (self , snapshot ):
2 basis_functions = self. reduced_problem . basis_functions
3 new_basis_function = self.GS. apply(snapshot , basis_functions )
4 basis_functions . basis_functions . enrich ( new_basis_function )
5 self. reduced_problem .N += 1
Listing 2.9. rbnics/reduction_methods/base/rb_reduction.py.
This section aims at providing explicit expressions for the error estimation employed in the
greedy Algorithms 2.2 and 2.4. Namely, we would like to employ a sharp and inexpensive a
posteriori error bound for kuN (µ) − uN (µ)kV . This is crucial in order to make the greedy
building phase efficient and nondependent on the high-fidelity dimension N . Exploiting an error
estimator in the greedy offline phase, indeed, leads to several advantages, such as the following:
• The time needed to evaluate a new parametric instance decreases in a reduced setting
without losing numerical accuracy, i.e., ∂tcomp (µ → uN (µ)) is small with a rigorous
assessment in the error kuN (µ) − uN (µ)kV for all µ ∈ P.
• A larger training set Σtrain can be used to better explore the solutions of the parametric
problem MN .
• A faster building phase can be performed, which will result in rapid convergence to VN .
In the following we provide explicit expressions of the error estimator ∆N (µ) for several
problems based on different equations, including elliptic, Stokes, and parabolic problems.
Furthermore,
r(v, µ)
||r(·; µ)||V N 0 = sup = ||ê(µ)||V . (2.55)
v∈V N ||v||V
In order to achieve our goal of making ∆N (µ) explicit, we still need the following parametric
quantities:
• The continuous and the discrete coercivity constants of the bilinear form a(·, ·; µ) are
defined, respectively, as
a(v, v; µ)
α(µ) = inf (2.56)
v∈V ||v||2V
and
a(v, v; µ)
αN (µ) = inf . (2.57)
v∈V N ||v||2V
For an elliptic coercive problem such as (2.53) the following relation holds:
N
0 < αLB (µ) ≤ αN (µ) ∀µ ∈ P
where
a(v, w; µ)
γ(µ) = sup sup .
v∈V w∈V ||v||V ||w||V
We now have all the needed quantities to allow us to define the error estimator
||ê(µ)||V
∆en
N (µ) = q . (2.58)
N (µ)
αLB
Concerning the quality of the considered error estimator, we can define the effectivity as
en ∆en
N (µ)
ηN (µ) = ,
||uN (µ)
− uN (µ)||µ
where kvk2µ = a(v, v; µ) is called the energy norm. The effectivity gives information about how
far we are from the exact error we commit by approximating the high-fidelity solution with a
reduced one. Before explaining its role in detail, we want to specify the relation between the
effectivity, coercivity, and continuity constants.
Proof. The proof is straightforward once we recall the symmetric property over the bilinear form
a(·, ·; µ). Indeed, it is a matter of simple computations to show that
Taking v = e(µ) = uN (µ) − uN (µ), applying the Cauchy–Schwarz inequality to a(e(µ), e(µ),
µ), and recalling the definition of the energy norm, we obtain the bilinear form
In addition, exploiting the coercivity property of the bilinear form a(·, ·; µ), we have
p q
||e(µ)||µ = a(e(µ), e(µ); µ) ≥ αN (µ))||e(µ)||V , (2.62)
N (µ)
or
en
ηN (µ) ≥ 1 (2.64)
thanks to the definition of the error estimator and the effectivity. We now want to show the other
side of the inequality (2.59). To this end, let us plug v = ê(µ) into (2.60). Thus, by applying
once again the Cauchy–Schwarz inequality, we have
and thus
||ê(µ)||µ ≤ γ(µ)||e(µ)||µ . (2.67)
It is now clear, exploiting (2.66) and (2.67), that the following relation holds:
∆en
N (µ)
2
||ê(µ)||µ ||e(µ)||µ γ(µ)||e(µ)||2µ
N (µ)
≤ N (µ)
≤ N (µ)
,
αLB αLB αLB
and thus, by definition, we obtain the bound
s p
en ∆en
N (µ)
2 γ(µ)
ηN (µ) = N (µ)||e(µ)||2
≤ q .
αLB µ N
αLB (µ)
The effectivity measures the quality of the considered error estimator. Inequality (2.64) guar-
antees the rigor, while the second inequality guarantees its sharpness: in order to achieve better
performance, the effectivities are desired to be close to unity. It remains to answer to the fol-
lowing question: is the error estimator inexpensive to compute? The main feature to guarantee
an efficient way to build and compute the error estimator is its parameter affine decomposition.
Indeed, the residual can be written as
Moreover, if we consider the equivalent expression exploiting the related Riesz representation
together with the linear superposition property, we obtain
Qa X
X N
(ê(µ), v)V = f (v) − θaq (µ)(uN (µ))n aq (ζn , v).
q=1 n=1
2.4. A Posteriori Error Estimation for Certified RB Methods 47
Qa X
X N
ê(µ) = C + θaq (µ)(uN (µ))n Lqn ,
q=1 n=1
where
(C, v)V = f (v) ∀v ∈ V N ,
while
(Lqn , v)V = −aq (ζn , v) ∀v ∈ V N , 1 ≤ n ≤ N, 1 ≤ q ≤ Qa .
• (C, C)V , (C, Lqn )V , and (Lqn , Lrm )V are assembled and stored for 1 ≤ n, m ≤ Nmax and
1 ≤ q, r ≤ Qa .
We recall that the offline phase still depends on the high-fidelity dimension N ; however, it is
performed only once. After this possibly costly phase, the online stage is performed for each
new parametric instance µ. In this specific stage of the reduction process, the complexity de-
pends on N and Qa but not on N , so that its computational cost scales as O(Q2a N 2 ), and, thus,
many evaluations of the residual norm can be performed in a small amount of time. However, it
still remains to take care of the lower bound of the coercivity constant, which does not depend
on N . Indeed, we now discuss some strategies to compute the lower bound for the coercivity
constant. The lower bound is necessary since, by definition (2.57), the discrete version of the
coercivity constant is equivalent to the solution of the following generalized eigenvalue problem:
find (λ, w) ∈ R × V N such that
The first approach we present is the min-θ approach. It applies to parametrically coercive prob-
lems. Indeed, we first recall that the bilinear form we are dealing with is affine decomposed and
can be written as
XQa
a(v, w; µ) = θaq (µ)aq (v, w),
q=1
where > 0 for all µ ∈ P and a (v, v) ≥ 0 for all v ∈ V and 1 ≤ q ≤ Qa . Thus, taking
θaq (µ) q
w = v, we obtain
Qa
X
a(v, v; µ) = θaq (µ)aq (v, v)
q=1
Qa
X θaq (µ) q 0 q
= θ (µ )a (v, v)
q=1
θaq (µ0 ) a
Qa
θaq (µ) X q 0 q (2.71)
≥ min q θa (µ )a (v, v)
q∈[1,Qa ] θa (µ0 )
q=1
θq (µ)
≥ min qa 0 a(v, v; µ0 )
q∈[1,Qa ] θa (µ )
θq (µ)
= min qa 0 ||v||2V ;
q∈[1,Qa ] θa (µ )
This procedure provides a positive lower bound for the discrete coercivity constant, coinciding
with the exact coercivity constant when µ = µ0 . However, it results in an unsharp bound.
where
Qa
obj
X aq (v, v)
J (µ; v) = θq (µ) .
q=1
||v||2V
2.4. A Posteriori Error Estimation for Certified RB Methods 49
The main idea is to interpret the search for the lower bound as a minimization problem, as
described in (2.72). To this end, we write
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
where
Qa
X
J obj (µ; y) = θq (µ)y q
q=1
and ( )
Qa N aq (vy , vy )
q
Y= y∈R | ∃vy ∈ V : y = , 1 ≤ q ≤ Qa
||vy ||2V
is the set of admissible solutions. The space of admissible solutions can be built using a lower
and an upper bound to the set. Namely, we want to find YU B ⊂ Y ⊂ YLB to restrict or enlarge
the minimization space we are dealing with. Before starting this construction argument, we still
need to define some quantities:
• the box
Qa
" #
Y aq (v, v) aq (v, v)
B= inf 2 , sup 2 ,
v∈V N ||v||V v∈V N ||v||V
q=1
given by the computation of the smallest and the largest eigenvalues of the generalized
eigenvalue problem related to each aq (·, ·), which can be performed only once, in an offline
fashion;
• a finite set of parameters for which we know the value of the coercivity constants, i.e.,
CJ = {µ1SCM , . . . , µJSCM } ⊂ P;
We can now focus on building a lower bound for the admissible set. Indeed, we can define YLB
as
Qa
( )
.
θq (µ0 )y q > αN (µ) ∀µ0 ∈ CJM,µ .
X
YLB = YLB (µ; CJ , M ) = y ∈ RQa | y ∈ B,
q=1
Y ⊂ YLB (µ; CJ , M ) ∀µ ∈ P.
Proof. We show that if y ∈ Y, then y ∈ YLB . Indeed, since y ∈ Y, by definition, there exists
vy ∈ V N such that
aq (vy , vy )
yq = ∀µ ∈ P.
||vy ||2V
Moreover, it is straightforward to see that
namely, y ∈ B. Lastly,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Qa
X a(vy , vy ; µ)
θq (µ)y q = ≥ αN (µ) ∀µ ∈ P,
q=1
||vy ||2V
The next step is to define a lower bound for the discrete coercivity constant. Consider the fol-
lowing linear optimization problem:
N
αLB (µ; CJ , M ) = min J obj (µ; y). (2.73)
y∈YLB
Proposition 2.9. Given CJ ⊂ P and M ∈ N, we can define the coercivity lower bound
N
αLB (µ) ≤ αN (µ) ∀µ ∈ P.
Proof. The statement directly follows by definitions and by Lemma 2.8. Indeed,
N
αLB (µ) = min J obj (µ; y)
y∈YLB (µ)
≤ min J obj
(µ; y) (2.74)
y∈Y
N
= α (µ).
Now we can define the other subset we are interested in, the upper bound for the admissible set,
i.e.,
.
YU B = YU B (µ; CJ , M ) = {y ∗ (µ0 ) ∈ Y | µ0 ∈ CJM,µ },
where
y ∗ (µ) = arg inf J obj (µ; y).
y∈Y
The only point left to clarify is how to build the set of parameters CJ . As specified in Algo-
rithm 2.9, a greedy approach is applied in order to iteratively enrich CJ , building an approxima-
tion of the coercivity constant which is more precise at each step of the procedure. As already
specified, the whole SCM approach can be efficiently divided into an offline and an online phase.
• Offline: We build, in a greedy fashion, the lower and upper bounds for the spaces. Here,
Jmax problems of the form (2.73) are performed, resulting in 2Qa + Jmax eigenproblems
over V N . This procedure allows us to build the box B and the set of coercivity constants
{αN (µ0 ) |µ0 ⊂ CJmax } which are the elements defining YLB . It still remains to construct
the upper bound YLB , which is given by Jmax Qa inner-product evaluations over V N .
2.4. A Posteriori Error Estimation for Certified RB Methods 51
• Online: Given µ ∈ P, the set CJmax is sorted in order to find the M µ-nearest parameters
CJM,µ
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
max
through (M + 1)Qa evaluations of the map µ0 → θq (µ0 ) and the related M evalua-
tions of αN (µ0 ). In the end, the lower bound of the coercivity constant is computed using
(2.73), and its cost does not depend on N .
In this subsection, we focused on error estimators for linear coercive problems. However,
other results can be proved for several state equations. Next, we will extend the concept of the
error estimator to more complex problems, such as Stokes and parabolic equations.
• the Brezzi inf-sup constant β N (µ) and its lower bound βLB
N
(µ) defined over the continuity
equation as
. b(v, q; µ)
β N (µ) = inf sup N
≥ βLB (µ) > 0;
q∈Q v∈V kvkV kqkQ
• the separated residuals with respect to the momentum and the continuity equations:
All these ingredients combined will result in an error estimator for the Stokes problem. Indeed,
the following proposition holds [223].
52 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
Then
k(uN , pN )(µ) − (uN , pN )(µ)kV ×Q ≤ ∆Ba
N (µ). (2.77)
The proof follows from standard Banach–Nečas–Babuška theory [412], i.e., an extension of
Lax–Milgram techniques to noncoercive problems. We recall that the Stokes case is a special
case of noncoercive problems. However, the treatment of more complicated and general equa-
tions goes beyond the scope of this contribution. Focusing on Stokes equations, the estimator
∆BaN (µ) is typically used in the greedy algorithm and allows us to extend the techniques to a
very common and useful model in fluid dynamics. However, it gives no insight into the total
error, since it is divided among velocity and pressure components. Thus, an alternative can be
employed.
Let us define
• a lower bound αLB (µ) of the coercivity constant α(µ) of a(·, ·; µ), and
Then
kuN (µ) − uN (µ)kV ≤ ∆uN (µ),
kpN (µ) − pN (µ)kQ ≤ ∆pN (µ).
We do not report the proof of this statement; however, once again, the interested reader may
refer to [223].
uN ,k N ,k−1
N (µ) − uN
(µ)
rk (v, µ) = g(tk )f (v, µ) − m , v; µ − a(uN ,k
N (µ), v; µ)
∆t
(2.78)
for all v ∈ V N and for all k ∈ K. Then, if we consider for each time instance
v
u k
u ∆t X
∆kN (µ) = t LB LB
(N (tj ; µ))2 ,
α (µ)σ (µ) j=1
2.5. Nonaffine and Nonlinear Problems 53
||uN ,k (µ) − uN ,k k
N (µ)||L2 (Ω) ≤ ∆N (µ).
In this subsection we proposed an overview on a posteriori error bounds that are crucial quantities
in the efficient application of the greedy algorithm. Here we focus on their expression based,
essentially,
This guarantees an efficient offline-online decomposition for several problems governed by dif-
ferent equations, such as
• elliptic problems,
• parabolic problems.
In the next section we address more complicated parametric problems. Indeed, many applica-
tions, from sciences to industry, rely on models which can be nonlinear or that might fail the
affine decomposition assumption (2.34). Thus, to tackle them, we present some advanced tech-
niques aimed at solving these more involved models in a faster way and adapting the classical
knowledge about ROMs to these frameworks.
Q Qf
X X
a(u, v; µ) = θqa (µ)aq (u, v), F (v; µ) = θqf (µ)Fq (v).
q=1 q=1
This allows us to factor out any µ-dependent terms θq· (·) and the integrals aq (·, ·) or fq (·) on
the domain, so that, after the offline phase, a Galerkin projection of the parameter-independent
quantities aq (·, ·) or fq (·) can be performed, precomputing reduced order matrices AqN and vec-
tors FNq .
During the online stage, once given a specific value of µ, the corresponding reduced order
left-hand and right-hand sides can be obtained as
Q Qf
θqa (µ)AqN , θqf (µ)FNq .
X X
AN (µ) = FN (µ) =
q=1 q=1
Due to its importance, this form is usually assumed in many model reduction methods (e.g., the
separability requirements in PGD [123]).
54 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
Z
F (v; µ) = g(x; µ)v(x) dΩ. (2.79)
Ω
It is clear that, if the integrand satisfies g(x; µ) = g1 (x)g2 (µ), we easily obtain an affine
(parameter-separable) term
Z
F (v; µ) = g2 (µ) g1 (x)v(x) dΩ. (2.80)
Ω
However, there are several relevant cases in which this assumption does not hold, such as non-
affine (non-parameter-separable) and nonlinear integrands, respectively:
Z
F (v; µ) = g(x; µ)v(x) dΩ for a generic g(x; µ) (2.81)
Ω
and Z
F (v; u(µ), µ) = g(x; u(x; µ), µ)v(x) dΩ. (2.82)
Ω
We remark that in some situations one can rely on a workaround to prevent the lack of affine
expansion. If we consider polynomial nonlinearities such as
A similar formula holds for the nonlinear term of the Navier–Stokes system with p = 2 and
is commonly used to recover the computational efficiency. Let us describe this process in more
detail. Let j = [j1 , . . . , jp ] be a multi-index of size p, and ζi the ith RB function (RBF), where
each index i, j1 , . . . , jp varies in the range {1, . . . , N }. At the end of the offline stage, one can
preassemble the tensor of order p + 1, given by
We remark that the tensor is dense, so it might be a challenge to store it in memory if p is very
large or if N is not small enough. Instead, during the online stage, as usual, we can expand the
reduced solution uN as
XN
uN (µ) = [uN ]j ζj
j=1
where we have exploited the multilinearity of t in the first p arguments and then the multilinearity
of t in its (p + 1)th argument to choose v = ζi for any i = 1, . . . , N . Finally, we can simply
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
recover the offline-online decomposition by using the (p + 1)th order tensor Ti,j .
Even if the case of polynomial nonlinearities occurs frequently (e.g., incompressible fluid
dynamics, optimal control problems) and allows us to obtain an exact parameter-separable rep-
resentation, it is certainly not the most general one.
Therefore, in the general case, we will resort to approximate parameter-separable represen-
tations such as
Z Qf
X
F (v; µ) = g(x; µ)v(x) dΩ ≈ θqf (µ)Fq (v).
Ω q=1
Building an approximation like the one above is an active field of research in the ROM com-
munity, which goes under the name “hyperreduction.” One of the most widespread of these
hyperreduction techniques is the empirical interpolation method (EIM) [40], which allows us to
obtain an efficient offline-online decoupling when dealing with nonaffine and nonlinear systems.
In the next subsection, we will give further details on this affine recovery technique.
Once this representation is known, one can easily obtain an affine representation of the integral
form Z
F (v; µ) = g(x; µ)v(x) dΩ, (2.85)
Ω
i.e.,
Q
X Z
F (v; µ) = hq (µ)F q (v), where F q (v) = cq (x)v(x) dx. (2.86)
q=1 Ω
Of course, the same decomposition can be formally extended to the nonlinear case with
where cq are parameter-dependent coefficients and the basis functions hq are selected as a linear
combination of snapshots of the parametrized functions g(µ1 ), . . . , g(µQ ).
56 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
This means that, once the basis functions hq (x) are set, the problem of finding the coefficients
c (µ) is solved by imposing the interpolation condition, i.e.,
q
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Q
X
IQ g(xq ; µ) = cq (µ)hq (xq ) = g(xq ; µ), q = 1, . . . , Q. (2.89)
q=1
We remark that the above problem can be recast in matrix form as T c(µ) = g(µ), with
1: k = 1
2: I0 g(x; µ) = 0 . Initialization
3: while (k < Nmax ) . Build space hierarchy
4: µk = arg max kg(·; µ) − Ik−1 g(·; µ)kL∞ (Ω) . Greedy parameter selection
µ∈P
5: r(x) = g(x; µk ) − Ik−1 g(x; µk ) . Residual assembly
6: xk = arg max |r(x)| . Find interpolation point
x∈Ω
7: hk (x) = r(x)/r(xk ) . Residual normalization
8: εk = krkL∞ (Ω) = r(xk ) . Tolerance threshold computation
9: if (εk < ε) . Stopping criterion
10: break
11: else
12: k =k+1
Finally, in order to obtain a practical version of the EIM, it is common to replace the param-
eter space P with the discrete training set Ptrain . In the same spirit, the L∞ (Ω) computations
(three times) are replaced by evaluations of maximums over a discrete set (e.g., mesh nodes or
quadrature points).
Moreover, g(·; µ) is usually precomputed for every µ ∈ Ptrain , since this evaluation does
not change with the iteration counter k, while the right-hand sides of interpolation conditions are
usually computed on a reduced mesh that contains only the patches of xk .
Let us remark that the greedy computation of basis functions hk may be replaced by a POD
compression of snapshots g(µ1 ), . . . , g(µQ ), similarly to RB generation.
The algorithm can easily be adapted to the discrete case, where g is not an analytical function
but the result of the discretization of a linear function (vector) or a bilinear form (matrix). The
role of x will be played by an index (vector) or a pair of indices (matrix). The resulting method
is known in the literature as the discrete empirical interpolation method [116].
Figure 2.15. Left: RB temperature solution. Right: computational mesh for Ω = (−1, 1)2 with
EIM interpolation points.
conductivity coefficient is fixed at one, while the heat source is characterized by the expression
The parameter vector µ, given by µ = (µ0 , µ1 ) ∈ [−1, 1]2 , affects the center of the Gaussian
source g(x; µ), which could be located at any point in Ω. In order to more quickly (and provably
accurately) evaluate the problem, we propose to use a certified RB approximation for the prob-
lem. In order to preserve the affinity assumption (for the sake of performance), the EIM will be
used on the forcing term g(x; µ).
Thus, let u(µ) be the temperature in the domain Ω. The weak formulation for this problem
is
a (u(µ), v; µ) = f (v; µ) ∀v ∈ V,
and
Z
f (v; µ) = g(µ)v dx.
Ω
In Figure 2.15 we show the computational mesh used for the high-fidelity simulation, the
solution corresponding to a particular value in the parameter space P, and the interpolation points
selected by the EIM.
Finally, we show in Figure 2.16 the behavior of the average relative error with respect to
the number of EIM and RB functions. As we can see, even in the nonaffine case we obtain
exponential convergence of the RB method.
58 Chapter 2. Finite Element–Based Reduced Basis Method in CFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 2.16. Convergence of the RB method for the elliptic problem with Gaussian source, with
respect to the number of basis functions, for EIM (left) and RB (right).
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 3
Certified Smagorinsky
Reduced Basis
Turbulence Model
3.1 Introduction
In this chapter we start by presenting a reduced basis (RB) Smagorinsky turbulence model (cf.
[541]) for steady flows, which is the basic large eddy simulation (LES) turbulence model, in
which the effect of the subgrid scales on the resolved scales is modeled by eddy diffusion terms
(cf. [111, 515]). It is an intrinsically discrete model, since the eddy viscosity term depends
on the mesh size. Although it is well known that the Smagorinsky model is overdiffusive, the
construction of this RB Smagorinsky turbulence model represents a first step that is complex
enough and essential for the RB models presented later on.
We start by defining the finite element (FE) problem, and then, by properly selecting snap-
shots with the greedy algorithm, we obtain the low-dimensional spaces for which we define the
Smagorinsky RB model. The eddy viscosity is modeled by a nonlinear term that depends on the
mesh size and the modulus of the velocity gradient. We approximate this nonlinear term using the
empirical interpolation method (EIM; cf. [40, 245]) in order to obtain a linearized decomposition
of the RB Smagorinsky model.
Then, we present an RB Boussinesq model, with variational multiscale (VMS)-Smagorinsky
modeling for both the eddy viscosity and the eddy diffusivity. This model takes into account
the buoyancy forces present in natural convection problems. This model differs from the ones
presented in Section 3.2 because it considers the energy equation coupled with the momentum
and continuous equations. In this section, we assume that the temperature of the fluid becomes
essential in the flow of the fluid considered. We perform the RB model of Boussinesq VMS-
Smagorinsky equations with the same strategy presented in Section 3.2, developing an a poste-
riori error bound, which is essential for the greedy algorithm used in snapshot selection. Again,
we take into account the EIM in order to approximate the nonlinear terms for the eddy viscosity
and eddy diffusivity.
Finally, we consider the application of the RB Boussinesq VMS-Smagorinsky in a variable-
height cavity domain. The variability of the cavity height is considered through a geometrical
parametrization on the domain. Since we are interested in efficiently solving the parameter-
dependent problem, we need to reformulate the Boussinesq VMS-Smagorinsky model in a pa-
rameter-independent domain with a change of variables. This change of variables leads us to
59
60 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
obtain operators that depend on the geometrical parameter in the same way that they depend on
the physical parameters. This setting lets us decompose affinely the operators with respect to the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
parameters, both physical and geometrical, letting us store parameter-independent matrices and
tensors in the offline phase.
The chapter is structured as follows: in Section 3.2 we first present the Smagorinsky RB
model, and Subsection 3.2.1 is devoted to the FE problem. Then, in Subsection 3.2.2, we present
the greedy algorithm used to construct the RB spaces, whereas in Subsection 3.2.3 we describe
the EIM for the approximation of the nonlinear terms. To finish this first section, we present
the numerical analysis for the well-posedness of the FE problem in Subsection 3.2.4 with the
construction of an a posteriori error estimator in Subsection 3.2.5 and the numerical test for
this problem in Subsection 3.2.6. Then, in Section 3.3, we introduce the VMS-Smagorinsky
model, with the FE model in Subsection 3.3.1 and the RB model in Subsection 3.3.2. Again, we
present the well-posedness analysis (Subsection 3.3.3), the construction of the a posteriori error
estimator (Subsection 3.3.4), and the numerical results (Subsection 3.3.5). Finally in Section 3.4,
we describe the geometrical parametrization of the VMS-Smagorinsky model, with the problem
setting in Subsection 3.4.1 and the numerical tests in Subsection 3.4.2.
3.2.1 FE Problem
To formulate the Smagorinsky turbulence model, let Ω be a bounded polyhedral domain in
Rd , (d = 2, 3). We assume that its boundary is split into Γ = ΓD ∪ ΓN , where ΓD = ΓDg ∪ ΓD0
is the boundary relative to the homogeneous and nonhomogeneous Dirichlet boundary conditions
and ΓN is the boundary relative to the Neumann conditions.
Let {Th }h>0 a family of affine-equivalent and conforming triangulations of Ω, formed by
triangles or quadrilaterals (d = 2) or tetrahedra or hexahedra (d = 3). As usual, the parameter
h is the maximum diameter hK among the elements K ∈ Th . Given an integer l ≥ 0 and an
element K ∈ Th , we denote by Rl (K) either Pl (K), the space of Lagrange polynomials of
degree ≤ l, defined on K, if the grids are formed by triangles (d = 2) or tetrahedra (d = 3),
or Ql , the space of Lagrange polynomials of degree ≤ l on each variable, defined on K, if the
family of triangulations is formed by quadrilaterals (d = 2) or hexahedra (d = 3).
Although the Smagorinsky model is intrinsically discrete, it can be interpreted as a discretiza-
tion of a continuous model. We next present this model to clarify its relationship with the Navier–
Stokes one. In this way, the “continuous” Smagorinsky turbulence model is formulated as
1
w · ∇w + ∇p − ∇ · + ν (w) ∇w = f in Ω,
T
Re
in Ω,
∇·w=0
w = gD on ΓDg , (3.1)
on ΓD0 ,
w=0
1 ∂w
on ΓN ,
−pn + + νT (w) =0
Re ∂n
3.2. RB Smagorinsky Turbulence Model 61
where here Re is the Reynolds number, w is the velocity field, and p is the pressure, with the
latter two depending on the Reynolds number. Moreover, the eddy diffusion term is given by
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
X
νT (w) = CS2 h2K ∇w|K χK , (3.2)
K∈Th
where · denotes the Frobenius norm in Rd×d and CS is the Smagorinsky constant (cf. [512]).
For this problem, the parameter considered is the Reynolds number, denoted in the following
by µ ∈ D.
Let us consider the spaces Y = {v ∈ H 1 (Ω) : v|ΓD = 0} for velocity and M = L2 (Ω) for
pressure. We assume that there exists a lift function uD ∈ (H 1 (Ω))d such that uD |ΓDg = gD ,
uD |ΓD0 = 0, and ∇ · uD = 0 in Ω. With these conditions, we ensure that the reduced velocity
u = w − uD is still incompressible and satisfies the homogeneous Dirichlet boundary conditions
on ΓD . We will assume that f ∈ (L2 (Ω))d and gD ∈ (H 1/2 (Ω))d .
Let Yh ⊂ Y and Mh ⊂ M be two finite subspaces of Y and M . We consider the following
variational discretization of problem (3.1), actually, the “true” Smagorinsky model:
the trilinear form c(·, ·, ·; µ) and the nonlinear Smagorinsky term aS (·; ·, ·; µ) are given by
Z Z
c(z, u, v; µ) = (z · ∇u)v dΩ, aS (z; u, v; µ) = νT (z)∇u : ∇v dΩ. (3.5)
Ω Ω
where h·, ·i stands for the duality pairing between Y 0 and Y , with Y 0 being the dual space of Y .
The solution of problem (3.3) is intended to approximate the large-scale component of the
solution of the Navier–Stokes problem (i.e., problem (3.1) with νT = 0).
Let us define the norms relative to the spaces Y and M . For the velocity space Y , we consider
a weighted inner product (·, ·)T defined as
Z
1
(u, v)T = + νT∗ ∇u : ∇v dΩ ∀u, v ∈ Y, (3.6)
Ω µ
where νT∗ = νT (w(µ)), µ = arg minµ∈D K∈Th (CS hK )2 minx∈K |∇w(µ)|(x)χK (x), and
P
w(µ) is the velocity solution of (3.1).
1/2
This inner product induces a norm linked to the eddy diffusion term k · kT = (·, ·)T . Since
the functions of Y vanish on ΓD , this norm is equivalent to the usual H -norm. This norm will
1
turn out to be crucial to apply our error estimator in the RB construction by the greedy algorithm.
For the pressure space M , we will use the usual L2 -norm.
62 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
q
kU kX = kuk2T + kpu k20,2,Ω ∀U = (u, pu ) ∈ X. (3.7)
With this notation, we can rewrite the variational problem (3.3) as follows:
A2 (U ; V ) = (u · ∇u)v dΩ,
ZΩ
A3 (U ; V ) = νT (u + uD ) ∇(u + uD ) : ∇v dΩ.
Ω
In practice, the error kUh (µ) − UN (µ)k may be hard to compute numerically because we
have to compute the solution of the FE problem, uh (µ), for all µ ∈ Dtrain . Thus, rather than
computing the exact error, we consider an a posteriori error bound, ∆N (µ), that is cheaper to
compute than the exact error.
The algorithm resulting from the substitution of the real error by the a posteriori error bound
is usually called the weak greedy algorithm (cf. [60, 470]) and is the one that is commonly used
for the RB method.
In each iteration of the (weak) greedy algorithm, we actualize the set of parameter values
SN = {µ1 , . . . , µN } and the reduced spaces MN = span{ξkp := puh (µk ), k = 1, . . . , N } and
3.2. RB Smagorinsky Turbulence Model 63
YN = span{ζ2k−1 v
:= uh (µk ), ζ2kv
:= Tpµ ξkp , k = 1, . . . , N } for N = 1, . . . , Nmax , with
Nmax the maximum number of bases that we consider in our problem. With this procedure, we
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
3. Compute Uh (µk ), define the reduced spaces Mk and Yk by adding the newly computed
snapshots, and orthonormalize them.
4. Stop if max ∆k (µ) < εRB . If not, go back to step 2.
µ∈Dtrain
Here σk (µ) for k = 1, . . . , M is the solution of a lower-triangular linear system, where the
second member is the value of g(x; wh (µ)) at certain points xi . Furthermore,
X Z
âS (qk , wN , vN ) = (CS hK )2 qk ∇w : ∇v dΩ. (3.15)
K∈Th K
This technique allows us to linearize the eddy viscosity term. Thus, the RB problem with this
last approximation of the Smagorinsky term is defined as follows:
Find (uN , pN ) ∈ YN × MN such that
a(uN , vN ; µ) + b(vN , pN ; µ) + âS (wN ; vN ; µ)
+ c(uD , uN , vN ; µ) + c(uN , uD , vN ; µ) (3.16)
+ c(u , u , v ; µ) = F (v ; µ) ∀v ∈ Y ,
N N N N N N
b(uN , qN ; µ) = 0 ∀qN ∈ MN .
64 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
The solution (uN (µ), pN (µ)) ∈ XN of (3.16) can be expressed as a linear combination of the
basis functions
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
2N N
p
X X
uN (µ) = uN v
j (µ)ζj , pN (µ) = pN
j (µ)ξj .
j=1 j=1
Taking into account this representation, for the bilinear terms in (3.16), during the offline phase
we store the parameter-independent matrices, as in [387], defined as
For the convective and Smagorinsky terms, we need to store a parameter-independent tensor
of order three during the offline phase, defined as
Thanks to this, we are able to solve problem (3.16), linearized by a semi-implicit evolution
approach. Note that the treatment of the approximation of the eddy viscosity term in the offline-
online phase is similar to the treatment of the convective term, as a consequence of the tensoriza-
tion.
For the well-posedness of the problem, we have to guarantee the uniform coerciveness and
boundedness of ∂1 A in the sense that, for any solution Uh (µ) of (3.8), there exist β0 > 0 and
3.2. RB Smagorinsky Turbulence Model 65
γ0 ∈ R such that, ∀µ ∈ D,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Then, according to the BRR theory (cf. [76, 93]), it will follow that in a neighborhood of Uh (µ)
the solution of (3.8) is unique and bounded in k · kX in terms of the data.
We have the following results.
Proposition 3.2. Let C ? = CT2 (Cµ + 1). Suppose that k∇uD k0,2,Ω < 1
C? and k∇uh k0,2,Ω ≤
˜
C ? − k∇uD k0,2,Ω . Then, there exists βh > 0 such that
1
b(vh , ph ; µ)
αkqh k0,2,Ω ≤ sup ,
vh ∈Yh kvk1,2,Ω
thanks to Proposition 3.2, we can prove that the operator ∂1 A satisfies the inf-sup condition in
(3.19). See [106] for more details.
Observe, also, that since kgD k1/2,ΓD ≤ kwh k1,Ω ≤ CΩ k∇wh k0,2,Ω , the condition needed
CΩ
in Proposition 3.2, k∇wh k0,2,Ω ≤ C 2 (C1µ +1) , will only be possible if kgD k1/2,ΓD ≤ C 2 (Cµ +1)
;
T T
thus the Dirichlet boundary data should be sufficiently small.
such that
∂1 A(UN (µ), Vh ; µ)(Zh )
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Theorem 3.4. Let µ ∈ D, and assume that βN (µ) > 0. If problem (3.8) admits a solution Uh (µ)
such that
βN (µ)
kUh (µ) − UN (µ)kX ≤ ,
ρT
then this solution is unique in the ball BX UN (µ), βNρT(µ) .
At this point, we are in a position to define the a posteriori error bound estimator by
βN (µ) h p i
∆N (µ) = 1 − 1 − τN (µ) , (3.26)
2ρT
Theorem 3.5. Assume that βN (µ) > 0 and τN (µ) ≤ 1 for all µ ∈ D. Then there exists a
unique solution Uh (µ) of (3.8) such that the error with respect to UN (µ), the solution of (3.16),
is bounded by the a posteriori error bound estimator, i.e.,
with effectivity
2γN (µ)
∆N (µ) ≤ + τN (µ) kUh (µ) − UN (µ)kX . (3.29)
βN (µ)
ΓDg
ΓD0 ΓD0
ΓD0
10 -1
max kg(µ) − IM [g(µ)]k∞
µ∈D
-2
10
10 -3
-4
10
0 5 10 15 20 25
M
In this numerical test, we need Mmax = 22 bases in the EIM algorithm until it reaches the
tolerance for the relative error of εEIM = 5 · 10−4 . In Figure 3.2, we show the convergence of
the EIM algorithm.
For the greedy algorithm, we prescribe a tolerance of εRB = 5 · 10−5 . This tolerance is
reached for Nmax = 12 basis functions. Note that, in this case, N = 8 bases are needed in order
to ensure that τN (µ) < 1 for all µ in D. In Figure 3.3 (left) we show the convergence of the
greedy algorithm, and in Figure 3.3 (right) we show the value of the error and the a posteriori
error bound for all µ in D.
In Figure 3.4 we compare the FE velocity solution (left) and the RB velocity solution (right)
for a chosen parameter value µ = 4521. Both images are practically equal, as the error between
the solutions is of order 10−7 .
Finally, we show in Table 3.1 a summary of the results obtained for several values of µ in D.
For this test, we observe a dramatic speedup in the computation of the numerical solution. These
large speedup factors are possibly due to the highly turbulent levels of viscosity introduced by
the Smagorinsky turbulence model. The offline phase of this test took approximately two days
to complete.
68 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
10 6 10 -3
max τN (µ) ∆N (µ)
µ∈D
kUh (µ) − UN (µ)kX
10
4 max ∆N (µ) 10 -4
µ∈D
-5
10
2
10
10 -6
0
10
-7
10
-2
10
-8
10
-4
10
10 -9
10 -6 10 -10
1 2 3 4 5 6 7 8 9 10 11 12 1000 1500 2000 2500 3000 3500 4000 4500 5000
N Re
Figure 3.3. Convergence of the greedy algorithm (left) and the value of ∆Nmax (µ) and the exact
error between the FE solution and the RB solution (right).
Table 3.1. Computational time for FE solution and RB online phase, with the speedup and the
relative error.
In this section we present a Boussinesq model for natural convection flow problems. We deal
with a square cavity problem, for which the top and the bottom are isolated walls and the vertical
walls have a prescribed temperature. The difference in temperatures in the vertical wall produces
a circulation of the fluid, which is supposed to be viscous, Newtonian, and incompressible, and
the buoyancy forces are modeled with a Boussinesq approximation. The proof of the main results
of this section can be found in [30, 151].
3.3.1 FE Problem
We next present the Navier–Stokes equations with the Boussinesq approximation for the energy
equation (cf. [52, 66, 67, 109]), for which we also include the modeling of the eddy viscosity
and the eddy diffusivity by the Smagorinsky approach (see [512, 541]). We present a continuous
form of the Smagorinsky eddy viscosity and eddy diffusivity:
u · ∇u − P r∆u − ∇ · (νT (u)∇u) + ∇p − P r Ra θ ed = f in Ω,
in Ω,
∇·u=0
in Ω,
u · ∇θ − ∆θ − ∇ · (KT (u)∇θ) = Q
(3.30)
u=0 on Γ,
on ΓD ,
θ = θD
on ΓN .
∂n θ = 0
Here, u is the velocity field, p is the pressure, and θ is the temperature. In addition, ed
is the last vector of the canonical basis of Rd , while Ra and P r are the Rayleigh and Prandtl
dimensionless numbers, respectively. Both the external body forces f and the heat source term Q
are given data for the problem. In (3.30), νT (u) is the eddy viscosity term defined in (3.2), and
KT (u) is the eddy diffusivity, given by
1
KT (u) = νT (u). (3.31)
Pr
The Prandtl number is defined as
ν0
P r := ,
k0
with ν0 a reference kinematic viscosity and k0 a reference thermal diffusivity. Since the Prandtl
number depends on the physics of the fluid, we consider it as a fixed value and not as a parameter.
Thus, for this problem, we consider the Rayleigh number as the only parameter for the model,
denoted by µ ∈ D ⊂ R.
To define the variational form of problem (3.30), let us define the spaces Y = (H01 (Ω))d
for velocity, Θ = H01 (Ω) for temperature, and M = L20 (Ω) for pressure. We consider the H 1 -
seminorm for the velocity and temperature spaces and the L2 -norm for the pressure space. In
addition, let us define the Sobolev embedding constants Cu and Cθ associated with these norms,
such that
kvk0,4,Ω ≤ Cu k∇vk0,2,Ω ∀v ∈ Y (3.32)
and
kθk0,4,Ω ≤ Cθ k∇θk0,2,Ω ∀θ ∈ Θ. (3.33)
Moreover, we consider the tensor space X = Y × Θ × M , with the following associated norm:
kU k2X = k∇uk20,2,Ω + k∇θk20,2,Ω + kpk20,2,Ω ∀U = (u, θ, p) ∈ X. (3.34)
70 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
To set the VMS framework, we decompose the velocity, pressure, and temperature spaces,
Y , M , and Θ, respectively, as
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Y = Yh ⊕ Y 0 , M = Mh ⊕ M 0 , Θ = Θh ⊕ Θ0 ,
where Yh , Mh , and Θh are, respectively, the large-scale finite dimensional spaces for velocity,
pressure, and temperature, while Y 0 , M 0 , and Θ0 are the small-scale complementary spaces. We
suppose that the large and small scales are separated by assuming that the sum is direct, i.e.,
Yh ∩ Y 0 = {0}, Mh ∩ M 0 = {0}, and Θh ∩ Θ0 = {0}.
The VMS modeling is a discretization of a set of macro-micro scale equations derived from
the previous decomposition (see [111] for more details), based upon the following procedure:
(i) Approximate the small-scale spaces Y 0 , M 0 , and Θ0 by finite-dimensional subspaces of
small resolved scales Yh0 , Mh0 , and Θ0 , respectively. Then, Y 0 = Yh0 ⊕Y 00 , M 0 = Mh0 ⊕M 00 ,
and Θ0 = Θ0h ⊕Θ00 , where Y 00 , M 00 , and Θ00 are complementary spaces of small unresolved
scales of finite dimension. This yields the unique decompositions
We can also define Y h = [VHl (Ω)]d , where VHl (Ω) is a subspace of Vhl (Ω) with a larger grid
size H > h; typically, H = 2h or H = 3h.
With that notation, we are identifying Y h as the large-scale velocity space and Yh0 = (Id −
Πh )Yh as the resolved small-scale velocity space and analogously for the temperature and pres-
sure spaces. Denoting the Rayleigh number by µ ∈ R, we obtain the following variational
formulation of problem (3.30):
Here, the bilinear forms au (·, ·; µ), aθ (·, ·; µ), b(·, ·; µ), and f (·, ·; µ) are defined by
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Z Z
au (u, v; µ) = P r ∇u : ∇v dΩ, aθ (θu , θv ; µ) = ∇θu · ∇θv dΩ,
Ω Ω
Z Z (3.37)
f (θu , v; µ) = −P r µ θu v2 dΩ, b(u, pv ; µ) = − (∇ · u)pv dΩ;
Ω Ω
and the nonlinear VMS-Smagorinsky term for eddy viscosity, a0Su (·; ·, ·; µ), is given by
Z
a0Su (z; u, v; µ) = νT (Π∗h z)∇(Π∗h u) : ∇(Π∗h v) dΩ, (3.39)
Ω
with
νT (u) = (CS hK )2 |∇u|K |.
For the eddy diffusivity term, let us first introduce a mollifier φ ∈ Cc∞ (R), with supp(φ)⊂
B(0, 1), φ ≥ 0, kφk0,1,R > 0, and φ even, i.e., φ(−x) = φ(x). Let us consider the mollifier
sequence {φn (x)}n≥1 , with φn ∈ Cc∞ (R), supp(φn )⊂ B(0, 1/n), φn ≥ 0, kφn k0,1,R = 1,
defined by
n
φn (x) = φ(n x).
kφk0,1,R
The following result for mollifiers can be found in [73].
Proposition 3.6. Assume f ∈ Lp (Rd ) with 1 ≤ p < ∞. Then (φn ∗ f ) converges to f when
n → ∞ in Lp (Rn ).
with
νT,n (u) = (CS hK )2 (φn ∗ |∇u|K |).
Thanks to Proposition 3.6, it holds that a0Sθ,n converges uniformly to a0Sθ , with
Z
a0Sθ (uh ; θhu , θhv ; µ) = νT (Π∗h uh )∇(Π∗h θhu ) · ∇(Π∗h θhv ) dΩ.
Ω
The eddy diffusivity a0Sθ,n is considered for the well-posedness analysis and the development
of an a posteriori error bound estimator. In practice, we consider the eddy diffusivity term in the
Boussinesq VMS-Smagorinsky model as a0Sθ .
Writing Xh = Yh × Θh × Mh , we rewrite problem (3.36) as follows:
Defining Vh = (vh , θhv , pvh ) ∈ Xh , the µ-independent operators in (3.42) are given by
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Z Z
A0 (Uh , Vh ) = Pr ∇uh : ∇vh dΩ + ∇θhu · ∇θhv dΩ
Z Ω Z Ω
3.3.2 RB Problem
In this section, we present the RB method for the model presented in Subsection 3.3.1. The RB
problem is as follows:
YN = span{ζ2k−1
v
:= uh (µk ), ζ2k
v
:= Tpµ ξkp , k = 1, . . . , N }, (3.46)
where here, again, we are denoting by Tpµ ξkp the inner pressure supremizer that assures the inf-sup
stability of the RB problem.
With this notation, the RB solutions of problem (3.44) can be represented as
2N N N
p
X X X
uN (µ) = uN v
j (µ)ζj , θN (µ) = θjN (µ)ϕθj , pN (µ) = pN
j (µ)ξj .
j=1 j=1 j=1
The snapshots for the construction of the RB spaces are given by the greedy algorithm explained
in Subsection 3.2.2.
in (3.43), we obtain
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Z
∂1 A3 (U ; V )(Z) = νT (Π∗h u)∇(Π∗h z) : ∇(Π∗h v) dΩ
Ω
X Z ∇(Π∗h u) : ∇(Π∗h z)
+ (CS hK )2 ∇(Π∗h u) : ∇(Π∗h v) dΩ
K |∇(Π∗h u)|
K∈Th
1
Z
+ νT,n (Π∗h u) ∇(Π∗h θhz ) · ∇(Π∗h θhv ) dΩ
Pr Ω
1
Z
+ ∂1 νT,n (Π∗h u)(Π∗h z) ∇(Π∗h θu ) · ∇(Π∗h θv ) dΩ,
Pr Ω
where
∂1 νT,n (u)(z) = (CS hK )2 (φ0n ∗ |∇u|K |) : ∇z .
The existence of γ0 ∈ R and β0 > 0 satisfying (3.48) and (3.49), respectively, is given by
the following results.
Proposition 3.8. If the Dirichlet boundary data is sufficiently small, then there exists β̃(µ) > 0
such that
b(vh , ph ; µ)
αkph k0,2,Ω ≤ sup ,
vh ∈Yh k∇vh k0,2,Ω
we can prove that the inf-sup (3.49) is satisfied thanks to Proposition 3.8.
74 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
In this section, we present the development of the a posteriori error bound estimator, needed in
the greedy algorithm to construct the reduced spaces for velocity, temperature, and pressure. For
this purpose, we deal with the BRR theory [76].
With the following result, we prove that the directional derivative ∂1 A is Lipschitz con-
tinuous.
Lemma 3.9. There exists a positive constant ρn such that ∀Uh1 , Uh2 , Zh , Vh ∈ Xh ,
βN (µ) h p i
∆N (µ) = 1 − 1 − τN (µ) , (3.52)
2ρn
Theorem 3.10. Let µ ∈ D, and assume that βN (µ) > 0. If problem (3.41) admits a solution
Uh (µ) such that
βN (µ)
kUh (µ) − UN (µ)kX ≤ ,
ρn
then this solution is unique in the ball BX UN (µ), βNρn(µ) .
Moreover, assume that τN (µ) ≤ 1 for all µ ∈ D. Then there exists a unique solution Uh (µ)
of (3.41) such that the error with respect to UN (µ), the solution of (3.44), is bounded by the a
posteriori error bound estimator, i.e.,
with effectivity
2γN (µ)
∆N (µ) ≤ + τN (µ) kUh (µ) − UN (µ)kX . (3.55)
βN (µ)
(0, 1) ∂n θ = 0 u = 0 (1, 1)
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
u=0 u=0
θ=1 θ=0
(0, 0) ∂n θ = 0 u = 0 (1, 0)
Figure 3.5. Unit square cavity domain Ω, with the different boundaries identified, for problem (3.30).
vertical linear contouring for the temperature and a recirculating motion in the core of the re-
gion. As we increase the value of the Rayleigh number in D, the flow is stretched to the walls,
especially to the vertical walls, and the heat transfer becomes driven mainly by convection. The
isotherms become horizontal in a domain inside the cavity, far from the walls, that increases
as the Rayleigh number increases. When we consider the second scenario, where the Rayleigh
number range is higher, the velocity in the center of the cavity is practically zero and has large
and normal gradients near the vertical walls. The temperature isolines are horizontal in a large
domain inside the cavity, except near the vertical walls. This behavior agrees with the results
presented in several works, e.g., [109, 146, 601].
We consider different meshes depending on the scenario. For µ √ ∈ [103 , 105 ] we consider a
uniform mesh, with 50 divisions on each square side, i.e., h = 0.02 2. For µ ∈ √ [10 , 10 ] we
5 6
consider a finer mesh, with 70 divisions on each square side, i.e., h = 1/70 · 2, in order to
efficiently reproduce the eddies near the vertical walls appearing in this Rayleigh number range.
In Figure 3.6, we show some snapshots for different parameter values.
In the RB framework, we perform an EIM for both the eddy viscosity and the eddy diffusivity.
Although for the numerical analysis performed in this work we have considered a regularized
eddy diffusivity, the numerical tests are done with the eddy diffusivity defined in (3.31). Since
the eddy diffusivity is proportional to the eddy viscosity, we only need to perform one EIM. With
the EIM we are able to decouple the parameter dependence of the nonlinear eddy viscosity and
eddy diffusivity terms. For this test, we need M = 42 bases to reach a prescribed tolerance of
εEIM = 5 · 10−3 , when we consider the case when µ ∈ [103 , 105 ], and M = 150 basis functions
when we consider the second scenario with µ ∈ [105 , 106 ]. In this last case, the Smagorinsky
eddy viscosity and eddy diffusivity terms become more relevant, and for this reason we take a
lower tolerance for this test than for the previous one, considering εEIM = 10−4 . In Figure 3.7
we show the evolution of this error for both scenarios.
For the greedy algorithm we prescribe a tolerance of εRB = 10−4 for both scenarios. For
the first scenario, when µ ∈ [103 , 105 ], we need Nmax = 22 bases to reach this tolerance. When
N = 15, the condition of Theorem 3.10 holds and τN (µ) < 1 for all µ in D. In the second
scenario, when µ ∈ [105 , 106 ], we need N = Nmax = 64 basis functions to reach the tolerance
previously prescribed, making τN (µ) smaller than one when we get N = 52 basis functions. In
both cases, when τN (µ) > 1 and the a posteriori error bound is not defined, we use as a posteriori
error bound the proper τN (µ). In Figure 3.8 we show the convergence for the greedy algorithm,
76 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 3.6. FE solution, velocity magnitude (top) and temperature (bottom), for µ = 4523,
µ = 55732, and µ = 642639 (left to right).
10 3 104
max kνT (µ) − IM [νT (µ)]k∞ max kνT (µ) − IM [νT (µ)]k∞
µ∈D
103
µ∈D
10 2
102
10 1 101
100
10 0
10-1
10
-1
10-2
10-3
10 -2
10-4
10 -3 10-5
0 5 10 15 20 25 30 35 40 45 0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150
M M
Figure 3.7. Error evolution for the EIM for µ ∈ [103 , 105 ] (left) and µ ∈ [105 , 106 ] (right).
and in Figure 3.9 we show the value of the a posteriori error estimator for N = Nmax for both
scenarios.
Finally, in Table 3.2, we compare the FE and RB solutions for several Rayleigh values in both
scenarios. We show the computational time for solving an FE solution and an RB solution in the
online phase. As can be observed, the speedup rate of the computational time is larger than three
orders of magnitude when µ ∈ [103 , 105 ], while when µ ∈ [105 , 106 ] the speedup rate is close to
300. The difference in the speedup magnitude between both cases is due to the greater number of
EIM and RB functions computed in each case. In addition, we show the errors in the H 1 -norm
for velocity and temperature and in the L2 -norm for pressure, for which we observe that the RB
solution is close enough to the FE solution, with the errors always being below of 105 in both
cases. For this test, the offline phase when µ ∈ [103 , 105 ] took approximately two days. For
the case when µ ∈ [105 , 106 ], the offline phase took approximately three weeks. In this offline
3.4. Geometrical Parametrization 77
10 10
10 10
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
5 5
10 10
10 0 100
10 -5 10-5
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 1 4 7 10 13 16 19 22 25 28 31 34 37 40 43 46 49 52 55 58 61 64
N N
Figure 3.8. Evolution of the a posteriori error bound in the greedy algorithm for µ ∈ [103 , 105 ]
(left) and µ ∈ [105 , 106 ] (right).
10 -3 10-3
-4
10
10-4
10 -5
10-5
10 -6
10 -7 10-6
10 -8
10-7
10 -9
10-8
10 -10
10 -11 10-9
10 3 10 4 10 5 105 106
Ra Ra
Figure 3.9. A posteriori error bound for N = Nmax for µ ∈ [103 , 105 ] (left) and µ ∈ [105 , 106 ] (right).
computational time we consider either the EIM or the greedy algorithm with the computation of
the a posteriori error estimator.
Table 3.2. Computational time for FE and RB solutions, with the speedup and the error, for
problem (3.30): Ra ∈ [103 , 105 ] (top) and Ra ∈ [105 , 106 ] (bottom)
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
where the subscripts x and y denote the addend of the corresponding operator relative to the
partial derivative with respect to x or y, respectively.
Here we suppose that we consider a uniform mesh in the reference domain Ωr , with Nh
partitions on each side. Since the mesh size, hK , in the VMS-Smagorinsky eddy viscosity and
3.4. Geometrical Parametrization 79
eddy diffusivity terms is related to the parameter-dependent original domain, we adapt it to the
reference domain by applying the change of variables map T defined in (3.56), obtaining
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
µ2g + 1
hK = ∀K ∈ Th .
Nh2
Note that problem (3.41) is a particular case of problem (3.58), taking µg = 1. The well-
posedness analysis of this problem can be performed in an analogous way as in Section 3.3,
considering again the mollifier for the thermal eddy diffusivity in order to obtain that the deriva-
tive operator is locally Lipschitz.
Figure 3.10. FE snapshots for µg = 0.5 (top left), µg = 1 (bottom left), µg = 1.5 (middle), and
µg = 2 (right).
80 Chapter 3. Certified Smagorinsky Reduced Basis Turbulence Model
10 3 10 2
µg ∈D µ∈D
2
10 1
10
10 1
10 0
10 0
10 -1
10 -1
10 -2
10 -2
10 -3 10 -3
10 -4
10 -4
0 10 20 30 40 50 60 70 0 20 40 60 80 100 120 140
M M
Figure 3.11. Error evolution for the EIM and for the Boussinesq VMS-Smagorinsky model with
µ ∈ [103 , 104 ] × [0.5, 2].
and physical parametrization we need M = 138 basis functions with a prescribed tolerance of
εEIM = 10−3 . In Figure 3.11 we show the evolution of the infinity norm of the error between
the eddy viscosity νT (µg ) and its EIM approximation, for both cases.
For the greedy algorithm, we prescribe a tolerance for the a posteriori error bound of εRB =
10−4 in the case of the geometrical parameter only, while when we consider both parameters,
we prescribe a tolerance of εRB = 10−3 . We need N = 23 basis functions to guarantee the
conditions of Theorem 3.10 and get τN (µ) < 1. Then, we reach the prescribed tolerance when
N = Nmax = 32 for the case of a single geometrical parametrizations and N = Nmax = 46
when we consider both physical and geometrical parametrizations. In Figure 3.12, we show the
maximum value for all µ ∈ D of the a posteriori error bound estimator, as well as τN (µ) in
each iteration of the greedy algorithm for both cases. Moreover, in Figure 3.13, also for both
cases, we show the a posteriori error bound for all µ ∈ D in the last iteration of the greedy
algorithm.
10 10 10 8
10 5 10 4
10 2
10 0 10 0
-2
10
10 -5 10 -4
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 32 1 4 7 10 13 16 19 22 25 28 31 34 37 40 43 46 49 52 54
N N
Figure 3.12. Evolution of the a posteriori error bound in the greedy algorithm for only the
geometrical parameter (left) and for both physical and geometrical parameters (right)
Finally, in Table 3.3, we summarize the results for several parameter values, for only the ge-
ometrical parametrization test. We compare the time for computing a FE solution and the online
phase computational time. We obtain a speedup rate of several hundred in the computational
time. The RB solution accuracy is fairly good, since the error is approximately of order 10−6 for
velocity, 10−8 for temperature, and 10−5 for pressure.
3.4. Geometrical Parametrization 81
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
10 -4
10 -5
10 -6
10 -7
10 -8
0.5 1 1.5 2
Figure 3.13. A posteriori error bound for only the geometrical parameter (left) and for both
physical and geometrical parameters (right), for N = Nmax .
Moreover, in Table 3.4 we summarize some results obtained for some values of the parameter
when we consider both the geometrical and the physical parametrization. There we show that
the error between the FE solution and the RB solution is of order 10−5 for velocity, 10−7 for
temperature, and 10−5 for pressure. For this test, the speedup rate obtained in the computation
of the RB solution in the online phase with respect to the computation of the FE solution is
around 50. Again, we have obtained a good accuracy in the RB solution with respect to the FE
solution, with a decrease of the computational time that is worthy of consideration.
Table 3.3. Computational time for FE and RB solutions, with the speedup and the error, for the
Boussinesq VMS-Smagorinsky model with µg ∈ [0.5, 2].
Table 3.4. Computational time for FE and RB solutions, with the speedup and the error, for the
Boussinesq VMS-Smagorinsky model with µ ∈ [103 , 104 ] × [0.5, 2].
3.5 Conclusion
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In this chapter we presented two different models: the Smagorinsky model and the VMS-
Smagorinsky Boussinesq model. We constructed the RB problem associated with them and
developed numerically the a posteriori error estimator. In the case of the VMS-Smagorinsky
Boussinesq model we defined a mollifier in order to ensure that the derivative operator is locally
Lipschitz. Moreover, for the nonlinear terms, we dealt with the EIM in order to be able to store
the matrices for the RB problem. We also considered a variation of the RB VMS-Smagorinsky
Boussinesq model, introducing in the last section a geometrical parameter. We presented numer-
ical tests for all the models, in which we can observe a high speedup in the computation of the
RB solutions compared with the computational time of the FE solution.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 4
Finite Element–Based
Reduced Basis Method
for Optimal Flow Control
4.1 Introduction
This chapter is a brief introduction to parametrized optimal control problems (OCP(µ)s) gov-
erned by parametrized partial differential equations (PDE(µ)s). They represent a versatile math-
ematical tool, able to bridge the gap between equations and collected data. Indeed, they have
been employed in many applications including fluid dynamics [148, 413, 461, 145], biomedicine
[33, 359, 562, 638], and environmental sciences [473, 475, 557, 562]. Despite the ubiquitous role
they play, OCP(µ)s are complex and challenging not only from a theoretical viewpoint but also
from a numerical one. Indeed, their goal is to change the expected state solution using control
variables, which steer the system toward a desired configuration, representing a more reliable
prediction of a known phenomenon or an observable realization of it.
Specifically, we will deal with a parametrized setting, which will change physical and ge-
ometrical features of the optimization processes by µ varying in a parametric space. Our goal
is to provide an overview of methodologies which can be used to treat the OCP(µ)s in a many-
query and real-time context, where several parametric instances are required in, possibly, a small
amount of time. Indeed, the applicability of optimal control strategies is limited by the computa-
tional costs they carry with them, most of all in the time-dependent or nonlinear framework. For
these kinds of systems, the needed effort for a parametric analysis can be too expensive to tackle
using standard discretization techniques. To overcome this issue, we will present reduced order
method (ROM) approaches [281, 464, 506, 505] adapted to this specific mathematical model
[25, 24, 149, 223, 302, 328, 329, 352, 413, 414, 475], moving from the well-known literature for
steady linear problems to time-dependent and nonlinear applications.
The contribution is outlined as follows. In Section 4.2, we introduce the linear problem struc-
ture in its saddle-point formulation and we show how to tackle it in a ROM fashion. Furthermore,
a first environmental application is shown: an optimal control on a pollutant release in the Gulf
of Trieste. Section 4.3 deals with the extension of the proposed structure and related ROM tech-
niques to nonlinear state equations, with a second environmental application in oceanographic
weather forecasting. Then, we move toward time dependency in Section 4.4, where the saddle-
point structure in the space-time formulation [235, 590, 631] is briefly described and space-time
83
84 Chapter 4. Finite Element–Based Reduced Basis Method for Optimal Flow Control
ROM techniques are presented [558]. Their performance has been tested with a boundary control
for a Graetz flow. Finally, some conclusions and perspectives follow in Section 4.5.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
y yd
1 α
J (y, u) = m(y − yd , y − yd ) + n(u, u). (4.4)
2 2
To solve the optimization problem, we define an auxiliary variable z ∈ Y , the adjoint variable,2
and the Lagrangian functional
.
L(y, u, z; µ) = J (y, u) + hF (y, u; µ) , zi . (4.5)
Namely, in this way, the constrained optimization problem at hand is equivalent (see, e.g., [290])
to the unconstrained minimization of (4.5), which can be solved thanks to the following system
of equations: given µ ∈ D, an observation yd ∈ Yd , and f (µ) ∈ Y ∗ , find the triplet (y, u, z) ∈
Y × U × Y which verifies
(adjoint equation),
Ly (y, u, z; µ)[ζ] = 0 ∀ζ ∈ Y
Lu (y, u, z; µ)[τ ] = 0 ∀τ ∈ U (optimality equation), (4.6)
Lz (y, u, z; µ)[ξ] = 0 ∀ξ ∈ Y (state equation),
where the subscripts y, u, and z indicate the Fréchet differentiation with respect to the state,
control, and adjoint variables, respectively.
We now want to recast the system (4.6) in a saddle-point formulation. From this perspective,
.
we define the state-control variable x = (y, u) ∈ X = Y ×U , the bilinear forms A : X×X → R
and B : X × Y → R, and the functionals H ∈ Y ∗ as
.
A(x, ω; µ) = m(y, ζ) + αn(u, τ ) ∀ω = (ζ, τ ) ∈ X,
B(x, ξ; µ) = a(y, ξ; µ) − c(u, ξ) ∀ξ ∈ Y, (4.7)
hH, ωi = m(yd , ζ) ∀ζ ∈ Y.
It is a matter of computation to show that, thanks to definitions (4.7), the optimality system
(4.6) is equivalent to the following mixed problem: given µ ∈ D, an observation yd ∈ Yd , and
f (µ) ∈ Y ∗ , find the pair (x, z) ∈ X × Y which verifies
A(x, ω; µ) + B(ω, z; µ) = hH, ωi ∀ω ∈ X,
(4.8)
B(x, ξ; µ) = hf (µ), ξi ∀ξ ∈ Y.
The well-posedness of the system (4.8) relies on the following theorem.
Theorem 4.1 (Brezzi theorem). The parametrized problem (4.8) admits a unique solution
(x, z) ∈ X × Y if the following conditions hold:
• A(·, ·; µ) is continuous and weakly coercive on the set X0 which denotes the kernel of
B(·, ·; µ); i.e., for a positive constant βA ,
A(x, ω; µ) A(x, ω; µ)
inf sup ≥ βA > 0 and inf sup ≥ 0.
x∈X0 ω∈X0 kxkX kωkX ω∈X0 x∈X0 kxkX kωkX
• B(·, ·; µ) is continuous and infsup-stable; i.e., there exists a positive constant βB such that
B(x, ξ; µ)
inf sup ≥ βB > 0. (4.9)
ξ∈Y x∈X kxkX kξkY
2 The adjoint variable is chosen in the state space Y in order to guarantee the well-posedness of the problem in the
The choice of z ∈ Y makes Theorem 4.1 provable and thus gives us the chance to study the
discretized solution of saddle-point type OCP(µ)s. In the next subsection, we will briefly discuss
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the finite element (FE) technique and an ad hoc reduction algorithm for OCP(µ)s that we employ
in the numerical results presented in Section 4.2.2.
where Pr is the space of polynomials of degree at most r. In this discretized framework, defining
. N
X Nx = Y Ny × U Nu , the FE OCP(µ) reads as follows: given µ ∈ D, an observation yd ∈ Yd y ,
and f (µ) ∈ Y , find the pair (x , z ) ∈ X × Y
Ny Nx Ny Nx Ny
which verifies
System (4.11) is well posed, once again, thanks to Theorem 4.1. However, exploiting the FE
.
approximation of global dimension N = 2Ny + Nu might lead to high computational costs,
which can be unacceptable since the optimization process is usually performed for many different
parameter values (many-query context), and, eventually, for a given new value of µ ∈ D, we
want to compute the solution quickly (real-time context). Supposing that the system at hand is
affine decomposed [281], i.e., that all the bilinear forms considered in (4.11) can be written as a
summation of µ-dependent functions times µ-dependent forms and bilinear forms, we can build
an efficient ROM model based on two stages:
1. The first is an expensive offline procedure which is performed only once. It deals with all
the µ-independent quantities. Here Nmax snapshots, i.e., FE solutions at properly chosen
parameters, are computed and manipulated in order to build the reduced model;
2. The second is a cheap online phase performed for every new µ in a low-dimensional frame-
work dependent on a dimension value N N .
Let us suppose we have built some low-dimensional spaces YyN , U N , and YzN for the state,
.
control, and adjoint variables, respectively. Moreover, let us define, consequently, X N = YyN ×
U N . Namely, in order to solve the optimality system more quickly, one performs a Galerkin
projection in this reduced framework. Thus, the reduced problem reads as follows: given µ ∈ D,
an observation yd ∈ YdN , and f (µ) ∈ YyN , find the pair (xN , z N ) ∈ X N × YzN which verifies
Also in this case, the well-posedness of system (4.12) relies on the Brezzi Theorem 4.1. The con-
tinuity of the involved bilinear forms and the weak coercivity of A(·, ·; µ) are directly inherited
by the high-fidelity discretization. However, this is not the case for the reduced inf-sup stability
condition
B(x, ξ; µ)
inf sup ≥ βBN > 0. (4.13)
ξ∈YzN x∈X N kxkX kξkY
4.2. Linear OCP(µ)s 87
In order to satisfy condition (4.13), we will exploit the aggregated space technique [149, 328,
329, 413, 414]. This strategy is based on postprocessing the basis functions provided by the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
which verifies
A(xN , ω; µ) + B(ω, z N ; µ) = hH, ωi ∀ω ∈ X N ,
(4.14)
B(xN , ξ; µ) = hf (µ), ξi ∀ξ ∈ Y N ,
which is well posed since the relative reduced inf-sup condition
B(x, ξ; µ)
inf sup ≥ βBN > 0 (4.15)
ξ∈Y N x∈X N kxkX kξkY
is provable. Despite the higher dimensionality of the reduced system, enlarging the state and
adjoint spaces assures the well-posedness of the problem in the reduced setting. In the next
subsection, we propose a first linear application in environmental sciences: control of a pollutant
release in a marine ecosystem in the reduced setting.
1 α
Z Z
min (y − yd )2 dΩd + u2 dΩu (4.16)
Y ×U 2 Ω 2 Ωu
obs
Figure 4.2. Left: zoom of the mesh overlapping satellite images (Trieste harbor). Right: subdo-
mains and boundaries. Orange: observation domain Ωobs . Green: control domain Ωu .
Nmax = 100 uniformly distributed snapshots for a fixed α = 10−5 . The global dimension of the
FE system is 5939, where linear polynomials have been exploited for all the involved variables.
In order to build the reduced space, we used a value of Ny = 20 bases for state and adjoint
variables and Nu = 1 basis for the control variable, since U = R, leading to a global reduced
space of dimension N = 4Ny + Nu = 81. The number of bases ensures a good recovery of the
parametrized solution: this can be seen both from the plots of some representative solutions in
Figure 4.3 and from the average relative log-error of the variables with respect to the value of N ,
over 100 uniformly distributed parameters. Indeed, the FE and ROM simulations coincide and
the average log-error shows that we reach a very good approximation even with a small number
of basis functions. Furthermore, defining the speedup index as the number of reduced problems
one can perform in the time of an FE simulation, we reach values of around 250 for a reduced
space of dimension N = 81.
Figure 4.3. Left and center: optimal FE and ROM state pollutant concentration for µ =
(1., −1., 1.), representing the Bora wind action. Right: averaged relative log-error for the variables.
i.e., the action of the state equation can be divided into linear and nonlinear contributions, repre-
sented by a` (·, ·; µ) and an` (·, ·; µ), respectively. In order to achieve the goal yd ∈ Yd , we rely
once again on the Lagrangian formulation to solve the minimization problem, i.e., one can build
the Lagrangian functional (4.5) and solve the nonlinear optimality system (4.6). At the contin-
uous level, the saddle-point structure we presented in the linear case is not preserved and the
4.3. ROMs for Nonlinear OCP(µ)s 89
Brezzi theorem cannot be exploited in this general case. For the sake of brevity, we will not fo-
cus on the well-posedness of the problem; however, the interested reader may refer to [290, 452].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
From now on, we will always assume that the nonlinear OCP(µ) at hand is well posed.
The main goal is to recover the saddle-point structure at the discrete level.3 Indeed, once the
FE spaces Y Ny and U Nu are considered, as in Section 4.2.1, the problem to be solved is as fol-
lows: given µ ∈ D, an observation yd ∈ Yd , and f (µ) ∈ Y ∗ , find the triplet (y Ny , uNu , z Ny ) ∈
Y Ny × U Nu × Y Ny which verifies
Ly (y Ny , uNu , z Ny ; µ)[ζ] = 0 ∀ζ ∈ Y Ny ,
Algebraically, one has to deal with (4.19) through Newton’s method. Namely, we itera-
tively solve a linearized version of (4.19) until a convergence criterion is reached. It is a matter
of computation to show that the structure of the linearized system has a saddle-point nature
[289, 452, 560]. This concept is fundamental to justify the employment of the aggregated space
technique for this specific case, since the peculiar saddle-point structure is also preserved at a re-
duced level and, consequently, the technique for linear problems adapts to nonlinear ones. Thus,
applying the partitioned POD algorithm and space aggregation as in Subsection 4.2.1, the re-
duced problem reads as follows: given µ ∈ D, an observation yd ∈ YdN , and f (µ) ∈ Y N , find
the triplet (y N , uN , z N ) ∈ Y N × U N × Y N which verifies
Ly (y N , uN , z N ; µ)[ζ] = 0 ∀ζ ∈ Y N ,
Lu (y N , uN , z N ; µ)[τ ] = 0 ∀τ ∈ U N , (4.20)
Lz (y N , uN , z N ; µ)[ξ] = 0 ∀ξ ∈ Y N .
The next subsection describes an application in environmental sciences, where the OCP(µ)
is meant as a tool to understand and forecast atmospheric oceanographic phenomena.
• On the one hand, the general ocean circulation model describing large-scale flow dynamics
is a very complicated coupled system of ocean and atmosphere actions [103, Chapter 3].
Despite its completeness as an equation system, it can give results which are not compara-
ble with collected data.
• On the other, oceanography suffers from a lack of data. The data is scattered, difficult to
interpret, noisy, and, furthermore, complicated to collect.
Our aim is to add data to the model using optimal control, exploiting the following formulation:
given µ ∈ D = [0.073 , 1] × [10−4 , 1] × [10−4 , 0.0452 ], and yd ∈ L2 (Ω), find (ψ, q) ∈ Y =
H01 (Ω) × H01 (Ω) and u ∈ U = L2 (Ω) which solve
1 α
Z Z
2
min (ψ − ψd ) dΩ + u2 dΩ (4.21)
Y ×U 2 Ω 2 Ω
3 This is why we also restrict ourselves to the case z ∈ Y in the nonlinear scenario.
90 Chapter 4. Finite Element–Based Reduced Basis Method for Optimal Flow Control
in Ω,
q = ∆ψ
∂ψ
= u − µ1 q + µ2 ∆q − µ3 Q(ψ, q) in Ω,
∂x (4.22)
ψ = 0 on ∂Ω,
q=0 on ∂Ω,
∂ψ ∂q ∂ψ ∂q
Q(ψ, q) = − . (4.23)
∂x ∂y ∂y ∂x
Using such a data assimilation model will give an optimal solution close to the physical data
and could prevent having to collect it, a process which might be expensive, in terms of time and
money. The parametric setting is essential since the dynamic is influenced by several seasonal
phenomena and the physical parameters can help us understand and analyze them. While µ1 and
µ2 represent a diffusive action, the value of µ3 affects the local nonlinear dynamics. The model
describes both the North Atlantic Ocean large-scale currents and the location and intensity varia-
tions of the gyres which are typical of this geographic area, depicted in Figure 4.4, which shows
the Florida peninsula, where the motion of the Gulf Stream starts, eventually warming the coasts
of northern Europe. In the weather forecasting field, it is very important to run many simulations
for several parametric values in order to better study and predict climatological phenomena. In
particular, for this experiment, the desired state ψd is an FE simulation obtained by the uncon-
trolled governing equation with a right-hand side given by f = − sin(πy), with µ1 = 10−4 ,
µ2 = 0.073 , and µ3 = 0.072 , and represents the seasonal Gulf Stream. Figure 4.5 shows the
desired solution profile together with the high-fidelity and reduced solutions. They all match.
Furthermore, the same figure also represents the averaged log-error norm between FE and re-
duced variables over 100 randomly distributed parametric evaluations. It indicates the ability of
the ROM solutions to recover the field with an aggregated reduced system of global dimension
N = 180. About time performances, despite solving a much lower dimensional system than the
original one, we do not reach great speedup; it is around two. This issue is related to the nonlinear
nature of the system. Indeed, the nonlinear form still depends on the FE dimension N = 6490,
obtained by using P1 elements for all the variables, and the online phase involves the assembly
(and projection) of the high-fidelity solutions. Nevertheless, one can rely on such hyperreduc-
tion techniques as the empirical interpolation method (EIM) to recover an affine decomposed
structure; see, e.g., [40].
Figure 4.4. Left: zoom of the mesh overlapping satellite images (Florida peninsula). Right:
triangulation of the spatial domain Ω, representing the North Atlantic Ocean, from the Florida peninsula
to northern Europe.
4.4. ROMs for Space-Time OCP(µ)s 91
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 4.5. Top: comparison between desired state, FE and ROM solutions for µ =
(10−4 , 0.073, 0.0452), representing the Gulf Stream. Bottom: averaged relative log-error for all the vari-
ables.
where s : Y × Y → R is a bilinear form describing the time evolution of the state equation (i.e.,
it depends on the time derivative yt ) in the time interval [0, T ] and
.
Y = {y ∈ L2 (0, T ; Y ) such that yt ∈ L2 (0, T ; Y ∗ )}.
The action of s(·, ·; µ) is
s(y, ζ; µ) = hS(µ)yt , ζiY ∗ Y . (4.25)
For a more detailed explanation, see, e.g., [559]. In the time-dependent setting, we have not only
spatial boundary condition but also an initial time evolution condition: in this contribution we
will restrict ourselves to the case of a state variable that verifies the spatial boundary conditions
and y(t = 0) = 0 in Ω. Here, guided by the numerical test we present in the following subsec-
tion, we deal with geometrical parametrization, and the forms involved in (4.24) are the possible
trace-backs in a reference domain that we call, with abuse of notation, Ω. The trace-back tech-
nique is used to deal with geometrical parameters in a ROM setting in [505]. We specify that
S(µ), in the nongeometrical case, is the identity map. Now, we can define the objective func-
tional and, in this specific setting of time-dependent problems, it is of the form
1 T α T
Z Z
J (y, u) = m(y − yd , y − yd )dt + n(u, u)dt, (4.26)
2 0 2 0
92 Chapter 4. Finite Element–Based Reduced Basis Method for Optimal Flow Control
.
with yd taken in Yd = L2 (0, T ; Yd ). Also in this case, we can exploit the Lagrangian formalism,
defining the adjoint variable z ∈ Y; solve the problem by building a functional of the form
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
.
L(y, u, z; µ) = J (y, u) + hFt (y, u; µ) , zi ; (4.27)
and, differentiating with respect to the three different variables, obtain a time-dependent optimal-
ity system. The main feature of this system is that, provided we have a suitable discrete approx-
imation, the saddle-point structure is preserved even in the time-dependent setting. To this end,
we introduce the space-time approximation. Extending what was already proposed for parabolic
problems [235, 590, 631] to the optimal control context, we define the finite-dimensional function
spaces YNy and UNu . The global high-fidelity dimension is N = 2Ny +Nu , with Ny = NΩy ·Nt
and Nu = NΩu · Nt , where NΩy , NΩu , and Nt are the state-adjoint spatial, the control spatial, and
the time discretization dimensions, respectively. In the test case we will present, we used the
FE-Euler approach. Namely, the time interval [0, T ] is divided into Nt equispaced subintervals
of length ∆t. The generic discrete time instance is tk = k∆t for k = 0, . . . , Nt . Thus, the
Ny Nu Ny
discrete variables at time tk are yk Ω , uk Ω , and zk Ω . Now, given an FE basis for the spatial state
Ny Nu
space Y and the spatial control space U , say {φi }i=1
Ω
and {ϕi }i=1
Ω
, respectively, we can expand
the variables at tk as follows:
y u y
NΩ NΩ NΩ
Ny Nu Ny
X X X
yk Ω = yki φi , uk Ω = uik ϕi , and zk Ω = zki φi . (4.28)
1 1 1
In other words, each time instance is an FE variable, which drastically increases the dimension
of the algebraic problem at hand. We underline that the adjoint version of a parabolic equation
is a backward parabolic one. Indeed, the adjoint of s(y, ζ; µ) is actually −s(z, ξ; µ), exploiting
the final time condition z(t = T ) = 0. For the parabolic state equation, we used a backward
Euler method, while the adjoint equation is discretized backward in time using the forward Euler
method. In [558], the authors show that, given a common discretization technique for state and
adjoint, the whole space-time optimality system reads as follows: given µ ∈ D, an observation
N
yd ∈ Yd y , and f (µ) ∈ YNy , find the pair (xNx , z Ny ) ∈ XNx × YNy which verifies
Z T Z T Z T
A(xNx , ω; µ)dt + B(ω, z Ny ; µ)dt = hH, ωi dt ∀ω ∈ XNx ,
Z0 T Z 0T 0 (4.29)
B(xNx , ξ; µ)dt = hf (µ), ξidt ∀ξ ∈ YNy ,
0 0
.
where XNx = YNy × UNu . Here, the bilinear forms A : X × X → R and B : X × Y → R and
the functionals H ∈ Y ∗ are defined as
.
A(x, ξ; µ) = m(y, ω; µ) + αn(u, κ; µ) ∀ξ = (ω, κ) ∈ X,
B(x, ζ; µ) = s(y, ζ) + a(y, ζ; µ) − c(u, ζ; µ) ∀ζ ∈ Y, (4.30)
H(ξ; µ) = m(yd , ω; µ) ∀ω ∈ Y,
where the involved bilinear forms were defined in Section 4.2. However, here, as already spec-
ified, they can represent their possible trace-back. This framework justifies using a partitioned
space-time POD: namely, we separately applied the POD algorithm [85, 114, 281] to the three
variables containing the solutions at all the temporal steps. After this first compression step, the
reduced spaces were built using the aggregation technique, due to the saddle-point nature of the
system at hand.
4.4. ROMs for Space-Time OCP(µ)s 93
and
.
Ωd (µ3 ) = [1, 1 + µ3 ] × [0, 0.2] ∪ [1, 1 + µ3 ] × [0.8, 1].
The physical domain Ω is depicted in Figure 4.6. Besides the geometrical parameter, we will
deal with µ1 and µ2 , which represent the diffusivity coefficient of the system and the desired
profile we want to reach in Ωd (µ3 ), respectively. In this specific case, we define y ∈ Y, where
Y = HΓ1D (Ω(µ3 )), and u ∈ U, where U = L2 (ΓC (µ3 )), where ΓD = ({0} × [0, 1]) ∪ ([0, 1] ×
{0}) ∪ ([0, 1] × {1}) is the portion of the domain where Dirichlet boundary conditions have been
applied.
ΓD Ω1 Ω2 (µ3 ) ΓN (µ3 )
Ω4 (µ3 )
(0, 0) (1, 0) (1 + µ3 , 0)
Figure 4.6. Domain Ω. Observation domain: Ωd (µ3 ) = Ω3 (µ3 ) ∪ Ω4 (µ3 ). Control domain:
ΓC (µ3 ) (red solid line). Blue dashed line: Dirichlet boundary conditions. The reference domain Ω is given
by µ3 = 1.
The problem we consider reads as follows: given µ ∈ D, find the state-control variable
(y, u) ∈ Y × U which solves
1 T α T
Z Z Z Z
2
min (y − yd (µ2 )) dxdt + u2 dxdt (4.31)
Y×U 2 0 Ω3 (µ) 2 0 ΓC
Figure 4.7. Left: FE state solutions for t = 1 s, 2 s, 3 s for µ = (12.0, 2.5, 2.0). Right: ROM
state solutions for t = 1 s, 2 s, 3 s for µ = (12.0, 2.5, 2.0).
where x1 and x2 denote the spatial components, y0 is the null function in the domain which
respects the boundary conditions, yd (µ2 ) ≡ µ2 , and T = 5. With ΓN (µ3 ) = {1 + µ3 } × [0, 1],
we indicate the portion of the domain where Neumann conditions apply. First of all, in order
to solve this OCP(µ), we traced back the problem in the reference domain corresponding to
µ3 = 1, following the pullback strategy presented in [505]. We used ∆t = 1/6 over the time
interval, employing Nt = 30 time steps in an Euler-based scheme. Furthermore, we exploited
P1 elements for all the variables involved, resulting in a global space-time dimension of N =
313830. The reduced function spaces have been built by compressing the information of Nmax =
70 snapshots, retaining N = 35 bases for each variable. After the aggregation technique, we
obtained a global reduced system of dimension 5N = 175. The reliability of the approach is
shown in Figure 4.7, which depicts representative FE solutions and ROM solutions at different
times. The space-time POD is able to reproduce the fields, no matter the temporal instance
considered. Furthermore, we ensure the accuracy of the strategy used with the log-error plot
in Figure 4.8. The considered error is averaged over a testing set of 50 uniformly distributed
parameters: the minimum value reached is around 10−4 for the state and the control variable,
while the adjoint relative log-error is around 10−3 . In this case, we have a remarkable speedup
of order O(105 ), indicating the convenience of using ROMs in a time-dependent optimization
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
setting. The saved computational time can be used to solve many parametric instances and to
completely and deeply analyze the system.
4.5 Conclusions
This chapter summarized ROM approaches for OCP(µ)s and presented them as a reliable and fast
tool to deal with several state equations, including steady linear, nonlinear, and time-dependent
problems. The problems were described in their continuous and discrete FE formulation (steady
or space-time), and the ROM strategies were validated with several test cases in fluid dynamics
and environmental sciences. We presented the following examples:
5.1 Introduction
This chapter is concerned with analyzing and developing efficient reduced order models (ROMs)
to investigate complex bifurcating phenomena represented by nonlinear parametrized partial dif-
ferential equations (PDEs) in many physical contexts, including continuum and quantum me-
chanics and fluid dynamics [446].
Indeed, reconstructing the bifurcation diagrams, which highlight the singularities of the equa-
tions and the possible coexisting states, requires a huge computational effort, especially in the
multiparameter context. To overcome this issue, we developed a reduced order branchwise al-
gorithm for the efficient investigation of such complex behavior, with a focus on the stability
properties of the solutions. We applied our approach to the Von Kámán equations for buckling
plates [451], the Gross–Pitaevskii equations in Bose–Einstein condensates [450], the hyperelastic
models for bending beams [449], and the Navier–Stokes model for flow in a channel [452, 447].
Despite the huge discrepancies among the aforementioned models, we considered a unique
abstract framework in which to analyze all the different complex physical behaviors. The pres-
ence of bifurcating phenomena usually requires fine mesh discretization of the domain and, to-
gether with the parameter dependence, leads to a many-query context that could be a bottleneck
for a deep study of the problem under investigation.
In fact, when dealing with bifurcations, one needs to compute the numerical approximation
of the solution for many instances of the parameter in order to discover the critical points of the
model and its postbifurcating behavior. Moreover, the analysis becomes even more complicated
if the goal is a complete reconstruction of all the possible solutions that the model could admit.
Finally, a key aspect in these contexts is the stability analysis of the discovered solutions, which
can be studied by means of specific eigenvalue problems.
The numerical approximation of such models requires a combination of different method-
ologies. To obtain a high-fidelity approximation of the problem, we considered as a full-order
model (FOM) the finite element method (FEM), which we combined with a continuation method,
to follow the solution by varying the parameter, and with the Newton method, which is used to
linearize the equation.
97
98 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
We tackled the huge computational cost of this investigation using ROMs. In particular, we
developed a branchwise algorithm based on the proper orthogonal decomposition (POD) tech-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
nique, which aims to efficiently reconstruct the bifurcating behavior by projecting the governing
equations into a low-dimensional manifold spanned by global basis functions (with respect to the
parameter space).
Moreover, several issues and questions arose when dealing with the approximation and re-
duction of bifurcating phenomena. In particular, we developed a reduced order approach for
the deflated continuation method to efficiently discover new solution branches [454]. We pro-
posed and discussed different optimal control problems (OCPs) to steer the bifurcating behavior
toward desired states [452]. We investigated the evolution of bifurcations in the multiphysics
context by means of fluid-structure interaction (FSI) problems [337]. Finally, we used a neural
network approach based on the proper orthogonal decomposition (POD-NN) as an alternative to
the empirical interpolation method (EIM) to develop a reduced manifold–based algorithm for the
efficient detection of the bifurcation points [447, 448]. For the sake of exposition, we will not
treat these latest advances here, and we redirect the reader to the references reported above.
and the weak formulation of equation (5.1) reads as follows: given µ ∈ P, find X(µ) ∈ X such
that
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
g(X(µ), Y ; µ) = 0 ∀ Y ∈ X. (5.3)
In order to completely describe the problem formulation in (5.1), we need some suitable
boundary conditions (BCs) on the domain Ω. Here, we will cover the cases of both homo-
geneous and inhomogeneous Dirichlet and Neumann conditions, depending on the test case at
hand. Moreover, these BCs will automatically be embedded in (5.3) by an appropriate choice of
the space in which to seek the solution.
Having introduced the mathematical formulation of the problem, a starting point for the
analysis is to investigate its well-posedness. A PDE is said to be a well-posed problem if it
is characterized by the existence and uniqueness of the solution. Indeed, these properties are
also fundamental for the numerical approximation of the model. Although these assumptions are
often required, dealing with more realistic models can complicate the setting. In fact, one has to
find a trade-off between the complexity of the model and its effective tractability.
As we will see later on, complex nonlinear and parametric PDEs are proved to be more
accurate in describing the physics, but the lack of good mathematical properties makes them
difficult to approach, both theoretically and numerically.
Hence, in the following, we investigate when such properties are conserved and thus when
the problem admits a unique solution. On the contrary, in the next subsection, we will discuss
the presence of critical points in which the well-posedness is lost, understanding how the model
behaves and evolves in such cases.
Let us assume that the map G in equation (5.1) is continuously differentiable with respect to
X and µ. Let (X; µ) ∈ X × P be a known solution, i.e., G(X; µ) = 0. Then, we denote by
DX G(X; µ) : X → X0 and Dµ G(X; µ) : P → X0 the partial derivatives of G on a generic point
(X, µ) ∈ X × P. When dealing with the approximation of (5.1), the strong assumption usually
found in literature, which ensures the well-posedness of the problem, is that DX G(X; µ) : X →
X0 is bijective. Indeed, the following result holds.
Theorem 5.1. Let Br (X), Br (µ) be two balls of radius r and r around X and µ, respectively.
Let G : X × P → X0 be a C 1 map, and assume that
(1) G(X; µ) = 0 for the solution (X, µ) ∈ X × P, and
(2) DX G(X; µ) : X → X0 is bijective.
Then there exist r, r > 0 and a unique solution X(µ) ∈ Br (X) ∩ X such that
G(X(µ); µ) = 0 ∀µ ∈ Br (µ) ∩ P.
While this is a straightforward application of the implicit function theorem [129, 642], it
ensures the well-posedness of the problem and thus the existence of a local branch of nonsingular
solutions. Furthermore, we can rewrite the Fréchet partial derivatives of G on (Z, µ) ∈ X × P
with respect to X as
dg[Z](X, Y ; µ) = hDX G(Z; µ)X, Y i ∀X, Y ∈ X, (5.4)
where we have introduced the parametrized variational form dg[Z](·, ·; µ). We now observe that
assumption (2) in Theorem 5.1 can be reformulated in terms of the variational form dg[Z](·, ·; µ)
itself, which is said to be
(2a) continuous on X × X if there exists a continuity constant γ > 0 such that
dg[Z](X, Y ; µ)
γ(µ) = sup sup >γ ∀µ ∈ P, (5.5)
X∈X Y ∈X kXkX kY kX
100 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
(2b) inf-sup stable on X × X if there exists an inf-sup constant β > 0 such that
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
dg[Z](X, Y ; µ)
β(µ) = inf sup ≥β ∀µ ∈ P (5.6)
X∈X Y ∈X kXkX kY kX
and
dg[Z](X, Y ; µ)
inf sup >0 ∀µ ∈ P. (5.7)
Y ∈X X∈X kXkX kY kX
We remark that conditions (5.6) and (5.7) are equivalent to the injectivity and the surjec-
tivity of the Fréchet derivative DX G(Z; µ), respectively. In particular, condition (5.6) can be
reformulated as
∃β(µ) > 0 : kDX G(Z; µ)XkX0 ≥ β(µ)kXkX , (5.8)
i.e., the Fréchet derivative of G is bounded below. It is clear that (5.8) holds if and only if
DX G(Z; µ) is injective. On the other hand, condition (5.7) can be easily restated as the injec-
tivity of the adjoint operator of DX G(Z; µ), which can be expressed in terms of its variational
form dg as
Y ∈ X, dg[Z](X, Y ; µ) = 0 ∀X ∈ X ⇒ Y = 0. (5.9)
Therefore, we observe that condition (5.7) is actually stating that DX G(Z; µ) is surjective, since
the injectivity of the adjoint of an operator is equivalent to the surjectivity of the operator itself.
Moreover, when dealing with linear parametrized PDEs, it is clear that (e.g., for an unforced
problem)
dg[Z](X, Y ; µ) = g(X, Y ; µ) ∀X, Y, Z ∈ X, ∀µ ∈ P. (5.10)
Hence, the general assumptions in (5.5), (5.6), and (5.7) fall back on the standard ones required
by the well-known theorems of Lax–Milgram and Nečas [186, 520].
In conclusion, the bijectivity assumption in Theorem 5.1, and equivalently the injectivity and
surjectivity conditions (5.8) and (5.9), are the key ingredients of a well-posed problem. In the
following, we will see what happens when these hypotheses are no longer valid at some point in
the parameter space.
the magnitude of the load and on the specific material properties of the beam. Hence, we expect
that a small perturbation of µ leads to a configuration with the same qualitative features.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
At the same time, we empirically know that there exists a critical value for the load, which
we call µ∗ , after which the model does not preserve the original qualitative behavior. Practically
speaking, if the load is big enough, the beam cannot sustain the compression and it buckles. This
is a clear example of a qualitative change.
We call bifurcation phenomena the qualitative changes that are not stable under small per-
turbations of a parameter µ, and bifurcation points the points µ∗ at which such changes occur
[535, 342, 355]. Usually, we can distinguish between different qualitative states by means of
their geometrical shape or pattern configuration. Indeed, two of the most relevant features that
may change in the presence of a bifurcation phenomenon are the stability and the symmetry of
the resulting states.
The aim is to investigate the solutions of a specific model by varying a one-dimensional
parameter, i.e., the bifurcation parameter, which is responsible for the bifurcating phenomena.
Correlated with the qualitative change of the solutions is the nonuniqueness behavior; in fact,
as we will see, in such situations the model admits different solutions for the same value of the
parameter. We will refer to the set of solutions with the same qualitative properties as a branch.
As we said before, when the parameter changes slightly, we expect a stable solution to evolve
continuously in a unique manner.
The situation changes drastically if the inf-sup stability of the model, ensured by (5.6) and
(5.7), is lost. This happens when the bifurcation parameter reaches a critical value µ∗ for which
the system admits the existence of a qualitatively different solution that bifurcates from the pre-
vious stable one.
Following [17, 16], we consider the problem (5.1) and assume that X = 0 is a solution for
every parameter µ ∈ P, namely
G(0; µ) = 0 ∀µ ∈ P,
which will be referred to as the trivial solution. Then, we can define the set of nontrivial solutions
of (5.1) as
S = {(X, µ) ∈ X × P | X 6= 0, G(X; µ) = 0} . (5.11)
For the sake of simplicity, we chose to present here the bifurcations that originate from the
trivial solution. Indeed, many times, and in particular among the test cases we treated, we have
a nontrivial ground state solution, which loses its uniqueness by branching to another nontrivial
state.
It is remarkable to note that this situation can happen in a nested fashion. Increasing in
complexity, one can find a bifurcation emerging from an already bifurcated configuration. This
phenomenon is usually called secondary bifurcation; an example will be shown in Subsection
5.4.3.
As we understood from the example reported at the beginning, there could exist some values
for the parameter µ such that one or more nontrivial solutions branch off from the trivial one.
These are the critical points that we have introduced before, and in the context of bifurcation
theory we define them as in [17, 16].
Definition 5.2. A parameter value µ∗ ∈ P is a bifurcation point for (5.1) if there exists a
sequence (Xn , µn ) ∈ X × P, with Xn 6= 0, such that
◦ G(Xn ; µn ) = 0,
◦ (Xn , µn ) → (0, µ∗ ).
102 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
Bifurcation is thus a paradigm for nonuniqueness in nonlinear analysis, and a necessary con-
dition is the failure of the implicit function theorem.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In fact, we highlight that the inf-sup stability property, which should guarantee the well-
posedness of the PDE, is actually no longer valid for bifurcating phenomena. Indeed, at the bi-
furcation point µ∗ the inf-sup constant β becomes zero, and the Fréchet derivative DX G(X; µ)
fails to be invertible. Therefore, unless otherwise specified, we will talk about well-posed prob-
lems as long as µ 6= µ∗ .
The aim of bifurcation theory is thus to provide a mathematical description of the bifurcating
scenario that can be observed in physical systems and experiments. The understanding of the
global behavior, from both the theoretical and the numerical viewpoint, is usually supported by
reconstructing some scalar characteristic value, which we will denote by s(X(µ)).
Given (X(µ), µ), a solution of (5.1), for each instance of the parameter, one can draw a plot
of s(X(µ)) versus µ, which in the literature is known as a bifurcation diagram. Therefore, the
existence of different solutions for the same values of the parameter will result in the presence
of multiple branches on the diagram. Different choices of s(·) result in possibly different plots;
thus, to avoid confusion, the function has to be chosen carefully and has to represent the main
features while changing the qualitative behavior of the resulting configuration.
It is clear that there exist many different bifurcation phenomena, and usually complex non-
linear problems are characterized by one (or more) of those. In fact, the literature is full of
examples of physical systems with qualitatively changing behavior. Each one of these bifurcat-
ing phenomena has a specific peculiarity, but in the field of bifurcation theory they are usually
categorized by some common features. Among others, the most known ones are turning point
(or fold) bifurcations, transcritical bifurcations, pitchfork bifurcations, and Hopf bifurcations.
While the latter are usually of interest when dealing with time-dependent problems, here we will
focus on models that exhibit pitchfork bifurcations.
We remark that when µ ∈ / P, the case k = 1 can be seen as the well-posed problem, and X is
often referred to as the solution manifold [281].
Since we are interested in recovering the full bifurcation diagram for the model under consid-
eration, the numerical approximation deals with the computation of the discretized counterpart
5.3. Numerical Approximation of Bifurcating Phenomena 103
of (5.13). To this end, we pursue a branchwise approach. In other words, we aim to reconstruct
one fixed branch M ⊂ X at time, corresponding to some i ∈ {1, . . . , k}. The global behavior is
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
thus obtained by approximating its solutions for all the values in P and then moving to the next
solution branch in X .
We will refer to these steps as the high-fidelity approximation, since we are dealing with high-
dimensional and time-consuming procedures. In order to avoid this computational complexity, in
the second part, we will present the ROMs, or low-fidelity approximations, that we have applied
to reduce the dimensionality of the system, while allowing for an efficient approximation of the
bifurcating phenomena using a reduced version of the branchwise algorithm.
(j)
and denote the solution vector by XN (µ) = {XN (µ)}N
j=1 . We then go back to the study of the
solution XN (µ) ∈ RN of the nonlinear system
GN (XN (µ); µ) = 0 in RN , (5.16)
where the high-fidelity residual vector GN is defined as
(GN (XN (µ); µ))i = g(XN (µ), E i ; µ) ∀ i = 1, . . . , N .
The nonlinear solver chosen to linearize the weak formulation in (5.14) is the well-known
Newton–Kantorovich method [129, 470], which reads as follows: given µ ∈ P and choosing
an initial guess XN0
(µ) ∈ XN , for every k = 0, 1, . . . , we seek the variation δXN ∈ XN such
that
k k
dg[XN (µ)](δXN , YN ; µ) = g(XN (µ), YN ; µ) ∀ YN ∈ XN . (5.17)
Then we update the solution at the iteration k + 1 as
k+1 k
XN (µ) = XN (µ) − δXN
and repeat these steps until an appropriate stopping criterion is verified.
104 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
applied to (5.16) reads as follows: given fixed µ ∈ P, find δXN ∈ RN such that
Therefore, once we have fixed µ ∈ P, we are able to approximate a solution XN (µ) and pos-
sibly discover its stability properties using some suitable eigenvalue problems [451, 452], but
our aim is to investigate the evolution of the solutions by following its branching behavior.
Moreover, detecting the solution path is the key to recovering the bifurcation diagram, since
it provides a global picture of the model behavior. In view of this need, we introduce the con-
tinuation methods [14, 170, 169, 334], which allow us to generate a sequence of solutions cor-
responding to the selected values of the parameter in order to construct branches of possible
configurations.
The main ingredients for a continuation method are (i) the selection of a good initial guess,
(ii) a rule to select the step size ∆µ, and (iii) a detection tool for the bifurcation points. In
particular, we are interested in predictor-corrector methods, which represent a wide class of
techniques that help the branch tracing by splitting the approximation in two steps, as follows.
Let us start from a solution XN (µj ) to (5.16) for the parameter value µj ; we want to approximate
the solution XN (µj+1 ) for the corresponding next value µj+1 . Instead of directly computing the
pair (XN (µj+1 ), µj+1 ), a predictor-corrector method acts as follows: in the first step the pair
(Xe N (µ e j ) is constructed from (XN (µj ), µj ), and then the former, without being a solution
e j ), µ
for µj+1 , merely serves as an initial guess in the second step.
Since we always deal with pitchfork bifurcations, we will restrict ourselves to the simple
continuation method, which can be seen as a basic predictor-corrector scheme where the pair
(Xe N (µ e j ) is given by (XN (µj ), µj + ∆µj ). We can finally present the algorithm we have
e j ), µ
developed to deal with the branchwise reconstruction of the bifurcating behavior of the models.
Algorithm 5.1 is the implementation result of the building blocks needed to linearize, discretize,
and continue each solution branch; thus, let us now review the combination of these method-
ologies. At the very beginning, one implicitly chooses the branch to approximate by choosing
the initial guess to start the iteration of the Newton method. Indeed, the simplest way to guide
the nonlinear solver to a desired branch is by the choice of the initial guess. Hence, in order
to fully recover the branching behavior, when dealing with the simple continuation method (in
which the parameter step length is fixed and already prescribed), we consider a discrete version
of the parameter space PK = [µ1 , . . . , µK ] ⊂ P of cardinality K and loop over this ordered set.
The loop serves to mimic the predictor-corrector method, assigning the solution obtained for a
given parameter µj as the initial guess for the nonlinear solver at the next iteration of µj+1 . This
allows us to follow the bifurcating behavior of the model.
For the actual linearization and discretization of equation (5.16), we then apply the Newton–
Kantorovich method (5.18) in combination with the Galerkin FEM. The former helps to linearize
the system around the approximation of the solution at the kth iteration, while the latter projects
the problem into the finite-dimensional space XN . As we said, having computed a solution Xj of
the problem for the parameter µj , we can also investigate its stability properties by solving the
eigenvalue problem, which involves the Jacobian matrix JN and the inner product matrix MN .
This will allow us to understand the physical stability of the approximated solutions and to detect
the bifurcation points connected to the qualitative changes of the model.
5.3. Numerical Approximation of Bifurcating Phenomena 105
Among the others, the POD and the greedy techniques [281, 440, 470] are the most well
known and used ones. Let us assume for now that such a basis is already built, and let us see how
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the reduced problem is constructed through the projection on the subspace spanned by it.
For the online phase, it is the efficient and reliable part where the solutions are computed
through the projection on XN . The complexity reduction, which enables the efficiency, is based
on two main assumptions:
(i) The affine decomposition holds, i.e., one can rewrite the weak formulation as a linear com-
bination of µ-independent forms and µ-dependent coefficients. Hence, the contribution of
the parameter is entirely encoded by the coefficients, which usually allows us to rapidly
assemble the system, relying on the precomputed parameter-independent quantities.
(ii) We can accurately approximate the discrete manifold XN with a much lower dimension
space XN , i.e., we need only a small number N N of basis functions.
Therefore, the reduced computational cost comes mainly from avoiding projecting on the large
FE manifold, while relying on the small RB one. It is clear that, in our context, assumption
(i) is difficult to fulfill because of the µ dependence of the solution around which we linearize
the nonlinear weak formulation (5.16). Assumption (ii) is instead usually linked with the con-
cept of Kolmogorov n-width [393, 133], which expresses the ability of a reduced manifold to
approximate the high-fidelity one.
Let us now describe the details of the online phase. We consider the discrete weak formula-
tion (5.14) and, projecting it into the reduced space XN , we obtain the following problem: given
µ ∈ P, we seek XN (µ) ∈ XN such that
XN = span{Σ1 , . . . , ΣN };
N
(m)
X
XN (µ) = XN (µ)Σm . (5.22)
m=1
(m)
From an algebraic standpoint, we can denote by XN (µ) = {XN (µ)}N m=1 ∈ R the reduced
N
N
!
(m)
X
g XN (µ)Σm , Σn ; µ =0, n = 1, . . . , N . (5.23)
m=1
5.3. Numerical Approximation of Bifurcating Phenomena 107
N
!
(m)
X
(GN (XN (µ); µ))n = g XN (µ)Σm , Σn ; µ , (5.24)
m=1
which we will refer to as the reduced residual vector. Moreover, we will denote by V ∈ RN ×N
the transformation matrix whose elements
(V)jm = Σm
(j) (5.25)
are the nodal evaluation of the mth basis function at the jth node. With this notation, we can
rewrite problem (5.23) as
VT GN (VXN (µ); µ) = 0. (5.26)
Finally, the algebraic form of the Newton method, combined with the RB technique, provides
the following formulation: at every iteration k we seek δXN ∈ RN such that
For a general nonlinear problem, equation (5.27) still involves the degrees of freedom of
the high-fidelity problem N . Because of this, the repeated assembly of the reduced Jacobian
compromises the efficiency of the reduced order method during the online phase. As we will see
later, this issue can be overcome by adopting a class of affine-recovery techniques, which allow
consistent speedups by interpolating the nonlinear part of the variational form.
Having described the online simulation, we now go back to the building process of the re-
duced manifold XN .
POD
The key point for the application of the RB method is to construct the basis. Here we focus on
POD [34, 316], which is a compression strategy, closely related to singular value decomposition
(SVD) and principal component analysis (PCA), that serves to reduce the dimensionality of a
given dataset. Indeed, starting from sufficiently rich information about the system, it allows us
to extract the main features by a lower-dimensional representation given by the first computed
modes, which ideally retain most of its energy.
In our applications, we use the POD to generate the RB space XN . Moreover, this is
proved to be optimal in the l2 (RN ) sense. Therefore, in order to construct XN , we consider
Ptrain = {µ1 , . . . , µtrain } ⊂ P and build from the corresponding solutions {XN (µj )}N train
j=1
a symmetric and linear correlation operator C : XN → XN . Studying the associated eigenvalue
problem and sorting the eigenvalue-eigenfunction pairs in descending order, we can use the first
N modes to build the reduced space XN . Moreover, we can choose the minimal integer N such
that the retained energy from the last (Ntrain −N ) snapshots is less than a fixed tolerance P OD ;
this means that we are finding the basis which minimizes, over all possible N -dimensional or-
thonormal bases ZN , the error between the snapshots and their projection through ZN .
It is clear that one of the disadvantages of this technique is the huge number of high-fidelity
solutions that have to be computed to obtain a fairly accurate representation of the high-fidelity
manifold. Indeed, if we miss some information at the first discretization level, the RB, after the
compression step given by the POD, will usually not contain it.
108 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
Hyperreduction Strategies
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
As we said, one of the main assumptions to obtain an efficient reduction method is that the
forms in (5.14) have an affine parametric dependence. This is crucial since it serves to ob-
tain an N -independent online phase, in which we can efficiently compute the solution of the
parametrized PDE at each new instance of the parameter µ. Despite this, when dealing with
nonlinear problems, this assumption is still not sufficient because the residual (5.16) depends on
the parametrized solution itself. Indeed, it is clear that we cannot recover an efficient offline-
online decomposition of
QG
q
(µ)VT GqN (VXN ),
X
GN (XN ; µ) = θG (5.29)
q=1
since its assembly and projection still involve the N degrees of freedom, compromising the
efficiency.
Among others, reliable alternatives to achieve an efficient online phase come from a hyperre-
duction approach. These techniques include EIM [40, 385] and its variant, the discrete empirical
interpolation method (DEIM) [116], which aim to make the online computations independent of
the number of degrees of freedom of the chosen full-order discretization method. The hyperre-
duction techniques serve to approximate the reduced residual vector as
QG
cqθ (XN ; µ)VT hq ,
X
GN (XN ; µ) ≈ (5.30)
q=1
where {hq }Q q
q=1 represents a suitable basis and cθ are the interpolation coefficients, which include
G
Here, we want to describe the main differences between the high-fidelity Algorithm 5.1 and its
reduced variant for the efficient reconstruction of the bifurcation diagram.
First of all, a hyperreduction technique can be applied for the affine-recovery step. Then, to
construct the reduced manifold, we implemented the POD approach described above. To help the
convergence and prevent unnecessary Newton iterations, we select the snapshot location using
the continuation method, which in its simplest variant results in an equispaced sampling of P.
A huge number of snapshots have to be computed (as usual when dealing with POD, but here
it depends also on the number of branches to be approximated). Once the offline phase is finished,
we compute a (global or branch-specific) basis for the reduced manifold; thus we combine the
projection step on XN , the Newton method (5.21), and the simple continuation method for the
reduced vector solution XN .
For the theoretical rationale behind the RB method [470, 281], it has been proved that one of
the main ingredients to have good approximation properties is the regularity of the solution as a
function of the parameter, rather than its regularity in space.
In fact, this can be an issue for bifurcation problems, where, e.g., for pitchfork phenomena
the critical points represent a discontinuity in the parametric sensitivity ∂X(µ)/∂µ. Hence, we
expect that the error analysis in the parameter space P will show higher peaks at bifurcation
points. We remark that in a black-box approach, one can utilize this statistic to forecast the
location of bifurcating phenomena.
5.4. Von Kármán Equations for Structural Buckling of Plates 109
Plates
In the context of continuum mechanics, von Kármán proposed a model to describe all the possible
configurations that a plate under compression can assume [598].
The von Kármán model is often used to describe buckling phenomena. A buckling phe-
nomenon is an example of bifurcating behavior, where the system suddenly changes its configu-
ration. This section is mainly based on the work done in [451].
(0, 1) (L, 1)
(0, 0) (L, 0)
Figure 5.2. An elastic and rectangular plate compressed along the edges parallel to the y direction.
Then, the displacement from its flat state and the Airy stress potential, respectively u and φ,
satisfy the von Kármán equations
∆ u = [µh + φ, u] + f in Ω,
2
(5.31)
∆2 φ = − [u, u] in Ω,
where h and f are the external forces acting on the plate, ∆2 is the biharmonic operator, and
. ∂ 2u ∂ 2φ ∂ 2u ∂ 2φ ∂ 2u ∂ 2φ
[u, φ] = 2 2
−2 +
∂x ∂y ∂x∂y ∂x∂y ∂ y 2 ∂ x2
is the bracket of Monge–Ampére. The von Kármán model is of fourth order, due to the presence
of the biharmonic operator; nonlinear, due to the product of second derivatives in the brackets;
and parametric, due to the buckling coefficient µ varying in a proper range of real numbers.
In order to completely describe the physics involved, we must supplement the system of
PDEs with some opportune BCs for both of the unknowns. Although they can be imposed in
many different ways (see, e.g., [128]), we present only the simply supported BCs
u = ∆u = 0 in ∂Ω,
(5.32)
φ = ∆φ = 0 in ∂Ω,
which are physically complex to reproduce but also the most used for the simulations because
of their versatility in the weak formulation. We remark that, despite the simple BCs chosen, the
goal of this work is to understand the bifurcation behavior of the von Kármán plate equation,
regardless of the numerical constraints that a conforming method for more involved boundary
conditions could impose.
110 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
where the following bilinear and trilinear forms have been introduced:
Z Z Z
a(η, ω) = ∇η · ∇ω dΩ , b(η, ω) = ηω dΩ , c(η, ω, ζ) = [η, ω] ζ dΩ .
Ω Ω Ω
Now we recast the weak formulation (5.33) in the finite-dimensional space XN , as we did in
general in (5.14).
Moreover, in order to embed the simply supported BCs at the discrete level, we set XN as
the space of test functions which vanish on the boundary. In this case, the matrix formulation of
the Newton method (5.27) reads as follows: given µ ∈ P and an initial guess XN 0
∈ XN , for
k = 0, 1, . . . until convergence, we seek δXN = (δuN , δUN , δφN , δΦN ) ∈ XN such that
AN ukN + BN UkN
AN BN 0 0 δuN
C2 + µC0 C1N 0 δUN k 1 k 0 k
AN = AN UN + C N uN + µCN uN
N N k k
,
0 0 AN BN δφN AN φN + BN ΦN
−C1N − C3N 0 0 AN δΦN AN ΦkN − C1N ukN
k+1
where, defining XkN = (ukN , UkN , φkN , ΦkN ), we set XN k
= XN − δXN . As we said, the RB
technique has the same structure as the FEM; thus at the reduced level we will have the same
formulation with the blocks obtained through the projection.
s(u)
Figure 5.3. Left: high-fidelity bifurcation diagram for the rectangular plate with L = 2. Right:
high-fidelity displacements linked to the five branches at µ = 65.
phase, while for the reduced manifold we chose N = 8 bases. For the branch reconstruction, we
designed the continuation method with K = 60 equispaced points.
We can now present the bifurcation diagram in Figure 5.3, where the buckling phenomenon
from the trivial solution happens for µ1 ' 39.47.
Moreover, by properly choosing the initial guess, we were able to detect the second and the
third bifurcations for the plate, respectively, for values of µ2 ' 46.5 and µ3 ' 61.68. The
physical symmetry issue is evident in all buckling points; in fact every solution from the upper
branches is just the reflection with respect to the plane of the plate of the lower ones.
As reported in Figure 5.3, we can distinguish these solutions, and the branch to which they
belong, by their cell-like displacement. Moreover, we can observe that the buckling near µ3 has
a qualitatively different nature since two branches depart from the same point; for this reason it
is said to be a double bifurcation.
An interesting property of the von Kármán model is mode jumping. Usually, while recon-
structing a bifurcating branch, the wave number (or equivalently the number of cells) remains
constant, but as we observed in Figure 5.3 the state corresponding to the second bifurcation
undergoes a sudden change in its behavior at µ4 ' 52.25, where the buckling occurs from an al-
ready buckled state. This is an example of secondary bifurcation, and its reconstruction requires
a more involved continuation method based on arclength algorithms. Finally, we show in Fig-
ure 5.4 that the RB method works well at efficiently approximating the solution. In fact, we can
see that the RB error contour plot for the reconstruction of the first branch, and the corresponding
error on P, show the maximum peaks at the bifurcation points µ1 .
Figure 5.4. Left: RB error on P for the first branch. Right: RB contour error plot of the
displacement u at µ = 65.
112 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
Condensates
In this section, we consider the Gross–Pitaevskii equation in quantum mechanics [456]. Of-
ten referred to as a nonlinear Schrödinger equation, the Gross–Pitaevskii equation models cer-
tain classes of Bose–Einstein condensates (BECs), a special state of matter formed by identical
bosons at ultralow temperatures. Despite the complexity of the model, here we stick to the simple
Gross–Pitaevskii equation and present a multiparameter study varying the chemical potential and
the normalized trap strength. This section is mainly based on work done with A. Quaini [450].
coordinate, and W (r) = 21 τ 2 r2 is the external parabolic potential, with τ being the normalized
trap strength, i.e., the ratio of trappings along and transverse to the plane.
The construction of the steady solution is based on the assumption that the wave-function
Φ(r, t) = φ(r) exp(−iµt), with φ(r) : Ω → C, where µ is the chemical potential, which can be
seen as a measure of the density at the center of the trap and has to satisfy µ ≥ τ . This way, we
obtain the nonlinear problem given by the parametrized PDE
. 1
G(φ; µ) = − ∆φ + |φ|2 φ + W (r)φ − µφ = 0. (5.35)
2
The solutions of the one-dimensional problem (5.35) exhibit a not particularly rich bifurcating
behavior, while its two-dimensional version is far more interesting, including symmetry-breaking
bifurcations and vortex-bearing states [397, 398, 115, 212].
We will analyze it using its bifurcation diagram, which plots some characteristic quantities
of the solution, such as the number of bosons NB in the BEC or a different norm of the particle
density ρ, defined as Z
NB = ρ dr, ρ = |Φ|2 , (5.36)
Ω
with respect to the chemical potential µ. The critical values of the associated eigenvalue problem
.
are given by µ∗m,n = (m + n + 1)τ .
Thus, given an initial energy µ at the linear limit, we increase the chemical potential (and
therefore the number of atoms NB ) in order to approach the strongly nonlinear regime that can
lead to the discovery of new states. For the numerical characterization of the stability for each
state, we refer the reader to [115].
the chemical (bifurcating) parameter and the trap strength, and Ω ⊂ R2 the domain in which we
will approximate the solution.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
We recall that the solution φ to equation (5.35) is a complex function, so we will denote
its real and imaginary parts respectively by ϕ and ψ, which we will treat as the two com-
ponents of the solution. Having represented the model with the parametrized mapping G :
2
X × P → X0 with X = H01 (Ω) , its weak formulation reads as follows: given µ ∈ P,
.
find X(µ) = (ϕ(µ), ψ(µ)) ∈ X such that (omitting the µ dependency)
.
g(X, Y ; µ) = a(X, Y ) + n(X, Y ) + b(X, Y ; τ ) − d(X, Y ; µ) = 0 ∀ Y ∈ X, (5.37)
AN = VT AN V, BN (τ ) = VT BN (τ )V,
N
(n)
X
DN (µ) = VT DN (µ)V, CN = XN VT CN (Σn )V,
n=1
In this case, an extensive study of the associated eigenproblem has not been taken into ac-
count given the infinite symmetry group that acts on the space of solutions. Instead, Hermite
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
guesses are used to initialize the continuation procedure, but more advanced strategies, such as
the deflation technique, can be applied in these scenarios [189, 454]. We chose the parame-
ter space as P = [0, 0.8] × [0.1, 0.3], varying the chemical potential and the concavity of the
parabolic trap. To recover the first bifurcation at µ∗ = τ , corresponding to the ground state,
we applied Algorithm 5.1 for three equispaced values of the trap strength and K = 160 eq-
uispaced values of the chemical potential. The resulting POD basis consists of only N = 12
modes, with a tolerance fixed to P OD = 10−8 . Figure 5.5 shows the evolution of this branch
in a two-parameter bifurcation diagram obtained with the RB version of Algorithm 5.1, using
∆τ = 0.01 and a continuation step of ∆µ = 1.25 · 10−3 . As expected, we see that, as τ in-
creases, the critical value of µ for the first bifurcation increases linearly. As for the efficiency,
due to the lack of a hyperreduction strategy, the speedup is only 1.52 with almost tRB = 2.73(h)
needed for the RB reconstruction of the full three-dimensional diagram in Figure 5.5, in contrast
to the tF E = 4.15(h) of the FE method. Here we were also able to build a unique ROM which
approximates with good accuracy all the first branches by varying τ ∈ [0.1, 0.3]. In fact, as we
can see from Figure 5.5, the reduced error for the density shows a maximum error of order 10−3 ,
with peak located at the current critical value of τ .
Figure 5.5. Left: RB bifurcation diagram for the multiparameter test case with infinity norm of
the density as output. The black dotted line shows the critical values of µ. Right: RB error on P for density
in the L2 - and H01 -norms for τ = 0.18.
Similarly, we recovered the reduced order bifurcation diagram for the first three emerging
branches (ground state for µ∗ = τ , single charge vortex and 1-dark soliton stripe for µ∗ = 2τ ),
shown in Figure 5.6. In this case we obtained a basis with dimension N = 41; in fact now it
has to encode different configurations. Some representative solutions for the three branches at
µ = 0.8 are depicted in Figure 5.6, where it is clear that the more one increases the trap strength,
the more the solutions remain confined in a smaller region around the center of the domain.
We remark that, while the left plot in Figure 5.5 has to be considered as an application of the
branchwise procedure, sampling offline one single qualitative configuration, the right plot in Fig-
ure 5.6 was obtained through a global approach, where all the bifurcating modes are compressed
by means of a single POD. We highlight that in these cases the speedups are still dependent on
N . For this reason we now present an application of the hyperreduction techniques to recover the
efficiency of the whole methodology. We focus on the reconstruction of a single branch, say the
single charge vortex in Figure 5.5 emerging from µ∗ = 2τ . Within the same setting as before, we
consider the EIM and DEIM techniques with greedy tolerance Gr = 10−7 together with a basis
5.6. Hyperelastic Models for Bending Beams 115
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 5.6. Left: RB density functions at µ = 0.8 with τ ∈ {0.15, 0.25} for the first three
branches. Right: RB bifurcation diagram for the multiparameter test case with infinity norm of the density
as output. The blue dashed lines show the critical values of µ for both branching phenomena.
for the reduced manifold of dimension N = 6. We show in Figure 5.7 the reduced order error
on the number of bosons NB . These increased errors are a small price to pay for a considerable
computational speedup. In fact, while it takes tHF = 246(s) to reconstruct the whole branch
with FE, the RB with hyperreduction technique requires only tEIM = 55(s) and tDEIM = 7(s),
respectively, for EIM and DEIM. Finally, we remark that since we are dealing with bifurcating
phenomena, the reduced algorithm can converge to a different configuration of the system. This
is actually the case reported in Figure 5.7, where we show the high-fidelity real part φ of X and
its RB approximation, which is rotated by a certain angle but is still a solution for the system.
Figure 5.7. Left: error between the branches of the bifurcation diagram computed with FE and
RB with EIM/DEIM in the µ-NB plane. Right: high-fidelity and RB real part φ of the solution X for the
1-dark soliton stripe branch, left and right, respectively, at µ = 0.5.
∂φ(X)
F = = ∇u + I and J = det(F ).
∂X
Thus, the equilibrium equation derived from the equation of motion [577, 127] is
−div(P ) = B in Ω, (5.39)
1 T
ψ(F ) = λ1 E : E + λ2 (tr(E))2 /2, and E = (F F − I), (5.41)
2
where λ1 and λ2 are the Lamé constants related to the material properties through the Young
modulus E and the Poisson ratio ν as follows:
E Eν
λ1 = , λ2 = .
2(1 + ν) (1 + ν)(1 − 2ν)
The second hyperelastic constitutive relation we will consider is the neo-Hookean (NH) model,
which can be expressed using the strain energy function
λ1 λ2
ψ(F ) = (I1 − 3) − λ1 ln J + (ln J)2 , (5.42)
2 2
where I1 = tr(C) is the first principal invariant and λ1 , λ2 are the Lamé constants as before.
We remark that the words buckling and bifurcation will be used here interchangeably, but we
will be interested only in the approximation of the first buckling mode, properly following the
postbuckling behavior as the target branch.
5.6. Hyperelastic Models for Bending Beams 117
Starting from the equilibrium equation (5.39), we can consider the boundary value problem, in
the reference domain Ω, given by
in Ω,
−div(P (u)) = B
u = uD in ΓD , (5.43)
P (u)n = T in ΓN ,
where u : Ω → Rd is the in-plane displacement, B is the body force per unit reference area, and
T is the traction force per unit reference length. Thus, our aim is to study the deformation of the
domain Ω, subject to a prescribed displacement uD on the Dirichlet boundary ΓD together with
body and traction forces. We remark that, due to the Dirichlet nonhomogeneous BCs, the func-
.
tion space in which we set the problem is the space HD 1
(Ω) = v ∈ H 1 (Ω) | v = uD ∈ ΓD .
Taking the dot product with a test function v ∈ X = H0 (Ω) and integrating over the reference
1
Furthermore, the boundary value problem in (5.43) for hyperelastic media can also be expressed
as a minimization problem by means of the theorem of virtual work [255]. In fact, we can define
the potential energy of the beam in terms of the strain energy function ψ as
Z Z Z
.
Π(u) = ψ(u) dΩ − B · u dΩ − T · u dΓ. (5.45)
Ω Ω ΓN
Thus, at the minimum point of Π(u), the directional derivative of Π with respect to the change
in u is equal to zero for all v ∈ X, i.e.,
. dΠ(u + δv)
g(u, v) = Dv Π(u) = |δ=0 , (5.46)
dδ
.
and the projection onto the RB space XN reads as follows: given µ ∈ P, we seek uN = uN (µ) ∈
XN that satisfies
JN (ukN (µ); µ)δuN = GN (ukN (µ); µ), (5.47)
updating the solution as uk+1
N = ukN − δuN until convergence.
which correspond to a clamped condition on the left end of the beam and an increasing uni-
form uniaxial compression on the other end. For the high-fidelity setting we choose Ntrain =
118 Chapter 5. Reduced Basis Approaches to Bifurcating Nonlinear Parametrized PDEs
1000 points in the parameter space P = [0, 0.2], and an online continuation method based on
K = 2000 equispaced points in P, in order to properly detect the critical point and follow the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
postbuckling behavior.
We obtained an RB space of dimension N = 5, from which we recover the bifurcation
diagram in Figure 5.8, with s(u) = kuy k∞ as functional output. We can clearly observe that
the beam buckles at the value µ∗ ≈ 0.03. In fact, at this point the vertical component of the
displacement u changes suddenly from being trivial to detecting the buckling. Despite the fact
that we are focusing only on the first branch, which is the most common cause of failure, different
buckling configurations still exist, such as the ones reported in Figure 5.8.
Figure 5.8. Left: RB bifurcation diagrams for the SVK unforced beam. Right: high-fidelity zeroth
and second mode displacement u for the SVK beam with B = (0, 0) at µ = 0.2.
Moreover, we also investigated the multiparameter case, defining µ = (µ, E, ν), where µ
is the bifurcating compression parameter and E and ν are respectively the Young modulus and
the Poisson ratio. For the analysis, we chose a parameter space given by P = [0.0, 0.2] ×
[105 , 107 ] × [0.25, 0.42] and a perpendicular body force B = (0, −1000). Since the parameter
space is three-dimensional, we expected that a much greater number of basis functions would be
needed; indeed, using a POD tolerance P OD = 10−8 we obtained an RB space of dimension
N = 43. The online continuation method was based on K = 201 equispaced µ-values in P,
which corresponds to a continuation step ∆µ = 10−3 , and the bifurcation diagram in Figure 5.9
was depicted for five random pairs of the elasticity constants (E, ν). We remark that we chose
the nontrivial body force was to avoid sharp gradients in the bifurcation diagrams and thus larger
errors in the reduced approximation. From the computational speedup point of view, the plot of
the reduced diagram costs exactly the same as its high-fidelity version: tHF ≈ tRB = 900(s);
despite this we were able to detect the buckling and the related postbuckling behavior for a wide
range of three-dimensional parameter space. We also considered parametrized geometries, where
we tried to understand how this influences the buckling property of the beam. Within the same
physical setting, we considered the SVK model for a beam with parametrized semilength l ∈
[0.5, 1]. We used the online continuation method to reconstruct the three-dimensional bifurcation
diagram in Figure 5.9 with N = 12 bases, using K = 201 equispaced points in P = [0, 0.2].
As we can see from Figure 5.9, the reduced manifold was also able to approximate the buckling
for nonsampled geometries with good approximation accuracy, and as expected we observed that
the longer the beam, the sooner it buckles.
Finally, we introduce a real test case scenario coming from the Norwegian petroleum indus-
try [425], investigating the deformation of a three-dimensional tubular geometry with annular
section. From the analysis of the parametrized geometries, we understood that, as before, the
5.7. Navier–Stokes Flow and the Coandă Effect 119
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 5.9. Left: RB bifurcation plot for SVK beam with B = (0, −1000) for five random pairs
(E, ν) ∈ [105 , 107 ] × [0.25, 0.42]. Right: RB bifurcation plot for SVK beam with B = (0, −1000) and
l ∈ [0.5, 1].
length of the tube plays a fundamental role in the bifurcation, also varying the type of buckling
involved. An example is given by the tubular domain Ω isolated by the annulus section AR r with
r = 0.28 and R = 0.30. We chose SVK material and trivial body and traction forces, and we
fixed the material properties E = 2.1 · 105 and ν = 0.3. In Figure 5.10 we observe the huge dis-
crepancy of the modes’ behavior corresponding to l = 2 and l = 20. Even within this complex
multiparameter setting, we were able to recover the bifurcation diagram using the RB approach.
Figure 5.10. Left: RB bifurcation diagram for tubular unforced geometries with l ∈ [10, 20].
Right: representative solutions of the three-dimensional SVK model with l ∈ {2, 20}.
The problem described above translates to a parametrized PDE of the form (5.1) that loses the
uniqueness of the solution while decreasing the viscosity µ below a certain critical value µ∗ .
We expect two qualitatively different configurations: (i) the symmetric solution—a physically
unstable configuration with a symmetric jet flow, and (ii) the asymmetric solution—a physically
stable configuration with a wall-hugging jet. These coexisting solution branches form the pitch-
fork bifurcation we want to approximate.
We consider a two-dimensional planar straight channel with a narrow inlet and a sudden
expansion, depicted in Figure 5.11, which represent a simplification of the left atrium and the
mitral valve, respectively. We define the inflow and outflow boundaries respectively as Γin =
{0} × [2.5, 5] and Γout = {50} × [0, 7.5], and the boundaries representing the walls as Γ0 =
∂Ω \ (Γin ∪ Γout ). The steady and incompressible NSEs for a viscous flow in Ω, with stress-
free condition on Γout , no-slip homogeneous Dirichlet on Γ0 , and nonhomogeneous Dirichlet
.
vin = [20(5 − x2 )(x2 − 2.5), 0] on Γin , is
in Ω,
−µ∆v + v · ∇v + ∇p = 0
in Ω,
∇·v =0
v = vin on Γin , (5.49)
v = 0 on Γ ,
0
on Γout ,
−pn + (µ∇v)n = 0
where v is the velocity of the fluid, p is its pressure, and µ represents the kinematic viscosity.
For later convenience, we introduce the dimensionless Reynolds number as Re = U w/µ, which
represents the ratio between inertial and viscous forces, where U and w are the characteristic
velocity (maximum inlet velocity U = 31.25) and the characteristic length of the domain (length
of the inlet section w = 2.5), respectively.
40
10
Γ0
Figure 5.11. Domain Ω, which represents a straight channel with a narrow inlet.
As usual, we can project the model by means of the FEM into the finite-dimensional space
XN = VNv × QNp . Thus, the combination of the Galerkin FE and the generic kth step of
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
k k
a(δvN , ψN ; µ) + s(vN , δvN , ψ) + s(δvN , vN , ψN ) + b(ψN , δpN )
= a(vN , ψN ; µ) + s(vN , vN , ψ) + b(ψN , pkN ),
k k k
(5.52)
k
b(δvN , πN ) = b(vN , πN ) ∀ ψN ∈ V0,N , ∀ πN ∈ QN ,
k+1
and then we set XN = XN k
− δXN .
Equivalently, we can project (5.50) onto the RB space XN = VNv × QNp of dimension
N = Nv +Np . This way, the kth reduced step of Newton’s method reads as follows: given µ ∈ P
and an initial guess X0N = (vN 0
, p0N ) ∈ XN for k = 0, 1, . . . we seek δXN = (δvN , δpN ) ∈ XN
that satisfies (5.27), where the reduced Jacobian and residual are given respectively by
k k
) BTN
AN (µ) + S1,N (vN ) + S2,N (vN
JN (XkN (µ); µ) = and
BN 0
k k k T
(5.53)
AN (µ)vN + S1,N (vN )vN + BN pN
GN (XkN (µ); µ) = .
BN vN
Once again, we have introduced in the expression above the transformation matrices with respect
to the different components of the solution, Vv and Vp , respectively, for velocity and pressure,
defining the reduced matrices
Nv
(n)
X
AN (µ) = VvT AN (µ)Vv , BN = VpT BN Vv , Si,N = vN VvT Si,N (Σnv )Vv (5.54)
n=1
Figure 5.12. Left: RB bifurcation diagram for the NS system. Right: velocity magnitude of the
asymmetric branch varying the viscosity µ.
responsible for the change in stability properties of the model. Indeed, the unique symmetric
solution remains stable until it encounters the critical value µ∗ , where it becomes unstable, while
at the same time the wall-hugging solution evolves as the physically stable one. Some repre-
sentative solutions for the lower and middle branches are presented in Figure 5.13, where we
show the symmetric and asymmetric velocity and pressure fields for the same viscosity param-
eter µ = 0.5. Here, the pressure for the lower branch decreases near the bottom left corner of
the expansion, causing the velocity to deflect, hugging the lower wall. Finally, thanks to the
no-slip boundary condition, the flux goes back to the midline, ending with a non-axis-symmetric
outflow. As we can see from Figure 5.14, we were able to recover online a good approximation
of the bifurcating asymmetric branch regarding the errors on the velocity and pressure fields. In
fact, we can observe once again that the RB error has its maximum at the critical value µ∗ , but it
has an average value on P of order 10−7 , confirming the accuracy of the methodology.
The stability is investigated by a standard eigenvalue analysis, analyzing the behavior of the
first Neig = 100 eigenvalues while varying the viscosity of the fluid. These eigenvalues are
plotted in Figure 5.15 for the stable lower branch and unstable middle branch. Note that since
the NS operator is not symmetric, we have both real and complex eigenvalues. Here we are just
interested in pitchfork bifurcation; thus in the zoom we follow the behavior of the biggest real
eigenvalue. Looking at the stability of the lower branch, all the eigenvalues of the NS system have
negative real part, and from this we can assert the stability of the wall-hugging branch. Indeed,
the zoom in the left plot of Figure 5.15 shows no crossing of negative real part eigenvalues. On
the contrary, the closeup in the right plot for the symmetric flow shows a sign change and thus
indicates instability.
Figure 5.13. Representative solutions for the NS system at µ = 0.5, velocity and pressure fields,
lower and middle branch, top and bottom respectively.
5.7. Navier–Stokes Flow and the Coandă Effect 123
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 5.14. RB errors with respect to µ ∈ P for the velocity and pressure of the asymmetric branch.
Figure 5.15. Eigenvalues of the state eigenproblem in the complex plane for the NSE: stable and
unstable solutions, left and right panels, respectively.
Chapter 6
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Reduced Basis
Stabilization for
Convection-Dominated
Problems
6.1 Introduction
In this chapter we present stabilized reduced basis (RB) models for convection-dominated prob-
lems. We start by defining a stabilized RB model for the advection-diffusion equation. In this
problem, the instabilities appear when the Péclet number is higher than one. Different stabiliza-
tion procedures are well known for the finite element (FE) problem [477], and we extend it for
the RB framework, following the work performed by Pacciarini and Rozza in [433].
Then we present a stabilized RB model for both the Stokes and Navier–Stokes problems.
The consideration of a nonstable pair of FE spaces for velocity and pressure leads to a nonstable
FE solution. The discrete spaces must verify a discrete inf-sup condition in order to guaran-
tee the well-posedness of the discrete solution for the Stokes and Navier–Stokes problems. For
the Stokes problem, we consider the Brezzi–Pitkäranta stabilization [75], while for the Navier–
Stokes problem we consider the streamline upwind Petrov–Galerkin (SUPG) stabilization pro-
cedure [78]. In the RB framework, different stabilization strategies are compared for both Stokes
and Navier–Stokes problems. We see that the offline-only stabilization is not consistent and the
RB solutions computed are far from the FE one. But with the offline-online stabilization for
the RB problem, the RB solution is computed accurately, either considering the inner pressure
supremizer [471] for the velocity space enrichment or not.
After that, we present a stabilized variational multiscale (VMS)-Smagorinsky model. In
this VMS-Smagorinsky model, the eddy viscosity term acts only on the resolved small scales.
We consider a local projection stabilization term for the pressure, derived from the high-order
term-by-term stabilization method (cf. [105, 108]). This methodology allows us to obtain more
accuracy than with the classic penalty stabilization procedure, with a reduced computational cost
[108]. The existence of a pressure stabilization term lets us consider a nonstable pair of FEs,
such as P2-P2, with the consequent increase of accuracy in both velocity and pressure.
The development of the a posteriori error bound estimator for the RB VMS-Smagorinsky
model with local projection stabilization on the pressure is done using the Brezzi–Rappaz–
Raviart (BRR) theory. In the online phase, we will also need the empirical interpolation method
(EIM) to approximate the nonlinear and nonaffine terms, with respect to the parameter, that
take part in the model considered in this chapter. Thus, we need to approximate using the EIM
125
126 Chapter 6. Reduced Basis Stabilization for Convection-Dominated Problems
the eddy-viscosity term and the pressure stabilization coefficient. Finally, we present numerical
results for the two-dimensional cavity steady problems, for which we obtain speedup rates of
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
several thousand.
The chapter is structured as follows: in Section 6.2 we present the stabilized RB model for
the advection-diffusion problem. After that, we present the Stokes problem in Section 6.3, where
we define both the FE problem in Subsection 6.3.1 and the stabilized RB problem in Subsection
6.3.2. Then, in Section 6.4, we present the Navier–Stokes problem, defining both the stabilized
FE problem in Subsection 6.4.1 and the stabilized RB model in Subsection 6.4.2. Finally, we
present a stabilized VMS-Smagorinsky model in Section 6.5. We present the FE problem for this
model in Subsection 6.5.1 and the RB formulation in Subsection 6.5.2. Moreover, we present the
numerical analysis for the well-posedness of the VMS-Smagorinsky model in Subsection 6.5.3
and the development of the a posteriori error estimator in Subsection 6.5.4. Some numerical
experiments for the VMS-Smagorinsky model are presented in Subsection 6.5.5.
where here we are considering that ε and β are functions Ω → [0, +∞) and Ω → R2 , respec-
tively, both depending on a given parameter value µ ∈ D. For problem (6.1) we consider suitable
Dirichlet, Neumann, or mixed boundary conditions.
To study problem (6.1), let us consider V ⊂ H 1 (Ω) to be a suitable Sobolev space. For
simplicity of the analysis, let us consider V = H01 (Ω), corresponding to homogeneous Dirichlet
boundary conditions. Thus, we define the variational formulation of problem (6.1) as follows:
(
Find u ∈ V such that
(6.2)
a(u, v) = F (v) ∀v ∈ V,
with
a(u, v) = (ε(µ)∇u, ∇v)Ω + (β(µ) · ∇u, v)Ω ,
F (v) = hf, vi ,
where we denote by h·, ·i the duality pairing between V and V 0 , where V 0 is the dual space of
V . Moreover, considering Vh ⊂ V as an FE internal approximation of V , we define the discrete
problem as follows:
(
Find uh (µ) ∈ Vh such that
(6.3)
a(uh , vh ; µ) = F (vh ; µ) ∀vh ∈ Vh .
It is well known that the advection-diffusion discrete problem has instabilities when the advection
term dominates the diffusive one. To analyze this issue, let Th be an admissible triangularization
of the domain Ω, and let K be an element of Th . We consider that problem (6.3) is advection
dominated in K if
|β(µ)|hK
Pe(µ) := > 1 ∀µ ∈ D, (6.4)
2ε(µ)
where here hK is the diameter of K. When we are dealing with advection-dominated problems,
we need to consider some additional terms in the discrete problem in order to avoid instabilities
6.3. Steady Stokes Equations 127
in the numerical solutions. There are several stabilization methods for advection-dominated
problems in the literature, e.g., [78, 312, 477].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
We consider a stabilized model, with a stabilization term in the discrete problem (6.3), de-
fined in a general way as
X hk
astab (uh , vh ; µ) = δK ε(µ)∆uh + β(µ) · ∇uh , (ρε(µ)∆uh + β(µ) · ∇uh ) ,
|β(µ)| K
K∈Th
(6.5)
where considering different values for ρ leads us to different stabilization methods. Here δk are
some suitable coefficients for the different stabilization procedures. The SUPG method corre-
sponds to ρ = 0, while ρ = 1 or ρ = −1 corresponds to the Galerkin least-squares (GLS) and
Douglas–Wang (DW) methods, respectively. For a detailed explanation of stabilization methods
we refer the reader to [477] and the references therein. Thus, the stabilized FE problem can be
defined as follows:
(
Find uh (µ) ∈ Vh such that
(6.6)
a(uh , vh ; µ) + astab (uh , vh ; µ) = F (vh ; µ) ∀vh ∈ Vh .
In the rest of this section, we consider the SUPG method for the stabilization procedure.
Let SN = {µ1 , . . . , µN } be a set of properly selected parameter values, e.g., with a greedy
algorithm [465, 471, 505], using a suitable a posteriori error estimator. We define the RB space
as VN = span{uh (µi ), i = 1, . . . , N }, where uh (µi ) are the high-fidelity solutions of problem
(6.6) for each µi ∈ SN .
We can consider two stabilization procedures for the RB model: the offline-only stabilization
and the offline-online stabilization. The main difference between methods is that in the offline-
only stabilization method we do not consider the stabilization term defined in (6.5) in the RB
model, while we do consider it in the offline-online stabilization model. Thus, the offline-only
stabilization method reads as follows:
(
Find uN (µ) ∈ VN such that
(6.7)
a(uN , vN ; µ) = F (vN ; µ) ∀vN ∈ VN .
Recall that in both cases, the RB space VN is constructed with the solution of the stabilized
FE problem (6.6). In [433] there is a comparison of both procedures, where the authors conclude
that the consistency of the offline-online stabilization model is necessary for the good behavior
of the RB solution. They observe how with the offline-online stabilization model they are able
to obtain a reduced solution close to the FE solution, while using the offline-only stabilization
model, they are not able to approximate in a good way the RB solution with respect to the FE
solution.
in Ω,
−ν∆u + ∇p = f
∇·u=0 in Ω, (6.9)
on ∂Ω,
u = gD
where u is the velocity, p is the pressure, f is a given force, and ν is the viscosity. In the RB
framework, we can consider as a parameter, e.g., the viscosity of the fluid or a deformation of the
domain. In this last case, we should define a reference domain and perform all the computations
in this reference domain. For more details, see, e.g., [12, 30, 470]. We consider that ∂Ω is
decoupled into ∂Ω = ΓD0 ∪ ΓDg , where ΓD0 and ΓDg , respectively, are the homogeneous and
nonhomogeneous Dirichlet boundaries.
For the sake of simplicity, in the following we consider that gD = 0. In the case of nonhomo-
geneous boundary conditions, it is enough to define a lift function uD such that uD |ΓDg = gD .
To define the variational formulation of problem (6.9), let us introduce the spaces V =
(H01 (Ω))2 and M = L20 (Ω) on the reference domain. Moreover, let us consider Vh ⊂ V and
Mh ⊂ M to be two FE internal approximations. Thus, the discrete weak formulation is
a(uh , vh ; µ) + b(vh , ph ; µ) = hf, vh i ∀vh ∈ Vh ,
(
(6.10)
b(uh , qh ; µ) = 0 ∀qh ∈ Mh ,
where
a(u, v; µ) = ν(∇u, ∇v)Ω , b(v, q; µ) = −(∇v, q)Ω .
Problem (6.10) must verify the discrete inf-sup condition
b(vh , qh ; µ)
∃β0 > 0 : β0 ≤ βh (µ) = inf sup , (6.11)
qh ∈Mh vh ∈Vh kvh k1 kqh k0
where we are denoting by k · k0 and k · k1 the standard L2 -norm and H 1 seminorm, respectively.
There are different possibilities for the choice of the stabilization term in problem (6.12),
depending on the stabilization term spres . The simplest is the Brezzi–Pitkäranta stabilization,
introduced in 1984 in [75], in which the stabilization term is given by
X
spres (qh ; µ) = h2k (∇ph , ∇qh )K . (6.13)
K∈Th
For a more detailed explanation of stabilization procedures for the Stokes problem, we refer the
reader to [477]. Due to its simplicity and reliability, in the following, we will consider the Brezzi–
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Pitkäranta stabilization as the stabilization procedure. The stabilized model (6.12) satisfies a new
inf-sup condition:
b(vh , qh ; µ)
0 < β0 ≤ βh = sup + h2K k∇qh k20 . (6.15)
vh ∈Vh k∇vh k0
leads to an accurate approximation for both cases: with and without the inner pressure suprem-
izer, although the case of not considering the inner pressure supremizer gives a better approxima-
tion, with less computational effort. This better approximation might be due to the “pollution”
of the velocity space when we enrich it by adding the inner pressure supremizer.
1
− ∆u + u · ∇u + ∇p = f in Ω,
Re
∇·u=0 in Ω, (6.17)
u=g on ∂Ω,
D
where, again, u is the velocity, p is the pressure, f is a given force, and, here, Re is the Reynolds
number. In the RB framework, we can consider as a parameter, e.g., the Reynolds number or a
deformation of the domain.
For the sake of simplicity, again in this section, we consider that gD = 0. In the case
of nonhomogeneous boundary conditions, it is enough to define a lift function uD such that
uD |ΓDg = gD , leading to a similar analysis of the problem.
To define the weak formulation of problem (6.17), let us recall the spaces V = (H01 (Ω))2
and M = L20 (Ω) on the reference domain. Moreover, let us consider Vh ⊂ V and Mh ⊂ M to
130 Chapter 6. Reduced Basis Stabilization for Convection-Dominated Problems
where a(·, ·; µ) and b(·, ·; µ) are defined in the same way as in the Stokes problem and
In order to obtain a suitable solution for problem (6.18), the discrete spaces Vh and Mh must
verify the discrete inf-sup condition (6.11).
Here, the stabilization terms for the convective term sconv , pressure spres , and grad-div stabi-
lization sdiv are given, respectively, by
X 1
sconv (vh ; µ) = δ h2K − ∆uh + uh · ∇uh + ∇ph , uh · ∇vh ,
Re K
K∈Th
X 1
spres (qh ; µ) = δ h2K − ∆uh + uh · ∇uh + ∇ph , ∇qh ,
Re K
K∈Th
and X
sdiv (vh ; µ) = γ (∇ · uh , ∇ · vh )K ,
K∈Th
where δ and γ are suitable stabilization coefficients. We refer the reader to [78, 429, 578] and
the references therein.
This stabilized Navier–Stokes model corresponds to the SUPG model [78]. Other possibil-
ities for stabilization are the GLS [298] and the stabilization model introduced by Franca and
Frey in [205].
With this SUPG stabilization of the Navier–Stokes problem, we ensure that the spaces Vh
and Mh satisfy the discrete inf-sup condition
b(vh , qh ; µ)
0 < β0 ≤ βh = sup + h2K k∇qh k20 . (6.20)
vh ∈Vh k∇vh k0
an offline-online stabilization with and without the inner pressure supremizer velocity space
enrichment.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Here, the snapshots for the construction of spaces VN and MN are selected by a weak greedy
algorithm (cf. [179]) by constructing an a posteriori error estimator based on the BRR theory.
See [162, 164, 387] for more details about the construction of the a posteriori error estimator for
the Navier–Stokes equations.
In [13] the reader can find numerical studies of this problem where the different stabilization
procedures for the RB problem are compared. There, the authors conclude that, again, the option
of an offline-only stabilization procedure is not valid since the RB solutions have a big error
with respect to the FE solutions. They conclude that considering the offline-online stabilization
preserves the consistency of the problem and leads to more accurate RB solutions. With respect
to the enrichment of the velocity space with the inner pressure supremizer, in both cases they
obtain accurate RB solutions, with the advantage in the case of not considering it of requiring
less computational effort in the computation of the RB solution.
6.5.1 FE Problem
Let Ω be a bounded domain of Rd (d = 2, 3), with Lipschitz-continuous boundary Γ, which we
suppose is split into Γ = ΓD ∪ ΓN . Here, ΓD is decoupled into ΓD = ΓD0 ∪ ΓDg , where ΓD0
and ΓDg , respectively, are the homogeneous and nonhomogeneous Dirichlet boundaries, while
ΓN is the Neumann boundary.
To set the VMS-Smagorinsky model, as in Chapter 3, we decompose the velocity and pres-
sure spaces, Y and M , respectively, as
Y = Yh ⊕ Y 0 , M = Mh ⊕ M 0 ,
where Yh and Mh are respectively the large-scale finite-dimensional spaces for velocity and
pressure, while Y 0 and M 0 are the small-scale complementary spaces. We suppose that the large
and small scales are separated by assuming that the sum is direct, i.e., Yh ∩ Y 0 = {0} and
Mh ∩ M 0 = {0}.
We define Yh = Y h ⊕Yh0 and Mh = M h ⊕Mh0 . Thus, uh = uh +u0h ∈ Yh and ph = ph +p0h .
Let us define the discrete space Vhl (Ω) = {r ∈ C 0 (Ω) : r|K ∈ Pl (K) ∀K ∈ Th }. Thus, we
define the velocity and pressure FE spaces as Yh = (Vhl (Ω) ∩ H01 )d and Mh = Vhm ∩ L2 . For the
subgrid eddy viscosity modeling in the VMS-Smagorinsky model, let us consider Y h = Πh Yh ,
with Πh a uniformly stable interpolation operator on Y h , where
Y h = [Vhl−1 (Ω)]d . (6.22)
132 Chapter 6. Reduced Basis Stabilization for Convection-Dominated Problems
With this notation, we are identifying Y h as the large-scale velocity space and Yh0 = (Id−Πh )Yh
as the resolved small-scale velocity space. We denote by µ ∈ D ⊂ R the parameter considered
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
for the RB problem, which is again the Reynolds number, leading to the following variational
formulation problem:
Find (uh , ph ) = (uh (µ), ph (µ)) ∈ Yh × Mh such that
a(uh , vh ; µ) + b(vh , ph ; µ) + a0S (wh ; wh , vh ; µ)
+ c(uh , uh , vh ; µ) + c(uD , uh , vh ; µ) (6.23)
+ c(uh , uD , vh ; µ) = F (vh ; µ) ∀vh ∈ Yh ,
b(uh , qh ; µ) = 0 ∀qh ∈ Mh ,
where here again, we are defining wh = uh + uD , with uD a lift function. The bilinear forms in
(6.23), a(·, ·; µ) and b(·, ·; µ), are defined as
a(uh , vh ; µ) = (∇uh , ∇vh )Ω , b(vh , ph ; µ) = −(ph , ∇ · vh )Ω ,
and the trilinear form c(·, ·, ·; µ), is defined as
c(wh , uh , vh ; µ) = (wh · ∇uh , vh )Ω .
The nonlinear form a0S (·; ·, ·; µ) is a multiscale Smagorinksy modeling for the eddy viscosity
term, and it is given by
Z
0
aS (zh ; uh , vh ; µ) = νT (Π∗h zh )∇(Π∗h uh ) : ∇(Π∗h vh ) dΩ, (6.24)
Ω
where Π∗h = Id − Πh . In VMS terminology, this corresponds to the small-small setting eddy
viscosity term (cf. [299]). Other possibilities are the large-small setting that models the turbulent
viscosity by a function of the whole resolved velocity, taking
Z
0
aS (zh ; uh , uh ) = νT (zh )∇(Π∗h uh ) : ∇(Π∗h vh ) dΩ. (6.25)
Ω
Depending on the stabilization procedure, we can distinguish two different kind of methods:
residual-based and penalty. The residual-based stabilization methods include GLS (cf. [298]) and
modifications of this method such as the SUPG method (cf. [78, 477]) or the adjoint-stabilized
method (cf. [204]). On the other hand, for penalty methods we particularly mention the penalty
term-by-term stabilized method (cf. [105]), which is an extension of the penalty method intro-
duced in [75]. In the latter method, the penalty term acts on the discretization of the pressure
gradient.
The stabilization procedure that we consider in this work is a local projection-stabilization
(LPS), introduced in [132], on the pressure gradient. The stabilization procedure considered
here is based on the high-order term-by-term stabilized method (cf. [108]) that stabilizes each
operator, e.g., the convection term or pressure gradient. We consider the stabilization on the latter
next.
Thus, we consider the LPS-VMS-Smagorinsky model, with stabilization on the pressure gra-
dient, as follows:
Find (uh , ph ) = (uh (µ), ph (µ)) ∈ Yh × Mh such that
a(uh , vh ; µ) + b(vh , ph ; µ) + a0S (wh ; wh , vh ; µ)
+ c(uh , uh , vh ; µ) + c(uD , uh , vh ; µ) (6.26)
+ c(u , u , v ; µ) = F (v ; µ) ∀v ∈ Y ,
h D h h h h
b(uh , qh ; µ) + spres (ph , qh ; µ) = 0 ∀qh ∈ Mh ,
6.5. Stabilized VMS-Smagorinsky Turbulence Model 133
where here spres (·, ·; µ) is the projection-stabilization term for the pressure, defined as
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
X
spres (ph , qh ; µ) = τp,K (µ)(σh∗ (∇ph ), σh∗ (∇qh ))K . (6.27)
K∈Th
Here, τp,K (µ) is the stabilization coefficient in (6.27), and it must verify the following hypothe-
sis: there exist two positive constants α1 , α2 , independent of h, such that
In the following, we will use the stabilization pressure coefficient τp,K (µ) proposed by Cod-
ina in [132] and used in [110, 512]:
−1
1/µ + ν T|K (µ)
UK (µ)
τp,K (µ) = c1 + c2 , (6.29)
h2K hK
where ν T|K is some local eddy viscosity, UK is a local velocity, and c1 , c2 are some positive
experimental constants. Defining τp,K (µ) in this way, we ensure (6.28).
By defining Xh = Yh × Mh , we rewrite problem (6.23) as follows:
where
1
A(Uh (µ), Vh ; µ) = A0 (Uh , Vh ) + A1 (Uh , Vh ) + A2 (Uh ; Vh )
µ (6.31)
+ A3 (Uh ; Vh ) + Apres (Uh , Vh ),
where, with U = (u, pu ), V = (v, pv ), and w = u + uD , the various terms are defined by
Z
A0 (U, V ) = ∇u : ∇v dΩ,
Ω
Z Z
A1 (U, V ) = (∇ · u)p dΩ − (∇ · v)pu dΩ
v
Ω
Z ΩZ
6.5.2 RB Problem
In this section we present the RB model derived from the FE problem (6.30). The reduced spaces
are computed using the greedy algorithm. The RB problem reads as follows:
Here, the reduced space is defined as XN = YN × MN , where the reduced velocity space YN
and the reduced pressure space MN are given by
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Both the small-small setting of the Smagorinsky eddy diffusion term defined in (6.31),
νT (Π∗h (∇w)) := νT (µ), and the pressure stabilization coefficient, τp,K (µ), defined in (6.27)
have a nonlinear representation with respect to the parameter and consequently need to be lin-
earized with the EIM.
We consider an approximation of the eddy viscosity term and the pressure stabilization term.
For this purpose, we need to build two RB spaces, WM S
1
= {q1S (µ), . . . , qM
S
1
(µ)} and WMP
2
=
{q1P (µ), . . . , qM
P
2
(µ)}, by a greedy selection procedure, with W S
M1 and W P
M2 the EIM reduced
spaces associated with the eddy viscosity term and the pressure stabilization term, respectively.
Thus, we approximate them by the following trilinear forms:
where
M1
X
â0S (wN , vN ; µ) = σkS (µ)s(qkS , wN , vN ),
k=1
M2
(6.36)
X
ŝpres (pN , qN ; µ) = σkP (µ)r(qkP , pN , qN ),
k=1
with X
s(qkS , wN , vN ) qkS ∇(Π∗h wN ), ∇(Π∗h vN )
= K
,
K∈Th
X (6.37)
r(qkP , pN , qN ) qkP σh∗ (∇pN ), σh∗ (∇qN ) K .
=
K∈Th
Here we are considering that the approximations given by the EIM for νT (µ) and τK,p (µ) are,
respectively,
M1
X
IM1 [νT (µ)] = σkS (µ)qkS ,
k=1
M2
(6.38)
X
IM2 [τK,p (µ)] = σkP (µ)qkP .
k=1
This representation leads to a linearization of the RB problem (6.32), with an affine depen-
dence with respect to the parameter, given by the following:
We can express the solution (uN (µ), pN (µ)) ∈ XN of (6.39) as a linear combination of the
basis functions:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
N N
p
X X
uN (µ) = uN
j (µ)ζ j
v
, pN (µ) = pN
j (µ)ξj .
j=1 j=1
The tensor representations associated with the eddy viscosity term and the pressure stabiliza-
tion coefficient are defined as follows:
X
(S0N (qsS ))ij = (qsS ∇(Π∗h ζjv ), ∇(Π∗h ζiv ))K , i, j = 1 . . . , N, s = 1, . . . , M1 , (6.40)
K∈Th
With this tensor representation for the nonlinear terms in (6.32) in the offline phase, it holds that
M1 M2
ŝpres (ξjp , ξip ; µ) =
X X
â0S (ζjv ; ζiv ; µ) = σsS (µ)S0N (qsS ) and σsP (µ)PN (qsP ).
s=1 s=1
Problem (6.39) is solved by a semi-implicit evolution approach. Note that thanks to that
representation, we are able to solve efficiently the online phase due to the linearization of the
eddy viscosity and pressure stabilization terms.
In this case, we no longer consider the supremizer for the velocity space enrichment as in
the previous sections, due to the offline-online stabilization proposed for the RB problem (6.32).
With this offline-online stabilization, the inner pressure supremizer operator is no longer nec-
essary to recover the pressure. Moreover, considering the supremizer operator for the velocity
space enrichment is counterproductive because the RB solution is no longer consistent with the
FE solution, i.e.,
UN (µk ) = Uh (µk ) ∀µk ∈ S = {µ1 , . . . , µN }, (6.42)
where S = {µ1 , . . . , µN } is the set of parameter values chosen in the greedy algorithm. The fact
of losing the consistency causes the greedy algorithm to rechoose a parameter value previously
chosen, making the matrix of the RB problem singular.
1/2
and its associated norm kf kτp = (f, f )τp . We also consider a modification in the weight of the
scalar product. Let
Z
1 0∗
(uh , vh )T = + νT ∇uh : ∇vh dΩ ∀uh , vh ∈ Yh , (6.44)
Ω µ
136 Chapter 6. Reduced Basis Stabilization for Convection-Dominated Problems
X
µ = arg min (CS hK )2 min |∇(Π∗h wh )(µ)|(x)χK (x),
µ∈D x∈K
K∈Th
with wh (µ) the velocity solution of (6.26). Here again, we define the norm of space Xh as
q
kUh kX = kuh k2T + kph k20,2,Ω ∀Uh = (uh , ph ) ∈ Xh . (6.45)
The demonstration of this inf-sup condition can be derived from [108]. This result can be
extended by considering that the triangulation {Th }h>0 is regular rather than uniformly regular.
For further details, see [108].
Let us consider the directional derivative of the operator A(·, ·; µ). If we derive each operator
term in (6.31), we get
According to the BRR theory (cf. [76, 93]), the well-posedness problem (6.30) is guaranteed
by the following continuity and inf-sup conditions:
1 1
k∇uD k0,2,Ω < and k∇uh k0,2,Ω ≤ − k∇uD k0,2,Ω .
µCµ2 C ? µCµ2 C ?
The inf-sup condition (6.48) is verified because the operator b(vh , qh ; µ) satisfies the discrete
inf-sup condition (6.11). Moreover, the condition of Proposition 6.3 is verified when the Dirichlet
boundary data is sufficiently small.
Lemma 6.4. There exists a positive constant ρT such that, ∀Uh1 , Uh2 , Zh , Vh ∈ Xh ,
∂1 A(Uh1 , Vh ; µ)(Zh ) − ∂1 A(Uh2 , Vh ; µ)(Zh ) ≤ ρT kUh1 − Uh2 kX kZh kX kVh kX . (6.50)
We define the following inf-sup and continuity constants, associated with the well-posedness
of the RB problem (6.32):
∂1 A(UN (µ), Vh ; µ)(Zh )
0 < βN (µ) ≡ inf sup , (6.51)
Zh ∈Xh Vh ∈Xh kZh kX kVh kX
∂1 A(UN (µ), Vh ; µ)(Zh )
∞ > γN (µ) ≡ sup sup . (6.52)
Zh ∈Xh Vh ∈Xh kZh kX kVh kX
We define the a posteriori error bound estimator as
βN (µ) h p i
∆N (µ) = 1 − 1 − τN (µ) , (6.53)
2ρT
where τN (µ) is given by
4N (µ)ρT
τN (µ) = 2 (µ) , (6.54)
βN
with N (µ) the dual norm of the residual. The a posteriori error bound estimator is stated by the
following result.
Theorem 6.5. Let µ ∈ D, and assume that βN (µ) > 0. If problem (6.30) admits a solution
Uh (µ) such that
βN (µ)
kUh (µ) − UN (µ)kX ≤ ,
ρT
then this solution is unique in the ball BX UN (µ), βNρT(µ) .
Moreover, assume that τN (µ) ≤ 1 for all µ ∈ D. Then there exists a unique solution Uh (µ)
of (6.30) such that the error with respect UN (µ), the solution of (6.32), is bounded by the a
posteriori error bound estimator, i.e.,
kUh (µ) − UN (µ)kX ≤ ∆N (µ), (6.55)
with effectivity
2γN (µ)
∆N (µ) ≤ + τN (µ) kUh (µ) − UN (µ)kX . (6.56)
βN (µ)
138 Chapter 6. Reduced Basis Stabilization for Convection-Dominated Problems
In this section we show the numerical results of the pressure LPS-Smagorinsky RB model
for the lid-driven cavity two-dimensional problem. The numerical results were obtained using
FreeFrem++ (cf. [268]).
We consider the Reynolds number as a parameter, ranging in the interval D = [1000, 5100].
We consider a regular mesh with 5000 triangles and 2601 nodes. The FE pair chosen is the pair
P2-P2. We impose a nonhomogeneous boundary condition uD = 1 on the top of the geometry,
while on the rest of the boundaries, we impose homogeneous Dirichlet conditions.
For this test, considering the supremizer enrichment for the velocity space leads to a reselec-
tion of a parameter value already selected in the greedy algorithm. In Figure 6.1 (left) we show
that in this case, in the fourth iteration, the greedy algorithm breaks down.
In the EIM considered in this test, we compute M1 = 20 bases for the eddy-viscosity ap-
proximation and M2 = 25 bases for the pressure stabilization constant approximation to reach
a prescribed tolerance of εEIM = 5 · 10−4 . In Figure 6.1 (right) we show the evolution of the
error for both EIMs.
For the greedy algorithm we prescribe a tolerance of εRB = 10−4 , which is reached for
N = 16 bases. In Figure 6.2 (left) we show the evolution of the a posteriori error bound estimator
during the greedy algorithm. Then, in Figure 6.2 (right) we show the comparison of the error
and the a posteriori error bound estimator. In this case, the a posteriori error bound estimator is
about two orders of magnitude greater than the exact error.
105 100
max τN (µ) without supremizer kνT (µ) − I[νT (µ)]k∞
µ∈D
max τN (µ) with supremizer kτK,p (µ) − I[τK,p (µ)]k∞
µ∈D
104
10-1
103
10-2
2
10
-3
10
101
10
0
10-4
1 2 3 4 5 6 7 8 5 10 15 20 25
N M
Figure 6.1. Comparison of the stabilization with/without supremizer (left) and evolution of the
error in the EIM (right).
105 10-2
10-5
101
100 10-6
10-1
10-7
10-2
10-8
10-3
10-9
10-4
10-10
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 1000 1500 2000 2500 3000 3500 4000 4500 5000
N Re
Figure 6.2. Evolution of the error in the greedy algorithm (left) and a posteriori error estimator
for N = 16, with error (right).
6.6. Conclusions 139
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In Figure 6.3, we show two solutions for a prescribed value of the Reynolds number µ =
2751, for the FE problem (right) and the RB problem (left). We can observe the similarity of
both solutions, since the error between them is of order 10−6 .
Finally, in Table 6.1 we summarize results obtained for several values of the Reynolds num-
ber. For this test, we obtain speedup rates that can reach the order of tens of thousands, with the
error between the FE and RB solutions quite small.
Table 6.1. Computational time for FE solution and RB online phase, with the speedup and the error.
6.6 Conclusions
In this chapter we have presented different RB stabilization models for convection-dominated
problems. We started with an advection-diffusion problem, then we defined the different sta-
bilization procedures for the Stokes and Navier–Stokes problems, and, finally, we presented a
stabilized VMS-Smagorinsky model, with the numerical analysis for the construction of the a
posteriori error estimator used in the weak greedy algorithm for the snapshot selection. We have
seen that the offline-online procedure is the best option in the models presented in this chapter
since it is the one that preserves the consistency in the RB problem. Moreover, since the inner
pressure supremizer for the enrichment of the velocity space does not influence the Stokes and
Navier–Stokes accuracy of the solution, not considering it decreases the computational effort of
computing the solution. Finally, in the VMS-Smagorinsky model, we have seen that considering
the inner pressure supremizer is no longer valid.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Part II
Order Models
and Discontinuous
Galerkin-Based Reduced
Finite Volume, Spectral Element,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 7
Finite Volume–Based
Reduced Order Models
for Laminar Flows
7.1 Introduction
This chapter focuses on reduced order models (ROMs) for fluid problems discretized with the
finite volume method (FVM) [408, 596]. Currently, the use of the FVM for solving industrial
problems is predominant in many different fields, e.g., fluid dynamics and continuum mechanics
applications. In fact, most of the computational fluid dynamics (CFD) solvers used to tackle
real-life applications are based on the FVM, e.g., Fluent [391] and STAR CCM+ [3] (commercial
codes) and OpenFOAM [612] (open-source code). These CFD solvers are well equipped with
the computational tools needed to tackle problems coming from various engineering fields.
The demand for ROMs which could reduce computational costs has been increasing in recent
decades. These observations and demands motivate investing in efforts to construct ROMs for fi-
nite volume–based discretization. In this work, the full-order solver is chosen to be OpenFOAM,
a choice justified by the fact that it is open-source code. In addition, OpenFOAM offers well-
established documentation and tutorials for various benchmark problems. As a consequence, the
ROMs proposed in this chapter have taken into consideration the full-order modeling techniques
utilized by OpenFOAM.
143
144 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
obtain a stable-pressure ROM field. The supremizer stabilization method ensures that a reduced
order version of the inf-sup condition, needed by saddle-point formulations, such as the NSEs, is
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
met; initially it was constructed for finite element (FE) POD-Galerkin ROM in [34]. The POD-
Galerkin ROMs in [380, 548, 550] differ in the methodology adopted to reduce the NSEs. Further
details on the approaches of these works are recalled in the sections related to U-ROM and PPE-
ROM. We mention works in which numerical analysis is used to study ROMs for turbulent flows
and ROMs for pressure approximations [513, 333, 91]. The last works present different ROM
frameworks for computing the reduced pressure field.
Recently, the work [551] tackled the issue of geometrical parametrization for FVM-based
POD-Galerkin ROM. The authors in [551] propose a ROM which is fully consistent with the
SIMPLE algorithm [438]. The SIMPLE algorithm is the full-order solver approach for steady
flows in OpenFOAM and is a segregated approach. Other works which deal with geometrical
parametrization in FVM-based ROM include [369, 644], which focus on inviscid Euler equa-
tions. Also addressing the issue of geometrical parametrization, the contributions [609] and
[637] deal with turbulent compressible NSEs and partial differential equation (PDE)-constrained
optimization problems, respectively. Extension of the FVM-based POD-Galerkin ROM for ther-
mal mixing problems is presented in [219]. Another POD-Galerkin ROM for the problem of
buoyancy-driven enclosed flows is developed in [552].
where Γ is the boundary of the fluid domain Ω ∈ Rd , with d = 1, 2, or 3. u is the flow velocity
vector field; t is the time; ν is the fluid kinematic viscosity; p is the normalized pressure field,
which is divided by the fluid density ρf ; and the time window under consideration is [0, T ].
We remark that in this work the parameter µ could be a physical parameter such as the fluid
kinematic viscosity or a geometrical one such as the dimension of a certain part of the domain.
We would like to emphasize that the velocity and pressure fields are functions of time, space, and
the parameter, i.e., u = u(t, x; µ), p = p(t, x; µ). These dependencies have been dropped in
the equations above for the sake of keeping the notation concise.
The incompressible NSEs in (7.1) are solved by the FVM [408]. The first step in this method
is to choose a suitable polygonal tessellation, and then the PDE system is written in integral
form over each control volume. Denote by Nh the dimension of the FOM, which is basically
the number of degrees of freedom of the discretized problem. The momentum and continuity
equations are solved with the help of a segregated approach which adapts the Rhie and Chow
7.2. Computational Fluid Dynamics—Laminar NSEs 145
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
uP
d
uN
uf Sf
interpolation [486]. The discretization process starts with the momentum equation by writing it
in integral form over each control volume Vi as follows:
∂
Z Z Z Z
T
udV + ∇ · (u ⊗ u)dV − ∇ · ν ∇u + (∇u) dV + ∇pdV = 0. (7.2)
Vi ∂t Vi Vi Vi
The discretization procedure of all terms in the momentum equation is explained in what follows.
The pressure gradient term is discretized using Gauss’s theorem:
Z Z X
∇pdV = pdS ≈ Sf pf , (7.3)
Vi Si f
where pf is the value of the pressure at the center of the faces and Sf is the area vector of each
face of the control volume.
The convective term is discretized by exploiting Gauss’s theorem as follows:
Z Z X X
∇ · (u ⊗ u)dV = (dS · (u ⊗ u)) ≈ Sf · uf ⊗ uf = Ff u f . (7.4)
Vi Si f f
In the above equation uf is the velocity vector evaluated at the center of each face of the control
volume. Note that the velocity values are initially computed at the cell centers and therefore the
values at the center of the faces have to be deduced from the ones calculated at the cell centers.
Consequently, these uf values are interpolated using the values computed at the cell centers.
Plenty of interpolation schemes can be used, such as central, upwind, second-order upwind, and
blended differencing schemes. The mass flux Ff is computed using the previous converged
velocity in the first iteration and then updated by Ff = uf · Sf to remove the nonlinearity.
As for the diffusion term, it is discretized as
Z Z X
T T
∇ · ν ∇u + (∇u) dV = dS · ν ∇u + (∇u) ≈ νf Sf · (∇u)f , (7.5)
Vi Si f
where (∇u)f is the gradient of u at the faces. In similar fashion to the computation of the
pressure gradient in (7.3), one may compute (∇u)f . As for the computation of the term Sf ·
(∇u)f in (7.5), this depends on one particular feature of the mesh: the orthogonality. The mesh
in Figure 7.1 is orthogonal if the line that connects two cell centers is orthogonal to the face that
divides these two cells. If the mesh is orthogonal, and assuming that the two cell centers are
equally distanced from the face, then the term Sf · (∇u)f is computed as
uN − uP
Sf · (∇u)f = |Sf | , (7.6)
|d|
146 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
where uN and uP are the velocities at the centers of two neighboring cells and d is the distance
vector connecting the two cell centers; see Figure 7.1. On the other hand, a nonorthogonal
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
correction term is needed for the case of nonorthogonal meshes. The correction term [309] is
computed by the equation
uN − uP
Sf · (∇u)f = |∆| + J · (∇u)f , (7.7)
|d|
where the relation Sf = ∆ + J holds. The first vector, ∆, is chosen parallel to Sf . The
term (∇u)f is obtained by interpolating the values of the gradient at the cell centers (∇u)N
and (∇u)P , in which the subscripts N and P indicate the values at the centers of the two
neighboring cells. In the next section, we address the discretized equations (written in matrix
form) for velocity and pressure. Additionally, we present the segregated pressure-based solver
approaches.
system.
The momentum equation in (7.8) can be written as
u + [A−1
u ][∇(·)][p] = 0, (7.9)
where A−1u denotes the inverse of the momentum matrix in the discretized form. By applying
the divergence operator to the last equation, one obtains
[∇ · (·)]u + [∇ · (·)][A−1
u ][∇(·)][p] = 0. (7.10)
Then, by exploiting the free divergence constraint on the velocity [∇ · (·)]u = 0, one may derive
the pressure equation
[∇ · (·)][A−1
u ][∇(·)][p] = 0. (7.11)
Thus the matrix A−1 u acts as the diffusivity in the Laplace equation for the pressure. For compu-
tational reasons, solving (7.11) in this form is inconvenient. This is because A−1u is likely to be
dense. In addition, the product of three matrices could result in a dense matrix, which increases
the difficulty of solving the linear system. As a result, a different approach is used in which the
momentum matrix is decomposed into diagonal and off-diagonal matrices:
where [Du ] is a diagonal matrix and therefore can be easily inverted and [LUu ] is the matrix
containing the off-diagonal part of [Au ]. Inserting (7.12) into (7.8) yields the modified system
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[Du ] [∇(·)] u −[LUu ][u]
= . (7.13)
∇ · (·) [0] p 0
The last equation is a Poisson equation for pressure with the diagonal part of the momentum
equation acting as a diffusivity, while the divergence of the velocity is on the right-hand side.
The derivation of the pressure equation in (7.14) is referred to as a Schur complement.
In the next part of this subsection, the same derivation will be done without having to use the
assembly of Schur’s complement. The first step is to write the discretized momentum equation
for each control volume, namely
X
au
P uP + au
N uN = rp − ∇p, (7.15)
N
where P represents a generic cell center and N is the set of neighboring points around it (7.1),
uN and uP are the velocities at the centers of two neighboring cells, au P is the vector of diagonal
coefficients of the equations, au
N is the vector that consists of off-diagonal coefficients, and rp
is a term that contains any contributions of the right-hand side of the momentum equation of the
NSEs. For the sake of simplicity, one may introduce the operator H(u) = rp − N au N uN ,
P
which contains the off-diagonal part of the momentum matrix and any right-hand side contribu-
tions. Thus,
au
P uP = H(u) − ∇p (7.16)
and
−1
uP = (au
P) (H(u) − ∇p). (7.17)
Inserting the last expression for uP into the continuity equation ∇ · u = 0 gives
−1 −1
∇ · [(au
P) ∇p] = ∇ · [(au
P) H(u)]. (7.18)
The last pressure equation is equivalent to the one in (7.14), where it can be seen that au P is
a coefficient in the diagonal matrix [Du ] and that H(u) is the product [LUu ][u]. Note that
one of the strongest assumptions in the SIMPLE algorithm is H(u) = 0. This means that this
algorithm is not consistent. If a more consistent solution is needed, the SIMPLEC algorithm is
to be preferred.
The mass flux through each face of the control volume is denoted by Ff . Equations (7.17)
and (7.18) are used together with the discretized version of the continuity equation to update the
mass fluxes Ff :
−1 −1
Ff = uf · Sf = −(au
P) Sf · ∇p + (au
P) Sf · H(u). (7.19)
|Sf |
(au
P)
−1
Sf · ∇p = (au
P)
−1
(pN − pP ) = apN (pN − pP ), (7.20)
|d|
|Sf |
where pP and pN are pressure values at the centers of two neighboring cells and apN = (au
P)
−1
|d|
is the off-diagonal matrix coefficient in the pressure equation.
148 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
At this point, the SIMPLE algorithm may be introduced. The first step in this algorithm is
called the momentum predictor step, where one solves the momentum equation (7.16) for an
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
initially guessed pressure field p? . Then, using the obtained value of the velocity u? from the
last step, one may assemble the new off-diagonal vector H(u? ). This is followed by solving the
Poisson equation for pressure (7.18) to correct the pressure field. The next step is to assemble the
conservative flux fields Ff before going through the cycle again and carrying out the momentum
predictor: fluxes are needed to update both au P and H(u). This procedure is repeated till con-
vergence is achieved. However, the pressure correction equation in its current form is prone to
divergence [596]. In order to ensure convergence, underrelaxation is employed.
The pressure field which results from solving the Poisson equation in (7.18) is p? + p0 , the
sum of the initial guess p? and a correction term denoted by p0 . The underrelaxed pressure field
is modified as follows:
p?? = p? + αp p0 , (7.21)
where αp is the pressure underrelaxation factor. If αp = 1, this means that no underrelaxation
is introduced and this choice of αp is too large for stable computations, particularly when p? is
far away from the final solution. A zero value of αp means that no correction is introduced at all
and the pressure is not updated, which is obviously unwanted. Therefore, the value of αp should
be taken in the range (0, 1) to allow one to add a correction term which is a fraction of p0 . This
term should be large enough to move the algorithm forward toward convergence and at the same
time small enough to ensure stable computations.
Unlike pressure, the underrelaxation for the velocity is not explicit; instead, the momentum
underrelaxation is implicit, where the discretized momentum equation in (7.16) is modified by
adding artificial terms as follows:
1 − αu u ?? ?? ? ? 1 − αu u ?
aP uP + au
P uP = H(u ) − ∇p + aP uP . (7.22)
αu αu
The new added terms consist of the new underrelaxed velocity values, denoted by u?? P , and the
old ones, denoted by u?P . These terms cancel each other in the case of convergence. Here αu is
the velocity underrelaxation factor, and the above equation simplifies to
1 u ?? 1 − αu u ?
a u = H(u? ) − ∇p? + aP uP . (7.23)
αu P P αu
As for the guidelines for choosing underrelaxation factors, there is no straightforward way to set
their optimum values, where in fact their values depend on the type of problem and the flow.
However, in the literature it is recommended that
0 < αu < 1,
0 < αp < 1,
αu + αp ≈ 1.
Let us summarize the important aspects of the SIMPLE algorithm. In the SIMPLE algorithm,
the role of pressure in the momentum equation is to ensure that the velocity field is divergence
free. After carrying out the momentum predictor step, the velocity field does not satisfy the
divergence constraint since the pressure field is a guessed one. Consequently, the pressure field
is corrected, which gives a pressure field which consists of two parts. The first one is physical
and thus consistent with the global flow field, while the other one represents the correction term,
which guarantees the continuity. Only the first component of the pressure field is desired and used
to build the physical pressure field. To build this physical pressure field, the SIMPLE algorithm
resorts to the underrelaxing approach explained in this subsection. The full SIMPLE algorithm
is outlined in Algorithm 7.1.
7.2. Computational Fluid Dynamics—Laminar NSEs 149
The SIMPLE algorithm is used in OpenFOAM to solve the velocity-pressure coupled system
for the case of steady flows. This algorithm may be extended to transient simulations; however,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the most used transient solver in CFD codes is PISO. In the remaining part of this section, the
PISO algorithm is addressed. The PISO algorithm is based on the idea that for low values of
the Courant number (small time steps), the pressure-velocity coupling is much stronger than
the nonlinear u-u coupling in the convective term. Thus it suggests that a repeated number of
pressure corrections can be carried out without having to rediscretize the momentum equation.
The PISO algorithm consists of two correction steps for the pressure field. The first yields a
conservative velocity field, while the second gives back a more physical pressure field. The use
of more than one pressure corrector implies that it is not necessary any more to underrelax the
pressure field. However, one thing that could be seen as a drawback of the PISO algorithm is
that it is assumed that the momentum discretization is not needed when the pressure correctors
are applied. This last assumption is only true for small time steps, which forces the maximum
Courant number [138, 139] to be under the value of Comax = 0.9.
1: Start with an initial guess of the pressure field p? and the velocity field u? .
2: Momentum predictor step: assemble au ? ?
P , H(u ), and ∇p , and solve the discretized momentum equation for the guessed
pressure field p? and velocity field u? :
u ? ? ?
aP uP = H(u ) − ∇p .
3: Calculate the off-diagonal vector H(u?? ) after obtaining the velocity field u?? from u?P .
4: Correct the pressure: the new pressure field is computed based on the obtained velocity field from the last step:
u −1 u −1 ??
∇ · [(aP ) ∇p̃] = ∇ · [(aP ) H(u )].
u −1 ??
ũP = (aP ) (H(u ) − ∇p̃).
5: Relax the pressure field and the velocity equation with the prescribed underrelaxation factors αp and αu , respectively. The under-
relaxed fields are called pur and uur .
6: Assemble the conservative face flux Ff :
u −1 ur p ur ur
Ff = uf · Sf = (aP ) Sf · H(u ) − aN (pN − pP ).
The PISO algorithm starts in a similar way to SIMPLE, using a guessed pressure field to carry
out the momentum predictor step. Then it assembles the H(u? ) vector and corrects the pressure
and velocities. The next step is the key difference between the two algorithms, where at this
point the velocity field has been corrected after solving the pressure equation. This means that
the H vector has been changed and so has the source term in the pressure equation, which makes
the pressure field no longer correct, so one has to update the H vector. To address this issue,
the SIMPLE algorithm goes all the way back to the momentum equation and performs another
momentum predictor, and thus it obtains a new velocity value, which will be used to update the
H vector to resolve the pressure equation. On the other hand, in the PISO loop the momentum
predictor is done just once. In more detail, rather than solving the momentum equation for the
second time, it uses the corrected velocity field to directly update the H vector and then performs
the second pressure correction step. The PISO complete procedure is structured in Algorithm 7.2.
Other transient solvers could be used to solve the coupled velocity-pressure system. We
mention here the PIMPLE algorithm, which merges PISO and SIMPLE. The PIMPLE algorithm
150 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
contains an outer loop which iterates over the PISO procedure, resulting in a more stable algo-
rithm. The number of times the outer loop is entered is known as the number of outer correctors
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
of the PIMPLE; if it is set to one then the PIMPLE replicates the PISO.
2: Solve the second pressure correction, and obtain the second pressure correction component p00 .
3: Correct the pressure and velocities; this step gives p??? and u??? :
??? ? 0 00
p =p +p +p .
where ur and pr are the reduced velocity and pressure fields, respectively, while φi (x) and
χi (x) are the reduced modes for velocity and pressure, respectively. These modes do not depend
on µ and t. The reduced order degrees of freedom are denoted by ai (t; µ) for both velocity and
pressure; the number of degrees of freedom is Nr . The temporal coefficients ai (t; µ) depend on
time t and on the parameter vector µ. Note that in (7.24) the velocity and pressure fields have
7.2. Computational Fluid Dynamics—Laminar NSEs 151
the same temporal coefficients. This last assumption makes the projection-based ROM simpler,
as mentioned earlier. However, this assumption has several limitations and drawbacks as may be
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
anticipated; in the next subsections these drawbacks will be addressed in greater detail.
The velocity POD modes are computed using the standard procedure. On the other hand,
the pressure modes in this formulation are computed using the eigenvectors and the eigenvalues
obtained in the SVD procedure for the velocity POD modes. Thus, the pressure modes are
expressed as
Ns
1 X
χi = S j V u. (7.25)
Ns λui j=1 p ij
The next step in building the POD-Galerkin ROM is to project the momentum equation of
(7.1) onto the velocity POD basis [φi ]N
i=1 as follows:
r
∂u
T
φi , + ∇ · (u ⊗ u) − ∇ · ν ∇u + (∇u) + ∇p = 0. (7.26)
∂t L2 (Ω)
The following reduced order dynamical system is obtained after inserting the decomposition
assumptions of (7.24):
ȧ = ν(B + BT )a − aT Ca − Ha, (7.27)
where a is the vector of reduced order degrees of freedom, and each of B, BT , C, and H is
either a reduced order matrix or a tensor. These terms are computed as follows:
(B)ij = (φi , ∇ · ∇φj )L2 (Ω) ,
(BT )ij = φi , ∇ · (∇φTj ) L2 (Ω) , (7.28)
The POD-Galerkin ROM presented in the last subsection has shown a lack of accuracy in
reconstructing the pressure field at the reduced order level. An accurate pressure approximation
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
is imperative since many outputs of interest depend greatly on the pressure field. Consequently,
stabilization methods have been considered in order to accurately reproduce the pressure field.
In this work, two stabilization techniques are addressed. The first exploits the Poisson equation
for pressure (PPE) at the reduced order level, which makes it possible to separate the pressure
reduced degrees of freedom from those of the velocity. The PPE can be obtained by taking the
divergence of the momentum equation in (7.1) and then taking advantage of the fact that the
divergence of the velocity field is null. The resulting system is
∂u h
T
i
+ ∇ · (u ⊗ u) = ∇ · −pI + ν ∇u + (∇u) in Ω × [0, T ], (7.31a)
∂t
∆p = −∇ · (∇ · (u ⊗ u)) in Ω. (7.31b)
We remark that the ROM developed in this section cannot be used to reduce steady flows [313,
376]. In fact, among the possible choices for the pressure boundary condition required to render
the PPE formulation of the NSEs equivalent to the original one, we selected one which ensures
that the velocity is divergence free only in the unsteady setting (see the remark in Section 2 in
[313]).
The reduced order decomposition assumptions in this case are
PNu
u(x, t; µ) ≈ ur (x, t; µ) = i=1 ai (t; µ)φi (x),
PNp (7.32)
p(x, t; µ) ≈ pr (x, t; µ) = i=1 bi (t; µ)χi (x),
where one can see that new temporal coefficients denoted by bi (t; µ) have been introduced to ap-
proximate the reduced pressure field. It is worth mentioning that the way the pressure modes are
computed now is different from in the previous section. Here the SVD of the pressure correlation
matrix C p ∈ RNs ×Ns has to be done. The entries of the matrix C p are
(C p )ij = Spi , Spj 2 , (7.33)
L (Ω)
where λp is the matrix containing in its diagonal the eigenvalues of the pressure correlation
matrix C p and V p is the corresponding matrix of eigenvectors. At this point one may perform
Galerkin projection of the equations. The momentum equation is projected onto the velocity
POD modes, while the PPE is projected onto the pressure POD space. These projections yield
i
∂u h
T
φi , + ∇ · (u ⊗ u) − ∇ · −pI + ν ∇u + (∇u) = 0, (7.35a)
∂t L2 (Ω)
ȧ = ν(B + BT )a − aT Ca − Hb,
(7.36)
Db + aT Ga − νN a − L = 0,
7.2. Computational Fluid Dynamics—Laminar NSEs 153
(D)ij = (∇χi , ∇χj )L2 (Ω) , (G)ijk = (∇χi , ∇ · (φj ⊗ φk ))L2 (Ω) , (7.37a)
1: Start with an initial guess of the reduced pressure field b? and the reduced velocity field a? .
2: Reconstruct both pressure and velocity fields p? and u? .
3: Momentum predictor step: assemble the momentum equation using those fields, project it using the velocity modes φi , and solve it
to obtain a new reduced velocity a?? :
u ? ? ?
(φi , aP uP − H(u ) + ∇p )L2 (Ω) = 0.
4: Reconstruct the new full velocity u?? and calculate the off-diagonal vector H(u?? ).
5: Correct the pressure: the new reduced pressure b?? field is computed after having projected the pressure equation using the pressure
modes:
u −1 u −1 ??
(χi , ∇ · [(aP ) ∇p̃] − ∇ · [(aP ) H(u )])L2 (Ω) = 0.
??
Then correct the velocity explicitly after having reconstructed the new pressure p .
6: Relax the pressure field and the velocity equation with the prescribed underrelaxation factors αp and αu , respectively. The under-
ur ur
relaxed fields are called p and u .
7: Assemble the conservative face flux Ff :
u −1 ur p ur ur
Ff = uf · Sf = (aP ) Sf · H(u ) − aN (pN − pP ).
Convergence can be checked in different ways. The simplest is to check over the residual;
however, its order of magnitude could be very far from the full-order one and, thus, a threshold
may be difficult to set. On the contrary, it is possible to set the jump of the residual by comparing
two consecutive steps of the algorithm, but the possibility of being locked near a local minima
could become high. The best solution is to use the two aforementioned choices together so that
one achieves a convergence criterion which is as safe as possible.
Of course the reduced algorithm inherits all the instability issues affecting the full-order one.
For this reason the initial guesses a? and b? have to be properly chosen and preferably not set to
zero.
154 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
To simulate physical fluid dynamics problems, we must often impose nonhomogeneous Dirich-
let boundary conditions on certain parts of the boundary. This might be the case in inlet-outlet
problems, where it is natural to have nonhomogeneous Dirichlet velocity conditions at the in-
let. In the context of parametric PDEs, it may very well be the case that this nonhomogeneous
boundary velocity vector is the parameter under study. These aspects make the treatment of the
nonhomogeneous Dirichlet boundary conditions essential for building an accurate ROM.
The methods employed to enforce nonhomogeneous boundary Dirichlet condition include
the penalty method [63, 23, 42, 319, 539] and the lifting function method [240, 251, 285].
Before entering into the details of both methods, we introduce the following notation: let
ΓD ⊂ Γ be the Dirichlet boundary that might be composed by separate boundaries, i.e., ΓD =
ΓD 1 ∪ΓD 2 . . .∪ΓD K . Let NBC be the number of velocity boundary conditions we would like to
impose on some parts of the Dirichlet boundary. It is important to clarify that each nonzero scalar
component value of the velocity field that has to be set at one part of the boundary is counted as
one boundary condition. As an example, in a two-dimensional problem, let UΓD 1 = (Ux1 , 0) and
UΓD 2 = (Ux2 , 0) be the velocity vectors that must be imposed at the Dirichlet boundaries ΓD 1
and ΓD 2 , respectively. In this case there are two nonhomogeneous Dirichlet boundary conditions
to set and thus NBC = 2. Let UBC,i,j be the nonzero value of the ith component of the velocity
vector to be imposed at the reduced order level at the jth part of the Dirichlet boundary ΓD j .
Denote by UBC the vector of all scalar velocities UBC,i,j —this vector has a dimension of NBC ,
and its kth element is called UBC k .
ũ = u − UBC · φL , (7.38)
h
where φL ∈ RNBC ×Nu is a matrix of the lifting functions φL i,j . Each lifting function φL i,j
has homogeneous Dirichlet boundary conditions in all parts of the Dirichlet boundary except in
the ith component at ΓD j , where it has unitary value. At this point, one can apply the POD
method on the snapshot matrix
h
Sũ = {ũ(x, t1 ; µ1 ), . . . , ũ(x, tNT ; µM )} ∈ RNu ×Ns . (7.39)
In the online stage, the given boundary velocity vector is called UBC
∗
, which may contain values
different from the ones in the original velocity snapshots. We want to approximate the reduced
order velocity field corresponding to UBC
∗
. The ROM velocity field is approximated as
Nu
X
∗ ∗
u(x, ·; UBC ) ≈ UBC · φL + ai (·)φi (x). (7.40)
i=1
It can be seen above that the boundary velocity vector UBC is assumed to be the parameter
under question. However, in the presence of nonhomogeneous boundary condition(s) for the
velocity field at a part of the boundary, the same described boundary treatment procedure has to
7.2. Computational Fluid Dynamics—Laminar NSEs 155
• when building a ROM for the reproduction and extrapolation in time without parameters
(nonparametrized ROM);
• with the parametrized case where the boundary velocity vector is one of the parameters;
• with the parametrized case where the boundary velocity vector is not a parameter.
The way to choose a suitable lifting function is problem dependent. To reduce unsteady
nonparametrized cases, where reduction aims to reproduce time snapshots and potentially ex-
trapolate in time, a possible choice for the lifting function is the average of the offline velocity
snapshots. A more general approach for generating appropriate lifting functions is to solve linear
potential flow problems with a unitary boundary condition for each nonhomogeneous boundary
condition to be set. These problems are steady ones; therefore, an iterative procedure with a ten-
tative velocity field is carried out till convergence. While solving each of these potential flows,
the value of the initial velocity field at the Dirichlet boundary has to be zero everywhere except
for one scalar entity where the lifting function is sought. The converged velocity field will be
considered as the lifting function corresponding to the nonhomogeneous boundary condition at
the aforementioned entity. Besides the requirement of having unitary boundary condition, the
lifting functions have to be divergence-free fields.
The lifting/control function method can be adapted to work with both the PPE-ROM and the
SUP-ROM. On the other hand, adapting this method with the ROM mentioned in Subsection
7.2.3 (U-ROM) is not straightforward. This is because the reduced order degrees of freedom of
the velocity are the same as those of the pressure. Therefore, the homogenization of the veloc-
ity snapshots in (7.38) has to be accompanied with a procedure that obtains the corresponding
pressure snapshots. This procedure is not easily defined.
As a consequence of the additional lifting function mode/modes, the dimension of the ve-
locity POD space VuP OD will be Nu + NBC , or Nu + NS + NBC for the PPE-ROMs, when a
supremizer stabilization is absent or needed, respectively.
The nonhomogeneous boundary conditions in the penalty method are enforced by presenting a
constraint in the reduced order dynamical system. This is done by adding a term in the reduced
momentum equation which has zero value everywhere except on the Dirichlet boundary. This
method was initially implemented in the context of the FEM in [23, 42]. In [240], the authors
utilized the penalty method to enforce nonhomogeneous time-dependent Dirichlet boundary con-
ditions for the case of the flow around a circular cylinder at Reynolds number 100. The POD-
Galerkin ROM in [539] introduced the penalty method in order to account for the time-dependent
Dirichlet boundary conditions—the authors presented a study of the accuracy and the stability of
the method depending on the penalty parameter values. The penalty method has also been used
in other ROMs, e.g., [318, 319, 307, 380, 548].
The method can be employed for all the ROMs mentioned in the previous sections. For
example, when a supremizer stabilization is employed for the SUP-ROM (see Subsection 1.2.4
for more details on the supremizer stabilization strategy), the architecture is modified as follows:
N
!
X BC
T
M ȧ = ν(B + BT )a − a Ca − Hb + τ k k
(UBC k D − E a) , (7.41a)
k=1
P a = 0, (7.41b)
156 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
where τ is a penalization factor whose value is set by sensitivity analysis or by automatic tun-
ing, as recently presented in [554], where an iterative penalty method is presented. Generally
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
speaking, having a higher value of τ leads to a stronger enforcement of the boundary conditions
but might ill-condition the dynamical system. The newly introduced boundary matrices E k and
vectors D k are defined as
As mentioned earlier, having full decoupling between the offline and the online stages is a vital
feature of efficient ROM. One may approximate the reduced order forces Fr by simply comput-
ing the following integral after reconstructing the reduced fields:
Z
Fr = (2µ∇ur − pr I)nds, (7.44)
∂Ωf
but this would require accessing the original mesh, which means that the computational cost of
carrying out this integral will depend on Nh (the number of degrees of freedom of the FOM).
Therefore, an alternative approach is needed to efficiently reconstruct the fluid dynamics forces.
The approach used throughout this work involves inserting the reduced order approximations
in (7.43). These approximations could be the uniform ones in (7.24) or the nonuniform ones in
(7.32). Assuming the latter approximations, the forces are computed as follows:
! Np
Z Nu
X X
Fr = 2µ∇ ai (t; µ)φi (x) − bi (t; µ)χi I nds, (7.45)
∂Ωf i=1 i=1
Z Nu Z Np
X X
Fr = 2µ ai (t; µ)∇φi (x)nds − bi (t; µ)χi nds, (7.46)
∂Ωf i=1 ∂Ωf i=1
Nu Z Np Z
X X
Fr = ai (t; µ) 2µ∇φi (x)nds − bi (t; µ) χi nds. (7.47)
i=1 ∂Ωf i=1 ∂Ωf
Let Z
δi = 2µ∇φi (x)nds for i = 1, . . . , Nu , (7.48a)
∂Ωf
Z
θj = χj nds for j = 1, . . . , Np , (7.48b)
∂Ωf
7.3. Numerical Experiments 157
where each of ∇φi (x) and χj can be viewed as a velocity and a pressure field, respectively. The
terms (7.48a) and (7.48b) can be precomputed during the offline stage and then stored.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In the online stage, when new time and parameter values of t∗ and µ∗ , respectively, are
introduced, the full-order forces are computed as
Z
∗ ∗
F (t ; µ ) = (2µ∇u(x, t∗ ; µ∗ ) − p(x, t∗ ; µ∗ )I)nds, (7.49)
∂Ωf
Nu Np
X X
Fr (t∗ ; µ∗ ) = ai (t∗ ; µ∗ )δi − bj (t∗ ; µ∗ )θj . (7.50)
i=1 j=1
It is important to underline the fact that the above formulas for the reduced forces are valid only
in the case of affine parameter dependency.
ΓD Γ0
u u = (1, 0) u = (0, 0)
p ∇p · n = 0 ∇p · n = 0
Figure 7.2. Sketch of the mesh for the lid-driven cavity problem together with the boundary
subdivisions and boundary conditions.
158 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
Table 7.1. The table contains the cumulative eigenvalues for the lid-driven cavity test. The first,
second, and third columns report the cumulative eigenvalues for the velocity, pressure, and supremizer
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
fields, respectively. The last column contains the value of the inf-sup constant, in the supremizer stabi-
lization case, for different numbers of supremizer modes and with a fixed number of velocity and pressure
modes (10 modes for velocity and 10 modes for pressure).
N Modes u p s β
1 0.978946 0.975406 0.980260 9.264e-05
2 0.994184 0.991528 0.995232 9.264e-05
3 0.997737 0.995385 0.997912 7.175e-04
4 0.998990 0.998116 0.999400 7.175e-04
5 0.999483 0.999270 0.999844 7.175e-04
10 0.999971 0.999971 0.999997 1.551e-02
is equal to ν = 1 × 10−4 m2 /s, which leads to a Reynolds number of 1000. For the full-
order simulation, the time discretization is treated using a second-order backward differencing
scheme, while the discretization in space is performed with a fourth-order interpolation scheme.
The time step is kept constant and equal to ∆t = 5 × 10−4 , and the simulation is run till
T = 10 s. The snapshots are acquired every 0.01 s, giving a total number of snapshots equal
to 1000. For the ROM, the dimensions of the reduced spaces for velocity and pressure are set,
for both the presented methodologies, equal to Nur = 10 and Npr = 10. In the SUP-ROM
the reduced space for velocity is enriched with 10 additional supremizer modes. This selection
is done according to Table 7.1, where it is possible to observe that this number of modes is
sufficient to retain more than 99.9% of the energy for both velocity and pressure. We remark
that in this numerical experiment any kind of parametrization can be introduced. The ROM
is in fact used to simulate the same conditions tested in the full-order setting, and the results
are compared to the full-order simulation results. The time discretization, at the reduced order
level, is treated by using a first-order backward Newton method. Figure 7.8 depicts a comparison
between the high-fidelity simulation and the ROM one, for both velocity and pressure fields, at
different time instants. As one can see from the figure, both models are capable of reproducing
the main flow pattern for both fields. Figure 7.3 reports the evolution in time of the L2 relative
error for velocity and pressure, respectively. Figure 7.4 shows a qualitative comparison of the
velocity and pressure fields for the high-fidelity, the SUP-ROM and the PPE-ROM. The plots
also report the error without any type of stabilization. It is clear that, without stabilization,
105
104
100
||uH F − uROM ||/ ||uH F ||
103
100
10−1
10−2
0 2 4 6 8 10 0 2 4 6 8 10
Time(s) Time(s)
Figure 7.3. Error analysis for the velocity field. The L2 -norm of the relative error is plotted
over time for three different models: with supremizer stabilization (USUP, continuous red line), with PPE
stabilization (UPPE, dotted blue line), and without stabilization (PNOS, dashed green line). The ROMs are
obtained with 10 modes for velocity, pressure, and supremizer.
7.3. Numerical Experiments 159
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 7.4. Comparison of the velocity and pressure fields for high-fidelity (UHF, column 1;
PHF, column 4), SUP-ROM (USUP, column 2; PSUP, column 5), and PPE-ROM (UPPE, column 3; PPEE,
column 6). The fields are depicted for different time instants equal to t = 0.2 s, 0.5 s, 1 s, and 5 s,
respectively, and increasing in the image from top to bottom. The ROM models are obtained with 10 modes
for velocity and pressure and for the SUP-ROM only with 10 additional supremizer modes. The velocity
and pressure magnitudes are shown in the image legends.
even though the ROM does not diverge, both the velocity and pressure fields are completely
unreliable. For this particular numerical test, the SUP-ROM produces, with respect to the PPE-
ROM, worse results for the velocity field but better results for the pressure field. This difference
can be justified by the fact that, within a supremizer stabilization technique, the POD velocity
space is enriched by unnecessary (for the correct reproduction of the velocity field) supremizer
modes. During the initial transient, both fields present a higher relative error due to the relatively
low number of snapshots acquired during the initial transient. The snapshots, as highlighted
above, are equally distributed in time, and, to enhance the performance of the ROM, one should
concentrate the snapshots in the timespan where the system exhibits the most nonlinear behavior.
For the SUP-ROM, according to [34], the number of supremizer modes is chosen equal to the
number of pressure modes. Table 7.1 also reports the value of the inf-sup constant β, obtained
by keeping the number of velocity and pressure modes (10 modes for velocity and 10 modes for
pressure) constant and varying the number of supremizer modes. As one can observe from the
table, increasing the number of supremizer modes leads to a remarkable increase in the inf-sup
constant.
160 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
ΓIn Γ0 Γs ΓOut
u u = (1, 0) u = (0, 0) u·n=0 ∇u · n = 0
p ∇p · n = 0 ∇p · n = 0 ∇p · n = 0 p=0
Figure 7.5. The figure shows a general overview of the mesh with dimension and boundaries
(upper left), a table with the imposed values at the boundaries (bottom), and a zoom of the mesh near the
cylinder (upper right).
Table 7.2. The table contains the cumulative eigenvalues for the cylinder problem. In the first,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
second, and third columns are reported the cumulative eigenvalues for the velocity, pressure, and supremizer
fields, respectively, as a function of the number of modes. In the last column is reported the value of the
inf-sup constant for the supremizer stabilization case for different numbers of supremizer modes with a fixed
number of velocity and pressure modes (15 modes for velocity and 10 modes for pressure).
N Modes u p s β
1 0.390813 0.793239 0.921046 2.608e−04
2 0.598176 0.85809 0.941746 4.492e−04
3 0.802176 0.911636 0.961438 7.869e−03
4 0.879096 0.934997 0.978072 1.662e−02
5 0.949519 0.955578 0.98669 1.662e−02
10 0.986025 0.992347 0.998307 1.098e−01
15 0.995922 0.997994 0.999732 1.199e−01
010
USUP ν = 0 005
UPPE ν = 0 005
kinetic energy relative error
USUP ν = 0 005625
005 UPPE ν = 0 005625
000
−005
−010
0 10 20 30 40 50 60 70 80
Time(s)
Figure 7.6. Kinetic energy relative error in the cylinder example for ν = 0.005 and ν =
0.005625. The kinetic energy relative error is plotted over time for the two values of viscosity and for
the two different models: with supremizer stabilization (USUP and PSUP) and PPE stabilization (UPPE
and PPPE). The ROM solutions are obtained with 15 modes for velocity, 10 modes for pressure, and 12
modes for supremizers. The time window in this case is wider than the one used to generate the RB spaces
(∆T = 10 s).
Figure 7.7. First four basis functions for velocity (first row), pressure (second row), and suprem-
izers (third row).
162 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 7.8. Comparison of the velocity field for high-fidelity (UHF, first row), supremizer stabi-
lized ROM (USUP, second row), and PPE stabilized ROM (UPPE, third row). The fields are depicted for
different time instants equal to t = 195 s, 200 s, 230 s, and 270 s and increasing from left to right. The
ROM solutions are obtained with 15 modes for velocity and 10 modes for pressure, and for the SUP-ROM
only with 12 additional supremizer modes. The velocity magnitude is shown in the image legends.
equal dimensions for the pressure and the supremizer spaces do not automatically guarantee the
fulfillment of the inf-sup condition.
To test the ROMs, the results are compared against the full-order results for an intermediate
value of the viscosity, ν = 0.005625, which is not included in the values of viscosity (ν =
[0.005, 0.00625, 0.0075, 0.00875, 0.01]) employed to generate the snapshots used to create the
RB spaces. The results were compared on two different time windows. The first one covers 10 s
of simulation and coincides with the time window used to generate the snapshots; the second one
covers 80 s of simulation and therefore is much wider than the time window used to generate
the snapshots. Figures 7.8 and 7.9 show the comparison, for velocity and pressure, respectively,
between the results obtained with the full-order model, the SUP-ROM, and the PPE-ROM. The
fields are depicted at four different time instants equal to t = 195 s, 200 s, 230 s, and 270 s.
The first two time instants are, respectively, in the middle and at the end of the time window
used to generate the snapshots, while the other two time instants are outside of it. Figure 7.10
Figure 7.9. Comparison of the pressure field for high-fidelity (PHF, first row), supremizer stabi-
lized ROM (PSUP, second row), and PPE stabilized ROM (PPPE, third row). The fields are depicted for
different time instants equal to t = 195 s, 200 s, 230 s, and 270 s and increasing from left to right. The
ROM solutions are obtained with 15 modes for velocity and 10 modes for pressure, and for the SUP-ROM
only with 12 additional supremizer modes. The pressure magnitude is shown in the image legends.
7.3. Numerical Experiments 163
0.040 0.20
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
USUP PSUP
0.035 UPPE PPPE
uROM ||/ ||uH F ||
0.030 0.15
0.020 0.10
0.015
||pH F
||uH F
0.010 0.05
0.005
0.000 0.00
0 2 4 6 8 10 0 2 4 6 8 10
Time(s) Time(s)
Figure 7.10. Error analysis for the velocity (left plot) and pressure (right plot) fields in the cylin-
der example with ν = 0.005625. The L2 -norm of the relative error is plotted over time for the two different
models: with supremizer stabilization (USUP and PSUP, continuous red line) and PPE stabilization (UPPE
and PPPE, dot-dashed blue line). The ROM solutions are obtained with 15 modes for velocity, 10 modes
for pressure, and 12 modes for supremizers.
reports the L2 -norm of the relative error for velocity and pressure in the 10-s-wide time window
for ν = 0.005625. The figure also confirms the behavior observed in the cavity example: the
SUP-ROM produces worse results for the velocity field but better results for the pressure field.
Figure 7.11 shows the same plots in a wider time window, and also for one of the values of
viscosity (ν = 0.005) used to generate the full-order snapshots. For both ROMs, the relative
error is increasing in time. Cross-referencing the data of Figure 7.11 with the plots of Figures 7.8
and 7.9, one can deduce that the increase in the error is given, for the SUP-ROM, by the numerical
instabilities that occur for long time integrations. In the last time step (t = 270 s), it is in fact
possible to observe, for both velocity and pressure, a completely incorrect and nonphysical flow
pattern. For the PPE-ROM, instead, the flow pattern still looks regular and sufficiently similar to
1.0 3.0
USUP ν = 0.005 PSUP ν = 0.005
UPPE ν = 0.005 PPPE ν = 0.005
USUP ν = 0.005625 2.5 PSUP ν = 0.005625
0.8
uROM ||/ ||uH F ||
2.0
0.6
1.5
0.4
1.0
||pH F
||uH F
0.2
0.5
0.0 0.0
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
Time(s) Time(s)
Figure 7.11. Error analysis for the velocity (left plot) and pressure (right plot) fields in the
cylinder example with ν = 0.005 and ν = 0.005625. The L2 -norm of the relative error is plotted over
time for the two values of viscosity and for the two different models: with supremizer stabilization (USUP
and PSUP) and PPE stabilization (UPPE and PPPE). The ROMs are obtained with 15 modes for velocity,
10 modes for pressure, and 12 modes for supremizers. The time window is in this case wider than the one
used to generate the RB spaces (∆T = 10 s).
164 Chapter 7. Finite Volume–Based Reduced Order Models for Laminar Flows
the high-fidelity one but it is possible to observe a phase shift between the high-fidelity and the
ROM solution. The PPE-ROM, in fact, even though it continues to produce a regular and physical
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
pattern, has a period of vortex shedding which is slightly longer than in the HF solution. To show
a better idea about the behavior of the different ROMs for long time integrations, Figure 7.6
depicts the relative error of the total kinetic energy. It is well known that POD-Galerkin models
are affected, in fact, by a blow-up energy issue [137, 196]. It is not the objective of this work
to deal with long time integration instabilities, but it is worth checking which kind of pressure
stabilization method is likely prone to this issue. From the figure it is clear that the PPE-ROM
accurately preserves the total kinetic energy of the system. On the other hand, the SUP-ROM
exhibits an oscillating behavior with an increase in the total kinetic energy.
Table 7.3. The table contains the computational time for the supremizer (SUP) and the PPE
stabilization techniques. In the cavity experiment, the SUP-ROM is obtained with 10 modes for velocity,
pressure, and supremizers, while the PPE-ROM is obtained with 10 modes for pressure and supremizers.
In the cylinder experiment, the SUP-ROM is obtained with 15 modes for velocity, 10 for pressure, and 12
for supremizers, while the PPE-ROM is obtained with 15 modes for velocity and 10 for pressure.
HF SUP-ROM PPE-ROM
Cavity Exp. 25min 7.64s 4.86s
Cylinder Exp. 18.5min × 6proc. 3.14s 0.971s
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 8
Finite Volume–Based
Reduced Order Models
for Turbulent Flows
8.1 Introduction
This chapter focuses on turbulence modeling for incompressible flows at both the full- and the
reduced order level.
In the previous chapter, we dealt with the full-order incompressible Navier–Stokes equations
(NSEs) for both steady and unsteady flows within a laminar regime by referring to the solution
algorithms available in OpenFOAM. Here, we will follow an analogous work flow for turbulence
modeling. Several models are implemented in OpenFOAM. In this work we mainly rely on two
of them: the Reynolds-averaged Navier–Stokes (RANS) equations and the large eddy simulation
(LES).
On the other hand, at the reduced order level, we mainly propose a hybrid projection/data-
driven strategy: a classical proper orthogonal decomposition (POD)-Galerkin projection ap-
proach for the reconstruction of the velocity and the pressure fields and a data-driven reduc-
tion method based on radial basis functions (RBFs) to approximate the turbulent eddy viscosity.
However, for the sake of completeness, we also review other reduced order modeling (ROM)
approaches.
165
166 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
expressed as the sum of mean and fluctuating parts. In the majority of cases, the fluctuating
component appears as a vibration around an equilibrium, or average flow solution. In a common
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
case, one is interested in solving only for the time-averaged part of the fluid dynamic variables.
To achieve this, following the procedure introduced by Reynolds, after decomposing all the vari-
ables, the momentum and continuity equations are time averaged. So one obtains a new system
in which the unknowns include the mean components of all the flow fields. However, the pres-
ence of the nonlinearity associated with the convective term leads to residual terms in which the
fluctuations are still present in the new system, where the number of equations is not equal to the
number of unknowns. This is called the closure problem.
If we consider a generic fluid dynamics scalar field called σ(x, t), then the Reynolds average
operation for σ(x, t) gives
σ(x, t) = σ(x, t) + σ 0 (x, t), (8.1)
where σ(x, t) and σ 0 (x, t) are the mean and the fluctuating parts, respectively. For steady flows,
the mean part σ is a spatial field without dependence on time, i.e., σ = σ(x). In this case, the
mean field is computed as
t+T
1
Z
σ(x) = lim σ(x, τ )dτ. (8.2)
T →∞ T t
As for the unsteady case, let T1 and T2 be the time scales of the fluctuating and the mean fields,
respectively. Time averaging is computed by the integral
t+T
1
Z
σ(x, t) = σ(x, τ )dτ, T1 ≤ T ≤ T2 . (8.3)
T t
Hereinafter, we will use the overbar notation to indicate any time-averaged quantity. We
recall that the time average of the mean field σ(x, t) is the mean field itself, while the time
average of the fluctuating part σ 0 (x, t) is zero.
We consider the time average of the product of two scalar quantities named φ and ψ:
where one exploits the fact that the product of a mean quantity and a fluctuating quantity has zero
mean. On the other hand, the quantity φ0 ψ 0 is not zero, which basically means that the product
of the means φψ is not equal to the mean of the products φψ. In fact the two quantities φ and ψ
are called uncorrelated if φ0 ψ 0 = 0; otherwise they are correlated.
Now we are ready to derive the RANS system. We start from the NSEs (7.1) written in scalar
form:
∂ui
= 0, (8.5)
∂xi
∂ui ∂ui ∂p ∂Sji
+ uj =− + 2ν , (8.6)
∂t ∂xj ∂xi ∂xj
where the Einstein summation convention has been used and where S is the strain-rate tensor,
given by
1 ∂ui ∂uj
Sij = + . (8.7)
2 ∂xj ∂xi
The convective term can be written as
∂ui ∂ (ui uj ) ∂uj ∂ (ui uj )
uj = − ui = , (8.8)
∂xj ∂xj ∂xj ∂xj
8.2. RANS Equations 167
where (8.5) has been employed. By using (8.8) we can recast the momentum equation (8.6) as
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
The application of Reynolds time averaging to (8.5) and (8.9) leads to the RANS equations,
namely
∂ui
= 0, (8.10)
∂xi
∂ S − u0 u0
∂ui ∂ui ∂p ji i j
+ uj =− + 2ν . (8.11)
∂t ∂xj ∂xi ∂xj
We point out that the averaged continuity equation is identical to the original one, which also
means that the fluctuating velocity component u0i has zero divergence. As for the momentum
equation, not all the fluctuating terms appearing in it have vanished. In fact, the term u0i u0j still
appears in the momentum equation because of the nonlinearity of the convective term. The term
u0i u0j is often referred to as the Reynolds stress tensor R, with Rij = u0i u0j . The Reynolds stress
tensor R is a symmetric tensor with six unknown components. Therefore, in the RANS system
(8.11) and (8.10) there are six additional unknowns with respect to the Navier–Stokes system,
without any additional equation. The bottom line in turbulence modeling is to deal with the
Reynolds stress tensor in order to close the system so as to compute the mean flow fields. Such
an issue in turbulence modeling is the aforementioned closure problem.
8.2.2 EVMs
Over the years the closure problem has been addressed by several methodologies characterized
by different computational costs and accuracy of the results. Among others, we refer to the meth-
ods based on the derivation of a transport equation for the Reynolds stress tensor components.
Such closure models are obtained by multiplying the NSEs by a fluctuating quantity and then
taking the mean. However, this procedure will result in additional nonlinear terms containing
fluctuation product averages, which furthermore increase the number of unknowns. Specifically,
these additional unknowns include a correlation term involving fluctuating velocity components
coming from the convective term, and also other correlation terms from the diffusive contribution
and a term which amounts to a fluctuating velocity-pressure correlation. Overall, the procedure
brings in 22 new unknowns. In order to avoid such complexity, several algorithms in RANS tur-
bulence modeling are based on the Boussinesq eddy-viscosity assumption [71]. The Boussinesq
assumption (or hypothesis) states that the Reynolds stress tensor is proportional to the traceless
strain rate tensor S, namely
2
Rij = 2νt Sij − kδij , (8.12)
3
where k is the turbulent kinetic energy per unit mass and νt is the so-called artificial or eddy vis-
cosity. The turbulence models which employ the Boussinesq eddy viscosity assumption are often
called eddy viscosity models (EVMs). In these models, the eddy viscosity is approximated by
making use of either algebraic relations or transport-diffusion-reaction partial differential equa-
tions (PDEs) for other quantities which have an algebraic relationship with the eddy viscosity.
The first group of EVMs is the algebraic models or the zero-equation models. As an example we
mention the mixing length turbulence model [462]. The first example of a PDE-based model is
the Spalart–Allmaras model [546]. In this model one solves for a viscosity-like variable called
ν̃. Richer EVMs are based on solving two additional PDEs for specific turbulent flow variables,
168 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
among which we mention the k- and the k-ω models [314, 347]. In these models and ω rep-
resent the specific turbulent dissipation rate of the turbulent kinetic energy into internal thermal
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
energy. The turbulent kinetic energy from the turbulent velocity fluctuations u0i is given by
1 0 0 1 2 2 2
k= u u = (u01 + u02 + u03 ). (8.13)
2 i i 2
As for , we have
∂u0i ∂u0i
=ν , (8.14)
∂xk ∂xk
and for ω we get
ω= , (8.15)
kβ ∗
where β ∗ is a constant equal to 0.09.
The first form of a two-equation turbulence model was presented by Kolmogorov in [347] as
a k-ω model. Later on, other forms of the k-ω model were proposed with significant improve-
ments, such as in [514, 617, 618]. In [361] the standard k- model is presented.
The RANS equations have been utilized extensively in the literature for modeling turbulence
in different application fields. In [207] one may find an application of the RANS in aerodynam-
ics, in particular for modeling a high-speed aerodynamic flow transition. In [331], numerical
simulations for steady turbulent problems in automotive engineering using the Spalart–Allmaras
model are performed. The RANS equations have also been used to study wind dynamics in
civil engineering: in [405] the authors present a three-dimensional study involving the use of
steady RANS to predict mean wind pressure distributions on windward and leeward surfaces of
a medium-rise building with and without balconies. In [200], the k- model is used to describe
the hydrodynamic flow around a boat. The work [141] presents an application of the RANS
equations for the design of sailing yachts.
Another two-equation model called the SST k-ω is presented in [395]: this turbulence model
merges the k-ω model and the k- model by combining their advances. The SST k-ω is used
extensively in the computational fluid dynamics (CFD) community because it is a good compro-
mise for flows with detachment and recirculation. This model has also been used in this work,
so we have chosen to report its complete mathematical formulation.
The turbulent kinetic energy k equation is given by
∂k ∂k ∂ ∂k
+ uj = P − β ∗ ωk + (ν + σk νt ) , (8.16)
∂t ∂xj ∂xj ∂xj
As for the terms and the constants which appear in the two equations above, they are defined as
follows:
∂ui 1 ∂ui ∂uj
, 10β ∗ ωk , Wij =
p
P = min Rij − , Ωs = 2Wij Wij ,
∂xj 2 ∂xj ∂xi
a1 k
νt = , φ = F1 φ1 + (1 − F1 )φ2 ,
max(a1 ω, Ωs F2 )
" √ ! #
4
k 500ν 4σω2 k
F1 = tanh arg1 , arg1 = min max , , ,
β ∗ ωy y 2 ω CDkω y 2
8.3. LES 169
1 ∂k ∂ω −10
CDkω = max 2ρσω2 , 10 ,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
ω ∂xj ∂xj
√ !
2
k 500ν
F2 = tanh arg2 , arg2 = max 2 ∗ , 2 ,
β ωy y ω
σk1 = 0.85, σw1 = 0.65, β1 = 0.075,
σk2 = 1.00, σw2 = 0.856, β2 = 0.0828,
β ∗ = 0.09, a1 = 0.31.
Hence the RANS equations coupled with the SST k-ω turbulence model can be written as
h i
T
∂u
∂t + ∇ · (u ⊗ u) = ∇ · −pI + (ν + νt ) ∇u + (∇u) in Ω × [0, T ],
∇ · = 0 in Ω × [0, T ],
u
νt = max(aa11ω,Ω k
in Ω × [0, T ],
s F2 ) (8.18)
∂t + u · ∇k = P − β ∗ ωk + ∇ · [(ν + σk νt ) ∇k] in Ω × [0, T ],
∂k
∂ω
+ u · ∇ω = αΩ2s − βω 2 + ∇ · [(ν + σω νt ) ∇ω]
∂t
+2(1 − F1 ) σωω2 ∇ω · ∇k in Ω × [0, T ].
8.3 LES
Large eddy simulation (LES) was introduced for the first time in 1963 by Joseph Smagorinsky
to simulate atmospheric flows [541]. The basic idea of LES is to obtain a significant reduction
of the computational cost required by a DNS by neglecting the smallest length scales via low-
pass filtering of the NSEs. Then the information related to such scales is recovered by proper
additional modeling. So one can assert that in a LES framework the largest length scales are
simulated while the smallest length scales are modeled.
A LES filter can perform a spatial and/or temporal filtering operation. The filtered field
ϕ(x, t) associated with a field ϕ(x, t) is defined as
Z ∞Z ∞
ϕ(x, t) = ϕ(r, τ )G(x − r, t − τ )dτ dr, (8.19)
−∞ −∞
∂u 1
+ ∇ · (u ⊗ u) = − ∇p + ν∆u − ∇ · τ SGS , (8.25)
∂t ρ
where
τ SGS = u ⊗ u − u ⊗ u (8.26)
is the subgrid scale stress tensor that takes into account the effect of filtered (namely solved)
scales and small (namely unsolved) scales as well as their interaction and needs to be properly
modeled. By using (8.21) we obtain the following expression for τ SGS :
τ SGS = u ⊗ u − u ⊗ u + u ⊗ u0 − u0 ⊗ u + u (8.27)
0 u0 ,
| {z } | {z } |{z}
R
L C
where L is the Leonard stress tensor involving only the resolved quantities (and therefore it can
be computed). However, C, the cross-term stress tensor, and R, the subgrid scale Reynolds stress
tensor (which can be interpreted as an analogous term to the Reynolds stress tensor in the RANS
equations), are related to unresolved scales. Then the filtered NSEs are given by (8.22)–(8.25).
A LES filtering approach that has recently been successfully used in the context of numerical
simulations of incompressible flows at moderately large Reynolds numbers is the Leray model
[371]. It couples the NSEs with a differential filter, resulting in the following system:
∂u 1
+ ∇ · (u ⊗ u) = − ∇p + ν∆u, (8.28)
∂t ρ
∇ · u = 0, (8.29)
−δ 2 ∇ (a(u)∇u) + u + ∇λ = u, (8.30)
∇ · u = 0, (8.31)
where δ can be interpreted as the filtering radius (namely the radius of the neighborhood where
the filter extracts information from the unresolved scales), the variable λ is a Lagrange multiplier
to enforce the incompressibility constraint for the filtered velocity u, and a(·) is the so-called
indicator function such that a(u) ≈ 0 where the flow does not need regularization and a(u) ≈ 1
where the flow does need regularization. The indicator function plays a crucial role in the success
of the Leray model. The choice a = 1 corresponds to the classic Leray model [371] that has the
advantage of making the operator in the filter equations linear and constant in time, but its ef-
fectivity is rather limited because it introduces the same amount of regularization in each region
of the computational domain, which results in overdiffusion in many cases. Different choices of
a(·) have been proposed and compared in [72, 599, 300, 362]. One of the most mathematically
convenient indicator functions is a(u) = |∇u| [69] because of its strong monotonicity prop-
erties. With this choice, we recover a Smagorinsky-like model, which is however known to be
not sufficiently selective. Recently, a class of deconvolution-based indicator functions has been
investigated:
a = |v − F (v)|, (8.32)
where F is typically the linear Helmholtz filter operator FH defined by
F = FH = (I − α2 ∆)−1 . (8.33)
8.4. Hybrid Projection-Based/Data-Driven ROM for Turbulent Flows 171
ṽ − α2 ∆ṽ = v. (8.34)
A monolithic approach to solving the problem (8.28)–(8.31) would be associated with a high
computational cost. In order to decouple the Navier–Stokes system (8.28)–(8.29) from the filter
system (8.30)–(8.31) at time tn+1 , a solve-then-filter approach can be employed: (i) solve equa-
tions (8.28)–(8.29) first, replacing the advection field un+1 with a suitable extrapolation, and (ii)
then solve the filter problem (8.30)–(8.31). In this framework, the evolve-filter-relax (EFR) algo-
rithm has recently been proposed, providing very promising results; see, e.g., [362, 58, 229, 231].
It was shown in [430] that the EFR algorithm is equivalent to a generic viscosity model in LES.
The two main advantages of the EFR method over other LES models are (i) the modularity, which
in fact introduces a differential problem to the Navier–Stokes problem unlike, e.g., the variational
multiscale approach [49], where extra terms are added directly in the NSEs, and (ii) the ease of
implementation; in fact the filter problem can be interpreted as a generalized Stokes problem and
can therefore be employed in a legacy Navier–Stokes solver by allowing it to simulate higher
Reynolds number flows without major modifications to the software core.
For a complete validation of the EFR method for Reynolds numbers up to 6500 in fixed
domains, we refer the reader to [58, 229].
• The approach complicates the task of solving the reduced order system in the online stage.
This is due to the additional equations, which have to be treated at the reduced order level,
where the additional turbulent equations are characterized by high nonlinearity.
• The approach forces the creation and the maintenance of a ROM for each turbulence clo-
sure model available at the full-order level. This customization of the ROM is clearly not
practical. For instance, a popular library such as OpenFOAM is well supplied with various
closure models, all of which would require a separate and specific ROM.
172 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Since the goal here is to design unified reduction methodologies which work with multiple full-
order model solvers, each with its own specific implementation of several different turbulence
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
models, we prefer to opt for a different strategy. To this end, the proposed approach involves
extending the decomposition assumption only to the eddy viscosity νt without doing the same
for the other turbulence variables, such as k, , ω, or ν̃. This essentially means that a reduced
order version of the eddy viscosity (νtr ) is introduced, namely
Nνt
X
νt (x, t; µ) ≈ νtr (x, t; µ) = gi (t, µ)ηi (x), (8.35)
i=1
where ηi (x) is the ith eddy viscosity POD mode and gi (t, µ) is the ith coefficient of the POD
expansion. If one considers the above expansion as an extension to those used for the supremizer
(SUP)-ROM and the Poisson pressure equations (PPE)-ROM, then it can be seen that the reduced
eddy viscosity assumes a different set of degrees of freedom for ai (t, µ) and bi (t, µ), which are
the reduced velocity and pressure solutions, respectively. Indeed, in principle, g varies over
time, responding to variations of a and b. This reflects at the reduced order level the fact that
the turbulent viscosity in every EVM depends on the mean flow field variables. At this point,
Nνt
one has to find a suitable way to compute the eddy viscosity reduced solution [gi (t, µ)]i=1 .
Since the specific turbulence equations of each EVM will not be used, the approach chosen to
compute the reduced eddy viscosity coefficients is based on data-driven methods and in particular
interpolation with RBFs [363, 396]. In detail, the reduced turbulent model treatment starts by
computing the eddy viscosity modes by making use of the method of snapshots, as explained in
Subsection 1.2.3. First, we define the eddy viscosity snapshot matrix
h
Sνt = {νt (x, t1 ; µ1 ), . . . , νt (x, tNT ; µM )} ∈ RNνt ×Ns , (8.36)
where Sνi t is the ith column of the eddy viscosity snapshot matrix Sνt . Then the eddy viscosity
correlation matrix is computed as
(C νt )ij = Sνi t , Sνjt L2 (Ω) . (8.37)
where the matrix λνt contains in its diagonal the eigenvalues of the matrix C νt and V νt is the
matrix whose columns are the corresponding eigenvectors of C νt .
The next step consists of training the RBF using the data of the snapshots acquired. Later,
Nνt
during the online stage, the coefficients [gi (t, µ)]i=1 are interpolated.
We will extend the SUP-ROM presented in Chapter 7 for the turbulent case with the de-
composition assumption (8.35). This means that the momentum equation is projected onto the
modes of the velocity and the continuity equation is projected onto the modes of pressure, with
the use of the supremizer stabilization approach. These projections will result in the following
differential-algebraic system of equations (DAEs):
M ȧ = ν(B + BT )a − aT Ca + g T (CT 1 + CT 2 )a − Hb,
(8.39)
P a = 0,
where g is the vector containing the reduced order degrees of freedom of the eddy viscosity and
the new terms in the DAEs above are defined as
(CT 1 )ijk = (φi , ηj ∇ · ∇φk )L2 (Ω) ,
(CT 2 )ijk = φi , ∇ · ηj (∇φTk ) L2 (Ω) .
8.4. Hybrid Projection-Based/Data-Driven ROM for Turbulent Flows 173
The system (8.39) has Nu + Np equations, while the number of unknowns is Nu + Np + Nνt .
This problem is solved, as mentioned, by employing an interpolation technique to obtain the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
vector g. Thus for the vector g we are using a data-driven approach known in the literature as
POD-I [608, 600, 519]. We remark that the system in (8.39) will have additional penalty terms
if the penalty method for the treatment of the nonhomogeneous boundary conditions is used.
The ROM put forward makes use of the projected momentum and continuity equations to
obtain the reduced solution of the velocity and the pressure. On the other hand, it employs data-
driven techniques to approximate the eddy viscosity solution manifold. These kinds of ROMs are
called hybrid ROMs because they merge the classical projection-based methods with data-driven
techniques. It is common to label the projection-based methods with the term intrusive, while
the term nonintrusive is used to describe methods which employ only the data without using the
governing equations. In this contribution we will refer to the ROM given by the reduced DAE in
(8.39) as H-SUP-ROM.
Data-driven methods are helpful in approximating the maps between various terms and quan-
tities which appear in the full-order model (FOM) formulation of the NSE. In this work, the
data-driven method selected is, as mentioned, interpolation using RBF. This method is used to
Nνt
approximate the coefficients [gi (t, µ)]i=1 in (8.39). In order to do this, one has to use the data
available for the eddy viscosity. This task can be done by properly choosing the maps that need
to be approximated or, in other words, by choosing a suitable independent variable of the RBF
interpolation. In the upcoming subsections, we present two different methodologies for carrying
out the interpolation step with their corresponding hybrid ROM.
where Xµ,t is the Cartesian product of the set containing the time instants at which snapshots
were acquired and the discretized parameter set. The set Xµ,t has a cardinality Ns , and its ith
member is denoted by xiµ,t . It can be seen that there is a unique correspondence between the
elements of Xµ,t and the columns of the matrix of snapshots (for all fields) used in the offline
stage to construct the reduced basis (RB).
The parameter sample introduced in the ROM at the online stage is denoted by µ∗ . In order
to have an accurate ROM result, the value of µ∗ should be close enough to the values of the
parameter samples used in the offline phase. Let t∗ be the time instant at which the ROM solution
is sought, where t1 6 t∗ 6 tNT . We refer to z ∗ = (t∗ , µ∗ ) as the online time-parameter
combined vector.
Let gr,l be the offline L2 projection coefficient vector which results from the projection of
the rth eddy viscosity snapshot Sνrt onto the lth eddy viscosity mode ηl , as follows:
At this point, one may define the interpolation problem as follows. Given the set Xµ,t , the cor-
Ns ,Nνt
responding eddy viscosity snapshots [Sνi t ]Ni=1 , and the corresponding coefficients [gr,l ]r=1,l=1 ,
s
approximate the vector g in (8.39) for the online time-parameter vector z . In this case one
∗
Nνt
extrapolates the scalar coefficients of the reduced eddy viscosity expansion [gi (t∗ , µ∗ )]i=1 (de-
174 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
N
noted briefly by [gi (z ∗ )]i=1
νt
), which means that one has to perform one interpolation for each of
the Nνt modes used in the online stage.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
It is worth remarking that the procedure described implies that each of the mode’s coefficients
are interpolated independently from the others. Therefore, we may fix the eddy viscosity mode
to be the Lth one, ηL , where L is an arbitrary integer index, 1 6 L 6 Nνt . The goal of the
interpolation is to approximate the map [GL ] in gL = GL (t, µ):
GL : Rq+1 → R for any L = 1, 2, . . . , Nνt . (8.42)
In order to approximate this map, firstly, we form the set of observations which consists of the
coefficients of the L2 projections of all the offline snapshots onto the eddy viscosity mode ηL ,
namely YL = [gr,L ]N r=1 ∈ R
s Ns
. The goal is to approximate the value of gL (z ∗ ) = gL (t∗ , µ∗ ).
Interpolation using RBF is based on the formula
Ns
wL,j ζL,j kz − xjµ,t kL2 (Rq+1 )
X
GL (z) = for L = 1, 2, . . . , Nνt , (8.43)
j=1
where ζL,j for j = 1, . . . , Ns are the RBF functions; z = (t, µ), where t refers to any time
instant inside the snapshot window and µ is a parameter value which lies in the parameter space
P; and wL,j are the weights of the RBFs which have to be computed during the training stage.
In order to compute the weights, we will use the data obtained by the FOM:
GL (xiµ,t ) = gi,L , for i = 1, 2, . . . , Ns , (8.44)
which implies that
Ns
wL,j ζL,j kxiµ,t − xjµ,t kL2 (Rq+1 ) = gi,L
X
for i = 1, 2, . . . , Ns . (8.45)
j=1
where (AζL )ij = ζL,j (kxiµ,t − xjµ,t kL2 (Rq+1 ) ). The linear system is then solved to obtain the
weights wL , which are stored in order to use them during the online stage. In this work, the
linear system above is solved using a QR decomposition with column pivoting.
Nνt
In the online stage, the value of g(z ∗ ) = [gi (z ∗ )]i=1 may be computed as
Ns
wi,j ζi,j kz ∗ − xjµ,t kL2 (Rq+1 )
X
∗ ∗
gi (z ) ≈ Gi (z ) = for i = 1, 2, . . . , Nνt . (8.47)
j=1
The strategy of the RBF interpolation in this subsection has the deficiency of not being accurate
for values of time t∗ which lie outside the snapshot window [0, T ] (the RBF interpolant is trained
with time values which lie inside the aforementioned range, and, therefore, it is not logical to
expect that it could accurately approximate values of t∗ that do not fall in [0, T ]). This could
be considered as a major issue in several applications in ROM since extrapolation is generally
regarded as a more complex task than interpolation. In addition, the strategy of treating time as a
part of the interpolation independent variable might not make physical sense. In fact, in several
applications the flow could be periodic, which could make the absolute value of the time variable
irrelevant to the dynamics of the fluid fields. For these reasons, we have tried to circumvent this
problem by adopting a different methodology for interpolating the RBF, as explained in the next
subsection.
8.4. Hybrid Projection-Based/Data-Driven ROM for Turbulent Flows 175
Coefficient Values
The RBF interpolation methodology discussed in the last subsection is quite simple and straight-
forward. However, this methodology only works with steady and unsteady flows, without being
able to extrapolate over time (outside the offline time snapshot window [0, T ]). H-SUP-ROM
with the methodology presented in the last subsection is introduced in [284], where it is applied
to a steady problem. This subsection suggests a different way of carrying out the RBF inter-
polation, in which the velocity projection coefficients (or a form of them as we will see) are to
be the independent variable of the interpolation. This choice will be justified in the following
discussion. The approach carried out in this subsection is presented in [286].
Before addressing the choice of independent variable for the interpolation, it is important to
introduce a modification to the way the eddy viscosity field was decomposed at the reduced order
level in (8.35). Primarily, we suggest spliting the FOM eddy viscosity field as follows:
where basically we have divided the eddy viscosity into two contributions. The first term corre-
sponds to the time-averaged viscosity field for the parameter µ, while the second term represents
the time-varying part. The rationale for this proposal is based on the fact that the parameter
variation is largely seen in the mean part νt (x; µ). On the other hand, parameter changes do
not affect to a great extent the small oscillations in the field νt0 (x, t; µ). In fact, we have noticed
that, in the numerical examples discussed in this work, we may drop the dependency on the pa-
rameter for the time-varying field, i.e., νt0 (x, t; µ) ≈ νt0 (x, t). This decomposition will turn out
to be beneficial for approximating the reduced order eddy viscosity. Splitting the eddy viscosity
into two terms, one which is time dependent and the other which is parameter dependent, makes
approximating νt an easier task.
The reduced order approximation of the νt is now modified as follows:
M Nνt
X X
νt (x, t; µ) ≈ νtr (x, t; µ) = g i (µ)η i (x) + gi (t)ηi (x), (8.49)
i=1 i=1
where the averaged part νt (x; µ) is approximated by the first sum in the above decomposition,
and the second sum approximates the time-varying part νt0 (x, t). In this formulation, one may
look at the fields [η i ]M
i=1 as additional eddy viscosity modes. The number of these fields is M ,
which is the number of parameter samples used in the offline stage (or the dimension of the dis-
cretized parameter set PM ). Each of these fields is computed by taking the time average over the
offline snapshots which correspond to only one of the parameter samples inside PM . In other
words, the field η i is computed by taking the average of the eddy viscosity fields acquired at dif-
ferent time instants but corresponding to the ith parameter in PM . As for [g i ]M i=1 , they represent
the degrees of freedom corresponding to the averaged fields and are important for periodic regime
problems. These are parameter dependent, and if we are testing the ROM in the online stage for
the ith parameter inside the offline parameter set (i.e., µ∗ = µi ), which is a reproduction test
in the parameter space, then the vector g = eM i , where ei is the unit vector of dimension M
M
consisting of zeros except for the its ith component, which it has the value one. The value of g
for cross-validation tests (extrapolation in the parameter space) can be fixed by linear interpola-
tion; e.g., if µ∗ lies at the distances d∗1 and d∗2 in the parameter space from its closest two offline
parameter samples µ1 and µ2 , respectively (i.e., kµ∗ − µ1 kL2 (Rq ) = d∗1 , kµ∗ − µ2 kL2 (Rq ) =
d∗2 , d∗1 < kµ∗ − µi kL2 (Rq ) and d∗2 < kµ∗ − µi kL2 (Rq ) for all i = 3, 4, . . . , M ), then
d∗ d∗
g can be approximated as d∗
2 M
∗ e1 + d∗
1
∗ e2 .
M
Finally, the dynamical system of the H-SUP-
1 +d2 1 +d2
176 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
P a = 0,
(8.50)
where the two new tensors are defined as
(CT 1 )ijk = φi , η j ∇ · ∇φk L2 (Ω) ,
(CT 2 )ijk = φi , ∇ · η j (∇φTk ) L2 (Ω) .
At this point, we may come back to tackle the main issue in this subsection, which is how we
could choose the independent variable of the interpolation in such a way that it leads to a hybrid
ROM which we could extrapolate over time. The suggested solution is to take the velocity
L2 projection coefficients as the independent variable of the RBF interpolation. This choice is
inspired by the fact that the eddy viscosity ultimately depends on the mean velocity field and its
gradient. The spirit of this choice is established on the fact that the eddy viscosity νt is a function
of the time history of the velocity field u. In other words, if we denote the FOM eddy viscosity
and the velocity at time tn by νtn and un , respectively, then we may express the eddy viscosity as
νtn = νt (u1 , u2 , . . . , un ). In order to mimic the dependency between the velocity field and the
eddy viscosity field at the reduced order level, we have to take into account the time evolution
aspect of the last equation. This is done by assuming that the eddy viscosity reduced vector g is
a function of the velocity coefficients a, along with their time derivative ȧ, i.e.,
Formula (8.51) represents an approximation of the relationship that exists between the eddy
viscosity and velocity fields. This approximation has been successfully used in the numerical
examples that we have considered in this work. However, the eddy viscosity evolves through a
PDE, albeit in many cases not directly, but passing through an algebraic relationship with turbu-
lence variables, which in turn evolve through PDEs. Thus, it is possible that the approximation
(8.51) could eventually be improved by including other fluid dynamics variables in the turbulence
model PDEs.
We proceed now to the training phase of the RBF interpolation. This phase is conducted
in the ROM offline stage. First, the L2 projection coefficients of the velocity modes onto the
snapshots are computed. Then, those corresponding to the supremizer are dropped, and the time
derivatives of the projection coefficients are calculated using a backward Euler scheme. The
velocity projection coefficients with their corresponding derivatives are put together to train the
RBF. In more detail, the procedure starts from the solution snapshots, namely
Sµ1 ,u Sµ1 ,p Sµ1 ,νt
Sµ2 ,u Sµ2 ,p Sµ2 ,νt
Su = .. , Sp = .. , Sν t = .. , (8.52)
. . .
SµM ,u SµM ,p SµM ,νt
where the snapshot matrices for all the variables have been expressed as M vertically aligned
submatrices, with each submatrix containing the time snapshots corresponding to a single pa-
rameter sample. Then one may define the L2 velocity projection coefficients arµk ,L2 ∈ RNu :
arµk ,L2 = [(Sµr k ,u , φ1 )L2 (Ω) , . . . , (Sµr k ,u , φNu )L2 (Ω) ] (8.53)
for r = 1, 2, . . . , NT , k = 1, 2, . . . , M.
8.4. Hybrid Projection-Based/Data-Driven ROM for Turbulent Flows 177
Let
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
At this point, we can compute the time derivative vectors needed for the RBF interpolation by
simply using the backward differentiation scheme as follows:
In order to build the matrix containing all the data observations needed to train the RBF, one has
to merge the L2 projection coefficients of velocity, starting from the second time snapshot, with
the time derivative coefficients. This gives the following matrix:
As for the eddy viscosity, one has to compute the projection coefficients of the eddy viscosity
modes onto the snapshots corresponding to the time-varying part in (8.48). In other words, the
time-averaged part corresponding to the kth parameter in the training set PM , which is νt (x; µk ),
has to be subtracted from each eddy viscosity snapshot, as follows:
r
gµ k ,i,L
r
2 = (Sµ ,ν − νt (x; µk ), ηi )L2 (Ω)
k t
for r = 2, 3, . . . , NT , (8.58)
i = 1, 2, . . . , Nνt , and k = 1, 2, . . . , M.
Let G̃i,k ∈ R(NT −1) be the vector containing the coefficients in (8.58) for fixed i and k. The
combined matrices and vectors for all parameter samples are denoted by à and G̃i , respectively,
and are defined as follows:
Ã1 G̃i,1
Ã2 G̃i,2
à = . ∈ R(Ns −M )×2Nu , G̃i = .. ∈ R(Ns −M ) . (8.59)
. . .
ÃM G̃i,M
N
The maps to be interpolated are [Gi ]i=1
νt
in gi = Gi (a, ȧ), where
This approximation is based on the interpolation points given in each row of the matrix à and
the vector G̃i . The interpolation procedure relates the mth row of Ã, which is a vector called
ãm
L2 ∈ R
2Nu
, to the mth element of G̃i , denoted by [G̃i ]m ∈ R.
178 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
NX
s −M
Gi (ãonline ) = wi,j ζi,j kãonline − ãjL2 kR2Nu for i = 1, 2, . . . , Nνt , (8.61)
j=1
where ãonline := [aonline , ȧonline ] ∈ R2Nu is a combination of the reduced velocity vector and its
vector derivative. The FOM data is used to establish the relation
Gi (ãm
L2 ) = [G̃i ]m for m = 1, 2, . . . , Ns − M, (8.62)
NX
s −M
j
wi,j ζi,j kãm
L2 − ãL2 kR2Nu = [G̃i ]m for m = 1, 2, . . . , Ns − M. (8.63)
j=1
N
In the online stage one can approximate g(a∗ , ȧ∗ ) = [gi (a∗ , ȧ∗ )]i=1
νt
:
NX
s −M
gi ((a∗ , ȧ∗ )) ≈ Gi ([(a∗ , ȧ∗ )]) = wi,j ζi,j k[(a∗ , ȧ∗ )] − ãjL2 kR2Nu . (8.64)
j=1
We would like to remark that the H-SUP-ROM in its different versions discussed in this section
is compatible with both the penalty method and the lifting function method. The interpolation
discussed in this subsection could be done with the velocity coefficients coming from the lifting
Nνt
modes when the lifting function approach is employed. If that is the case, then the maps [Gi ]i=1
in gi = Gi (a, ȧ) are
It is also important to underline that the homogenized set of velocity snapshots Sũ in Chapter 7
has to replace Su in (8.52) if the lifting function method is used.
The H-SUP-ROM in (8.50) can be solved by time-integrating the dynamical system in the
case of unsteady flows, or by simply solving the algebraic system in the case of steady flows.
In this work, we used a Newton method to solve the reduced dynamical system. The Newton
method computes the Jacobian numerically. In the general unsteady case, it is recommended to
choose a time step which is consistent with the one used during the FOM simulations. The time
advancement scheme for computing the derivative of the reduced vector a is chosen as a first- or
second-order backward Euler scheme.
was done with the pressure field. In other words, the reduced eddy viscosity is given by
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Nr
X
νt (x, t; µ) ≈ νtr (x, t; µ) = ai (t, µ)ηi (x), (8.66)
i=1
The corresponding DAE for this turbulent ROM, which is based only on the projection of the
momentum equation, is
ȧ = ν(B + BT )a − aT (C − CT 1 − CT 2 )a − Ha, (8.68)
where the vector a represents the degrees of freedom of the velocity, pressure, and eddy viscosity.
This ROM can be used to reduce both steady and unsteady flows. It also has the advantage of
being of low computational online cost, since the system (8.68) is relatively small. However, as
mentioned earlier, this formulation undermines the reduced pressure approximation. In addition,
numerical evidence has demonstrated that this ROM is prone to long-term instabilities when the
system (8.68) is integrated over long times (see Subsection 8.6.2). Furthermore, the U-ROM
is limited by not being compatible with the lifting function method described in Subsection
7.2.5. The lifting function method requires a homogenization process for the velocity snapshots.
The proposal of having a unique set of reduced coefficients [ai (t, µ)]N i=1 is directly affected by
r
the homogenization process of the velocity snapshots. The process of modifying the velocity
snapshots suggests that this process should be followed for the pressure and the eddy viscosity
snapshots, which is not well defined. In the next chapter, we compare the turbulent U-ROM
and the H-SUP-ROM with the penalty method used to enforce the nonhomogeneous boundary
conditions.
At this point, one may proceed to the extension of the PPE-ROM presented in Subsection
7.2.3. There are two different ways of extending this ROM. The first is based on assuming
that the reduced eddy viscosity field has the same temporal coefficients as the velocity field, as
proposed in (8.67). In this formulation the velocity and the eddy viscosity fields will have the
vector a as their reduced vector, while the pressure field will still be computed using a different
set of coefficients represented by the vector b. This ROM is called the semiuniform-PPE-ROM
or SU-PPE-ROM. We recall the reduced approximations of this ROM:
Nu Np
X X
u(x, t; µ) ≈ ai (t, µ)φi (x), p(x, t; µ) ≈ bi (t; µ)χi (x), (8.69)
i=1 i=1
Nu
X
νt (x, t; µ) ≈ ai (t, µ)ηi (x), (8.70)
i=1
The FOM momentum and Poisson equations for the RANS turbulent modeling are as follows:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
h i
T
∂u
∂t + ∇ · (u ⊗ u) = ∇ · −pI + (ν + νt ) ∇u + (∇u) in Ω × [0, T ],
h i
T
in Ω, (8.73)
∆p = −∇ · (∇ · (u ⊗ u)) + ∇ · ∇ · νt ∇u + (∇u)
+ Boundary conditions on Γ × [0, T ],
+ Initial conditions in (Ω, 0).
The Galerkin projections of the momentum and the Poisson equations give the following DAE:
ȧ = ν(B + BT )a − aT (C − CT 1 − CT 2 )a − Hb,
(8.74)
Db + aT (G − CT 3 − CT 4 )a − νN a − L = 0.
The additional tensors appearing in the reduced Poisson equation come from the divergence of
the eddy viscosity term in the FOM momentum equation in (8.18). Their entries are defined as
The SU-PPE-ROM cannot be used to reduce steady flows. This is related to the boundary con-
ditions, which have to be satisfied at the full-order level. We mentioned this issue in Subsection
7.2.3, where it is related to the way the PPE is derived from the momentum and continuity equa-
tions of the NSEs or the RANS. The derivation of an equivalent system to the one formed by the
NSEs requires an additional boundary condition for the divergence of the velocity or the pressure
[313, 376]. In this work, the PPE formulation adopted involves an additional Neumann bound-
ary condition for the pressure field. This latter condition is needed in order to make sure that the
velocity field is divergence free for all time instants. This condition can be fulfilled only in the
general unsteady setting (see the remark in Section 2 in [313]).
The additional tensors CT 3 and CT 4 account for the parameter contribution of the eddy viscosity
in the Poisson equation and are calculated as
(CT 3 )ijk = ∇χi , η j ∇ · ∇φk L2 (Ω) ,
(CT 4 )ijk = ∇χi , ∇ · η j (∇φTk ) L2 (Ω) .
Like the case of the PPE-ROM, its hybrid turbulent extension works only for nonstationary
flows. The H-PPE-ROM has an advantage over the H-SUP-ROM in having fewer unknowns in
the dynamical system because the supremizer reduced degrees of freedom are not present in the
H-PPE-ROM formulation. This often results in a slight advantage in terms of speedup during the
online stage.
8.6. Application of the H-SUP-ROM to Turbulent Problems 181
In this section, we present the results obtained by applying the proposed POD-Galerkin H-SUP-
ROM to two test cases that are very common benchmark problems in the development of tur-
bulent flow simulation tools. The first problem is that of turbulent flow past a backward-facing
step. This classical benchmark in the turbulence modeling community is used here to evaluate
the performance of the proposed ROM when applied to parametrized problems based on steady
RANS equations. The second problem is that of turbulent flow past a circular cylinder. This case
is instead able to characterize the POD-Galerkin H-SUP-ROM behavior in the presence of un-
steady turbulent flows, which also depend on physical parameters—the Reynolds number in this
case. For both test cases we will compare the H-SUP-ROM results with the ones obtained by the
ROM developed in [380]. In that work, the authors proposed a ROM which considers a single set
of reduced coefficients for the velocity, pressure, and eddy viscosity field. The main advantage
of this methodology is that it allows for a unified approach to dealing with laminar or turbulent
flows, with a consequently simple implementation. From now on, in all the comparisons of the
results of the methodology proposed here, the ROM developed in [380] will be referred to as
U-ROM.
The finite volume C++ library OpenFOAM [612] has been used to solve all the full-order
problems involved in the numerical tests presented in this section. As for the ROM simulations,
the reduced order problems shown have been reduced and resolved using the C++-based library
ITHACA-FV [549].
Figure 8.1. The computational domain used in the numerical simulations; all lengths are de-
scribed in terms of the characteristic length D, which is equal to 1 m.
182 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
[1 × 103 , 2.5 × 104 ]. As for the numerical setup of the full-order simulations, the Gauss linear
scheme was selected to approximate the gradients, and the Gauss linear scheme with nonorthog-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
onal correction was selected to approximate the Laplacian terms. A second-order bounded Gauss
upwind scheme was instead used to approximate the convective term. Finally, a first-order
bounded Gauss upwind scheme is used to approximate all terms involving the turbulence model
parameters k, , and ω.
The modes of the velocity, pressure, and eddy viscosity fields were obtained by POD analysis
of the respective snapshot matrices. Figure 8.2 shows the cumulative eigenvalue decay for the
velocity, pressure, and eddy viscosity fields. As can be appreciated, a relatively small number of
modes is sufficient to recover most of the energy information contained in the snapshots.
Once the reduced model was trained and the POD modes were obtained, a set of H-SUP-
ROM and U-ROM simulations were executed on a set of sampling points in the parameter space,
which was not contained in the original set used for training. More specifically, the online sam-
ple values for U , denoted by Ui∗ , where i = 1, . . . , Nonline-samples , were chosen as 80 equally
distributed samples in the interval [3, 20]. This set of samples is selected to include both sam-
ples close to the offline training ones and others which lie almost midway between two offline
samples. Clearly, the test is aimed at assessing how accurate the reduced approximation is for
parameter values that were not in the training set.
The correct inflow velocity in the reduced simulations is enforced by means of the penalty
method, as in Subsection 7.2.5. In this regard, we must remark here that the simulation results
appeared quite sensitive to the penalization factor τ . A sensitivity analysis was performed to se-
lect a value of τ that would allow satisfactory results for both the k- and the SST k-ω turbulence
models considered.
The first step of the online stage is to interpolate the eddy viscosity coefficients with respect
to the values of the inflow velocity parameter considered. More specifically, the result of the
interpolation is the reduced eddy viscosity vector g, which is then introduced in the reduced
system (8.39) to close the mathematical problem and finally obtain the vectors of coefficients a
and b. The RBF interpolation methodology used in this work to obtain g based on the parameter
value is implemented in the C++ library SPLINTER [247].
Figure 8.3 depicts the velocity fields corresponding to the velocity value U ∗ = 7.0886 m/s
computed via the FOM, the U-ROM, and the H-SUP-ROM for the k- turbulence model. A
similar comparison is presented in Figure 8.4 for the pressure field and in Figure 8.5 for the
Figure 8.2. Cumulative ignored eigenvalue decay. In the plot, the solid black line refers to the
velocity eigenvalues, the dashed magenta line indicates the pressure eigenvalues, and the dash-dotted blue
line refers to the eddy viscosity eigenvalues.
8.6. Application of the H-SUP-ROM to Turbulent Problems 183
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
(c)
Figure 8.3. k- turbulence model case, velocity fields for the value of the parameter U = 7.0886
m/s: (a) shows the FOM velocity, while in (b) one can see the U-ROM velocity, and finally in (c) we have
the H-SUP-ROM velocity.
eddy viscosity field. At the online stage level, all the solutions presented in the aforementioned
figures were generated using 10 modes to approximate the velocity, pressure, supremizer, and
eddy viscosity fields, for both the H-SUP-ROM and the U-ROM. The images indicate that the
hybrid projection/data-driven-based approach allows for qualitatively accurate approximations
of the FOM solutions for velocity, pressure, and eddy viscosity fields. However, this is not
completely the case when the U-ROM approach is employed. In fact, despite being able to obtain
a qualitatively accurate result for the velocity and eddy viscosity fields, the U-ROM computes a
pressure field that does not correctly reproduce its FOM counterpart. To quantitatively measure
the performance of both ROMs, we evaluate the relative L2 error for the velocity and pressure
fields, which, respectively, read
in which u∗ and p∗ are general reduced order velocity and pressure fields, respectively. The
relative L2 errors between the FOM and the H-SUP-ROM velocity and pressure fields presented
(a) (b)
(c)
Figure 8.4. k- turbulence model case, pressure fields for the value of the parameter U = 7.0886
m/s: (a) shows the FOM pressure, while in (b) one can see the U-ROM pressure, and finally in (c) we have
the H-SUP-ROM pressure.
184 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
(c)
Figure 8.5. k- turbulence model case, eddy viscosity fields: (a) shows the FOM eddy viscosity,
while in (b) one can see the U-ROM eddy viscosity, and finally in (c) we have the H-SUP-ROM eddy
viscosity.
in Figures 8.3 and 8.4 are respectively u = 0.4444% and p = 0.3654%, confirming the positive
result of the qualitative assessment. As for the U-ROM results, the corresponding errors are
u = 0.6522% and p = 20.9441%, respectively. The results of this test case suggest that the
H-SUP-ROM is able to overcome the main limit of the U-ROM approach, which is represented
by its inability to obtain accurate pressure field approximations.
A further simulation campaign was carried out with a different SST k-ω turbulence model
to evaluate how responsive the results of the hybrid H-SUP-ROM and of the U-ROM are with
respect to the turbulence model employed for the FOM simulations. Thus, a new set of SST
k-ω FOM simulations was executed. The inflow velocity parameter values used for this simu-
lation campaign were the same as in the k- model case previously described. The snapshots
generated were once again used for the modal decomposition–based training of both reduced
models considered. Figures 8.6, 8.7, and 8.8 show the velocity, pressure, and eddy viscosity
fields obtained by the FOM, the U-ROM, and the H-SUP-ROM for the inflow velocity value
U ∗ = 7.0886, respectively. Again, a visual inspection of the plots suggest that the H-SUP-ROM
velocity, pressure, and eddy viscosity fields appear in good qualitative agreement with their SST
(a) (b)
(c)
Figure 8.6. SST k-ω turbulence model case, velocity fields for the value of the parameter U =
7.0886 m/s: (a) shows the FOM velocity, while in (b) one can see the U-ROM velocity, and finally in (c)
we have the H-SUP-ROM velocity.
8.6. Application of the H-SUP-ROM to Turbulent Problems 185
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
(c)
Figure 8.7. SST k-ω turbulence model case, pressure fields for the value of the parameter U =
7.0886 m/s: (a) shows the FOM pressure, while in (b) one can see the U-ROM pressure, and finally in (c)
we have the H-SUP-ROM pressure.
k-ω FOM counterparts. As for the U-ROM results, although the velocity and, to some extent,
the eddy viscosity, field appear close to the k-ω FOM result, the U-ROM pressure field is visibly
different from the FOM one. From a quantitative standpoint, the L2 relative errors between the
FOM and the H-SUP-ROM velocity and pressure fields are, respectively, u = 0.8088% and
p = 0.7329%. As for the U-ROM results, the corresponding errors are u = 0.8177% and
p = 22.3972%, respectively. These results once again confirm that the H-SUP-ROM is able to
improve the pressure predictions obtained with the U-ROM approach.
The fact that the same test was run with both the k- and the SST k-ω models offers the
opportunity to measure how sensitive the H-SUP-ROM results are to the turbulence models em-
ployed at the offline FOM stage. Since the most appreciable differences between the k- and
k-ω solutions occur in the pressure fields, in Figure 8.9 we present the pressure field values on a
horizontal line cutting the domain at a vertical coordinate x2 = 5D6 located at half the maximum
height. The curves in the plot, which represent the pressure predicted by each model on the afore-
mentioned line as a function of the horizontal coordinate x1 , show a clear difference between the
k- and k-ω full-order model results. Quite remarkably, the curves referring to the H-SUP-ROM
(a) (b)
(c)
Figure 8.8. SST k-ω turbulence model case, eddy viscosity fields: (a) shows the FOM eddy
viscosity, while in (b) one can see the U-ROM eddy viscosity, and finally in (c) we have the H-SUP-ROM
eddy viscosity.
186 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 8.9. The pressure fields obtained using both k- and SST k-ω turbulence models and the
H-SUP-ROM ones for the value of the parameter U = 7.0886 m/s. The plot is for the pressure value along
the x1 direction, keeping the value of x2 fixed at half the maximum height.
pressure prediction each fall very close to their FOM counterparts. The H-SUP-ROM trained
with the k- model (green dots) is in fact very close to the black line representing the FOM k-
result, while the H-SUP-ROM trained with the SST k-ω model (blue dots) is instead very close
to the red line representing the FOM SST k-ω prediction. Thus, the investigation presented in
the plots suggests that the H-SUP-ROM approach is able to adapt to the turbulence model used
in the FOM simulations without losing accuracy. This accomplishes one of the main goals of the
H-SUP-ROM development, which was to obtain a ROM model sensitive to the turbulence model
used at the FOM level, without involving its specific turbulence model equations in the Galerkin
projection.
Finally, the convergence analysis for the H-SUP-ROM results is shown in Figure 8.10. The
plots show the mean L2 relative error over all the 80 samples used in the cross-validation test
in the online stage as a function of the number of modes used. As previously mentioned, in
these preliminary tests we used a uniform number of modes for velocity (Nu ), pressure (Np ),
supremizer (NS ), and eddy viscosity (Nνt ). The plots indicate that for the problem considered,
the H-SUP-ROM results exhibit fast convergence to the FOM solution for both k- and SST k-ω.
Yet, after less than 10 modes, the convergence appears to stall, as the error settles on nonzero,
but fairly acceptable, values. This is likely because, as the number of modes grows, the gain in
accuracy becomes only marginal compared to the νt field interpolation error.
(a) (b)
Figure 8.10. The mean of the L2 relative errors for all the online samples versus the number of
modes used in the online stage. The convergence analysis is done for both H-SUP-ROM models obtained
with two different turbulence models at the full-order level, which are k- and SST k-ω. The errors are
reported in percentages—in (a) we have the mean error of the velocity fields, while in (b) we have the mean
error of the pressure fields.
8.6. Application of the H-SUP-ROM to Turbulent Problems 187
The present subsection evaluates the performance of the reduction methodologies on an unsteady
turbulent test case. The classical computational fluid dynamics (CFD) benchmark considered
is that of unsteady flow past a cross-flow circular cylinder. We refer the interested reader to
[640, 641, 526] for a deep analysis of the main physical aspects of this phenomenon, which is
of particular relevance in the design of several engineering artifacts, such as electrical power
lines, gas and oil pipelines and risers, suspension bridge cables, and chimneys. Aside from their
physical relevance, vortex-shedding models are also of great interest among applied mathemati-
cians, because they represent a particular case of unsteady PDE problems with steady boundary
conditions. For these reasons, in the recent past several researchers in the model order reduction
community have considered vortex shedding past cylinders as an ideal playground for testing
reduction algorithms. Among others, we mention [548, 402, 628, 91, 253].
Problem Setup
Given the high cylinder length/diameter ratios encountered in all the aforementioned application
fields, tip effects typically have marginal relevance in this kind of flow. Thus, vortex shedding
past circular cylinders can be considered an inherently two-dimensional phenomenon. Although
there are of course notable exceptions to this assumption (see for instance [403]), we based the
validation campaign reported in the present subsection on a two-dimensional simulation setup.
A sketch of the flow domain and details of the computational grid composed of 11644 cells are
depicted in Figure 8.11. All the distances indicated in the diagram are expressed in terms of the
characteristic length of the problem, i.e., the cylinder diameter D = 1 m. The figure also reports
the boundary conditions imposed on the velocity and the pressure fields in correspondence with
all the domain edges. In all the simulations, the fluid kinematic viscosity ν is set equal to 10−4
m2 /s, and the horizontal velocity at the inlet is set to the uniform value Uin . Through variations
of this inlet velocity magnitude Uin , the test case is parametrized based on the Reynolds number
Re. In more detail, Uin was varied in the range [7.5, 12] m/s throughout the simulation cam-
paign, corresponding to Re ∈ [7.5 × 104 , 1.2 × 105 ]. Each simulation evolved over time until
a final periodic regime was fully developed, and a minimum of two periods were available for
postprocessing.
The full-order simulations were carried out using the nonsteady solver pimpleFoam imple-
mented in the framework of the OpenFOAM software library. This solver merges the SIM-
PLE and the PISO algorithms described in Subsection 7.2.2. The pimpleFoam solver in Open-
FOAM is able to adapt the time step length to ensure that the maximum Courant number CF L
[138, 139] does not exceed a prescribed threshold, which in this case has been set to CF Lmax =
0.9. A backward Euler time-advancing scheme was selected, resulting in first-order accurate
time derivatives of the velocity field. A similarly accurate first-order Gauss linear scheme was
(a) (b)
Figure 8.11. (a) The OpenFOAM mesh used in the simulations for the unsteady case of the flow
around a circular cylinder. (b) A picture of the mesh zoomed near the cylinder.
188 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
employed for the spatial gradients. As for the convective term, the selected approximation was
a second-order bounded Gauss upwind divergence scheme which utilizes upwind interpolation
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
weights, with explicit correction based on the local cell gradient. The diffusive term was also
discretized with a Gauss linear approximation scheme. The values of the relaxation factors αu
and αp (see Subsection 7.2.2) were set at 0.7 and 0.3, respectively. At each time step, a single
nonorthogonal corrector iteration was used to deal with the nonorthogonality of the mesh. As for
the linear solvers, a Gauss–Seidel smoother solver was used for the velocity equation, whereas
the GAMG solver with Gauss–Seidel smoother was selected for the pressure equation. Finally,
turbulence treatment at the full-order level was carried out by means of the SST k-ω turbulence
model available in the OpenFOAM library.
A first necessary step of the offline phase was to identify the sample points in the parameter
space. Ten uniformly spaced values of inlet horizontal velocity Uin were then taken inside the
range [7.5, 12] m/s. In each of the resulting simulations, snapshots were collected at each time
step after making sure that the flow reached the final periodic regime. We emphasize that in
such a numerical example, the objective of the reduction is to reproduce the final periodic regime
solution for each field and other related quantities. Thus, for each inflow velocity considered
the time snapshots need to be collected wisely by covering 1.5–2 solution cycles in the periodic
regime. This is vital because it allows one to compute POD modes able to fully represent the fluid
dynamics field features across the periodic solution cycle. Thus, in order to properly choose the
time instants at which snapshots had to be acquired, it was necessary to obtain the time period
of the final periodic regime. This was done by means of Fourier analysis, computing the fast
Fourier transform (FFT) of the FOM time signal of lift and drag fluid dynamics forces acting on
the cylinder’s surface. To illustrate this procedure, Figure 8.12 depicts the time signal of the lift
coefficient (obtained from the lift force L as Cl = 1 ρUL2 D ) for the case where Uin = 10 m/s.
2
Since this time signal was computed with nonuniform time steps, an interpolation procedure
was used to obtain the CL values on the uniform time grid required by the FFT algorithm. The
FFT analysis resulted in a vortex-shedding period of 0.4299 s, which corresponds to a Strouhal
number [563] of St = fVUS D = 0.2326, where fV S is the frequency of vortex shedding. We
point out that the value obtained is relatively close to the well-assessed experimental one of
approximately 0.20 [64]. Once the time period was computed, it was possible to simulate enough
cycles of the solution and collect the correct number of snapshots for each inflow velocity tested.
In the case presented in Figure 8.12, 1.2 additional seconds were simulated with a fixed time step
of 0.0003 s, while snapshots of the fluid dynamics fields were saved every 0.006 s. This resulted
in a total of 200 snapshots collected for the parameter sample point discussed here.
Figure 8.12. The lift coefficient curve for parameter sample Uin = 10 m/s.
8.6. Application of the H-SUP-ROM to Turbulent Problems 189
The procedure of acquiring the time step snapshots was then repeated for each parameter
sample in the training set. The number of snapshots taken for each parameter sample was NT =
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
200, which resulted in a grand total of Ns = 2000 snapshots. Table 8.1 reports the time steps
and the time intervals in which snapshots were taken for each value of the sampled parameter.
The table suggests that the time step and the time window in which snapshots were saved had
to be varied to adapt to the changes in the frequency of vortex shedding driven by modifications
of the asymptotic velocity Uin . The correct value of Uin was introduced in the offline full-order
simulations by means of a nonhomogeneous Dirichlet boundary condition at the inlet, enforced
by the penalty method [240, 539].
Once the results of all the offline simulations were collected, the POD procedure was directly
applied to the snapshot matrices of the velocity, pressure, and eddy viscosity fields. Figure 8.13
shows the cumulative ignored eigenvalue decay of the correlation matrices of the three fluid
dynamics variables. Once the POD modes were obtained, the supremizer problem could be
solved for each pressure mode [34], resulting in the supremizer modes used to enrich the velocity
POD space.
A further step of the offline phase was to set up the interpolation procedure to approximate the
eddy viscosity coefficient vector g. As discussed in Subsections 8.4.1 and 8.4.2, the value of
g used at each time step is interpolated based on either the current time and parameter values
(t∗ , Uin
∗
) or on the reduced order velocity coefficient vectors of a and ȧ as in (8.51). The main
limit of the application of the former interpolation approach for the present unsteady case is that
it restricts the online time integration to the window where snapshots were acquired (for instance,
1.2 s for the case of Uin = 10 m/s or the window of 1-s length for Uin = 12 m/s). For the
present test case we selected the latter interpolation strategy (described in Subsection 8.4.2) be-
cause it is designed precisely to allow for time extrapolation in the online phase. This is based
on the assumption that as the values of the components of the vector time signal a(t) oscillate
between a minimum and a maximum over time, the replacement of t by a(t) as the RBF inter-
Figure 8.13. Cumulative ignored eigenvalue decay. In the plot, the solid black line refers to the
velocity eigenvalues, the dashed magenta line indicates the pressure eigenvalues, and the dash-dotted blue
line refers to the eddy viscosity eigenvalues.
190 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
polation independent variable has in fact the advantage of allowing extrapolation. Nonetheless,
this is true if the values of the vector a components obtained during the ROM time integration
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
fall within the bounds of the FOM snapshots. Consequently, it is evident that the accuracy of this
interpolation outside the offline snapshot window is highly dependent on how close the current
solution vector a is to the vectors of the L2 projection coefficients (see (8.53)) used in the offline
stage to train the RBF. To further optimize the interpolation strategy, in this problem we applied
the splitting assumption presented in equation (8.48). Thus, the RBF interpolation was only used
to obtain the coefficients of the fluctuating part of the reduced eddy viscosity field. As a result,
the contribution of the parameter Uin only affects the time-averaged part νt (x; µ). At the re-
duced order level, the vector g is constant over time and is dependent on Uin . In the framework
of the present test case we had M = 10 corresponding to the number of eddy viscosity time-
averaged fields, which were computed as the average of the 200 time snapshots collected for
each parameter sample tested (see Table 8.1). At the online stage, the vector g(Uin ∗
) was com-
puted by linear interpolation, while the parameter-independent vector g(t ) was obtained from
∗
the RBF interpolation with respect to a and ȧ. Finally, the initial values for all vectors a(0, Uin
∗
),
b(0, Uin ), and g(0) were obtained from the inlet velocity parameter using a linear interpolation
∗
(based on the values of the initial L2 projection vectors of a(0, Uin ), b(0, Uin ), and g(0)). We
finally report on the tuning of the shape parameter required by the radial basis (RB) chosen for
the RBF interpolation. For all the tests in this work, the bases selected were Gaussian radial
functions, namely ζi,j (k · kR2Nu ) = exp(−r2 k · k2R2Nu ). The value of the RBF shape parameter
r was selected using cross-validation algorithm called leave-one-out cross-validation (LOOCV)
[190]. This choice resulted in satisfactory results for the tests conducted in this chapter.
Table 8.1. Offline parameter samples and the corresponding snapshot data.
Parameter sample: Uin in m/s FOM time step in s Snapshot acquiring time in s
7.5 0.0004 0.008
8 0.0004 0.008
8.5 0.00035 0.007
9 0.0003 0.006
9.5 0.0003 0.006
10 0.0003 0.006
10.5 0.0003 0.006
11 0.0003 0.006
11.5 0.00025 0.005
12 0.00025 0.005
Here, we first report the results of the reduction strategy based on the supremizer approach. We
will compare the U-ROM and the H-SUP-ROM on the unsteady turbulent test case considered
here. At a later stage, we will also report the results of the turbulent ROMs based on the Pois-
son pressure equation (PPE). However, most of the results presented focus on the H-SUP-ROM
because of its general reduction formulation (it may in fact be used for both steady and unsteady
flows).
At this point, we illustrate the details of the first numerical test, which is a cross-validation or
an extrapolation test in the parameter space. The test is aimed at reconstructing the time history
of the fluid dynamics fields for the online parameter value Uin ∗
= 7.75 m/s, which was not
8.6. Application of the H-SUP-ROM to Turbulent Problems 191
used to train the ROM. Also, one of the objectives of this test is to obtain the fluid dynamics
forces for time values which go far beyond the offline snapshot time window. After carrying out
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the offline phase, all the reduced vectors, matrices, and tensors which appear in the dynamical
systems of the U-ROM and the H-SUP-ROM (in (8.68) and (8.50), respectively) were saved for
later use at the online stage. The online stage involved integrating the aforementioned dynamical
systems over time to obtain the vector a in the case of the U-ROM and the two vectors a and b
in the case of the H-SUP-ROM. The initial reduced vectors were approximated as the weighted
average of the ones corresponding to the first two offline parameter samples since Uin ∗
= 7.75
m/s lies between them. The fields obtained by solving the reduced systems of both ROMs
were then compared to the FOM ones. The FOM simulator was run for enough time to reach the
periodic regime, before extending the simulation at a fixed time step of 0.0004 s. FOM snapshots
were taken every 0.008 s, and the total simulation time reached was 8 s, which encompassed 13
solution periods. We emphasize that the last FOM snapshots mentioned were computed just for
the sake of comparing them with the obtained ROM fields, and these FOM fields were obviously
not used during the offline stage compute the POD modes. Given the periodic time dependence
of the variables considered, and the objective of comparing different models for cross-validation,
it was also important to pay special attention to the phase of the FOM snapshots. In particular,
the phase of the first snapshot acquired for all the training parameter samples was chosen to be
the same. This was also the case for the FOM testing snapshots computed to assess the ROM
accuracy. It was also assumed that the starting time of the online simulations was equal to zero
in all the tests considered here.
The velocity, pressure, and eddy viscosity fields computed by the FOM, the U-ROM, and
the H-SUP-ROM at the time instant t = 2.8 s are depicted in Figures 8.14, 8.15, and 8.16,
respectively. The results show clearly that the FOM velocity field is accurately approximated by
both the U-ROM and the H-SUP-ROM. This visual assessment is also confirmed by quantitative
evaluation. The L2 relative errors for both the U-ROM and the H-SUP-ROM are in fact 1.3553%
and 0.6954%, respectively. As for the pressure field, upon visual inspection the accuracy of the
U-ROM is noticeably lower than that of the hybrid ROM. This is again confirmed by quantitative
evaluation, because the L2 relative error for the U-ROM and H-SUP-ROM pressure reduced
fields are 33.0963% and 4.8085%, respectively. The modes used in the online stage in this test
case were Nr = 14 in the case of the U-ROM and Nu = 20 and Np = NS = Nνt = 10
(a) (b)
(c)
Figure 8.14. Velocity fields for the parameter value Uin = 7.75 m/s at t = 2.8 s: (a) shows the
FOM velocity, while in (b) one can see the U-ROM velocity with Nr = 14, and finally in (c) we have the
H-SUP-ROM velocity with Nu = 20 and Np = NS = Nνt = 10.
192 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
(c)
Figure 8.15. Pressure fields for the parameter value Uin = 7.75 m/s at t = 2.8 s: (a) shows the
FOM pressure, while in (b) one can see the U-ROM pressure with Nr = 14, and finally in (c) we have the
H-SUP-ROM pressure with Nu = 20 and Np = NS = Nνt = 10.
for the H-SUP-ROM. This online mode setup results for both ROMs in the best results in terms
of the fluid dynamics field’s L2 norm relative error with respect to the FOM solution. The
reduced pressure field plots suggest that the main limit of the U-ROM approach is its inability
to obtain accurate pressure field predictions. This lack of accuracy is most noticeable in the
region close to the surface of the cylinder, which clearly affects the accuracy of the reduced
approximation of the lift and drag forces. This shortcoming is clearly associated with the lack of
specific pressure and turbulence treatments in the U-ROM formulation. On the other hand, the
H-SUP-ROM features additional variables that are included to reproduce at the ROM level the
pressure-velocity coupling and the turbulence model used in the FOM. As a result, the accuracy
of its pressure field prediction is satisfactory.
A further assessment of the accuracy of both ROMs is given by the time evolution of the L2
relative error associated with the approximation of each field. The plot corresponding to the time
(a) (b)
(c)
Figure 8.16. Eddy viscosity fields for the parameter value Uin = 7.75 m/s at t = 2.8 s: (a)
shows the FOM eddy viscosity, while in (b) one can see the U-ROM eddy viscosity with Nr = 14, and
finally in (c) we have the H-SUP-ROM eddy viscosity with Nu = 20 and Np = NS = Nνt = 10.
8.6. Application of the H-SUP-ROM to Turbulent Problems 193
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
Figure 8.17. The time evolution of the L2 relative errors of the velocity-reduced approximations
for both the U-ROM and the H-SUP-ROM. The curves correspond to the case run with the parameter value
Uin = 7.75 m/s: (a) shows the error curve for the U-ROM, while (b) depicts the case of the H-SUP-ROM.
The error values in both graphs are in percentages.
evolution of the velocity field error is depicted in Figure 8.17, while Figure 8.18 shows the error
curve for the pressure field. In ROM, it is expected that the error is a growing function of time.
This is especially true when the ROM simulation extends well beyond the time window in which
the training snapshots were collected. Thus, ROMs which are able to contain the growth of this
error during the integration of the dynamical system are considered more accurate and favorable.
In this regard, the error plots presented suggest that the H-SUP-ROM is more effective at curbing
the error values. The velocity field error diagrams also indicate that increasing the number of
velocity modes Nu (Nr ) results in lower values of u for the H-SUP-ROM (U-ROM). As for the
pressure field errors observed in Figure 8.18, an increasing number of modes Nr does not lead
to significant error reduction in the U-ROM. Instead, increasing the number of modes associated
with all the fields generally lowers the pressure error, which can be kept to satisfactory values for
the whole time window considered.
(a) (b)
Figure 8.18. The time evolution of the L2 relative errors of the pressure-reduced approximations
for both the U-ROM and the H-SUP-ROM. The curves correspond to the case run with the parameter value
Uin = 7.75 m/s: (a) shows the error curve for the U-ROM, while (b) depicts the case of the H-SUP-ROM.
The error values in both graphs are in percentages.
194 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
pend on the values of the velocity and pressure fields in the specific domain region surrounding
the boundary corresponding to such a body. Thus, evaluating the accuracy of the ROM using
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the global error evaluators discussed so far only provides marginal information on how well the
reduced solution is able to lead to satisfactory force predictions. In the case of the present nu-
merical test, for instance, a considerable pressure or velocity error localized in the small region
around the cylinder might have a substantial impact on the fluid dynamics forces acting on the
cylinder, while having little effect on the global field errors. For this reason, the following anal-
ysis considers the time evolution of the lift coefficient Cl , i.e., the nondimensionalized vertical
component of the fluid dynamics force acting on the cylinder. It is important to point out that
the lift and drag forces exerted by the fluid on the cylinder are not a direct result of the H-SUP-
ROM computations. The reduced system solution consists in fact of the modal coefficients of
the velocity and pressure fields at each time instant, which are in turn used to obtain the H-SUP-
ROM approximation of the full-rank flow field. This approximation can obviously be used to
obtain—by integrating the pressure and skin friction on the cylinder surface—the reduced order
approximation of the fluid dynamics force components and the corresponding nondimensional
force coefficients. Yet, in the ROM community it is recommended to avoid possibly expensive
operations which involve evaluating the full-rank flow field. For this reason, the lift and drag
coefficients in this work are computed in a fully reduced order fashion, based on the offline com-
putation of suitable matrices which are then used in the online stage. The detailed procedure for
the online fluid dynamics forces computation was laid out in Subsection 7.2.6.
The resulting lift coefficient curves computed by both the U-ROM and the H-SUP-ROM
are compared to the FOM. Figure 8.19 shows the time signal of the lift coefficient obtained
with the FOM, the U-ROM, and the H-SUP-ROM, when the inflow velocity parameter value
is set to Uin
∗
= 7.75 m/s. Plot (a) presents the full history of the lift coefficient for the [0, 8]
s time interval of this test. Plot (b) represents instead a closeup of the same curves’ behavior
in the last 2 s of the simulation. The diagrams—plot (b) in particular—indicate that the H-
SUP-ROM predictions lead to improved force predictions with respect to the U-ROM. Thus,
the lower overall L2 error norms on pressure and velocity fields obtained by the H-SUP-ROM
indeed result in more accurate force predictions with respect to the U-ROM. Moreover, the long-
time integration used in the present test suggests that the U-ROM solution presents a slightly
unstable behavior, since its force amplitude grows as the simulation time increases, along with
the prediction error. On the other hand, the H-SUP-ROM Cl curve matches its FOM counterpart
to a good degree, even in the final time instants of the simulation. The number of reduced modes
employed in the online stage is the same as reported for the fluid dynamics field error results
presented previously.
(a) (b)
Figure 8.19. Lift coefficient curves for the cross-validation test done for the parameter value
Uin = 7.75 m/s for the time range [0, 8] s; the figure shows the FOM, the U-ROM, and the H-SUP-ROM
lift coefficient histories: (a) the full range is shown, and (b) the last 2 s of Cl is shown.
8.6. Application of the H-SUP-ROM to Turbulent Problems 195
In order to have a better quantitative assessment of lift coefficient prediction accuracy by both
ROMs, we introduce the L2 relative percentage error
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
where Cl (t) is the time signal of the FOM lift coefficients between T1 and T2 . Cl ∗ (t) is the
corresponding lift coefficient time signal computed by the ROM—either the U-ROM or the H-
SUP-ROM. Referring back to Figure 8.19, we calculated Cl for the full time range of [0, 8] s and
also for its first half [0, 4] s. The values of the U-ROM for these time intervals are 16.7095% and
5.1252%, respectively. On the other hand, the H-SUP-ROM Cl approximation error is 3.5792%
for the full simulation time range [0, 8] s and 4.0504% for its first half [0, 4] s. The higher values
of the Cl error in the second part of the simulation confirm the U-ROM’s unsteady behavior,
which particularly affects problems requiring long-time integration. Such instability has not
been observed in the case of the hybrid model, in which instead the errors measured in the full
time interval and in its first half are very close. Figure 8.20 shows the error mCl as a function
of the number of modes used in the online stage. The plot suggests that the H-SUP-ROM force
coefficient error is minimized, considering 20 velocity modes and 10 modes each for pressure,
supremizers, and eddy viscosity fields. In this condition, the relative force coefficient error is as
low as 3%. As for the U-ROM Cl error, it is consistently above 16%, with a minimum value
obtained with 14 velocity modes.
Figure 8.21 displays the ROM results in terms of a further error metric devised in an attempt
to attribute how much of the L2 error is caused by imprecise reproduction of the amplitude rather
than the phase of the lift coefficient oscillations. The relative peak error peak is
P Kn,F OM − P Kn,∗
n,peak = × 100%, (8.78)
P Kn,F OM
where P Kn,F OM is the value of the nth FOM Cl peak and P Kn,∗ is the value of the nth U-ROM
or H-SUP-ROM Cl peak. The figure shows the relative peak error for both ROMs, computed for
(a) (b)
Figure 8.20. The graph of the L2 relative errors for the lift coefficient curve versus the number
of modes used in the online stage in the U-ROM and the H-SUP-ROM. The curves correspond to the case
run with the parameter value Uin = 7.75 m/s. The error is computed between the lift coefficient curve
obtained by the FOM solver and the one reconstructed from both the U-ROM and the H-SUP-ROM for the
time range [0, 8] s: (a) shows the error curve for the U-ROM, where Nr is the number of modes used in the
online stage for all variables (by construction of the U-ROM, it is not possible to choose a different number
of online modes for the reduced variables). (b) depicts the case of the H-SUP-ROM, where one can see
the error values by varying the number of modes used for the pure velocity, with different fixed settings for
the three other variables (the pressure, the supremizers, and the eddy viscosity). The error values in both
graphs are in percentages.
196 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
Figure 8.21. The graph of the peak relative errors for the lift coefficient curves for varied values
of the number of modes used in the online stage in the U-ROM and the H-SUP-ROM. The curves correspond
to the case run with the parameter value Uin = 7.75 m/s. The error is computed between the peak values
of the lift coefficient curve obtained by the FOM solver and the ones reconstructed from both the U-ROM
and the H-SUP-ROM for the time range [0, 8] s: (a) shows the error curve for the U-ROM, where Nr is
the number of modes used in the online stage for all variables (by construction of the U-ROM it is not
possible to choose a different number of online modes for the reduced variables). (b) depicts the case of the
H-SUP-ROM. The error values in both graphs are in percentages.
each of the 29 peaks in the time interval [0, 8] s. The different curves in the plot refer to the dif-
ferent combinations of modal truncation values employed for each field considered in the ROM
simulations. Figure 8.21 indicates that the U-ROM relative error increases at each peak until it
reaches values as high as 10–20%. On the other hand, the hybrid ROM presents lower values
of the relative peak errors, which are not monotonically growing at each lift coefficient peak
throughout the simulation. The peak error is lower than 3.5% for several combinations of the
modal truncation order fields tested. The cross-validation tests presented so far only involved the
U-ROM and the H-SUP-ROM. They were clearly aimed at establishing how well such reduced
order methodologies are able to perform when employed in a significant engineering application.
We now turn our focus to the performance of the two other turbulent ROMs described in Sec-
tion 8.5. To this end, we report on the results of a further cross-validation simulation campaign
carried out on the same cross-flow cylinder test case considered in the previous tests.
The ROMs we consider now are referred to as the SU-PPE-ROM and the H-PPE-ROM. At the
online level, both ROMs use the Poisson equation for pressure to obtain a stable pressure-velocity
coupling. Given the formulation of the Poisson equation for pressure boundary conditions, these
two ROMs are only suited for the reduction of unsteady flows, and they cannot be extended
to the steady case. The SU-PPE-ROM is an extension of the U-ROM, obtained by adding a
specific Poisson equation–based treatment for the pressure field. The H-PPE-ROM is a further
extension, obtained by adding a turbulence treatment equivalent to the H-SUP-ROM, based on
the data-driven approximation of the eddy viscosity field. Thus, whereas the SU-PPE-ROM is
an intrusive ROM, by our previous definition the H-SUP-ROM is a hybrid ROM.
As mentioned above, the cross-validation test for the SU-PPE-ROM and the H-PPE-ROM is
identical to the one conducted for the U-ROM and the H-SUP-ROM. Therefore, the online pa-
rameter value selected for the comparisons was Uin ∗
= 7.75 m/s. In the online simulations, the
SU-PPE-ROM was applied by using the penalty method to treat the nonhomogeneous Dirichlet
boundary conditions on the velocity field. On the other hand, the lifting function method was
used to enforce the boundary conditions in the H-PPE-ROM online solution. The lifting function
8.6. Application of the H-SUP-ROM to Turbulent Problems 197
chosen in this case corresponds to the solution of a potential flow problem with the inlet velocity
fixed at Uin = 1 m/s. Such an auxiliary problem was solved by running an iterative steady
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
solver. The lifting velocity field obtained was considered as the additional lifting mode/function
which was added to the velocity POD modes. The next step of the offline phase was to homoge-
nize the velocity snapshots using the velocity lifting function, which resulted in the 2000 velocity
snapshots processed to obtain the velocity POD modes. All the reduced matrices, vectors, and
tensors which appear in the dynamical systems of both ROMs and are reported in (8.75) and
(8.74) were then computed and stored. The initial velocity, pressure, and eddy viscosity reduced
vectors were set in the exact way as in the previous test.
The first results shown refer to the valuation of the lift coefficient Cl in the same [0, 8] s
time interval considered in the previous test. Equation (8.22) depicts the lift coefficient curves
obtained by the PPE ROM and the FOM. The plot indicates that both ROMs lead to acceptable
force coefficient predictions. If on one hand the accuracy of the hybrid H-PPE-ROM prediction
is once again superior to that of the SU-PPE-ROM, the latter model leads to significantly better
Cl results than the U-ROM ones presented in Figure 8.19. The main difference between the
U-ROM and the SU-PPE-ROM lies in the specific pressure treatment included in the latter ROM
online equations. Thus, its improved performance over the U-ROM indicates the accuracy gains
resulting from including suitable pressure-velocity coupling methodologies at the online stage.
In addition, the further performance improvement obtained with the H-PPE-ROM is an indirect
measure of the accuracy gains associated with the H-PPE-ROM-specific turbulence treatment,
which is the main algorithmic difference between the two ROMs.
We also point out that the combination of modal truncation levels for each fluid dynamics
field that was used to generate Figure 8.19 represents for both ROMs the best combination in
terms of accuracy measured by the error Cl (defined in (8.77)). In a similar fashion, Figure
8.22, which shows Cl as a function of time, was generated using the SU-PPE-ROM with the
number of velocity and pressure modes—nine—minimizing the error. Also, the H-PPE-ROM
curve in the plot was obtained with the most successful combination of modal truncation levels
tested, in which the number of velocity, pressure, and eddy viscosity modes is equal to 15. From
a quantitative perspective, the results in Figure 8.22 show that the SU-PPE-ROM has an error
Cl = 5.0525% for the time range [0, 4] s and Cl = 12.4255% for the full time range. On the
other hand, the corresponding H-PPE-ROM errors are 2.2445% and 4.3678%. This evaluation
confirms that, with the addition of a specific pressure-velocity coupling methodology, the SU-
PPE-ROM represents a significant improvement over the U-ROM. Nevertheless, its main source
of residual inaccuracy is associated with its turbulence treatment. In particular, the constraint of
having temporal coefficients shared between velocity and eddy viscosity could be problematic
(a) (b)
Figure 8.22. Lift coefficient curves for the cross-validation test done for the parameter value
Uin = 7.75 m/s for the time range [0, 8] s. The figure shows the FOM, the SU-PPE-ROM, and the
H-PPE-ROM lift coefficient histories: (a) the full range is shown, and (b) the last 2 s of Cl is shown.
198 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
in specific settings. For example, if 20 and 12 modes are needed to accurately reconstruct the
velocity and the eddy viscosity fields, respectively, then this formulation would force us either to
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
work with a nonoptimal set of velocity shape functions or to include a set of eight unneeded eddy
viscosity modes in the solution of the reduced DAE. Since these additional eddy viscosity modes
are typically affected by a high signal to noise ratio, this is likely to cause stability problems.
This is further confirmed by Figure 8.23, which presents the convergence analysis for the
ROMs based on the Poisson equation for pressure. The plots, which depict the SU-PPE-ROM
(left) and the H-PPE-ROM (right) lift coefficient error as a function of the velocity modal trun-
cation level, present different curves, associated with the different truncation level combinations
selected for the other fields in the tests. The left plot clearly shows that the SU-PPE-ROM error
reaches a minimum when Nu = Np = 9, which is the case depicted in Figure 8.22, for a total
number of degrees of freedom of 18. The SU-PPE-ROM error pattern is similar to the U-ROM
one depicted in Figure 8.20. Both curves present in fact two minima located at truncation values
of 9 and 14 for the velocity and the pressure modes, and the error values remain consistently
above 10%. This can once again be explained by pointing out that after nine velocity modes
are reached, the accuracy gains granted by adding further modes are spoiled by the presence of
unneeded and noisy eddy viscosity modes in the system.
As for the H-PPE-ROM, it was instead able to fully exploit the potential accuracy gains
associated with higher velocity truncation levels. The lift coefficient error in the right plot is in
fact still decreasing when the number of velocity modes grows up to 13–15 modes. In addition,
the error values are lower than 10% for several truncation settings. For example, Cl ' 8%
when 13 modes are employed for all reduced variables. We must however remark that among the
Nu H-PPE-ROM modes, we also include the lifting function mode added to treat the boundary
conditions at the reduced order level. Figure 8.24 reports the error for each peak of the Cl
obtained for the PPE ROMs for different settings of the online modes. The left plot substantially
confirms that the SU-PPE ROM improved results with respect to the U-ROM, which speaks to
the necessity of having a separate set of reduced coefficients for the pressure field. The values of
the SU-PPE ROM peak errors observed in the left diagram are in fact as low as 2% for certain
(a) (b)
Figure 8.23. The graph of the L2 relative errors for the lift coefficient curve versus number of
modes used in the online stage in the SU-PPE-ROM and the H-PPE-ROM. The curves correspond to the
case run with the parameter value Uin = 7.75 m/s. The error is computed between the lift coefficient
curve obtained by the FOM solver and the one reconstructed from both the U-ROM and the H-PPE-ROM
for the time range [0, 8] s: (a) shows the error curve for the SU-PPE-ROM, where Nu is the number of
modes used in the online stage for both the velocity and the eddy viscosity, while Np is the number of modes
used for the pressure field. (b) depicts the case of the H-PPE-ROM, where one can see the error values
varying the number of modes used for the velocity (including the lifting velocity mode) with different fixed
settings for the two other variables (the pressure and the eddy viscosity). The error values in both graphs
are in percentages.
8.6. Application of the H-SUP-ROM to Turbulent Problems 199
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
Figure 8.24. The graph of the peak relative errors for the lift coefficient curves for varied values
of the number of modes used in the online stage in the SU-PPE-ROM and the H-PPE-ROM. The curves
correspond to the case run with the parameter value Uin = 7.75 m/s. The error is computed between the
peak values of the lift coefficient curve obtained by the FOM solver and the ones reconstructed from both the
SU-PPE-ROM and the H-PPE-ROM for the time range [0, 8] s: (a) shows the error curve for the U-ROM,
where Nu is the number of modes used in the online stage for both the velocity and the eddy viscosity, while
Np is the number of modes used for the pressure field. (b) depicts the case of the H-PPE-ROM. The error
values in both graphs are in percentages.
choices of the online modal truncation levels. The image on the right illustrates that the H-PPE-
ROM results succeeded in recovering the peaks of the Cl time signal, with errors lower than 2%
for online modal truncation orders Nu = 13, 14, 15. A further result is reported in Figure 8.25
and Figure 8.26, which depict the time evolution of the ROMs’ L2 prediction error for velocity
and pressure fields, respectively. The plots further confirm the H-PPE-ROM’s superior ability to
obtain more accurate pressure field approximation than the SU-PPE-ROM. The SU-PPE-ROM
pressure field error values at the end of the 8-s time integration interval are above 10% for all
the different modal truncation orders tested. On the other hand, the corresponding error value
for the H-PPE-ROM reaches values as low as 4% at the final time of the reduced simulations,
for the best combinations of the online mode numbers tested. A comparison with Figure 8.18
also suggests that the H-PPE-ROM has a greater ability to suppress the error growth than the H-
SUP-ROM. A possible justification for this finding is that the use of the PPE at the reduced level
has introduced a higher level of consistency between the ROM formulation and the FOM one.
In fact, the finite volume FOM solver features a pressure equation obtained from the momentum
(a) (b)
Figure 8.25. The time evolution of the L2 relative errors of the velocity-reduced approximations
for both the SU-PPE-ROM and the H-PPE-ROM. The curves correspond to the case run with the parameter
value Uin = 7.75 m/s: (a) shows the error curve for the SU-PPE-ROM. (b) depicts the case of the H-PPE-
ROM. The error values in both graphs are in percentages.
200 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
Figure 8.26. The time evolution of the L2 relative errors of the pressure-reduced approximations
for both the SU-PPE-ROM and the H-PPE-ROM. The curves correspond to the case run with the parameter
value Uin = 7.75 m/s: (a) shows the error curve for the SU-PPE-ROM. (b) depicts the case of the H-PPE-
ROM. The error values in both graphs are in percentages.
and the continuity equation, as explained in Subsection 7.2.2. This procedure is mimicked by
the H-PPE-ROM, possibly allowing for an improvement in accuracy. Nonetheless, we have to
mention that in general the FOM solution algorithm differs from the ROM one, in that the FOM
one is based on an iterative segregated approach, while the ROM one uses a coupled approach.
To summarize the results given by the four ROMs tested in the current problem, Table 8.2
reports the performance of all the ROMs in terms of their computational efficiency and of the
accuracy indicators discussed. In the table SU stands for the speedup, which is calculated as
tof f
SU = tonline , with tof f being the wall time needed to run the full-order simulations and export
the snapshots and tonline the corresponding online time. In the cross-validation test considered
for Uin = 7.75 m/s, the value of tof f is 1540.766 s. This corresponds to the wall time of running
the simulations in parallel on six processors, along with the time needed to reconstruct the fields.
The accuracy results reported in the table are obtained with each ROM’s best combinations of
online modal truncation orders and refer to the error CL in the full time integration range [0, 8] s.
Table 8.2. Summary of the accuracy and the efficiency results for the ROMs considered in the
problem of flow around a cylinder.
The objective of the next test is to measure the accuracy of the hybrid ROM (in this case the H-
SUP-ROM was chosen) for higher Reynolds number and for longer time extrapolation intervals.
Therefore, we select the inlet parameter online sample Uin ∗
= 11.75 m/s (corresponding to Re =
117500) and choose a ROM time integration interval of [0, 10] s. This time interval features 27
periodic solution cycles, which is almost double the number of solution cycles contained in the
[0, 8] s considered in the previous case. The results include the time history of the lift coefficient
Cl in addition to its corresponding L2 relative error CL and also the peak error and the reduced
approximation of the time period.
8.6. Application of the H-SUP-ROM to Turbulent Problems 201
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(a) (b)
Figure 8.27. Lift coefficient curves for the cross-validation test done for the parameter value
Uin = 11.75 m/s for the time range [0, 10] s. The figure shows the FOM and the H-SUP-ROM lift
coefficient histories: (a) the full range is shown; (b) the last 3-s history of Cl is shown.
As in the first test, the FOM simulator was run for enough time to reach the periodic regime.
Then the simulation was extended for another 10 s, starting from a phase which is equivalent to
the one set for the offline samples. The additional 10 s was simulated with time step equal to
0.00025 s. The formulation and the reduction strategy of the H-SUP-ROM are the same as the
ones set in the first numerical test in this section. The Cl H-SUP-ROM curve is reproduced using
12 modes for the velocity and 10 modes for each of pressure, supremizers, and eddy viscosity.
Figure 8.27 shows the results for the Cl curves over the whole time range considered in plot (a).
Diagram (b) depicts a closeup of the lift curves in the last 3 s simulated. The images clearly
suggest that the approximately 50% Reynolds number increase does not affect the accuracy of
the H-SUP-ROM. They also prove the ability of the hybrid ROM to tackle long-time integration,
since at a visual investigation the ROM solution accuracy does not appear to deteriorate over
time. The value of the overall error CL (as defined in (8.77)) is 1.9486%. The peaks of the FOM
Cl curve were recovered with maximum error of 2.0526% among all 55 peaks present in the
time integration interval. Finally, the average time period computed by the FOM solver is about
0.3641 s, while the average time period approximated by the hybrid ROM is about 0.3642 s.
The results presented in this section fully characterize the accuracy and efficiency of the turbulent
ROMs developed. The final test we present is instead aimed at assessing their flexibility with
respect to the turbulent FOM. An important aim in generating these reduction algorithms was
to obtain ROMs that were sensitive to the turbulence model used at the FOM level and yet did
not require resolving the turbulent model equations at the ROM level. Thus, the present test will
evaluate the hybrid ROM’s ability to obtain accurate approximations of FOM solutions obtained
with different turbulence models. To this end, we carried out a further FOM simulation in which
the turbulence models selected were both the k- model and the SST k-ω model previously
considered. The test case was still the cylinder benchmark case. No parameter dependence was
considered in this test. So, the goal of the reduction was just to reproduce the time snapshots and
extrapolate the time evolution of the system. The Reynolds number selected is Re = 105 , and
the FOM simulation was carried out for both turbulence models until the regime solution was
fully developed. After that, snapshots were acquired covering a time window of 1.2 s and 1.6
s for the k- and SST k-ω models, respectively. The hybrid ROM chosen is the H-SUP-ROM
with the penalty method–based treatment of the boundary conditions. The H-SUP-ROM was
run for 8 s for both turbulence models. The FOM lift coefficients obtained with both closure
models are depicted in Figure 8.28, along with their ROM approximations. It is evident from the
202 Chapter 8. Finite Volume–Based Reduced Order Models for Turbulent Flows
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 8.28. The lift coefficient curves obtained using both the k- and SST k-ω turbulence mod-
els and the H-SUP-ROM ones. The case considered is nonparametrized, with Uin = 10 m/s corresponding
to Re = 105 . The plot is for the time range t ∈ [6, 8]; the H-SUP-ROM achieved relative L2 errors (over
the range t ∈ [0, 8]) less than 5% in both cases.
graph that the H-SUP-ROM proves sensitive to the specific turbulence model used in the FOM
solver, although no additional PDEs for the turbulent quantities are solved at the reduced level.
The value of the errors defined in (8.77) for both turbulence models are well below 5%, and the
reduced approximation of the average time period is accurate in both cases, with relative errors
below 1%.
Chapter 9
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Nonintrusive Data-Driven
Reduced Order Models
in Computational Fluid
Dynamics
9.1 Introduction
In the previous chapters we discussed a wide range of aspects of the reduced basis (RB) ap-
proach, presenting different applications and different couplings with the traditional discretiza-
tion methods for the resolution of partial differential equations (PDEs). Such applications have
in common that, independent of the solver, the construction of the reduced order model (ROM)
requires knowledge of the actual mathematical model and access to the operators describing the
PDE in order to project them onto the reduced space.
This projection is often impossible when dealing with commercial software, since the source
code is not accessible by the user, making the standard intrusive ROM approach infeasible. More-
over, nonlinear terms are particularly challenging to treat with classical projection-based meth-
ods, and the computational gain obtained by dimensionality reduction is not as remarkable as for
linear problems.
To address these issues, nonintrusive data-driven methods can be used. This approach retains
the typical offline-online decomposition, but instead of computing the reduced order operators,
it computes the low-dimensional representation of the snapshots. These new latent variables are
computed in an equation-free and data-driven framework, making this approach very versatile
and suitable for model discovery, system identification, and efficient future state predictions, and
in general for dealing with experimental data. These features allowed the spread of nonintrusive
ROMs in many industrial contexts, since they can be easily integrated into existing computational
pipelines without disrupting the industrial workflow, while reducing the overall computational
burden [517, 508, 571].
This chapter focuses on such nonintrusive methods, presenting in Section 9.2 the general
framework for parametric ROMs in a data-driven fashion and in Section 9.3 the dynamic mode
decomposition (DMD) technique for reducing dynamical systems. Lastly, we provide in Sec-
tion 9.4 a summary of the numerical results obtained by applying the methods to three different
engineering problems.
203
204 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
ROMs
In this section we present a general framework for nonintrusive data-driven ROMs. We assume
we do not know the equations behind the computation of the solution snapshots, or, even if we
know the model being used, we do not have access to the solver, so we cannot perform intrusive
model order reduction.
From an abstract point of view we can identify three main steps to constructing this kind
of ROM: create the solutions database, compute a basis of reduced dimension, and choose a
regression method to reconstruct the solution manifold. As we can already see, there are multiple
choices for every one of these steps. We can use, e.g., different sampling methods to collect
the snapshots, linear or nonlinear techniques for modes selection, and finally any method from
polynomial interpolators to artificial neural networks to reconstruct the parameters into a modal
coefficients map.
The framework described in this section, with its modular structure, can be found in the
open-source Python package4 called EZyRB [158]; the reader can also refer to Chapter 19.
n represents the number of degrees of freedom of our system, while p is the number of input
parameters. We remark that the snapshots can represent data coming from experiments, from
simulations, or even from sensors in real time. We arrange the snapshots by column in X as
| | |
X = x1 x2 . . . x m . (9.1)
| | |
There are several possible initial sampling strategies for selecting the parameters’ location,
depending on the specific task for which we construct the ROM. For example, to cover as much
of the parameter space as possible the most used approaches are Latin hypercube sampling (LHS)
[555]; Sobol sequences [544], which are quasi-random low-discrepancy sequences; or other
space-filling strategies, such as dynamic propagation sampling [339]. If the parameters repre-
sent data coming from sensors or other components with an associated probability distribution, a
weighted approach is suggested (see Chapter 12). For optimization tasks, any a priori knowledge
about the possible region of interest and/or the simulated process should be incorporated within
the sampling method. We remark that some of the aforementioned sampling schemes will not
work if the parameters can assume only a finite set of possible values.
4 Available at https://github.com/mathLab/EZyRB/.
9.2. A General Framework for Nonintrusive Parametric ROMs 205
for some modes ϕk ∈ Rn and for some modal coefficients cki ∈ R. We can rewrite (9.2) in
matrix form, arranging all the data by column, as done in equation (9.1) obtaining
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
| | | | | | | | |
x1 x2 . . . xm ≈ ϕ1 ϕ2 . . . ϕr c1 c2 . . . cm (9.3)
| | | | | | | | |
X = U ΣV ∗ , (9.4)
where ∗ denotes the conjugate transpose, and retain only the first r components. The columns of
U are the so-called POD modes. They are orthogonal and they span the optimal low-dimensional
subspace in the least squares sense. They can also be computed as the eigenvectors (up to a nor-
malization factor) of the Gram matrix X T X. The diagonal matrix Σ = diag(λ1 , . . . , λr ) is
composed by the singular values, arranged in descending order, which indicate the energy con-
tribution of the corresponding modes. The error introduced by the truncation r can be measured
as [470]
kX − U ΣV ∗ k22 = λ2r+1 , (9.5)
v
u m
u X
∗
kX − U ΣV kF = t λ2i , (9.6)
i=r+1
where the subscripts 2 and F refer to the Euclidean norm and the Frobenius norm, respectively.
The singular values λi are arranged in descending order. In Figure 9.1 we can see an example of
POD modes for a flow past a cylinder modeled with a parametrized Navier–Stokes problem.
We remark that given the first r modes, we only need r coefficients to reconstruct the whole
state xi . To compute the modal coefficients, we project the data onto the POD subspace:
C = U T X. (9.7)
Given these coefficients we can only reconstruct the original snapshots which make up the initial
database. For a fast and accurate prediction of the state x∗ , corresponding to an unseen parameter
µ∗ , we need to reconstruct the parameters–to–modal coefficients map. Then by performing a
matrix multiplication with the POD modes we obtain the whole field of interest. This method
is called POD with interpolation (PODI). In the next subsection we show some possible choices
for reconstructing such a map.
206 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 9.1. First three POD modes obtained from the resolution of a parametrized Navier–Stokes
problem, describing the flow past a cylinder.
x∗ = U s(µ∗ ). (9.8)
To build the interpolant s we can use several techniques, such as linear interpolation [519,
213], nonlinear Gaussian process regression (GPR) [622, 252, 431, 160], multifidelity methods
[494, 575], radial basis function (RBF) interpolation [82, 571], inverse distance weighting (IDW)
[623, 31], and artificial neural networks (ANNs) [602, 448], to cite a few. Here we will not review
all the possibilities; some of them are extensively presented in Chapter 17 in the context of shape
morphing, while in Chapter 20 an ANN approach is presented. We will focus only on GPR due
to its versatility and spread and to make clear the applications in the following sections.
We start by assuming the m input-output pairs to be noise-free observations. We assign
a prior to s with mean m(µ) and covariance function k(µ, µ0 ; θ), which means that s(µ) ∼
GP(m(µ), k(µ, µ0 ; θ)). Our beliefs about the function before looking at the observed values are
expressed by the prior. We consider zero mean GP, i.e., m(µ) = 0. We define the covariance
matrix K ∈ Rm×m as Ki,j = k(µi , µj ; θ). To make predictions using the Gaussian process we
9.3. DMD for Time-Dependent Problems 207
need to find the optimal values for the hyperparameter vector θ by maximizing the log likelihood:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
1 1 m
log p(c|µ, θ) = − cT K −1 c − log |K| − log 2π. (9.9)
2 2 2
Let µ∗ be the test samples and Km∗ = k(µ, µ∗ ; θ) be the matrix of the covariances evaluated
at all pairs of training and test samples. In a similar way, K∗m = k(µ∗ , µ; θ) and K∗∗ =
k(µ∗ , µ∗ ; θ). The predictions s∗ are obtained by conditioning the joint Gaussian distribution on
the observed values and by sampling the posterior as
s∗ |µ∗ , µ, c ∼ N (K∗m K −1 c, K∗∗ − K∗m K −1 Km∗ ). (9.10)
This solution minimizes the error in the Frobenius norm kX 0 − AXkF . Since the usual CFD
problem is characterized by high-dimensional state snapshots with n m, we cannot solve this
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
X ≈ U ΣV ∗ , (9.14)
A = X 0 V Σ−1 U ∗ . (9.15)
From a computational point of view it is more efficient to project the operator A onto the POD
modes, obtaining its low-rank representation
à = U ∗ AU = U ∗ X 0 V Σ−1 . (9.16)
With this operator we can express the evolution of the low-dimensional projection of the snap-
shots as x̃k+1 = Ãx̃k , with x̃k = U ∗ xk . The final step is to compute the eigenpairs of à with its
eigendecomposition
ÃW = W Λ, (9.17)
where Λ stands for the diagonal matrix of the eigenvalues and the columns of W are the eigen-
vectors. We remark that the eigenvalues of A are given by Λ, while the eigenvectors Φ, also
called exact DMD modes [588], are given by
Φ = X 0 V Σ−1 W. (9.18)
In the literature can also be found the so-called projected DMD modes [527] expressed by Φ =
U W , which tend to converge to the exact ones if the matrices in (9.11) share the column space.
The ith eigenvalue λi provides information about the frequency ωi of the corresponding
mode, defined as
Im(log λi )
ωi = , (9.19)
2π∆t
and also about its stability. In Figure 9.2 we can see the DMD modes corresponding to different
eigenvalues for a toy problem: if the eigenvalue is on the unit circle we have a stable mode; if it
is inside the unit circle it is convergent; otherwise it diverges.
Finally, to predict the future state for every time, we use
r
X
x(t) ≈ φk eψk t bk = ΦeΨt b, (9.20)
k=1
where Ψ = diag(ψk ) = diag(ln(λk )/∆t) and bk denotes the initial amplitude of the corre-
sponding kth DMD modes φk . To compute amplitudes we use the best fit in the least-squares
sense:
b = Φ† x1 , (9.21)
where we use the first snapshot x1 . We remark that this can be done with respect to every snap-
shot, and of course it will improve the accuracy in the neighborhood of that specific time instant.
The amplitudes can also be computed by minimizing the error over all the snapshots [315].
9.3. DMD for Time-Dependent Problems 209
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 9.2. In the left column are the eigenvalue positions with respect to the unit circle, and in the
right column are the corresponding DMD modes. From top to bottom we have a stable, a convergent, and
a divergent mode, corresponding to an eigenvalue on the unit circle, inside it, and outside it, respectively.
When dealing with actuated systems, DMD fails to properly reconstruct the underlying dynamics
because it is unable to incorporate the contribution of the forcing term. DMD with control
(DMDc) [463] has been developed to overcome this issue. It includes the actuation snapshots in
the analysis and thus it is able to provide reduced order representations for input-output systems.
The DMDc method quantifies the effect of control inputs on the state of the system and computes
the underlying dynamics without being confused by the effect of external control.
In mathematical terms, and using the notation introduced in the previous subsection, we
collect the input control snapshots in the matrix I as
| | |
I = u1 u2 . . . um−1 . (9.22)
| | |
210 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
Given the state xk for the kth time instant and the corresponding control uk , we seek the follow-
ing approximation for the successive state xk+1 :
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Multiresolution DMD
Multiresolution DMD (mrDMD) [354] integrates multiresolution analysis into the classical DMD.
This method allows one to handle transient phenomena and structures which rotate or translate
in the domain, overcoming a known weakness of every SVD-based method.
The main idea is to iteratively select the slow modes for a given time window and then split
each subdomain along the temporal axis. This results in the slowest modes being retained at the
first level, while the fastest are in the last level. We can see the different performance of DMD
and mrDMD with a simple example where the dataset contains many features at varying time
scales, like oscillating sines and cosines, one-time events, and random noise. First we apply
the classical DMD without any SVD rank truncation, and then we try to reconstruct the data.
In Figure 9.3, in the middle panel, the results clearly show that all the transient time events are
missing. Applying instead mrDMD, we are able to properly reconstruct the whole dataset thanks
to its hierarchical nature (right panel of Figure 9.3).
For an extension and generalization of this particular method, the interested reader can refer
to [177].
Figure 9.3. On the left is the initial dataset containing features at different time scales. In the
middle is the DMD reconstruction, and on the right is the mrDMD reconstruction.
Compressed DMD
In the case of very high dimensional data, DMD can be computed on a compressed version of
the data, thus achieving great computational savings. This extension is called compressed DMD
(cDMD) [183].
9.3. DMD for Time-Dependent Problems 211
The algorithm is very similar to the classical one, with the main difference just before the
computation of the modes, where we use a measurement matrix to compress the snapshot data.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
This is very useful, e.g., when dealing with video streams or in general when we have a prescribed
set of sensors. Despite the reconstruction error being slightly higher, the computational savings
are remarkable. Thus this DMD extension is useful even when we do not have any sensors, and
the compression matrix is randomly generated. In Figure 9.4 we compare the reconstruction
performance of DMD and cDMD.
Figure 9.4. On the left is the initial dataset, in the middle is the DMD reconstruction, and on the
right is the cDMD reconstruction.
In order to access the computational savings, we perform a simple test with an increasing
dimension of the snapshot matrix and compare the time needed to compute the DMD and the
cDMD modes, respectively. The results are shown in Figure 9.5. For further details, please refer
to tutorial6 number 4 of the PyDMD Python package.
14 Exact DMD
Compressed DMD
12
10
8
Seconds
6
4
2
0
0 20000 40000 60000 80000 100000
Snapshots dimension
Figure 9.5. Computational savings of cDMD with respect to the exact DMD for an increasing
dimension of the snapshot matrix.
In this section we present different engineering problems where we applied the nonintrusive
ROMs introduced in the previous sections.
PyDMD
DMDPyDMD
by PyDMD ⎡ | | ⎤ GA by DEAP
⎢ ⎥
= 0 …
⎢ ⎥
⎣ | | ⎦
APPROXIMATED
Validation and enrichment OPTIMUM
∗ ∗
( , )
ROM
Figure 9.6. Representation of the computational pipeline used for the shape optimization problem.
The naval hull is parametrized by applying the free-form deformation (FFD) technique [533]
(see also Chapter 17) to the region of interest of the ship. A visual example of the deformations
obtained by morphing the bulb is presented in Figure 9.7, where the preservation of surface
derivative continuity is evident.
Once the deformation map has been constructed and the parameter bounds are set, we sample
the parameter space to obtain a limited set of deformed configurations, which are used to create
the database of high-fidelity snapshots. To compute such snapshots we numerically solve the full-
order model: since for these applications a Reynolds-averaged Navier–Stokes (RANS) model
9.4. Applications in CFD 213
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 9.7. Example of bulbous bow deformations using the FFD method.
with VoF and turbulence is used, we apply a finite volume discretization using the OpenFOAM
library [612]. As already mentioned, the computational cost of these simulations is reduced by
involving the DMD to accelerate the convergence to regime state. The finite volume simulation
is carried out for a limited set of time steps, collecting the system states and using them to feed
the DMD algorithm. Assuming convergent or periodic behavior, DMD is able to capture such
dynamics and we can exploit the linear operator to predict a future state. Again, we remark that
this procedure is applied to all parametric samples independently.
After DMD, we obtained a snapshot for all the parametric samples and for any field of inter-
est. For this particular application, we collected the pressure, velocity, and VoF variable fields,
since they are required in order to compute the objective function. We applied the PODI algo-
rithm by decomposing the snapshot matrix by means of SVD and by approximating the solution
manifold in the reduced space. We construct this approximation using an RBF interpolation
[154] or a GPR [152]. From the latter work, we report in Figure 9.8 a summarizing comparison
between the different approximation techniques we tested: linear interpolator, RBF (with two
different values of the smoothing parameter, which controls the smoothness and thus the interpo-
latory nature of the regression), and GPR. GPR and (smoothed) RBF are the methods that show
the smallest error, but a remarkable difference is the higher ROM accuracy for GPR with few
input snapshots, allowing for a computationally cheaper offline phase.
The PODI algorithm was applied for the three aforementioned fields, resulting in a faster
evaluation of the total drag for any new geometrical deformation. Finally, this parametric model
can be used within an optimization cycle: the objective function is not computed any more on
0.1 0.1
relative L2 error
8 · 10−2
8 · 10−2
6 · 10−2
6 · 10−2
0 20 40 60 80 20 40 60 80
# modes # snapshots
GP linear RBF0 RBF0.01
Figure 9.8. Comparison of the average relative L2 error of the ROM in a naval shape optimiza-
tion problem as a function of the number of modes (left) and the number of snapshots used (right).
214 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
the simulation output, but using the PODI. The computational gain is huge: instead of requiring
several hours (using a high-performance computing facility) for a single iteration, the PODI
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
lasts only a few seconds for a new query. The aforementioned references [154, 152] show the
successful use of the genetic algorithm for the optimization task. The large number of evaluations
usually required by such evolutive procedures is compensated for by the extremely fast response.
We remark again that, thanks to the generality of the model, any optimization method can be
used as well.
Due to the missing error estimation, we have validated the optimal shape resulting from the
optimization procedure using the FOM. If the discrepancy error is above a certain threshold, we
refine the solution manifold by adding to the solution database the snapshot computed during the
validation. The ROM is then enriched with additional information, resulting in a more accurate
output: iteratively, a new optimization is performed using the new model until the PODI error
results are acceptable.
Data acquisition
High-fidelity fluid dynamics FOM in parameter space
Basis functions
Modal decomposition using SVD
Construct ROM
DMD/POD + continuous modal coefficients
Midcast
Interrogate intermediate snapshots in time as parameter.
Acoustic evaluation
Analyze acoustic performance from prediction data.
spanned by POD modes are used to construct an interpolant function. Also in the DMD case
the operators of the dynamical system are constructed using only the training set. In both the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
DMD and the PODI cases the accuracy of the ROM procedure is checked on the testing set. The
FOM is based on the incompressible Navier–Stokes equations coupled with a LES filtering tech-
nique, and the physical problem consists of the flow around a sphere of diameter D = 0.01 m,
immersed in a water stream with constant streamwise velocity U0 = 1 m/s. The kinematic
viscosity is ν = 2.0 × 10−6 m2 /s, so that the Reynolds number based on the sphere diameter
is Re = 5000.
The computational domain, depicted in Figure 9.10, consists of a box with dimensions 16D×
16D × 16D along the x-, y-, and z-axes, respectively. The sphere is centered along the y- and
z-axes. Along the x-axis it is at distance 12D from the outlet boundary. The boundary conditions
for the pressure field are set to zero-gradient on all boundaries with the exception of the outlet
boundary, where we set a uniform null value. The velocity is set to U0 at the inlet, the stress-free
condition is set at the lateral boundaries, and the zero-gradient condition is set for the velocity
components at the outlet. The grid, unstructured and body-fitted, consists of about five million
cells. Since no wall functions are adopted, the grid spacing normal to the wall for the densest
layer of
pcells (indicated by R5) has its first cell center within a wall unit y (y = uτ y/ν with
+ +
uτ = τw /ρ0 and τw the mean shear stress). The grid has five different refinement zones, as
depicted in Figure 9.10.
Figure 9.10. Sketch of the computational domain used for the FOM. The sphere diameter is
D = 0.01 m. Successive mesh refinement layers (R2, R3, R4) are performed using the cell splitting
approach until the finest grid spacing 0.001D is reached in the region R5.
The time integration is carried out using a constant time step of ∆t = 10−5 s, which is
enough to keep the Courant number under the threshold of 0.5. The flow around the sphere is
completely developed after about 80 characteristic times D/U0 .
Once the snapshots have been acquired, the techniques reported in Sections 9.2 and 9.3 are
used for PODI and DMD, respectively. In particular, for the PODI case we used a cubic spline
interpolation approach, while for the DMD case we used the classical algorithm. We then used
the PODI and DMD approaches to reconstruct the snapshots of the testing set, and we compared
216 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Modal Index
Figure 9.11. Normalized singular values for streamwise velocity and pressure snapshots.
the results with the high-fidelity data by changing the dimension of the POD and DMD spaces.
The eigenvalue decay resulting from the SVD of the training snapshots matrix is reported in Fig-
ure 9.11. As one can deduce from the picture, the problem is not characterized by a sudden drop
in the eigenvalues. This suggests the presence of a large number of different spatial scales. This
feature is a particular characteristic of high-fidelity LES simulations. A visual representation of
the identified POD and DMD modes for different index modes is reported in Figures 9.12 and
9.13 for velocity and pressure, respectively. In order to assess the accuracy of the procedure for
predicting hydroacoustic analysis, we postprocessed the reconstructed results using both a dipole
and a quadrupole approach [125, 126] and we compared the results coming from full-order snap-
shots. Figure 9.14 shows a quantitative comparison of the error of changing the dimension of
the POD and DMD spaces. In the picture we depict the time evolution of the L2 relative er-
ror for both the velocity and the pressure fields. As expected, for both approaches, increasing
the number of employed modes decreases the error. For this particular case the PODI approach
performs the best for both fields. Since the aim of this test was to assure the capability of this
type of ROM to be used for hydroacoustic analysis, in Figure 9.15 we compare the spectrum of
the hydroacoustic analysis for both the dipole and the quadrupole case. Both ROMs prove to be
accurate in predicting acoustic noise.
Figure 9.12. Isosurfaces of the DMD modes (left) at a value of −15 × 10−5 m/s versus the
isosurfaces of POD modes (right) at a value of 6.5 m/s for the streamwise velocity field. Modes: 1, 2, 8,
and 36.
9.4. Applications in CFD 217
Figure 9.13. Isosurfaces of the DMD modes (left) at a value of −5 × 10−4 Pa versus the isosur-
faces of POD modes (right) at a value of 300 Pa for the pressure field. Modes: 1, 2, 8, and 36.
Figure 9.14. Relative error percentage in the Frobenius norm for velocity and pressure (top and
bottom, respectively) fields, where half-sample-rate snapshots are used to train the reduced model, DMD
(left) or PODI (right), for a particular truncation (r) while predicting the intermediate snapshots. Plots
show only the snapshotwise prediction error, while disregarding errors in the training set, which are almost
null.
218 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 9.15. Linear dipole (top) and nonlinear quadrupole (bottom) terms of FWH equation
evaluated from LES data and compared to corresponding DMD (left) and PODI (right) reduced models at
different truncation ranks (r).
The physical problem is the one of flow around a wing profile. The geometry of the wing
is parametrized using the parameter vector µ ∈ R10 , and the undeformed configuration corre-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
sponds to the four-digit NACA 4412 wing profile. The deformed configuration is obtained by
summing five bump functions on the upper and lower part of the profile [283]. Defining the upper
and lower y coordinates of the wing profile by yu and yl and the upper and lower coordinates in
the undeformed configuration by yu and yl , the deformed configuration is given by
5
X 5
X
yu = yu + µi ri , yl = yl − µi+5 ri , (9.25)
i=1 i=1
where ri are bump functions as defined in [283], which are also reported in Figure 9.19. The
parameters used for the training are obtained with a uniform probability distribution in the pa-
rameter space D := [0, 0.03]10 ⊂ R10 . Once the coordinates of the deformed configuration are
obtained, the mesh is deformed using an RBF interpolation technique. We refer the reader to
[573] for more details on the deformation approach. The high-fidelity computational model is
based on the incompressible Navier–Stokes equations with RANS turbulence modeling. Specifi-
cally, the Spalart–Allmaras turbulence model has been employed [546]. The discretized problem
is obtained by means of the finite volume method using the OpenFOAM [612] and the ITHACA-
FV volume packages [550]. A schematic view of the computational domain, together with a
zoom on the mesh near the wing profile, is reported in Figure 9.17. The mesh counts 46500
hexahedral cells. At the inlet we have a constant and uniform velocity uin = 1 m/s. The kine-
matic viscosity equals ν = 2 × 10−5 m2 /s, which considering a chord length of D = 1 m,
produces a Reynolds number of Re = 50000. The pressure velocity coupling was resolved using
the PIMPLE algorithm as implemented in the OpenFOAM library. To advance the simulations
in time we used a constant time step of ∆t = 10−3 s. Each configuration is simulated for a total
time of T = 30 s. In Figure 9.18 we report the temporal evolution of the lift coefficient for nine
different values of the input parameter.
7.5C C 16C
16C
24.5C
Figure 9.17. Sketch of the computational domain used to solve the fluid dynamics problem in its
reference configuration. The left plot is a zoom on the mesh near the wing. The right picture is a schematic
view of the domain with the main geometrical dimensions.
The possibility of exploring the solution manifold with a limited number of simulations during
the training stage is highly correlated with the number of parameters. By increasing the number
of parameters it is in fact easy to fall victim to the curse of dimensionality. It is then crucial to
keep the number of parameters as low as possible. In this section we show a possible approach
220 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 9.18. The temporal evolution of the lift coefficient from 1 s to 30 s for nine different parameters.
NACA profile
0.2 r1
r2
0.1 r3
y
r4
0.0 r5
−0.1
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
x
Figure 9.19. Airfoil shape functions with respect to the profile abscissa. The leading edge corre-
sponds to x = 0. The bump functions are rescaled by a factor of 0.2 for illustrative reasons.
to reducing the number of parameters using the property of active subspaces (ASs) [134]. In
Figure 9.20 we show the DyAS results for the lift coefficient output that aim to show the effec-
tiveness of the AS over time. The left diagrams show, at different time instants, the lift coefficient
at each sample point as a function of the first active variable, which is obtained by a linear combi-
nation of the sample point coordinates in the parameter space. We underline that the eigenvector
components of all the time instants presented, corresponding to the coefficients c1 , c5 , d1 , and
d5 , are almost zero. This means that on average the lift coefficient is almost flat along these
directions. We will exploit this fact by freezing these parameters and constructing a GPR on a
reduced parameter space.
With the previous analysis we were able to discover the presence of several input parameters
with minimal influence on the output function (the lift coefficient). This information was used
to construct a surface response which is based only on the relevant input parameters. The ap-
proximation was constructed using a GPR technique with RBF kernel exploiting the GPy library
[239]. The regression with GPR was constructed using both the full and the reduced parameter
space, and results were compared over a test set of 100 samples. The training set is the same for
both the reduced and the full parameter space case and consists of 70 samples. For the first 20 s
9.4. Applications in CFD 221
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 9.20. On the left is the sufficiency summary plot for the lift coefficient at times t =
10.0, 14.0, 18.0 s (top to bottom). On the right are the first eigenvector components at the corresponding
parameters.
we used the high-fidelity data, while the last 10 seconds of the simulation were approximated
using DMD to speed up convergence. The DMD model was constructed using the high-fidelity
results in the temporal interval [12, 20] s. The DMD approximation was constructed using 8000
temporal pieces of information (∆t = 0.001 s) with 10 DMD modes. Figure 9.21 summarizes
the different configurations analyzed. In particular we have the GPR regression in the full param-
eter space and in the reduced parameter space, and the same results using a DMD approximation
for the last 10 s of the simulation.
222 Chapter 9. Nonintrusive Data-Driven ROMs in CFD
·10−2
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
8 GPR
GPR reduced
GPR + DMD
6 GPR + DMD reduced
Relative Error
0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
Time [s]
Figure 9.21. The relative error of the approximated outputs at different times. The relative error
is computed on 100 test samples, using the high-fidelity lift coefficient to train the regression for t ≤ 20 s,
while for t > 20 s the DMD forecasted states are used for the training.
Chapter 10
Spectral Element
Method–Based Model
Order Reduction
The spectral element method (SEM) uses high-order polynomials defined over a coarse spatial
grid to numerically resolve a function. This is in contrast to, e.g., the widely adopted finite
element method (FEM), which uses low-order polynomials but a very dense spatial discretization
grid. This chapter introduces the main concepts of model reduction with the SEM, which was
first introduced in [439]. Beginning with the polynomial spaces used as ansatz spaces, it also
shows the global assembly and points out some particularities of the technical implementations
with the partial differential equation (PDE) solver Nektar++ and the model reduction software
ITHACA-SEM. This is an introduction with several references to the literature, where the topic
is analyzed more deeply. Overview books are [96, 97, 166, 330].
Jacobi Polynomials
Jacobi polynomials Pnα,β (x) (with α, β > 1) are the solutions to the singular Sturm–Liouville
problem
d 1+α 1+β d
(1 − x) (1 − x) P (x) = λn (1 − x)1+α (1 − x)1+β Pnα,β (x), (10.1)
α,β
dx dx n
λn = −n(α + β + n + 1). (10.2)
223
224 Chapter 10. Spectral Element Method–Based Model Order Reduction
Jacobi polynomials can be computed with the Rodriguez formula and the following two
recursion relations.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Rodriguez Formula
(−1)n dn
Pnα,β (x) = (1 − x)−α (1 + x)−β n (1 − x)α+n (1 + x)β+n . (10.3)
n
2 n! dx
Recursion Relations
d α,β α,β
b1n (x) P (x) = b2n (x)Pnα,β (x) + b3n (x)Pn−1 (x), (10.11)
dx n
b1n (x) = (2n + α + β)(1 − x2 ), (10.12)
b2n (x) = n(α − β − (2n + α + β)x), (10.13)
b3n (x) = 2(n + α)(n + β). (10.14)
Orthogonality
Jacobi polynomials are orthogonal with respect to (1−x)1+α (1−x)1+β over the interval [−1, 1]:
Z 1
(1 − x)1+α (1 − x)1+β Ppα,β (x)Pqα,β (x)dx = Cδpq . (10.15)
−1
Derived Polynomials
Special cases of derived polynomials are, e.g.,
• Legendre polynomials (α = β = 0), i.e., Ln (x) = Pn0,0 (x), and
1 1
22n (n!)2 − 2 ,− 2
• Chebyshev polynomials (α = β = − 12 ), i.e., Tn (x) = (2n)! Pn (x).
Quadrature Rules
Quadrature rules approximate the integral of a function in the elemental domain (here the interval
[−1, 1]) as
Z 1 Q−1
X
u(ξ)dξ ≈ wi u(ξi ). (10.16)
−1 i=0
In contrast to the FEM, the integrand u(ξ) is a high-order polynomial and thus requires
quadrature rules which are sufficiently accurate.
There are three different types of Gauss quadrature, known as Gauss, Gauss–Radau, and
Gauss–Lobatto. The difference between the three types of quadrature lies in the choice of the
10.1. Basic Notions and Functions of the SEM 225
zeros. Gauss quadrature uses zeros which have points that are interior to the interval, −1 < ξi <
1 for i = 0, . . . , Q − 1. In Gauss–Radau the zeros include one of the endpoints of the interval,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
usually ξ = −1, and in Gauss–Lobatto the zeros include both endpoints of the interval, i.e.,
α,β
ξ = +1, −1. Let ξi,P denote the P zeros of the P th-order Jacobi polynomial PPα,β (x).
Gauss–Legendre Quadrature
0,0
ξi = ξi,Q , i = 0, . . . , Q − 1, (10.17)
−2
2 d
wi0,0 = (L Q (ξ))| ξ=ξi , i = 0, . . . , Q − 1. (10.18)
1 − (ξi )2 dξ
Gauss–Radau–Legendre Quadrature
−1, i = 0,
ξi = 0,1 (10.19)
ξi−1,Q−1 , i = 1, . . . , Q − 1,
1 − ξi
wi0,0 = , i = 0, . . . , Q − 1. (10.20)
Q2 [L Q−1 (ξi )]
2
Gauss–Lobatto–Legendre Quadrature
−1, i = 0,
ξi = ξ 1,1 , i = 1, . . . , Q − 2, (10.21)
i−1,Q−2
1, i = Q − 1,
2
wi0,0 = i = 0, . . . , Q − 1. (10.22)
Q(Q − 1)[LQ−1 (ξi )]2
Aliasing
Aliasing occurs when the integrand is of higher order than the quadrature rule can resolve. An
approximation error will thus occur. In many implementations only the linear terms are integrated
exactly. Nonlinear terms, like in the Navier–Stokes equations, are then underresolved in favor of
computational speed.
chosen and the numerical efficiency and the condition numbers of resulting linear systems are
considered. In particular, a boundary-interior decomposition is used in the SEM, where “bound-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
ary” also refers to internal boundaries given by the h-refinement and not just the boundary of the
actual domain.
Two frequently used terms in the field of SEM are nodal and modal expansions.
Modal Expansions
Given a set of polynomial functions {Φp , 0 ≤ p ≤ P }, a modal expansion set refers to a hierar-
chical expansion set, i.e., the expansion set of order P contains the set of order P − 1.
An example of a modal expansion is the set of monomials {xp , p = 0, . . . , P }. It is called
the moment expansion.
Nodal Expansions
In a nodal expansion {Φp , 0 ≤ p ≤ P }, the expansion coefficients are the values at certain nodes
{xp , 0 ≤ p ≤ P }, i.e., Φp (xq ) = δpq .
An example of a nodal expansion is the Lagrange polynomials
ΠP
q=0,q6=p (x − xq )
Φp (x) = , p = 0, . . . , P. (10.23)
ΠP
q=0,q6=p (xp − xq )
Let the Jacobi polynomials be denoted by Pnα,β (x). The reference quadrilateral in reference
coordinates ξ1 , ξ2 is defined with the midpoint (ξ1 , ξ2 ) = (0, 0). The vertices, i.e., corners are
(ξ1 , ξ2 ) = (−1, −1), (ξ1 , ξ2 ) = (−1, +1), (ξ1 , ξ2 ) = (+1, −1), and (ξ1 , ξ2 ) = (+1, +1).
A modal basis over the reference quadrilateral is given by vertex-related expansion bases
1 − ξ1 1 − ξ2
Φ= , (10.26)
2 2
1 − ξ1 1 + ξ2
Φ= , (10.27)
2 2
1 + ξ1 1 − ξ2
Φ= , (10.28)
2 2
1 + ξ1 1 + ξ2
Φ= ; (10.29)
2 2
edge-related expansion bases
1 − ξ1 1 + ξ1 1,1 1 − ξ2
Φ= Pp−1 (ξ1 ) , 0 < p < P1 , (10.30)
2 2 2
1 − ξ1 1 + ξ1 1,1 1 + ξ2
Φ= Pp−1 (ξ1 ) , 0 < p < P1 , (10.31)
2 2 2
1 − ξ1 1 − ξ2 1 + ξ2 1,1
Φ= Pp−1 (ξ2 ), 0 < p < P2 , (10.32)
2 2 2
1 + ξ1 1 − ξ2 1 + ξ2 1,1
Φ= Pp−1 (ξ2 ), 0 < p < P2 ; (10.33)
2 2 2
10.2. Assembly of System Matrices 227
1 − ξ1 1 + ξ1 1,1 1 − ξ2 1 + ξ2 1,1
Φ= Pp−1 (ξ1 ) Pp−1 (ξ2 ), 0 < p < P1 , 0 < q < P2 . (10.34)
2 2 2 2
More explicitly given expansion bases can be found in Appendix D of [330].
A change to cylindrical coordinates shows that for ξ1 = −1, ξ2 = 1, we have η1 ∈ [−1, 1].
L(u) = 0, (10.37)
Galerkin Orthogonality
The test space is orthogonal to the residual, which is the interpretation of (10.39).
for discrete functions uδ in the Sobolev space H 1 , which implies continuity, i.e., uδ ∈ C 0 . The
mapping between reference simplex and actual element is denoted by χe (ξ).
Since global modes Φi (x) can be expressed in terms of local expansions φe (ξ), it holds that
Ndof Nel X
P Nel X
P
X X X
uδ (x) = ûi Φi (x) = ûep φep (ξ) = ûep φep ((χe )−1 (x)). (10.41)
i=1 e=1 p=0 e=1 p=0
The relationship between local and global degrees of freedom is expressed with a scattering
matrix A as ûl = Aûg from global to local degrees of freedom and a corresponding assembly
matrix, which maps from local to global degrees of freedom.
An important feature of ansatz functions is the sparsity pattern generated. It can be assessed
by looking at the projection problem. The projection problem reads as follows: find uδ ∈ χδ
such that
(v δ , uδ ) = (v δ , f ) ∀v δ ∈ χδ . (10.42)
This leads to the linear system M û = f with Mpg = (Φp , Φq ) and fp = (Φp , f ). For a moment
expansion, i.e., the set of monomials, the mass matrix M is 50% full; in a Lagrange expansion
the mass matrix M is full; and in a Legendre expansion the mass matrix M is diagonal.
A boundary-interior decomposition requires uδ ∈ H 1 to impose continuity across element
boundaries. But an interface-matching condition potentially couples all degrees of freedom:
P
X P
X
ûep φep (1) = ûe+1 e+1
p φp (−1). (10.43)
p=0 p=0
Instead, a division into boundary modes and interior modes, such as that given by the Lagrange
expansion
φp (−1) = 1, p = 0, (10.44)
φp (−1) = 0, p 6= 0, (10.45)
φp (1) = 1, p = P, (10.46)
φp (1) = 0, p 6= P, (10.47)
such that
ûeP φeP (1) = ûe+1 e+1
0 φ0 (−1) (10.48)
ROM work, e.g., in [457, 458, 273, 270]. The SEM solver used in the work presented here is the
spectral/hp element software Nektar++ [95].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
is solved for the next iterate uk+1 = u. A common stopping criterion is that the relative change
between iterates in the H 1 -norm falls below a given tolerance. The initial solution u0 (ν0 ) is
computed by time advancement of (10.49)–(10.50) from zero initial conditions at a parameter
value ν0 . From this starting point, solutions on the whole parameter domain can be found.
The discretized system solved in each step of the Oseen iteration is given by (10.58) as
T
A −Dbnd B vbnd fbnd
−Dbnd 0 −Dint p = 0 , (10.58)
T T vint fint
B̃ −Dint C
where vbnd and vint denote velocity degrees of freedom on the boundary and in the interior,
respectively. The forcing terms fbnd and fint refer to the boundary and interior, respectively.
The matrix A assembles the boundary-boundary terms, B the boundary-interior terms, B̃ the
interior-boundary terms, and C the interior-interior terms of elemental velocity ansatz functions.
In a Stokes system, B = B̃ T , but this is not the case here for the Oseen equation, since the
230 Chapter 10. Spectral Element Method–Based Model Order Reduction
linearization term (uk · ∇)u is present in (10.54). The matrices Dbnd and Dint assemble the
pressure-velocity boundary and pressure-velocity interior couplings, respectively.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
The linear system (10.58) is assembled in local degrees of freedom, leading to block matrices
A, B, B̃, C, Dbnd , and Dint , with each block referring to a single spectral element. It follows that
the system is singular in this form. To solve the system, the local degrees of freedom need to be
gathered into the global degrees of freedom. Since C contains the interior-interior contributions,
it is invertible and the system can be statically condensed into
A − BC −1 B̃ T BC −1 Dint
T T
− Dbnd 0 vbnd
Dint C −1 B̃ T − Dbnd −1 T
−Dint C Dint 0 p
B̃ T T
−Dint C vint
fbnd − BC −1 fint
Taking the top left 2×2 block and reordering the degrees of freedom, such that the mean pressure
mode of each element is inserted into the corresponding block of Â, leads to
fˆbnd
 B̂ b
= , (10.60)
Ĉ D̂ p̂ fˆp
with D̂ invertible. This means that a second level of static condensation can be used. It follows
that
The vector b contains the velocity boundary degrees of freedom and the mean pressure modes,
while the remaining solution components are computed by reverting the transformations of the
static condensations. The largest computational effort is in solving the final system:
The matrices C and D̂ also have to be inverted, which because of the elemental block struc-
ture requires inverting submatrices of the size of the degrees of freedom per element in each
submatrix.
The offline-online decomposition allows for fast input-output solves, because they are inde-
pendent of the original model size Nδ . It is an important part of efficient ROM, but since the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
static condensation includes the inversion of the parameter-dependent matrix C, the projection is
applied to the first system, (10.58). Other options are possible by gathering part of (10.58) into
global degrees of freedom; see [276]. In the offline phase, snapshot solutions have been gathered
over the parameter domain, which now serves as a projection space to define the reduced order
setting. To have fast reduced order solves, the offline-online decomposition expands (10.58) in
the parameter of interest and stores the parameter-independent projections as small matrices of
order N × N . Since during the Oseen iteration each matrix is dependent on the previous iterate,
the submatrix corresponding to each basis function is assembled and then formed online with
the RB coordinate representation of the current iterate. This is analogous to reduced order as-
sembly of the nonlinear term in the Navier–Stokes equations. For more details and applications,
see [273].
Consider a variable kinematic viscosity in the interval ν ∈ [0.05; 10] for a channel flow, as
depicted in Figures 10.1 and 10.2. A parabolic inflow profile is prescribed on the left wall at y =
0, a natural outflow boundary is prescribed at y = 8, and the other walls are no-slip boundaries.
Figures 10.1 and 10.2 show the extreme cases considered here for very small viscosity and very
large viscosity.
Figure 10.1. Full-order, steady-state solution for ν = 10: velocity in the x direction (top) and y
direction (bottom).
Sample solutions are computed at 13 values of ν in the interval of interest. The POD is
computed and the POD energy reaches a threshold of 99% with three modes and a threshold of
99.99% with seven modes. Ideally, the exponential decay in POD energy should translate into
an exponential decay in relative approximation error. Figure 10.3 shows the mean and maximum
relative L2 (Ω) error in the velocity as an increasing number of POD modes is considered. A
mean relative error of 1% is reached with four basis functions, and a maximum relative error of
1% is reached with five basis functions.
232 Chapter 10. Spectral Element Method–Based Model Order Reduction
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 10.2. Full-order, steady-state solution for ν = 0.05: velocity in the x direction (top) and
the y direction (bottom).
Mean error
Maximum error
relative L2 (Ω) error in velocity
−1
10
10−3
10−5
10−7
10−9
0 2 4 6 8 10 12
ROM basis size
Figure 10.3. Mean and maximum relative L2 (Ω) error in the velocity with increasing basis size.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 11
Discontinuous
Galerkin-Based Reduced
Order Models
11.1 Introduction
In this chapter the theory and applications of intrusive reduced order methods based on the
discontinuous Galerkin method (DGM) are presented. We focus on presenting the numerical
method and on showing how techniques which are used for the finite volume method (FVM)
and the finite element method (FEM) can also be employed in this framework in order to design
reduced order models (ROMs).
The DGM is a numerical method for the approximation of partial differential equations
(PDEs) and it is well suited for conservative laws and, in particular, for predicting phenom-
ena in which discontinuities and steep gradients are involved. For example, convection-diffusion
problems may present boundary layers in the immediate proximity of bounding surfaces char-
acterized by steep gradients. Propagating shock waves with sharp discontinuities may appear
in the solutions of nonlinear conservation laws, e.g., in the case of the Navier–Stokes equations
describing compressible flows.
The DGM can be considered as a generalization of the FVM and the FEM since it tries to
combine the best features of both.
Two key features of the DGM are the local conservativeness when approximating conserva-
tion laws, achieved by introducing the numerical flux, and the local formulation of the method.
Each element can be thought of as autonomous from the others, so the order of the method and
the relative number of nodes may change independent of the neighboring elements; this makes
the implementation relatively simple and suitable for p-adaptivity with less effort. In fact, unlike
the FEM, continuity is not imposed on the elemental boundaries, which in the DGM formulation
can have different nodes depending on the side considered. Moreover, all the basis functions
of the discretized finite space of the solutions have support in only one single element and the
fluxes can be evaluated with just the information from the neighboring cells that constitute the
compact stencil; this property of the DGM makes it efficient to run in parallel, since the amount
of information that has to be exchanged between elements is limited. The DGM is also suit-
able for implementing solvers for problems with complex geometries, nonstructured meshes,
and curved elements.
The main drawback of the DGM is the increment of the number of degrees of freedom used
to obtain the same accuracy as the FEM, but this additional cost is well balanced by its benefits.
233
234 Chapter 11. Discontinuous Galerkin-Based Reduced Order Models
In fact, with respect to the FEM with stabilization, the DGM introduces less numerical viscosity,
thus following the physical solution more tightly.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In addition, FEM may suffer from the Gibbs phenomenon if no advanced stabilization tech-
nique is involved, while DGM counters this problem with techniques like flux limiters. Like
the FVM, the DGM searches for piecewise-continuous solutions defined on every element of
the computational domain. This allows a better resolution of shock waves and contact discon-
tinuities. FVM, unlike DGM, uses only low-order approximations with a consequent loss of
accuracy; on the other hand, DGM, as a high-order method, is more susceptible to numerical
instabilities: underresolved simulations typically blow up.
A complete introduction to the DGM and a dissertation on its stability properties can be
found in [282].
The first formulation of the DGM dates back to 1973, proposed by Reed and Hill [483]
to solve equations which describe neutron transport. The first applications of the DGM were
mainly related to hyperbolic systems of equations. Lasaint and Raviart [357] provided a first a
priori error analysis for linear hyperbolic systems, while Cockburn and Dong [131] later analyzed
nonlinear systems. The Euler set of equations was solved using a Runge–Kutta (RK) method for
the time integration, while the discontinuities of the solution on the elemental interfaces were
treated by evaluating the fluxes between neighboring cells with the aid of techniques originally
developed in the finite volume framework. Then, Hu et al. [294] analyzed the properties of
the DGM for linear problems (i.e., wave propagation phenomena). In the mid-nineties DGM
was used to solve problems involving compressible flows; notable were the works of Bassi and
Rebay, which considered both the Euler equations [43] and the Navier–Stokes equations [44].
A complete analysis of DGM for applications related to compressible flows can be found in
[171]. Some later works by Cockburn et al. provided error estimates, numerical stability, and
convergence analysis for the DGM for convection-diffusion problems [131]. Current efforts are
focused on efficient strategies for implementations and on the development of new algorithms in
order to reduce the computational costs of DGM-based codes [441].
Up to now, only a few works have dealt with model reduction for codes based on the DGM,
and in general techniques which were originally proposed for either FVM or FEM have been
adapted to this different framework. Nevertheless, some notable works about ROMs for DGM
can be found in the literature [632, 19].
The chapter is structured as follows: the following three sections are devoted to the presen-
tation of the nodal DGM; in Subsection 11.3.2 a reduced order method is introduced; results are
presented in the last section.
∂U
+ ∇ · F (U, Q) = S, (11.1)
∂t
where U is the state variable vector, Q is the vector containing the derivatives of the state vari-
ables, F is the matrix containing the terms related to the fluxes, and S is the source term.
Many different sets of equations can be written in the form reported above, including the
Euler equations, the Maxwell equations, the Navier–Stokes equations, the linearized Navier–
Stokes equations, and the linearized Euler equations.
11.2. Nodal DGM 235
Proper initial conditions U0 and boundary conditions UΣ must be set in order to have a
well-posed problem:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
U = UΣ on Σ,
U = U0 for t = 0.
The domain Ω is approximated by a partition of nonoverlapping elements Ωk as
M
[
Ω ≈ Ωh = Ωk , (11.2)
k=1
and within each element Ωk the solution is approximated by the following nodal expansion [280]
for i ∈ {1, 2, 3}:
N
X N
X
Ukh (x, t) uk (xl , t) i Lkl (x) i = ukl (t) i Lkl (x) i , (11.3)
i
=
l=0 l=0
where N is the number of nodes inside the element Ωk ; Lkl (x) are the multivariate Lagrange
interpolation polynomials defined by the points xl in the element Ωk ; and ukl (t), the expansion
coefficients, are the values of the solution at the nodal points.
Applying a Galerkin projection in (11.1) onto each member of the basis set Lki , the weak
form of the problem can be written for each element Ωk as
∂Ukh
Z Z
k k k
Li · + ∇ · Fh dΩ = Lki · Skh dΩk , (11.4)
Ωk ∂t Ωk
where the source term is approximated using the basis sets, and likewise for the conservative
variables (11.3), and the flux term is approximated with a polynomial matrix basis Mlk (x) for
i, j ∈ {1, 2, 3}:
N
X N
X
Skh skl Lkl Fhk Flk Mlk (11.5)
i
= i i
, ij
= ij ij
.
l=1 l=1
where Σk indicates the boundary limiting Ωk and n is the outward-pointing unit normal to each
edge.
With expansions (11.3) and (11.5), setting
Nedge
k
X
Fhk ·n= (F · n)l Lkl , (11.7)
l=1
where Nedge is the number of nodes on the element edges and Lkl is the restriction of the poly-
nomial basis onto the edge, the weak formulation (11.6) reads
N k
! N
! Nedge
du
Z X Z X Z X
Lki · l
Lk dΩk − ∇Lki · Flk Lkl dΩk + Lki · Flk Lkl · n dΣk
Ωk dt l Ωk Σk
l=1 l=1 l=1
N
Z !
X
= Lki · skl Lkl dΩk , (11.8)
Ωk l=1
The standard DGM is formulated to approximate first-order derivatives of the primitive vari-
ables. The viscous components of the fluxes of the Navier–Stokes equations show a dependence
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
up to the second-order derivatives of the conservative variables. Among the several extensions
of the DGM to convection-diffusion problems with higher-order derivatives, following Yan and
Shu [630], it is possible to introduce an auxiliary Q = ∇U. The global system of equations for
the variables U and Q on the computational domain Ω is
Q − ∇U = 0 , (11.9)
∂U
+ ∇ · F (U, Q) = S . (11.10)
∂t
Similarly, (11.9) can be written in a weak formulation using the same spatial discretization
and basis functions, and one eventually ends up with the following equation:
N N
Z ! Z !
X X
Mik : qlk Mlk dΩk + k
ukl Lkl dΩk
∇ · Mi ·
Ωk l=1 Ωk l=1
Z Nedge
X
ukl Lkl · Mik n dΣk = 0 . (11.11)
−
Σk l=1
For triangular elements, the master element is the unit right triangle with vertices (ξ1 , η1 ) =
(0, 0), (ξ2 , η2 ) = (1, 0), and (ξ3 , η3 ) = (0, 1). For quadrangular elements, the master element is
a square centered at the origin of the (ξ, η) system, with sides parallel to the reference axes and
length lside = 2. The master elements are shown in Figure 11.2.
11.2. Nodal DGM 237
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
N NX
X −m
x(ξ, η) = ϑmn ξ m η n , (11.12)
m=0 n=0
N NX
X −m
y(ξ, η) = φmn ξ m η n , (11.13)
m=0 n=0
where the coefficients ϑmn and PφNmn P are the solution of a linear system which is always well
N −m
conditioned because the terms m=0 n=0 ξ m η n form a linearly independent set of bases.
The mapping expressions (11.12) and (11.13) enable a fast evaluation of the transformation
Jacobian J and its determinant ||J||. The mapping does not impose any limits on the geometric
shape of the physical cell, so it is suited for describing curvilinear elements.
Moreover, (11.12) permits one to evaluate simply the Jacobian J at each point of the element:
" #
∂x ∂x
∂ξ ∂η
J= ∂y ∂y , ||J|| = det(J), (11.14)
∂ξ ∂η
and, therefore,
N NX−m
∂x(ξ, η) X
= ϑmn mξ m−1 η n ,
∂ξ m=0 n=0
N NX−m
∂x(ξ, η) X
= ϑmn nξ m η n−1 ,
∂η m=0 n=0
N NX−m
∂y(ξ, η) X
= φmn mξ m−1 η n , (11.15)
∂ξ m=0 n=0
N NX−m
∂y(ξ, η) X
= φmn nξ m η n−1 ,
∂η m=0 n=0
where the coefficients δmn and γmn depend only on the physical values of the element. With
respect to a numerical evaluation of the Jacobian, the approach that has just been described is
much more efficient even if this advantage is limited to the initialization process.
238 Chapter 11. Discontinuous Galerkin-Based Reduced Order Models
In this section we provide a brief overview of the fluxes implemented in the full-order solver,
starting from the discretized numerical solution rather than the exact one. A complete mathe-
matical dissertation on the influence of the fluxes on the stability and accuracy of the solution is
beyond the scope of this work.
Since the numerical solution takes two distinct values in general at the interfaces between
neighboring elements, a numerical procedure must be adopted in order to evaluate flux exchanges
between contiguous cells, see Figure 11.3.
In particular, the third term of the left-hand side of (11.6) can be written as
Z Z
k k k
Lk+ k+ + k−
· Fhk− n− dΣk , (11.16)
Li · Fh n dΣ = i · Fh n + Li
Σk Σk
where + and − distinguish the values of the discontinuous quantities on either side of an inter-
face. The + sign corresponds to the element Ωk , with the normal direction consistent with the
edge orientation. Defining the jump operator [[A]] = A+ · n+ + A− · n− = (A+ − A− ) · n+
and the average operator hAi = (A+ + A− ) /2, relation (11.16) can be rewritten as
Z Z
Lki Fhk · n dΣk =
k k
Li hFh i + Fhk hLki i dΣk . (11.17)
Σk Σk
In the DGM, the argument of integral (11.17) is replaced by an interface flux function Γ depend-
ing on the test function, the local normal, and the values of the vector U and its derivatives of
both elemental edges delimiting the interface:
Γ U+ , U− , Lk+ k−
(11.18)
i , Li , n ;
it is stressed that the numerical flux Γ is evaluated from quantities that could be expanded in a
polynomial basis with different order Lk+ k−
i , Li .
To ensure consistency, Γ must satisfy the condition
To make the DGM elementwise conservative, the numerical flux must satisfy
Γ U+ , U− , Lk+ k− − + k− k+
(11.20)
i , Li , n = −Γ U , U , Li , Li , −n .
Γ = Γc + Γd . (11.21)
The convective discretization is stabilized by using an approximate Riemann solver H for Γ. The
approximate Riemann solvers H are based on the ones developed in the finite volume frame-
work. Among the most common are the HLL scheme [582], the Lax–Friedrichs scheme, and the
Rusanov scheme [581]:
F (U+ ) + F (U− ) supV∈Vad |∇(F · n)(V)|
HLF U+ , U− , Lk+ i , L k−
i , n = + [[U]] ,
2 2
(11.22)
where Vad is the space of admissible states and U+ and U− are expanded in the possibly different
polynomial bases Lk+i and Lk−i , respectively.
On the other hand, the diffusive flux exchange between neighboring elements depends both
on the values of the conservative variables and on their derivatives.
Many techniques can be adopted to evaluate the diffusive fluxes; one of the most common is
the one known as interior penalty [174].
Usually, diffusive splitting schemes also provide the splitting for the evaluation of the last
term of the left-hand side of (11.11), which depends again on the values of the variables on the
edges delimiting the interface. The reader can refer to [20, 375] for extensive studies of fluxes.
where µ is a set of parameters which the solution depends on, ϕ(x) is a set of global basis
functions, and ai are the unknown coefficients.
As with the other numerical methods, the reduced order methods for DGM rely on an online-
offline paradigm:
• offline phase: the snapshots are collected, modes are extracted, and matrices for reduced
order computations are computed;
• online phase: solutions are obtained in near real time for a given combination of parameters
with the aid of the matrices computed in the previous phase.
In the following subsections, a brief recap of the main steps needed to obtain the reduced
models is reported for the sake of clarity.
where µk is a set of parameters, P is the space of parameters, and Nk is the size of the
training set.
• Accurate Ns solutions of the flow field are computed for given sets of parameters µ in
order to obtain the snapshots:
• Reduced bases of dimension Nvr with Nvr N × M (i.e., the total number of degrees of
freedom of the full order case),
(11.26)
Lv = ϕ1 , ϕ2 , . . . , ϕNur ,
are obtained through a proper orthogonal decomposition (POD). These bases are used to
represent the flow field state variables, and they contain information about the dominant
dynamics which characterize the problem. They are ordered according to the magnitude
of their eigenvalues, whose values are proportional to the average energy of the projection
of the ensemble
λ1 > λ2 > · · · > λn > · · · > λN , (11.27)
and only the most energetic modes are retained. It is interesting to note that we are intro-
ducing a set of global bases here, while in a DGM formulation, bases are defined only on
the single elements of the domain.
The POD modes are obtained by solving a constrained optimization problem of the form
Rφ = λφ , (11.28)
• Once the modes have been extracted, the governing equations describing the problem at
hand can be projected onto the reduced space distinguished by the first modes computed
in the previous step.
The details of how the reduced matrices are obtained by projecting the governing equations onto
the reduced basis are presented in Subsection 11.4.
11.4. Projection of the Governing Equations 241
During the online phase, the reduced order system is solved for a given set of parameters µ in
near real time, given the small size of the matrices involved in the computation. The fields are
approximated at each time step as a linear combination of modes multiplied by proper weights,
which are the unknowns of the reduced order method. In general, a reliable approach consists
of solving the reduced system with the same time integration scheme adopted for the full-order
set in order to guarantee consistency between the full-order method and the reduced one. Both
explicit and implicit strategies can be adopted.
Putting together the system of equations just written and (11.23), we obtain the following ex-
pressions
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Nv Nv
X γ0 v̄j − α0 vjn − α1 vjn−1 D V V E X
φj ,φI = −β0 vjn vkn
j
δt
jk
D E
− β1 vjn−1 vkn−1 φV
j ∇φ V
k , φV
I ,
Np Nv
D
p p
E γ0 D V E
v̄j φj , ∇φpI ,
X X
pn+1
j ∇φ j , ∇φ I = −
j j
δt
Nv Np n+1 D
D E γ0 D V VE pj E
∇φpj , φV
X X
vjn+1 V V
∇φj , ∇φI + n+1
vj − v̄j φj , φI = − I ,
j
νδt j
ν
where φa and φb , φa and φb , and Φa and Φb are two generic scalar, vectorial, and matrix-valued
functions, respectively, defined over the domain Ω. An L2 -norm was chosen in this context
because this particular type of projection minimizes the energy loss in the system.
This approach is effective for describing flows in which strong nonlinear phenomena do not
occur.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
The linearized compressible Navier–Stokes set of equations projected on the domain Ω can
be written as
where the term depending on the pressure gradient is rewritten in terms of temperature T and
density using the gas equation
p = ρRT , (11.33)
where R is the gas constant. Moreover, in the mass conservation equation, the term describing
the density fluxes has been split into two components, applying the chain rule for derivatives.
The two equations have been multiplied by Rρ̄T̄ and ρ̄, respectively. This has been demon-
strated to be an effective way to further stabilize the reduced order system, and it is equivalent to
adopting a stabilized inner norm:
D E Z
φa , φb = φa Hφb dΩ , (11.34)
H Ω
where H is the stabilization matrix. Further details are reported in [317, 318].
The terms in (11.32) are evaluated in a way similar to the one adopted for the incompressible
flows, as follows:
D dρ0 E XNm D E
, ψIρ = ċj ψjρ , ψIρ ,
dt j=1
D E XNm D E
∇ρ0 , ψIρ = cj ∇ψjρ , ψIρ ,
j=1
D E XNm D E
∇ · V0 , ψIρ = ak ∇ · ψkV , ψIρ ,
k=1
D dV0 E Nm
X D E
, ψIV = ȧj ψjV , ψIV ,
dt j=1
D E XNm D E
∇ρ0 , ψIV = cj ∇ψjp , ψIV ,
j=1
D E XNm D E
∇ · V0 , ψ V = aj ∇ · ψlV , ψIV .
j=1
The terms reported above are all linear, which has beneficial effects in terms of storage and com-
putational efficiency. The integration in time is carried out with a strong stability-preserving
algorithm both for the full-order solver and the reduced order one in order to guarantee consis-
tency between full and reduced orders.
244 Chapter 11. Discontinuous Galerkin-Based Reduced Order Models
In this section, two different test cases are described, one for the incompressible case and one at
a moderate Mach number, in order to show both the effectiveness and the flexibility of the ROMs
based on DGM for modeling cases characterized by very different Mach and Reynolds number.
In order to train the reduced order methods, 10 simulations were carried out with different
values of viscosity, varying between 0.004 and 0.04 to match the Reynolds number, and 24
different snapshots were saved at different time instants for each simulation. Then both pressure
and velocity modes were extracted using a procedure based on the SVD. The first few modes for
the velocity are depicted in Figure 11.5.
Different numbers of modes Nu = Np were considered, and the equations were projected
onto the reduced space created by the modes.
Simulations for different Nu were run, imposing a viscosity equal to 0.01, a value which did
not belong to the training set. The final time of simulation was set to 5 s. The results obtained
with the ROM were compared to the ones obtained with the high-fidelity model, and they are
in good agreement for both pressure and velocity. Figure 11.6 shows the fields obtained when
different numbers of modes are taken into account. As expected, one can notice that the ones
with a higher number of modes provide a better estimation of the flow field. In Figures 11.7 and
11.8, the point-by-point differences between the field obtained with the ROM with Nu = 12 and
the one computed with the discontinuous Galerkin–based solver are reported.
Figure 11.5. First four velocity modes for the square cylinder test case.
Corresponding to the cavity, the width of the duct is 1.3 m. Solid wall no-slip boundary condi-
tions were imposed on the bottom wall, while on the upper part, slip boundary conditions were
used; the inlet and outlet boundaries were modeled with nonreflecting boundary conditions based
on the conservation of the Riemann invariants in order to avoid the reflection of outgoing waves
back into the computational domain. The discretization of the domain is depicted in Figure 11.9,
consisting of approximately 6200 unstructured straight-edged triangles. The polynomial order
is constant and equal to three in the whole computational domain. The Mach number of the
unperturbed flow at the inlet is fixed and set to 0.3, while the Reynolds number varies in a range
between 9000 and 12000.
During the offline phase, nine full-order simulations were run, corresponding to different
linearly spaced values of the Reynolds number, and 40 snapshots were saved for each of them
every 0.015 s; after that, the behavior became periodic. Once all the snapshots were collected,
the snapshot matrix was built and the modes were extracted with the aid of an SVD procedure;
the first ones are reported in Figure 11.10. Figure 11.11 shows a qualitative comparison between
a full-order solution and the corresponding reduced one for a Reynolds number equal to 10000,
a value which was not included in the previous set of simulations. The ROM results are in good
accordance with the full-order ones; some differences can be observed in particular where the
nonlinear dynamics are stronger, e.g., close to the corners and in the area where the cavity is
placed. These discrepancies can be explained by the fact that in this case a linearized version of
the Navier–Stokes equations was implemented in the reduced order solver.
Figure 11.11. Comparison between FOM (top) and ROM (bottom) solutions.
247
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Part III
Fluid Dynamics
Models for Computational
Advances in Reduced Order
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 12
12.1 Introduction
Partial differential equations (PDEs) are an effective tool to model phenomena in applied sci-
ences. Realistic problems usually depend on several physical and geometrical parameters that
can be calibrated by using real data. In real scenarios, however, these parameters are affected by
uncertainty due to measurement errors or scattered data information. To deal with more reliable
models which take this issue into account, stochastic PDEs can be numerically approximated. In
the uncertainty quantification (UQ) context, many simulations are run to better understand the
system at hand and to compute statistics of outcomes over quantities of interest. In particular, the
input parameters of the stochastic PDEs are assumed to be random finite-dimensional variables.
Classical numerical approximations, i.e., high-fidelity solutions, can lead to untenable com-
putational costs for computing statistical momenta. These are the leading motivations of this
contribution. Indeed, we will focus on projection-based reduction techniques which can lighten
this issue: reduced order models (ROMs) [281, 505]. In many applications, the parametrized so-
lutions can be sought in a low-dimensional subset of the solution space. If one applies a Galerkin
projection in this reduced space, the problem is solved more rapidly than the original discretiza-
tion and is still accurate. These approaches, thus, might accelerate standard statistical analysis
techniques, such as the Monte Carlo methods. (For alternative methods for forward UQ with
random inputs, e.g., stochastic Galerkin, stochastic collocation, and Karhunen–Loève approxi-
mation, we refer the reader to [565].) More precisely, this chapter focuses on weighted ROMs for
forward uncertainty problems, where the reduced algorithms are modified to comply with some
previous knowledge of the distribution of the parameters and to use this information to accelerate
the reduced simulations even more; see, e.g., [117, 592] and the references therein. Moreover,
the proposed weighted ROMs are not intrusive with respect to the classical ROM approaches;
i.e., given a reduced model the weighted ROMs just weigh its outcome without changing its
form. We refer the reader to [406] for inverse problem applications and to [285] for polynomial
chaos expansion-reduced techniques.
251
252 Chapter 12. Weighted Reduced Order Methods for Uncertainty Quantification
The contribution is outlined as follows: in Section 12.2 we present stochastic PDEs and their
approximation. Then, we move toward weighted ROMs, describing standard algorithms such
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
as weighted proper orthogonal decomposition (POD) [117] and weighted reduced basis [592]
in Section 12.3 and the related sampling strategies in Section 12.4. Then, in Section 12.5, we
validate the proposed algorithms in several contexts, including heat transfer, Stokes problems,
advection-dominated phenomena, and optimal control for environmental sciences. Conclusions
follow in Section 12.6.
In this chapter we will focus on how the offline phase of different algorithms can be modified to
take into account the knowledge of the distribution of the random variable µ. This is done mainly
by means of weighted algorithms that are derived from classical ROM algorithms [281, 505]. We
will present the weighted reduced basis (wRB) method [117] and the weighted proper orthogonal
decomposition (wPOD) [592]. Moreover, we will focus only on the offline modification of such
algorithms, because the online phase is unmodified with respect to the original algorithms. The
difference is that now we are interested not only in one evaluation for a given µ but also in the
momenta of the random variable un (µ). We refer the reader to [281, 504] for the development
of those parts of the algorithms.
In the stochastic context, we want to give different importance to different parameters µ, i.e.,
different realizations, according to the underlying probability distribution. This must be done
by modifying the error measure that we used in the deterministic setting, introducing a different
norm k · kw that modifies the search of the worst-represented snapshot in the greedy algorithm.
In particular, the new norm is defined as
The weight function w : Γ → R+ depends on the target norm that we want to minimize. As an
example, if we want to minimize the expected value of the 2-norm of the error, i.e.,
Z
E[kuN − un k2VN ] = kuN (µ(ω)) − un (µ(ω))k2VN dP (ω) (12.6)
D
Z
= kuN (µ) − un (µ)k2VN ρ(µ)dµ, (12.7)
Γ
254 Chapter 12. Weighted Reduced Order Methods for Uncertainty Quantification
the natural choice for the weight function is w(µ) = ρ(µ) [117]. Consequently, the error
p
estimator also reflects this modification by introducing a new error bound ηnw (µ) = w(µ)ηn (µ),
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
resulting in
Z Z
2 w 2
E[kuN − un kVN ] ≤ ηn (µ) dy = ηn (µ)2 ρ(µ)dµ. (12.8)
Γ Γ
Other choices may be more meaningful in other contexts, e.g., different norms or output error
minimization [580, 593]. The modified algorithm is reported in Algorithm 12.2.
end
Another important aspect in this algorithm is the choice of the sampling strategy of the parameter
training set Ξt . In Section 12.4, we will investigate in more detail the different possibilities and
their effects on the overall results.
12.3.2 wPOD
The other well known algorithm in the ROMs community is the proper orthogonal decomposition
(POD) [281]. Its weighted version (wPOD), originally introduced in [592], is useful when an
error estimator ηn is not available.
The POD Algorithm 12.3 finds the global minimum of the mean square error
Z
kuN (µ) − un (µ)k2VN dµ (12.9)
Γ
over all the possible reduced spaces Vn ⊂ VN of dimension n. In practice, a tolerance on the
error can be set and the dimension n will be provided by the algorithm. In the discrete situation
(12.9) becomes
1 X
kuN (µ) − un (µ)k2VN , (12.10)
|Ξt |
µ∈Ξt
with Ξt being the training set. The reduced space Vn is defined by the eigenvectors correspond-
ing to the n leading eigenvalues of the operator C : VN → VN defined by
|Ξt |
X
C(v) := hv, uN (µi )iVN uN (µi ). (12.11)
i=1
In practice, this is computed by an eigenvalue analysis on the correlation matrix Ĉij := huN (µi ),
P|Ξt | i
uN (µj )iVN . The bases of the reduced space Vn are defined by ξ i = j=1 ψj uN (µj ), where ψj
are the eigenvectors of Ĉ.
12.4. Sampling Strategies 255
for µ ∈ Ξt do
Solve (12.2) for µ to get uN (µ)
end
Assemble the matrix Ĉij = huN (µi ), uN (µj )iVN
Compute the biggest n eigenvalues λk and the eigenvectors ψk of Ĉ for k = 1, . . . , n
P|Ξt | i
Define Vn := span{ξ 1 , . . . , ξ n }, where ξ i = j=1 ψj uN (µj ).
The stochastic setting modifies the mean square error (12.9) as the expected value (12.7).
This leads to a modified correlation matrix
w
Ĉij := w(µi )huN (µi ), uN (µj )iVN . (12.12)
We remark that Ĉ w is not diagonalizable in the usual sense, but it is with respect to the scalar
product induced by C, which allows us to obtain n orthogonal leading eigenvectors. The modi-
fied algorithm is reported in Algorithm 12.4.
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Figure 12.1. Comparison of multivariate quadrature rules with order of accuracy 15.
of the function and the accuracy degree of the quadrature rule, and K is the dimension of the
(parameter) space Γ.
Sparse quadrature rules can improve this by putting the quadrature points in different loca-
tions. The Smolyak quadrature rule [292, 48, 610] allows us to not fall victim to the curse of
dimensionality, using different refinement levels of the grids on the different parameters. Con-
(j)
sider Γ := ΠK j=1 Γj , and define by Ui the univariate quadrature rule at the refinement level i on
the interval Γj ⊂ R, introducing the differences operators on Γj as
(j) (j) (j) (j)
∆0 = 0, ∆i := Ui+1 − Ui for i ≥ 0. (12.13)
The Smolyak quadrature rule of order q in Γ is defined as
X K
O
QK
q :=
(i)
∆α i
. (12.14)
|α|1 ≤q, α∈NK i=1
The difference can be observed in Figure 12.1, where a two-dimensional example shows the
tensor product and the Smolyak rule.
In the next section, we will alternate different quadrature strategies to show the advantages of
choosing samples distributed with the underlying probability law and the cost reduction without
loss of quality of sparse quadratures.
12.5 Applications
In this section we apply the strategies presented above to different problems. In the first subsec-
tion we apply the presented strategies directly to a heat equation problem and to Stokes problems.
In Subsection 12.5.2 we use the proposed methods for a selective stabilization technique in the
context of uncertainty quantification. Finally, in Subsection 12.5.3 we apply the weighted algo-
rithms to parametrized optimal control problems OCP(µ)s in environmental sciences.
where µ : [0, 1]2 → R is defined piecewise constant on nine subsquares defined by Dij =
[i/3, (i+1)/3]×[j/3, (j+1)/3], i, j = 0, . . . , 2, by nine parameters µij ∈ R with i, j = 0, . . . , 2.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In particular, we consider these heat transfer coefficients to behave like some Beta(20,10) random
variables, i.e., µij ∼ Beta(20,10) rescaled on the interval [1, 10]. The Beta distribution is defined
R1
using the Beta function B(α, β) = 0 µα−1 (1 − µ)β−1 dµ with the probability density function
µα−1 (1 − µ)β−1
ρ(µ) = . (12.16)
B(α, β)
We compare the reduced algorithms presented before in their weighted and classical formu-
lations, and we compare the uniform Monte Carlo sampling strategy with the Beta(20,10) Monte
Carlo sampling one, in both cases with |Ξt | = 100.
To assess the quality of the algorithms, we compute reduced solutions on a Beta(20, 10)
distributed test set and we average the error obtained with such solutions.
Test error on Beta distributed samples
10 2
10 3
10 3
Uniform MC, Uniform POD Uniform MC, Uniform RB
10 4 Uniform MC, Beta POD Uniform MC, Beta RB
Beta MC, Uniform POD Beta MC, Uniform RB
Beta MC, Beta POD Beta MC, Beta RB
2 4 6 8 10 12 14 2 4 6 8 10 12 14
n reduced basis n reduced basis
Figure 12.2. Heat transfer problem: error decay with respect to the number of reduced ba-
sis functions, comparison between weighted and not weighted algorithms and between uniform and
Beta(20,10) distributed training samples. Left: POD; right: greedy algorithm.
In Figure 12.2, we notice that for the POD the weighting of the algorithm does not lead to
great improvement in the decay of the error, while sampling of the training set with the underlying
distribution greatly reduces the error, by as much as a factor of 10. Conversely, for the greedy
algorithm we see a great improvement with only the weighted algorithm of factor five. Here the
sampling strategy also plays a bigger role in the training phase and gives an overall improvement
of at least a factor of 10. The combination of the two strategies slightly improves this result. We
remark that a uniform sample considers parameters which are not at all well represented in the
reduced space, leading to worse average errors.
We now focus on the stochastic steady Stokes equations results shown in [218]: here we
extend the deterministic problem presented in [211] following the strategies first discussed in
[118]. In this case, we will deal with physical and geometrical parametrization. Namely, we
consider the random domain D(µ(ω)) : A → R2 and a random inlet condition gin (x, µ(ω)) :
∂Din ×A → R2 . From now on, for the sake of notation, we will use µ and µ(ω) interchangeably.
The random boundary of the domain is given by the partition
where ∂DD,0 (µ) and ∂Din (µ) represent the portions of the boundary with homogeneous and
nonhomogeneous Dirichlet conditions, respectively, while ∂DN (µ) is characterized by Neumann
boundary conditions. The stochastic Stokes problem reads as follows: given an outcome ω ∈ A,
258 Chapter 12. Weighted Reduced Order Methods for Uncertainty Quantification
Figure 12.3. Reference domain D. The reference parameter is µ = (1., 1.5, 1., 1.5, µ4 ).
12.5. Applications 259
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 12.4. Stokes problem: error decay with respect to the number of RB functions, compar-
ison between unweighted (left) and weighted (right) algorithms and between the Smolyak, tensor product,
uniform, and Beta(75,75) distributed training samples. For the weighted algorithm, the Smolyak rule, the
tensor product rule, and Beta(75,75) sampling coincide.
on the domain D = [0, 2] × [0, 1] with Dirichlet boundary condition equal to one on the bound-
aries with x1 ≤ 1 and equal to zero on the boundaries with x1 > 1. We suppose that the
parameter µ is distributed as a Beta(5,3) rescaled on the interval [0, 6]. The considered problem
will be advection dominated as µ increases, and a standard finite element method (FEM) would
fail to provide an accurate solution because it would be led by oscillations due to the instability
of the problem. In order to stabilize the method, we introduce in the offline phase a stabilization
term, represented by streamline upwind Petrov–Galerkin (SUPG) [477], which maintains the
consistency of the method but adds a stabilization effect.
In this simple example the stabilization is not computationally expensive, but in more realistic
cases the stabilization term might include more elaborate nonlinear terms.
In the online phase we can choose whether we want to apply the stabilization again or not.
As for the high-fidelity model, it is necessary to use it in the advection-dominated regime. We
see this behavior by running an online error decay study with respect to the dimension of n in
Figure 12.5. There, it is clear that, both with POD and greedy without online stabilization, the
error has a plateau at 10−1 , which is the distance between the nonstabilized and the stabilized
solutions [580].
In Figure 12.5 for stabilized methods, we observe the same order of convergence but smaller
errors for weighted algorithms and distributed training samples. Hence, for the rest of this section
we consider only weighted algorithms with distributed training samples.
We can observe in Figure 12.6 that the unstabilized and stabilized online solutions show
differences in the error, particularly in the advection-dominated regime, while, in the diffusive
regime, the two solutions show similar errors. Hence, one may be more interested in solving
these solutions without the stabilization terms to save computational time.
260 Chapter 12. Weighted Reduced Order Methods for Uncertainty Quantification
10 2
10 3
10 4 10 1
0 5 10 15 20 25 30 0 5 10 15 20 25 30
n reduced basis n reduced basis
10 3
100
10 4
10 5
10 6 10 1
0 10 20 30 40 50 0 10 20 30 40 50
n reduced basis n reduced basis
Figure 12.5. Graetz problem: error decay with respect to the number of RB functions, and
comparison between weighted and unweighted algorithms and between uniform and Beta(5,3) distributed
training samples. With online stabilization (left), without online stabilization (right). POD algorithm (top),
greedy algorithm (bottom).
Error
Selective Online
10 4
Stabilization Line
10 3
10 5
10 4
10 6
10 5 10 7
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Advection parameter Advection parameter
Figure 12.6. Graetz problem: weighted algorithms and Beta(5,3) distributed training samples.
Error of stabilized and nonstabilized online with respect to the advection parameter µ. Comparison with
selective online strategy. Left: POD; right: Greedy.
The task that we aim to solve in this section is to compute some statistical quantities of the
stochastic solution u(µ), e.g., the average. A classical strategy could be to approximate the
average via a Monte Carlo algorithm, i.e.,
N MC
u(µi )ρ(µi )
Z Z X
E(u(µ)) = u(µ(ω))dP (ω) = u(µ(ω))ρ(µ(ω))dω ≈ , (12.19)
A A i=1
NM C
where ρ is the probability density function of the underlying distribution of µ and µi are Monte
Carlo samples uniformly distributed in the domain of interest.
12.5. Applications 261
Table 12.1. Tables for the selective stabilization approach. Given a certain threshold advection
coefficient ε, we obtain the mean error by computing only a percentage of all the stabilization terms.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Computing many high-fidelity simulations uN (µi ) is too expensive for this task, so we con-
sider the reduced solutions un (µi ) instead. Now, the stabilization term is important to guarantee
an accurate solution in the advection-dominated regime, i.e., µ 0. Conversely, in the diffusion-
dominated regime, the stabilization term is not necessary and would be time-consuming.
What we present here is a selective approach to the reduced solutions, which will be com-
puted with the stabilization term or not according to a rule led by a threshold value ε. Namely,
while computing (12.19), for every ui we will use un (µi ) with the stabilization term if µi > ε;
else we will use un (µi ) without the stabilization term. This idea is illustrated in Figure 12.6 by
the selective online plot.
In Table 12.1 we see the result of this approach of changing the value of the threshold ε. In
the POD case we see that for ε = 1.5 we save the computational costs of 22% of the stabilization
terms and we obtain a negligible worsening of the error. With ε = 2 the error is already a bit
larger though increasing to almost the 30% of stabilizations not computed. Higher thresholds
make less sense. For the greedy approach we observe similar results, even if the error increases
a bit more with small thresholds.
constrained to
(12.21)
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
P(y(µ), u(µ); µ) = 0,
where P is a linear stochastic constraint in the weak formulation. The interested reader may refer
to [587] for a survey on optimal control theory. This problem formulation has been used in data
assimilation contexts, where the variable u(µ) steers the system toward the desired configuration
yd , making the solution y(u(µ)) the most similar to the observation. The stochastic parameter
still plays an important role since both the data and the system may be affected by uncertainty. In
the following, we report some results concerning this framework applied to a hypothetical loss
of pollutant in the Gulf of Trieste. Let us indicate this geographical region with D ⊂ R2 , repre-
sented in Figure 12.7. We call Du the subdomain where the pollutant loss is happening (green
subdomain in Figure 12.7). The goal is to determine the maximum loss allowed in Du to keep
harmless the pollutant concentration in Dobs , the Miramare natural reserve (orange subdomain
in Figure 12.7). The boundary is partitioned into ∂DD , the coastline, and ∂DN , the open sea,
where Dirichlet and Neumann boundary conditions have been applied, respectively. The state
concentration is y ∈ Y := H∂D 1
D
, u ∈ R, and yd = 0.2 over Dobs . The observation represents a
safe threshold for the pollutant concentration in the Miramare area.
Figure 12.7. Domain D, the Gulf of Trieste. Orange: Miramare reserve Dobs . Pollutant spill Du .
Here, the random parameter µ = (µ1 , µ2 , µ3 ) ∈ (0.5, 1) × (−1, 1) × (−1, 1) models the sea
dynamics under several meteorological phenomena. The constant L = 1000 is used to make
the system nondimensional [557]. Furthermore, we impose α = 10−7 . We want to compare
the performances of informed sampling and standard POD. We will show how exploiting data
information in this context might help monitor and solve in real time a potentially dangerous
situation. Thus, we exploit the unweighted version of the POD with aggregated space strategy
[328, 329, 414], since, as we already said in the aforementioned applications, the POD is only
slightly affected by the weights, while the sampling strategy is of great importance. In this
experiment |Ξt | = 100. In Figure 12.8 we present the average relative error for the state and
the control variables, where the sampling for the test coincides with that of the offline phase.
This ensures good performances for all the samplings. However, we would like to underline
that adding distribution information can be crucial when dealing with random variables. Indeed,
thanks to some previous knowledge about parameter distribution, one can employ a much lower
number of bases. This is due to the nature of the algorithm that focuses on a very specific
parameter setting both in creating the reduced spaces and in assessing their accuracy. This is the
12.6. Conclusions 263
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 12.8. Optimal control problem: error decay with respect to the number of RB functions;
comparison between state variable (left) and control variable (right) with an unweighted POD algorithm
between uniform, Beta(20, 5), Beta(5, 5), and Beta(75,75) distributed training samples.
case for µ ∼ Beta(75,75) (rescaled on the respective parameter ranges), where n = 4 is sufficient
to reach error values around 10−6 for both the variables, gaining two orders of magnitude with
respect to uniform sampling. Fewer bases translate to a gain of computational time that can be
exploited in UQ analysis even for very complicated problems such as OCPs, which are usually
characterized by high computational costs due to the minimization framework.
12.6 Conclusions
This chapter focused on wROMs for a broad class of stochastic PDE-based models. First, we
introduced stochastic PDEs with random inputs and their standard discretizations. Then, we
introduced ROM strategies in order to deal with them more quickly, following the POD and the
RB model. The related weighted approaches were also described. Furthermore, we highlighted
the role of several sampling strategies based on different (possibly sparse) quadrature rules. We
validated this setting in many contexts, such as the heat transfer equation, Stokes problems, the
stabilization of advection-dominated problems, and optimal control for environmental sciences.
The results show the better performances of weighted approaches and distributed quadratures
than a standard algorithm based on a deterministic viewpoint.
Chapter 13
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Reduced Basis,
Embedded Methods, and
Parametrized Level-Set
Geometry
−∆T = g in D, with T = gD on Γ.
We are looking for a variational formulation that is satisfied by the weak solution T ∈ H 1 (D),
i.e., it is consistent, it is symmetric, and it has a unique solution—the bilinear form is coercive.
We start by taking the strong form of the equation, multiply by a test function v ∈ H 1 (D), and
265
266 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
integrate by parts. Starting with the right-hand side, we add the productive zero 0 = T − gD on
the boundary and obtain
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Z Z
(g, v)D = (−∆T, v)D = (∇T, ∇v)D − n · ∇T v ds − (T − gD )n · ∇v ds.
Γ Γ
Separating linear and bilinear forms, for all v ∈ H 1 (D), the variational equation—the symmetric
bilinear formulation—becomes
Z Z Z Z
(∇T, ∇v)D − n · ∇T v ds − T n · ∇v ds = − gD n · ∇v ds + gv dx.
Γ Γ Γ D
We note that in the above formulation the symmetric bilinear form is not coercive because it
cannot be bounded from below for T = v by ckvk2H 1 (D) . So, we add a symmetric term that
vanishes for the true solution: η Γ (T − gD )v ds for some η > 0 large enough. This leads to
R
the following symmetric, consistent, coercive weak formulation: find T ∈ H 1 (D), ∀v ∈ H 1 (D)
such that
Z Z Z Z
(∇T, ∇v)D − n · ∇T v ds − T n · ∇v ds + η T v ds = − gn · ∇v ds
Γ Γ Γ Γ
Z Z
+η gD v ds + gv dx. (13.1)
Γ D
For finite element (FE) discrete approximations Th , vh ∈ Vh ⊂ H 1 (Ω), and for the Nitsche
penalty chosen as η = ch−1 with c > 0 sufficiently large, one can manage a stable discrete
problem with respect to a suitable mesh-dependent norm.
or, equivalently,
Z Z Z
(∇T, ∇v)D̃ − ñ · ∇T v ds − T ñ · ∇v ds + η T v ds
ZΓ̃ Γ̃
Z Γ̃
Z
=− (T + ∇T · d)ñ · ∇v ds + η (T + ∇T · d)v ds + gv dx,
Γ̃ Γ̃ D̃
where g̃D denotes an extension of the Dirichlet boundary condition gD to the boundary Γ̃ of the
surrogate domain based on a second-order accurate Taylor expansion, T + ∇T · d ≈ g̃D , or, if
we let x = M(x̃), then T (x̃) + ∇T (x̃) · d ≈ gD (x).
13.1. Introduction and Overview 267
(i) CutFEM
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(ii) SBM
Figure 13.1. Embedded FEM geometrical tools for (i) CutFEM and (ii) SBM.
which, using an SBM weak formulation, can be transformed into the system of linear equations
(rewritten in matrix form)
A(µ)T(µ) = Fg (µ), (13.3)
where A(µ) ∈ RNh ×Nh and Fg (µ) ∈ RNh ×1 correspond to the bilinear and linear forms,
respectively.
Online stage. This stage is often performed on a system with reduced computational power and
storage capacity; the reduced system can be solved for any new value of the input parameters
with predicted accuracy and reduced computational and time cost. For the interested reader we
refer to [281].
268 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
13.1.5 POD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Subsequently, with a POD one can generate the RB space after the full-order model (FOM) has
been solved for each µ ∈ K = {µ1 , . . . , µNs } ⊂ Rk , where K is a finite-dimensional training
set of parameters inside the parameter space P, Ns is the number of snapshots, Nh denotes the
number of degrees of freedom for the discrete full-order solution, and the snapshot matrix S is
given by Ns full-order snapshots properly extended to a fixed background mesh and defined on
µi parameter-dependent domains:
Ns NX 2
X P OD
with (ϕi , ϕj )L2 (D) = δij , i, j = 1, . . . , Ns . This is equivalent to solving the eigenvalue problem
CQ = Qλ for Cij = (Ti , Tj )L2 (D) , i, j = 1, . . . , Ns , where C is the correlation matrix
obtained by starting from the snapshots S, Q is a square matrix of eigenvectors, and λ is a
diagonalPmatrix of eigenvalues. The basis functions can then be obtained with the formula ϕi =
Ns Nh ×N r
j=1 Tj Qij , and the POD space L = [ϕ1 , . . . , ϕN ] ∈ R for N r < Ns is chosen
1 r
1/2
Ns λii
according to the eigenvalue decay of λ (see [505]).
r
for which the reduced solution vector a ∈ RN ×1 depends only on the parameter values. The
basis functions ϕi depend only on the physical space, and the unknown vector of coefficients a
can be obtained by using a Galerkin projection of the full-order system of the equations onto the
POD RB space, resulting in the reduced algebraic system
(i)
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(ii)
(iii)
Figure 13.2. (i) Parametrized geometry. (ii) Some RB components for µ ∈ [−0.5, 0.5]
parametrized geometry. (iii) The full/reduced order solution and the absolute error (µ = −0.015).
We assume that the embedded domain consists of a rectangle of size 0.8 × 0.7, and its position
inside the domain is parametrized with a geometrical parameter µ. The position (0, µ) of the
rectangular embedded domain depends on its parametrized y center, while the horizontal coor-
dinate of the center of the box is not parametrized and is located at the x center of the domain.
The whole configuration is immersed in a background domain of size D = [−2, 2] × [−1, 1].
The ROM has been trained with 400 samples, which is the dimension of the offline FOM, for
µ ∈ [−0.5, 0.5] chosen randomly inside the parameter space, while the ROM results have been
compared with the FOM for 50 additional random samples. The mean relative errors, the eigen-
value decay, and the behavior of the results for one parameter sample are visualized in Fig-
ures 13.2 and 13.3 and Table 13.1. The time savings are reported in Table 13.2, where the com-
putation time includes assembling the full-order matrices, their projection, and the resolution of
the reduced problem, while we have avoided (a) using a reference domain, (b) remeshing, and
(c) adaptive refinement.
270 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(i)
1 Eigenvalues Decay
10
4
10
Eigenvalues magnitude
7
10
10
10
13
10
16
10
(ii)
1%
0.1%
5 10 15 20 25 30 35 40 45 50
N of modes
Figure 13.3. Heat exchange problem: (i) eigenvalue decay and (ii) mean relative errors.
Table 13.2. Heat exchange problem: execution time, savings, and speedup.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Savings Speedup
Modes Execution time(s) tFOM −tRB tFOM
tRB tRB
2 4.119470 × 10−2 96.399% 27.770
10 4.168334 × 10−2 96.356% 27.445
20 4.243647 × 10−2 96.290% 26.957
30 4.353909 × 10−2 96.194% 26.275
40 4.449359 × 10−2 96.110% 25.711
50 4.494564 × 10−2 96.071% 25.452
100 4.992923 × 10−2 95.635% 22.912
FOM 1.14540 × 100 – –
where (u) = 1/2(∇u + ∇uT ) is the velocity strain tensor (i.e., the symmetric gradient of the
velocity), p is the pressure, g is a body force, gD is the value of the velocity on the Dirichlet
boundary, and gN is the normal stress on the Neumann boundary.
Based on [326] we define the following shifted boundary weak formulation: find u ∈ Vh
and p ∈ Qh such that, ∀w ∈ Vh and ∀q ∈ Qh ,
and
−(∇ · u, q)D̃(µ) + hu · ñ, qiΓ̃(µ) + hqd ⊗ ñ, ∇uiΓ̃(µ) = hḡD · ñ, qiΓ̃(µ) ,
where n and τi are the unit normal vector and unit tangential vectors to the boundary Γ and can be
272 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 13.4. Stokes with SBM experiment: sketch of the mesh, the embedded domain, and the
parameters considered in the numerical examples.
extended to the boundary Γ̃, namely ñ(x̃) ≡ n(M(x̃)) and τ̄i (x̃) ≡ τi (M((x̃)), ḡD (x̃) =
gD (M(x̃)). We clarify that µ is defined similarly in Subsection 13.1.5 and h is a characteristic
length of the elements.
where Nur , Npr < Ns are chosen according to the eigenvalue decay of the vectors of eigenvalues
λu and λp . Furthermore, for the best approximation we employ supremizer enrichment as in the
work of [511] and we manage a solvable and stable problem that satisfies a reduced and also
parametric version of the inf-sup condition. With this approach, the velocity supremizer basis
N h ×Nsup
r
functions Lsup , with Lsup = [η1 , . . . , ηNsup
r ] ∈ R u , are computed and added to the reduced
velocity space, which is transformed into L̃u :
h r r
N ×(Nu +Nsup )
r ] ∈ R u
L̃u = [ϕ1 , . . . , ϕNur , η1 , . . . , ηNsup . (13.9)
(i)
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(ii)
(i)
(ii)
Figure 13.5. Stokes system ROM-SBM POD velocity and pressure components for (i) µ1 geome-
try and (ii) (µ0 , µ1 ) geometry.
the two-dimensional case the relative errors are in Table 13.5 and Figure 13.7(ii). Focusing on
the supremizer stabilization approach, the improved results for pressure are obvious and the plots
clearly show at a glance that the high-fidelity and ROM solutions cannot be easily distinguished.
Related to the execution time investigation for the one-dimensional geometrical parametrization,
we compare online stage execution times against the full-order computational times and FOM
solutions. Namely, for all 1024 snapshots and supremizers, the solutions cost 1 h 42 min 54 s.
Table 13.3. Stokes (SBM): relative errors between the full-order solution and the RB solution,
one-dimensional geometrical parametrization.
(i)
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(ii)
(i)
(ii)
Figure 13.6. Stokes SBM: the full and the reduced solution and the absolute error for velocity
and pressure for the (i) one-dimensional and (ii) two-dimensional parametrization case.
Table 13.4. Execution time, at the reduced order level, for the case with one-dimensional geo-
metrical parametrization.
no supr. supr.
Modes exec. time (s) exec. time (s)
8 7.3858961 7.710907
12 7.6042165 8.091225
16 7.9584049 8.290780
20 8.0206915 9.036709
25 8.2229143 9.495323
30 8.9529275 9.972288
35 9.0867916 10.47633
40 9.6555775 11.13931
45 9.8934008 11.49422
50 10.302459 11.92024
13.2. Parametrized Steady Stokes Equations 275
u
u, supremizers
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
p
p, supremizers
100%
relative error
(i)a
10%
0.1%
10 20 30 40 50
N
truth
reduced
reduced (supr.)
40
30
execution time
(i)b 20
10
10 20 30 40 50
N
u, 1024 snapshots
u 1024 snapshots, supremizers
u, 900 snapshots
u 900 snapshots, supremizers
100%
relative error U
(ii)a 10%
1%
20 40 60 80 100
N
p, 1024 snapshots
p, 1024 snapshots, supremizers
p, 900 snapshots
p, 900 snapshots, supremizers
100%
relative error P
(ii)b
10%
20 40 60 80 100
N
Figure 13.7. Stokes (SBM) parametrization: (i) µ1 geometry, with the relative errors for velocity
and pressure in (i)a and the execution times in (i)b ; (ii) (µ0 , µ1 ) geometry, with the velocity and pressure
fields with and without supremizer stabilization.
276 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
Table 13.5. Stokes SBM (µ0 , µ1 ) geometry: Supremizer basis enrichment and the relative error.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
This is an expensive stage, but fortunately it is executed only once at the beginning. We clarify
that the online stage computation time includes assembling the full-order matrices and product-
ing the ROM and its resolution (see Table 13.4). Ten different values of the input parameter
are considered, with reference time the time execution at the FOM level, which for each one-
parameter solution is equal to ≈ 37 s. Obviously, the ROM leads to a considerable speedup
for all the different analyzed configurations and for cases both with and without supremizer en-
richment. The interest in the latter experiments was more about testing the feasibility and the
accuracy of a ROM, constructed starting from a shifted boundary FOM, and we did not employ
any hyperreduction techniques, which means that, at the reduced order level, we also assembled
the full-order discretized differential operators. Focusing on the (µ0 , µ1 ) geometry experiment
with range µ = (µ0 , µ1 ) = [−1.5, −1.0] × [−0.15, 0.15], which is a more complex and demand-
ing scenario, we can easily notice that the supremizer enrichment is necessary for convergence
stability and reliable pressure results, and without supremizer enrichment and 900 snapshots, the
best achieved—and most disappointing—relative errors were equal to 0.0263821 and 0.1854861,
respectively, for velocity and pressure. Figure 13.7 and Table 13.5 also give an overview of the
dependence on the number of snapshots for two choices of 900 and 1024 used in the offline stage.
Figure 13.8. Shape parametrization with large deformations and a zoom into the embedded
cylinder visualizing the extended solution.
The boundary value problem is formulated on a domain T (µ) that contains D(µ) ⊂ T (µ),
whose mesh Th (µ) is not fitted to the domain boundary, where Gh (µ) := {K ∈ Th (µ) : K ∩
Γ(µ) 6= ∅} is the set of elements that are intersected by the interface, DT (µ) := {K ∈ Th (µ) :
K ∩ D(µ)} ∪ Gh (µ); the background domain is typified by B, while its corresponding mesh is
denoted by Bh , such that DT (µ) ⊂ B and Th (µ) ⊂ Bh for all µ ∈ K. See also Figure 13.1(ii).
We remark that Th (µ), Gh (µ), and DT (µ) depend on µ through D(µ) or its boundary, while the
background domain B and its mesh Bh do not depend on µ. Furthermore, the set of element faces
FG (µ) associated with Gh (µ) is defined as follows: for each face F ∈ FG (µ), there exist two
simplices K 6= K 0 such that F = K ∩K 0 and at least one of the two is a member of Gh (µ). Note
that the boundary faces of Th (µ) are excluded from FG (µ). On a face F ∈ FG (µ), F = K ∩K 0 ,
the jump of the gradient of v ∈ C 0 (DT ) is defined by [[nF · ∇v]] = nF · ∇v|K − nF · ∇v|K 0 ,
where nF denotes the outward-pointing unit normal vector with respect to K. So the CutFEM
discretization is as follows: we seek a discrete solution uh (µ) in the FE space
+ (γN hnΓ · ∇uh (µ), nΓ · ∇vh (µ))ΓN (µ) + j(uh (µ), vh (µ)),
h
` (vh (µ)) = (g(µ), υh (µ))D(µ) − (gD , nΓ · ∇υh (µ))Γ(µ)
+ γD h−1 gD , vh (µ) Γ(µ) + (gN (µ), vh (µ)+γN hnΓ · ∇vh (µ))ΓN (µ) ,
extends the coercivity from the physical domain D(µ) to the µ-dependent mesh domain DT . γD ,
γN , and γ1 are positive penalty parameters. The coefficients γD and γN account for a Nitsche
weak imposition of boundary conditions. We set up some experiments again for embedded
278 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
Figure 13.9. Six classical (i) and six improved (ii) POD modes.
FEMs and ROMs, emphasizing improvements in cases with large geometrical deformations. The
parametrized domain D(µ) ⊂ R2 is an ellipse described by the level set φ(x, y; µ1 , µ2 , µ3 , µ4 ) =
µ22 (x − µ3 )2 + µ21 (y − µ4 )2 − µ21 µ22 R, where the reference radius R = 0.05, the length of
the axes of the ellipse is (µ1 , µ2 ) ∈ [0.3, 1.8]2 , and the position of the center of the ellipse is
(µ3 , µ4 ) ∈ [−0.85, 0.85]2 . A corresponding background domain B = [−1.2, 1.2]2 is chosen so
that the ellipse is strictly contained in B for any µ = (µ1 , µ2 , µ3 , µ4 ) in the parametric range
K = [0.3, 1.8]2 × [−0.85, 0.85]2 . The value µ = (1, 1, 0, 0), corresponding to a circle of radius
R centered at the origin, is chosen for the transport method. The data of the problem described
by (13.2) comprises the force g(x, y; µ) = 20 and the Dirichlet boundary force gD (x, y; µ) =
0.5 + xy.
In Figure 13.10(i) for the CutFEM Poisson system, the reduced solution obtained from the
zero extension is inaccurate even for N = 140 modes, being affected by relative errors of the
order of 10−1 . A nonzero extension is beneficial, resulting in relative errors on the order of 10−2 ,
for the maximum value of N . The combination with inverse transportation allows us to further
improve the results, up to errors of 10−4 for N = 140 modes in the case of a POD basis obtained
13.3. Looking for a Better ROM with CutFEM 279
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
1
10
4
10
7
10
N / max
(i) 10 10
13 zero extension
10
harmonic extension
natural smooth extension
16
10 transport and harmonic extension
transport and natural smooth extension
19
10
zero extension
harmonic extension
natural smooth extension
transport and harmonic extension
22% transport and natural smooth extension
13%
Relative error
2%
0.8%
0.2%
0.04%
(ii) 35%
17%
10%
3%
0 10 20 30 40 50
N
Figure 13.10. (i) Poisson system (CutFEM): eigenvalue decay and error analysis between re-
duced order and high-fidelity approximations with and without transport; (ii) steady Stokes (CutFEM):
relative errors with and without transport.
280 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
from transport and natural smooth extension. Thus, the pivotal role of snapshot transportation can
be inferred from these results, being capable of improving the accuracy of almost three orders of
magnitude compared to the simplest zero extension. Figure 13.9 shows the first six POD modes
with and without preprocessing. Nonetheless, all methods reach a plateau after which no further
improvement is shown. We claim that this is due to integration errors occurring on Γ(µ) and
Nitsche weak imposition of the Dirichlet boundary conditions, e.g., the maximum values of the
error are consistently attained on the boundary.
Further investigation has shown similar good behavior for fluid flow systems, namely steady
Stokes and RB with transportation and CutFEM [322], with relative errors without and with
transport improvements visualized in Figure 13.10(ii). Once more, the transported snapshots and
the inverse transported modes appear beneficial for both velocity and pressure, needing very few
modes; in particular it manages to reach its best accuracy with only three basis components.
The equilibrium state of the considered mixture minimizes the above Ginzburg–Landau en-
ergy subject to mass conservation: ∂t
∂
= 0. Hence, the parametrized Cahn–Hilliard
R
D(µ)
u(µ)dx
system can be described as
∂u(µ) 1
= −ε2 ∆2 u(µ) + 2 ∆F 0 (u(µ)) in D(µ) × [0, T ], (13.10)
∂t ε
1
∂n u(µ) = ∂n (−ε2 ∆u(µ) + 2 F 0 (u(µ))) = gN (µ) on Γ(µ) × [0, T ], (13.11)
u(·, 0) = u0 (·) in D(µ), (13.12)
where n is the unit outward normal vector of Γ and F is a double-well function of u, usually a
fourth-power polynomial:
Figure 13.12. Cahn–Hilliard (CutFEM): ROM basis results, the first six modes.
Figure 13.13. Results for the embedded circle geometrical parametrization µ = 0.4261 and
t = [28, 36, 46, 100]dt.
Starting from the above initial form and based on a splitting method and Nitsche boundary
enforcement and with efficient CutFEM stabilization, we manage a proper weak form of H 1
space regularity; for more details we refer the reader to [324]. The CutFEM mesh and level-
set geometry can be seen in Figure 13.11, while some ROM basis components can be seen
in Figure 13.12, and the concentration field in the FOM and the ROM and the absolute error
level after parametrization of the embedded circle are visualized for the time instances t =
[28, 36, 46, 100]dt and for a randomly selected parameter µ = 0.4261 in Figure 13.13. The full-
order and the reduced mass evolution with respect to time verifies the conservation of mass for
the proper number of modes and for the truth solver and parameter µ = 0.4261. The experiment
took place for time instances t = ndt, n = 1, . . . , 100; see Figure 13.14. We mention that we
trained our basis with 900 snapshots in the parameter range for the diameter µtest ∈ [0.36, 0.48].
282 Chapter 13. RB, Embedded Methods, and Parametrized Level-Set Geometry
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
1
5.38 × 10
1
5.375 × 10
1
5.37 × 10
Mass(t)
3 modes ROM
5 modes ROM
1
5.365 × 10 7 modes ROM
11 modes ROM
21 modes ROM
1
5.36 × 10
29 modes ROM
41 modes ROM
50 modes ROM
1
5.355 × 10
FOM
0 20 40 60 80 100
Time
Figure 13.14. Cahn–Hilliard (CutFEM): the FOM mass evolution with respect to time, together
with its RB approximation and the conservation of mass.
Chapter 14
14.1 Introduction
This chapter is entirely devoted to the presentation of some reduced order model (ROM) pro-
cedures that are tailored for fluid–structure interaction (FSI) problems. Despite their intrinsic
complexity (see [185, 242]), FSI problems are a widespread topic in the applied mathematics
community: in naval engineering, FSI can be used to study and simulate the interaction between
the water and the hull of a ship (see, e.g., [379]); in a biomedical framework, multiphysics can
model the interaction between blood flow and the deformable walls of a vessel, as well as the be-
havior of blood flow in the presence of serious pathologies (as an example of the implementation
of FSI in the medical field, see [607, 466, 32, 469, 476, 384]); in aeronautical engineering, FSI
models the interaction of the air with a plane or with (parts of) a shuttle (see [478, 165, 187, 373]).
These are just some of the many applications for FSI, and each of these problems has different
characteristics that lead to different dynamics of the coupled system: e.g., the aerospace ap-
plications are in the setting of high Reynolds numbers and a turbulent regime; in addition, the
fluid is usually compressible and has a behavior that requires a three-dimensional model. On the
other hand, for biomedical applications, we are usually in a laminar regime, so the models can
be either two- or three-dimensional, and blood is usually modeled as an incompressible fluid.
Another very important difference in all these problems is represented by the physical quantities
describing the solid and the fluid, and mainly the solid density ρs and its ratio with respect to
the density ρf of the fluid under consideration: there are some algorithms for solving FSI prob-
lems that may work very well for a model that couples air with a very stiff material but cease
to work when the problem couples blood flow with biological tissue. Other difficulties come
from the different assumptions on the structure, and hence the different dynamics that one can
obtain; see, e.g., [47]. To summarize, FSI problems give rise to a wide range of different situ-
ations and different dynamics, requiring different computational tools, depending on the setting
in which we are working: the aforementioned difficulties are just some of the many peculiarities
of multiphysics problems that contribute to making it a very difficult yet extremely fascinating
topic of research.
In this chapter, according to the approach adopted to address a multiphysics problem, we
will see how we can adapt, modify, and improve the reduced basis method (RBM); all the ROMs
283
284 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
presented are based on a proper orthogonal decomposition (POD) [249, 351, 482, 620]. In the
following, we will consider two different problems of interest: in the first test case we model the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
behavior of two leaflets that bend under the influence of a fluid jet, whereas in the second test
case the solution of the FSI problem exhibits a transport-dominated behavior.
• Continuity of the displacement: This condition imposes continuity on the solid displace-
ment and the fluid displacement at the fluid-structure interface. This geometrical condition
is the mathematical translation of the hypothesis that the fluid and the solid domains do
not overlap.
• Continuity of the velocity: This is a kinematic condition that represents the hypothesis
that the fluid sticks to the moving fluid-structure interface. This condition resembles in
some ways the no-slip boundary condition, which is very common in viscous fluid dy-
namics. In aeroelasticity applications, the continuity of the velocity at the fluid-structure
interface is relaxed and substituted with the nonpenetrating condition, which prescribes
the motion in the direction normal to the fluid-structure interface.
• Balance of the stresses: This is a classical action-reaction principle that imposes the
balance between the fluid and the solid stresses at the fluid-structure interface.
These coupling conditions are imposed at the fluid-structure interface: one of the big challenges
of coupled problems is the fact that this interface is not fixed; instead it deforms in time, and
the deformation is not known a priori. Another difficulty is the fact that the fluid domain in
FSI applications is a moving domain: in solid mechanics it is common to deal with moving
domains, where the deformation is usually the unknown of the problem. For fluid dynamics, on
the other hand, one usually considers fixed domains. This different point of view is an intrinsic
characteristic of FSI problems, and it gives rise to a formalism that is very well known and widely
used in the community, called the arbitrary Lagrangian Eulerian (ALE) formulation.
x
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Ω̂ Ω(t)
Figure 14.1. Example: domain reference configuration Ω̂ (left) and domain configuration at time
t, Ω(t) (right). In blue we have the fluid domain; in orange the solid domain. In green is the fluid-structure
interface Γ̂F SI in the reference configuration.
the “initial position” of a given point inside the solid physical domain. On the contrary, when
describing the behavior of a fluid, the definition of “domain reference configuration” or “original
position of a particle” is not clear, and therefore the Eulerian formalism is used instead: the
conservation laws are formulated on the configuration Ωf (t) at the current time t. In order to be
able to describe both the fluid and the solid behavior, a mixed formulation (the ALE formulation)
is used. The underlying idea is that we can pull back the fluid equations, rewriting them in an
arbitrary time-independent configuration Ω̂f : one possible choice for Ω̂f is Ω̂f = Ωf (t = 0),
the domain at the initial time. In Figure 14.1 we can see an example of a reference configuration
and the configuration of the domain at the current time t. Let [0, T ] be a time interval, and let Ω̂f
be a reference configuration for the fluid.
Definition 14.1. The ALE mapping Af (t) for every t ∈ [0, T ] is defined as follows:
Here dˆf (t): Ω̂f 7→ Ω̂f is the mesh or fluid displacement. The definition of dˆf usually changes
depending on the kind of fluid problem we want to model, and also depending on the degree of
regularity that we would like to see in the solution of the coupled system; usually the mesh
displacement is defined as an extension of the solid displacement to the whole fluid domain.
Let Ω̂s ⊂ R2 be the reference configuration for the solid, and let dˆs (t): Ω̂s 7→ R2 be the solid
displacement: one possible way to define dˆf is by a harmonic extension of dˆs :
ˆ dˆf = 0 in Ω̂f ,
−∆
(14.2)
df = dˆs on Γ̂F SI ,
ˆ
where Γ̂F SI is the fluid-structure interface in the reference configuration and ∆ ˆ denotes the
gradient with respect to the coordinates in the reference configuration. For other alternatives on
how to define dˆf , we refer the reader to [487].
We underline that dˆf represents the displacement of the grid points, so it is not a quantity
with a real physical meaning but rather a geometrical quantity that describes the deformation of
the mesh according to the deformation of the physical domain. It is also important to underline
that ∂t dˆf 6= ûf ; in fact, while ûf represents the velocity of the fluid, ∂t dˆf is again a geometrical
quantity that can be interpreted as the velocity with which the mesh moves.
Great attention has to be paid to the definition of the mesh displacement because different
definitions for dˆf lead to different levels of regularity: if we lose regularity due to the mesh
displacement, then we lose regularity at the fluid-structure interface, which is exactly where the
coupling between the two physics takes place. It is beyond the scope of this chapter to discuss
286 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
the regularity of different definitions of the mesh displacement; we refer the interested reader to
Subsection 5.3.5 of [487].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
• Explicit algorithms: After having discretized the FSI problem in time, the conditions
on the continuity of the displacement and on the continuity of the velocity at the inter-
face are treated explicitly at every time step. These algorithms, also known as weakly
or loosely coupled algorithms [84], are successfully applied in aerodynamics applications
(see [188, 455]), but some studies (see [102, 305, 365]) show that they are unstable under
some physical and geometrical conditions due to the added mass effect, as we previously
mentioned.
• Implicit algorithms: In these algorithms, also known as strongly coupled algorithms,
the coupling conditions are treated implicitly at every time step (see, e.g., [605, 606]).
This implicit coupling represents a way to circumvent the instability problems due to the
added mass effect; nevertheless, an implicit treatment of the coupling conditions leads to
algorithms that are more expensive in terms of computational time.
• Semi-implicit algorithms: In these algorithms (see [27, 28, 45]), the continuity of the dis-
placement is treated explicitly, whereas the other coupling conditions are treated implicitly.
This alternative represents a trade-off between the computational cost of the algorithm and
its stability in relation to the physical and geometrical properties of the problem. In this
chapter we will see a reduced order method that is based on this kind of partitioned ap-
proach.
14.4. Partitioned RBM 287
In a monolithic algorithm (see, e.g., [36, 487, 216, 201, 29]), the fluid and the solid problem
are solved simultaneously. This usually results in algorithms that are more stable, which is
highly desirable, especially if we wish to use large time steps in our simulations. The main
drawback is that they deeply rely on the availability of ad hoc software that can be used to solve
the fluid problem and the solid problem; in this sense, monolithic algorithms are less flexible and
more tailored to the particular problem at hand. Moreover, the coupling conditions at the fluid-
structure interface are more complicated to treat; in order to pursue a Galerkin discretization of
the original problem of interest, we will use two Lagrange multipliers to impose the continuity
of the displacement and of the velocity at the interface: this results in the introduction of two
new unknowns in the coupled problem.
ΓD
s Γtop
Γin Γout
µs ←
}µg
ΓD
s Γbottom
Figure 14.2. Original configuration at time t. Blue domain: the original fluid configuration
Ωf (t). Orange leaflets: the original solid configuration Ωs (t). ΓD
s : the part of the leaflets that does not
move.
288 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
Let us denote by µ = (µg , µs ) the parameter vector, with µ ∈ P ⊂ R2 , where P := [µming , µmax
g ]
× [µs , µs ] is the parameter domain. Let us denote by Ω(t; µg ) := Ωf (t; µg ) ∪ Ωs (t; µg )
min max
the current physical domain: we now have a time dependence and a parameter dependence.
We introduce the time-independent intermediate configuration Ω̃(µg ) := Ω̃f (µg ) ∪ Ω̃s (µg ),
where we are considering the reference configuration of both physics, still taking into account the
parameter dependence. Finally, we have the time-independent, parameter-independent reference
configuration Ω̂ := Ω̂f ∪ Ω̂s .
We call T the shape parametrization map: for every µg ∈ [µming , µmax
g ], we have a map Tµg
defined as follows:
Tµg : Ω̂ 7→ Ω̃(µg ),
x̂ 7→ x̃ = Tµg (x̂).
We then have the ALE map Af (t; µg ), introduced in Subsection 14.2.1, which is now a map from
the current parametrized fluid configuration Ω̂f (t; µg ) to the intermediate fluid configuration
Ω̃f (µg ):
˜ +∇
F̃ (x̃; µg ) = Id ˜ d˜f , ˜ x̃; µg ) = detF̃ ,
J(
where now ∇ ˜ represents the gradient with respect to the coordinates x̃ of the intermediate con-
figuration. We can pull back the gradient F̃ (x̃; µg ) to the reference domain Ω̂f , and we obtain
F (x̂; µg ) = Id + ∇ ˆ dˆf G−1 (x̂, µg ). With this notation, we can conclude that the gradient of
the deformation map from the reference configuration to the current configuration is given by
F (x̂, µg )G(x̂, µg ); for the sake of simplicity of notation let us denote by Fµg and Gµg the
gradients F (x̂, µg ) and G(x̂, µg ), respectively.
Problem Formulation
With the notation introduced in the previous paragraph we are now ready to state the strong
formulation in the current configuration: for every t ∈ [0, T ] and for every µ ∈ P, find the
fluid velocity uf (t; µ): Ωf (t; µ) 7→ R2 , the fluid pressure pf (t; µ): Ωf (t; µ) 7→ R, the mesh
displacement d˜f (t; µ): Ω̃f (µ) 7→ R2 , and the solid displacement d˜s (t; µ): Ω̃s (µ) 7→ R2 such
that
˜ d˜f = 0 in Ω̃f (µ) × [0, T ],
−∆
(14.3)
˜
df = d˜s on Γ̃F SI × [0, T ]
and
where ∂t df |x̃ (x; t) = ∂t d˜f (x̃; t) is the ALE derivative (and similar for ∂t uf |x̃ ), ρf is the fluid
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
density, ρs is the solid density, and bf and b̃s are the external volume forces acting on the fluid
and on the solid, respectively, on the domains Ωf (t; µ) and Ω̃s (µ). In system (14.4), div() is the
divergence with respect to the coordinates x, whereas div() ˜ is the divergence with respect to the
intermediate coordinates x̃. σf is the fluid Cauchy stress tensor in the configuration Ωf (t; µ):
with I the 2 × 2 identity matrix and µf the fluid viscosity. P̃ is the second Piola–Kirchoff stress
tensor expressed in the intermediate configuration:
with λs the first Lamé constant. System (14.4) is completed by the coupling conditions
df = ds on ΓF SI (t; µg ) × [0, T ],
uf = dt d
ds on ΓF SI (t; µg ) × [0, T ], (14.7)
J σ̃f (ũf , p̃f )F̃ −T ñf = P̃ (d˜s )ñs on ΓF SI (t; µg ) × [0, T ]
˜
where ñf and ñs are the normal vectors to the fluid-structure interface Γ̃F SI in the intermediate
configuration, out of the fluid domain and the solid domain, respectively. Moreover,
˜ ũf F̃ −1 + F̃ −T ∇
σ˜f (ũf , p̃f ) := µf (∇ ˜ T ũf ) − p̃f I (14.9)
ˆ
div(JKG µg Fµg uf ) = 0 in Ω̂f × [0, T ],
−1 −1
where
ˆ ûf G−1
σ̂f (ûf , p̂f ) = µf (∇ −1 −T −T ˆ T
µg Fµg + Fµg Gµg ∇ ûf ),
The coupling conditions (14.7) are pulled back onto the reference fluid-structure interface Γ̂F SI :
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Here n̂f and n̂s are the normals to the fluid-structure interface in the reference configuration,
out of the fluid domain and the solid domain, respectively.
We remark that, from now on, in order to ease the notation and the exposition, we will drop
the notation with the ∧, since everything is intended to be taken in the reference configura-
tion Ω̂.
endowed with the H 1 -norm (V (Ωf ) and E f (Ωf )) and the L2 -norm, respectively, and the func-
tion space for the solid, E s (Ωs ) = [H 1 (Ωs )]2 , endowed with the H 1 -norm. We discretize in
space the FSI problem, using second-order Lagrange FEs for the fluid velocity, the fluid dis-
placement, and the solid displacement, resulting in the discrete spaces Vh ⊂ V , Ehf ⊂ E f , and
Ehs ⊂ E s , while the fluid pressure is discretized with first-order Lagrange FEs, resulting in the
discrete space Qh ⊂ Q.
After space discretization, the problem of interest is defined as follows: for every µ =
(µg , µs ) ∈ Ptrain , and for i = 0, . . . , NT , we have the following:
f f
• Extrapolation of the mesh displacement: Find di+1
f,h ∈ Eh such that ∀ef,h ∈ Eh ,
( R
K∇di+1 −1 −1
f,h Gµg · ∇ef,h Gµg dx = 0,
Ωf
(14.12)
di+1
f,h= dis,h on ΓF SI .
ui+1 −ui
· vh dx + ρf Ωf JK[∇ui+1 −1 −1 i+1
R f,h
R
f,h Gµg Fµg ]uf,h · vh dx
f,h
ρ f Ωf
JK ∆T
i+1 i+1
− ρf Ωf JK[∇uf,h G−1 −1
R
µg Fµg ]Dt df,h · vh dx
i+1 −T −T (14.13)
R
+ µf Ωf JKε(uf,h )Fµg Gµg : ∇vh dx
−T −T i
R R
+ Ωf JKFµg Gµg ∇pf,h · vh dx = Ωf JKbf · vh dx,
i+1 i+1
uf,h = Dt df,h on ΓF SI ,
ρf
Z Z
− −1 −1 i+1
div(JKGµg Fµg uf,h )qh dx + αROB pi+1,j
f,h qh ds
∆T Ωf ΓF SI
Z Z
i+1,j i+1,j+1
− ρf (Dtt ds,h ) · JKFµ−T g
G−T
µg nf qh ds = αROB pf,h qh ds
ΓF SI ΓF SI
Z
i+1,j+1
+ JKG−T −T
µg Fµg ∇pf,h · G−T −T
µg Fµg ∇qh dx,
Ωf
Steps 1 and 2 are iterated until a convergence criterion is satisfied. In our case, this stopping
criterion is given by
i+1,j+1 i+1,j
||Q ||di+1,j+1 − di+1,j
( )
||pf,h − pf,h s,h s,h ||E
max i+1,j+1
; i+1,j+1
< ε, (14.14)
||pf,h ||Q ||ds,h ||E
where ε is a tolerance of our choice. To enhance the stability of the fluid-solid subproblem
(implicit step), we used a Robin coupling condition in the pressure Poisson formulation:
ρf
The constant αROB is defined as αROB = zp ∆T , where zp is called the solid impedance:
zp = ρs cp ,
s
λs + 2µs
cp = .
ρs
Introducing the Robin coupling between the fluid and the solid problem improves the stability
properties of the implicit step; computing the Robin coupling can be straightforward, e.g., in the
coupling of a fluid with a thin, one-dimensional structure (see, e.g., [26, 22]), or it may be more
complicated, such as in our test case, where the structure is thick and two-dimensional and the
292 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
influence of the fluid on the solid does not appear immediately from the beginning in the solid
problem formulation. For this second situation, we refer the reader to [39].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
The main idea here is to introduce a change of variable in the fluid problem, in order to transform
condition (14.132 ) into a homogeneous boundary condition. The motivation for this choice is
that, to impose condition (14.132 ), we could use a Lagrange multiplier λ, but unfortunately
introducing a new variable leads to an increased dimension of the system to be solved in the
online phase. Therefore, in order to avoid this and design a more efficient reduced method, we
choose to transform the nonhomogeneous coupling condition into a homogeneous one. In order
i+1
to do this, we define a new variable zf,h :
i+1
zf,h := ui+1 i+1
f,h − Dt df,h . (14.15)
With this change of variable, (14.132 ) is equivalent to the homogeneous boundary condition for
the new variable:
i+1
zf,h = 0 on ΓF SI , (14.16)
for which no imposition of the Lagrange multiplier is needed. Therefore, during the offline phase
of the scheme, at every iteration i + 1, after we have computed the velocity ui+1
f,h , we compute
i+1
the change of variable zf,h . We then consider the snapshot matrix
h
1 NT
S z = [zf,h , . . . , zf,h ] ∈ RNu ×NT , (14.17)
For this test case, the idea is to first perform a standard POD on the set of snapshots computed
for every value of the parameter µ in the training set Ptrain ; subsequently, we take the modes
computed with the standard POD, weighted with the corresponding eigenvalue, and we perform
a final outer run of POD; for a more detailed discussion we refer the reader to [415].
We start by constructing, for each parameter µi,j = (µig , µjs ) ∈ Ptrain , the snapshot matrices
S z (µi,j ) for the fluid change of variable zf , S p (µi,j ) for the fluid pressure pf , and S ds (µi,j )
for the solid displacement ds :
h
S z (µi,j ) = {zf,h (t0 , µi,j ), . . . , zf,h (tNT , µi,j )} ∈ RNu ×NT ,
h
S p (µi,j ) = {pf,h (t0 , µi,j ), . . . , pf,h (tNT , µi,j )} ∈ RNp ×NT ,
h
S ds (µi,j ) = {ds,h (t0 , µi,j ), . . . , ds,h (tNT , µi,j )} ∈ RNds ×NT ,
where Nph = dimQh and Ndhs = dimEhs . We then perform a standard POD on each snapshot ma-
Nz Np Nd
trix and we extract the basis functions {Φkzf (µi,j )}k=1
i,j
, {Φkp (µi,j )}k=1
i,j
, and {Φkds (µi,j )}k=1
i,j
.
Nz Np Nd
Let us also call {λ(µi,j )zk }k=1
i,j
, {λ(µi,j )pk }k=1
i,j
, and {λ(µi,j )dk }k=1
i,j
the eigenvalues, in decreas-
ing order of magnitude, returned by the POD on each snapshot matrix S z (µi,j ), S p (µi,j ), and
S ds (µi,j ). Afterwards, we perform a second run of POD in the following way: we start by
14.4. Partitioned RBM 293
building the snapshot matrices, normalizing each snapshot with the corresponding eigenvalue
given by the standard POD:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
q z
z
q NM
λ(µ1,1 )1 Φ1zf (µ1,1 ), . . . , λ(µMg ,Ms )zN z
,M
Sz = Φzf g s (µMg ,Ms ) ,
Mg ,Ms
p
q
p NM
λ(µ1,1 )1 Φ1p (µ1,1 ), . . . , λ(µMg ,Ms )pN p
q ,M
Sp = Φp g s (µMg ,Ms ) ,
Mg ,Ms
(q )
r d
d NM
S ds = λ(µ1,1 )1 Φ1ds (µ1,1 ), . . . , λ(µMg ,Ms )dN d Φds g ,Ms
(µMg ,Ms ) .
Mg ,Ms
We then perform a second POD on the previous snapshot matrices, and finally we obtain a set of
z
k Np k Nd
basis functions {Φkzf }N
k=1 , {Φp }k=1 , and {Φds }k=1 .
In order to generate the reduced basis for the fluid displacement df (µ), we pursue the idea
d
presented in [37]. Once we have obtained the reduced basis {Φkds }Nk=1 for the solid displace-
ment, we employ a harmonic extension of each one of the reduced bases Φkds on the whole fluid
domain:
(
−∆Φkdf = 0 in Ωf ,
(14.18)
Φkdf = Φkds on ΓF SI .
We can then define the reduced space for the fluid displacement as
f
EN := span{Φkdf }N
k=1 .
d
(14.19)
This choice is extremely efficient: the first advantage is that we can avoid introducing another
Lagrange multiplier to impose the nonhomogeneous boundary condition (14.122 ) because it is
automatically satisfied by every Φkdf and hence by every linear combination of these basis func-
tions. In addition, with our method, we avoid solving the reduced system related to (14.12):
instead of solving a harmonic extension problem at every time step in the online phase, we solve
all Nd harmonic extension problems at once in the expensive offline phase.
The reduced problem then reads as follows: for every i = 0, . . . , NT and for µ = (µg , µs ) ∈ P,
do the following:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
i+1
• Fluid explicit step (with change of variable): Find zf,N (µ) ∈ VN such that ∀vN ∈ VN ,
z i+1 (µ) − ui
f,N (µ)
Z Z
f,N i+1
ρf JK · vN dx + µf JKε(zf,N (µ))Fµ−T G−T
µg : ∇vN dx
Ωf ∆T Ωf
g
Z
i+1 −1 i+1
+ ρf JK∇zf,N (µ)G−1
µg Fµg zf,N (µg ) · vN dx
Ωf
Z
+ ρf JK∇Dt di+1 −1 −1 i+1
f,N (µ)Gµg Fµg zf,N (µ) · vN dx
Ωf
Z Z Dt di+1 (µ)
f,N
+ JKFµ−T G−T i
µg ∇pf,N (µ) · vh dx = −ρf JK · vN dx
Ωf
g
Ωf ∆T
Z Z
− µf JKε(Dt di+1 −T −T
f,N (µ))Fµg Gµg : ∇vN dx + JKbf,N · vN ,
Ωf Ωf
where bf,N is the projection of the fluid external force bf onto the reduced space VN . We
then restore the reduced fluid velocity: ui+1 i+1 i+1
f,N (µ) = zf,N (µ) + Dt df,N (µ).
• Implicit step: For any j = 0, . . . until convergence do the following:
ρf
Z Z
− div(JKG−1 −1 i+1
µg Fµg uf,N (µ))qN dx + αROB pi+1,j
f,N (µ)qN ds
∆T Ωf ΓF SI
Z Z
i+1,j
− ρf Dtt ds,N (µ) · JKFµ−T
g
G−T
µg nf qN ds = αROB pi+1,j+1
f,N (µ)qN ds
ΓF SI ΓF SI
Z
i+1,j+1
+ JKFµ−T
g
G−T
µg ∇pf,N (µ) · Fµ−T
g
G−T
µg ∇qN dx,
Ωf
where bs,N is the projection of the solid external force bs onto the reduced space EN
s
.
14.4. Partitioned RBM 295
Tµg
x
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Ω̂ Ω̃(µg )
↓ Af (t; µg )
Ω(t; µg )
Figure 14.3. Domains: reference configuration Ω̂ (top left), parametrized reference configuration
Ω̃(µg ) (top right), and original configuration Ω(t; µg ) (bottom).
and Tin = 0.4 s. The tolerance chosen for the implicit step of the whole algorithm is ε = 10−6 ,
and the time step used in the problem time discretization is ∆T = 10−4 .
For the simulation, we use a time step of ∆t = 10−4 for a maximum number of time steps
NT = 800; thus T = 0.08 s. Table 14.1 summarizes the details of the offline stage and of the
finite element (FE) discretization.
Figure 14.4 shows three different examples of the behavior of the leaflets, according to the
change of the physical and/or of the geometrical parameters: all the pictures represent the online
displacement of the solid at the final time step of the simulation, namely for t = 0.08 s. The
results were obtained using Nd = 100 reduced bases for the displacement; as we can see, in-
dependent of the length of the leaflets, an increase in the shear modulus leads to a material that
is much harder to deform. On the contrary, for a fixed value of the properties of the material
under consideration, an increase in the length of the leaflets leads to an increase in the displace-
ment. As the results show, our ROM is able to capture small deformations of the solid, and it is
also able to reproduce variations and changes in the behavior of the leaflets, also at very small
magnitudes.
296 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
Table 14.1. Physical and geometrical constants and parameters for the geometrically and physi-
cally parametrized leaflets test case.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 14.4. Reduced order solid displacement dN s (µ). Comparison of different behaviors of the
material, for different values of the geometrical and physical parameters. From left to right: same leaflet
length µg = 0.8 cm and increased shear modulus (µs = 100000, 800000); same leaflet length µg = 1 cm
and increased shear modulus (µs = 100000, 800000); increased leaflet length (µg = 0.8, 1.0 cm), and
same shear modulus µs = 100000.
variety of situations, there is still a challenging task for these procedures: advection-dominated
problems; see, e.g., [428, 244, 181]. In the last years it has therefore become clear that a modifi-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
cation of the standard RBM is necessary, especially in order to obtain a small basis set in these
more challenging situations.
Before going any further, we will try to explain the difficulties that arise when we are in the
presence of transport-dominated problems.
N (z(η); η) = 0, (14.25)
Definition 14.2. Let z(η) be a component of z(η), and let Mz be the corresponding solution
manifold, embedded in some normed linear space (Xz , || · ||Xz ), defined as
From this definition we understand that, given a problem of interest, the Kolmogorov n-width
Dn represents the entity of the error that we commit by approximating any element f of Mz with
an element g of a linear space En of dimension n. Of course this means that the faster Dn decays
as we let n grow, the better chance we have to find a good linear approximation space for Mz
of small dimension. In the majority of problems there is no explicit analytic formula for Dn , yet
there are some situations where we can compute good bounds on it; see, e.g., [393, 133]. Since
it is very difficult to provide bounds on Dn , one can only hope that the n-width of the solution
manifold of the problem of interest is small. A heuristic way to check that this hypothesis is
satisfied is to run a POD on a set of snapshots and check the rate of decay of the eigenvalues
{λi }i returned by the POD: if the {λi }i decay fast, then we can expect to be able to build a good
low-dimensional linear approximation space for Mz . Unfortunately, the problem is that this
assumption often fails in transport-dominated problems, which usually show a very slow decay
298 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
of the eigenvalues: it therefore becomes clear that the standard RBM is unable to reconstruct any
element of Mz by using a small number of basis functions.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In recent years a growing number of works have appeared that focus on constructing alter-
native (nonlinear) model order reduction techniques, that can be effectively applied to transport-
dominated problems: we refer the interested reader to some examples of these works, such as
[9, 221, 222, 181, 99, 18, 442, 303, 57, 88, 87, 613, 614, 411, 567, 322, 427, 500, 59]. In
[367], the authors attempt to design a reduction procedure that is not strictly related to the phys-
ical phenomenon under investigation, e.g., transportation of a quantity or translation of a wave.
The authors propose a methodology which focuses on optimal projection of general dynamical
systems onto arbitrary nonlinear trial manifolds in the online stage, where this nonlinear trial
manifold is computed from snapshot data alone in the offline stage, thanks to the employment of
convolutional autoencoders. The reader interested in the use of machine learning techniques in
the model order reduction framework is referred to in [206, 264]. A methodology that is light,
simple, more natural, based on the definition of some transport maps is introduced and applied to
some toy problems by Cagniart [86] and Cagniart, Maday, and Stamm [88]. The goal of this sec-
tion is to present an application of model reduction based on transported snapshots for problems
in fluid dynamics and FSI, focusing in particular on comparing results of the standard offline
phase and the novel one, which relies on transport.
Maps in the family Fz depend on the same η appearing in (14.25) and are essentially problem
specific in the sense that different maps (translation, dilatation) should be employed in different
settings: translation maps will be enough in some simple periodic settings, while dilatation maps
are more suited in the case of nonperiodic boundary conditions. In any case, the main goal of
these maps is to align some feature of the solution of the problem under consideration (e.g., a
shock or a peak) with a fixed point. In [88], e.g., a family of translations Fη (x) = x − γ(η),
x ∈ Ω ⊂ R, γ(η) ∈ Ω, is employed for the viscous Burgers equation, parametrized by the
parameter η: there, γ(η) is chosen so that the steepest points of the solutions coincide.
Before going any further, let us make the following assumption: the problem of interest is time
dependent, and no other parameter appears in the formulation; thus η = t and E = [0, T ]. Given
the family Fz of smooth and invertible mappings, we can introduce the preprocessed solution
manifold.
Assuming that Fz is carefully chosen, MFz has a smaller Kolmogorov n-width than that of
Mz . The practical realization of this preprocessing procedure is incorporated in the offline phase.
Given a time discretization of the original problem [0 = t0 , . . . , tNT = T ], we compute each
solution component z(ti ) with i = 0, . . . , NT . The discrete approximations of the corresponding
14.5. A Monolithic RBM for a Transport-Dominated FSI Problem 299
Mtr i
z = {z(t ), i = 0, . . . , NT },
Mtr
Fz = {z(t ) ◦ Ft−1
i
i , i = 0, . . . , NT },
provide snapshots for compression by a POD. The compression is applied here to both Mtr z
and Mtr Fz to compare the standard offline phase and one with preprocessing, but in practical
computations one would neglect the compression of Mtr z because it is understood that MFz
tr
would result in a POD basis set {Φk }N k=1 of lower dimension. Assume now that an accurate
approximation zN (ti ) of the solution component z(ti ) at time ti is known as a linear combination
of a pullback of the basis functions Φk :
N
X
zN (ti ) = αki Φk ◦ F̃ti , (14.30)
k=1
the idea proposed in [88] is to iterate between the search for the reduced coordinates αki+1 and the
search for a suitable pullback map F̃ti+1 . This procedure is carried out by means of a minimiza-
tion problem of the L2 -norm of some residual. In this book, we do not follow a minimization
approach: we employ instead a polynomial interpolation method to learn the best-suited pullback
map F̃ti+1 to be used in the online phase.
To summarize, here is an outline of the general RBM with preprocessing of the snapshots:
PN
• during the online phase, at time step ti+1 , find zN (ti+1 ) = k=1 αki+1 Φk ◦ F̃ti+1 , where
F̃ti+1 is a suitable pullback map.
Problem Formulation
A two-dimensional rectangle of height hf and length L is filled with a Newtonian fluid. The
top and the bottom walls of the rectangle represent the structure, which is considered to be
deformable, and its thickness is negligible with respect to the height of the rectangle. Since the
300 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
%
Γin Γout
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Ω̂ Σ̂ &
Figure 14.5. Physical domain (reference configuration): fluid subdomain (blue) and structure
subdomain (orange). The fluid-structure interface coincides with the structure in our case; the structure
has been magnified for visualization purposes.
structure is thin, it is described by a one-dimensional model. We further assume that the displace-
ment of the walls in the horizontal direction is negligible, and hence the structure undergoes only
vertical motion: the behavior of the compliant walls is therefore described by the generalized
string model [469, 476]. We want to study the behavior of the solution in the time interval [0, T ].
Let Σ(t) and Ωf (t) be, respectively, the structure and the fluid domain at time t. Let Σ̂ be the
solid reference configuration (undeformed walls), and let Ω̂f be an arbitrary fluid reference do-
main: for convenience we take Ω̂f = Ωf (t = 0) (the blue fluid domain in Figure 14.5). Thanks
to the introduction of the ALE map Af (t) defined in Subsection 14.2.1, we can map the fluid
equations back to the reference domain Ω̂f . Let F be the gradient of Af (t), and let J be the
Jacobian. The FSI problem, formulated on the reference configuration, reads as follows: find the
fluid velocity ûf (t): Ω̂f → R2 , the fluid pressure p̂f (t): Ω̂f → R, and the structure displacement
dˆs (t): Σ̂ → R such that
ρs hs ∂tt dˆs − c0 ∂xx dˆs + c1 dˆs = −σ̂ f (ûf , p̂f )n̂ · n̂ in Σ̂ × [0, T ].
(14.32)
Here ρf is the fluid density, ρs is the structure density, hs is the structure height (or thickness),
c0 and c1 are the structure constitutive parameters, and n̂ is the outward normal to Σ̂. σ̂ f is
the Cauchy stress tensor for the fluid, in the reference configuration, which was introduced in
Subsection 14.4.1.
In order to ease the notation, we drop the ∧, since everything will be formulated in the
reference configuration. The coupling conditions are
df = ds n in Σ × [0, T ],
(14.33)
uf = ∂t ds n in Σ × [0, T ].
We further assume that at time t = 0 the system is at rest; finally, the boundary conditions can
be summarized in the following system:
where the third condition says that the structure is fixed at its extremities.
where
H01 (Ωf ) := {ef ∈ [H 1 (Ωf )]2 such that ef = 0 on Γin ∪ Γout } (14.38)
and
H01 (Σ) := {es ∈ H 1 (Σ) such that es = 0 on ∂Σ}, (14.39)
and where V is endowed with the H 1 -norm; Q with the L2 -norm; and E f , E s , Lu , and Ld with
the H1 seminorm. Let us now consider some finite-dimensional subspaces Vh ⊂ V , Qh ⊂ Q,
Ehf ⊂ E f , Ehs ⊂ E s , Luh ⊂ Lu , and Ldh ⊂ Ld of dimension N u , N p , Nfd , Nsd , Nuλ , and Ndλ ,
respectively. We discretize the time interval [0, T ] with the following partition:
for all vh ∈ VH , qh ∈ Qh , ef,h ∈ Ehf , es,h ∈ Ehs , muh ∈ Luh , and mdh ∈ Ldh . To obtain a stable
approximation of the fluid pressure during the online phase of the method, we adopt a supremizer
enrichment technique. In this case, the supremizer sih ∈ Vh is defined by the following problem:
for any i = 0, . . . , Nt , find sih ∈ Vh such that
Z Z Z
∇sih : ∇vh dx + sih · vh dx = − pif,h divvh dx (14.42)
Ωf Ωf Ωf
u p N p
for all vh ∈ Vh . Let {ϕuj }N
j=1 be a set of basis functions for Vh , {ϕj }j=1 a set of basis functions
d Nfd Nd
for Qh , {ϕj f }j=1 a set of basis functions for Ehf , {ϕdj s }j=1
s
a set of basis functions for Ehs ,
302 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
λu λd N d λ
{ϕλj u }N
j=1 a set of basis functions for Lh , and {ϕj }j=1 a set of basis functions for Lh , where
u d
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
f
N u = dimVh , N p = dimQh , Nfd = dimEh , Nsd = dimEhs , N λu = dimLuh , and N λd =
dimEhd . We introduce the vector of degrees of freedom associated with the velocity components
of the solution:
u
N
(j) u u X (j)
uf := (uf,h )N
j=1 ∈ R
N
⇐⇒ uf,h = (uf,h )ϕuj ∈ Vh , (14.43)
j=1
and similarly for the other components of the solution. Then the system (14.41) is equivalent to
the following algebraic system:
ϕdj s n · ϕλk u
Z
(L2u )j,k = dx,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
∆T
ZΣ
d
(L1d )j,k = ϕλj d · ϕkf dx,
ZΣ
(L2d )j,k =ϕλj d · nϕdks dx,
Σ
Z Z
(Fu )j = − pin (ti )n · ϕuj dx + bi · ϕuj dx. (14.44)
Ωf Ωf
Table 14.2. Problem data for the test case of a flow in a channel with deformable walls.
Figure 14.6. Fluid pressure behavior: the solution is pictured here at times t = 0.001, t = 0.005,
and t = 0.015. The peak of the wave is propagating into the domain, creating a transport phenomenon.
304 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 14.7. Vertical component of fluid displacement behavior: again the solution is pictured at
times t = 0.001, t = 0.005, and t = 0.015. The peak of the wave is still very small at the beginning, it
grows for some time, and then it starts to propagate.
Figure 14.8. Decay of the eigenvalues for the POD on the fluid pressure (blue line), fluid dis-
placement (green line), and fluid velocity (magenta line).
In order to make the following exposition clearer and easier to read, from now on we focus
only on a particular component of the solution of our problem, namely the fluid pressure pf . In
any case, it is important to keep in mind that, based on our simulation, all the components of the
solution to the FSI problem are subject to a transport phenomenon, and hence every consideration
that we make on pf can be easily applied to any of the other components of the solution.
Preprocessing Step
We apply the preprocessing procedure to the fluid pressure solution manifold Mpf . Figure 14.6
shows how the peak of the pressure is transported in the domain. The idea is to move, at every
time step, the peak of the wave to a fixed position of our choice; in this way, we obtain a set of
snapshots where the pressure wave is not moving at all.
Starting from this observation, we build a one-parameter family of mappings
such that for every t in [0, T ], the peak of pf (Ft−1 (·), t) is located not at its original position
but is instead moved to the middle of the domain. A suitable map to be used in this problem
(nonperiodic setting) is a dilatation map Ft−1 :
3xγ
Ft−1 (x) = , (14.46)
x(γ − 3) + 3(6 − γ)
where γ = γ(t) is the abscissa of the position of the peak of the wave at time t. We assume that
the abscissa of the points on the inlet boundary Γin is x = 0 and the length of the domain Ωf is
L = 6; in addition, the position to which we are moving the peak of the wave at every time t is
exactly in the center of the domain. We remark that the map Ft−1 is just a stretching in the hori-
zontal direction: this is because the solution of the original problem does not exhibit any transport
14.5. A Monolithic RBM for a Transport-Dominated FSI Problem 305
phenomena in the vertical direction, and hence there is no need for a transformation in the y-axis.
So, with an abuse of notation, we can think of Ft−1 (x, y) as Ft−1 (x, y) = (Ft−1 (x), y).
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Our family of mappings Fpf is a one-parameter family; therefore to identify the stretching
map Ft−1 we have to identify the parameter γ(t) at time t. One simple way of doing this is to use
polynomial interpolation: we approximate the physical law governing the position of the peak of
the wave γ(t) with a polynomial qk (t) of degree k. Of course, when the physical phenomenon
under investigation becomes particularly complex, or in the presence of multiple parameters,
other techniques may become more useful and better performing, such as polynomial regression
or neural networks, presented in [579]. For this toy problem, however, a polynomial of degree
k = 1 is sufficient to obtain very good results in the online phase.
First of all we compute all the snapshots p1 , . . . , pNT , where NT = ∆t
T
, pi = pf (ti ), and
t = i∆t. Once we have the snapshots, we compute the peak γTin and γNT . The reason for
i
choosing γTin instead of a more straightforward γ0 is that usually in computational fluid dy-
namics (CFD) applications, complex phenomena such as vortex formation and propagation of a
quantity require some time to develop, unlike the wave propagation problem presented in [88] or
[579], e.g. For our problem, Tin is a reasonable starting time for the preprocessing procedure:
recall the expression of pin from Table 14.2; as we can see, pin represents a given pulse at the
inlet boundary, which is nonzero up to time t = Tin = 0.0025. So, up to time t = 0.0025
the pulse makes the wave grow; when the pulse is null, the wave starts to propagate into the
domain. At the reduced level, this translates to the fact that up to t = Tin we adopt a standard
reduction technique, using a standard reduced basis obtained with a standard POD on the first
Tin snapshots. From the snapshot Tin to the last one, on the other hand, we first perform the
preprocessing step and then perform a POD.
In Figure 14.9 (top) we can compare the decay of the eigenvalues for the POD on the pressure
with and without this preprocessing procedure. As we can see, with the preprocessing technique
we do actually get an improvement in the decay of the eigenvalues, and in fact with less than 15
modes we reach a level of 10−3 , which is one order of magnitude less than the one we get with
less than 15 modes in the standard case. We get the same results for the extended displacement
df ; see Figure 14.10 (top). As we can see from Figure 14.9 (bottom) and 14.10 (bottom), with
the preprocessing procedure, the first modes are able to capture almost 70% more energy than
the modes without preprocessing. Figure 14.11 represents some snapshots for the fluid pressure
before and after the preprocessing, in order to help the reader to better visualize the effect that
this procedure has on the transport phenomenon.
Figure 14.9. Top: comparison between the rate of decay of the eigenvalues for the pressure with
and without preprocessing. Bottom: retained energy as a function of the number N of POD modes.
Figure 14.10. Top: comparison between the rate of decay of the eigenvalues for the displacement
with and without preprocessing. Bottom: retained energy as a function of the number N of POD modes.
14.5. A Monolithic RBM for a Transport-Dominated FSI Problem 307
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 14.11. Original snapshots for pf at times t = 0.001, t = 0.005, and t = 0.015 (left
column) and corresponding preprocessed snapshots (right column).
where F̃ti is the inverse map of the dilatation map introduced previously, with the important dif-
ference that now we are replacing the exact position of the peak γi = γ(ti ) with γ̃n , a prediction
of the position of the peak of the wave at time ti , obtained by simply evaluating q1 (ti ).
We remark that, in order to have a more stable approximation of the pressure at the reduced
order level, we enrich the reduced space associated with the fluid velocity with some supremizer
u
modes; therefore VN f = span{Φ1u , . . . , ΦN 1 Ns
u , Φs , . . . , Φs }.
u
We introduce the matrix Z p , which is a rectangular matrix whose entries are the FE degrees of
freedom of the reduced bases Φpi , returned by a POD on the preprocessed solution manifold, and
the matrix Z u,s , which is the matrix of FE degrees of freedom of the basis of the reduced space
u
VN f . Then the online system to be solved reads as follows: for every i = 0, . . . , Nt , find the
u p i Np df i N df
vector of coefficients αu (ti ) := (αku (ti ))N
k=1 , α (t ) := (αk (t ))k=1 , α (t ) := (αk (t ))k=1 ,
p i df i
ds λ u i N λu λd i N λd
αds (ti ) := (αkds (ti ))N
k=1 , α (t ) := (αk (t ))k=1 , and α (t ) := (αk (t ))k=1 such that
λu i λd i
d
1
(MRB,s ) = ZF̃u,T Ms1 ZF̃f ,
ti ti
2
(MRB,s ) = ZF̃ds ,T Ms2 ZF̃λu ,
ti ti
d ,T
C(·)RB = ZF̃f C(·)ZF̃u i ,
ti t
d ,T d
DRB = ZF̃f DZF̃f ,
ti ti
d ,T
(L1RB,d ) = ZF̃f L1d ZF̃λd ,
ti ti
d
and the other matrices ZF̃u , ZF̃f , ZF̃ds , ZF̃λu , ZF̃λd are defined similarly. Therefore in our
ti ti ti ti ti
algorithm, in order to obtain the reduced pressure at every time step, we use the basis matrix
Z p obtained by performing a POD on the pressure preprocessed solution manifold MFpf , and
we obtain a new matrix ZF̃p by a simple composition with F̃ti : ideally, the new matrix ZF̃p
ti ti
should represent a set of basis functions Φ̃pj (ti ) := Φpj ◦ F̃ti which is sufficient to express the
reduced pressure pN (ti ) in the original configuration, i.e., with the peak of the wave moving in
the domain.
It is important to underline that, following the definition of the matrix ZF̃p , it is clear that
ti
the set of basis functions used in the online step of the algorithm is time dependent: at every time
step of our simulation we compute a new set of basis functions Φ̃pj (ti ), according to the chosen
deformation map F̃ti .
Figure 14.12. Pressure snapshots (left column) at time t = 0.005 (top) and final time t = 0.015
(bottom). Reduced order pressure simulation (right column) at time t = 0.005 and time t = 0.015. The
reduced simulation was obtained with N = 4 basis functions for each component of the solution of the FSI
problem.
Figure 14.13. Displacement snapshots (left column) at time t = 0.005 (top) and final time
t = 0.015 (bottom). Reduced order displacement simulation (right column) at time t = 0.005 and time
t = 0.015. The reduced simulation was obtained with N = 4 basis functions for each component of the
solution of the FSI problem.
Figure 14.14. Analysis of the behavior of the relative error for the fluid pressure approximation.
Dashed lines were obtained by employing the preprocessing procedure, continuous lines were obtained
using a standard model order reduction.
by using just N = 4 modes for each component of the solution. Since in total we have six
unknowns for our FSI problem (remember the two Lagrange multipliers used to impose the
coupling conditions), we use Ntot = 24 basis functions. To obtain similar results, but without the
preprocessing procedure, we need to use at least 10 basis functions for each component, meaning
at least 60 modes in total. Thus, during the online phase we are decreasing the dimension of
the system to be solved by at least 40. These considerations are finally confirmed by an error
behavior analysis, as shown in Figures 14.14 and 14.15. In these figures, we can see that the
overall behavior of the relative error is improved by the preprocessing procedure. In Figure
14.14 the oscillating behavior of the relative error without preprocessing shows how the set of
basis functions struggles to reproduce the exact location of the peak of each component of the
solution. On the other hand, Figure 14.15 shows that the mean relative error is very low, with
just four basis functions for each component of the solutions.
310 Chapter 14. Reduced Order Methods for Fluid-Structure Interaction Problems
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 14.15. Behavior of the mean relative error for the fluid pressure, depending on the number
N of basis functions employed, with preprocessing (red) and without preprocessing (blue).
14.5.7 Conclusions
In this chapter, we presented two reduced order models that can be adopted when addressing
FSI problems. We first introduced the ALE formalism, adapting it to the case where a shape
parametrization is considered, and we set the formalism of multiphysics problems. We then
constructed two RBMs, both based on a POD to generate the reduced basis. The first method is
a partitioned procedure, where the coupling conditions are treated in a semi-implicit way with
respect to time, and where we added a Robin coupling condition to improve the stability of
the algorithm. At the reduced order level, we employed a change of variable and a harmonic
extension of the displacement basis functions in order to control the dimensions of the systems
to be solved online.
The second algorithm is a monolithic procedure that has been adapted in order to obtain good
performance in the case of transport-dominated problems. Here we added a preprocessing of the
snapshots during the offline phase, where we try to somehow get rid of the transport phenomenon.
At the online level, after having obtained the reduced basis functions with a standard POD, we
obtain the reduced solution as a linear combination of a pullback of these reduced bases. In this
case, a supremizer enrichment technique is used to obtain stable approximations of the pressure.
Both procedures perform very well and show the ability of the RBM to handle problems as
complex as FSI problems, even in the presence of shape parametrization.
Chapter 15
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
15.1 Introduction
This chapter will present the development and analysis of an efficient reduced order model
(ROM) for the study of bifurcating phenomena governed by nonlinear partial differential equa-
tions (PDEs) in fluid-structure interaction (FSI) problems. In particular, the main ingredients for
this investigation are those presented in Chapter 5 on the numerical framework for the analy-
sis of the bifurcation phenomenon, and Chapter 14 for the arbitrary Lagrangian Eulerian (ALE)
formulation to be used for FSI problems [418].
We focus on a particular bifurcating phenomenon known as the Coandă effect (see Section
18.3). We deal with this phenomenon in the context of a symmetry-breaking bifurcation for the
solutions of the steady Navier–Stokes equations in a planar contraction-expansion channel. We
have chosen to deal with this particular geometry because it represents a simplified setting aimed
at analyzing a heart disease known as mitral valve regurgitation [112]. The latter occurs when the
blood flows abnormally from the left ventricle to the left atrium, due to the defective closure of
the mitral valve. It may happen that this flow, in addition to flowing in an anomalous direction,
exhibits a wall-hugging behavior, i.e., an asymmetric jet that attaches to the left atrium wall
[457]. This behavior is an example of the aforementioned Coandă effect. Moreover, from the
medical point of view, the problem is that the wall-hugging jet leads to incorrect measurements
of the blood volume through echocardiography.
Several studies have been conducted on the phenomenon. In particular, the literature is ex-
tensive in experimental [122] and numerical [543] analyses which aim at identifying the main
characteristics of the flow. The most recent investigations use reduced models [467, 457, 446,
448, 452] to alleviate the computational complexity of this type of study.
The results of these analyses show that the Coandă effect occurs even for relatively low
Reynolds numbers in the range [40, 160]. This can be partially justified by the fact that although
the Coandă effect occurs more intensely in a turbulent regime, it is amplified in two-dimensional
flows [586]. Specifically, it has been shown that, working with a fixed orifice width and decreas-
ing the kinematic viscosity, the inertial effects become progressively more important, breaking
the symmetry of the solution for a certain critical value. Therefore, below this value, one phys-
ically unstable configuration of symmetric flow and two stable asymmetric ones can be found.
311
312 Chapter 15. ROMs for Bifurcating Phenomena in Fluid-Structure Interaction Problems
The peculiarity is that these solutions coexist for the same parameter values, as is typical of
bifurcating phenomena.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
We point out that all previous works have addressed the problem in a purely fluid dynamics
context. For this reason, in this chapter, we want to introduce a multiphysics model that allows
us to study the coupling between the fluid and an elastic solid [420] and the effect that this can
have on the bifurcation. To do this, we analyze both linear and nonlinear constitutive relations
for the structure.
Furthermore, since reconstructing the bifurcation diagrams related to a nonlinear PDE re-
quires considerable computational effort, we mitigated this cost by using a reduced order mod-
eling (ROM) technique. To do so we employed the reduced branchwise algorithm presented
in Chapter 5, which is based on a monolithic proper orthogonal decomposition (POD) coupled
with Galerkin finite elements (FEs). This approach optimizes the computational resources and
provides results that can be applied in a multiparameter context. For a more exhaustive de-
scription of the techniques and results obtained for the present investigation, we recommend
consulting [337].
where ρs (respectively ρf ) is the density of the solid (respectively fluid), while b̂s (respectively
bf ) represents the external volume forces for the solid (respectively fluid).
In system (15.1) the symbol ∧ is used to denote operators and fields taken with respect to the
reference variable x̂.
The Cauchy stress tensor σf and the second Piola–Kirchoff stress tensor P are prescribed by
means of the following constitutive relations:
T
σf (u f , pf ) = ρ f µ ∇uf + ∇u f − pI,
ˆ (15.2)
P(ds ) = λs tr
Es (ds )I + 2µs Es (ds ),
Es dˆs = 1 ∇ ˆ dˆs + ∇
ˆ T dˆs ,
2
15.2. Problem Formulation 313
where λs and µs are the Lamé parameters. The kinematic viscosity of the fluid µ has a prominent
role in the present chapter since it is the bifurcation parameter under investigation.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In order to solve the former problem, we need to pull back the Navier–Stokes equations on
the reference configuration Ω̂f . To do so, we first introduce the gradient of the ALE map and its
determinant, namely,
F := ∇ ˆ x̂ + dˆf and J := det F, (15.3)
and then we use the former quantities to give a reference representation of σf , i.e.,
σ̂f = µ ∇ûˆ f F−1 + F−T ∇ ˆ T ûf − p̂I. (15.4)
We exploit the Piola transformation rule, which leads to the following system:
ρf J +∇ûˆ f F−1 (ûf ) − divˆ (J σ̂f (ûf , pf )F−T ) = Jbf in Ω̂f ,
ˆ (JF−1 uf ) = 0 in Ω̂f ,
div
(15.5)
ˆ
∆df = 0 in Ω̂f ,
−divˆ P(dˆs ) = bs in Ω̂s .
• Boundary conditions:
where n̂ the normal outgoing vector from the fluid resolution domain. See Figure 15.1 for
a sketch of the different portions of the boundaries.
Vf = H 1 (Ω̂f ; Rn ),
Qf = L2 (Ω̂f ),
(15.8)
E = H 1 (Ω̂f ; Rn ),
f
Vs = H 1 (Ω̂s ; Rn ).
We multiply each equation of system (15.5) by an appropriate test function belonging to the
corresponding functional spaces (15.8) and apply the divergence theorem. This leads to the
following equations:
314 Chapter 15. ROMs for Bifurcating Phenomena in Fluid-Structure Interaction Problems
Γ̂D
s
Γ̂0
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Γ̂F SI
Γ̂F SI
Ω̂s
Γ̂D
s Γ̂0
Jρf [∇uf F−1 ]uf · vf dx̂ + Ω̂f Jσf F−T : ∇vf dx̂
R R
Ω̂R
f
(15.9)
− ∂ Ω̂f Jσf F−T n · vf dAx̂ − Ω̂f Jbf · vf dx̂ = 0 for every vf ∈ Vf .
R
• Extension equation:
Z
∇df : ∇ef dx̂ = 0 for every ef ∈ Ef . (15.11)
Ω̂f
A first simplification comes from summing up the two equations (15.9) and (15.12). In fact, by
doing so one can directly exploit the dynamic coupling condition in the following way:
Therefore, by using the former result we can join (15.9) and (15.12) in a single fluid-structure
momentum equation as follow:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Jρf [∇uf F−1 ]uf · vf dx̂ + Ω̂f Jσf F−T : ∇vf dx̂
R R
Ω̂R
f
Λu = H 1 (Γ̂F SI ; Rn ),
(15.15)
Λd = H 1 (Γ̂F SI ; Rn ).
basis method (RBM). To do so, we will follow the approach of [389]. The first critical observa-
tion is that we will only present a basic strategy of projecting the nonlinear governing equations
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
onto the reduced basis. This, in principle, does not guarantee substantial computational savings
because the complexity of the procedure still depends directly on N (the cardinality of the high-
fidelity model). An alternative to the presented methodology is hyperreduction strategies such
as the empirical interpolation method (EIM) [40, 245], which can be used to achieve an online
computation cost independent of N . However, such strategies require further complexity in the
offline phase and seem to lack robustness when applied to bifurcating problems [446, 450]. Let
us denote by XN the span of the reduced basis obtained via POD (see Chapter 2), where the car-
dinality of this basis is N . The reduced problem therefore reads as follows: find XN (µ) ∈ XN
such that
f (XN (µ), YN ; µ) = 0 ∀ YN ∈ XN . (15.19)
This reduces, as for the full-order model (FOM), to a system of N nonlinear equations, which
can be solved by means of the Newton method, namely, gien an initial guess XN 0
∈ XN for
k = 0, 1, . . . , do the following:
1. Solve the linear variational problem:
k k
(15.20)
df XN (µ) (δXN , YN ; µ) = −f XN (µ); YN ; µ .
From the discussion in Chapter 2 we know that problem (15.20) can be cast in matrix form as
One can therefore prove that the linear system (15.22) is equivalent to
At this point, we recall that our goal, as discussed in Chapter 5, is to be able to follow a single
branch while changing the value of the parameter we are dealing with. This task is not trivial
since, after the bifurcation has occurred, we lose the uniqueness of the solution, and thus the
convergence of the Newton method is strongly tied to the initial hypothesis. Consequently, we
need a strategy to follow the solution as the parameter varies. For this reason, we will implement
a continuation method, i.e., a procedure to generate a succession of solutions associated with
the same branch [170]. We refer the reader to Subsection 5.3.1 for details on the algorithm
for full-order branchwise reconstruction. In order to use the RBM, it is therefore necessary to
use Algorithm 5.1 by operating on a discrete version of the parameter space, i.e., Ptrain =
{µ1 , µ2 , . . . , µNtrain } ⊂ P.
15.3. Application to the Coandă Effect 317
Thus, having the set of solutions {X(µj ) | µ ∈ Ptrain }, one can extract the reduced space
XN using the POD. At this point, the branch is reconstructed using the reduced order counterpart
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
of Algorithm 5.1, which is almost identical to its FOM version. The only differences lie in
substituting the FOM quantities by their reduced order equivalent and taking account of the
previous considerations for the Jacobian matrix. The result is presented in Algorithm 15.1.
6: Xk+1 k
N,j = XN,j + δXN
7: end while
8: end for
15.3.1 Preliminaries
Let us start by noting that we have used the centimeter-gram-second unit system for all the
equations. In the following, we omit the units of measurement of the quantities we will be
working with to simplify the notation. The dimensions of the computational domain are reported
in Figure 15.2. In particular, the leaflets (in magenta) represent the solid resolution domain Ω̂s ,
whereas the blue interstice is the fluid resolution domain Ω̂f .
For the particular case under consideration, we imposed the following boundary conditions:
5. 1. 44.
2.5
2.5 7.5
2.5
50.
where
Γ̂in = 0 × [0, 7.5],
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(15.26)
Γ̂out = 50 × [0, 7.5].
The previous conditions are equivalent to enforcing a stress gradient between the outlet and the
inlet of the channel, which therefore animates the fluid flow, with direction from Γ̂in to Γ̂out .
The FOM problem was solved by employing the FEM. In particular, the meshes we used for the
spatial discretization are unstructured and composed of triangular elements. The discretization
performed is second order for uf , df , and ds , while it is first order for p, λd , and λu . This
choice was made to avoid known stability issues typical of saddle-point problems [65, 74].
The software employed to solve the FOM problem is multiphenics [6], which has been cou-
pled with RBniCS [38] for the reduced order phase.
In particular, the previous constitutive relation is completely characterized by the Lamé con-
stants, which in our case assume the following values:
For the bifurcation analysis, we analyzed the behavior of the solution for the kinematic viscosity
interval P = [0.5, 2.0], discretized by using 51 points. In the following, we will denote this
discretization by Ph .
The first phase was conducted by computing the set of FOM solutions for each value of
the kinematic viscosity µj in Ph . In this regard, Figure 15.3 shows the velocity and pressure
fields obtained for some of the parametric instances considered during this step. Figure 15.4
reports the displacement field of the solid df for certain values of the kinematic viscosity. Each
of these fields is associated with the fictitious displacement of the fluid mesh df , which is its
extension, as prescribed by the harmonic problem (see Figure 15.5). At this point we want to plot
the bifurcation diagram for the problem under consideration (see Chapter 5). In particular, the
significant scalar output we choose is the vertical velocity of the fluid at a point of the channel
symmetry axis upstream of the expansion, i.e., uy (x), with x = (14, 3.75). The diagram in
question is shown in Figure 15.6. The result obtained is consistent with the analysis presented
in Subsection 5.7.3, in which the problem was solved in the absence of the elastic structure, and
therefore only for the fluid phase. In that case, a pitchfork bifurcation was identified, which is
also present for the FSI problem. In particular, in our case we focused on obtaining the lower
stable branch, which was obtained by recursively decreasing the value of the kinematic viscosity.
The next step is to obtain a bifurcation diagram for the solid phase. Since the particular
bifurcation faced is caused by the nonlinearity of the Navier–Stokes system, it is interesting to
understand how the structure undergoes it. In this case, we used the scalar indicator
∆d = | max dˆs − max dˆs |, (15.28)
x̂∈Ω̂1 x̂∈Ω̂2
15.3. Application to the Coandă Effect 319
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
f_234 Magnitude
0.0e+00 5 10 15 20 25 2.9e+01
Figure 15.4. Solid’s deformation snapshots: the solutions are given in a spatial frame using the df map.
0.0
0.014
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
− 0.5
0.012
− 1.0
0.010
u y (14, 3.5)
− 1.5
0.008
Δd
− 2.0
0.006
− 2.5
0.004
− 3.0
0.002
− 3.5
0.000
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
μ μ
Figure 15.6. Bifurcation diagram for the Figure 15.7. Bifurcation diagram for the
fluid phase. solid phase.
where Ω̂1 is the upper leaflet and Ω̂2 is the lower one. This quantity is significant in plotting
the bifurcation diagram presented in Figure 15.7 because it represents an indirect measure of the
imbalance of the forces acting on the two leaflets, following the onset of the asymmetric flow
downstream of the expansion.
σi
σ̂i = . (15.29)
σ1
Furthermore, if we take into account the cumulative sum of the eigenvalues, we get another
important indicator known as the retained energy:
Pn
σi
En = PNi=1
train
, (15.30)
i=1 σi
with Ntrain = 51 being the number of snapshots computed during the offline phase. The trend
of both these indicators is presented in Figure 15.8.
We note an exponential decay of the eigenvalues as a function of the cardinality of the basis.
In particular, the two displacement fields decay faster than the pressure and the velocity of the
fluid. This is also confirmed by the trend of retained energy, whose values for the first mode are
presented in Table 15.1.
uf p df ds
En 92.9% 90.1% 99.9 % 99.9%
15.3. Application to the Coandă Effect 321
10
0
100 p
uf
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
df
9.8 × 10
−1
ds
10−2
−1
9.6 × 10
10−4
En
σn̂
−1
9.4 × 10
10−6
9.2 × 10
−1
p
uf
df 10−8
9 × 10
−1 ds
10
0
10
1
0 10 20 30 40 50
n n
Figure 15.8. Retained energy. Figure 15.9. Decay of the normalized eigenvalues.
Thus, a single mode would be sufficient to retain 99.9% of the energy (variance) of the dis-
placement’s snapshots. In fact, we recall that the problem associated with ds is that of linear
elasticity, while that for df is a harmonic extension, which is also linear. Consequently, the
problems’ linearity justifies the lower complexity of their solution manifold.
We decided to investigate the trend of the mean error by varying the number of bases used
by the RBM. The reconstruction efficiency was measured by the metric
NX
train
1 kXh (µj ) − Xrb (µj )kX
Errrb = . (15.31)
Ntrain j=1
kXh (µj )kX
In Figure 15.10 we observe that, as expected, the errors are higher than their direct projection
counterparts presented in Figure 15.11. Moreover, we can observe that the decay, although not
monotone, maintains an average exponential trend. To refine the reconstruction of the bifurcation
diagram, we decided to use 101 equispaced points within the interval P. These kinematic vis-
cosity values were employed to reconstruct the lower stable branch using Algorithm 15.1. Each
of the reduced solutions was subsequently compared to its FOM counterpart by calculating the
relative reconstruction error (Figure 15.12). In particular, despite the fluctuations due to the local
effect of approximation improvement by each of the basis functions, we observe that the average
trend is in line with what was previously observed in Figure 15.10 for the mean errors. In this
case, the reconstructed pressure worse than the other fields. Finally, we note the increase of the
error both for low-viscosity regimes, around µ = 0.5, and at the bifurcation point.
10−1 p 10−1 p
df df
ds 10−2 ds
uf uf
10−2 10−3
10−4
Errproj
10−3
Errrb
10−5
10−6
10−4
10−7
10−5 10−8
2 4 6 8 10 12 14 16 2 4 6 8 10 12
n n
Figure 15.10. Average reconstruction er- Figure 15.11. Average projection error
ror in logarithmic scale. in logarithmic scale.
322 Chapter 15. ROMs for Bifurcating Phenomena in Fluid-Structure Interaction Problems
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
• The rigid structure case represents the incompressible fluid dynamics problem already
presented in Chapter 5. In this case, the Navier–Stokes equations are solved, removing the
structure’s description from the computational domain (see Figure 15.2).
• The Saint Venant–Kirchoff (SVK) model is a nonlinear hyperelastic model for which the
constitutive relation for the Piola tensor is
For the FSI problem, addressed with both the linear and SVK models, we decided to use as
independent parameters, within the constitutive relation, Young’s modulus (E) and Poisson’s
ratio (ν), which can be used to express the Lamé constants as
E Eν
µs = , λs = . (15.33)
2(1 + ν) (1 + ν)(1 − 2ν)
In particular, the results of the previous section for the linear model refer to the test case where
E = 2.89e5 and ν = 0.44. In that case, introducing the structure does not seem to alter the
bifurcation phenomenon, since the high stiffness results in small deformations. For this reason,
we decided to work in a regime of large bifurcations by reducing the structure stiffness under
consideration for both the linear and the SVK models, as reported in Table 15.2. In fact, both
the linear (A) and the SVK cases lead to deformations of the same order of magnitude as the
resolution domain.
15.4. Model Comparison 323
Table 15.2. Values of Young’s modulus and Poisson’s ratio used for the test cases.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Case E ν
linear (A) 4.67e4 0.21
linear (B) 2.89e5 0.44
SVK 4.67e4 0.21
0 SVK
linear (A)
linear (B)
rigid body
−1
uy(14, 3.75)
−2
−3
−4
Figure 15.15. Comparison of the bifurcation diagram for the fluid phase for the different test cases.
A first comparison, presented in Figure 15.13, concerns the maximum displacement of the
two leaflets for the linear (A) and SVK model. We notice that although the SVK model generates
solid deformations of higher magnitude, it reduces the asymmetry of the behavior between the
two leaflets in the postbifurcation regime, as evidenced by the trend of ∆d in Figure 15.14. As a
result, the response of the linear model in front of the bifurcation is accentuated compared to its
nonlinear counterpart, although the deformation field is of lower intensity. To understand how
this difference in behavior is reflected in the fluid phase, we finally examined the bifurcation
diagram relative to the already employed quantity uy ((14, 3, 75)). Figure 15.15 shows how the
324 Chapter 15. ROMs for Bifurcating Phenomena in Fluid-Structure Interaction Problems
different constitutive relation for the solid has an important influence on the bifurcation point
for the fluid. In particular, we have two linear models (A and B) that relate to cases of more
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
or less rigid structure. The increase in structural stiffness in the linear case leads to a delay
in bifurcation, and in fact the linear model (B) seems to not modify the bifurcation point with
respect to the rigid structure case. The SVK model, on the other hand, delays the bifurcation with
respect to the rigid case but anticipates it when compared to its linear (A) counterpart. Finally, it
can be seen that the SVK and the linear models (A) differ most at the bifurcation point, showing
similar behavior once the bifurcation has occurred.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Part IV
and Applications
Further Advances, Perspectives,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 16
Reduction in Parameter
Space
16.1 Introduction
Parameter space reduction has been established in recent years as a valuable tool to fight the curse
of dimensionality, which means that the complexity of the algorithms grows exponentially with
the dimension of the input parameter space. Reduced order methods [281, 517, 508, 503] have
been developed to deal with complex parametric problems, but the limit for high-dimensional
parameter spaces still remains. Nowadays the problem is even more emphasized thanks to the
spread of high-performance computing facilities which enable the study of highly parametrized
systems. On the other hand, if we consider modern industrial optimization we also find the
presence of many input parameters in order to analyze a wide range of possible designs.
The active subspace (AS) [134, 136] method has emerged as a reliable and explainable tech-
nique for linear dimensionality reduction of the input parameter space. This success is also
testified to by the range of successful applications, including shape optimization [383, 161, 157],
turbomachinery [534], chemistry [311, 310, 597], automotive engineering [432, 494], and naval
and nautical engineering [574, 571, 404, 155, 572, 575].
The methodology has been innovated through the development of several extensions, e.g.,
ASs for vector-valued functions [635], multifidelity approximations [356], localized models
[495], and kernel-based ASs [493]. Moreover, other nonlinear techniques are emerging, such
as nonlinear level-set learning [643] and the active manifold method [77]. We will present all
these methods in the following sections, highlighting their characteristics and providing a gen-
eral picture for parameter space reduction. Many of the techniques described in this chapter are
implemented in the open-source Python package8 called ATHENA [496]; the reader can also
refer to Chapter 19.
This chapter is organized as follows: in Section 16.2 we introduce a general framework for
the ridge approximation problem and we present several techniques based on the AS method,
such as kernel-based active subspaces (KASs), local active subspaces (LASs), some gradient-
free extensions, and a multifidelity setting; in Section 16.3 we briefly present other nonlinear
approaches to parameter space dimensionality reduction; Section 16.4 is devoted to showing a
8 Available at https://github.com/mathLab/ATHENA/.
327
328 Chapter 16. Reduction in Parameter Space
wide range of possible applications of the described techniques, with some comparisons. Finally,
in Section 16.5 we draw some conclusions and describe some future perspectives for parameter
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
space reduction.
where a ∈ Rn \ {0} is called the direction and f is a univariate real-valued function, i.e.,
f : R → R. We note that F is constant on the hyperplanes a · x = c, where c ∈ R. This
concept can be generalized by considering more than one direction. We call the generalized
ridge function any real-valued multivariate function F : Rn → R such that
where A ∈ Rn×d is a tall matrix, with 1 ≤ d < n, and f : Rd → R. The goal of ridge recovery
is to estimate the matrix A using only point queries of f .
The methods presented below seek to approximate a target function of interest with a gener-
alized ridge function.
h iT
∂f ∂f
where Eρ denotes the expected value, ∇x f = ∇f (x) = ∂x 1
, . . . , ∂xn is the column vector of
partial derivatives of f , and dL is the Lebesgue measure. Since Σ is symmetric positive definite,
it has an eigendecomposition with real eigenvalues Σ = WΛWT .
We can retain only a small number r n of eigenpairs: r becomes the AS dimension, and
the span of the corresponding eigenvectors defines the AS. The partition is
Λ1 0
Λ= , W = [W1 W2 ] , (16.4)
0 Λ2
The AS dimension r can be selected a priori, or by looking at the presence of a spectral gap
[134], or by imposing a cumulative energy threshold for the eigenvalues.
16.2. AS-Based Methods 329
We remark that the projection map onto the AS is not an injective map because W1T is defined
as a linear projection onto a subspace. So the back-mapping from the AS onto X is not trivial.
Let x∗M be a point in the AS; we can find K ∈ N+ points in the full parameter space which are
mapped onto x∗M by W1T . Following from the decomposition introduced above we have
with the additional constraint coming from the rescaling of the input parameters needed to apply
the AS: −1 ≤ x ≤ 1. By 1 we denote the vector in Rn with all elements equal to one. To
sample the inactive variable y we need to satisfy
−1 ≤ W1 x∗M + W2 y ≤ 1 (16.7)
or, equivalently,
W2 y ≤ 1 − W1 x∗M − W2 y ≤ 1 + W1 x∗M . (16.8)
These inequalities define a polytope in R from which we want to uniformly sample K points.
n−r
The inactive variables are in general sampled from the conditional distribution p(y|x∗M ); here
we show how to perform it for the uniform distribution. For more general distributions one could
use Hamiltonian Monte Carlo. Here we start with a rejection sampling scheme, which finds a
bounding hyperbox for the polytope, draws points uniformly from it, and rejects points outside
the polytope. If this method does not return enough samples, we use a hit and run method
[542, 50, 381] to sample from the polytope. This method starts from the center of the largest
hypersphere within the polytope and then selects a random direction and identifies the longest
segment lying inside the polytope. The new sample is randomly drawn along this segment. This
procedure continues by iterating the same steps starting from the last sample until K samples are
found. If that does not work, we select K copies of the Chebyshev center [70] of the polytope.
In Figure 16.1 we depict the sampling strategy at different stages.
New samples
Figure 16.1. Illustration of the inactive variable sampling strategy. Successive steps are depicted
all at once. We highlight the Chebyshev center, the selection of the next sample using the hit and run method,
and the polytope defined by (16.8).
Sometimes the function of interest does not directly admit an accurate ridge approxima-
tion but can nevertheless be approximated by it after a nonlinear transformation. Even if, in-
stead of (16.2), the function of interest is factorized through an optimal profile [134], f (x) ≈
Eρ [f |W1 W1T ](W1 W1T x), i.e., the random variable f (X) conditioned on W1 W1T X, this may
still not be enough to provide an accurate approximation.
A fairly simple example is radial symmetric functions, with generatrix g : R+ → R, of the
form
f (x) = g(kxk) ∀x ∈ X ⊂ Rn (16.9)
in a radial domain X with respect to a fixed origin. For each chosen active direction, the optimal
profile is the average of the radial symmetric function over the orthogonal inactive direction
p(xM , y)
Z
Eρ [f |W1T ](xM ) = f (x) dy,
p(y)
but a more natural factorization would associate with each input its norm (as an encoded feature)
and then choose the generatrix as profile.
In order to overcome these problematic situations, nonlinear dimension reduction is em-
ployed. The best way to achieve nonlinear dimension reduction with a one-dimensional reduced
parameter space could be realized by discovering a curve in the space of parameters that cuts the
level sets of f transversely; this variation is presented in Subsection 16.3.2 as an active manifold.
Based on similar reasoning, nonlinear level-set learning (NLL) [643] is introduced in Subsec-
tion 16.3.1.
Feature space
mapping
Active
subspace
Regression
Projection
Active variable
Figure 16.2. Illustration of the design of a response surface using KAS and Gaussian process
regression (GPR). The factorization through an RKHS is emphasized.
{wi · φ}i∈{1,...,d} . In comparison, in the AS method we use the following linear encoding of the
random variable X:
W1T X = w1 (w1 · X) + · · · + wd (wd · X),
Since for AS only the samples of the Jacobian matrix of the model function are employed, we
can ignore the definition of the new map f˜ : φ(X ) ⊂ H → R and focus only on the computation
of the Jacobian matrix of f˜ with respect to the new input (latent) variable y := φ(X). The
correlation matrix becomes
Z Z
H̃ = ˜ T ˜
(Dy f ) (y)(Dy f )(y) dµ(y) = (Dy f˜)T (Dy f˜) dLX (x),
φ(X ) X φ(x) φ(x)
M
1 X
Z
H̃ = (Dy f˜)T (Dy f˜) dLX (x) ≈ (Dy f˜)T (Dy f˜) .
X φ(x) φ(x) M i=1 φ(Xi ) φ(Xi )
332 Chapter 16. Reduction in Parameter Space
The gradients of f˜ with respect to the new input random variable y are computed from the
known values DX f with the chain rule: if the Hilbert space H has finite dimension H ∼ RD ,
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
this procedure leaves us with an underdetermined linear system to solve for Dy f˜:
with the usual notation for the singular value decomposition (SVD) of Dφ(X),
Dφ(X) = U ΣV T ,
and Σ† ∈ M(r × r) equal to the diagonal matrix with the inverse of the singular values as
diagonal elements.
The choice for the map φ is linked to the theory of RKHS [56] and in particular is linked to the
definition of random Fourier features [479]. It is defined as
r
2
y = φ(X) = σf cos(W X + b),
D
1
cos(W X + b) := √ (cos(W [1, :] · X + b1 ), . . . , cos(W [D, :] · X + bD ))T , (16.10)
D
D
r !
∂z j 2 X
=− σf sin Wik xk + bk Wij (16.11)
∂xi D
k=1
cross-validation. For a more in depth description of the hyperparameter tuning procedure, the
reader can refer to [493].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
where Γ is the diagonal matrix of the eigenvalues of the correlation matrix of the gradients of the
function of interest and xi and xj are a couple of input points.
In Figure 16.3 we consider a quartic function defined as f (x) = x41 − x42 in the domain
[0, 1]2 , where we highlight the direction of the global AS, which induces the cluster division in
Figure 16.4, depending on the chosen technique. As we can see, K-means produces uninformed
subsets, since it uses the standard Euclidean metric to find the centroids. By exploiting the
distance in (16.12), instead, the clusters are more aligned with the inactive space direction.
0.4 0.25
0.3
0.50
0.2 0.6
0.75
0.9
0.0 1.00
0.0 0.2 0.4 0.6 0.8 1.0 1.0 0.5 0.0 0.5 1.0
x1 Active variable W1Tx
Figure 16.3. Illustration of the global AS direction, highlighted in orange in the left panel. On
the right is the corresponding sufficient summary plot.
In general, applying the AS response surface design to many smaller subdomains is not
guaranteed to achieve better accuracy than employing the global AS. This is due mainly to
two reasons. When data is scarce, since the response surface design with AS is based on a
Monte Carlo approximation, the local surrogate models will be even more penalized. Secondly,
if the function of interest has a relatively high total variation, it is not guaranteed that the lo-
cal AS is a better minimizer of the local ridge approximation error with respect to the global
AS [495].
334 Chapter 16. Reduction in Parameter Space
1.0
0.8
0.6
x2
0.4
0.2
0.0
0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00
x1 x1 x1
Figure 16.4. Comparison between K-means, K-medoids, and hierarchical top-down clustering
using the AS-induced distance metric for the quartic example of Figure 16.3.
N
Sq := {xiq , yqi }i=1
q
⊂ X ×R ⊂ Rd ×R, π1 (Sp ) ⊂ π1 (Sp−1 ) ⊂ · · · ⊂ π1 (S1 ) ⊂ X , (16.13)
N
where {xiq , yqi }i=1
q
represent the input-output pairs of the qth fidelity model and π1 : X × R →
X is the projection onto the d-dimensional input components. The final multifidelity model
integrates recursively all the previous ones. Since all q-fidelity models fq that we consider are
Gaussian processes, they are completely defined by the mean field (which is supposed to be the
0 constant field as before) and by their kernel
where is a noise term, we employed the notation x̄ := (x, fq−1 (x)) ∈ Rd × R for q > 1 and
16.2. AS-Based Methods 335
kq ((x, fq−1 (x)), (x0 , fq−1 (x0 )); θq ) = kqρ (x, x0 ; θqρ ) · kqf (fq−1 (x), fq−1 (x0 ); θqf )
+ kqδ (x, x0 ; θqδ ) ,
in which the hyperparameters to be tuned are represented by θq = (θqρ , θqf , θqδ ). So, apart from
the lowest fidelity, all the GP models have the input domain X ⊂ R augmented by the predic-
tions of the previous fidelity. The hyperparameters θq are optimized recursively with maximum-
likelihood estimation for each GP model GP(fq |0, k q (θq )):
1 1
arg max log p(yq |xq , θq ) ∝ − log |k q (θq )| − yqT (k q (θq ))−1 yq ; (16.15)
θq 2 2
this is why a hierarchical dataset is needed. After tuning, the predictive mean and variance for a
new input x∗ at fidelity q are integrated recursively with Monte Carlo over the distribution of the
previous fidelity Lfq−1 (x∗ ) :
Z
p(fq (x∗ )|x∗ , xq , yq ) = p(fq (x∗ )|f∗q−1 , x∗ , xq , yq )p(f∗q−1 )dz
ZX
= p(fq (x∗ )|s, x∗ , xq , yq )dLfq−1 (x∗ ) (s)
fq−1 (X )
M
X
≈ p(fq (x∗ )|si , x∗ , xq , yq ).
i=1
We now introduce how ASs are involved in the design of a two-fidelity multifidelity model.
The incorporation of prior knowledge, biasing the learning process, is inevitable for the success
of learning algorithms. In our methodology this is realized in two ways: with the kernel of the
Gaussian process (GP; type of spatial correlation, degree of smoothness of the model) and with
the low-fidelity intrinsic dimensionality assumption (presence of a dominant AS).
In our paradigm, the low-fidelity model is represented by the response surface built with the
AS (computed from the high-fidelity data) on the whole domain X , and the high-fidelity model
is the GPR of the input-output data available (usually in small quantity). Due to the low intrinsic
dimensionality assumption, we expect the multifidelity model to perform better than the high-
fidelity one, thanks to the nonlinear integration of the data with the information of the presence
of an AS.
We provide a simple example. The function of interest is a hyperbolic paraboloid
The high-fidelity inputs are sampled uniformly: we sample and evaluate 50 high-fidelity and
200 low-fidelity input-output pairs. As previously said, the AS used to build the low-fidelity
model is obtained after computing the analytic gradients for the 50 high-fidelity input-output
pairs (generally the gradients could also be approximated directly from the input-output pairs).
The low-fidelity samples are cheap, and the cost of their evaluation is not comparable with the
cost of a high-fidelity output evaluation. The low-fidelity response surface is represented in
Figure 16.5.
Usually the high-fidelity model is not accurate enough due to the relatively scarce data avail-
able, while the multifidelity model, by exploiting the inductive bias of an intrinsic low dimen-
sionality, performs better. The outputs predicted on the 1000 test inputs for the three models
(low-fidelity, high-fidelity, multifidelity) are qualitatively compared in Figure 16.5. In the high-
fidelity model, the actual input-output pairs used for training are depicted in white.
336 Chapter 16. Reduction in Parameter Space
Mean
Data
1.0
Confidence
0.5
f (x)
0.0
0.5
High-fidelity data
1.0 0.5 0.0 0.5
Active variable W1Tx
Figure 16.5. Illustrative scheme of a multifidelity regression procedure that employs AS as a low-
fidelity model. The function to be approximated is the hyperbolic paraboloid from (16.16). Starting from
10 high-fidelity data points (depicted in blue and in white) we construct as a low-fidelity model a response
surface which is constant along the inactive subspace.
The l2 error of the multifidelity model against the 1000 test samples is 0.00101, while the l2
error for the high-fidelity model is 0.00125. Even if in this test case (employed only to clarify
the procedure) the gain is only marginal, in practice there may be noticeable improvements,
especially when the dimension of the parameter space is higher and the data is still scarce for the
complexity of the problem at hand [494].
where X is the training data matrix. It is clear how it simultaneously learns the link function and
the AS projection matrix. With a similar approach, in [215] they use a fully Bayesian approach
in order to obtain full posterior probability distributions for both the hyperparameters θ and the
projection matrix W in (16.17). Thus, for a given new unseen input, the predictive uncertainty
obtained from the queried surrogate model accounts for all the uncertain model parameters. The
authors also impose that the projection matrix is an element of the Stiefel manifold.
Another approach is presented in [626], where the authors propose a closed-form expression
for the uncentered covariance matrix defining the AS for a GP for the popular covariance kernels,
a technique usually executed via Monte Carlo finite differencing. This allows one to create
acquisition functions enabling evaluation-efficient estimation of such subspaces via sequential
16.3. Other Nonlinear Techniques 337
design, with fewer black-box evaluations. A gradient-free approach exploiting Morris screening
elementary effects can be found in [372].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
where u and v are partitions of the states, h is a scalar time step, the matrices K contain the
weights, b represent the biases, and σ is the activation function. We remark that the original
coordinates and the transformed ones are split in two in u and v. In Figure 16.6 we depict a
two-dimensional example where we can actually plot the transformed input parameter space.
We implemented this technique in the ATHENA [496] Python package.
Figure 16.6. NLL application two-dimensional example. On the left we have the target func-
tion against the first transformed coordinate, in the middle the loss function decay, and on the right the
transformed parameter space.
all the values f can attain. This allows for more accurate regressions over the low-dimensional
manifold with respect to the AS method. The drawback is the computational overhead. An
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
16.4 Applications
In this section we review a wide range of possible applications of the techniques presented above.
We provide a comparison in the construction of response surfaces, we introduce the activity
scores for sensitivity analysis, and we show how to integrate ASs into optimization strategies.
Finally, we present some insights into the treatment of inverse problems and coupling with other
reduced order methods.
N
1 X
≈ f (W1 W1T X + (Id − W1 W1T )Yi ) =: ĝ(W1 W1T X),
N i=1
16.4.2 Regression
We show some examples of response surface design with GPR [622]: we compare four response
surfaces built by exploiting an AS, a KAS, an LAS, and the NLL method. The chosen test case
models the spread of the Ebola virus:9 the output of interest is the basic reproduction number
R0 of the susceptible-exposed-infectious-removed (SEIR) model, described in Table 16.1, which
reads
β1 + β2 ρω1 γ1 + βγ23 ψ
R0 = , (16.19)
γ1 + ψ
with parameters distributed uniformly in Ω ⊂ R8 . The parameter space Ω is defined by the lower
and upper bounds summarized in Table 16.1.
We collected 200 input-output gradient triples as the training set. They are employed to find
the AS, to tune the feature map for KAS, to cluster the parameter space with hierarchical top-
down clustering for LAS, and to train the recursive neural network (RNN) for NLL. The spectral
Table 16.1. Parameter ranges for the Ebola model. Data taken from [167].
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
β1 β2 β3 ρ1 γ1 γ2 ω ψ
Lower bound 0.1 0.1 0.05 0.41 0.0276 0.081 0.25 0.0833
Upper bound 0.4 0.4 0.2 1 0.1702 0.21 0.5 0.7
measure chosen for KAS is a Beta distribution whose two shape parameters are tuned with log-
grid search and cross-validation [493]. The clustering algorithm employed for LAS is hierarchi-
cal top-down clustering, which applies K-medoids with the AS-informed metrics (16.12) at each
level of the clustering refinements. The RNN of the NLL method has seven layers and is trained
for 1000 epochs with the Adam stochastic optimization method [344].
Figure 16.7 shows the response surfaces for the AS, KAS, and NLL methods. The local
response surfaces for each cluster of LAS are reported in Figure 16.8.
3.0
Active Subspaces Kernel-based Active Subspaces Non-linear Level-set Learning
2.5 2.5
2.5
2.0 2.0 2.0
mean
1.5
R0
Figure 16.7. On the y-axis: basic reproduction number R0 of the SEIR model for the Ebola
epidemic. Left: GPR on the one-dimensional AS. Middle: GPR on the one-dimensional KAS. Right: GPR
on the one-dimensional active variable of the NLL method. In the plots are also represented the 200 training
points used to find the reduced parameter space and build the GPRs.
1.25
1.5 0.8 1.00
0.8 0.6 0.4 0.2 0.0 0.2 0.0 0.2 0.4 0.6 0.75 0.50 0.25 0.00 0.25 mean
LAS cluster 3 LAS cluster 4 LAS cluster 5 68% confidence
2.0 train data
1.50 3.0
1.8
1.25
1.6 2.5
1.00
R0
1.4
2.0
0.75 1.2
0.50 1.0 1.5
0.0 0.5 1.0 0.6 0.4 0.2 0.0 0.2 1.0 0.8 0.6 0.4 0.2
reduced parameter space reduced parameter space reduced parameter space
Figure 16.8. Each one of the six plots represents the GPR on the reduced space of one out of six
partitions of the parameter space found with LAS. On the y-axis is the basic reproduction number R0 of
the SEIR model for the Ebola epidemic. On the x-axis is the reduced one-dimensional AS of the relative
cluster. In the plots are also represented the 200 training points scattered among all six clusters and used
to find the partition of the parameter space with hierarchical top-down clustering and build the GPRs.
340 Chapter 16. Reduction in Parameter Space
Each response surface is tested on an independent set of 500 samples. The R2 scores are
shown in Table 16.2. The KAS, LAS, and NLL methods present an evident gain with respect to
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the use of the AS to design a global response surface. Usually the overhead of the KAS, LAS,
and NLL training procedures over the standard AS method is negligible with respect to the cost
for obtaining the input-output datasets.
Table 16.2. Comparison of the R2 scores of the response surfaces for the Ebola model built with
the AS, KAS, LAS, and NLL methods.
where m is the total number of parameters and wij , λj are the elements of the eigenvector matrix
and the eigenvalues, respectively.
Conversely to the first eigenvector, which provides information through the sign of its com-
ponents, the activity scores remove this information by squaring the components. This results
in a much more comparable metric with respect to other widespread sensitivity metrics such as
Sobol’s total sensitivity indices and derivative-based global sensitivity measures.
16.4.4 Optimization
A novel application of ASs for optimization tasks is the coupling with the classical genetic algo-
rithm (GA), called the active subspace-based genetic algorithm or ASGA [161]. The main idea
is to perform the typical reproduction and mutation steps on the lower-dimensional AS of the
current generation. In Figure 16.9 is depicted the scheme of the ASGA procedure, where the
yellow boxes denote the parts performed for ASGA which are not done in the classical GA. The
AS back-mapping is performed as described in Subsection 16.2.1. Of course, given a sample in
the AS, in general, there are many points in the full space which are mapped onto the same point
on the reduced space. The new algorithm allows for the selection of more than one point in the
full space in order to explore different parts of the domain simultaneously.
ASGA has been tested on classical high-dimensional benchmark functions, exhibiting great
performance in terms of both better convergence to the actual minimum f (x∗ ) for a given com-
putational budget and better spatial convergence to the location of x∗ . The method has also been
tested for shape optimization in a naval engineering context [157].
The coupling of AS with optimization algorithms is not limited to GA. We cite [226, 340]
for application of ASs to Bayesian optimization for aerostructural design of an aircraft wing for
fuel-burn minimization and for materials design, respectively.
16.4. Applications 341
3 individuals
First generation
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
with N genes
Accordingly to
Selection best fitness
From N genes
AS Projection to M genes
Crossover
Reproduction
Random
Mutation mutation
From M genes
AS back mapping to N genes
Offsprings +
New generation best individuals
Figure 16.9. ASGA scheme. The main steps of the classical GA are depicted from top to bottom.
Projections onto and from a lower-dimension AS are highlighted with yellow boxes, which are specific to
ASGA.
Sampling from the posterior distribution of the latent variables x with Monte Carlo is costly
due to the evaluations of the model m:
In order to avoid this we can approximate the model with a surrogate obtained by employing an
342 Chapter 16. Reduction in Parameter Space
It is important to observe that in this way the inactive variable z has a multivariate Gaussian
distribution and is independent of the random variable f .
ASs are also employed to design a surrogate model of the parameters-to-observable map of
a PDE model in the context of optimal experimental design [624]. In a Bayesian framework, the
expected information gain is maximized in order to find the best sensor locations. In this case, the
function f of (16.21) is the output of a complex PDE model with an intrinsic low dimensionality;
thus reduction in the parameter and output spaces is beneficial for decreasing the computational
cost of the optimization.
16.5 Conclusions
In this chapter we presented a general overview of different parameter space reduction tech-
niques, both linear and nonlinear, to fight the curse of dimensionality, together with a wide range
of possible applications.
In the first part we focused on AS-based methods in a ridge approximation framework. We
started with classical AS, and then we introduced novel extensions such as KAS and LAS. Both
techniques try to move from linear to nonlinear manifold approximation. We then presented a
multifidelity framework to enhance GPR through a low intrinsic dimensionality bias coming from
parameter space reduction. To complete the overview we briefly mentioned some gradient-free
approaches and nonlinear methods not based on AS, such as NLL and active manifolds.
In the second part we moved to the application side, showing how it is possible to use the
techniques presented in the first part to solve common numerical analysis problems. We provided
16.5. Conclusions 343
a comparison in terms of the regression accuracy of response surfaces built over the reduced pa-
rameter space identified by AS, KAS, LAS, and NLL. We used a common benchmark dataset
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
created to model the spread of Ebola. We then briefly touched on how to exploit AS for sen-
sitivity analysis with activity scores, and how to enhance genetic algorithms by performing the
reproduction and mutation steps over the AS of the current population. Finally, we showed
with a simple example how to treat inverse problems and commented on possible coupling of
AS with other reduced order methods. Other interesting applications, not touched on in this
work, are the reduction of highly parametrized artificial neural networks (ANNs); in this direc-
tion, in [394], the authors propose an AS approach to decreasing the memory consumption of
modern ANNs for image recognition and classification. For a deeper presentation, please refer
to Chapter 20.
More work has to be done to better integrate parameter space reduction with other ROMs,
especially for highly parametrized systems and outer-loop applications such as shape optimiza-
tion. Future research lines should also focus on the development of reliable nonlinear methods,
possibly incorporating physical constraints, without losing the interpretability characterizing the
linear methods.
Chapter 17
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Geometrical
Parametrization and
Morphing Techniques
with Applications
17.1 Introduction
As repeatedly pointed out, the present book focuses for the most part on parametrized partial
differential equation (PDE) problems. In many cases of interest, the input parameters of such
PDE problems happen to be associated with the shape of the computational domain, or of one
of its portions. For instance, virtual prototyping applications very often investigate the influence
that the shape of a given craft or artifact has on its predicted performance. In some case, through
repetitive resolution of PDE systems, such applications go as far as identifying the optimal shape
of the craft or artifact under study, or quantifying the sensitivity of its performance to variations
of its geometry. In these last cases, the virtual prototyping application pipelines will include op-
timization or uncertainty quantification tools. Naturally, such mathematical algorithms can only
be fed with numbers. Thus, the main purpose of shape parametrization is to map any geometric
variation of the object under study to variations of numbers—the parameters—which can then
be processed by any other mathematical algorithm used in the virtual prototyping pipeline.
To allow for completely automated virtual prototyping and optimization pipelines, it is very
important that the shape parametrization algorithm be able to automatically obtain the deformed
domain and computational grid once the shape parameter values are prescribed. This constraint
not only ensures that repeated solution of parametrized PDE problems is possible when shape
parameters are included among the inputs but also significantly limits the time required for mesh-
ing, which in such a framework typically represents the most time-consuming activity in terms
of human operator hours.
A further constraint for shape parametrization algorithms arises from their application in the
context of model order reduction. As discussed with greater detail in several chapters of the
present book, many modal decomposition strategies used in reduced order models (ROMs) are
based on a snapshot matrix which is obtained under the assumption that the number of degrees
of freedom and the order of the PDE problem remain unaltered as input parameter values are
modified. For this reason, the shape parametrization algorithm must not alter the topology of
the computational grid used to resolve the PDE. Thus, a simple approach consisting of creating
a new mesh after the shape of the domain has been altered is not suited for ROM applications.
345
346 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
A more comprehensive strategy which deforms the computational mesh along with the domain
geometry is needed.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
For this reason, in the present chapter we first analyze algorithms that are suited for deforming
codimension-one subsets of R2 or R3 , which are used to define the geometry of the computa-
tional domain by describing their boundary representation. In two dimensions, codimension-one
subsets are typically described by the points of piecewise linear curve representations or by
the control points and nodes of nonuniform rational basis spline (NURBS) curves included in
computer aided design (CAD) files. In three dimensions the most common representations of
codimension-one subsets are given by the points and connectivity of surface triangulations, or
by the control points and nodes of NURBS saved into CAD files. After this, we will focus on
the deformation of codimension-zero computational domains in R2 or R3 , which are typically
used to define computational grids. In both two and three dimensions, these domains are typi-
cally described by the nodes and the connectivity of polygonal or polyhedral tessellations. It is
important to stress that an additional advantage of deforming both the domain geometry and the
computational grid in a consistent way is that it allows for a strong interaction between design
and simulation teams. In fact, the shape modifications validated by PDE-based simulations will
be readily available for production with no further processing.
In this chapter, we first consider parametrization algorithms which attempt to come up with
a general law to deform at once the whole two- or three-dimensional space in which the mor-
phed object is embedded. Such general purpose algorithms typically operate by extending or
interpolating the displacement of a relatively small set of control points to all the points of the
two- or three-dimensional space. Thus, the displacements of such control points become the user-
prescribed shape parameters. The position of each point of the objects of interest is then displaced
by interpolating the control point displacement to obtain the modified shape. This paradigm is
common to free-form deformation (FFD), originally proposed in [533], and to the application of
radial basis function (RBF) interpolation [82] and inverse distance weighting (IDW) interpola-
tion [537] to shape parametrization. A considerable advantage of these algorithms is that they
can indifferently operate on the points of domain geometries or on the nodes of computational
grids. For this reason they have been easily applied to CAD geometries, as for instance in [574],
and to computational meshes, as in [157]. On the other hand, one of the possible drawbacks of
these tools is that the shape parameters identified are not necessarily related to the specific phys-
ical behavior of the object under study. For instance, a simple scaling parameter can affect the
aspect ratio of a ship hull and have a clear effect on its hydrodynamic performance. On the other
hand, the effect of the prescribed displacement of one of many shape control points on the hull
performance is more difficult to establish. Ideally, in several engineering applications the shape
parameter should be readily associated with qualitative variations in the performance of the de-
signed object. Moreover, in some cases general purpose algorithms might violate engineering
constraints on the shapes to be deformed. This is for instance the case for aircraft and ship pro-
pellers, which are objects with a highly engineered shape and strict design rules [101]. Thus, we
will also discuss cases in which the shape parametrization strategy is tailored to certain bodies to
alter their shape, preserving the geometric characteristics they have been designed to possess.
The content of this chapter is organized as follows. Section 17.2 introduces and discusses
the mathematical background of different algorithms for shape parametrization. After general
purpose methods (FFD, RBF, IDW) are presented, we will also describe examples of different
object-specific deformation strategies. In Section 17.3, we discuss the application of all these
strategies to the deformation of both domain geometries and computational meshes. In this
context, we will also present several applications of the proposed shape parametrization method-
ologies to industrial and real-life problems. Finally, Section 17.4 provides some concluding
remarks, as well as some interesting perspectives on the field of shape parametrization in the
next few years.
17.2. Parametrization Techniques 347
In this section, we present a diverse range of shape parametrization tools. We divide such de-
formation techniques into two main categories based on their objective: general purpose shape
parametrizations and object-specific ones.
FFD
FFD is a well-known morphing technique for both local and global deformations of a given
object. Originally presented in [533], in the last decades it has gained popularity in the context
of parametric PDEs and model order reduction for shape optimization studies (see [521, 507, 35,
519, 154]), optimal flow control problems [509], and sailing boats [379], in naval engineering
[574, 155, 168, 569, 572, 156], CFD applications [160], and isogeometric analysis [518, 213],
to cite a few. Since the deformation induced by the FFD does not depend on the topology of the
object to be morphed, it is very versatile and nonintrusive, especially for complex geometries or
in industrial contexts [508, 517, 571].
FFD deforms the physical domain Ω by propagating the displacements of a lattice of points,
called FFD control points, placed in Ω. The map M is the composition of three different geo-
metric deformations, i.e.,
where the maps ψ and T̂ are depicted on a simple bidimensional example in Figure 17.1. The
function ψ represents the affine mapping from the physical domain Ω to its reference configura-
tion Ω̂. The FFD control points P are moved, and the morphing is propagated to all the points
within the lattice, usually by using Bernstein polynomials, but other choices are also possible.
The displacements of such control points along the cardinal directions define the parameter vec-
tor µ. The map T̂ (·, µ) : Ω̂ → Ω̂(µ) for a general three-dimensional case with a lattice of
L × M × N control points is then defined as
L X
X M X
N
T̂ (x, µ) = blmn (x)Plmn (µlmn ) ∀x ∈ Ω̂, µ ∈ P, (17.2)
l=1 m=1 n=1
where blmn stands for the Bernstein polynomials of degree l, m, n in each direction, and
Plmn (µlmn ) := Plmn + µlmn , with Plmn indicating the coordinates of the FFD control point
with indices l, m, n. By exploiting the good properties of such polynomials it is also possible to
impose continuity constraints by adding control points. This is particularly useful at the interface
348 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 17.1. Sketch of the three deformation maps which compose the FFD morphing.
between the unmorphed part of the geometry and the lattice of control points, for which physical
constraints usually have to be satisfied.
In the classical offline-online framework, we emphasize that the terms blmn (x) can be com-
puted in the offline phase, while in the online phase we just perform linear combinations of
precomputed quantities depending on x- and µ-dependent control point locations. Moreover,
since ψ and ψ −1 are affine maps, the two phases remain separated.
RBF Interpolation
By RBF [82] we mean any smooth real-valued function ϕ̄ : Rd → R whose value depends only
on the distance between the input and some fixed point. This means that there exists ϕ : R+ → R
such that ϕ̄(x) = ϕ(kx − ck), where k · k stands for the Euclidean distance and c is some fixed
point.
To illustrate the many possible choices among all the RBFs, in Figure 17.2 we consider
a simple univariate example where we approximate the function f (x) = 0.5 sin(4πx) +
exp(x cos(2πx)) using six different types of RBFs:
2
• Gaussian [82], defined as ϕ(r) = e−(rε) ;
2
y
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
x x
Figure 17.2. Interpolations corresponding to six different RBFs. The exact function is depicted
with a red dashed line. The parameter ε is equal to the inverse of the average distance between the sampling
points.
0.0
0.1
0.2
0.0 0.2 0.4 0.6 0.8 1.0
x
Figure 17.3. Interpolation error for different types of RBFs. Refer to Figure 17.2 for the actual
interpolations.
We remark that the list does not cover all the possible RBFs. Figure 17.3 reports the interpolation
error over the entire domain for the different RBFs described above.
We follow [407, 390] for a brief presentation of the RBF interpolation method in order to
define M(·, µ) for this technique. Let {xCi }N i=1 be the control points placed over the surface
C
of the object we want to parametrize. We can induce a deformation by moving these control
points and interpolating them with a new surface. The parameter vector µ is composed by the
displacements of the RBF control points. Let ϕ be the selected RBF and {γi }N i=1 be the weights
C
NC
associated with the RBF centered in {xCi }i=1 , respectively. We can define the deformation map
350 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
NC
X
M(x, µ) := s(x, µ) + γi (µ)ϕ(kx − xCi k) ∀x ∈ Ω, µ ∈ P, (17.3)
i=1
where s(·, µ) is a polynomial term of degree one, which means s(x, µ) = S(µ)x + c(µ). The
total unknowns are d + d2 for the polynomial term and d × NC for the weights, where d = 2 or 3,
depending on the dimension of Ω. Regarding the γi , let {yCi (µ)}N i=1 be the new control points
C
obtained from xCi after applying the displacements defined by µ. The interpolation constraints
we impose are
M(xCi , µ) = yCi (µ) ∀i ∈ [1, . . . , NC ]. (17.4)
For the remaining unknowns due to the polynomial term, we impose the conservation of the total
force and momentum (see [82]) as
NC NC
(j)
X X
γi (µ) = 0, γi (µ)xCi = 0 ∀j ∈ [1, . . . , d], (17.5)
i=1 i=1
IDW Interpolation
IDW is an interpolation method originally proposed in [537]. Let {(xCi , f (xCi )}N i=1 , where
C
such that IIDW [f ](xCi ) = f (xCi ) for i = 1, . . . , NC . The functions wi : Rd → R are defined
as
kx−xCi k−p
(
if x 6= xCi ,
(17.7)
PNC
wi (x) = j=1 kx−xCj k
−p
1 if x = xCi for i = 1, . . . , NC ,
where the power −p controls the rate of influence of the ith input point and k · k is the Euclidean
norm. Such functions require one to compute all the distances between the point to interpolate
x and the input data {xCi }Ni=1 . Assuming the weight function becomes negligible after a certain
C
where for the sake of notation we define the distance function d : Rd × Rd → R+ as d(x, y) =
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
kx − yk.
This interpolation can be adopted as an efficient tool for shape parametrization problems.
Among all the contributions in the literature, we suggest [623] for a parametric computational
grid by means of IDW, [202] for application of IDW to FSI problems, and [31] for an example
of domain parametrization in a ROM framework.
The main idea for shape parametrization using interpolation techniques is to extend the de-
formation we impose to a limited set of control points by interpolating their displacements. We
define this parameter-dependent displacement at point xCi as fµ (xCi ) ≡ f (xCi , µ) ∈ R for
i = 1 . . . , NC . Interpolating the displacements with IDW, we obtain
NC
X
IIDW [fµ ](x) = wi (x)f (xCi , µ), x ∈ Rd , µ ∈ P. (17.9)
i=1
We remark that we have considered until now scalar displacement, but the IDW extension to
vector functions is trivial: assuming all components are independent, we simply apply IDW to
each component. Thus, considering a d-dimensional vector as the displacements, we can derive
the deformation map M(·, µ) : Rd → Rd from (17.9):
NC
X
M(·, µ) := wi (·)f (xCi , µ), x ∈ Rd , µ ∈ P. (17.10)
i=1
Equation (17.10) shows the decoupling between the spatially dependent terms (wi (·)) and the
parameter-dependent ones (f (xCi , µ)). This implies that for any x, we can precompute (and
store) the weight functions, requiring us to compute only the parameter-dependent term for any
new evaluation of M(·, µ) in the online phase.
In addition, there are several situations in which general purpose algorithms cannot be em-
ployed without a more invasive and complex adaptation. As an example, consider the design
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
of ship or yacht hulls. Typically, the lateral walls of a hull are constrained to be vertical or at
least straight because this results in easier management and arrangement of internal passengers
or cargo areas. Compared to a global constraint, such as the aforementioned fixed-volume one,
this local constraint is more difficult to fulfill. Both RBF and FFD would in fact require signifi-
cant modifications to be effective for generating parametrized hulls with vertical or straight side
walls. An even more important case of shapes with local constraints limiting the effectiveness of
general purpose shape parametrization algorithms is that of aircraft wings and propeller or turbo
machinery blades. These objects were originally designed to have airfoil sections with standard
and tabulated shapes with well-documented fluid dynamics performance. Because remarkably
efficient estimates of the overall wing or blade performance can be obtained from that of its air-
foil section, it is very important that any parametrization algorithm applied on these artifacts not
lead to sections with nontabulated shape or performance. It is clear that in this case RBF and
FFD, as they have been described, also cannot generate shapes with this property.
In light of all the problems illustrated by these examples, in this subsection we will discuss
shape parametrization algorithms adapted to specific design problems. In particular, making use
of some of the aforementioned design cases (i.e., planing hull yachts and naval propeller blades),
we will illustrate possible approaches for when general purpose shape parametrization methods
cannot be directly employed.
We now focus on the case of the planing hull yacht illustrated in Figure 17.4. The geometry of
this kind of hull is typically characterized by a V-shaped lower part, surmounted by a lateral wall
which is typically straight and—except in the bow region—has a constant angle with respect to
the vertical direction. The V-shaped bottom part of the hull and its lateral walls are typically
divided by a sharp edge called the chine. The example in Figure 17.4 has a double chine line
which delimits a narrow chine surface.
Figure 17.4. Front view (top) and side view (bottom) of a planing hull featuring a double chine
line and substantially straight walls above the chine line.
17.2. Parametrization Techniques 353
In the hydrodynamic design process of this kind of hull engineers typically want to investigate
how shape modifications affect performance in terms of lift, drag, and stability (or seakeeping).
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
It is quite clear that deforming the V-shaped bottom part of the hull, which is in contact with
the water throughout the navigation, will have the most impact on the overall hydrodynamic
behavior of the hull. For this reason, design engineers are mainly interested in modifying the
lower part of the hull and—to a less extent—the chine surface, while keeping the top part of the
hull substantially nondeformed. Under these requirements, FFD can hardly be used because its
smooth shape functions would not allow one to sharply separate what happens above and below
the chine lines. As for RBF, it is still possible to employ its control point displacements to impose
the desired deformation on the bottom part of the hull, and a somewhat rigid displacement on
the top part, to accommodate for the bottom part deformation. Yet this opens the problem of
identifying the RBF control point displacements that would result in the desired deformations of
the bottom and top hull regions. So, rather than solving the latter problem, it is convenient to deal
with the more intuitive problem of developing a parametrized law to displace any point on the hull
surface and generate admissible overall deformations. As we will see in Subsection 17.3.3, RBF
is still used in this context as an extremely valid tool to extend the hull surface deformations to
the surrounding volume, adapt the volumetric computational meshes, and carry out parametrized
fluid dynamics simulations.
The parametrized law to displace any point on the hull is in this case developed by taking
advantage of the main design features of the object at hand. As suggested in Figure 17.5, the
transverse vertical sections have the same topology for most of the hull length. Each section is
in fact characterized by the bottom vertex coinciding with the section plane intersection with the
keel line and by two vertices resulting from the section plane intersection with the chine lines.
So, the devised parametrization strategy acts by modifying each section with the same strategy
based on these easily identifiable points. More specifically, the V angle of the hull bottom is
altered by rigidly rotating all the section points located between the intersection of the internal
chine line and the keel line around the keel line intersection by the desired angle θ. This rotation
Figure 17.5. Vertical section of a planing hull perpendicular to its longitudinal axis. The bottom
vertex of the section is the intersection of the section plane and the keel line. The intersections of the two
chine lines are also visible on the bottom left side of the section. The first step of the parametrization
strategy is a rigid rotation of the bottom part of the section around the keel intersection. This is combined
with a rigid translation of the chine and wall surface sections needed to keep the top part of the section in
contact with the displaced bottom part.
354 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
results in a net displacement of the internal chine line intersection point. This displacement is
used to rigidly translate the section of the chine and wall surfaces and ensure that the top part
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
of the section remains connected to the rotating bottom. Under the same principle, once these
two steps are completed, it is possible to add rotations to the chine surface and, if desired, to the
wall surface. Finally, a further parametrized function can be introduced to modulate the sectional
rotations introduced along the hull’s longitudinal coordinate.
In summary, the present example suggests that when a highly engineered object such as a
planing yacht hull is to be parametrized by fulfilling some specific shape constraints, a viable
possibility is to exploit the main topological features of the original geometry. Such topologi-
cal features are often a result of the characteristics that designers want their object to possess.
So, as in the case of the keel and chine lines for a planing hull, they can be used to build a
parametrization law that is useful to engineers.
Propellers
As previously mentioned, aircraft wings and propeller or turbo machinery blades are further ex-
amples of objects for which RBF, IDW, and FFD are not ideal shape parametrization algorithms.
This is mainly because such objects are designed starting from a set of two-dimensional airfoil
sections, which are placed in the three dimensional space according to some strict design rule. If
we consider for instance a propeller blade, a specific design table first reports the list of airfoil
section shapes to be used in its generation. In such a design table, each airfoil section, originally
described in the R2 space, is also complemented with a set of parameters (radius, chord length,
pitch angle, skew angle, and rake) which indicate how the airfoil section should ultimately be
placed in the R3 space. As depicted in the left part of Figure 17.6, the airfoil sections are first
laid on cylinders with axes coinciding with that of the propeller. Thus, each airfoil section is
associated with a specific radius. The chord length parameter represents the distance between
the leading and trailing edges of the airfoil, projected on the section cylinder. The skew and
Figure 17.6. Front view (left) and lateral view (right) of a ship propeller. In both images, one of
the five propeller blades’ surface is transparent, and a set of airfoil cylindrical sections is indicated in red
to show how the blade surface is generated as the envelope surface of the single sections.
17.2. Parametrization Techniques 355
pitch angles represent instead the rotations prescribed for each section around the blade axis and
the entire propeller axis, respectively. Finally, the rake is the translation along the propeller axis
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
prescribed for each section. For more details about propeller blade design, the interested reader
is referred to in [101].
It should at this point appear clear that the design procedure described automatically leads
to parametrized propeller blades. In fact, the chord length, pitch angle, and skew angle at each
radius corresponding to the airfoil sections available are indeed parameters of the generated
propeller blade. Moreover, several different families of parametrized airfoil sections are available
across the engineering community. For instance, the NACA airfoils [8] are parametrized based
on quantities such as maximum thickness, maximum deflection of the camber line, and location
along the chord of the maximum deflection point.
Despite the fact that the previous description suggests that the geometry of propeller blades—
as well as turbo machinery blades and aircraft wings—possesses an inherent parametrization, it
is still possible to modify the parametrization to have better control over the finalized shape. As
depicted in Figure 17.7 for the case of pitch radial distribution, it is possible to use B-splines
to interpolate the blade parameter variation along the radial coordinates. This results in the
possibility of switching the shape parameters to the control points of the single chord length,
pitch, skew, and rake radial distribution curves. Given the smoothness of B-splines, this ap-
proach results in a smoother parameter dependence, especially considering that the CAD mod-
elers used to build the three-dimensional surfaces passing through the airfoil sections are also
based on B-splines and NURBS surfaces. A further advantage of this approach is that if the
two-dimensional airfoil parameters are also interpolated with this strategy, the number of air-
foil sections can be arbitrarily increased to provide the CAD modeler with more sections and
allow them to have more control over the generated three-dimensional surfaces. For an efficient
selection of such parameters using the active subspace (AS) method, please refer to [404] and
to Chapter 16.
The standard generation of the three-dimensional propeller blade from its design table and
from B-spline interpolated radial distributions of airfoil parameters have been implemented in
the Python package BladeX [210]. Given sectional values of airfoil section, radii, chord length,
pitch and skew angle, and rake, BladeX can directly generate CAD representations of the cor-
responding blade surface. In addition, the software supports the B-spline representation of the
airfoil section parameter variations along the relative radial coordinate, which as mentioned can
be used to reparametrize the blade shape or to enhance the smoothness of the resulting CAD
surface.
1.50
1.45
1.40
1.35
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
r/R
Figure 17.7. The airfoil section pitch as a function of relative radial coordinate. The plot shows
both an initial pitch distribution (red continuous line) and the control points associated with its B-spline
interpolation (red piecewise linear line). A new, smooth pitch radial distribution (blue continuous line) is
then obtained using a displaced set of pitch control points (blue piecewise linear line).
356 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
The parametrization strategies described in the previous section have been applied to different
design problems. In this section, we will describe a wide spectrum of shape parametrization ex-
amples, with particular focus on practical details such as the file format containing the geometric
information of the considered objects. In addition, we will distinguish between cases in which
the parametrization tools only act on the surfaces describing the object of interest [574, 404] and
those that are used to deform a three-dimensional computational grid [32]. In the former case,
the deformation of a codimension-one, R2 surface embedded in R3 is studied. In the latter case,
the complete morphing of a codimension-zero computational domain Ω ⊂ R3 is instead carried
out. Recall that the codimension of a subspace U of a vector space V is the dimension of the
quotient space V /U . It is indicated by the symbol codimV U and is equal to the dimension of the
orthogonal complement of subspace U in V , namely
Triangulated Surfaces
Figure 17.8. FFD application to the STL triangulation representing the bulbous bow of a cruise
ship hull. The left panel, which represents the original bow geometry, also displays the FFD undeformed
control point lattice. On the right, the visible displacement of the control point results in a deformed STL
geometry.
Figure 17.9. Two examples of a triangulated surface of a carotid artery deformed with RBF. The
original geometry is emphasized by the blue dots, while the RBF control points are depicted in red (the
deformed ones) and in green (the undeformed ones).
discover possible occlusions after the bifurcation of the carotid by simulating the blood inside
the vessel.
CAD Surfaces
We now describe some examples of the application of the shape parametrization strategies dis-
cussed in Section 17.2 to NURBS surfaces [530]. NURBS are an important component of numer-
ous industry and computer graphics CAD standards, such as IGES [336], STEP [4], and ACIS.
A full mathematical description of NURBS surfaces and their application to analytic geom-
etry is beyond the scope of the present work (we refer the interested reader to [488, 489]). Yet,
for our purposes it is at least important to recall here that NURBS curves and surfaces are de-
358 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
fined by their order, by a set of weighted control points, and by a knot vector. Control points
in particular either are always directly linked to the curve or surface or behave as if they were
connected by a spring to it. For this reason, they offer the possibility of modifying the NURBS
surface by altering the position of the control points. This very intuitive editing mechanism—
which is among the features that made this mathematical instrument very popular in the CAD
and computer animation industries—is exploited here to apply any parametrization strategy to
NURBS surfaces. As illustrated in Figure 17.10, a NURBS curve is deformed using FFD by di-
rectly applying the parametrization algorithm to the NURBS control points and then computing
the resulting modified shape.
It is important to point out here that the deformed surface obtained by moving the CAD
control points is identical to the one obtained by moving every surface point only if the trans-
formation law applied is affine [488]. We also recall that most of the parametrization algorithms
discussed in the present work do not, in general, lead to affine transformations. Thus, we must
remark that deforming a CAD shape by displacing its control points might result in slight dif-
ferences compared to deforming it by a surface triangulation. We report that, for all the exam-
ples discussed in the present chapter, the difference between applying the same deformation to
NURBS surfaces or to their triangulated equivalent is quite modest. Yet, we stress that this issue
must be taken into account if consistency between NURBS and triangulated surface deforma-
tions must be enforced up to a very strict tolerance. Ideally, in the latter case it is possible to
resort to least squares fitting to generate a novel NURBS surface passing as close as possible to
the deformed triangulation. We warn the reader, however, that this procedure would still bring in
the possibly significant source of error associated with the least squares minimization.
Figures 17.11 and 17.12 depict an example of the application to CAD geometries of the
parametrization algorithms described in Section 17.2 and presented in [155]. The shape consid-
ered is that of the DTMB-5415 hull [401], which was originally conceived for the preliminary
design of a US Navy Combatant ship. FFD is used here in the bow regions to displace the control
points of the hull NURBS surfaces and create a family of hulls which feature deformed sonar
domes. The images display the front and side views of both the original DTMB-5415 IGES
geometry (right and bottom), and of the IGES geometry of a specific modification in which the
sonar dome has been enlarged and inclined downward (left and top).
As can be appreciated in the pictures, the DTMB-5415 hull portrayed in Figures 17.11
and 17.12 is composed of three NURBS surfaces. Such faces border each other through shared
edges in which the shape is C 1 or C 0 continuous only up to a tolerance specified in the IGES
file containing the hull data. We use this very common example to remark that an additional
problem associated with deforming NURBS surfaces by displacing their control points is that
17.3. Application to Different Geometry Specifications 359
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 17.11. An application of the FFD algorithm to the NURBS surface specified by an IGES
file. A front view of the DTMB-5415 navy combatant hull featuring a modified and enlarged bow is displayed
on the left; the original hull surface is displayed on the right for reference.
Figure 17.12. A side view of the IGES geometry of the DTMB-5415 navy combatant hull shown
in Figure 17.11. The top picture depicts the CAD surface of the hull, modified via FFD. The bottom picture
shows the original hull surface.
it can introduce discontinuities in shapes composed of several CAD patches. Despite the fact
that all parametrization strategies described in this chapter are designed to be continuous maps
from R3 to R3 , the control point distribution in two bordering faces can in principle be so dif-
ferent that the deformation affects the two surfaces unequally. This might potentially alter the
continuity tolerance of the original CAD file and result in small gaps, overlaps, or normal vec-
tor jumps appearing between the NURBS surfaces which were originally smoothly joined. This
loss of C 1 and/or C 0 continuity is more pronounced when the deformation map gradients are
high relative to the local NURBS control point density. Thus, a solution to mitigate the problem
is to add control points to the NURBS surfaces. In the vast majority of cases considered, this
procedure proved effective in obtaining small continuity tolerances in composite CAD surface
deformations, although the increased number of control points results in larger CAD files [570].
The ship propeller family shown in Figure 17.13 represents a different strategy of obtaining
parametrized IGES geometries. The algorithm for propeller design described in Section 17.2 was
used for the bottom-up generation of parametrized ship propeller blades. In this specific case,
each blade is generated as the NURBS surface passes through 12 airfoil sections, each defined
by a different curve in the xy plane and different values of radial coordinate, chord length, pitch
angle, skew angle, and rake. The procedure implemented in the library BladeX [210] allows
360 Chapter 17. Geometrical Parametrization and Morphing Techniques with Applications
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 17.13. Representation of a family of ship propellers generated with bottom-up construc-
tion. The different propellers are obtained in this case by scaling the skew angle curve. In particular, in the
image an original propeller blade shape is presented along with two modifications resulting from scaling
the skew curve by factors of 0.8 and 1.1, respectively.
the airfoil section shape to be specified by means of a NACA four- or five-digit identifier or by
means of a CSV file. Once the radial coordinate, chord length, pitch angle, skew angle, and
rake associated with every section are specified by the user, the airfoil section points are placed
in the three-dimensional space, and a series of NURBS curves are generated, passing through
all the points of each section. Finally, the NURBS surface passing through all the sectional
curves is generated and joined to the propeller hub. The latter operations are carried out using
functions of the CAD Python library PythonOCC.10 In the example reported in Figure 17.13, the
original blade shape is portrayed along with two variants generated by altering the skew angle
distribution along the radial coordinate. We recall that the skew angle designs the position of
each radial blade section along the circumferential coordinate of the propeller disc plane. Thus,
the blade generated by scaling the skew values by a factor of 1.1 appears more curved along the
radial direction than in the original design. On the other hand, scaling the skew values by a factor
of 0.8 results in a propeller with blades that appear straighter along the radial direction.
We finally point out that the specific deformation displayed in Figures 17.11 and 17.12 is
part of a family of 130 IGES hulls generated via FFD in the framework of a CFD simulation
campaign presented in [155]. In that work, the geometric deformation was concentrated in the
region surrounding the bulbous bow. A further set of 130 hulls was obtained by deforming the
full length of the hull and was used to carry out the CFD simulations presented in [574]. A family
of 500 ship propellers like the one depicted in Figure 17.13 was also designed in the framework
of an optimization campaign presented in [404]. In all these works, active subspaces (ASs) were
successfully used to detect possible redundancies in the geometric parameters introduced with
FFD (please refer to Chapter 16). On this note, we point out that AS analysis is an extremely
useful postprocessing tool for complementing shape parametrization algorithms because it allows
a reduced number of combined parameters to be identified which have the most effect on the
output of interest. In addition, it can also be used to evaluate the output sensitivity to variations
of the original geometric parameters and eliminate inconsequential parameters.
10 Available at https://github.com/tpaviot/pythonocc-core.
17.3. Application to Different Geometry Specifications 361
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 17.14. Examples of computational domain deformation using FFD: the first picture (left)
shows the edges of the original mesh, while the others illustrate two deformed domains.
A first example, presented in Figure 17.14, illustrates the application of FFD to deforming the
structured surface mesh surrounding a square in a two-dimensional application. The images,
which refer to the original configuration (left) and to two deformed domains (center and right),
depict a closeup of the grid region surrounding the square.
Figure 17.15 shows instead an application of RBF in the context of a CFD simulation cam-
paign [639]. The pictures show two possible modifications of the Ahmed body [10] obtained
by altering the inclination of the diagonal surface at the rear end of the body. The computa-
tional grid for the CFD simulations is deformed by distributing a series of control points on the
boundaries of the three-dimensional domain and by displacing a portion of those located on the
modified diagonal surface. The vertical cut of the resulting volumetric grids is shown in the
plots on the top. The corresponding bottom images display the path lines of the mean velocity
field resulting from the unsteady Reynolds-averaged Navier–Stokes (RANS) simulations carried
out with the deformed grid. As can be appreciated, the deformation strategy is able to alter the
original computational mesh without impairing the convergence and effectiveness of the CFD
simulations.
Figure 17.15. The application of RBF shape parametrization to the Ahmed body and the sur-
rounding volumetric mesh for CFD simulations. In the test illustrated, the inclination of the diagonal
surface at the rear end of the Ahmed body is modified to investigate its effect on aerodynamic performance.
Two configurations, including a vertical cut of the volumetric mesh, are shown in the top images. For ref-
erence, path lines of mean velocity field resulting from the corresponding unsteady RANS simulations are
shown in the bottom pictures.
Figure 17.16 shows the deformation of the CFD computational mesh past a Potsdam propeller
test case (PPTC) ship propeller, following a variation of the blade pitch in the tip region obtained
with the strategy described in Subsection 17.2.2. In the plot, the blue lines represent the edges
of a cut of the volumetric mesh obtained by means of a plane perpendicular to the propeller axis.
The red lines refer to the edges of the deformed mesh in correspondence with the same sectional
plane. The closeup of the blade tip region displayed on the right shows how the deformation of
the blade surface is smoothly propagated throughout the internal nodes of the volumetric mesh,
avoiding the presence of inverted or highly skewed cells.
Finally, we want to stress that for practical reasons the RBF extension to volumetric meshes
is often convenient for surfaces originally deformed with FFD. As previously discussed, in prin-
ciple FFD allows for simultaneous deformation of both the boundary surface and the volumetric
grid. Yet, when deforming volumetric meshes, the aim is to distribute the deformations over a
high number of cells, rather than concentrating all the displacements in a very confined region
in which cells can get distorted or even inverted. But because FFD only affects points located
within the control point lattice, the latter must extend to a bigger volume. In addition, to max-
imize the volumetric mesh quality, the user must include more control points in the lattice to
make sure that different deformation magnitudes are imposed in regions close to the deformed
surface and far from it. Such manual control over the local mesh deformation can often become
quite cumbersome.
For these reasons, in several cases it will be convenient to employ FFD only to deform the
boundary surface mesh or a triangulated format of its geometry. Then, after the surface mesh has
17.4. Conclusions and Future Perspectives 363
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 17.16. Example of computational domain deformation applied to a propeller design prob-
lem, seen at the radial plane. The blue edges refer to the original mesh, whereas the red ones show the
deformed grid.
been modified using FFD, it is possible to resort to RBF to propagate the boundary displacements
to the internal nodes of the volumetric mesh for CFD simulations. In a broader sense, RBF is
an interpolation algorithm and is used here to estimate the node displacement function on every
point of a domain, based on the function values prescribed on the boundary. In the framework
of RBF, this is done by linear combinations of radial functions used to approximate at any point
of the domain a function with values prescribed only at a finite number of points. In the case of
interest, the displacement field function is prescribed only on the points of the boundary surface
and is interpolated in the positions corresponding to every node of the volumetric mesh. The
application of this procedure in a naval engineering test case (see [157] for further details) is
illustrated in Figure 17.17. The illustration displays a sectional view of the original (blue lines)
and deformed (red lines) volumetric mesh around the surface of a ship hull. As documented in
the original publication, the combined FFD/RBF strategy has been used to successfully generate
200 shape variations of a container ship hull and accordingly deform the volumetric grids to carry
out an optimization campaign based on the OpenFOAM [612] CFD solver library. The output
of the checkMesh utility of the OpenFOAM library confirms that the mesh deformation does not
significantly alter either the minimum face area or the minimum cell volume, while it only leads
to a modest—7.2% at most—increase in the average mesh nonorthogonality index. This increase
is well below a level that can significantly affect the result of the CFD simulations.
Figure 17.17. A sectional view of a volumetric grid deformation carried out using RBFs in a
naval engineering problem. The blue lines denote the original volumetric mesh edges, while the red lines
indicate the deformed configuration.
These artifacts also have to be actually manufactured, so general purpose parametrizations can
easily generate objects which are very costly to produce or infeasible.
In the second part of the chapter we presented methodologies to guarantee that any modifica-
tion of the geometry is reflected in the computational mesh. The advantage of these techniques is
twofold: first, the shapes validated by simulations will also be available for manufacturing; sec-
ond, we can easily apply nonintrusive reduced order methods (refer to Chapter 9) to the database
of solutions since we preserve the same number of degrees of freedom.
Future studies will address the possibility of applying the FFD technique while satisfying
some constraints. For example, we might want to preserve the volume of the deformed object, or
we may want to impose the presence of a second object near the one being morphed. Addition-
ally, some machine learning techniques can be exploited to propagate the surface morphing to the
volumetric computational mesh; in this chapter we proposed such a deformation extension using
RBF, which shows good results but may decrease the quality of the deformed mesh. To over-
come this issue, a neural network can interpolate the deformation on grid nodes, adding some
constraints during the neural network training step to produce high-quality deformed meshes.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 18
18.1 Introduction
Ischemic cardiovascular diseases, e.g., coronary artery disease (CAD), and diseases causing
blood regurgitation, such as mitral regurgitation (MR), are life-threatening events and there-
fore worth studying using computational fluid dynamics (CFD) tools. In CAD, plaque build-up
causes partial or complete blockage in blood perfusion to the heart muscle and risk of ischemia,
thrombosis, stroke, and heart attack. Likewise, severe MR leads to extra strain on the heart to
fulfill the blood supply demand and can result in arterial hypertension, atrial fibrillation, stroke,
and heart failure. Another cardiovascular problem worth exploring is the increased risk of aortic
thrombosis in the left ventricular assist devices (LVADs) used to provide mechanical circulatory
support during heart transplantation and destination therapy.
In studying these phenomena, one is interested in identifying the underlying causes, exploring
patency of surgical treatments and assisting devices, addressing the computational limitations,
and developing and implementing robust and efficient computational techniques. For instance,
coronary artery bypass graft (CABG) is a common surgical treatment for CAD and is needed by
the surgeons to identify the risk of restenosis in a noninvasive patient-specific manner. Thanks
to technological strides, the related techniques have advanced to patient-specific hemodynamics
studies; however, the challenge of computational efficiency and robustness remains. Similarly,
it is of interest in MR to identify the main features responsible for the blood flow conditions
caused by the nonsymmetric, wall-hugging jets that are responsible for the regurgitation, while
for LVADs, it is crucial to detect the optimal location and inclination of the anastomosis and to
tune the device flow rate. Corresponding computational modeling involves complex formulations
and thus reduced computational efficiency.
Full-order numerical techniques such as finite element methods (FEMs), finite volume meth-
ods (FVMs), and spectral element methods (SEMs), as is well known, provide computationally
stable and reliable solutions for complex problems such as the cardiovascular ones. Neverthe-
less, this accuracy depends on the degrees of freedom of the mathematical problem, which in turn
rely on the discretization size of the computational domain. Hemodynamics modeling in com-
365
366 Chapter 18. Reduced Order Methods for Hemodynamics Applications
plex cardiovascular geometries, even if restricted to a particular and small part of the heart or the
vessel of interest, requires a sufficiently dense mesh for stable solutions. Therefore, such mod-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
eling problems in the cardiovascular domain, especially with patient-specific anatomies, usually
comprise a large number of degrees of freedom and therefore have a high computational cost.
Furthermore, these problems are set in a many-query environment, meaning that cardiovascu-
lar studies cannot be beneficial unless repeated to tune some parameter values, which raises the
computational cost to an untenable level. This poses the challenge of developing and utilizing
reduced order modeling (ROM) techniques that keep the reliability of the full-order numerical
methods intact and increase the computational efficiency.
ROMs can be combined with other mathematical techniques to effectively address mathe-
matical limitations in cardiovascular hemodynamics problems. For instance, imposing accurate
and meaningful blood flow conditions at the boundaries of the cardiovascular domains is essen-
tial for realistic hemodynamics modeling. Multiscale modeling is used to face this challenge;
it uses lumped network and distributed network models to ensure the accuracy of the boundary
conditions using electrophysiology. The cost for these approaches is increased mathematical
complexity and lack of robustness. In other words, the physiological complexity of the biomed-
ical applications requires one to address several hemodynamics and mathematical challenges,
such as imposing accurate and meaningful boundary conditions and ever-changing hemodynam-
ics patterns.
This chapter illustrates the application of the ROM framework to the three cardiovascular
problems mentioned above: CABGs, the Coandă effect in regurgitant mitral valves, and LVADs.
These applications address several of the mentioned challenges regarding computational effi-
ciency and robustness.
18.2 CABGs
Coronary arteries have a diameter of only a few millimeters and are responsible for perfusing
the heart muscle with oxygenated blood. Atherosclerosis, or plaque build-up inside the arteries,
causes structural rigidity, and the resulting narrowing of the arteries reduces the blood supply.
Severe blockage leads to ischemia and can result in stroke, angina attack, or heart attack. In acute
stages, the disease requires invasive surgical treatments such as CABGs. CABGs are external
vessels anastomosed to the blocked coronary arteries and serve as alternative conduits for the
blood. Here we will illustrate the application of ROM to an optimal flow control problem (OFCP)
(Chapter 4) related to CABGs [638].
∇ · v (µ) = 0 in Ω. (18.2)
18.2. CABGs 367
Next, different boundary conditions are chosen for ∂Ω. The boundary of the walls, i.e., Γw , is
assumed to be rigid and nonpermeable and, therefore, no-slip boundary conditions are considered
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
there. Furthermore, the parametrization is introduced through the inflow conditions; i.e., the
Reynolds number defining the blood flow velocity at the inlets Γin is the parameter at hand. The
inflow velocity is imposed as a Poiseuille profile owing to the pipelike structure of the vessels
and the Dirichlet conditions. Lastly, at the outlets Γo , the unknown control variable denoted
by u is imposed using the Neumann conditions. The control variable accounts for different
outflows corresponding to different inflows generated by the tuning the Reynolds number. These
conditions are mathematically written as
v (µ) = 0 on Γw , (18.4)
−η (∇v (µ)) n + p (µ) n = u (µ) on Γo . (18.5)
The goal of an OFCP is to optimize (minimize/maximize) certain flow quantities while con-
trolling some unknown quantities that can affect the fluid flow behavior. Here, the quantity to be
optimized is set in a tracking-type manner, guided by the intended application at hand in patient-
specific data assimilation. This objective is motivated by the need to integrate patient-specific
physiological information in everyday health-care decisions and can be achieved in several ways,
one of which is to minimize the discrepancy between real-life physiological data and the output
of the above-defined mathematical model. Thus, the cost functional (also referred to as objective
functional, objective function, or cost function) J is defined as
1 α
Z Z
2
J (v, u; µ) = |v (µ) − vo | dΩ + |u (µ) |2 dΓo , (18.6)
2 Ω 2 Γo
where vo can be the patient-specific blood flow velocity read from patient-specific medical im-
ages, such as four-dimensional MRIs provided by the hospital or the clinic. In other words, J
in equation (18.6) is the distance between the blood flow velocity approximated using computa-
tional hemodynamics, i.e., v, and the real-life patient-specific blood flow velocity, i.e., vo . The
second term on the right-hand side of (18.6) denotes the energy of the control variable u, with
α > 0 as penalization parameter. The adjoint-based Lagrangian approach [468, 250], which is
commonly used in PDE-constrained OFCPs, is adopted here.
18.2.2 ROM
We use the Galerkin FEM approximation to discretize the full-order model (FOM) and proper
orthogonal decomposition (POD) [281, 352, 34] to construct the reduced order spaces. An alter-
native to POD is the greedy algorithm [281], which uses residual-based error estimators. In the
literature, both techniques have been applied to OFCPs [413, 414, 562, 150] and patient-specific
computational cardiovascular modeling [32, 33, 568]. Technical details can be found in Rozza
et al. [281, 472] and Ballarin et al. [34].
four-dimensional flow MRIs as the target velocity. In these cases, it is realistic to assume that
the velocity profile will vary from cross-section to cross-section. The target velocity vo in both
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
cases can be imposed in a parabolic shape pointwise along the center lines, which are defined to
be the lines between two sections of lumen such that their minimal distance from the boundary
of the vessel is maximal, using the expression
r2
vo = vconst 1 − 2 tc . (18.7)
R
Here, for a vessel, tc is the pointwise tangent along the corresponding center line in the axial
direction, R is the maximum radius of the vessel corresponding to points on the center line, and
r is the distance between mesh nodes and the nearest point on the center line.
To set the value of vo , it is observed that the maximum coronary flow velocity ranges from
1 m/s to 4 m/s [208, 524]. Therefore, one can choose any value for the target velocity that has
its maximum in this range. These numerical examples are run with the magnitude of vo equal
to 0.35 m/s or 350 mm/s 4 m/s. Moreover, the constant kinematic viscosity is assumed
to be η = 3.6 mm2 /s and the velocity at the inlets is imposed using the following parametric
expression, with the Reynolds number Re as the parameter:
r2
ηµ
vin (µ) = − 1 − 2 nin , µ := Re ∈ D. (18.8)
Rin Rin
Here, Rin is the maximum radius of an inlet, nin denotes the outward normal to the inlet, and r
is the distance between mesh nodes and the center of an inlet.
Furthermore, the average error Eδ for δ = s, z, u is defined to be the norm of the difference
between the FEM solution and the ROM solution for a fixed parameter value in the corresponding
continuous space, and the relative error (Erel ) is the average error relative to the truth solution.
Moreover, the absolute average error (ET ) is the sum of squares of the average errors, and the
absolute relative error is ETrel = (ksh k2 +kuh k2ET +kzh k2 )1/2 . Lastly, the difference be-
S(Ω) U (Γo ) Z(Ω)
tween the Galerkin FEM and the POD-Galerkin approximations of the objective functional is
calculated using the equation
EJ = |J (xh ; µ) − J (xN ; µ) |. (18.9)
These numerical examples are solved in RBniCS [38, 281] and multiphenics [6].
In the domain at hand, the graft connection is realized between the right internal mammary
artery (RIMA) and the stenosed left anterior descending artery (LAD). The parameter is assumed
3
to take values in D = [70, 80]. Moreover, vo ∈ L2 (Ωa ) and the parameter is fixed at µ = 80
for the reduced order problem. To solve the truth problem, Galerkin FEM is used with the
stable Taylor–Hood pair of spaces, i.e., P2-P1 for velocity and pressure and P2 for control. This
achieves 4 × 105 degrees of freedom, i.e., N = 4 × 105 . One such simulation requires 1213.3 s.
For the ROM problem, Nmax = 6 eigenvectors are used, which gets the reduced order
dimensions to N = 79. These reduced bases include 48 bases altogether for velocity and adjoint
velocity, 24 bases altogether for pressure and adjoint pressure, 6 bases for control, and 1 basis
for the lifting [501] for the nonhomogeneous parametrized Dirichlet inflow condition.
Figure 18.1 (right) shows that 99.99% of the energy of the full-order solution manifold is cap-
tured in six eigenvectors. Thus POD-Galerkin approximations retain the accuracy and reliability
of a FEM solution, while using only 79 bases as compared to the 4 × 105 degrees of freedom of
the latter (see Figure 18.2 (left) and (right)). Moreover, the magnitude of the Dirichlet boundary
control is reported in Figure 18.1 (left).
Figure 18.3 (top left) illustrates the average errors Eδ for δ = v, p, u, w, q. Ev decreases from
101 to approximately 10−5 , Ep decreases from 106 to 10−2 , and Eu decreases almost to 10−6
18.2. CABGs 369
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 18.1. CABG: Boundary control magnitude mm2 /s2 (left) and eigenvalue reduction (right).
Figure 18.2. CABG: Simulation velocity v: comparison between FE approximation (left) and
ROM approximation (right).
as the number of eigenvectors used goes from one to six. Error reduction for adjoint variable
approximations follows a similar pattern as error reduction for state variables. Total error and
total relative error are reported in Figure 18.3 (top right). ET decreases from approximately
106 to 10−2 for n = 1, . . . , Nmax , and the maximum ET demonstrates the same behavior. A
reduction from 10−2 to 10−10 is achieved for both ETrel and the maximum ETrel in this case.
Figure 18.3 (bottom) shows the difference between J computed using FOM and ROM. It is
calculated using equation (18.9), and a reduction from approximately 106 to 10−2 is observed,
with similar behavior for the corresponding relative error when n goes from one to six.
370 Chapter 18. Reduced Order Methods for Hemodynamics Applications
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 18.6. The velocity in the channel, zoomed in to the wall attachment region.
Figure 18.7. The mitral valve between the left atrium and the left ventricle produces the Coandă effect.
372 Chapter 18. Reduced Order Methods for Hemodynamics Applications
The book [586] uses the term entrainment, which is crucial to understanding the Coandă
effect. Entrainment means that the jet draws fluid into itself from the sides. It is due to a com-
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
bination of momentum conservation and energy dissipation; see [586]. Entrainment can occur
in laminar flows as well as turbulent flows. At the interface between turbulent and nonturbulent
flow, the turbulent flow spreads to or entrains the flow nearby. The rate of entrainment is much
higher for turbulent flows than for laminar flows. It is not only wall attachment but also the
attachment of two jets to each other which can be caused by the Coandă effect.
Moreover, a low-pressure area develops where the fluid that has been entrained used to be. In
the presence of a wall, there is no further fluid to flow into this area and cancel out the pressure
gradient. As a consequence, the jet itself is drawn to the wall. The low pressure area is sometimes
also included in the definition, e.g., according to the Merriam-Webster dictionary,12 the Coandă
effect is “the tendency of a jet of fluid emerging from an orifice to follow an adjacent flat or
curved surface and to entrain fluid from the surroundings so that a region of lower pressure
develops.”
A precise definition of the Coandă effect does not exist, which is why the two quotations are
descriptive in nature. The effect occurs due to the interplay of a jet stream, entrainment, and
the low-pressure area. When a jet stream is injected into a stationary fluid, entrainment occurs
naturally. From the interaction with the stationary fluid, the stream broadens and becomes slower;
if the Reynolds number is sufficiently high then it also becomes turbulent. In most cases there
will be more surrounding fluid at the sides, which will be sucked closer to the jet stream. In the
case of a wall nearby, the jet stream sucks itself toward the wall in a sense.
In hemodynamics, the Coandă effect can be observed in the blood flow close to the heart.
Immediately after pumping and opening of the mitral valve, a jet stream of blood is produced
which is fast enough to cause the Coandă effect. Computer simulation must thus be sensitive
enough to capture these effects, such that accurate simulations are provided.
A common misconception about the Coandă effect is that related to fluid flow in air. For
example, a spoon placed below a water tap will guide the water along its curvature, or a water
bottle which is skewed only slightly above 90 degrees might produce water flow along the bottle.
These occurrences do not relate to the Coandă effect but to surface tension and other physical
effects. In principle it would be very difficult for water to produce entrainment of the surrounding
air, because the air molecules are much more mobile.
length L:
UL
Re = . (18.15)
ν
In the simulations presented in [457, 454], the Coandă effect is computed as steady states,
i.e., solutions where ∂u ∂t vanishes. To observe the Coandă effect, a certain minimum Reynolds
number Re must be given; e.g., in the simulations of [454] the Reynolds number must be at least
35 to observe the Coandă effect. This shows that the Coandă effect can be produced either by fast
jet flows, i.e., by a large velocity U in (18.15), or by media with a low viscosity ν. If the process
of entrainment is to be studied, then time-dependent simulations are needed. Such simulations
allow various technical improvements, such as novel wind energy generators [214].
There is a large variety of numerical discretization methods to solve the incompressible
Navier–Stokes equations, such as FEM, FVM, SEM, and discontinuous Galerkin; see [96, 97,
468, 330]. The variational form of (18.10)–(18.11) is given by
∂u
,v + (v, u · ∇u)Ω = (∇ · v, p)Ω − ν (∇v, ∇u)Ω + (v, f )Ω , (18.16)
∂t Ω
(q, ∇ · u)Ω = 0, (18.17)
with (·, ·)Ω the L2 scalar product over Ω. The discrete velocity-pressure solution pair (u, p) is
then found by solving (18.16)–(18.17) for all velocity-pressure test functions pairs (v, q) with
zero boundary conditions.
In order to evaluate the flow patterns for changing Reynolds number, parametric model reduc-
tion can be used to quickly evaluate the model over a range of Reynolds numbers. For example,
when the change in Reynolds number is effected by a change in the viscosity ν, then the model
readily allows for an affine decomposition, and a ROM can be applied, as outlined in [270] or
Chapter 1 of this book.
As discussed in [457], the mitral valve of the human heart produces a jet flow which shows
the Coandă effect. Since the echocardiographic assessment of MR provides only limited data
about the flow profile, computer simulations are needed to assist medical doctors. A comparison
between experiments in vitro and simulations is given in [607].
18.3.3 ROM
The ROM approach followed in [457] uses an intrusive reduced basis method (RBM) with an
affine expansion in the Reynolds number and geometry. Full-order solutions are sampled over
a coarse grid of parameters, and the POD computes the ROM projection space. In the offline-
online procedure, the parameter-independent parts of the PDE operator are precomputed and
low-order matrices are formed online according to the affine form for fast evaluations. Besides
the reconstruction of the flow, bifurcation detection is performed, which determines the onset of
the Coandă effect.
The starting point of the investigation can be seen in Figure 1 of [457], where echocar-
diographic images of the mitral value are shown. However, it is insufficient to rely only on
the echocardiographic images, and simulation results provide additional insight, as discussed in
[457]. An intrusive ROM strategy is employed, where the discrete velocity space and the dis-
crete pressure space are projected onto a low-order space. In particular, the variational form
374 Chapter 18. Reduced Order Methods for Hemodynamics Applications
∂uN
+ v, uN · ∇uN = ∇ · v, pN − ν ∇v, ∇uN (18.18)
,v Ω Ω Ω
+ (v, f )Ω ,
∂t Ω
q, ∇ · uN (18.19)
Ω
=0
and is solved for all velocity-pressure test function pairs (v, q) in the low-order space. The
work of [457] employs the Piola transformation to keep the divergence-free condition (18.19)
for varying geometry.
18.4 LVAD
LVADs consist of a mechanical pump providing full or partial mechanical circulatory support
to the left ventricle of the heart for patients with critical cardiac failure. The device aspirates
oxygen-rich blood from the left ventricle to push it into the aorta through a graft (Figure 18.8);
see, e.g., [7, 540, 345]. This section briefly discusses a data-driven ROM based on the proper
orthogonal decomposition with interpolation (PODI) method used in order to investigate the
aortic blood flow patterns induced by the outflow cannula of a continuous LVAD in a patient-
specific framework. For further details about this work, the reader is referred to [228, 234].
where Ck is a compliance, pp,k and pd,k are the proximal and the distal pressure, Rp,k and
Rd,k are the proximal and distal resistance, and Qk is the flow rate. The total resistance can be
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
evaluated as P
Ak
Rk = Rp,k + Rd,k = SVR k , (18.22)
Ak
where Ak is the cross-sectional area and SVR is the systemic vascular resistance. For each outlet
R
k, we assumed Rp,kk
= 0.056 [358]. The aortic compliance can be estimated as
SBP − DBP
C= , (18.23)
SV
with SBP and DBP the systolic and the diastolic blood pressure, respectively, and SV the
stroke volume. Then the compliance Ck related to each outlet k was evaluated as
Ak
Ck = C P . (18.24)
k Ak
We note that the values of SVR, SV , SAP , and DAP were measured by the right heart
catheterization (RHC) and echocardiography (ECHO) tests [228].
18.4.2 ROM
We have used a nonintrusive data-driven approach based on a PODI method, where the POD is
used to extract a reduced basis space from a collection of parametric full-order snapshots, and
an interpolation procedure based on radial basis functions (RBFs) is adopted to compute the
reduced coefficients. So the reduced approximation of a variable ϕ related to a certain value of
the parameter π can be obtained by the linear combination
k
X
ϕ(π) ≈ αj (π)φj , (18.25)
j=1
where k denotes the cardinality of a reduced basis, φj are the POD modes, and αj (π) are the
reduced coefficients.
Thanks to its nonintrusive nature, this procedure can be applied both for primal quanti-
ties, pressure p and velocity u, and for derived quantities such as wall shear stress (WSS). Re-
garding the technical implementation of the PODI method, we used the Python package called
EZyRB [158].
Table 18.1. L2 -norm relative errors for pressure p, WSS, and velocity components ux , uy , and
uz for P F = 3.45 l/min.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
p WSS ux uy uz
Eϕ 0.2% 4.1% 5% 7.8% 5.8%
Table 18.2. L2 -norm relative errors for pressure p, WSS, and velocity components ux , uy , and
uz for P F = 4.35 l/min.
p WSS ux uy uz
Eϕ 0.5% 7.2% 9.7% 13.5% 9.3%
where XF OM is the value of a particular field in the FOM model and XROM is the one that is
calculated using the ROM in Tables 18.1 and 18.2, respectively.
Figures 18.9, 18.10, and 18.11 display a qualitative comparison between FOM and ROM
for pressure, WSS, and velocity for P F = 3.45 l/min. The very similar spatial distribution
demonstrates that the ROM is able to provide a good reconstruction for all the variables.
Finally, we comment on the efficiency of our ROM approach. The CPU time of the FOM is
72000 s, while that of the ROM is 0.01 s. This corresponds to a speedup of about six orders of
magnitude.13
13 The authors want to thank the Medical Doctors of Sacco Hospital (Milano, Italy), San Camillo Hospital (Roma,
Italy), Sunnybrook Hospital (Toronto, Canada), and Houston Methodist Hospital (Houston, USA) for fruitful collabora-
tions and discussions.
18.4. LVAD 377
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 18.10. LVAD: comparison of the FOM/ROM WSS (Pa) at P F = 3.45 l/min.
Scientific Software
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
19.1.1 RBniCS
The RBniCS Project14 contains an implementation in FEniCS of several ROM techniques (and,
in particular, the certified reduced basis method and the proper orthogonal decomposition (POD)-
14 https://www.rbnicsproject.org.
379
380 Chapter 19. Scientific Software Development and Packages for ROMs in CFD
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Galerkin methods) for parametrized problems. It is ideally suited for an introductory course on
reduced basis methods and ROM, thanks to an object-oriented approach and an intuitive and
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
versatile Python interface. To this end, it has been employed in several doctoral courses on
reduced basis methods for computational mechanics.
RBniCS can also be used as a basis for more advanced projects to assess the capability of
ROMs in their existing FEniCS-based software, thanks to the availability of several reduced or-
der methods (such as reduced basis and POD) and algorithms (such as the successive constraint
method and the empirical interpolation method (EIM)) in the library. Several tutorials are pro-
vided.
19.1.2 ITHACA-FV
ITHACA-FV15 [549] (In real Time Highly Advanced Computational Applications for Finite Vol-
umes) is a C++ library based on the finite volume solver OpenFOAM [2]. It implements several
ROM techniques for parametrized problems. The library includes POD-Galerkin [550] methods
with hyperreduction functionalities [551], nonintrusive methods using PODI techniques [209],
and neural network approximations [639].
ITHACA-FV has been used to solve a variety of different problems, including heat trans-
fer problems [551], laminar fluid dynamics problems [548], turbulent fluid dynamics prob-
lems [284], multiphysics problems [220, 553], and geometrically parametrized problems [551].
ITHACA-FV can also be used as a basis for more advanced projects to assess the capability of
ROMs in their existing OpenFOAM-based software, thanks to the availability of several reduced
order methods and algorithms.
ITHACA-FV is designed for OpenFOAM 1812, 1906, 1912, 2006, 2012, 2106, and 2112,
but it can be easily adapted to other versions of OpenFOAM.
19.1.3 ITHACA-SEM
ITHACA-SEM (In real Time Highly Advanced Computational Applications for Spectral Ele-
ment Methods) is a C++ library based on the spectral element solver Nektar++.16 It implements
several ROM techniques for parametrized problems. ITHACA-SEM can also be used as a ba-
sis for more advanced projects to assess the capability of ROMs. ITHACA-SEM periodically
merges in the current master branch of Nektar++ and uses Eigen for matrix decompositions.
Nektar++ and the Eigen header files are part of the ITHACA-SEM package.
The open-source software ITHACA-SEM17 currently has the ability to generate POD-based
ROMs for two-dimensional incompressible Navier–Stokes equations with parametric variation
in geometry and/or kinematic viscosity. The parametric dependency on geometry parameters is
assumed to be affine, and the user can specify the affine form in a header file. After another
compilation of ITHACA-SEM, the affine form is then available. It can thus serve as a head
start for a developer seeking to work in ROMs with the SEM and also to a practitioner within
the boundaries mentioned. A few test cases are part of the Nektar++ unit test, and additional
examples are available.
ITHACA-SEM has been used in numerous publications; see [273, 276, 274, 272] for basic
usage, [270, 275] for localized ROMs, [454] for deflation methods in bifurcation problems, and
[277] for work on sparse polynomial interpolation.
15 Available at https://github.com/mathLab/ITHACA-FV.
16 www.nektar.info.
17 https://mathlab.sissa.it/ITHACA-SEM.
382 Chapter 19. Scientific Software Development and Packages for ROMs in CFD
19.1.4 ITHACA-DG
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
19.1.5 EZyRB
EZyRB20 [158] is a Python package that performs a data-driven model order reduction for
parametrized problems using recent approaches. We refer the reader to Chapter 9 for a deeper
discussion of these methodologies. Here we summarize by saying that such techniques provide
a parametric model able to approximate the solution of a generic (potentially complex and non-
linear) problem in real time. The reduced model is totally built upon the numerical data obtained
by the original (to be reduced) model, without requiring any knowledge of the equations that
describe this model, resulting in a framework that is well suited for industrial contexts due to
its natural ability to integrate with commercial software. Figure 19.2 shows an approximation
obtained using EZyRB on an parametric automotive benchmark.
EZyRB offers several methods for dimensionality reduction, such as POD and autoencoders.
It also provides a wide range of regressors for manifold approximation, such as linear interpo-
lators, Gaussian processes, and artificial neural networks (ANNs), made accessible through an
18 https://mathlab.sissa.it/ITHACA-DG.
19 https://github.com/HopeFOAM/HopeFOAM.
20 Available at https://github.com/mathLab/EZyRB.
19.1. Scientific Open-Source Software for Reduced Order Models 383
intuitive interface. In this way, the user can easily test different methods for the problem at hand
and make the best choice, also thanks to the already available error estimators. The package
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
functionalities are illustrated in different tutorials, which together with the online documentation
help users take the first step with this package. EZyRB has been employed in several contribu-
tions, such as parametric heat conduction in an isogeometric analysis setting [213], naval shape
optimization problems [157, 152], hydroacoustic analysis [209], and hemodynamics problems
[234].
19.1.6 PyDMD
PyDMD21 [159] is a Python package that uses DMD for a data-driven model simplification based
on spatiotemporal coherent structures.
DMD is a model reduction algorithm developed by Schmid [527] that we already carefully
treated in Chapter 9. DMD relies only on high-fidelity measurements such as experimental data
and numerical simulations. Its popularity is due to the equation-free nature of the algorithm,
which does not require any assumptions about the underlying system. See [353] for a compre-
hensive overview of the algorithm and its connections to the Koopman-operator analysis, along
with examples in computational fluid dynamics (CFD). In the last years many variants have
arisen, such as multiresolution DMD, compressed DMD, forward-backward DMD, and higher-
order DMD, to deal with noisy data, big datasets, or spurious data e.g.
In PyDMD we implemented the majority of the variants mentioned above with a user-friendly
interface. The package is actively maintained, providing detailed documentation and several
tutorials and examples to show the common use cases, like the one sketched in Figure 19.3.
PyDMD won first prize in the DSWeb 2019 contest Tutorial on Dynamical Systems Software
(Junior Faculty Category). It is also possible to easily extend the package with a new DMD
variant, following a developer tutorial.
Regarding its usage for scientific publications, we cite its employment for industrial design
in CFD applications [573] and for speeding up numeric simulations of dynamical systems [572,
155] as some recent examples.
21 Available at https://github.com/mathLab/PyDMD.
384 Chapter 19. Scientific Software Development and Packages for ROMs in CFD
19.1.7 ATHENA
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
ATHENA22 [496] is an open-source Python package for parameter space reduction. It imple-
ments advanced numerical analysis techniques such as active subspaces (ASs) [134], kernel-
based active subspaces (KASs) [493], and nonlinear level-set learning (NLL) [643]. Refer to
Chapter 16 for a deeper description of these methods. The package is intended as a tool for
regression and sensitivity analysis, and in general to enhance existing numerical simulations’
pipelines by tackling the curse of dimensionality. In Figure 19.4, we depict a summary of the
implemented techniques. The package provides a widespread application programming interface
(API) following the scikit-learn conventions.
Feature space
mapping
Active
subspace
Regression
Projection
Active variable
Figure 19.4. Sketch of the techniques implemented in ATHENA for parameter space reduction.
The software has been successfully used for optimization tasks [161, 157], for local active
subspaces (LASs) [495], for multifidelity Gaussian process approximations [494, 497, 575], for
coupling with DMD [573], and to reduce the memory consumption of convolutional neural net-
works [394].
19.1.8 PyGeM
PyGeM23 [570] is a Python package using free-form deformation (FFD), radial basis functions
(RBFs), and inverse distance weighting (IDW) to parametrize and morph complex geometries.
These methodologies, as well as their applications in parametric systems, are discussed in detail
in Chapter 17. The package is organized to work with a large variety of different file formats,
dealing with discrete geometries and computational grids (e.g., STL or VTK files) and analyti-
cal geometries (CAD files). Moreover, we highlight that the capacity of operating over discrete
geometries can be in principle extended to any file format, since PyGeM only needs the coor-
dinates of the points describing the geometry. Such points are moved according to the chosen
deformation, without altering the topology of the input geometry. In this way, PyGeM can be
coupled with any file format, requiring only a way to extract the initial points and to save the de-
formed ones.
Thanks to its properties, this package is very useful for problems in the parametric domain, al-
lowing one not only to morph the studied object—e.g., industrial artifacts like the one illustrated
in Figure 19.5—but also to act directly on the computational grid, so that it is very powerful when
22 Available at https://github.com/mathLab/ATHENA.
23 Available at https://github.com/mathLab/PyGeM.
19.2. Best Practices for Scientific Programming 385
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
coupled to ROM. Documentation and tutorials are available online, helping the user to take the
first steps for using the package or to contribute to it.
19.1.9 BladeX
BladeX24 [210] is a Python package for geometrical parametrization and bottom-up construction
of propeller blades. It allows the user to generate and deform a blade based on the radial distri-
bution of its parameters, such as pitch, rake, and skew, and the sectional foils’ parameters, such
as chord and camber. The package is ideally suited for parametric simulations on a large number
of blade deformations. It provides an automated procedure for computer-aided design (CAD)
generation, reducing the time and effort required for modeling. A graphical example showing
some CAD files obtained with this package is shown in Figure 19.6. The main scope of BladeX
is to deal with propeller blades; however it is flexible enough to be applied to further applications
with analogous geometrical structures, such as aircraft wings, turbomachinery, and wind turbine
blades.
development. The aim is to provide young researchers with an introductory perspective about
software engineering, helping them to structure and accelerate the programming phase in their
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
research. If interested, the reader can refer to [545, 203] for a deeper discussion of software
engineering.
• Use functions: Rather than thinking about the source code as a (long) list of instructions,
try to semantically split it into minimal (and possibly hierarchical) pieces of code. For any
of these snippets, separate the expected input and output in order to create the functions.
The main benefit is the reusability of the code and the reduced number of duplicated lines,
but such splitting also helps readability, since several technical (and probably difficult to
understand) parts can be encapsulated in functions.
• Keep it simple: As already mentioned, scientific code should be easy to understand in or-
der to enable future methodological enhancements. Simplicity of the code is an extremely
nice feature, which we suggest preserving by avoiding overcomplicated structures, even
with minimal loss of efficiency. Production software of course needs a different approach,
but simplicity is much more important for collaborative and dissemination purposes—in
the first steps of scientific software development.
• Source code should be self-informative: Rather than adding a comment beside every
line of code, we suggest using meaningful names for variables, constants, functions, and
classes, such that reading the source code allows the user in principle to understand the
algorithm. Comments are of course very welcome in any snippet of code, but we suggest
using them to explain functions. Commenting on all the lines generally makes the source
more cryptic.
• Follow the standards: Especially in team development, the use of code standards helps
make the code uniform and, consequently, improves its readability. Several conventions
19.2. Best Practices for Scientific Programming 387
are available for any language, and their use is widespread. This aspect becomes even
more important when you plan to release the source code to allow new developers to easily
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
contribute.
As a final remark, we add that in any case the fundamental point to keep in mind during de-
velopment is common sense, probably the most important of the aspects mentioned above. The
developer has to proceed in the coding always thinking about possible extensions and refactor-
izations, trying to select the design that maximizes the future use of the code.
19.2.2 Testing
Producing data to be published in scientific journals means that the reliability of the software
is of course another important feature. Things can become very frustrating when you discover
that your novel implementation returns awesome results for a specific problem but is not able to
deal with well-known benchmarks. It doesn’t matter if this issue is related to a coding bug or
a methodological error, its resolution can cost several hours (or even days): high-quality source
code definitely helps to detect the issue, but it still results in a demanding procedure. To overcome
these problems, we suggest planning a testing procedure starting at the first development stage.
Intuitively, the procedure is composed of a collection of additional snippets of code that check
that all the implemented functions return, for a specific input, the expected output. If these tests
have already been developed, it is easy to run them after any modifications of the main code to
be sure no errors have been introduced in the source code.
19.2.3 Documentation
Now that we have discussed the structure of the code and its testability, we focus in this last
subsection on the importance of having and maintaining informative documentation for the code.
This becomes a key feature when the source code is shared with other scientists and developers
to improve its understandability. Practically, documentation is a manual that illustrates all the
details of the source code, such as the expected input and output of any functions, as well as
a guide for code employment for final users. Despite its usefulness, it is quite common to see
projects that lack this resource. It must be said that of course preparing such a document is a time-
consuming procedure, but by using some frameworks we can at least shorten the process. We
suggest postponing the creation of documentation until clean source code has been implemented
with the corresponding tests. Documentation helps new developers and users to get oriented
with the source code. It explains the code at its current status and its possible applications.
Even if writing everything from scratch is an option, the most efficient choice for creating the
documentation is usually to extract it from the comments of the code, the docstring. With the
aim of illustrating how any function works, such comments are usually placed near the function
prototype: in this way, the documentation can be manipulated by simply editing the source
code, turning it into a less onerous activity. Several frameworks are available, depending on
the language used, for this operation, giving the developers the ability to make aesthetic changes.
Among others we cite Sphinx25 and Doxygen.26
As already mentioned, the documentation should not be limited to a summary of all the func-
tions implemented in the source code. To increase the software dissemination, and consequently
also the dissemination of the algorithm implemented in the code, the developers should pro-
vide different use cases with the most common applications, such that new users can easily and
quickly understand the typical usage.
25 https://www.sphinx-doc.org/.
26 https://www.doxygen.nl/.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Chapter 20
A Deep Learning
Approach to Improving
Reduced Order Models
20.1 Introduction
This chapter is devoted to describing different approaches for reduced order modeling (ROM)
obtained by combining innovative deep learning techniques, artificial neural networks (ANNs),
and digital twins.
Deep learning is a thriving field with many practical applications and active research topics
[237]. It is widely used to tackle and solve difficult tasks, such as visual object recognition
[350, 130], speech recognition [243], natural language processing (NLP) [633], autonomous
vehicles [295, 306], digitalization [266], and regression problems [51, 547]. ANNs have achieved
impressive results in dealing with such problems thanks to their deep structures and their ability
to approximate functions.
Some of the aforementioned fields represent great challenges in academic research but also
in the industrial sector. In order to face the complexity of such topics, neural network (NN)
architectures are forced to go deep, leading to a high number of parameters to be calibrated.
In this case, developing a reduction technique based on methods widely used in ROM, such
as proper orthogonal decomposition (POD) and active subspaces (ASs), is a good solution for
overcoming the typical constraints of embedded systems [394, 142]. On the other hand, ROM is
applied to digital twins in order to obtain a better product by providing more accurate information
about the real counterpart.
A classical result in this context, the universal approximation theorem for neural networks
[143, 293], states that sufficiently wide (shallow) NNs can approximate any (continuous) func-
tion. This opened the possibility of applying ANNs to a wide range of fields and in particular to
ROM and inverse problems involving nonlinear partial differential equations (PDEs).
Therefore, in this chapter, we start by outlining in Sections 20.2 and in 20.3 the most im-
portant features of ANNs and of the usual NN architectures. This part is then followed by
Subsection 20.4.1, where a detailed description of the POD-neural network (NN) approach is
provided: the use of ANNs to approximate the functional relationship between input parame-
ters and the output solution field in ROM. Subsection 20.4.2 is then devoted to presenting the
389
390 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
learning framework for solving forward and inverse problems involving nonlinear PDEs. In Sec-
tion 20.6 we describe a technique to develop a reduced version of an ANN in order to overcome
the constraint of memory storage usually connected with embedded systems. Finally, we report
in Section 20.7 the importance of digital twins in modeling and simulations, with a focus on the
role of ROMs in this context.
20.2 ANNs
We start by briefly introducing the main notions about ANNs [237, 348, 92].
The structure of ANNs is inspired by that of the human brain, mimicking the way that bio-
logical neurons signal to one another. They are characterized by an input layer, a certain number
of hidden layers,27 and an output layer, as shown in Figure 20.1. Usually, each neuron in a layer
is connected to each neuron of the next layer, and the strength of this connection is described by
a parameter called the weight. It is also important to highlight that, as explained in Section 20.3
and depicted in Figure 20.1, not only do ANN inputs flow in one direction (from the input to the
output layer), but backward connections and loops may exist.
x1
.. ŷ1
x2
.. .. . ..
. . . ŷnout
xnin
More technically speaking, let x ∈ Rnin be the input vector and L the total number of hidden
layers of the ANN. The output vector ŷ ∈ Rnout is obtained by applying an activation function
to the weighted sum of all the inputs arriving at the output layer. The activation function is used
to introduce nonlinearity in the network, and some common choices are the ReLu function, the
sigmoid, the logistic function, and the radial activation functions [348, 92].
In order to better understand how to get the general formula (20.2) for computing NN output,
we start by considering a simple ANN with a single output and one hidden layer. In this case the
final output is given by
nin
!
X
ŷ = σ wi xi + bi , (20.1)
i=1
27 A priori there is no right number of hidden layers to use: it depends on the field of application of your net and on
where σ is the activation function, w ∈ Rnin represents the weights of the net, and b the bias.28
Thus, if we consider L layers, the final output can be seen as a weighted sum of its inputs
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
followed by the activation function, where each input can be rewritten in the same way:
nL nL nL−1
! !!!
(L+1) (L) (L+1)
X X X (L)
(L−1)
ŷj = σ wji yi =σ wji σ wim ym
i=1 i=1 m=1
nL nL−1 nin
!!!!!!!
(L+1)
X X (L) X (1)
= ... = σ wji σ wim σ . . . σ wsk xk ,
i=1 m=1 k=1
(20.2)
where nl , l = 0, . . . , L, represents the number of neurons29 in layer l; y (l) ∈ Rnl the output of
(l)
layer l for l = 1, . . . , L; and W l = (wji )ji , j = 1, . . . , nl , i = 1, . . . , nl−1 the weight matrix
related to layer l. Note that the first number in any weight’s subscript matches the index of the
neuron in the next layer and the second number matches the index of the neuron in the previous
layer.
Two important concepts are connected to NNs:
With the term forward propagation [92] we refer to the process we have described before, where
starting from some inputs we compute the outputs using (20.2). Usually the weights and the
other parameters of the network are randomly initialized. Thus, obviously we could not expect
to obtain the right results at the end of the forward pass. In order to adjust these parameters to gain
better results, a loss function and the backpropagation algorithm [490, 92] have to be introduced
to take into account the error made in predicting the outputs using the net with respect to the true
outputs.
Let {xi , yi }ni=1 be a set of n observations and L = L(y, f (x)) a loss function, where the ANN
in consideration is summarized as a generic function ŷ = f (x; W ) of the inputs and weights. The
problem that needs to be solved is a minimization:
n1 Xn o
min L(yi , f (xi ; W )) , (20.3)
W n i=1
where L is chosen based on the problem to be solved [15, 237]. A common choice is the mean
squared error (MSE) function, defined by
n n
1X 1X i
L({yi }, {ŷi }) = Li = (y − ŷi )2 , (20.4)
n i=1 n i=1
where Li = Li (yi , ŷi ), yi represents the expected output, and ŷi is the prediction made by our
ANN.
Another possibility is the mean absolute error, which measures the mean absolute value of
the elementwise difference between inputs:
n n
1X 1X i
L({yi }, {ŷi })) = Li = y − ŷi . (20.5)
n i=1 n i=1
28 For simplicity the bias is set to zero in the following discussion.
29 Note that for l = 0 we are considering the input vector and thus n0 = nin .
392 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
In order to solve the minimization problem (20.3), the gradient method is usually employed
[92, 490]. In this context, this method is also referred to as backpropagation because the gap
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
recorded at a certain data point through the loss function L is propagated backward in the network
to obtain the update formulas of the coefficients by computing the gradient of L. Computing
an analytical expression for the gradient is straightforward, but numerically evaluating such an
expression can be computationally expensive. Therefore, the backpropagation algorithm does so
using a simple and inexpensive procedure that we will now explain.
Given an observation (x, y), the backpropagation algorithm initially takes a step forward
along the entire network, propagating information along all the layers (feedforward step or for-
ward propagation). After this step, the loss function L is computed using the predicted output
ŷ and the expected output y. It is important to highlight that in this first step the weights are
randomly initialized and then adjusted during the backward pass.
Figure 20.2. The forward pass on the left calculates y as a function f (x1 , x2 ) using the input
variables x1 and x2 . The right side of the figure shows the backward pass. Receiving dL/dy, the gradient
of the loss function with respect to y from above, the gradients of x1 and x2 on the loss function can be
calculated by applying the chain rule, as shown in the figure.
In order to find the minimum that satisfies problem (20.3), the gradient of L needs to be
computed. Since the loss function depends on the output of the net ŷ = f (x; W ) and thus on the
architecture of the ANN itself represented by f , the chain rule needs to be exploited, as depicted
(l)
in Figure 20.2. Therefore, for a single weight wjk , with l = 1, . . . , L + 1, the gradient is given
by
(l) (l)
dL dL dyj dzj
(l)
= (l) (l) (l)
. (20.6)
dwjk dyj dzj dwjk
nl−1
!
(l) (l) (l−1) (l)
X
yj =σ wjk yk = σ(zj ), (20.7)
k=1
(20.6) becomes
dL dL (l) (l−1)
(l)
= (l)
σ 0 (zj )yk , (20.8)
dwjk dyj
where dL
(l) is usually defined as the local gradient δjl . The computation of the gradients then
dyj
20.3. NN Topologies 393
allows us to optimize the model’s parameters by adjusting the network’s weights as follows
[490, 92]:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(m),t (m),t−1 dL
wki = wki − (m) , (20.9)
dwki
where is the learning rate and is appropriately chosen according to the problem at hand; t
represents the training epoch under consideration, i.e., a time interval during which all training
examples are employed to update the network parameters.
The process described up to now is used in the training phase, where we take part of the
complete data set (usually 80%) to fit the parameters of the model. The remaining part is used in
the testing phase to evaluate the final performance of the network.
20.3 NN Topologies
We provide in this section an overview of the usual NN topologies [348, 370, 92], i.e., the con-
nections between neurons of different layers that define different types of ANN architectures.
x1
.. ŷ1
x2
.. .. . ..
. . . ŷnout
xnin
x1
.. ŷ1
x2
.. .. . ..
. . . ŷnout
xnin
Hence, these algorithms are commonly used for ordinal or temporal problems, such as language
translation, NLP, and speech recognition.
input image
output feature maps
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
multiplication (between the two matrices), and adding the multiplication outputs to get the
final integer, which forms a single element of the output matrix. In CNN terminology, the
output matrix is called a feature map and its size is controlled by three parameters that
have to be decided before the convolution step is performed: the depth D, i.e., the number
of filters that will also determine the number of output images at the end of the convolution
step; the stride S, i.e., the number of pixels by which the filter matrix is slid over the input
matrix; and the zero-padding P , i.e. adding zeros around the border of a matrix in order to
apply the filter to bordering elements of our input image matrix.
Therefore, given as input a matrix x with size W1 × H1 × D1 and D F × F filters ω, the
convolutional operation will output a W2 × H2 × D2 matrix, where
(W1 − F + 2P ) (H1 − F + 2P )
W2 = + 1, H2 = + 1, D2 = D. (20.10)
S S
In particular, the convolutional output will be obtained by summing up the components of
the input image weighted by the filter components:
D1 X
X F X
F
z(ij) = ω0 + ω(abc) x(i+a−1,j+b−1,c) ,
c=1 a=1 b=1
(20.11)
y(ij) = σ(z(ij) ),
where ω0 represents the bias.
• It has a pooling (or subsampling) layer. Its role is to downsample the output of a convo-
lution layer along both of the spatial dimensions of height and width. As can be seen in
Figure 20.7, if layer l is a pooling layer, the m(l−1) feature maps of the previous layer are
pooled individually by defining a fixed sliding window F × F and considering the average
or maximum within it. In other words, each unit in one of the pooled feature maps in layer
l represents the average or the maximum within the fixed window of the corresponding
feature map in layer (l − 1). Hence, given a feature map with dimensions W1 × H1 × D1
and a fixed filter F × F moving with stride s, the final dimensions of the pooled feature
map W2 × H2 × D2 are determined by
(W1 − F ) (H1 − F )
W2 = + 1, H2 = + 1, D2 = D1 . (20.12)
S S
As described before, after the feature extraction part, an FNN performs the task of classification
based on the input from the last convolutional block.
396 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
layer (l − 1) layer l
20.3.4 Autoencoder
An autoencoder (AE) is used for classification, clustering, and feature compression [237, 288,
528, 367]. It mainly consists of three components: encoder, code, and decoder. Therefore, to
build an AE we need three things: an encoding method, a decoding method, and a loss function
to compare the output with the target.
More technically speaking, both the encoder and the decoder are fully connected FNNs com-
posed of a sequence of fully connected layers plus the bias followed by a nonlinear function.
As shown in Figure 20.8, first the input x passes through the encoder to produce the lower-
dimensional code z, i.e., the AE learns in this way a dimensionality reduction. Then, the decoder,
which is typically the mirror image of the encoder, maps the code to a reconstruction of the input
x̃. Note that the dimensionalities of the input and output need to be the same. To obtain an out-
put almost identical to the input, AEs are trained the same way as ANNs via backpropagation,
input output
encoder decoder
minimizing the reconstruction loss L(x̃, x) (e.g., cross-entropy loss or MSE) that quantifies the
quality of the reconstruction x̃.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
In this context, we use FNN to learn the map f : Rd → RNh , which approximates the reduced
basis degrees of freedom at a given parameter tuple Ξ:
In order to create a training set for FNN, samples from parameter space P are drawn and
split into two sets: the training set {Ξi }ni=1tr
and the validation set {Ξj }nj=1
val
. At each sample
Ξi , the high-fidelity solution field uh (Ξi ) ∈ Vh and its projection urb
h (Ξi ) ∈ V rb
h are computed.
Hence, a set of known input-output pairs {Ξi , ζrb (Ξi )} is generated for training. Similarly, a
known set of input-output pairs {Ξj , ζrb (Ξj )} is generated for validation. FNN is trained to
minimize the loss function L(ζrb (Ξi ), ζNN (Ξi )) over the training set using backpropagation (see
Section 20.1). Since backpropagation is an iterative approach, it is important to ensure that
the initial guess of weights and biases does not inadvertently affect the outcome at the end of
backpropagation. Also, it should be noted that a larger weight assigned to a particular synaptic
connection may compromise the robustness of the method. The accuracy of the network can be
tested by sampling the parameter space P and measuring the error, in a suitable norm, between
the high-fidelity solution field and the output computed by the POD-ANN method.
Training an ANNs is an expensive procedure due to the trial-and-error approach used to
adjust its hyperparameters. During training, the following steps occur:
• Data for training the ANN is prepared by projecting high-fidelity solutions onto the re-
duced basis space.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
The higher offline cost, due to the training procedure, is disadvantageous for ANN-based meth-
ods as compared to the projection-based methods. One solution to this problem is to extract more
information from the high-fidelity solution. This can be done by modifying the loss function: loss
functions utilizing residuals can produce more accurate predictions [119]. Higher offline cost is
often associated with the complexity of the problem, which the ANN is required to learn. An
alternative to reduce this complexity is to assign an input-output map and use the ANN to de-
termine the parameters of this map. For example, fitting a linear regression line to approximate
the input-output map will require an ANN without any hidden layers. However, for real-world
applications, the underlying input-output relationship is complex, so such simple models may
have very limited applications.
Another important component of the training of the ANN is to avoid “overfitting.” Due to
overfitting, a trained ANN is accurate for predictions corresponding to the parameter tuples used
for training but fails to accurately predict output for parameters outside the training set. To avoid
overfitting, regularization techniques such as L2 regularization and early stopping criteria can be
used.
It is also pertinent to mention that the sampling strategy for generating training data for
an ANN should be chosen carefully. A good sampling strategy aims to take samples that are
distributed evenly over the parameter space. It needs to be noted that the sampling strategy used
for POD might not be the same as the sampling strategy used to create a training set. This is
mainly due to two reasons:
1. While POD aims to capture only dominant modes, an ANN tries to capture all patterns.
2. Often, the training set for an ANN is larger than the one used for POD.
During the online phase, the ANN only requires feedforward operation at a given parameter
tuple Ξ using (20.2). In the case of projection-based methods, the online time is higher due to the
time taken to assemble system equations if the online phase and offline phase are not completely
decoupled. Higher online time can also occur in coupled systems, where computing one solution
field requires solving more than one equation, or in the presence of too many geometric param-
eters [536]. The POD-ANN approach is a particular case of nonintrusive data-driven methods
(Chapter 9); i.e., system matrices corresponding to a discretized form of governing equations
are not needed. The nonintrusive nature of the method simplifies the interface between different
software libraries, such as the one used to compute high-fidelity solutions and the one used for an
ANN. It also needs to be noted that adding information about the governing equation certainly
helps to increase accuracy [119].
We now provide offline-online decomposition for the POD-ANN approach.
• Offline phase:
5. Repeat steps 2 to 4 for a different number of hidden layers, width of each hidden
layer, and number of sampled parameter tuples n to find the optimal configuration of
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
the ANN.
• Online phase:
1. For any given parameter tuple Ξ ∈ P, perform a forward pass using (20.2).
where ũ(x, t) is the shifted field and Ω and T are the spatial and temporal domains where the field
is defined. A visual example computed on a simple advection problem is depicted in Figure 20.9,
showing the benefits of the shifting paradigm in terms of the dimensions of the reduced space.
Machine learning extends this framework in order to generalize the method. The choice of
the translation to apply to all the snapshots is not always trivial and requires complete knowledge
of the problem at hand. A possible alternative is learning this shifting using only the collected
snapshots in an NN framework. This is the main idea of the neural network shifted proper
orthogonal decomposition (NNsPOD) [435, 436] algorithm, which is a fully nonintrusive, purely
data-driven machine learning method that seeks an optimal mapping of the various snapshots in
the low-rank linear subspace to a reference configuration via an automatic detection that does not
depend on the physical properties of the model. In this framework, the shifting operator becomes
where the translation of the field is not limited by a fixed value linearly depending on the time as
in (20.16) but depends on spatial and temporal coordinates. The shifting quantity b : Ω × T → Ω
is approximated using an FNN. The learning process is built upon the set of snapshots—the
discrete field—defined as
{V 3 u(ti )}N Ns
i=1 = {u(V, ti )}i=1 ,
s
(20.18)
400 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 20.9. Example of POD enhancement thanks to the shifting of the snapshots. A Gaussian
impulse advancing in time is a simple advection problem which shows the limitations of POD in these
contexts: the singular values obtained by the decomposition of the snapshots are not able to individuate
any reduced dimension (top). By shifting the snapshots, POD is able to detect the dimension needed to
linearly represent the solution manifold, as highlighted by the single nonzero singular value (bottom).
where ti are the Ns time steps where the corresponding snapshots have been collected and V
is the discretized spatial domain. The main idea is to transform these snapshots by minimizing
the Euclidean distance between them and a chosen reference configuration uref —e.g., the initial
condition, one of the snapshots. A graphical example showing a two-dimensional field in an
advection problem is presented in Figure 20.10: one snapshot translates toward the reference
configuration during the training of the network, demonstrating the ability to automatically detect
the optimal transformation. The loss function imposed during the learning process is therefore
Ns
1 X
J = kuref (x, ti ) − u(x − b(x, ti ), ti )k2 , x ∈ V. (20.19)
Ns i=1
In a discrete context, it is not guaranteed that x − b(x, t) still belongs to the discrete space
V, requiring an additional step to compute the mean distance between the snapshots and the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
reference configuration. To evaluate any shifted snapshots in the correct space, an additional NN
is employed to generate a continuous approximation of the reference configuration, such that it
is possible to evaluate it in the shifted coordinates. This methodology then relies on two different
networks:
• the ShiftNet computes the shifting quantity b(x, t) to obtain the new spatial coordinates of
the shifted snapshots;
• the InterpNet evaluates the reference configuration in the coordinates of the shifted snap-
shots to compute the distance.
The NNsPOD method results in a data-driven enhancement of the POD method thanks to the
combined employment of a complex neural architecture, overcoming several well-known issues
in the ROM community caused by advection.
Figure 20.11. Example of a PINN solution for a Poisson problem with parametric forcing term.
The two rows show the field of interest for two different parametric configurations, while in the columns the
analytical solution, the PINN approximation, and the absolute error are sketched (from left to right).
network, and their relation to the unknown field is learned by providing the equations describing
the temporal and parametric dependencies. A simple graphical example is shown in Figure 20.11,
comparing the PINN solution of a Poisson problem with a parametric forcing term against the
analytical solution.
In this case, the higher complexity due to the increased number of dimensions translates into
a higher computational cost to train the model. After the training, the network becomes a real-
time model able to approximate the output of interest for different time steps and parameters.
Even if it cannot be considered a reduced order method, it has an algorithmic structure similar
to the methods proposed in the previous chapters of this book: the learning step requires almost
all the computational cost, as typically happens for the offline phase. The network evaluation,
which means the field computation, is instead in real time, as in the large variety of ROMs.
The computational resources needed to train the NN make its application to parametric linear
problems quite inefficient compared to the consolidated techniques, but its nonlinear nature could
offer a valid approach to addressing more complex problems in the parametric context.
In this section, we propose strategies for the development of a reduced version of an ANN
and in particular of a CNN [394]. We will then present the results obtained using this reduced
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
• Network pruning [263, 377]: The main goal of these approaches is to prune redundant
channels in the weight matrices of a trained net and hence accelerate and compress the
whole model. There exist different sparsity constraints to select which channels have to be
pruned: the sum of the absolute value of weights or the average percentage of zeros. De-
spite the fact that these methods achieve parameter sparsity in the model, they are usually
iterative procedures that require a fine-tuning of the parameters, leading to a computation-
ally expensive method for deep NNs and for some applications.
• Parameter quantization [140]: This technique consists of substituting the floating points
with integers inside the network. The created network requires less memory, and, on cer-
tain hardware, the calculations are faster. On the other hand, this can lead to a significant
loss in accuracy, because there is no guarantee that approximating the parameters by sac-
rificing precision for a compressed representation will not compromise the final results.
As described, most of the existing methods do not change the network architectures but only
delete and change model parameters directly. Thus, starting from the idea explored in [142], we
propose a general framework for constructing a reduced version of an (already trained) ANN,
and in particular of a CNN, based on techniques widely adopted in the model order reduction
community, such as POD, described in Section 1.1, and the AS method, presented in Chapter 16.
The reduced network, described in Figure 20.12 and in [394], is constructed by retaining a certain
number of layers of the original ANN and replacing the remaining ones with an input-output
mapping. We are, thus, splitting the net into two parts connected by the reduced method, which
helps reduce the typically large dimensions of the intermediate layers by keeping only the most
important information. In this way, we are performing a smart selection of the main parameters
of the network, allowing us to reduce the required resources and the computing time to infer the
model.
31 This is the only assumption we are making regarding the original network.
404 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Figure 20.12. Graphical representation of the reduction method proposed for an ANN.
real outputs of the AN N . Given the index l, denoting the cutoff layer, the full net is split in two
different parts:
• the premodel, made of the first l layers, and
• the postmodel, formed by the last L − l layers.
It is important to specify that the cutoff layer l is chosen empirically as a trade-off between the
final accuracy and the compression ratio. This value plays a central role in the final outcome
since it is the parameter that controls how many layers of the original network are kept in the
reduced architecture and thus how much information of the original network we are discarding.
Describing the network as a composition of functions AN N ≡ fL ◦ fL−1 ◦ . . . ◦ f1 , the pre-
and postmodels can be formally described as
where fj : Rnj−1 → Rnj for j = 1, . . . , L are the functions representing the different layers
of the network—e.g., convolutional, fully connected, batch-normalization, ReLU, or pooling
layers. Therefore, we can rewrite the original model as a composition of the postmodel with the
premodel:
AN N (x0 ) = AN N lpost (AN N lpre (x0 )) (20.21)
for any 1 ≤ l < L and for any x0 ∈ D0 . The network is thus reduced by copying the premodel
from the original net and approximating the postmodel using the method we will describe.
The output of the premodel x(l) = AN N lpre (x0 ) usually lies in a high-dimensional space;
thus we aim to project x(l) onto a low-dimensional space of dimension r using the aforemen-
tioned reduction techniques:
• ASs: As described in Chapter 16 and in [142], we consider a function gl defined as the
composition of the loss function with the postmodel,
in order to construct the AS and thus extract the most important directions. In this way, we
will determine the matrix of the project W1 , which will reduce the premodel output xl in
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
• POD: As discussed in Section 1.1, the singular value decomposition (SVD) is used to
compute the projection matrix U and thus the reduced solution
After we obtain the reduced solution z, the last step of the method deals with the construction
of the mapping between z and the final output of the network y. Two different techniques have
been used to find that map:
• The polynomial chaos expansion (PCE) proposed by Wiener in [616] states that a real-
valued random variable can be approximated by a finite sum of orthogonal polynomials
weighted by unknown deterministic coefficients cα [308]. Therefore, in our context the
final output of the network y ∈ RnL can be approximated as
P
X
ŷ ≈ cα φα (z), |α| = α1 + . . . + αr , (20.25)
|α|=0
where φα (z) are the multivariate polynomial functions chosen based on the probability
density function ρ associated with z. The number of unknown coefficients in this summa-
tion is given by P + 1 = (p+r)!
p!r! [224], where p is the degree of the polynomial we are
considering in the r-dimensional space. Based on the work of Askey and Wilson [21],
we can provide the orthogonal polynomials for different distributions. One of the possible
choices is the Gaussian distribution with the related Hermite polynomials.
Hence, to determine the PCE, we need to estimate the PCE coefficients cα for |α| =
0, . . . , P , which can be carried out in different ways [564]: following a projection method
based on the orthogonality of the polynomials or following a regression method, the latter
of which we will describe here.
In order to determine the coefficients cα , the following minimization problem has to be
solved:
2
Ntrain p
1 X j
X
j
min y − cα φα (z ) . (20.26)
cα Ntrain
j=1 |α|=0
where if the matrix ΦT Φ is ill conditioned, as may happen, the SVD method should be
employed.
406 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
• With the FNN described in Section 20.3, the input matches the output of the reduction
layer z ∈ Rr , and thus by using (20.2) we obtain the fact that the final output ŷ of the
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
where nout corresponds to the number of categories that compose the dataset being exam-
ined and σ is the Softplus function:
1
Softplus(x) = log(1 + exp(βx)). (20.30)
β
Once the reduced network has been initialized, we need to train it using the input-output pairs
(xj0 , yj ) for j = 1, . . . , Ntrain and with x(0),j ∈ D0 ∀ j. The technique chosen to perform the
training process is that of knowledge distillation [287], as is done in [142]. This is an indirect
approach where a small student model, in our case AN N red , is trained to mimic the behavior
of a large pretrained model, called the teacher model, i.e., our full AN N , and thus to gain a
comparable or even superior performance.
The output of the last layer in a deep NN y is referred to in this framework as the vector of
logits. Each network input is then characterized by the probability that it belongs to the ith class,
which is defined by the softmax function
exp(yi /T )
pi = Pnclass , (20.31)
j=0 exp(yj /T )
where T , the temperature factor, is introduced to control the importance of each target, as ex-
plained in [287]. Note that if T → ∞, all classes have the same probability, whereas if T → 0
the targets pi become one-hot labels.
The final loss in this method is made of two different losses:
• The distillation loss LD matches the logits between the teacher and the student model.
Following [238, 287], LD is represented by the Kullback–Leibler (KL) divergence loss
[343]:
LD (p(yt , T ), p(ys , T )) = LKL ((p(yt , T ), p(ys , T )), (20.32)
where yt and ys indicate the logits of the teacher and student networks, respectively,
whereas the KL loss LKL is defined by
X pj (yt,j , T )
LKL ((p(ys , T ), p(yt , T )) = T 2 pj (yt,j , T ) log . (20.33)
j
pj (ys,j , T )
• The student loss LS is defined as the cross-entropy loss between the ground truth label and
the logits of the student network [238]:
X
LS (y, p(ys , T )) = LCE (ŷ, p(ys , T )) = −ŷi log(pi (ys,i , T )), (20.34)
i
where ŷ is a ground truth vector, characterized by having only the component correspond-
ing to the ground truth label on the training sample set to one, while the others are set to
zero.
32 Note that in this case the number of hidden layers is set to one since, as discussed in Subsection 20.6.3, we notice
that one hidden layer is enough to gain a good level of accuracy (see, e.g., Section 20.1).
20.6. A Reduced Approach for CNNs 407
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Both losses, (20.32) and (20.34), are characterized by the use of the logits of the student model
but with different temperatures: T = τ > 1 in the distillation loss and T = 1 in the student loss.
Therefore, the loss is obtained as a weighted sum of the aforementioned losses:
where λ is the regularization parameter, x0 ∈ D0 , and W are the parameters of the student model.
needed. For this reason, we kept the same value in the two different cases we are considering, i.e., for the two sets of
data, to have a fair comparison.
34 As explained in [142] and in the corresponding implementation, the indices 5, 6, and 7 correspond to the indices of
the convolutional layers in a list where only convolutional and linear layers are taken into consideration as possible cutoff
layers. Therefore, taking into account the whole net with all the different layers, these correspond to 11, 13, and 16.
408 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
Table 20.1. Results obtained for the reduced net POD+FNN (7) trained on CIFAR10 with differ-
ent structures for the FNN.
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Hidden layers
1 2 3 4
Epoch 0 81.39% 67.92% 75.52% 81.57%
10 Epoch 10 87.89% 87.59% 87.46% 87.26%
Hidden neurons
high-dimensional features, the output of the premodel, is set to 50 both for AS and for POD in
analogy with the structural analysis considerations presented in [142]. We also remark that to
compute the AS we employed the frequent direction method [225] implemented inside ATHENA.
Table 20.1 then reports different results obtained by training the reduced net for 10 epochs
using different FNN architectures,35 i.e., a different number of hidden layers and also hidden
neurons, which are kept constant in each hidden layer of the net. The comparison is made in
terms of memory allocation for the FNN and accuracy of the corresponding reduced net at epoch
0, i.e., after its initialization, and at epoch 10. Note that increasing the depth of the network
does not lead to a remarkable gain in accuracy. Therefore, the structures chosen for testing the
reduction method using an FNN as input-output map are the following:
• CIFAR10: 50 input neurons, 10 output neurons, and 1 hidden layer with 20 hidden neu-
rons;
• custom dataset: 50 input neurons, 4 output neurons, and 1 hidden layer with 10 hidden
neurons.
Tables 20.2 and 20.3 contain the results obtained by initializing the reduced nets and training
them for 10 epochs on the corresponding datasets. The data in the tables compares different
reduced networks, i.e., AS+PCE, AS+FNN, POD+FNN, using three different cutoff layers, 5,
6, and 7, in terms of accuracy (before and after the final training), memory storage, and time
needed for the initialization and training processes. We also highlight that the results we present
are consistent between the two different sets of data, as can be seen in Tables 20.2 and 20.3.
Regarding the problem of memory constraints in an embedded system, note that the stor-
age required for the proposed nets is less than that of the original VGG-16 in both cases (see
Tables 20.2 and 20.3). The checkpoint file for the reduced versions of VGG-16 now requires
less than 10 Mb in all cases, whereas before it occupied 56.14 Mb. Hence, if we compute the
35 Each of the FNNs used was trained for 500 epochs during its initialization process and thus before the retraining
compress_size
compression_ratio = 1 − , (20.36)
uncompress_size
we obtain that the savings in memory is around 90% in all cases. It can also be highlighted
that using an FNN instead of the PCE brings a gain in saving memory space of two orders of
magnitude: 10−4 against 10−2 .
From the point of view of the final accuracy, it can be observed that increasing the cutoff layer
index l leads to nets with higher accuracy since we are retaining more information (i.e., features)
from the original VGG-16, but on the other hand, there is a smaller compression ratio. Hence, the
index l should be chosen carefully, taking into account the field of application and also making
a trade-off between the final accuracy and the reduction. Despite this, for all the reduced CNNs,
the final level of accuracy is similar to (in most cases also greater than) the original one.
Using the reduced net POD+FNN leads to another gain in terms of time, since, as shown in
Table 20.2, in this case, additional training with the whole dataset is not required by this net.
From the table, note that after the initialization, i.e., at epoch 0, we already have a performing
network and thus there is a savings of time of the order of 5 h. This is true and also consistent with
the custom dataset (see Table 20.3), where for the three choices of l we have that the POD+FNN
net has an accuracy greater than that of the original VGG-16 even after the initialization process.
410 Chapter 20. A Deep Learning Approach to Improving Reduced Order Models
• The system integration digital twin can be used to simulate complex systems and test
settings on-premise. There is no direct data exchange between the virtual and the real
asset.
• The embedded digital twin receives data from the physical asset. It can be used for appli-
cation monitoring and virtual sensing.
• The companion digital twin exchanges data in a bidirectional way with its real counterpart.
The most advanced applications, which include dynamic optimization, optimal control,
and advanced diagnostics, belong to this category of digital twins.
Digital twin and physical assets, when bonded together, are usually referred to as cyber-physical
systems.
Hence, digital twins are opening new scenarios and possibilities by integrating virtual and
real assets, but at the same time, they propose new challenges. For example, there is a clear
need to find an effective way to exchange, collect, and analyze data coming from the real asset
and integrate it into the models which make up the digital twins. These challenges can be faced
by combining, in a unified framework, different techniques and technologies recently developed
in numerical methods, IoT, artificial intelligence, Big Data, and high-performance computing
(HPC). Some practical examples of applications of digital twin can be found in [266].
Intrinsic Complexity
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Modern systems are complex, and their complexity does not reside in the number of components
but rather in the number of physical domains which are involved. For example, cars up to 1980
were almost exclusively mechanical products, while a modern vehicle includes electric compo-
nents, electronic boards, air conditioning, and software. Digital twin must take into account all
the diverse domains to get a complete model so digital twins are intrinsically multiphysical.
Long-Time Consistency
Models implemented in digital twins must remain consistent with the corresponding physical as-
sets for their entire lifetimes, especially when the real system tends to degrade over time because
of wear, rusting, change of working conditions, and fatigue phenomena. To overcome this prob-
lem, models should be able to adapt their behavior by learning from the data generated by the
associated physical counterpart (see [321] for a probabilistic graphical model approach). Data
assimilation techniques are therefore mandatory in order to guarantee a tight consistency during
the life of the digital twin.
choice when one has to create accurate digital twins to deal with complex dynamics. ROMs
consist of two distinct phases:
Downloaded 12/19/23 to 222.29.66.238 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
• In the offline phase, high-fidelity simulations are performed for some given combinations
of input parameters, the dominant dynamics (or modes) are extracted, and the other infor-
mation needed for building the reduced order method is evaluated.
• In the online phase, the magnitudes of interest are constructed quickly by a linear super-
imposition of the modes multiplied by proper weights.
Two distinct ROM approaches exist: nonintrusive and intrusive. The latter relies on the
projection of the set of governing equations onto a reduced space individuated by the modes
computed during the offline phase. These methods are accurate, with a solid physical back-
ground, but they suffer from possible instabilities. On the other hand, the nonintrusive methods
use machine learning techniques to create mapping functions during the offline phase between
the weights which multiply the modes and the input parameters. Both of the approaches are
able to reconstruct complex phenomena in near real time, providing significant insight into the
phenomena.
ROMs in digital twin are relevant in two distinct phases. The most trivial application is to
apply digital twin to deployable digital twins to provide more accurate information about the real
counterpart. The second is related to the product development phase, when ROMs can be used to
obtain a better product by performing processes that are based on multiquery, e.g., optimization,
uncertainty quantification, and controls. In Figure 20.14 is an example of a model used for a
process of structural optimization, while Figure 20.15 shows the corresponding displacement
field for a given loading condition.
Figure 20.14. Example of a structural model in the naval engineering field. On the left is the
entire hull, while on the right is a sectional view.
Figure 20.15. A modern cruise ship example of displacement field for the hogging loading condition.
20.7. A Future Perspective on ROMs for Digital Twin 413
Digital twin is a relatively new technology that started attracting the attention of the industrial
world in the last few years, and many challenges are still open. The first challenge is to devise
a robust and general methodological approach for building digital twins; the first step in this
direction is some recent works by Kapteyn et al. [320], who propose an approach for enabling
data-driven digital twins based on libraries of components based on ROMs. Other relevant as-
pects which are the subject of research are to create big models, coupling data and equation-based
models, which would help to overcome problems that are related to Big Data in applications of
practical interest such as lack of data and data veracity, and it would allow uncertainties to be
quantified.
Some practical aspects are also under investigation, in particular the ones regarding the de-
velopment of general tools which allow an effective way to exchange information among models
generated with different software.
Thanks to the integration of data, ROMs, and models, the concept of digital twin can be
further expanded, and the frontiers of digital twin are related to the creation of complex virtual
copies of more complex systems [416]. These extensions would help to implement effective
strategies for sustainability and a circular economy and would be beneficial for mobility, logis-
tics, the economy, and the environment. Smart factories are one of the main goals of Industry
4.0, and they consist of a digital twin that is strongly integrated with the corresponding physical
plant. This would allow us to constantly monitor and optimize production processes, instanta-
neous power consumption, logistics, and product quality. The next wave of digital twins will
include the augmented digital twin, which interacts not only with its real asset but also with its
surroundings and other digital twins. The augmented digital twin will be used to create smart
infrastructure (harbors, freeways, etc.) and entire cities or ecosystems, e.g., to optimize mobility,
improve efficiency, and monitor pollution. Finally, the most ambitious application of the digital
twin technology is the one related to human digital twin to enable full life cycle health manage-
ment of a human being; given the complexity of biological systems, ROMs will be one of the
key enablers of the application of this new type of digital twins, even though a lot of work still
has to be done in terms of pure modeling of biological processes whose end is to understand
fundamental phenomena [147, 611].
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Bibliography
[1] ITHACA-DG—In real Time Highly Advanced Computational Applications for Discontinuous
Galerkin-ROMs for HopeFOAM. https://github.com/mathLab/ITHACA-DG. (Cited on p. 241)
[2] OpenFOAM website. https://openfoam.com/. (Cited on p. 381)
[3] STAR CCM+ Users Manual. http://www.cd-adapco.com/products/star-ccm/
documentation. (Cited on p. 143)
[4] STEP-file, ISO 10303-21. https://www.loc.gov/preservation/digital/formats/fdd/
fdd000448.shtml. (Cited on p. 357)
[5] STL (STereoLithography) File Format Family. https://www.loc.gov/preservation/digital/
formats/fdd/fdd000505.shtml. (Cited on p. 356)
[6] multiphenics—easy prototyping of multiphysics problems in FEniCS. http://mathlab.sissa.
it/multiphenics, 2016. (Cited on pp. 318, 368)
[7] K. A ARONSON , M. S LAUGHTER , L. M ILLER , E. M C G EE , W. C OTTS , M. ACKER , M. J ES -
SUP, I. G REGORIC , P. L OYALKA , O. F RAZIER , V. J EEVANANDAM , A. A NDERSON , R. KORMOS ,
J. T EUTEBERG , W. L EVY, D. NAFTEL , R. B ITTMAN , F. PAGANI , D. H ATHAWAY, AND S. B OYCE,
HeartWare Ventricular Assist Device (HVAD) Bridge to Transplant ADVANCE Trial Investigators.
Use of an intrapericardial, continuous-flow, centrifugal pump in patients awaiting heart transplan-
tation, Circulation, 125 (2012), pp. 3191–3200, https://doi.org/10.1161/CIRCULATIONAHA.
111.058412. (Cited on p. 374)
[8] I. A BBOTT AND A. VON D OENHOFF, Theory of Wing Sections, Including a Summary of Airfoil
Data, Dover Books on Aeronautical Engineering Series, Dover Publications, 1959. (Cited on p. 355)
[9] R. A BGRALL , D. A MSALLEM , AND R. C RISOVAN, Robust model reduction by L1 -norm mini-
mization and approximation via dictionaries: Application to nonlinear hyperbolic problems, Ad-
vanced Modeling and Simulation in Engineering Sciences, 3 (2016), https://doi.org/10.1186/
s40323-015-0055-3. (Cited on p. 298)
[10] S. R. A HMED , G. R AMM , AND G. FALTIN, Some salient features of the time-averaged ground ve-
hicle wake, SAE Transactions (1984), pp. 473–503, https://doi.org/10.4271/840300. (Cited
on p. 361)
[11] A. A JIT, K. ACHARYA , AND A. S AMANTA, A review of convolutional neural networks, in 2020 In-
ternational Conference on Emerging Trends in Information Technology and Engineering (ic-ETITE),
IEEE, 2020, pp. 1–5, https://doi.org/10.1109/ic-ETITE47903.2020.049. (Cited on p. 407)
[12] S. A LI, Stabilized Reduced Basis Methods for the Approximation of Parametrized Viscous Flows,
PhD thesis, SISSA, 2018. (Cited on p. 128)
[13] S. A LI , F. BALLARIN , AND G. ROZZA, Stabilized reduced basis methods for parametrized steady
Stokes and Navier–Stokes equations, Computers & Mathematics with Applications, 80 (2020),
pp. 2399–2416, https://doi.org/10.1016/j.camwa.2020.03.019. (Cited on pp. 129, 131)
415
416 Bibliography
[16] A. A MBROSETTI AND A. M ALCHIODI, Nonlinear Analysis and Semilinear Elliptic Problems, no. v.
10 in Cambridge Studies in Advanced Mathematics, Cambridge University Press, 2007. (Cited on
p. 101)
[17] A. A MBROSETTI AND G. P RODI, A Primer of Nonlinear Analysis, Cambridge Studies in Advanced
Mathematics, Cambridge University Press, 1995. (Cited on p. 101)
[18] D. A MSALLEM , M. J. Z AHR , AND C. FARHAT, Nonlinear model order reduction based on lo-
cal reduced-order bases, International Journal for Numerical Methods in Engineering, 92 (2012),
pp. 891–916, https://doi.org/10.1002/nme.4371. (Cited on p. 298)
[21] R. A SKEY AND J. A. W ILSON, Some basic hypergeometric orthogonal polynomials that generalize
Jacobi polynomials, Memoirs of the American Mathematical Society, 54 (1985), https://doi.
org/10.1090/memo/0319. (Cited on p. 405)
[22] M. A STORINO , F. C HOULY, AND M. A. F ERNÁNDEZ, Robin based semi-implicit coupling in fluid-
structure interaction: Stability analysis and numerics, SIAM Journal on Scientific Computing, 31
(2009), pp. 4041–4065, https://doi.org/10.1137/090749694. (Cited on p. 291)
[23] I. BABUŠKA, The finite element method with penalty, Mathematics of Computation, 27 (1973),
pp. 221–221, https://doi.org/10.1090/s0025-5718-1973-0351118-5. (Cited on pp. 154,
155)
[24] E. BADER , M. K ÄRCHER , M. G REPL , AND K. V EROY-G REPL, A certified reduced basis ap-
proach for parametrized linear-quadratic optimal control problems with control constraints, IFAC-
PapersOnLine, 48 (2015), pp. 719–720, https://doi.org/10.1016/j.ifacol.2015.05.167.
(Cited on p. 83)
[25] E. BADER , M. K ÄRCHER , M. A. G REPL , AND K. V EROY, Certified reduced basis methods for
parametrized distributed elliptic optimal control problems with control constraints, SIAM Journal on
Scientific Computing, 38 (2016), pp. A3921–A3946, https://doi.org/10.1137/16m1059898.
(Cited on p. 83)
[26] S. BADIA , F. N OBILE , AND C. V ERGARA, Fluid–structure partitioned procedures based on Robin
transmission conditions, Journal of Computational Physics, 227 (2008), pp. 7027–7051, https:
//doi.org/10.1016/j.jcp.2008.04.006. (Cited on p. 291)
[27] S. BADIA , A. Q UAINI , AND A. Q UARTERONI, Modular vs. non-modular preconditioners for
fluid–structure systems with large added-mass effect, Computer Methods in Applied Mechanics and
Engineering, 197 (2008), pp. 4216–4232, https://doi.org/10.1016/j.cma.2008.04.018.
(Cited on p. 286)
Bibliography 417
[28] S. BADIA , A. Q UAINI , AND A. Q UARTERONI, Splitting methods based on algebraic factorization
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
for fluid-structure interaction, SIAM Journal on Scientific Computing, 30 (2008), pp. 1778–1805,
https://doi.org/10.1137/070680497. (Cited on p. 286)
[29] S. BADIA , A. Q UAINI , AND A. Q UARTERONI, Coupling Biot and Navier–Stokes equations for mod-
elling fluid–poroelastic media interaction, Journal of Computational Physics, 228 (2009), pp. 7986–
8014, https://doi.org/10.1016/j.jcp.2009.07.019. (Cited on p. 287)
[31] F. BALLARIN , A. D’A MARIO , S. P EROTTO , AND G. ROZZA, A POD-selective inverse distance
weighting method for fast parametrized shape morphing, International Journal for Numerical Meth-
ods in Engineering, 117 (2018), pp. 860–884, https://doi.org/10.1002/nme.5982. (Cited on
pp. 206, 351)
[36] F. BALLARIN AND G. ROZZA, POD-Galerkin monolithic reduced order models for parametrized
fluid-structure interaction problems, International Journal for Numerical Methods in Fluids, 82
(2016), pp. 1010–1034, https://doi.org/10.1002/fld.4252. (Cited on pp. 286, 287, 299)
[37] F. BALLARIN , G. ROZZA , AND Y. M ADAY, Reduced-order semi-implicit schemes for fluid-structure
interaction problems, in Model Reduction of Parametrized Systems, Springer International Pub-
lishing, 2017, pp. 149–167, https://doi.org/10.1007/978-3-319-58786-8_10. (Cited on
pp. 286, 293)
[38] F. BALLARIN , A. S ARTORI , AND G. ROZZA, RBniCS—reduced order modelling in FEniCS. http:
//mathlab.sissa.it/rbnics, 2015. (Cited on pp. 26, 318, 368)
[39] J. BANKS , W. H ENSHAW, AND D. S CHWENDEMAN, An analysis of a new stable partitioned al-
gorithm for FSI problems. Part I: Incompressible flow and elastic solids, Journal of Computational
Physics, 269 (2014), pp. 108–137, https://doi.org/10.1016/j.jcp.2014.03.006. (Cited on
p. 292)
[41] A. BARRETT AND G. R EDDIEN, On the reduced basis method, ZAMM—Journal of Applied Math-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
ematics and Mechanics/Zeitschrift für Angewandte Mathematik und Mechanik, 75 (1995), pp. 543–
549, https://doi.org/10.1002/zamm.19950750709. (Cited on p. 3)
[42] J. W. BARRETT AND C. M. E LLIOTT, Finite element approximation of the Dirichlet problem using
the boundary penalty method, Numerische Mathematik, 49 (1986), pp. 343–366, https://doi.
org/10.1007/bf01389536. (Cited on pp. 154, 155)
[43] F. BASSI AND S. R EBAY, Accurate 2D Euler computations by means of a high order discontinu-
ous finite element method, in Fourteenth International Conference on Numerical Methods in Fluid
Dynamics, 453 (1995), pp. 234–240. (Cited on p. 234)
[44] F. BASSI AND S. R EBAY, A high-order accurate discontinuous finite element method for the numer-
ical solution of the compressible Navier–Stokes equations, Journal of Computational Physics, 131
(1997), pp. 267–279, https://doi.org/10.1006/jcph.1996.5572. (Cited on pp. 234, 236)
[45] S. BASTING , A. Q UAINI , S. Č ANI Ć , AND R. G LOWINSKI, Extended ALE method for
fluid–structure interaction problems with large structural displacements, Journal of Computational
Physics, 331 (2017), pp. 312–336, https://doi.org/10.1016/j.jcp.2016.11.043. (Cited on
p. 286)
[46] S. BASTING , A. Q UAINI , S. C ANIC , AND R. G LOWINSKI, On the implementation and benchmark-
ing of an extended ALE method for FSI problems, in Fluid-Structure Interaction, De Gruyter, 2017,
pp. 3–40, https://doi.org/10.1515/9783110494259-001. (Cited on p. 284)
[47] S. BASTING , A. Q UAINI , R. G LOWINSKI , AND S. C ANI Ć, Comparison of time discretization
schemes to simulate the motion of an inextensible beam, in Numerical Mathematics and Advanced
Applications - ENUMATH 2013, A. Abdulle, S. Deparis, D. Kressner, F. Nobile, and M. Picasso,
eds., vol. 103 of Lecture Notes in Computational Science and Engineering, Springer, 2015, pp. 175–
183. (Cited on p. 283)
[51] Y. B ENGIO, Deep learning of representations for unsupervised and transfer learning, in Proceedings
of 2011 ICML Workshop on Unsupervised and Transfer Learning, vol. 27, JMLR Workshop and
Conference Proceedings, 2012, pp. 17–36. (Cited on p. 389)
[52] M. B ENÍTEZ AND A. B ERMÚDEZ, A second order characteristics finite element scheme for natural
convection problems, Journal of Computational and Applied Mathematics, 235 (2011), pp. 3270–
3284, https://doi.org/10.1016/j.cam.2011.01.007. (Cited on p. 69)
[54] P. B ENNER , S. G UGERCIN , AND K. W ILLCOX, A survey of projection-based model reduction meth-
ods for parametric dynamical systems, SIAM Review, 57 (2015), pp. 483–531, https://doi.org/
10.1137/130932715. (Cited on p. 228)
Bibliography 419
vol. 45 of Lecture Notes in Computational Science and Engineering, Springer, 2005, https://doi.
org/10.1007/3-540-27909-1. (Cited on p. 1)
[56] A. B ERLINET AND C. T HOMAS -AGNAN, Reproducing Kernel Hilbert Spaces in Probabil-
ity and Statistics, Springer Science & Business Media, 2004, https://doi.org/10.1007/
978-1-4419-9096-9. (Cited on p. 332)
[57] F. B ERNARD , A. I OLLO , AND S. R IFFAUD, Reduced-order model for the BGK equation based on
POD and optimal transport, Journal of Computational Physics, 373 (2018), pp. 545–570, https:
//doi.org/10.1016/j.jcp.2018.07.001. (Cited on p. 298)
[58] L. B ERTAGNA , A. Q UAINI , AND A. V ENEZIANI, Deconvolution-based nonlinear filtering for in-
compressible flows at moderately large Reynolds numbers, International Journal for Numerical Meth-
ods in Fluids, 81 (2015), pp. 463–488, https://doi.org/10.1002/fld.4192. (Cited on p. 171)
[59] W.-J. B EYN AND V. T HÜMMLER, Freezing solutions of equivariant evolution equations, SIAM
Journal on Applied Dynamical Systems, 3 (2004), pp. 85–116, https://doi.org/10.1137/
030600515. (Cited on p. 298)
[61] C. B ISHOP, Pattern Recognition and Machine Learning, Information Science and Statistics,
Springer, 1 ed., 2006. (Cited on p. 330)
[62] D. A. B ISTRIAN AND I. M. NAVON, Randomized dynamic mode decomposition for nonintrusive
reduced order modelling, International Journal for Numerical Methods in Engineering, 112 (2017),
pp. 3–25, https://doi.org/10.1002/nme.5499. (Cited on p. 209)
[63] K. B IZON AND G. C ONTINILLO, Reduced order modelling of chemical reactors with recycle by
means of POD-penalty method, Computers & Chemical Engineering, 39 (2012), pp. 22–32, https:
//doi.org/10.1016/j.compchemeng.2011.10.001. (Cited on p. 154)
[65] D. B OFFI , F. B REZZI , AND M. F ORTIN, Mixed Finite Element Methods and Applications, vol. 44 of
Springer Series in Computational Mathematics, Springer-Verlag, 1 ed., 2013, https://doi.org/
10.1007/978-3-642-36519-5. (Cited on p. 318)
[66] J. B OLAND AND W. L AYTON, An analysis of the finite element method for natural convection
problems, Numerical Methods for Partial Differential Equations, 6 (1990), pp. 115–126, https:
//doi.org/10.1002/num.1690060202. (Cited on p. 69)
[67] J. B OLAND AND W. L AYTON, Error analysis for finite element methods for steady natural convec-
tion problems, Numerical Functional Analysis and Optimization, 11 (1990), pp. 449–483, https:
//doi.org/10.1080/01630569008816383. (Cited on p. 69)
[68] G. B ONCORAGLIO AND C. FARHAT, Active manifold and model reduction for multidisciplinary
analysis and optimization, in AIAA Scitech 2021 Forum, 2021, p. 1694, https://doi.org/10.
2514/6.2021-1694. (Cited on p. 338)
[69] J. B ORGGAARD , T. I LIESCU , AND J. P. ROOP, A bounded artificial viscosity large eddy simulation
model, SIAM Journal on Numerical Analysis, 47 (2009), pp. 622–645, https://doi.org/10.
1137/060656164. (Cited on p. 170)
420 Bibliography
[70] N. B OTKIN AND V. T UROVA -B OTKINA, An algorithm for finding the Chebyshev center of a convex
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[71] J. B OUSSINESQ, Des eaux courantes. Memoires présentés par divers savants à l’Académie des
Sciences de l’Institut National de France, Tome XXIII, (1877). (Cited on p. 167)
[73] H. B REZIS, Analyse fonctionnelle: Theorie et applications, Dunod, 2002. (Cited on p. 71)
[74] F. B REZZI, On the existence, uniqueness and approximation of saddle-point problems arising from
Lagrangian multipliers, Revue française d’automatique, informatique, recherche opérationnelle.
Analyse numérique, 8 (1974), pp. 129–151, https://doi.org/10.1051/m2an/197408R201291.
(Cited on pp. 8, 22, 64, 318)
[75] F. B REZZI AND J. P ITKÄRANTA, On the stabilization of finite element approximations of the Stokes
equations, in Efficient Solutions of Elliptic Systems, Vieweg+Teubner Verlag, 1984, pp. 11–19,
https://doi.org/10.1007/978-3-663-14169-3_2. (Cited on pp. 125, 128, 132)
[76] F. B REZZI , J. R APPAZ , AND P. A. R AVIART, Finite dimensional approximation of nonlinear prob-
lems, Numerische Mathematik, 36 (1980), pp. 1–25, https://doi.org/10.1007/BF01395985.
(Cited on pp. 64, 65, 72, 74, 102, 136, 137)
[79] S. L. B RUNTON AND J. N. K UTZ, Data-Driven Science and Engineering: Machine Learning,
Dynamical Systems, and Control, Cambridge University Press, 2019. (Cited on p. 207)
[80] S. L. B RUNTON , J. L. P ROCTOR , AND J. N. K UTZ, Discovering governing equations from data
by sparse identification of nonlinear dynamical systems, Proceedings of the National Academy of
Sciences, 113 (2016), pp. 3932–3937, https://doi.org/10.1073/pnas.1517384113. (Cited
on p. 222)
[81] A. B UFFA , Y. M ADAY, A. T. PATERA , C. P RUD ’ HOMME , AND G. T URINICI, A priori conver-
gence of the greedy algorithm for the parametrized reduced basis method, ESAIM: Mathematical
Modelling and Numerical Analysis, 46 (2012), pp. 595–603, https://doi.org/10.1051/m2an/
2011056. (Cited on pp. 5, 15, 31, 32)
[82] M. D. B UHMANN, Radial Basis Functions: Theory and Implementations, vol. 12 of Cambridge
Monographs on Applied and Computational Mathematics, Cambridge University Press, 2003,
https://doi.org/10.1017/CBO9780511543241. (Cited on pp. 10, 206, 346, 348, 350)
[83] T. B UI -T HANH , M. DAMODARAN , AND K. W ILLCOX, Aerodynamic data reconstruction and in-
verse design using proper orthogonal decomposition, AIAA Journal, 42 (2004), pp. 1505–1516,
https://doi.org/10.2514/1.2159. (Cited on p. 205)
Bibliography 421
tion in blood flow capturing non-zero longitudinal structure displacement, Journal of Computational
Physics, 235 (2013), pp. 515–541, https://doi.org/10.1016/j.jcp.2012.08.033. (Cited on
p. 286)
[85] J. B URKARDT, M. G UNZBURGER , AND H. L EE, POD and CVT-based reduced-order modeling
of Navier–Stokes flows, Computer Methods in Applied Mechanics and Engineering, 196 (2006),
pp. 337–355. (Cited on pp. 15, 87, 92)
[86] N. C AGNIART, A Few Non Linear Approaches in Model Order Reduction, PhD thesis, École Doc-
torale de Sciences Mathématiques de Paris Centre, 2018. (Cited on p. 298)
[87] N. C AGNIART, R. C RISOVAN , Y. M ADAY, AND R. A BGRALL, Model Order Reduction for Hyper-
bolic Problems: A New Framework. hal:01583224, 2017. (Cited on p. 298)
[88] N. C AGNIART, Y. M ADAY, AND B. S TAMM, Model order reduction for problems with large con-
vection effects, in Contributions to Partial Differential Equations and Applications, W. Fitzgibbon,
Y. Kuznetsov, P. Neittaanmäki, J. Periaux, and O. Pironneau, eds., vol. 47 of Computational Methods
in Applied Sciences, Springer, Cham., 2019, pp. 131–150. (Cited on pp. 298, 299, 305)
[89] J. W. C AHN AND J. E. H ILLIARD, Free energy of a nonuniform system. I. Interfacial free en-
ergy, The Journal of Chemical Physics, 28 (1958), pp. 258–267, https://doi.org/10.1063/1.
1744102. (Cited on p. 280)
[90] S. C AI , Z. M AO , Z. WANG , M. Y IN , AND G. E. K ARNIADAKIS, Physics-informed neural networks
(PINNs) for fluid mechanics: A review, Acta Mechanica Sinica, 37 (2021), pp. 1727–1738, https:
//doi.org/10.1007/s10409-021-01148-1. (Cited on p. 401)
[91] A. C AIAZZO , T. I LIESCU , V. J OHN , AND S. S CHYSCHLOWA, A numerical investigation of veloc-
ity–pressure reduced order models for incompressible flows, Journal of Computational Physics, 259
(2014), pp. 598–616, https://doi.org/10.1016/j.jcp.2013.12.004. (Cited on pp. 144, 187)
[92] O. C ALIN, Deep Learning Architectures: A Mathematical Approach, Springer Series in the Data
Sciences, Springer International Publishing, 2020. (Cited on pp. 390, 391, 392, 393)
[93] G. C ALOZ AND J. R APPAZ, Numerical analysis for nonlinear and bifurcation problems, Handbook
of Numerical Analysis, 5 (1997), pp. 487–637. (Cited on pp. 65, 72, 102, 136)
[94] E. J. C ANDÈS , X. L I , Y. M A , AND J. W RIGHT, Robust principal component analysis?, Journal of
the ACM (JACM), 58 (2011), pp. 1–37, https://doi.org/10.1145/1970392.1970395. (Cited
on p. 205)
[95] C. C ANTWELL , D. M OXEY, A. C OMERFORD , A. B OLIS , G. ROCCO , G. M ENGALDO , D. D E
G RAZIA , S. YAKOVLEV, J.-E. L OMBARD , D. E KELSCHOT, B. J ORDI , H. X U , Y. M OHAMIED ,
C. E SKILSSON , B. N ELSON , P. VOS , C. B IOTTO , R. K IRBY, AND S. S HERWIN, Nektar++:
An open-source spectral/hp element framework, Computer Physics Communications, 192 (2015),
pp. 205–219, https://doi.org/10.1016/j.cpc.2015.02.008. (Cited on p. 229)
[96] C. C ANUTO , M. H USSAINI , A. Q UARTERONI , AND T. Z HANG, Spectral Methods Fundamentals
in Single Domains, Scientific Computation, Springer, 2006. (Cited on pp. 223, 373)
[97] C. C ANUTO , M. H USSAINI , A. Q UARTERONI , AND T. Z HANG, Spectral Methods Evolution to
Complex Geometries and Applications to Fluid Dynamics, Scientific Computation, Springer, 2007.
(Cited on pp. 223, 373)
[98] G. C ARERE , M. S TRAZZULLO , F. BALLARIN , G. ROZZA , AND R. S TEVENSON, A weighted POD-
reduction approach for parametrized PDE-constrained optimal control problems with random in-
puts and applications to environmental sciences, Computers & Mathematics with Applications, 102
(2021), pp. 261–276, https://doi.org/10.1016/j.camwa.2021.10.020. (Cited on pp. 95,
261)
422 Bibliography
[99] K. C ARLBERG, Adaptive h-refinement for reduced-order models, International Journal for Numeri-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[100] K. C ARLBERG , C. FARHAT, J. C ORTIAL , AND D. A MSALLEM, The GNAT method for nonlinear
model reduction: Effective implementation and application to computational fluid dynamics and
turbulent flows, Journal of Computational Physics, 242 (2013), pp. 623–647, https://doi.org/
10.1016/j.jcp.2013.02.028. (Cited on p. 151)
[101] J. C ARLTON, Marine Propellers and Propulsion (Second Edition), Butterworth-Heinemann, Oxford,
2007, ch. 6—Propeller performance characteristics, pp. 88–135, https://doi.org/10.1016/
B978-075068150-6/50008-5. (Cited on pp. 346, 355)
[102] P. C AUSIN , J. F. G ERBEAU , AND F. N OBILE, Added-mass effect in the design of partitioned algo-
rithms for fluid–structure problems, Computer Methods in Applied Mechanics and Engineering, 194
(2005), pp. 4506–4527, https://doi.org/10.1016/j.cma.2004.12.005. (Cited on p. 286)
[103] F. C AVALLINI AND F. C RISCIANI, Quasi-geostrophic Theory of Oceans and Atmosphere: Topics in
the Dynamics and Thermodynamics of the Fluid Earth, vol. 45, Springer Science & Business Media,
2013, https://doi.org/10.1007/978-94-007-4691-6_2. (Cited on p. 89)
[105] T. C HACÓN R EBOLLO, A term by term stabilization algorithm for finite element solution of incom-
pressible flow problems, Numerische Mathematik, 79 (1998), pp. 283–319, https://doi.org/10.
1007/s002110050341. (Cited on pp. 125, 132)
[106] T. C HACÓN R EBOLLO, An analysis technique for stabilized finite element solution of incompressible
flows, ESAIM: Mathematical Modelling and Numerical Analysis, 35 (2001), pp. 57–89, https:
//doi.org/10.1051/m2an:2001107. (Cited on p. 65)
[108] T. C HACÓN R EBOLLO , M. G ÓMEZ M ÁRMOL , V. G IRAULT, AND I. S ÁNCHEZ M UÑOZ, A high
order term-by-term stabilization solver for incompressible flow problems, IMA Journal of Numeri-
cal Analysis, 33 (2012), pp. 974–1007, https://doi.org/10.1093/imanum/drs023. (Cited on
pp. 125, 132, 136)
[109] T. C HACÓN R EBOLLO , M. G ÓMEZ M ÁRMOL , F. H ECHT, S. RUBINO , AND I. S ÁNCHEZ M UÑOZ,
A high-order local projection stabilization method for natural convection problems, Journal of Sci-
entific Computing, 74 (2018), pp. 667–692, https://doi.org/10.1007/s10915-017-0469-9.
(Cited on pp. 69, 75)
[110] T. C HACÓN R EBOLLO , M. G ÓMEZ M ÁRMOL , AND S. RUBINO, Numerical analysis of a finite
element projection-based VMS turbulence model with wall laws, Computer Methods in Applied Me-
chanics and Engineering, 285 (2015), pp. 379–405, https://doi.org/10.1016/j.cma.2014.
11.023. (Cited on pp. 133, 135, 136)
[112] S. C HANDRA, A three-dimensional insight into the complexity of flow convergence in mitral regur-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
gitation: Adjunctive benefit of anatomic regurgitant orifice area, American Journal of Physiology.
Heart and Circulatory Physiology, 301 (2011), https://doi.org/10.1152/ajpheart.00275.
2011. (Cited on p. 311)
[114] D. C HAPELLE , A. G ARIAH , P. M OIREAU , AND J. S AINTE -M ARIE, A Galerkin strategy with
proper orthogonal decomposition for parameter-dependent problems: Analysis, assessments and
applications to parameter estimation, ESAIM: Mathematical Modelling and Numerical Analysis,
47 (2013), pp. 1821–1843, https://doi.org/10.1051/m2an/2013090. (Cited on pp. 15, 87,
92)
[116] S. C HATURANTABUT AND D. C. S ORENSEN, Nonlinear model reduction via discrete empirical
interpolation, SIAM Journal on Scientific Computing, 32 (2010), pp. 2737–2764, https://doi.
org/10.1137/090766498. (Cited on pp. 56, 108)
[117] P. C HEN , A. Q UARTERONI , AND G. ROZZA, A weighted reduced basis method for elliptic partial
differential equations with random input data, SIAM Journal on Numerical Analysis, 51 (2013),
pp. 3163–3185, https://doi.org/10.1137/130905253. (Cited on pp. 251, 252, 253, 254)
[118] P. C HEN , A. Q UARTERONI , AND G. ROZZA, Multilevel and weighted reduced basis method for
stochastic optimal control problems constrained by Stokes equations, Numerische Mathematik, 133
(2016), pp. 67–102, https://doi.org/10.1007/s00211-015-0743-4. (Cited on p. 257)
[122] W. C HERDRON , F. D URST, AND J. H. W HITELAW, Asymmetric flows and instabilities in symmetric
ducts with sudden expansions, Journal of Fluid Mechanics, 84 (1978), pp. 13–31, https://doi.
org/10.1017/s0022112078000026. (Cited on p. 311)
[123] F. C HINESTA , P. L ADEVEZE , AND E. C UETO, A short review on model order reduction based on
proper generalized decomposition, Archives of Computational Methods in Engineering, 18 (2011),
p. 395, https://doi.org/10.1007/s11831-011-9064-7. (Cited on pp. 43, 53)
[125] M. C IANFERRA , V. A RMENIO , AND S. I ANNIELLO, Hydroacoustic noise from different geometries,
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[126] M. C IANFERRA , S. I ANNIELLO , AND V. A RMENIO, Assessment of methodologies for the so-
lution of the Ffowcs Williams and Hawkings equation using LES of incompressible single-phase
flow around a finite-size square cylinder, Journal of Sound and Vibration, 453 (2019), pp. 1–24,
https://doi.org/10.1016/j.jsv.2019.04.001. (Cited on p. 216)
[128] P. G. C IARLET, Mathematical Elasticity: Volume II: Theory of Plates, Studies in Mathemat-
ics and Its Applications, Elsevier Science, 1997, https://doi.org/10.1016/s0168-2024(97)
x8001-4. (Cited on p. 109)
[129] P. G. C IARLET, Linear and Nonlinear Functional Analysis with Applications, Other Titles in Ap-
plied Mathematics, Society for Industrial and Applied Mathematics, 2013. (Cited on pp. 99, 103,
104)
[130] D. C. C IRE ŞAN , U. M EIER , L. M. G AMBARDELLA , AND J. S CHMIDHUBER, Deep, big, sim-
ple neural nets for handwritten digit recognition, Neural Computation, 22 (2010), pp. 3207–3220,
https://doi.org/10.1162/neco_a_00052. (Cited on p. 389)
[131] B. C OCKBURN AND B. D ONG, An analysis of the minimal dissipation local discontinuous Galerkin
method for convection–diffusion problems, Journal of Scientific Computing, 32 (2007), pp. 233–262,
https://doi.org/10.1007/s10915-007-9130-3. (Cited on p. 234)
[132] R. C ODINA, Stabilization of incompressibility and convection through orthogonal sub-scales in finite
element methods, Computer Methods in Applied Mechanics and Engineering, 190 (2000), pp. 1579–
1599, https://doi.org/10.1016/s0045-7825(00)00254-1. (Cited on pp. 132, 133)
[133] A. C OHEN AND R. D E VORE, Kolmogorov widths under holomorphic mappings, IMA Journal of
Numerical Analysis, 36 (2015), pp. 1–12, https://doi.org/10.1093/imanum/dru066. (Cited
on pp. 32, 106, 297)
[134] P. G. C ONSTANTINE, Active Subspaces: Emerging Ideas for Dimension Reduction in Pa-
rameter Studies, vol. 2 of SIAM Spotlights, SIAM, 2015, https://doi.org/10.1137/1.
9781611973860. (Cited on pp. 9, 220, 327, 328, 330, 338, 384)
[135] P. G. C ONSTANTINE AND P. D IAZ, Global sensitivity metrics from active subspaces, Reliability
Engineering & System Safety, 162 (2017), pp. 1–13, https://doi.org/10.1016/j.ress.2017.
01.013. (Cited on pp. 340, 341)
[136] P. G. C ONSTANTINE , E. D OW, AND Q. WANG, Active subspace methods in theory and practice:
Applications to kriging surfaces, SIAM Journal on Scientific Computing, 36 (2014), pp. A1500–
A1524, https://doi.org/10.1137/130916138. (Cited on p. 327)
[137] M. C OUPLET, P. S AGAUT, AND C. BASDEVANT, Intermodal energy transfers in a proper orthog-
onal decomposition-Galerkin representation of a turbulent separated flow, Journal of Fluid Me-
chanics, 491 (2003), pp. 257–284, https://doi.org/10.1017/S0022112003005615. (Cited on
p. 164)
[138] R. C OURANT, K. F RIEDRICHS , AND H. L EWY, Über die partiellen Differenzengleichungen der
mathematischen Physik, Mathematische Annalen, 100 (1928), pp. 32–74, https://doi.org/10.
1007/bf01448839. (Cited on pp. 149, 187)
Bibliography 425
[139] R. C OURANT, K. F RIEDRICHS , AND H. L EWY, On the partial difference equations of mathematical
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
physics, IBM Journal of Research and Development, 11 (1967), pp. 215–234, https://doi.org/
10.1147/rd.112.0215. (Cited on pp. 149, 187)
[141] G. W. C OWLES , N. PAROLINI , AND M. L. S AWLEY, Numerical simulation using RANS-based tools
for America’s Cup design, Proceedings of the 16th Chesapeake Sailing Yacht Symposium (2003),
https://doi.org/10.5957/CSYS-2003-007. (Cited on p. 168)
[145] J. C. DE LOS R EYES AND F. T RÖLTZSCH, Optimal control of the stationary Navier–Stokes equa-
tions with mixed control-state constraints, SIAM Journal on Control and Optimization, 46 (2007),
pp. 604–629, https://doi.org/10.1137/050646949. (Cited on p. 83)
[146] G. D E VAHL DAVIS, Natural convection of air in a square cavity: A bench mark numerical solution,
International Journal for Numerical Methods in Fluids, 3 (1983), pp. 249–264, https://doi.org/
10.1002/fld.1650030305. (Cited on p. 75)
[147] M. A. B. D EAKIN, Modelling Biological Systems, Birkhäuser Boston, 1990, pp. 2–16, https:
//doi.org/10.1007/978-1-4684-6784-0_1. (Cited on p. 413)
[148] L. D EDÈ, Optimal flow control for Navier–Stokes equations: Drag minimization, International Jour-
nal for Numerical Methods in Fluids, 55 (2007), pp. 347–366, https://doi.org/10.1002/fld.
1464. (Cited on p. 83)
[149] L. D EDÈ, Reduced basis method and a posteriori error estimation for parametrized linear-quadratic
optimal control problems, SIAM Journal on Scientific Computing, 32 (2010), pp. 997–1019, https:
//doi.org/10.1137/090760453. (Cited on pp. 83, 87)
[150] L. D EDÈ, Reduced basis method and error estimation for parametrized optimal control problems
with control constraints, Journal of Scientific Computing, 50 (2012), pp. 287–305, https://doi.
org/10.1007/s10915-011-9483-5. (Cited on p. 367)
[151] E. D ELGADO ÁVILA, Development of Reduced Numeric Models to Aero-thermic Flows in Buildings,
PhD thesis, University of Seville, 2018. (Cited on pp. 60, 69, 136)
[153] N. D EMO , M. S TRAZZULLO , AND G. ROZZA, An Extended Physics Informed Neural Network for
Preliminary Analysis of Parametric Optimal Control Problems, arXiv preprint arXiv:2110.13530
(2021). (Cited on pp. 95, 401)
426 Bibliography
means of proper orthogonal decomposition and dynamic mode decomposition, in Technology and
Science for the Ships of the Future: Proceedings of NAV 2018: 19th International Conference
on Ship & Maritime Research, IOS Press, 2018, pp. 212–219, https://doi.org/10.3233/
978-1-61499-870-9-212. (Cited on pp. 11, 212, 213, 214, 347, 356)
[155] N. D EMO , M. T EZZELE , A. M OLA , AND G. ROZZA, An efficient shape parametrisation by free-
form deformation enhanced by active subspace for hull hydrodynamic ship design problems in open
source environment, in Proceedings of ISOPE 2018: The 28th International Ocean and Polar Engi-
neering Conference, vol. 3, 2018, pp. 565–572. (Cited on pp. 327, 347, 358, 360, 383)
[156] N. D EMO , M. T EZZELE , A. M OLA , AND G. ROZZA, A complete data-driven framework for the
efficient solution of parametric shape design and optimisation in naval engineering problems, in
Proceedings of MARINE 2019: VIII International Conference on Computational Methods in Marine
Engineering, R. Bensow and J. Ringsberg, eds., 2019, pp. 111–121. (Cited on p. 347)
[157] N. D EMO , M. T EZZELE , A. M OLA , AND G. ROZZA, Hull shape design optimization with param-
eter space and model reductions, and self-learning mesh morphing, Journal of Marine Science and
Engineering, 9 (2021), p. 185, https://doi.org/10.3390/jmse9020185. (Cited on pp. 11, 327,
340, 346, 363, 383, 384, 411)
[158] N. D EMO , M. T EZZELE , AND G. ROZZA, EZyRB: Easy reduced basis method, The Journal of Open
Source Software, 3 (2018), p. 661, https://doi.org/10.21105/joss.00661. (Cited on pp. 204,
375, 382)
[159] N. D EMO , M. T EZZELE , AND G. ROZZA, PyDMD: Python dynamic mode decomposition, The
Journal of Open Source Software, 3 (2018), p. 530, https://doi.org/10.21105/joss.00530.
(Cited on pp. 207, 383)
[160] N. D EMO , M. T EZZELE , AND G. ROZZA, A non-intrusive approach for reconstruction of POD
modal coefficients through active subspaces, Comptes rendus mécanique de l’Académie des Sci-
ences, DataBEST 2019 Special Issue, 347 (2019), pp. 873–881, https://doi.org/10.1016/j.
crme.2019.11.012. (Cited on pp. 206, 342, 347)
[161] N. D EMO , M. T EZZELE , AND G. ROZZA, A supervised learning approach involving active sub-
spaces for an efficient genetic algorithm in high-dimensional optimization problems, SIAM Journal
on Scientific Computing, 43 (2021), pp. B831–B853, https://doi.org/10.1137/20M1345219.
(Cited on pp. 11, 327, 340, 350, 384)
[162] S. D EPARIS, Reduced basis error bound computation of parameter-dependent Navier–Stokes equa-
tions by the natural norm approach, SIAM Journal on Numerical Analysis, 46 (2008), pp. 2039–
2067, https://doi.org/10.1137/060674181. (Cited on p. 131)
[163] S. D EPARIS , D. F ORTI , AND A. Q UARTERONI, A rescaled localized radial basis function interpola-
tion on non-Cartesian and nonconforming grids, SIAM Journal on Scientific Computing, 36 (2014),
pp. A2745–A2762, https://doi.org/10.1137/130947179. (Cited on p. 350)
[164] S. D EPARIS AND G. ROZZA, Reduced basis method for multi-parameter-dependent steady Navier-
Stokes equations: Applications to natural convection in a cavity, Journal of Computational Physics,
228 (2009), pp. 4359–4378, https://doi.org/10.1016/j.jcp.2009.03.008. (Cited on pp. 5,
131)
[165] A. D ERKEVORKIAN , P. AVERY, C. FARHAT, J. R ABINOVITCH , AND L. P ETERSON, Effects of
Structural Parameters on the FSI Simulation of Supersonic Parachute Deployments, in AIAA Avia-
tion 2019 Forum, Dallas, Texas, https://doi.org/10.2514/6.2019-3276. (Cited on p. 283)
[166] M. D EVILLE , P. F ISCHER , AND E. M UND, High-Order Methods for Incompressible Fluid Flow,
Cambridge University Press, 2002, https://doi.org/10.1017/CBO9780511546792. (Cited on
p. 223)
Bibliography 427
model for the spread of Ebola in Western Africa and metrics for resource allocation, Applied Math-
ematics and Computation, 324 (2018), pp. 141–155, https://doi.org/10.1016/j.amc.2017.
11.039. (Cited on pp. xxxiv, 339)
[168] M. D IEZ , E. F. C AMPANA , AND F. S TERN, Design-space dimensionality reduction in shape opti-
mization by Karhunen–Loève expansion, Computer Methods in Applied Mechanics and Engineering,
283 (2015), pp. 1525–1544, https://doi.org/10.1016/j.cma.2014.10.042. (Cited on p. 347)
[169] E. J. D OEDEL, Lecture Notes on Numerical Analysis of Nonlinear Equations, Springer Netherlands,
2007, pp. 1–49, https://doi.org/10.1007/978-1-4020-6356-5_1. (Cited on p. 104)
[170] E. J. D OEDEL , H. B. K ELLER , AND J. P. K ERNEVEZ, Numerical analysis and control of bifurcation
problems (i): Bifurcation in finite dimensions, International Journal of Bifurcation and Chaos, 01
(1991), pp. 493–520, https://doi.org/10.1142/S0218127491000397. (Cited on pp. 104, 316)
[171] V. D OLEJŠÍ AND M. F EISTAUER, Discontinuous Galerkin Method, Analysis and Applications to
Compressible Flow. Springer Series in Computational Mathematics, 48, Springer, 2015, https:
//doi.org/10.1007/978-3-319-19267-3. (Cited on p. 234)
[174] J. D OUGLAS AND T. D UPONT, Interior penalty procedures for elliptic and parabolic Galerkin
methods, in Computing Methods in Applied Sciences, Springer, 1976, pp. 207–216, https:
//doi.org/10.1007/BFb0120591. (Cited on p. 239)
[175] M. D ROHMANN , B. H AASDONK , AND M. O HLBERGER, Reduced basis approximation for non-
linear parametrized evolution equations based on empirical operator interpolation, SIAM Journal
on Scientific Computing, 34 (2012), pp. A937–A969, https://doi.org/10.1137/10081157X.
(Cited on p. 143)
[177] D. DYLEWSKY, M. TAO , AND J. N. K UTZ, Dynamic mode decomposition for multiscale nonlinear
physics, Physical Review E, 99 (2019), p. 063311, https://doi.org/10.1103/PhysRevE.99.
063311. (Cited on p. 210)
[178] C. E CKART AND G. YOUNG, The approximation of one matrix by another of lower rank, Psychome-
trika, 1 (1936), pp. 211–218, https://doi.org/10.1007/BF02288367. (Cited on p. 4)
[179] J. E DMONDS, Matroids and the greedy algorithm, Mathematical Programming, 1 (1971), pp. 127–
136, https://doi.org/10.1007/BF01584082. (Cited on p. 131)
[182] J. L. E LMAN, Finding structure in time, Cognitive Science, 14 (1990), pp. 179–211, https://doi.
org/10.1207/s15516709cog1402_1. (Cited on p. 393)
[186] L. C. E VANS, Partial Differential Equations, Graduate Studies in Mathematics, American Mathe-
matical Society, 2010, https://doi.org/10.1090/gsm/019. (Cited on p. 100)
[189] P. E. FARRELL , A. B IRKISSON , AND S. W. F UNKE, Deflation techniques for finding distinct solu-
tions of nonlinear partial differential equations, SIAM Journal on Scientific Computing, 37 (2015),
pp. A2026–A2045, https://doi.org/10.1137/140984798. (Cited on p. 114)
[190] G. E. FASSHAUER AND J. G. Z HANG, On choosing “optimal” shape parameters for RBF
approximation, Numerical Algorithms, 45 (2007), pp. 345–368, https://doi.org/10.1007/
s11075-007-9072-8. (Cited on p. 190)
[194] M. A. F ERNÁNDEZ , J. M ULLAERT, AND M. V IDRASCU, Explicit Robin–Neumann schemes for the
coupling of incompressible fluids with thin-walled structures, Computer Methods in Applied Me-
chanics and Engineering, 267 (2013), pp. 566–593, https://doi.org/10.1016/j.cma.2013.
09.020. (Cited on p. 286)
Bibliography 429
[195] J. E. F FOWCS W ILLIAMS AND D. L. H AWKINGS, Sound generation by turbulence and surfaces
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
in arbitrary motion, Philosophical Transactions of the Royal Society A, 264 (1969), pp. 321–342,
https://doi.org/10.1098/rsta.1969.0031. (Cited on p. 214)
[196] L. F ICK , Y. M ADAY, A. T. PATERA , AND T. TADDEI, A stabilized POD model for turbulent flows
over a range of Reynolds numbers: Optimal parameter sampling and constrained projection, Journal
of Computational Physics, 371 (2018), pp. 214–243. (Cited on p. 164)
[197] T. L. F INE, Feedforward Neural Network Methodology, Springer Science & Business Media, 2006.
(Cited on p. 393)
[198] J. P. F INK AND W. C. R HEINBOLDT, On the discretization error of parametrized nonlinear equa-
tions, SIAM Journal on Numerical Analysis, 20 (1983), pp. 732–746, https://doi.org/10.
1137/0720049. (Cited on p. 3)
[199] J. P. F INK AND W. C. R HEINBOLDT, On the error behavior of the reduced basis technique
for nonlinear finite element approximations, ZAMM-Zeitschrift für Angewandte Mathematik und
Mechanik, 63 (1983), pp. 21–28, https://doi.org/10.1002/zamm.19830630105. (Cited on
p. 3)
[201] D. F ORTI , M. B UKAC , A. Q UAINI , S. C ANIC , AND S. D EPARIS, A monolithic approach to fluid–
composite structure interaction, Journal of Scientific Computing, 72 (2017), pp. 396–421, https:
//doi.org/10.1007/s10915-017-0363-5. (Cited on p. 287)
[202] D. F ORTI AND G. ROZZA, Efficient geometrical parametrisation techniques of interfaces for
reduced-order modelling: Application to fluid–structure interaction coupling problems, International
Journal of Computational Fluid Dynamics, 28 (2014), pp. 158–169, https://doi.org/10.1080/
10618562.2014.932352. (Cited on pp. 350, 351)
[203] M. F OWLER, Refactoring: Improving the Design of Existing Code, Addison-Wesley Professional,
2018. (Cited on p. 386)
[204] L. P. F RANCA AND C. FARHAT, Bubble functions prompt unusual stabilized finite element methods,
Computer Methods in Applied Mechanics and Engineering, 123 (1995), pp. 299–308, https://
doi.org/10.1016/0045-7825(94)00721-x. (Cited on p. 132)
[205] L. P. F RANCA AND S. L. F REY, Stabilized finite element methods: II. The incompressible Navier-
Stokes equations, Computer Methods in Applied Mechanics and Engineering, 99 (1992), pp. 209–
233, https://doi.org/10.1016/0045-7825(92)90041-H. (Cited on p. 130)
[206] B. A. F RENO AND K. T. C ARLBERG, Machine-learning error models for approximate solutions to
parameterized systems of nonlinear equations, Computer Methods in Applied Mechanics and Engi-
neering, 348 (2019), pp. 250–296, https://doi.org/10.1016/j.cma.2019.01.024. (Cited on
p. 298)
[207] S. F U AND L. WANG, RANS modeling of high-speed aerodynamic flow transition with consideration
of stability theory, Progress in Aerospace Sciences, 58 (2013), pp. 36–59, https://doi.org/10.
1016/j.paerosci.2012.08.004. (Cited on p. 168)
comparison of LES data-driven reduced order approaches for hydroacoustic analysis, Computers &
Fluids, 216 (2021), p. 104819, https://doi.org/10.1016/j.compfluid.2020.104819. (Cited
on pp. 214, 381, 383)
[210] M. G ADALLA , M. T EZZELE , A. M OLA , AND G. ROZZA, BladeX: Python blade morphing, The
Journal of Open Source Software, 4 (2019), p. 1203, https://doi.org/10.21105/joss.01203.
(Cited on pp. 355, 359, 385)
[211] G. G ALDI, An Introduction to the Mathematical Theory of the Navier-Stokes Equations: Steady-
State Problems, Springer Science & Business Media, 2011. (Cited on p. 257)
[214] A. G ARZOZI AND D. G REENBLATT, A pulsed Coandă-effect reciprocating wind energy generator,
Energy, 150 (2018), pp. 965–978, https://doi.org/10.1016/j.energy.2018.02.114. (Cited
on p. 373)
[216] M. W. G EE , U. K ÜTTLER , AND W. A. WALL, Truly monolithic algebraic multigrid for fluid–
structure interaction, International Journal for Numerical Methods in Engineering, 85 (2011),
pp. 987–1016, https://doi.org/10.1002/nme.3001. (Cited on p. 287)
[217] R. G EELEN AND K. W ILLCOX, Localized non-intrusive reduced-order modeling in the operator
inference framework, Philosophical Transactions of the Royal Society A, 380 (2022), p. 20210206,
https://doi.org/10.1098/rsta.2021.0206. (Cited on p. 222)
[218] J. G ENOVESE , F. BALLARIN , G. ROZZA , AND C. C ANUTO, Weighted Reduced Order Methods
for Uncertainty Quantification in Computational Fluid Dynamics, in preparation (2022). (Cited on
p. 257)
[219] S. G EORGAKA , G. S TABILE , G. ROZZA , AND M. J. B LUCK, Parametric POD-Galerkin model or-
der reduction for unsteady-state heat transfer problems, Communications in Computational Physics,
27 (2020), pp. 1–32, https://doi.org/10.4208/cicp.oa-2018-0207. (Cited on p. 144)
[220] S. G EORGAKA , G. S TABILE , K. S TAR , G. ROZZA , AND M. J. B LUCK, A hybrid reduced order
method for modelling turbulent heat transfer problems, Computers & Fluids, 208 (2020), p. 104615,
https://doi.org/10.1016/j.compfluid.2020.104615. (Cited on pp. 205, 350, 381)
[221] J.-F. G ERBEAU AND D. L OMBARDI, Approximated Lax pairs for the reduced order integration of
nonlinear evolution equations, Journal of Computational Physics, 265 (2014), pp. 246–269, https:
//doi.org/10.1016/j.jcp.2014.01.047. (Cited on p. 298)
[222] J. F. G ERBEAU , D. L OMBARDI , AND E. S CHENONE, Reduced order model in cardiac electrophys-
iology with approximated Lax pairs, Advances in Computational Mathematics, 41 (2015), pp. 1103–
1130, https://doi.org/10.1007/s10444-014-9393-9. (Cited on p. 298)
Bibliography 431
[223] A.-L. G ERNER AND K. V EROY, Certified reduced basis methods for parametrized saddle point
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[224] R. G. G HANEM AND P. D. S PANOS, Stochastic Finite Elements: A Spectral Approach, Courier
Corporation, 2003. (Cited on p. 405)
[227] V. G IRAULT AND P. A. R AVIART, Finite Element Approximation of the Navier-Stokes Equations,
Springer, 1981. (Cited on p. 8)
[229] M. G IRFOGLIO , A. Q UAINI , AND G. ROZZA, A finite volume approximation of the Navier-Stokes
equations with nonlinear filtering stabilization, Computers & Fluids, 187 (2019), pp. 27–45, https:
//doi.org/10.1016/j.compfluid.2019.05.001. (Cited on p. 171)
[230] M. G IRFOGLIO , A. Q UAINI , AND G. ROZZA, A POD-Galerkin reduced order model for a LES
filtering approach, Journal of Computational Physics, 436 (2021), p. 110260, https://doi.org/
10.1016/j.jcp.2021.110260. (Cited on pp. 171, 205)
[231] M. G IRFOGLIO , A. Q UAINI , AND G. ROZZA, Fluid-structure interaction simulations with a LES fil-
tering approach in solids4Foam, Communications in Applied and Industrial Mathematics, 12 (2021),
pp. 13–28, https://doi.org/doi:10.2478/caim-2021-0002. (Cited on p. 171)
[232] M. G IRFOGLIO , A. Q UAINI , AND G. ROZZA, A Hybrid Reduced Order Model for Nonlinear LES
Filtering. arXiv:2107.12933, 2021. (Cited on p. 171)
[233] M. G IRFOGLIO , A. Q UAINI , AND G. ROZZA, Pressure stabilization strategies for a LES filtering
reduced order model, Fluids, 6 (2021), https://doi.org/10.3390/fluids6090302. (Cited on
p. 171)
[235] S. G LAS , A. M AYERHOFER , AND K. U RBAN, Two Ways to Treat Time in Reduced Ba-
sis Methods, Springer International Publishing, 2017, pp. 1–16, https://doi.org/10.1007/
978-3-319-58786-8_1. (Cited on pp. 83, 92)
[236] G. H. G OLUB AND C. F. VAN L OAN, Matrix Computations, third edition, Johns Hopkins University
Press, 1996. (Cited on p. 4)
[237] I. G OODFELLOW, Y. B ENGIO , AND A. C OURVILLE, Deep Learning, MIT Press, 2016, http:
//www.deeplearningbook.org. (Cited on pp. 11, 389, 390, 391, 396)
[240] W. R. G RAHAM , J. P ERAIRE , AND K. Y. TANG, Optimal control of vortex shedding us-
ing low-order models. Part I–Open-loop model development, International Journal for Numer-
ical Methods in Engineering, 44 (1999), pp. 945–972, https://doi.org/10.1002/(sici)
1097-0207(19990310)44:7<945::aid-nme537>3.0.co;2-f. (Cited on pp. 154, 155, 189)
[241] C. G RANDMONT, V. G UIMET, AND Y. M ADAY, Numerical analysis of some decoupling tech-
niques for the approximation of the unsteady fluid–structure interaction, Mathematical Models
and Methods in Applied Sciences, 11 (2001), pp. 1349–1377, https://doi.org/10.1142/
S0218202501001367. (Cited on p. 286)
[242] C. G RANDMONT AND Y. M ADAY, Existence de solutions d’un probléme de couplage fluide-
structure bidimensionnel instationnaire, Comptes Rendus de l’Académie des Sciences—Series I—
Mathematics, 326 (1998), pp. 525–530. (Cited on p. 283)
[244] C. G REIF AND K. U RBAN, Decay of the Kolmogorov N -width for wave problems, Applied Math-
ematics Letters, 96 (2019), pp. 216–222, https://doi.org/10.1016/j.aml.2019.05.013.
(Cited on p. 297)
[246] M. A. G REPL AND A. T. PATERA, A posteriori error bounds for reduced-basis approximations of
parametrized parabolic partial differential equations, ESAIM: Mathematical Modelling and Numer-
ical Analysis, 39 (2005), pp. 157–181. (Cited on p. 53)
[247] B. G RIMSTAD, SPLINTER: a library for multivariate function approximation with splines. http:
//github.com/bgrimstad/splinter, 2015. (Cited on p. 182)
[248] A. G RUBER , M. G UNZBURGER , L. J U , Y. T ENG , AND Z. WANG, Nonlinear level set learning
for function approximation on sparse data with applications to parametric differential equations,
Numerical Mathematics: Theory, Methods and Applications, 14 (2021), pp. 839–861, https://
doi.org/10.4208/nmtma.OA-2021-0062. (Cited on p. 337)
[250] M. D. G UNZBURGER, Perspectives in Flow Control and Optimization, vol. 5 of Advances in Design
and Control, SIAM, 2002, https://doi.org/10.1137/1.9780898718720. (Cited on p. 367)
[252] M. G UO AND J. S. H ESTHAVEN, Reduced order modeling for nonlinear structural analysis us-
ing Gaussian process regression, Computer Methods in Applied Mechanics and Engineering, 341
(2018), pp. 807–826, https://doi.org/10.1016/j.cma.2018.07.017. (Cited on p. 206)
Bibliography 433
[253] M. G UO AND J. S. H ESTHAVEN, Data-driven reduced order modeling for time-dependent problems,
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Computer Methods in Applied Mechanics and Engineering, 345 (2019), pp. 75–99, https://doi.
org/10.1016/j.cma.2018.10.029. (Cited on p. 187)
[256] B. H AASDONK, Convergence rates of the POD–greedy method, ESAIM: Mathematical Modelling
and Numerical Analysis, 47 (2013), pp. 859–873, https://doi.org/10.1051/m2an/2012045.
(Cited on p. 5)
[257] B. H AASDONK, Reduced basis methods for parametrized PDEs—A tutorial introduction for station-
ary and instationary problems, in Model Reduction and Approximation: Theory and Algorithms,
P. Benner, A. Cohen, M. Ohlberger, and K. Willcox, eds., Computational Science and Engineering,
ch. 2, SIAM, 2017, pp. 65–136, https://doi.org/10.1137/1.9781611974829.ch2. (Cited on
p. 1)
[258] B. H AASDONK, Mor software, in Model Order Reduction: Volume 3, ch. 13, pp. 431–460, Applica-
tions, P. Benner, S. Grivet-Talocia, A. Quarteroni, G. Rozza, W. Schilders, and L. M. Silveira, eds.,
De Gruyter, 2020, https://doi.org/10.1515/9783110499001-013. (Cited on p. 3)
[259] B. H AASDONK AND M. O HLBERGER, Adaptive basis enrichment for the reduced basis method ap-
plied to finite volume schemes, in Proceedings of the 5th International Symposium on Finite Volumes
for Complex Applications, 2008, pp. 471–478. (Cited on p. 143)
[260] B. H AASDONK AND M. O HLBERGER, Reduced basis method for finite volume approximations of
parametrized linear evolution equations, ESAIM: Mathematical Modelling and Numerical Analysis,
42 (2008), pp. 277–302, https://doi.org/10.1051/m2an:2008001. (Cited on pp. 5, 143)
[261] B. H AASDONK AND M. O HLBERGER, Reduced Basis Method for Explicit Finite Volume Approx-
imations of Nonlinear Conservation Laws, American Mathematical Society, 2009, pp. 605–614,
https://doi.org/10.1090/psapm/067.2/2605256. (Cited on p. 143)
[262] E. H ABER AND L. RUTHOTTO, Stable architectures for deep neural networks, Inverse Problems, 34
(2017), p. 014004, https://doi.org/10.1088/1361-6420/aa9a90. (Cited on p. 337)
[263] S. H AN , H. M AO , AND W. J. DALLY, Deep compression: Compressing deep neural network with
pruning, trained quantization and Huffman coding, in 4th International Conference on Learning
Representations, 2016, arXiv:1510.00149. (Cited on p. 403)
[264] D. H ARTMAN AND L. K. M ESTHA, A deep learning framework for model reduction of dynam-
ical systems, in 2017 IEEE Conference on Control Technology and Applications (CCTA), 2017,
pp. 1917–1922, https://doi.org/10.1109/CCTA.2017.8062736. (Cited on p. 298)
[265] D. H ARTMANN , M. H ERZ , AND U. W EVER, Model order reduction a key technology for dig-
ital twins, in Reduced-Order Modeling (ROM) for Simulation and Optimization, W. Keiper,
A. Milde, and S. Volkwein, eds., Springer, 2018, pp. 167–179, https://doi.org/10.1007/
978-3-319-75319-5_8. (Cited on p. 411)
[266] D. H ARTMANN AND H. VAN DER AUWERAER, Digital twins, in Progress in Industrial Math-
ematics: Success Stories, M. Cruz, C. Parés, and P. Quintela, eds., Springer, 2021, pp. 3–17,
https://doi.org/10.1007/978-3-030-61844-5_1. (Cited on pp. 389, 410)
434 Bibliography
[267] S. H AYKIN, Neural Networks and Learning Machines, vol. 10 of Neural Networks and Learning
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[271] M. H ESS , S. G RUNDEL , AND P. B ENNER, Estimating the inf-sup constant in reduced basis methods
for time-harmonic Maxwell’s equations, IEEE Transactions on Microwave Theory and Techniques,
63 (2015), pp. 1–9, https://doi.org/10.1109/TMTT.2015.2473157. (Cited on pp. 5, 7)
[272] M. W. H ESS , A. L ARIO , G. M ENGALDO , AND G. ROZZA, Reduced Order Modeling for Spectral
Element Methods: Current Developments in Nektar++ and Further Perspectives. arXiv preprint
arXiv:2201.05404, 2022. (Cited on p. 381)
[273] M. W. H ESS , A. Q UAINI , AND G. ROZZA, A spectral element reduced basis method for Navier–
Stokes equations with geometric variations, in Spectral and High Order Methods for Partial Dif-
ferential Equations ICOSAHOM 2018, S. J. Sherwin, D. Moxey, J. Peiró, P. E. Vincent, and
C. Schwab, eds., Springer International Publishing, 2020, pp. 561–571, https://doi.org/10.
1007/978-3-030-39647-3_45. (Cited on pp. 229, 231, 381)
[274] M. W. H ESS , A. Q UAINI , AND G. ROZZA, Reduced basis model order reduction for Navier-Stokes
equations in domains with walls of varying curvature, International Journal of Computational Fluid
Dynamics, 34 (2020), pp. 119–126, https://doi.org/10.1080/10618562.2019.1645328.
(Cited on p. 381)
[275] M. W. H ESS , A. Q UAINI , AND G. ROZZA, A comparison of reduced-order modeling approaches us-
ing artificial neural networks for PDEs with bifurcating solutions, ETNA—Electronic Transactions
on Numerical Analysis, 56 (2022), pp. 52–65, https://doi.org/10.1553/etna_vol56s52.
(Cited on p. 381)
[276] M. W. H ESS AND G. ROZZA, A spectral element reduced basis method in parametric CFD, in
Numerical Mathematics and Advanced Applications—ENUMATH 2017, F. A. Radu, K. Kumar,
I. Berre, J. M. Nordbotten, and I. S. Pop, eds., vol. 126, Springer, 2019, pp. 693–701, https:
//doi.org/10.1007/978-3-319-96415-7_64. (Cited on pp. 231, 381)
[277] M. W. H ESS AND G. ROZZA, Model Reduction Using Sparse Polynomial Interpolation for the
Incompressible Navier-Stokes Equations. arXiv preprint arXiv:2201.03228, 2022. (Cited on p. 381)
[278] J. H ESTHAVEN , B. S TAMM , AND S. Z HANG, Efficient greedy algorithms for high-dimensional
parameter spaces with applications to empirical interpolation and reduced basis methods, ESAIM
Mathematical Modelling and Numerical Analysis, 48 (2014), https://doi.org/10.1051/m2an/
2013100. (Cited on p. 5)
[279] J. H ESTHAVEN AND S. U BBIALI, Non-intrusive reduced order modeling of nonlinear problems
using neural networks, Journal of Computational Physics, 363 (2018), pp. 55–78, https://doi.
org/10.1016/j.jcp.2018.02.037. (Cited on p. 397)
Bibliography 435
[280] J. H ESTHAVEN AND T. WARBURTON, Nodal high-order methods on unstructured grids: I. Time-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
domain solution of Maxwell’s equations, Journal of Computational Physics, 181 (2002), pp. 186–
221, https://doi.org/10.1006/jcph.2002.7118. (Cited on p. 235)
[281] J. S. H ESTHAVEN , G. ROZZA , AND B. S TAMM, Certified Reduced Basis Methods for Parametrized
Partial Differential Equations, Springer Briefs in Mathematics, 1 ed., Springer, 2015, https://
doi.org/10.1007/978-3-319-22470-1. (Cited on pp. 1, 5, 15, 83, 86, 87, 92, 102, 106, 108,
205, 251, 253, 254, 267, 327, 347, 367, 368)
[283] R. M. H ICKS AND P. A. H ENNE, Wing design by numerical optimization, Journal of Aircraft, 15
(1978), pp. 407–412, https://doi.org/10.2514/3.58379. (Cited on p. 219)
[284] S. H IJAZI , S. A LI , G. S TABILE , F. BALLARIN , AND G. ROZZA, The effort of increasing Reynolds
number in projection-based reduced order methods: From laminar to turbulent flows, in FEF 2017
Selected Contributions, H. van Brummelen, A. Corsini, S. Perotto, and G. Rozza, eds., vol. 132 of
Lecture Notes in Computational Science and Engineering, Springer International Publishing, 2020,
pp. 245–264, https://doi.org/10.1007/978-3-030-30705-9_22. (Cited on pp. 175, 381)
[285] S. H IJAZI , G. S TABILE , A. M OLA , AND G. ROZZA, Non-intrusive polynomial chaos method ap-
plied to full-order and reduced problems in computational fluid dynamics: A comparison and per-
spectives, in QUIET Selected Contributions, M. D’Elia, M. Gunzburger, and G. Rozza, eds., vol. 137
of Lecture Notes in Computational Science and Engineering, Springer International Publishing,
2020, pp. 217–240, https://doi.org/10.1007/978-3-030-48721-8_10. (Cited on pp. 154,
251)
[286] S. H IJAZI , G. S TABILE , A. M OLA , AND G. ROZZA, Data-driven POD–Galerkin reduced order
model for turbulent flows, Journal of Computational Physics, 416 (2020), p. 109513, https://
doi.org/10.1016/j.jcp.2020.109513. (Cited on pp. 175, 205)
[287] G. H INTON , O. V INYALS , AND J. D EAN, Distilling the Knowledge in a Neural Network, in NIPS
Deep Learning and Representation Learning Workshop, 2015. (Cited on p. 406)
[288] G. E. H INTON AND R. Z EMEL, Autoencoders, minimum description length and Helmholtz free
energy, in Advances in Neural Information Processing Systems, vol. 6, J. Cowan, G. Tesauro, and
J. Alspector, eds., Morgan-Kaufmann, 1994, pp. 3–10. (Cited on p. 396)
[289] M. H INZE , M. K ÖSTER , AND S. T UREK, A space-time multigrid method for optimal flow control,
in Constrained Optimization and Optimal Control for Partial Differential Equations, G. Leugering, S.
Engell, A. Griewank, M. Hinze, R. Rannacher, V. Schulz, M. Ulbrich, and S. Ulbrich, ed., Springer,
2012, pp. 147–170, https://doi.org/10.1007/978-3-0348-0133-1_8. (Cited on p. 89)
[290] M. H INZE , R. P INNAU , M. U LBRICH , AND S. U LBRICH, Optimization with PDE constraints,
vol. 23 of Mathematical Modelling: Theory and Applications, Springer Netherlands, 2009, https:
//doi.org/10.1007/978-1-4020-8839-1. (Cited on pp. 84, 85, 89)
[291] P. H OLMES , J. L. L UMLEY, AND G. B ERKOOZ, Turbulence, Coherent Structures, Dynamical Sys-
tems and Symmetry, Cambridge Monographs on Mechanics, Cambridge University Press, 1996,
https://doi.org/10.1017/CBO9780511622700. (Cited on p. 373)
[292] M. H OLTZ, Sparse Grid Quadrature in High Dimensions with Applications in Finance and Insur-
ance, vol. 77, Springer, 2010. (Cited on p. 256)
[293] K. H ORNIK , M. S TINCHCOMBE , AND H. W HITE, Multilayer feedforward networks are uni-
versal approximators, Neural Networks, 2 (1989), pp. 359–366, https://doi.org/10.1016/
0893-6080(89)90020-8. (Cited on pp. 389, 397)
436 Bibliography
method for wave propagation problems, Journal of Computational Physics, 151 (1999), pp. 921–
946, https://doi.org/10.1006/jcph.1999.6227. (Cited on p. 234)
[295] Y. H UANG AND Y. C HEN, Survey of state-of-art autonomous driving technologies with deep learn-
ing, in 2020 IEEE 20th International Conference on Software Quality, Reliability and Security Com-
panion (QRS-C), 2020, pp. 221–228, https://doi.org/10.1109/QRS-C51114.2020.00045.
(Cited on p. 389)
[297] T. J. H UGHES , L. P. F RANCA , AND M. BALESTRA, A new finite element formulation for computa-
tional fluid dynamics: V. Circumventing the Babuška-Brezzi condition: A stable Petrov-Galerkin
formulation of the Stokes problem accommodating equal-order interpolations, Computer Meth-
ods in Applied Mechanics and Engineering, 59 (1986), pp. 85–99, https://doi.org/10.1016/
0045-7825(87)90093-4. (Cited on p. 128)
[298] T. J. H UGHES , L. P. F RANCA , AND G. M. H ULBERT, A new finite element formulation for
computational fluid dynamics: VIII. The Galerkin/least-squares method for advective-diffusive
equations, Computer Methods in Applied Mechanics and Engineering, 73 (1989), pp. 173–189,
https://doi.org/10.1016/0045-7825(89)90111-4. (Cited on pp. 130, 132)
[299] T. J. R. H UGHES , A. A. O BERAI , AND L. M AZZEI, Large eddy simulation of turbulent channel
flows by the variational multiscale method, Physics of Fluids, 13 (2001), pp. 1784–1799, https:
//doi.org/10.1063/1.1367868. (Cited on p. 132)
[300] J. H UNT, A. W RAY, AND P. M OIN, Eddies Stream and Convergence Zones in Turbulent Flows,
tech. Report CTR-S88, CTR report, 1988. (Cited on p. 170)
[303] A. I OLLO AND D. L OMBARDI, Advection modes by optimal mass transfer, Physical Review E, 89
(2014). (Cited on p. 298)
[304] R. I SSA, Solution of the implicitly discretised fluid flow equations by operator-splitting, Journal
of Computational Physics, 62 (1986), pp. 40–65, https://doi.org/10.1016/0021-9991(86)
90099-9. (Cited on p. 146)
[305] J. F. G ERBEAU AND M. V IDRASCU, A quasi-Newton algorithm based on a reduced model for
fluid–structure interaction problems in blood flows, ESAIM: Mathematical Modelling and Numerical
Analysis, 37 (2003), pp. 631–647, https://doi.org/10.1051/m2an:2003049. (Cited on p. 286)
[306] J. JANAI , F. G ÜNEY, A. B EHL , AND A. G EIGER, Computer vision for autonomous vehicles: Prob-
lems, datasets and state-of-the-art, Foundations and Trends in Computer Graphics and Vision, 12
(2020), pp. 1–308, https://doi.org/10.1561/0600000079. (Cited on p. 389)
[307] A. JANON , M. N ODET, AND C. P RIEUR, Certified reduced-basis solutions of viscous Burgers equa-
tion parametrized by initial and boundary values, ESAIM: Mathematical Modelling and Numerical
Analysis, 47 (2013), pp. 317–348, https://doi.org/10.1051/m2an/2012029. (Cited on p. 155)
Bibliography 437
[308] C. JANYA -A NURAK, Framework for Analysis and Identification of Nonlinear Distributed Param-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
eter Systems using Bayesian Uncertainty Quantification Based on Generalized Polynomial Chaos,
vol. 31, KIT Scientific Publishing, 2017, https://doi.org/10.5445/KSP/1000066940. (Cited
on p. 405)
[309] H. JASAK, Error Analysis and Estimation for the Finite Volume Method with Applications to Fluid
Flows, PhD thesis, Imperial College, University of London, 1996. (Cited on p. 146)
[312] C. J OHNSON , U. N ÄVERT, AND J. P ITKÄRANTA, Finite element methods for linear hyperbolic
problems, Computer Methods in Applied Mechanics and Engineering, 45 (1984), pp. 285–312,
https://doi.org/10.1016/0045-7825(84)90158-0. (Cited on p. 127)
[313] H. J OHNSTON AND J.-G. L IU, Accurate, stable and efficient Navier–Stokes solvers based on explicit
treatment of the pressure term, Journal of Computational Physics, 199 (2004), pp. 221–259, https:
//doi.org/10.1016/j.jcp.2004.02.009. (Cited on pp. 152, 180)
[314] W. J ONES AND B. L AUNDER, The prediction of laminarization with a two-equation model of
turbulence, International Journal of Heat and Mass Transfer, 15 (1972), pp. 301–314, https:
//doi.org/10.1016/0017-9310(72)90076-2. (Cited on p. 168)
[316] M. K AHLBACHER AND S. VOLKWEIN, Galerkin proper orthogonal decomposition methods for
parameter dependent elliptic systems, Discussiones Mathematicae, Differential Inclusions, Control
and Optimization, 27 (2007), pp. 95–117, http://eudml.org/doc/271156. (Cited on p. 107)
[317] I. K ALASHNIKOVA AND S. A RUNAJATESAN, A stable Galerkin reduced order model (ROM) for
compressible flow, in 10th World Congress for Computational Mechanics (WCCM), 2012, pp. 2012–
18407, https://doi.org/10.5151/meceng-wccm2012-18407. (Cited on pp. 242, 243)
[318] I. K ALASHNIKOVA AND M. BARONE, On the stability and convergence of a Galerkin reduced order
model (ROM) of compressible flow with solid wall and far-field boundary treatment, International
Journal for Numerical Methods in Engineering, 83 (2010), pp. 1345–1375, https://doi.org/10.
1002/nme.2867. (Cited on pp. 155, 243)
[322] E. N. K ARATZAS , F. BALLARIN , AND G. ROZZA, Projection-based reduced order models for a
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
cut finite element method in parametrized domains, Computers & Mathematics with Applications,
79 (2020), pp. 833–851, https://doi.org/10.1016/j.camwa.2019.08.003. (Cited on pp. 265,
280, 298)
[323] E. N. K ARATZAS , M. N ONINO , F. BALLARIN , AND G. ROZZA, A reduced order cut finite element
method for geometrically parameterized steady and unsteady Navier-Stokes problems, Computers &
Mathematics with Applications, 116 (2022), pp. 140–160, https://doi.org/10.1016/j.camwa.
2021.07.016. (Cited on p. 265)
[324] E. N. K ARATZAS AND G. ROZZA, A reduced order model for a stable embedded boundary
parametrized Cahn–Hilliard phase-field system based on cut finite elements, Journal of Scientific
Computing, 89 (2021), https://doi.org/10.1007/s10915-021-01623-8. (Cited on pp. 37,
265, 281)
[328] M. K ÄRCHER AND M. A. G REPL, A certified reduced basis method for parametrized elliptic
optimal control problems, ESAIM: Control, Optimisation and Calculus of Variations, 20 (2014),
pp. 416–441, https://doi.org/10.1051/cocv/2013069. (Cited on pp. 83, 87, 262)
[329] M. K ÄRCHER , Z. T OKOUTSI , M. A. G REPL , AND K. V EROY, Certified reduced basis methods
for parametrized elliptic optimal control problems with distributed controls, Journal of Scientific
Computing, 75 (2018), pp. 276–307, https://doi.org/10.1007/s10915-017-0539-z. (Cited
on pp. 83, 87, 262)
[330] G. K ARNIADAKIS AND S. S HERWIN, Spectral/hp element methods for CFD, Numerical Mathe-
matics and Scientific Computation, Oxford University Press, 2005, https://doi.org/10.1093/
acprof:oso/9780198528692.001.0001. (Cited on pp. 223, 227, 373)
[333] K. K EAN AND M. S CHNEIER, Error analysis of supremizer pressure recovery for POD based
reduced-order models of the time-dependent Navier–Stokes equations, SIAM Journal on Numeri-
cal Analysis, 58 (2020), pp. 2235–2264, https://doi.org/10.1137/19m128702x. (Cited on
p. 144)
Bibliography 439
[334] H. B. K ELLER, Lectures on numerical methods in bifurcation problems, Applied Mathematics, 217
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[335] M. C. K ENNEDY AND A. O’H AGAN, Predicting the output from a complex computer code when
fast approximations are available, Biometrika, 87 (2000), pp. 1–13, https://doi.org/10.1093/
biomet/87.1.1. (Cited on p. 334)
[336] P. K ENNICOTT, IGES/PDES O RGANIZATION , AND U.S. P RODUCT DATA A SSOCIATION, Initial
Graphics Exchange Specification, IGES 5.3, U.S. Product Data Association, 1996. (Cited on p. 357)
[337] M. K HAMLICH , F. P ICHI , AND G. ROZZA, Model Order Reduction for Bifurcating Phenomena in
Fluid-Structure Interaction Problems, 2021, https://arxiv.org/abs/2110.06297. (Cited on
pp. 98, 312, 317)
[338] A. K HAN , A. S OHAIL , U. Z AHOORA , AND A. S. Q URESHI, A survey of the recent architectures
of deep convolutional neural networks, Artificial Intelligence Review, 53 (2020), pp. 5455–5516,
https://doi.org/10.1007/s10462-020-09825-6. (Cited on p. 402)
[339] S. K HAN AND E. G UNPINAR, Sampling CAD models via an extended teaching–learning-based
optimization technique, Computer-Aided Design, 100 (2018), pp. 52–67, https://doi.org/10.
1016/j.cad.2018.03.003. (Cited on p. 204)
[340] D. K HATAMSAZ AND D. L. A LLAIRE, Materials design using an active subspace batch Bayesian
optimization approach, in AIAA SCITECH 2022 Forum, 2022, p. 0075, https://doi.org/10.
2514/6.2022-0075. (Cited on p. 340)
[342] H. K IELHÖFER, Bifurcation Theory: An Introduction with Applications to PDEs, Applied Mathe-
matical Sciences, Springer New York, 2006. (Cited on p. 101)
[344] D. P. K INGMA AND J. BA, Adam: A Method for Stochastic Optimization, arXiv preprint
arXiv:1412.6980 (2014). (Cited on p. 339)
[346] D. J. K NEZEVIC AND A. T. PATERA, A certified reduced basis method for the Fokker–Planck equa-
tion of dilute polymeric fluids: FENE dumbbells in extensional flow, SIAM Journal on Scientific
Computing, 32 (2010), pp. 793–817, https://doi.org/10.1137/090759239. (Cited on p. 5)
[349] A. K RIZHEVSKY AND G. H INTON, Learning Multiple Layers of Features from Tiny Images, Mas-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
ter’s thesis, Department of Computer Science, University of Toronto (2009). (Cited on p. 407)
[350] A. K RIZHEVSKY, I. S UTSKEVER , AND G. E. H INTON, Imagenet classification with deep convolu-
tional neural networks, Advances in Neural Information Processing Systems, 25 (2012), pp. 1097–
1105. (Cited on pp. 389, 402)
[351] K. K UNISCH AND S. VOLKWEIN, Galerkin proper orthogonal decomposition methods for a general
equation in fluid dynamics, SIAM Journal on Numerical Analysis, 40 (2002), pp. 492–515, https:
//doi.org/10.1137/s0036142900382612. (Cited on p. 284)
[352] K. K UNISCH AND S. VOLKWEIN, Proper orthogonal decomposition for optimality systems,
ESAIM: Mathematical Modelling and Numerical Analysis, 42 (2008), pp. 1–23, https://doi.
org/10.1051/m2an:2007054. (Cited on pp. 83, 367)
[355] Y. K UZNETSOV, Elements of Applied Bifurcation Theory, Applied Mathematical Sciences, Springer
New York, 2004. (Cited on p. 101)
[357] P. L ASAINT AND P. R AVIART, On a finite element method for solving the neutron transport equation,
Mathematical aspects of finite elements in partial differential equations, in C. de Boor, ed., New York,
1974, pp. 89–123. (Cited on p. 234)
[360] T. L ASSILA , A. M ANZONI , A. Q UARTERONI , AND G. ROZZA, Model order reduction in fluid dy-
namics: Challenges and perspectives, in Reduced Order Methods for Modeling and Computational
Reduction, A. Quarteroni and G. Rozza, eds., vol. 9 of MS&A series, Springer, 2014, pp. 235–273,
https://doi.org/10.1007/978-3-319-02090-7_9. (Cited on p. 7)
[361] B. L AUNDER AND D. S PALDING, The numerical computation of turbulent flows, Computer Meth-
ods in Applied Mechanics and Engineering, 3 (1974), pp. 269–289, https://doi.org/10.1016/
0045-7825(74)90029-2. (Cited on p. 168)
[362] W. L AYTON , L. R EBHOLZ , AND C. T RENCHEA, Modular nonlinear filter stabilization of methods
for higher Reynolds numbers flow, Journal of Mathematical Fluid Mechanics, 14 (2012), pp. 325–
354, https://doi.org/10.1007/s00021-011-0072-z. (Cited on pp. 170, 171)
[363] D. L AZZARO AND L. B. M ONTEFUSCO, Radial basis functions for the multivariate interpolation of
large scattered data sets, Journal of Computational and Applied Mathematics, 140 (2002), pp. 521–
536, https://doi.org/10.1016/s0377-0427(01)00485-x. (Cited on p. 172)
Bibliography 441
[364] S. L E C LAINCHE AND J. M. V EGA, Higher order dynamic mode decomposition, SIAM Journal on
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[365] P. L E TALLEC AND J. M OURO, Fluid structure interaction with large structural displacements,
Computer Methods in Applied Mechanics and Engineering, 190 (2001), pp. 3039–3067, https:
//doi.org/10.1016/S0045-7825(00)00381-9. (Cited on p. 286)
[366] Y. L E C UN AND Y. B ENGIO, Convolutional Networks for Images, Speech, and Time Series,
vol. 3361, MIT Press, 1995, p. 1995. (Cited on p. 394)
[368] M.-Y. L. L EE, Estimation of the error in the reduced basis method solution of differential algebraic
equation systems, SIAM Journal on Numerical Analysis, 28 (1991), pp. 512–528, https://doi.
org/10.1137/0728028. (Cited on p. 3)
[369] P. L E G RESLEY AND J. A LONSO, Investigation of Non-linear Projection for POD Based Reduced
Order Models for Aerodynamics, in 39th Aerospace Sciences Meeting and Exhibit, American Insti-
tute of Aeronautics and Astronautics, Reno, NV, U.S.A., Jan. 2001, https://doi.org/10.2514/
6.2001-926. (Cited on p. 144)
[370] S. L EIJNEN AND F. V. V EEN, The neural network zoo, in Multidisciplinary Digital Publishing In-
stitute Proceedings, vol. 47, 2020, p. 9, https://doi.org/10.3390/proceedings2020047009.
(Cited on p. 393)
[371] J. L ERAY, Sur le mouvement d’un fluide visqueux emplissant l’espace, Acta Mathematica, 63 (1934),
pp. 193–248, https://doi.org/10.1007/BF02547354. (Cited on p. 170)
[372] A. L EWIS , R. S MITH , AND B. W ILLIAMS, Gradient free active subspace construction using Morris
screening elementary effects, Computers & Mathematics with Applications, 72 (2016), pp. 1603–
1615, https://doi.org/10.1016/j.camwa.2016.07.022. (Cited on p. 337)
[374] J. L. L IONS, Optimal Control of System Governed by Partial Differential Equations, vol. 170 of
Grundlehren der mathematischen Wissenschaften, Springer-Verlag Berlin Heidelberg, 1971. (Cited
on p. 84)
[375] H. L IU, Optimal error estimates of the direct discontinuous Galerkin method for convection-diffusion
equations, Mathematics of Computation, 84 (2015), pp. 2263–2295, https://doi.org/10.1090/
S0025-5718-2015-02923-8. (Cited on p. 239)
[376] J.-G. L IU , J. L IU , AND R. L. P EGO, Stable and accurate pressure approximation for unsteady
incompressible viscous flow, Journal of Computational Physics, 229 (2010), pp. 3428–3453, https:
//doi.org/10.1016/j.jcp.2010.01.010. (Cited on pp. 152, 180)
[377] Z. L IU , M. S UN , T. Z HOU , G. H UANG , AND T. DARRELL, Rethinking the Value of Network Prun-
ing, in International Conference on Learning Representations, 2019. (Cited on p. 403)
[378] A. L OGG , K. M ARDAL , AND G. W ELLS, Automated Solution of Differential Equations by the Finite
Element Method, Springer-Verlag, Berlin, 2012. (Cited on p. 15)
442 Bibliography
boats: Dynamics, FSI, and shape optimization, in Variational Analysis and Aerospace Engineering:
Mathematical Challenges for Aerospace Design, Springer, 2012, pp. 339–377, https://doi.org/
10.1007/978-1-4614-2435-2_15. (Cited on pp. 283, 347)
[380] S. L ORENZI , A. C AMMI , L. L UZZI , AND G. ROZZA, POD-Galerkin method for finite volume
approximation of Navier–Stokes and RANS equations, Computer Methods in Applied Mechanics
and Engineering, 311 (2016), pp. 151–179, https://doi.org/10.1016/j.cma.2016.08.006.
(Cited on pp. 143, 144, 150, 155, 181)
[381] L. L OVÁSZ AND S. V EMPALA, Hit-and-run from a corner, SIAM Journal on Computing, 35 (2006),
pp. 985–1005, https://doi.org/10.1137/S009753970544727X. (Cited on p. 329)
[384] Y. M ADAY, Analysis of coupled models for fluid-structure interaction of internal flows, in Cardio-
vascular Mathematics: Modeling and simulation of the circulatory system, L. Formaggia, A. Quar-
teroni, and A. Veneziani, eds., Springer Milan, 2009, pp. 279–306, https://doi.org/10.1007/
978-88-470-1152-6_8. (Cited on p. 283)
[386] Y. M ADAY, A. T. PATERA , AND G. T URINICI, Global a priori convergence theory for reduced-
basis approximations of single-parameter symmetric coercive elliptic partial differential equa-
tions, Comptes Rendus Mathematique, 335 (2002), pp. 289–294, https://doi.org/10.1016/
S1631-073X(02)02466-4. (Cited on p. 32)
[387] A. M ANZONI, An efficient computational framework for reduced basis approximation and a posteri-
ori error estimation of parametrized Navier-Stokes flows, ESAIM: Mathematical Modelling and Nu-
merical Analysis, 48 (2014), pp. 1199–1226, https://doi.org/10.1051/m2an/2014013. (Cited
on pp. 5, 64, 131)
[388] A. M ANZONI AND F. N EGRI, Heuristic strategies for the approximation of stability factors in
quadratically nonlinear parametrized PDEs, Advances in Computational Mathematics, 41 (2015),
pp. 1255–1288, https://doi.org/10.1007/s10444-015-9413-4. (Cited on p. 5)
[390] A. M ANZONI , A. Q UARTERONI , AND G. ROZZA, Model reduction techniques for fast blood flow
simulation in parametrized geometries, International Journal for Numerical Methods in Biomedical
Engineering, 28 (2012), pp. 604–625, https://doi.org/10.1002/cnm.1465. (Cited on pp. 11,
349)
[391] J. E. M ATSSON, An Introduction to ANSYS Fluent 2021, SDC Publications, 2021. (Cited on p. 143)
[392] L. R. M EDSKER AND L. JAIN, Recurrent neural networks, Design and Applications, 5 (2001),
pp. 64–67, https://doi.org/10.1201/9781420049176. (Cited on p. 393)
Bibliography 443
[393] J. M. M ELENK, On n-widths for elliptic problems, Journal of Mathematical Analysis and Applica-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[394] L. M ENEGHETTI , N. D EMO , AND G. ROZZA, A Dimensionality Reduction Approach for Convo-
lutional Neural Networks, 2021, https://arxiv.org/abs/2110.09163. (Cited on pp. 12, 343,
384, 389, 403, 407)
[395] F. R. M ENTER, Two-equation eddy-viscosity turbulence models for engineering applications, AIAA
Journal, 32 (1994), pp. 1598–1605, https://doi.org/10.2514/3.12149. (Cited on p. 168)
[396] C. A. M ICCHELLI, Interpolation of scattered data: Distance matrices and conditionally positive
definite functions, Constructive Approximation, 2 (1986), pp. 11–22, https://doi.org/10.1007/
bf01893414. (Cited on p. 172)
[400] L. M IRSKY, Symmetric gauge functions and unitarily invariant norms, Quarterly Journal of Mathe-
matics, 11 (1960), pp. 50–59, https://doi.org/10.1093/qmath/11.1.50. (Cited on p. 4)
[401] A. M OELGAARD, PMM-Tests with a Model of a Frigate Class DDG-51, tech. Report 01, University
of Zurich, Department of Informatics, 2000. (Cited on p. 358)
[402] M. M OHEBUJJAMAN , L. G. R EBHOLZ , X. X IE , AND T. I LIESCU, Energy balance and mass con-
servation in reduced order models of fluid flows, Journal of Computational Physics, 346 (2017),
pp. 262–277, https://doi.org/10.1016/j.jcp.2017.06.019. (Cited on p. 187)
[405] H. M ONTAZERI AND B. B LOCKEN, CFD simulation of wind-induced pressure coefficients on build-
ings with and without balconies: Validation and sensitivity analysis, Building and Environment, 60
(2013), pp. 137–149, https://doi.org/10.1016/j.buildenv.2012.11.012. (Cited on p. 168)
[407] A. M ORRIS , C. A LLEN , AND T. R ENDALL, CFD-based optimization of aerofoils using radial ba-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
sis functions for domain element parameterization and mesh deformation, International Journal for
Numerical Methods in Fluids, 58 (2008), pp. 827–860, https://doi.org/10.1002/fld.1769.
(Cited on p. 349)
[408] F. M OUKALLED , L. M ANGANI , AND M. DARWISH, The Finite Volume Method in Computational
Fluid Dynamics: An Advanced Introduction with OpenFOAM and Matlab, 1st ed., Springer, 2015,
https://doi.org/10.1007/978-3-319-16874-6. (Cited on pp. 143, 144, 146)
[410] C. M UREA AND S. S Y, A fast method for solving fluid–structure interaction problems numerically,
International Journal for Numerical Methods in Fluids, 60 (2009), pp. 1149–1172, https://doi.
org/10.1002/fld.1931. (Cited on p. 303)
[411] N. J. NAIR AND M. BALAJEWICZ, Transported snapshot model order reduction approach for para-
metric, steady-state fluid flows containing parameter-dependent shocks, International Journal for Nu-
merical Methods in Engineering, 117 (2019), pp. 1234–1262, https://doi.org/10.1002/nme.
5998. (Cited on p. 298)
[412] J. N E ČAS, Les méthodes directes en théorie des équations elliptiques, Masson, 1967. (Cited on
p. 52)
[413] F. N EGRI , A. M ANZONI , AND G. ROZZA, Reduced basis approximation of parametrized optimal
flow control problems for the Stokes equations, Computers & Mathematics with Applications, 69
(2015), pp. 319–336, https://doi.org/10.1016/j.camwa.2014.12.010. (Cited on pp. 83, 87,
367)
[414] F. N EGRI , G. ROZZA , A. M ANZONI , AND A. Q UARTERONI, Reduced basis method for
parametrized elliptic optimal control problems, SIAM Journal on Scientific Computing, 35 (2013),
pp. A2316–A2340, https://doi.org/10.1137/120894737. (Cited on pp. 83, 87, 262, 367)
[415] N. C. N GUYEN , G. ROZZA , AND A. T. PATERA, Reduced basis approximation and a posteriori
error estimation for the time-dependent viscous Burgers’ equation, Calcolo, 46 (2009), pp. 157–185,
https://doi.org/10.1007/s10092-009-0005-x. (Cited on p. 292)
[416] S. A. N IEDERER , M. S. S ACKS , M. G IROLAMI , AND K. W ILLCOX, Scaling digital twins from the
artisanal to the industrial, Nature Computational Science, 1 (2021), pp. 313–320, https://doi.
org/10.1038/s43588-021-00072-5. (Cited on pp. 12, 413)
[417] K. N ODA , H. A RIE , Y. S UGA , AND T. O GATA, Multimodal integration learning of robot behavior
using deep neural networks, Robotics and Autonomous Systems, 62 (2014), pp. 721–736, https:
//doi.org/10.1016/j.robot.2014.03.003. (Cited on p. 402)
[418] M. N ONINO, On the Application of the Reduced Basis Method to Fluid-Structure Interaction Prob-
lems, PhD thesis, Scuola Internazionale Superiore di Studi Avanzati, Trieste, 2020. (Cited on p. 311)
[419] M. N ONINO , F. BALLARIN , G. ROZZA , AND Y. M ADAY, Overcoming Slowly Decaying Kol-
mogorov n-Width by Transport Maps: Application to Model Order Reduction of Fluid Dynamics
and Fluid–Structure Interaction Problems. arXiv:1911.06598, 2019. (Cited on p. 399)
[420] M. N ONINO , F. BALLARIN , G. ROZZA , AND Y. M ADAY, Projection Based Semi–Implicit Parti-
tioned Reduced Basis Method for Non Parametrized and Parametrized Fluid–Structure Interaction
Problems. arXiv:2201.03236, 2022. (Cited on p. 312)
[421] A. K. N OOR, Reduced basis technique for nonlinear analysis of structures, AIAA Journal, 18
(1980), pp. 455–462, https://doi.org/10.2514/3.50778. (Cited on p. 3)
Bibliography 445
[422] A. K. N OOR, On making large nonlinear problems small, Computer Methods in Applied Me-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[438] S. PATANKAR AND D. S PALDING, A calculation procedure for heat, mass and momentum transfer
in three-dimensional parabolic flows, International Journal of Heat and Mass Transfer, 15 (1972),
pp. 1787–1806, https://doi.org/10.1016/0017-9310(72)90054-3. (Cited on pp. 144, 146)
[439] A. PATERA, A spectral element method for fluid dynamics: Laminar flow in a channel expan-
sion, Journal of Computational Physics, 54 (1984), pp. 468–488, https://doi.org/10.1016/
0021-9991(84)90128-1. (Cited on p. 223)
[440] A. PATERA AND G. ROZZA, Reduced Basis Approximation and a Posteriori Error Estimation for
Parametrized Partial Differential Equation, MIT Pappalardo Monographs in Mechanical Engineer-
ing, Copyright MIT (2007–2010). (Cited on p. 106)
[441] W. PAZNER AND P.-O. P ERSSON, Stage-parallel fully implicit Runge–Kutta solvers for discon-
tinuous Galerkin fluid simulations, Journal of Computational Physics, 335 (2017), pp. 700–717,
https://doi.org/10.1016/j.jcp.2017.01.050. (Cited on p. 234)
[442] B. P EHERSTORFER, Model reduction for transport-dominated problems via online adaptive bases
and adaptive sampling, SIAM Journal on Scientific Computing, 42 (2020), pp. A2803–A2836,
https://doi.org/10.1137/19M1257275. (Cited on pp. 298, 399)
[444] Z. P ENG , M. WANG , AND F. L I, A Learning-Based Projection Method for Model Order Reduction
of Transport Problems, arXiv:2105.14633 [cs, math] (2021). (Cited on p. 399)
[446] F. P ICHI, Reduced Order Models for Parametric Bifurcation Problems in Nonlinear PDEs, PhD
thesis, Scuola Internazionale Superiore di Studi Avanzati, Trieste. (Cited on pp. 25, 97, 311, 316)
[447] F. P ICHI , F. BALLARIN , G. ROZZA , AND J. S. H ESTHAVEN, An Artificial Neural Network Ap-
proach to Bifurcating Phenomena in Computational Fluid Dynamics (2021). arXiv:2109.10765.
(Cited on pp. 11, 97, 98, 119)
[448] F. P ICHI , F. BALLARIN , G. ROZZA , AND J. S. H ESTHAVEN, Artificial neural network for bifur-
cating phenomena modelled by nonlinear parametrized PDEs, PAMM, 20 (2021), p. e202000350,
https://doi.org/10.1002/pamm.202000350. (Cited on pp. 98, 206, 311)
[449] F. P ICHI , J. E FTANG , G. ROZZA , AND A. T. PATERA, Reduced Order Models for the Buckling of
Hyperelastic Beams. MIT-FVG “ROM2S" report, 2020. (Cited on pp. 97, 115)
[450] F. P ICHI , A. Q UAINI , AND G. ROZZA, A reduced order modeling technique to study bifurcating
phenomena: Application to the Gross–Pitaevskii equation, SIAM Journal on Scientific Computing,
42 (2020), pp. B1115–B1135, https://doi.org/10.1137/20M1313106. (Cited on pp. 97, 112,
316)
Bibliography 447
[451] F. P ICHI AND G. ROZZA, Reduced basis approaches for parametrized bifurcation problems held
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
by non-linear Von Kármán equations, Journal of Scientific Computing, 81 (2019), pp. 112–135,
https://doi.org/10.1007/s10915-019-01003-3. (Cited on pp. 97, 104, 109)
[452] F. P ICHI , M. S TRAZZULLO , F. BALLARIN , AND G. ROZZA, Driving Bifurcating Parametrized
Nonlinear PDEs by Optimal Control Strategies: Application to Navier-Stokes Equations with Model
Order Reduction. arXiv:2010.13506, 2020. (Cited on pp. 25, 89, 95, 97, 98, 104, 119, 311)
[453] A. P INKUS, Ridge Functions, vol. 205, Cambridge University Press, 2015, https://doi.org/10.
1017/CBO9781316408124. (Cited on p. 328)
[454] M. P INTORE , F. P ICHI , M. H ESS , G. ROZZA , AND C. C ANUTO, Efficient computation of bi-
furcation diagrams with a deflated approach to reduced basis spectral element method, Advances
in Computational Mathematics, 47 (2020), https://doi.org/10.1007/s10444-020-09827-6.
(Cited on pp. 25, 98, 114, 119, 229, 372, 373, 381)
[455] S. P IPERNO AND C. FARHAT, Design of efficient partitioned procedures for the transient solution
of aeroelastic problems, Revue Européenne des Éléments Finis, 9 (2000), pp. 655–680, https:
//doi.org/10.1080/12506559.2000.10511480. (Cited on p. 286)
[456] L. P. P ITAEVSKII AND S. S TRINGARI, Bose-Einstein Condensation, International Series of Mono-
graphs on Physics, Oxford University Press, 2003, https://books.google.it/books?id=
rIobbOxC4j4C. (Cited on p. 112)
[457] G. P ITTON , A. Q UAINI , AND G. ROZZA, Computational reduction strategies for the detection of
steady bifurcations in incompressible fluid-dynamics: Applications to Coanda effect in cardiology,
Journal of Computational Physics, 344 (2017), pp. 534–557, https://doi.org/10.1016/j.jcp.
2017.05.010. (Cited on pp. 229, 311, 372, 373, 374)
[458] G. P ITTON AND G. ROZZA, On the application of reduced basis methods to bifurcation problems
in incompressible fluid dynamics, Journal of Scientific Computing, 73 (2017), pp. 157–177, https:
//doi.org/10.1007/s10915-017-0419-6. (Cited on p. 229)
[459] T. P ORSCHING, Estimation of the error in the reduced basis method solution of nonlinear
equations, Mathematics of Computation, 45 (1985), pp. 487–496, https://doi.org/10.1090/
S0025-5718-1985-0804937-0. (Cited on p. 3)
[460] T. A. P ORSCHING AND M. L. L EE, The reduced basis method for initial value problems, SIAM
Journal on Numerical Analysis, 24 (1987), pp. 1277–1287, https://doi.org/10.1137/0724083.
(Cited on p. 3)
[461] M. P OŠTA AND T. ROUBÍ ČEK, Optimal control of Navier–Stokes equations by Oseen approxima-
tion, Computers & Mathematics with Applications, 53 (2007), pp. 569–581, https://doi.org/
10.1016/j.camwa.2006.02.034. (Cited on p. 83)
[462] L. P RANDTL, Uber die ausgebildete Turbulenz, vol. 5, ZAMM, 1925. (Cited on p. 167)
[463] J. L. P ROCTOR , S. L. B RUNTON , AND J. N. K UTZ, Dynamic mode decomposition with control,
SIAM Journal on Applied Dynamical Systems, 15 (2016), pp. 142–161, https://doi.org/10.
1137/15M1013857. (Cited on pp. 209, 210)
[464] C. P RUD ’ HOMME , D. V. ROVAS , K. V EROY, L. M ACHIELS , Y. M ADAY, A. T. PATERA , AND
G. T URINICI, Reliable real-time solution of parametrized partial differential equations: Reduced-
basis output bound methods, Journal of Fluids Engineering, 124 (2002), pp. 70–80, https://doi.
org/10.1115/1.1448332. (Cited on p. 83)
[465] C. P RUD ’ HOMME , D. V. ROVAS , K. V EROY, AND A. T. PATERA, A mathematical and com-
putational framework for reliable real-time solution of parametrized partial differential equa-
tions, ESAIM: Mathematical Modelling and Numerical Analysis, 36 (2002), pp. 747–771, https:
//doi.org/10.1051/m2an:2002035. (Cited on p. 127)
448 Bibliography
[466] A. Q UAINI, Algorithms for Fluid–Structure Interaction Problems Arising in Hemodynamics, PhD
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[467] A. Q UAINI , R. G LOWINSKI , AND S. C ANIC, Symmetry breaking and preliminary results about
a Hopf bifurcation for incompressible viscous flow in an expansion channel, International Journal
of Computational Fluid Dynamics, 30 (2016), pp. 7–19, https://doi.org/10.1080/10618562.
2016.1144877. (Cited on p. 311)
[468] A. Q UARTERONI, Numerical Models for Differential Problems, MS&A Series Vol. 16, Springer
International, 2017, https://doi.org/10.1007/978-3-319-49316-9. (Cited on pp. 8, 103,
367, 373)
[469] A. Q UARTERONI AND L. F ORMAGGIA, Mathematical modelling and numerical simulation of the
cardiovascular system, Handbook of Numerical Analysis, 12 (2004), pp. 3–127. (Cited on pp. 283,
300)
[470] A. Q UARTERONI , A. M ANZONI , AND F. N EGRI, Reduced Basis Methods for Partial Dif-
ferential Equations: An Introduction, vol. 92, Springer, 2015, https://doi.org/10.1007/
978-3-319-15431-2. (Cited on pp. 62, 103, 106, 108, 128, 205)
[471] A. Q UARTERONI , A. M ANZONI , AND G. ROZZA, Certified reduced basis approximation for
parametrized partial differential equations and applications, Journal of Mathematics in Industry,
1 (2011), https://doi.org/10.1186/2190-5983-1-3. (Cited on pp. 62, 125, 127, 129, 130)
[474] A. Q UARTERONI , G. ROZZA , AND A. M ANZONI, Certified reduced basis approximation for
parametrized partial differential equations and applications, Journal of Mathematics in Industry,
1 (2011), https://doi.org/10.1186/2190-5983-1-3. (Cited on p. 6)
[475] A. Q UARTERONI , G. ROZZA , AND A. Q UAINI, Reduced basis methods for optimal control of
advection-diffusion problems, in Advances in Numerical Mathematics, no. CMCS-CONF-2006-003,
RAS and University of Houston, 2007, pp. 193–216. (Cited on p. 83)
[479] A. R AHIMI AND B. R ECHT, Random features for large-scale kernel machines, in Proceedings of
the 20th International Conference on Neural Information Processing Systems, NIPS’07, Curran As-
sociates Inc., 2008, pp. 1177–1184. (Cited on p. 332)
Bibliography 449
learning framework for solving forward and inverse problems involving nonlinear partial differential
equations, Journal of Computational Physics, 378 (2019), pp. 686–707, https://doi.org/10.
1016/j.jcp.2018.10.045. (Cited on pp. 11, 401)
[481] A. R ASHEED , O. S AN , AND T. K VAMSDAL, Digital twin: Values, challenges and enablers from
a modeling perspective, IEEE Access, 8 (2020), pp. 21980–22012, https://doi.org/10.1109/
ACCESS.2020.2970143. (Cited on pp. 12, 410)
[482] S. S. R AVINDRAN, A reduced-order approach for optimal control of fluids using proper or-
thogonal decomposition, International Journal for Numerical Methods in Fluids, 34 (2000),
pp. 425–448, https://doi.org/10.1002/1097-0363(20001115)34:5<425::AID-FLD67>3.
0.CO;2-W. (Cited on p. 284)
[483] W. R EED AND T. H ILL, Triangular mesh methods for the neutron transport equation, Journal of
Scientific Computing, (1973). (Cited on p. 234)
[484] J. R EISS , P. S CHULZE , J. S ESTERHENN , AND V. M EHRMANN, The shifted proper orthogonal de-
composition: A mode decomposition for multiple transport phenomena, SIAM Journal on Scientific
Computing, 40 (2018), pp. A1322–A1344, https://doi.org/10.1137/17M1140571. (Cited on
pp. 11, 399)
[485] O. R EYNOLDS, IV. On the dynamical theory of incompressible viscous fluids and the determination
of the criterion, Philosophical Transactions of the Royal Society of London (A), 186 (1895), pp. 123–
164, https://doi.org/10.1098/rsta.1895.0004. (Cited on p. 165)
[486] C. M. R HIE AND W. L. C HOW, Numerical study of the turbulent flow past an airfoil with trailing
edge separation, AIAA Journal, 21 (1983), pp. 1525–1532, https://doi.org/10.2514/3.8284.
(Cited on p. 145)
[487] T. R ICHTER, Fluid–Structure Interactions. Model, Analysis and Finite Element, vol. 118 of Lecture
Notes in Computational Science and Engineering, Springer International, 2017. (Cited on pp. 284,
285, 286, 287)
[489] D. F. ROGERS AND J. A. A DAMS, Mathematical Elements for Computer Graphics (2nd ed.),
McGraw-Hill, Inc., USA, 1989. (Cited on p. 357)
[490] R. ROJAS, The backpropagation algorithm, in Neural Networks, Springer, 1996, pp. 149–182,
https://doi.org/10.1007/978-3-642-61068-4_7. (Cited on pp. 391, 392, 393)
[491] R. ROJAS, Neural Networks: A Systematic Introduction, Springer-Verlag, 1996. (Cited on p. 397)
[493] F. ROMOR , M. T EZZELE , A. L ARIO , AND G. ROZZA, Kernel-based Active Subspaces with
Application to CFD Parametric Problems Using Discontinuous Galerkin Method, arXiv preprint
arXiv:2008.12083 (2020). (Cited on pp. 10, 327, 329, 333, 339, 384)
[494] F. ROMOR , M. T EZZELE , M. M ROSEK , C. OTHMER , AND G. ROZZA, Multi-fidelity Data Fusion
through Parameter Space Reduction with Applications to Automotive Engineering, arXiv preprint
arXiv:2110.14396 (2021). (Cited on pp. 9, 206, 327, 334, 336, 384, 411)
450 Bibliography
[495] F. ROMOR , M. T EZZELE , AND G. ROZZA, A Local Approach to Parameter Space Reduction for
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
Regression and Classification Tasks, arXiv preprint arXiv:2107.10867 (2021). (Cited on pp. 10, 327,
333, 384)
[496] F. ROMOR , M. T EZZELE , AND G. ROZZA, ATHENA: Advanced techniques for high dimensional
parameter spaces to enhance numerical analysis, Software Impacts, 10 (2021), p. 100133, https:
//doi.org/10.1016/j.simpa.2021.100133. (Cited on pp. 9, 327, 330, 337, 384, 407)
[497] F. ROMOR , M. T EZZELE , AND G. ROZZA, Multi-fidelity data fusion for the approximation of scalar
functions with low intrinsic dimensionality through active subspaces, in Proceedings in Applied
Mathematics & Mechanics, vol. 20, Wiley Online Library, 2021, https://doi.org/10.1002/
pamm.202000349. (Cited on pp. 334, 384)
[498] F. ROSENBLATT, The perceptron: A probabilistic model for information storage and organization
in the brain, Psychological Review, 65 (1958), p. 386, https://doi.org/10.1037/h0042519.
(Cited on p. 393)
[501] G. ROZZA, On optimization, control and shape design of an arterial bypass, International Journal for
Numerical Methods in Fluids, 47 (2005), pp. 1411–1419, https://doi.org/10.1002/fld.888.
(Cited on p. 368)
[502] G. ROZZA, Reduced basis methods for Stokes equations in domains with non-affine parameter de-
pendence, Computing and Visualization in Science, 12 (2009), pp. 23–35, https://doi.org/10.
1007/s00791-006-0044-7. (Cited on p. 4)
[503] G. ROZZA , M. H ESS , G. S TABILE , M. T EZZELE , AND F. BALLARIN, Basic ideas and tools
for projection-based model reduction of parametric partial differential equations, in Model Or-
der Reduction, P. Benner, S. Grivet-Talocia, A. Quarteroni, G. Rozza, W. H. A. Schilders, and
L. M. Silveira, eds., vol. 2, De Gruyter, 2020, ch. 1, pp. 1–47, https://doi.org/10.1515/
9783110671490-001. (Cited on pp. 9, 10, 327, 347)
[504] G. ROZZA , D. H UYNH , N. N GUYEN , AND A. PATERA, Real-time reliable simulation of heat
transfer phenomena, in ASME—American Society of Mechanical Engineers—Heat Transfer Sum-
mer Conference Proceedings, San Francisco, 2009, https://doi.org/10.1115/HT2009-88212.
(Cited on p. 253)
[505] G. ROZZA , D. H UYNH , AND A. T. PATERA, Reduced basis approximation and a posteriori er-
ror estimation for affinely parametrized elliptic coercive partial differential equations, Archives of
Computational Methods in Engineering, 15 (2008), pp. 229–275, https://doi.org/10.1007/
s11831-008-9019-9. (Cited on pp. 3, 5, 83, 91, 94, 127, 251, 253, 258, 268)
[506] G. ROZZA , D. B. P. H UYNH , AND A. M ANZONI, Reduced basis approximation and a poste-
riori error estimation for Stokes flows in parametrized geometries: Roles of the inf-sup stabil-
ity constants, Numerische Mathematik, 125 (2013), pp. 115–152, https://doi.org/10.1007/
s00211-013-0534-8. (Cited on p. 83)
[507] G. ROZZA , A. KOSHAKJI , AND A. Q UARTERONI, Free form deformation techniques applied to
3D shape optimization problems, Communications in Applied and Industrial Mathematics, 4 (2013),
https://doi.org/10.1685/journal.caim.452. (Cited on p. 347)
Bibliography 451
Advances in reduced order methods for parametric industrial problems in computational fluid dy-
namics, in ECCOMAS ECFD 7—Proceedings of 6th European Conference on Computational Me-
chanics (ECCM 6) and 7th European Conference on Computational Fluid Dynamics (ECFD 7),
R. Owen, R. de Borst, J. Reese, and P. Chris, eds., Glasgow, UK, 2018, pp. 59–76. (Cited on pp. 11,
37, 203, 327, 347)
[509] G. ROZZA , A. M ANZONI , AND F. N EGRI, Reduction strategies for PDE-constrained optimization
problems in haemodynamics, in Proceedings of the 6th European Congress on Computational Meth-
ods in Applied Sciences and Engineering, Vienna Technical University, 2012, pp. 1748–1769. (Cited
on p. 347)
[510] G. ROZZA , N. N GUYEN , A. PATERA , AND S. D EPARIS, Reduced Basis Methods and a Pos-
teriori Error Estimators for Heat Transfer Problems, in ASME—American Society of Mechan-
ical Engineers—Heat Transfer Summer Conference Proceedings, San Francisco, 2009, https:
//doi.org/10.1115/HT2009-88211. (Cited on p. 253)
[511] G. ROZZA AND K. V EROY, On the stability of the reduced basis method for Stokes equations in
parametrized domains, Computer Methods in Applied Mechanics and Engineering, 196 (2007),
pp. 1244–1260, https://doi.org/10.1016/j.cma.2006.09.005. (Cited on pp. 39, 258, 272)
[512] S. RUBINO, Numerical Modeling of Turbulence by Richardson Number-Based and VMS Models,
PhD thesis, University of Seville, 2014. (Cited on pp. 61, 69, 133)
[514] P. G. S AFFMAN, A model for inhomogeneous turbulent flow, Proceedings of the Royal Society of
London. A. Mathematical and Physical Sciences, 317 (1970), pp. 417–433, https://doi.org/10.
1098/rspa.1970.0125. (Cited on p. 168)
[515] P. S AGAUT, Large Eddy Simulation for Incompressible Flows: An Introduction., Springer-Verlag,
2006, https://doi.org/10.1007/b137536. (Cited on pp. 59, 60)
[519] F. S ALMOIRAGHI , A. S CARDIGLI , H. T ELIB , AND G. ROZZA, Free-form deformation, mesh mor-
phing and reduced-order methods: Enablers for efficient aerodynamic shape optimisation, Interna-
tional Journal of Computational Fluid Dynamics, 32 (2018), pp. 233–247, https://doi.org/10.
1080/10618562.2018.1514115. (Cited on pp. 173, 205, 206, 347)
452 Bibliography
[520] S. S ALSA, Partial Differential Equations in Action: From Modelling to Theory, UNITEXT, Springer
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[522] O. S AN , R. M AULIK , AND M. A HMED, An artificial neural network framework for reduced order
modeling of transient flows, Communications in Nonlinear Science and Numerical Simulation, 77
(2019), pp. 271–287, https://doi.org/10.1016/j.cnsns.2019.04.025. (Cited on p. 397)
[527] P. J. S CHMID, Dynamic mode decomposition of numerical and experimental data, Journal of Fluid
Mechanics, 656 (2010), pp. 5–28, https://doi.org/10.1017/S0022112010001217. (Cited on
pp. 207, 208, 383)
[528] J. S CHMIDHUBER, Deep learning in neural networks: An overview, Neural networks, 61 (2015),
pp. 85–117, https://doi.org/10.1016/j.neunet.2014.09.003. (Cited on p. 396)
[529] E. S CHMIDT, Zur theorie der linearen und nichtlinearen integralgleichungen, Mathematische An-
nalen, 63 (1907), pp. 433–476, https://doi.org/10.1007/BF01449770. (Cited on p. 4)
[530] I. J. S CHOENBERG, Spline functions and the problem of graduation, Proceedings of the National
Academy of Sciences, 52 (1964), pp. 947–950, https://doi.org/10.1073/pnas.52.4.947.
(Cited on p. 357)
[531] B. S CHÖLKOPF, A. S MOLA , AND K.-R. M ÜLLER, Kernel principal component analysis, in
International Conference on Artificial Neural Networks, Springer, 1997, pp. 583–588, https:
//doi.org/10.1007/BFb0020217. (Cited on p. 205)
[532] R. S CHREIBER AND H. K ELLER, Driven cavity flows by efficient numerical techniques, Journal of
Computational Physics, 49 (1983), pp. 310–333, https://doi.org/10.1016/0021-9991(83)
90129-8. (Cited on p. 157)
[533] T. S EDERBERG AND S. PARRY, Free-form deformation of solid geometric models, in Proceedings
of SIGGRAPH—Special Interest Group on GRAPHics and Interactive Techniques, vol. 20, SIG-
GRAPH, 1986, pp. 151–159, https://doi.org/10.1145/15886.15903. (Cited on pp. 10, 212,
346, 347)
[535] R. S EYDEL, Practical Bifurcation and Stability Analysis, Interdisciplinary Applied Mathematics,
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[538] K. S IMONYAN AND A. Z ISSERMAN, Very Deep Convolutional Networks for Large-Scale Image
Recognition, in 3rd International Conference on Learning Representations, 2015. (Cited on p. 407)
[539] S. S IRISUP AND G. K ARNIADAKIS, Stability and accuracy of periodic flow solutions obtained
by a POD-penalty method, Physica D: Nonlinear Phenomena, 202 (2005), pp. 218–237, https:
//doi.org/10.1016/j.physd.2005.02.006. (Cited on pp. 154, 155, 189)
[541] J. S MAGORINSKY, General circulation experiments with the primitive equations. I. The basic
experiment, Monthly Weather Review, 91 (1963), pp. 99–164, https://doi.org/10.1175/
1520-0493(1963)091<0099:GCEWTP>2.3.CO;2. (Cited on pp. 59, 60, 69, 169)
[542] R. L. S MITH, The hit-and-run sampler: A globally reaching Markov chain sampler for generat-
ing arbitrary multivariate distributions, in Proceedings Winter Simulation Conference, IEEE, 1996,
pp. 260–264, https://doi.org/10.1109/WSC.1996.873287. (Cited on p. 329)
[543] I. J. S OBEY AND P. G. D RAZIN, Bifurcations of two-dimensional channel flows, Journal of Fluid
Mechanics, 171 (1986), pp. 263–287, https://doi.org/10.1017/S0022112086001441. (Cited
on p. 311)
[544] I. M. S OBOL’, On the distribution of points in a cube and the approximate evaluation of integrals,
Zhurnal Vychislitel’noi Matematiki i Matematicheskoi Fiziki, 7 (1967), pp. 784–802, https://
doi.org/10.1016/0041-5553(67)90144-9. (Cited on p. 204)
[545] I. S OMMERVILLE, Software Engineering, (7th ed.), Pearson Education India, 2004. (Cited on p. 386)
[546] P. S PALART AND S. A LLMARAS, A One-Equation Turbulence Model for Aerodynamic Flows, in
30th Aerospace Sciences Meeting and Exhibit, American Institute of Aeronautics and Astronautics,
January 1992, https://doi.org/10.2514/6.1992-439. (Cited on pp. 167, 219)
[547] D. F. S PECHT, A general regression neural network, IEEE Transactions on Neural Networks, 2
(1991), pp. 568–576, https://doi.org/10.1109/72.97934. (Cited on p. 389)
[548] G. S TABILE , S. H IJAZI , A. M OLA , S. L ORENZI , AND G. ROZZA, POD-Galerkin reduced order
methods for CFD using finite volume discretisation: Vortex shedding around a circular cylinder,
Communications in Applied and Industrial Mathematics, 8 (2017), pp. 210–236, https://doi.
org/10.1515/caim-2017-0011. (Cited on pp. 143, 144, 151, 155, 187, 381)
[549] G. S TABILE AND G. ROZZA, ITHACA-FV—In real Time Highly Advanced Computational Appli-
cations for Finite Volumes. http://www.mathlab.sissa.it/ithaca-fv. (Cited on pp. 181,
381)
454 Bibliography
[550] G. S TABILE AND G. ROZZA, Finite volume POD-Galerkin stabilised reduced order methods for the
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
parametrised incompressible Navier–Stokes equations, Computers & Fluids, 173 (2018), pp. 273–
284, https://doi.org/10.1016/j.compfluid.2018.01.035. (Cited on pp. 39, 143, 144, 219,
381)
[551] G. S TABILE , M. Z ANCANARO , AND G. ROZZA, Efficient geometrical parametrization for finite-
volume-based reduced order methods, International Journal for Numerical Methods in Engineering,
121 (2020), pp. 2655–2682, https://doi.org/10.1002/nme.6324. (Cited on pp. 144, 381)
[554] S. K. S TAR , G. S TABILE , F. B ELLONI , G. ROZZA , AND J. D EGROOTE, Extension and comparison
of techniques to enforce boundary conditions in finite volume POD-Galerkin reduced order models
for fluid dynamic problems, arXiv:1912.00825 (2019). (Cited on p. 156)
[555] M. S TEIN, Large sample properties of simulations using Latin hypercube sampling, Technomet-
rics, 29 (1987), pp. 143–151, https://doi.org/10.1080/00401706.1987.10488205. (Cited
on p. 204)
[556] G. S TRANG, Linear Algebra and Its Applications, Academic Press, New York, 3rd ed., 1988. (Cited
on p. 4)
[557] M. S TRAZZULLO , F. BALLARIN , R. M OSETTI , AND G. ROZZA, Model reduction for parametrized
optimal control problems in environmental marine sciences and engineering, SIAM Journal on
Scientific Computing, 40 (2018), pp. B1055–B1079, https://doi.org/10.1137/17M1150591.
(Cited on pp. 83, 261, 262)
[558] M. S TRAZZULLO , F. BALLARIN , AND G. ROZZA, POD-Galerkin model order reduction for
parametrized time dependent linear quadratic optimal control problems in saddle point formulation,
Journal of Scientific Computing, 83 (2020), https://doi.org/10.1007/s10915-020-01232-x.
(Cited on pp. 84, 92, 205)
[559] M. S TRAZZULLO , F. BALLARIN , AND G. ROZZA, A Certified Reduced Basis Method for Linear
Parametrized Parabolic Optimal Control Problems in Space-Time Formulation, arXiv:2103.00460.
(Cited on p. 91)
[560] M. S TRAZZULLO , F. BALLARIN , AND G. ROZZA, POD-Galerkin model order reduction for
parametrized nonlinear time dependent optimal flow control: An application to shallow water equa-
tions, Journal of Numerical Mathematics, 30 (2022), pp. 63–84, https://doi.org/10.1515/
jnma-2020-0098. (Cited on pp. 89, 91, 95)
[562] M. S TRAZZULLO , Z. Z AINIB , F. BALLARIN , AND G. ROZZA, Reduced order methods for
parametrized non-linear and time dependent optimal flow control problems, towards applications
in biomedical and environmental sciences, in Numerical Mathematics and Advanced Applications
ENUMATH 2019, F. J. Vermolen and C. Vuik, eds., Springer International, 2020, pp. 841–850,
https://doi.org/10.1007/978-3-030-55874-1_83. (Cited on pp. 83, 367)
Bibliography 455
[563] V. S TROUHAL, Ueber eine besondere art der tonerregung, Annalen der Physik und Chemie, 241
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[564] B. S UDRET, Global sensitivity analysis using polynomial chaos expansions, Reliability Engineering
& System Safety, 93 (2008), pp. 964–979, https://doi.org/10.1016/j.ress.2007.04.002.
(Cited on p. 405)
[566] S. S Y AND C. M UREA, Algorithm for solving fluid–structure interaction problem on a global moving
mesh, Coupled Systems Mechanics, 1 (2012), pp. 99–113, https://doi.org/10.12989/csm.
2012.1.1.099. (Cited on p. 303)
[567] T. TADDEI, A registration method for model order reduction: Data compression and geometry re-
duction, SIAM Journal on Scientific Computing, 42 (2020), pp. A997–A1027, https://doi.org/
10.1137/19M1271270. (Cited on p. 298)
[568] M. T EZZELE , F. BALLARIN , AND G. ROZZA, Combined parameter and model reduction of car-
diovascular problems by means of active subspaces and POD-Galerkin methods, in Mathematical
and Numerical Modeling of the Cardiovascular System and Applications, D. Boffi, L. F. Pavarino,
G. Rozza, S. Scacchi, and C. Vergara, eds., vol. 16 of SEMA-SIMAI Series, Springer International,
2018, pp. 185–207, https://doi.org/10.1007/978-3-319-96649-6_8. (Cited on pp. 11, 342,
350, 356, 367)
[569] M. T EZZELE , N. D EMO , M. G ADALLA , A. M OLA , AND G. ROZZA, Model order reduction
by means of active subspaces and dynamic mode decomposition for parametric hull shape de-
sign hydrodynamics, in Technology and Science for the Ships of the Future: Proceedings of NAV
2018: 19th International Conference on Ship & Maritime Research, IOS Press, 2018, pp. 569–576,
https://doi.org/10.3233/978-1-61499-870-9-569. (Cited on pp. 37, 347)
[570] M. T EZZELE , N. D EMO , A. M OLA , AND G. ROZZA, PyGeM: Python Geometrical Morphing, Soft-
ware Impacts, 7 (2021), p. 100047, https://doi.org/10.1016/j.simpa.2020.100047. (Cited
on pp. 10, 347, 359, 384)
[572] M. T EZZELE , N. D EMO , AND G. ROZZA, Shape optimization through proper orthogonal decom-
position with interpolation and dynamic mode decomposition enhanced by active subspaces, in Pro-
ceedings of MARINE 2019: VIII International Conference on Computational Methods in Marine
Engineering, R. Bensow and J. Ringsberg, eds., 2019, pp. 122–133. (Cited on pp. 327, 347, 383)
[573] M. T EZZELE , N. D EMO , G. S TABILE , A. M OLA , AND G. ROZZA, Enhancing CFD predictions
in shape design problems by model and parameter space reduction, Advanced Modeling and Sim-
ulation in Engineering Sciences, 7 (2020), https://doi.org/10.1186/s40323-020-00177-y.
(Cited on pp. 11, 217, 219, 342, 350, 383, 384)
[574] M. T EZZELE , F. S ALMOIRAGHI , A. M OLA , AND G. ROZZA, Dimension reduction in hetero-
geneous parametric spaces with application to naval engineering shape design problems, Ad-
vanced Modeling and Simulation in Engineering Sciences, 5 (2018), https://doi.org/10.1186/
s40323-018-0118-3. (Cited on pp. 327, 346, 347, 356, 360)
[576] T HORATEC C ORPORATION, HeartMate 3 Left Ventricular Assist System: Instructions for Use. 2017.
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
(Cited on p. 374)
[577] S. P. T IMOSHENKO AND J. M. G ERE, Theory of Elastic Stability, Dover Civil and Mechanical
Engineering, Dover Publications, 2009. (Cited on p. 116)
[578] L. T OBISKA AND R. V ERFÜRTH, Analysis of a streamline diffusion finite element method for the
Stokes and Navier–Stokes equations, SIAM Journal on Numerical Analysis, 33 (1996), pp. 107–127,
https://doi.org/10.1137/0733007. (Cited on p. 130)
[579] D. T ORLO, Model Reduction for Advection Dominated Hyperbolic Problems in an ALE Framework:
Offline and Online Phases, arXiv:2003.13735 (2020). (Cited on p. 305)
[580] D. T ORLO , F. BALLARIN , AND G. ROZZA, Stabilized weighted reduced basis methods for
parametrized advection dominated problems with random inputs, SIAM/ASA Journal on Uncer-
tainty Quantification, 6 (2018), pp. 1475–1502, https://doi.org/10.1137/17M1163517. (Cited
on pp. 254, 259)
[581] E. F. T ORO, Riemann Solvers and Numerical Methods for Fluid Dynamics: A Practical Introduction,
Springer Science & Business Media, 2013. (Cited on p. 239)
[582] E. F. T ORO , M. S PRUCE , AND W. S PEARES, Restoration of the contact surface in the HLL-Riemann
solver, Shock Waves, 4 (1994), pp. 25–34, https://doi.org/10.1007/BF01414629. (Cited on
p. 239)
[583] S. T RENN, Multilayer perceptrons: Approximation order and necessary number of hidden units,
IEEE Transactions on Neural Networks, 19 (2008), pp. 836–844, https://doi.org/10.1109/
TNN.2007.912306. (Cited on p. 390)
[584] R. T RIPATHY, I. B ILIONIS , AND M. G ONZALEZ, Gaussian processes with built-in dimensional-
ity reduction: Applications to high-dimensional uncertainty propagation, Journal of Computational
Physics, 321 (2016), pp. 191–223, https://doi.org/10.1016/j.jcp.2016.05.039. (Cited on
p. 336)
[585] R. K. T RIPATHY AND I. B ILIONIS, Deep UQ: Learning deep neural network surrogate models for
high dimensional uncertainty quantification, Journal of Computational Physics, 375 (2018), pp. 565–
588, https://doi.org/10.1016/j.jcp.2018.08.036. (Cited on p. 332)
[586] D. J. T RITTON, Physical Fluid Dynamics, The Modern University in Physics Series, Springer
Netherlands, 1977, https://doi.org/10.1007/978-94-009-9992-3. (Cited on pp. 119, 311,
370, 372)
[587] F. T RÖLTZSCH, Optimal control of partial differential equations, Graduate Studies in Mathematics,
112 (2010), https://doi.org/10.1090/gsm/112. (Cited on p. 262)
[589] M. U DELL AND A. T OWNSEND, Why are big data matrices approximately low rank?, SIAM
Journal on Mathematics of Data Science, 1 (2019), pp. 144–160, https://doi.org/10.1137/
18M1183480. (Cited on p. 205)
[590] K. U RBAN AND A. T. PATERA, A new error bound for reduced basis approximation of parabolic
partial differential equations, Comptes Rendus Mathematique, 350 (2012), pp. 203–207, https:
//doi.org/10.1016/j.crma.2012.01.026. (Cited on pp. 83, 92)
[592] L. V ENTURI , F. BALLARIN , AND G. ROZZA, A weighted POD method for elliptic PDEs with
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[593] L. V ENTURI , D. T ORLO , F. BALLARIN , AND G. ROZZA, Weighted Reduced Order Methods
for Parametrized Partial Differential Equations with Random Inputs, Springer International, 2019,
pp. 27–40, https://doi.org/10.1007/978-3-030-04870-9_2. (Cited on p. 254)
[594] K. V EROY AND A. T. PATERA, Certified real-time solution of the parametrized steady incompress-
ible Navier–Stokes equations: Rigorous reduced-basis a posteriori error bounds, International Jour-
nal for Numerical Methods in Fluids, 47 (2005), pp. 773–788, https://doi.org/10.1002/fld.
867. (Cited on p. 5)
[595] K. V EROY, C. P RUD ’ HOMME , D. ROVAS , AND A. T. PATERA, A Posteriori Error Bounds for
Reduced-Basis Approximation of Parametrized Noncoercive and Nonlinear Elliptic Partial Dif-
ferential Equations, in 16th AIAA Computational Fluid Dynamics Conference, Orlando, Florida,
American Institute of Aeronautics and Astronautics, June 2003, https://doi.org/10.2514/6.
2003-3847. (Cited on pp. 4, 7)
[599] A. V REMAN, An eddy-viscosity subgrid-scale model for turbulent shear flow: Algebraic theory
and applications, Physics of Fluids, 16 (2004), pp. 3670–3681, https://doi.org/10.1063/1.
1785131. (Cited on p. 170)
[600] S. WALTON , O. H ASSAN , AND K. M ORGAN, Reduced order modelling for unsteady fluid flow using
proper orthogonal decomposition and radial basis functions, Applied Mathematical Modelling, 37
(2013), pp. 8930–8945, https://doi.org/10.1016/j.apm.2013.04.025. (Cited on p. 173)
[601] D. WAN , B. PATNAIK , AND G. W EI, A new benchmark quality solution for the buoyancy-driven
cavity by discrete singular convolution, Numerical Heat Transfer, Part B: Fundamentals, 40 (2001),
pp. 199–228, https://doi.org/10.1080/104077901752379620. (Cited on p. 75)
[602] Q. WANG , J. S. H ESTHAVEN , AND D. R AY, Non-intrusive reduced order modeling of unsteady
flows using artificial neural networks with application to a combustion problem, Journal of Com-
putational Physics, 384 (2019), pp. 289–307, https://doi.org/10.1016/j.jcp.2019.01.031.
(Cited on p. 206)
[603] S. WANG , Y. T ENG , AND P. P ERDIKARIS, Understanding and mitigating gradient flow pathologies
in physics-informed neural networks, SIAM Journal on Scientific Computing, 43 (2021), pp. A3055–
A3081, https://doi.org/10.1137/20M1318043. (Cited on p. 401)
[604] S. WANG , H. WANG , AND P. P ERDIKARIS, Learning the solution operator of parametric partial
differential equations with physics-informed DeepOnets, Science Advances, 7 (2021), https://
doi.org/10.1126/sciadv.abi8605. (Cited on p. 401)
[605] W. Q. WANG AND Y. YAN, Strongly coupling of partitioned fluid–solid interaction solvers using
reduced-order models, Applied Mathematical Modelling, 34 (2010), pp. 3817–3830, https://doi.
org/10.1016/j.apm.2010.03.022. (Cited on p. 286)
458 Bibliography
[606] Y. WANG , A. Q UAINI , AND S. C ANI Ć, A higher-order discontinuous Galerkin/arbitrary La-
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
grangian Eulerian partitioned approach to solving fluid–structure interaction problems with incom-
pressible, viscous fluids and elastic structures, Journal of Scientific Computing, 76 (2018), pp. 481–
520, https://doi.org/10.1007/s10915-017-0629-y. (Cited on p. 286)
[609] K. M. WASHABAUGH , M. J. Z AHR , AND C. FARHAT, On the Use of Discrete Nonlinear Reduced-
Order Models for the Prediction of Steady-State Flows Past Parametrically Deformed Complex Ge-
ometries, in 54th AIAA Aerospace Sciences Meeting, American Institute of Aeronautics and Astro-
nautics, January 2016, https://doi.org/10.2514/6.2016-1814. (Cited on p. 144)
[610] G. W. WASILKOWSKI, Explicit cost bounds of algorithms for multivariate tensor product problems,
Journal of Complexity, 1 (1995), pp. 1–56, https://doi.org/10.1006/jcom.1995.1001. (Cited
on p. 256)
[611] S. W EI, Is human digital twin possible?, Computer Methods and Programs in Biomedicine Update,
1 (2021), p. 100014, https://doi.org/10.1016/j.cmpbup.2021.100014. (Cited on p. 413)
[613] G. W ELPER, Interpolation of functions with parameter dependent jumps by transformed snapshots,
SIAM Journal on Scientific Computing, 39 (2017), pp. A1225–A1250, https://doi.org/10.
1137/16M1059904. (Cited on p. 298)
[614] G. W ELPER, Transformed snapshot interpolation with high resolution transforms, SIAM Journal on
Scientific Computing, 42 (2020), pp. A2037–A2061, https://doi.org/10.1137/19M126356X.
(Cited on p. 298)
[615] N. W ESTERHOF, J.-W. L ANKHAAR , AND B. W ESTERHOF, The arterial Windkessel, Medical
& Biological Engineering & Computing, 47 (2008), pp. 131–41, https://doi.org/10.1007/
s11517-008-0359-2. (Cited on p. 374)
[616] N. W IENER, The homogeneous chaos, American Journal of Mathematics, 60 (1938), pp. 897–936,
https://doi.org/10.2307/2371268. (Cited on p. 405)
[617] D. C. W ILCOX, Numerical Study of Separated Turbulent Flows, in 7th Fluid and PlasmaDynamics
Conference, American Institute of Aeronautics and Astronautics, June 1974, https://doi.org/
10.2514/6.1974-584. (Cited on p. 168)
[618] D. C. W ILCOX, Reassessment of the scale-determining equation for advanced turbulence mod-
els, AIAA Journal, 26 (1988), pp. 1299–1310, https://doi.org/10.2514/3.10041. (Cited on
p. 168)
[619] K. W ILLCOX, Unsteady flow sensing and estimation via the gappy proper orthogonal decomposi-
tion, Computers & Fluids, 35 (2006), pp. 208–226, https://doi.org/10.1016/j.compfluid.
2004.11.006. (Cited on p. 205)
Bibliography 459
[620] K. W ILLCOX AND J. P ERAIRE, Balanced model reduction via the proper orthogonal decomposition,
Downloaded 12/19/23 to 103.172.41.151 . Redistribution subject to SIAM license or copyright; see https://epubs.siam.org/terms-privacy
[621] R. W ILLE AND H. F ERNHOLZ, Report on the first European Mechanics Colloquium, on the
Coanda effect, Journal of Fluid Mechanics, 23 (1965), pp. 801–819, https://doi.org/10.1017/
S0022112065001702. (Cited on p. 372)
[622] C. K. W ILLIAMS AND C. E. R ASMUSSEN, Gaussian Processes for Machine Learning, vol. 2, MIT
Press, 2006. (Cited on pp. 206, 338)
[623] J. W ITTEVEEN AND H. B IJL, Explicit Mesh Deformation Using Inverse Distance Weighting In-
terpolation, in 19th AIAA Computational Fluid Dynamics, AIAA, 2009, https://doi.org/10.
2514/6.2009-3996. (Cited on pp. 206, 351)
[625] Y. W U AND X. C. C AI, A fully implicit domain decomposition based ALE framework for three-
dimensional fluid–structure interaction with application in blood flow computation, Journal of Com-
putational Physics, 258 (2014), pp. 524–537, https://doi.org/10.1016/j.jcp.2013.10.046.
(Cited on p. 286)
[626] N. W YCOFF , M. B INOIS , AND S. M. W ILD, Sequential learning of active subspaces, Journal
of Computational and Graphical Statistics, 30 (2021), pp. 1–14, https://doi.org/10.1080/
10618600.2021.1874962. (Cited on p. 336)
[627] D. X IAO , F. FANG , A. B UCHAN , C. PAIN , I. NAVON , J. D U , AND G. H U, Non linear model re-
duction for the Navier–Stokes equations using residual DEIM method, Journal of Computational
Physics, 263 (2014), pp. 1–18, https://doi.org/10.1016/j.jcp.2014.01.011. (Cited on
p. 151)
[629] L. X U , Y. TANG , X. X U , Y. F ENG , AND Y. G UO, A high order discontinuous Galerkin method
based RANS turbulence framework for OpenFOAM, in Proceedings of the 2017 2nd International
Conference on Communication and Information Systems, ICCIS 2017, New York, 2017, Association
for Computing Machinery, pp. 404–408, https://doi.org/10.1145/3158233.3159368. (Cited
on p. 241)
[630] J. YAN AND C.-W. S HU, Local discontinuous Galerkin methods for partial differential equations
with higher order derivatives, Journal of Scientific Computing, 17 (2002), pp. 27–47, https://
doi.org/10.1023/A:1015132126817. (Cited on p. 236)
[631] M. YANO, A space-time Petrov–Galerkin certified reduced basis method: Application to the Boussi-
nesq equations, SIAM Journal on Scientific Computing, 36 (2014), pp. A232–A266, https:
//doi.org/10.1137/120903300. (Cited on pp. 5, 83, 92)
[632] M. YANO, Discontinuous Galerkin reduced basis empirical quadrature procedure for model re-
duction of parametrized nonlinear conservation laws, Advances in Computational Mathematics, 45
(2019), pp. 2287–2320, https://doi.org/10.1007/s10444-019-09710-z. (Cited on p. 234)
[633] T. YOUNG , D. H AZARIKA , S. P ORIA , AND E. C AMBRIA, Recent trends in deep learning based
natural language processing, IEEE Computational Intelligence Magazine, 13 (2018), pp. 55–75.
(Cited on pp. 389, 402)
460 Bibliography
[637] M. J. Z AHR AND C. FARHAT, Progressive construction of a parametric reduced-order model for
PDE-constrained optimization, International Journal for Numerical Methods in Engineering, 102
(2014), pp. 1111–1135, https://doi.org/10.1002/nme.4770. (Cited on p. 144)
[638] Z. Z AINIB , F. BALLARIN , S. F REMES , P. T RIVERIO , L. J IME ŃEZ -J UAN , AND G. ROZZA, Re-
duced order methods for parametric optimal flow control in coronary bypass grafts, towards patient–
specific data assimilation, International Journal for Numerical Methods in Biomedical Engineering,
37 (2021), p. e3367, https://doi.org/10.1002/cnm.3367. (Cited on pp. 83, 366)
[639] M. Z ANCANARO , M. M ROSEK , G. S TABILE , C. OTHMER , AND G. ROZZA, Hybrid neural network
reduced order modelling for turbulent flows with geometric parameters, Fluids, 6 (2021), p. 296,
https://doi.org/10.3390/fluids6080296. (Cited on pp. 361, 381)
[641] M. M. Z DRAVKOVICH, Flow around Circular Cylinders: Volume 2: Applications, Oxford Univer-
sity Press, 2003. (Cited on p. 187)
[642] E. Z EIDLER, Nonlinear Functional Analysis and its Applications, vol. I: Fixed-Point Theorems,
Springer-Verlag New York, 1985, https://doi.org/10.1007/BF00047050. (Cited on pp. 99,
104)
[643] G. Z HANG , J. Z HANG , AND J. H INKLE, Learning nonlinear level sets for dimensionality reduc-
tion in function approximation, in Advances in Neural Information Processing Systems Conference,
2019, pp. 13199–13208. (Cited on pp. 327, 330, 337, 384)
[644] R. Z IMMERMANN AND S. G ÖRTZ, Non-linear reduced order models for steady aerodynamics, Pro-
cedia Computer Science, 1 (2010), pp. 165–174, https://doi.org/10.1016/j.procs.2010.
04.019. (Cited on p. 144)
Index
461
462 Index
This book will be of interest to researchers and graduate students in the field of
Methods and
reduced order modeling.
Applications in
computing, and reduced order modeling and methods with a special focus on
viscous flows and complex geometrical parametrizations.
Computational
several aspects concerning the numerical approximation of partial differential
equations with a special emphasis on model order reduction techniques for
computational fluid dynamic problems. He has worked with industrial partners on a
variety of projects aimed at exporting advanced numerical methods to the industrial
world.
FRANCESCO BALLARIN
GIOVANNI STABILE
Society for Industrial and Applied Mathematics GIANLUIGI ROZZA
3600 Market Street, 6th Floor
Philadelphia, PA 19104-2688 USA
+1-215-382-9800
[email protected] • www.siam.org
CS27
ISBN: 978-1-61197-724-0
90000
CS27
Computational Science and Engineering
9 781611 977240