163
Coupling and Simulation of Acoustic Fluid-Structure Interaction Systems Using Localized Lagrange Multipliers by Mike R. Ross B.S., Colorado School of Mines, 1998 M.S., University of Colorado, Boulder, 2004 A thesis submitted to the Faculty of the Graduate School of the University of Colorado in partial fulfillment of the requirements for the degree of Doctor of Philosophy Department of Aerospace Engineering Science 2006

Coupling and Simulation of Acoustic Fluid-Structure … · 2006-06-30 · Interaction Systems Using Localized Lagrange Multipliers by Mike R. Ross ... Coupling and Simulation of Acoustic

  • Upload
    dophuc

  • View
    214

  • Download
    0

Embed Size (px)

Citation preview

Coupling and Simulation of Acoustic Fluid-Structure

Interaction Systems Using Localized Lagrange Multipliers

by

Mike R. Ross

B.S., Colorado School of Mines, 1998

M.S., University of Colorado, Boulder, 2004

A thesis submitted to the

Faculty of the Graduate School of the

University of Colorado in partial fulfillment

of the requirements for the degree of

Doctor of Philosophy

Department of Aerospace Engineering Science

2006

This thesis entitled:Coupling and Simulation of Acoustic Fluid-Structure Interaction Systems Using Localized Lagrange

Multiplierswritten by Mike R. Ross

has been approved for the Department of Aerospace Engineering Science

Carlos Felippa

K.C. Park

Date

The final copy of this thesis has been examined by the signatories, and we find that both the contentand the form meet acceptable presentation standards of scholarly work in the above mentioned

discipline.

iii

Ross, Mike R. (Ph.D., Aerospace Engineering Science)

Coupling and Simulation of Acoustic Fluid-Structure Interaction Systems Using Localized Lagrange

Multipliers

Thesis directed by Prof. Carlos Felippa

This thesis presents a new coupling method for treating the interaction of an acoustic fluid with

a flexible structure, with emphasis on handling spatially non-matching meshes. It is based on the

Localized Lagrange Multiplier (LLM) method. A frame is introduced as a ”mediator” or ”information

relay” device between the fluid and the structure at the interaction surface. The frame is discretized

in terms of kinematic variables. A Lagrange multiplier field is introduced between the frame and the

structure, and another one between the frame and the fluid. The function of the multiplier pair is

weak enforcement of kinematic continuity. This configuration completely decouples the structure and

fluid models, because each model communicates to the frame through node collocated multipliers and

not directly to each other.

In order to assure proper communication, energy formulations of the fluid and structure models

are in terms of displacements and associated time derivatives. A novel transformation of the fluid

displacement model into a fluid displacement potential model enforces the irrotational condition of

the acoustic fluid. This transformation reduces the number of degrees of freedom in two and three-

dimensions and is suitable for both vibration and transient analyses.

The LLM method facilitates the construction of separate discretizations using different mesh gen-

eration programs, as well as use of customized time integration methods. To advance the solution in

time, the LLM coupling method is combined with a partitioned solution procedure. The time-stepping

computations are organized in a way that eliminates the traditional prediction step characteristic of

staggered solution procedures. This is accomplished by solving for the interface variables: Lagrange

multipliers and frame states, and then feeding this solution back to the coupled components. This

sequence forestalls the well-known stability degradation caused by prediction, yet it retains the de-

sirable localization features of a partitioned analysis procedure. One consequence of this method is

that if two A-stable integration schemes, such as the trapezoidal rule, are chosen for the fluid and

structure, then the coupled system retains unconditional stability. Other time integration schemes,

such as central difference, for one or both components can be readily accommodated.

iv

Acknowledgements

Above all, I would like to thank my advisor, Prof. Carlos Felippa, for his patience and his

guidance. I express my thanks to my committee members: Prof. K.C. Park, Prof. Thomas Geers,

Prof. Kurt Maute, Prof. Stein Sture, and Prof. Michael Sprague, for their insight and amazing patience

with my constant interruptions. I am grateful to the Center for Aerospace Structures, particularly

Deborah Mellblom and my fellow graduate students for their assistances and entertainment. Special

thanks is given to Dr. Hiraku Sakamoto for his support with the vibration analysis, and to Christophe

Kassiotis for his assistances with the Bleich-Sandler plate problem. Finally, I thank all those intrepid

souls that went climbing with me and helped maintain my sanity. This research was funded by NSF

grant CMS 0219422.

Of course, I am deeply indebted to my wife for all her support, and willingness to put her dreams

on hold.

v

Contents

Chapter

1 Introduction 1

1.1 Thesis Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Main Thesis Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

1.3 Manuscript Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Localized Lagrange Multiplier Method for Acoustic FSI 9

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2 System’s Energy Functionals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

2.3 Finite Element Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.3.1 Discretization of the Partitioned Domains . . . . . . . . . . . . . . . . . . . . . 14

2.3.2 Lagrange Multiplier Boundary Discretization . . . . . . . . . . . . . . . . . . . 15

2.3.3 Interface Frame Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2.3.4 Lagrange Multiplier Shape Functions Revisited . . . . . . . . . . . . . . . . . . 22

2.3.5 Matrix Form of the Total System Functional . . . . . . . . . . . . . . . . . . . 23

2.4 Inclusion of Irrotational Assumption by the Displacement Potential . . . . . . . . . . . 25

2.4.1 Circulation Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.4.2 Gradient Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.5 The Variation of the Energy Functional . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

3 Transient Concepts Using Localized Lagrange Multipliers 32

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.2 Partitioned Transient Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.2.1 Scaling the Interface Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3.3 Inclusion of Concept for other FSI Methods (i.e. CASE) . . . . . . . . . . . . . . . . . 39

vi

3.3.1 Relation Between Fluid Pressure and Structure Force . . . . . . . . . . . . . . 39

3.3.2 Relation between Subsystems’ Boundary Displacements . . . . . . . . . . . . . 41

3.4 Error/ Accuracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3.4.1 Geer’s C-Error Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.4.2 Energy Error . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4 Validation of Fluid Code: Pressure on Dam Face 47

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.2 Problem Description and Analytical Solution . . . . . . . . . . . . . . . . . . . . . . . 47

4.3 Fluid Code Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

4.4 Incorporation of Silent Boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

4.5 Silent Boundary Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.6 Inclusion of Surface Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5 Validation of Concept: Infinite Piston Problem 61

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

5.2 Problem Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.3 Analytical Model for 1-D Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.4 Current Computational Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

5.5 LLM Method Results in 3-D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.5.1 Matching Meshes Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.5.2 Non-matching Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.5.3 Stability Verification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

6 Gravity Dam Benchmark Study: Vibration Analysis 73

6.1 Problem Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

6.2 Linear Vibration Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

6.2.1 Modal Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

6.2.2 Kinematic Continuity Issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

6.2.3 Frequency Content of Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . 81

6.2.4 Frequency Response Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

vii

6.3 Examples of Passing the Patch Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

6.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

7 Gravity Dam Benchmark Study: Transient Analysis, and Error 87

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

7.2 Load Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

7.2.1 Relative Displacement’s Load Vector . . . . . . . . . . . . . . . . . . . . . . . . 88

7.2.2 Total Displacement’s Load Vector . . . . . . . . . . . . . . . . . . . . . . . . . 89

7.3 Matching Mesh Comparison for CASE/CAFE and LLM . . . . . . . . . . . . . . . . . 90

7.4 Scaling the Interface Matrix Verification . . . . . . . . . . . . . . . . . . . . . . . . . . 90

7.5 Non-Matching Mesh Comparisons for both CASE and LLM . . . . . . . . . . . . . . . 93

7.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

8 Reduced Order Modeling and Cavitation with the Transient LLM method 99

8.1 Reduced Order Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

8.1.1 Computational Results for ROM . . . . . . . . . . . . . . . . . . . . . . . . . . 100

8.2 Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

8.2.1 Modification of the Fluid Element Stiffness for Cavitation . . . . . . . . . . . . 104

8.2.2 Frothing: Spurious Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

8.2.3 Bleich-Sandler Plate Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

8.2.4 Cavitation in Koyna Dam Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 109

8.3 Operational Count . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

8.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

9 Partitioned 3-D Problem with Curved Surface 116

9.1 Problem Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

9.2 Creation of the Connection Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

9.3 Non-matching Mesh Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

9.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

10 Conclusion 125

10.1 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

viii

Bibliography 128

Appendix

A CASE/CAFE Method 137

A.1 Fluid Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

A.1.1 Fluid Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

A.1.2 Spectral Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

A.1.3 Silent Boundary of the Fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

A.1.4 Explicit Time Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

A.1.5 Stability of Fluid Time Integration . . . . . . . . . . . . . . . . . . . . . . . . . 142

A.2 Structure, Fluid, and the Silent Boundary Coupling . . . . . . . . . . . . . . . . . . . 142

A.3 Staggered Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

B Mortar Method and the Consistent Interpolation Based Method 144

B.1 Mortar Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

B.2 Consistent Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

B.3 Relationship between LLM, Mortar, and Consistent Interpolation . . . . . . . . . . . . 147

C Discretization Error Bound for Linear Interpolation of the Field Variables at the Interface in

the LLM Method 149

ix

Tables

Table

3.1 Properties of well-known members of the Newmark method . . . . . . . . . . . . . . . 33

4.1 Parameters for the pressure on the dam face problem. . . . . . . . . . . . . . . . . . . 49

5.1 Parameters for the infinite piston fluid-structure system. . . . . . . . . . . . . . . . . . 62

6.1 Parameters for Koyna dam model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

8.1 Parameters for the Bleich-Sandler plate problem . . . . . . . . . . . . . . . . . . . . . 109

9.1 Parameters for Morrow Point dam model. . . . . . . . . . . . . . . . . . . . . . . . . . 119

x

Figures

Figure

2.1 LLM Concept: (a)Given a system. (b) Divide the system into subdomains. (c) Insert

an interface displacement frame that is linked by the Local Lagrange Multipliers. . . . 10

2.2 Comparison of the LLM method to the Mortar method. (a) Localized Lagrange multi-

pliers. (b) Classical Lagrange multiplier. . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.3 Lagrange Multipliers are collocated with subdomains boundary nodes. . . . . . . . . . 16

2.4 Graphical representation of the determination of the interface nodes in 2-D. . . . . . . 20

2.5 Integration over the straight line piece γi. . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.6 Bilinear quadrilateral element . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

2.7 Two connected bilinear quadrilateral elements . . . . . . . . . . . . . . . . . . . . . . . 29

3.1 Graphical representation of the time stepping procedure for the LLM method. . . . . . 35

4.1 Rigid dam problem for fluid code verification. . . . . . . . . . . . . . . . . . . . . . . . 48

4.2 Analytical pressure on the rigid dam. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

4.3 Fluid mesh with a characteristic length of 20 m. . . . . . . . . . . . . . . . . . . . . . 50

4.4 Pressure on dam face with a fluid characteristic length of 20 m. . . . . . . . . . . . . . 51

4.5 Pressure on dam face with a fluid characteristic length of 10 m. . . . . . . . . . . . . . 52

4.6 Pressure on dam face with a fluid characteristic length of 5 m. . . . . . . . . . . . . . 52

4.7 Viscous Damping Boundary concept [55]. . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.8 Pressure on dam face with a fluid characteristic length of 10 m with and without a

silent boundary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.9 Pressure on dam face with a fluid characteristic length of 5 m with damping added to

the fluid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.10 Pressure on dam face with a fluid characteristic length of 5 m with damping added to

the fluid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

xi

4.11 Free surface wave . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5.1 Infinite piston fluid-structure system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.2 Free body diagram of a mass-spring system with water. . . . . . . . . . . . . . . . . . 63

5.3 Displacement of the mass in the infinite piston fluid-structure system by an analytical

method. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.4 Displacement of the mass in the infinite piston fluid-structure system by existing com-

putational models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.5 Model with 3-D elements for the LLM method. . . . . . . . . . . . . . . . . . . . . . . 68

5.6 Comparison between analytical and LLM method with no visible error; C-error = 0.0007. 68

5.7 Model with non-matching meshes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

5.8 Interface frame for the non-matching mesh. . . . . . . . . . . . . . . . . . . . . . . . . 70

5.9 C-error for the infinite piston system with non-matching meshes. . . . . . . . . . . . . 70

5.10 Time step values and the error associated with the time steps. . . . . . . . . . . . . . 71

6.1 Downstream views of the Koyna dam: (a)Present day. (b) After the earthquake of

Decemeber 1967. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

6.2 Koyna dam-reservoir system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

6.3 El Centro earthquake May 18, 1940, horizontal component. . . . . . . . . . . . . . . . 75

6.4 Koyna dam-reservoir system with non-matching meshes . . . . . . . . . . . . . . . . . 78

6.5 Mode frequencies and mode shapes of the system . . . . . . . . . . . . . . . . . . . . . 79

6.6 5th Mode shape of the system for the Mortar method with the interface discretized as

the coarse mesh. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

6.7 30th Mode shape of the system for the LLM method with the interface discretized as

the refined mesh. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

6.8 Frequency content of the seismic acceleration. . . . . . . . . . . . . . . . . . . . . . . . 81

6.9 Frequency response of the partitioned subsystems. . . . . . . . . . . . . . . . . . . . . 82

6.10 Frequency response of the assembled system. . . . . . . . . . . . . . . . . . . . . . . . 83

6.11 Frequency response of the assembled system with the dam face as the output DOF. . 84

6.12 Frequency response of the assembled system with the dam Lagrange multiplier as the

input DOF. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

6.13 An example of a constant state of stress across the interface frame. . . . . . . . . . . . 84

6.14 An example of a constant state of stress across the interface frame, where ”x” is the

locations of the interface nodes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

xii

7.1 Relative displacement concept. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

7.2 Total displacement concept. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

7.3 Portion of Koyna dam mesh used as benchmark mesh for C-error. . . . . . . . . . . . 91

7.4 C-error values with matching meshes. . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

7.5 Crest Displacement of Koyna dam of converge results for the transient analysis by the

LLM method and the CASE method. . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

7.6 Energy difference caused by scaling at the interface of matching meshes. Energy on one

side of the interface is of the magnitude of 106 Nm on average. . . . . . . . . . . . . . 92

7.7 Dam crest displacement values for different interface meshing with LLM for the transient

analysis and non-matching meshes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

7.8 Transfer of forces with the LLM method. . . . . . . . . . . . . . . . . . . . . . . . . . 94

7.9 Energy difference across the interface frame. Average energy on structure side is

6.78x106Nm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

7.10 Dam crest displacement values for different interface meshing with CASE for the tran-

sient analysis and LLM for the non-matching meshes. . . . . . . . . . . . . . . . . . . 95

7.11 Dam crest displacement values for different interface meshing with CASE for the tran-

sient analysis and Mortar for the non-matching meshes. . . . . . . . . . . . . . . . . . 96

7.12 C-error for different Total DOF of non-matching meshes. . . . . . . . . . . . . . . . . . 97

8.1 Characteristic portion of the non-matching mesh used for the ROM test. . . . . . . . . 101

8.2 Reduction of system up to the frequency of the eigenvector. . . . . . . . . . . . . . . . 102

8.3 Dam crest displacement values for a reduced order model using 20% of the eigenvectors

of the structure and the fluid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

8.4 Density pressure relation for a bilinear fluid [114]. . . . . . . . . . . . . . . . . . . . . . 103

8.5 Schematic model of the Bleich-Sandler plate problem. . . . . . . . . . . . . . . . . . . 108

8.6 Bleich-Sandler plate problem results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

8.7 Koyna dam mesh used for cavitation analysis. . . . . . . . . . . . . . . . . . . . . . . . 110

8.8 The effects of cavitation on the relative displacement of the dam crest. . . . . . . . . . 110

8.9 Cavitation zone during selective time intervals with the Dam’s face on the right side. . 111

8.10 Cavitation zone during selective time intervals with the Dam’s face on the right side

without suppressing frothing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

8.11 The effects of cavitation on the relative displacement of the dam crest with the maximum

acceleration equal to 1.5g. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

xiii

8.12 The effects of cavitation on the relative displacement of the dam crest with the maximum

acceleration equal to 1.5g, but with the acceleration in the opposite direction than

previously. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

9.1 Morrow Point dam representation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

9.2 Ground motion and frequency content of the seismic acceleration recorded at the 1952

Taft Lincoln School Tunnel, California earthquake. . . . . . . . . . . . . . . . . . . . . 118

9.3 Morrow Point dam model used for this study. The fluid is not shown and the foundation,

shown in brown, is rigid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

9.4 Mapping of the subdomains nodes to an interface element for use in the zero-moment

rule. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

9.5 Interface nodes ”x” determined by the zero-moment rule. . . . . . . . . . . . . . . . . 121

9.6 Structure Mesh that showed Convergence. . . . . . . . . . . . . . . . . . . . . . . . . . 123

9.7 Relative displacement comparison for non-matching mesh versus matched mesh with

63 structure interface nodes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

9.8 The effects of cavitation on the relative displacement of the dam crest with the maximum

acceleration equal to 1 g. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

B.1 An example of how the LLM method, the Mortar method, and the Consistent Interpo-

lation method can produce the same relation matrices. . . . . . . . . . . . . . . . . . . 148

C.1 Discretization error for the portion of the constraint functional∫

ΓiλF i(uF i − uBi)dΓi . 150

Chapter 1

Introduction

The dynamic interaction between a fluid and a structure is a significant concern in many engi-

neering problems. These problems include systems as diverse as offshore and submerged structures,

storage tanks, biomechanical systems, inkjet printers, aircrafts, and suspension bridges. The interac-

tion can drastically change the dynamic characteristics of the structure and consequently its response

to transient, cyclic, and stochastic excitation. Therefore, it is desired to accurately model these diverse

systems with the inclusion of the fluid-structure interaction (FSI).

In the FSI problem, the fluid behavior can vary greatly among problems. A broad class of FSI

problems involves a fluid model without significant flow, and the main concern in the fluid is the

propagation of a pressure wave. A classic example of this type of FSI problem is a model of a dam

during seismic excitation. Dam failures are of particular concern because of the destructive power of

the flood wave that would be released by the sudden collapse of a large dam. The failure of the St.

Francis dam in 1928 resulted in a flood that destroyed over 1000 homes, and more then 400 people

perished. Recently, the citizens of Tauton, Massachusetts were concerned about the possibility of a

small wooden dam failing. Finally, the failure of the earth embankment levees of New Orleans has

caused devastating effects that will be felt for decades.

These examples were built before the advances in predictive computational simulations, and one

can only hope that through the advances in modern technology, today’s engineer will foresee potential

failures and design accordingly. One of the best tools at the fingertips of today’s engineer is the use

of a computational model. In the past few decades, the desire to efficiently design many systems has

resulted in a great surge in the creation of models for multi-physics phenomena, especially FSI. As

can be expected, the fluid model will have different characteristics than the structural model; thus,

complicating the computational model.

2

1.1 Thesis Overview

The focus of this study is to provide the engineer with a set of tools to accurately and efficiently

model the fluid-structure interaction phenomena with particular reference to the classic acoustic FSI

problem of a dam experiencing seismic excitation. The model-based simulation of this class of coupled

multi-physics systems presents three challenges.

The first is discretization heterogeneity. Effective space and time discretization methods for the

two interacting components, the structure and fluid, are not necessarily the same. This dilemma is

particularly pressing when one would like to use available but separate computer codes for the fluid

and the structure treated as individual entities, and use them to solve the coupled problem.

The second challenge is effective treatment of the interaction when the discrete structure and

fluid models do not necessarily match over the interface. Non-matching meshes can arise for several

reasons: one of the physical systems may require a finer mesh than another for accurate results; teams

using different programs generate the meshes separately; the systems were previously modeled for other

problems (i.e. incremental simulation of the structure construction process); or ensuring conformity

between meshes would require too much valuable time and effort in mesh generation.

The third challenge is forestalling performance degradation. Even if the separate models are

satisfactory with regard to stability and accuracy, the introduction of the interaction can have a

damaging effect on the coupled response. Furthermore, if the coupled components have significantly

different physical characteristics (stiffness, mass density, etc.) the coupled system may be scale-

mismatched by orders of magnitude. A poorly scaled discrete model can be the source of unacceptable

errors, particularly under long-term periodic or cyclic loading.

This thesis presents a new coupling method for treating the interaction of an acoustic fluid

with a flexible structure, with emphasis on handling spatially non-matching meshes. It is based on

the Localized Lagrange Multiplier (LLM) method [103]. The LLM method maintains the kinematic

continuity at the interface of the fluid and the structure by enforcing the jump of the displacements

at the interface with Lagrange multipliers to a third field boundary displacement. With the use of

the LLM method, an algorithm is developed for the dynamic transient analysis that does not require

the traditional predictor step that is common in most partitioned staggered time integration methods.

This is referred to as the LLM transient method. The removal of the prediction step is advantageous,

because of the stability issues that are associated with the prediction step. By using the LLM transient

method, an implicit time integration method can be used in order to have accuracy imposing the time

step and not stability, yet the outstanding benefits of a partitioned system are still realized. However,

3

different time integration methods can be easily adopted and applied in the time integration scheme

presented in this study.

In essence, the LLM transient method for FSI has three main modules: a fluid module, a

structure module, and an interface module. The interface module determines the necessary traction

forces (Lagrange multipliers) to maintain the kinematic continuity between the fluid and the structure

given the input force and the previous known state variables of the two subsystems. This assures a

conservative system as the Lagrange multipliers on one side of the interface are set equal an opposite

to the tractions on the other side; thus, obeying Newton’s third law. After the interface module has

determined the Lagrange multipliers, they are then used in the structure and fluid modules to advance

the state variables of these two systems. The structure and fluid modules are solved separately with

the interaction being communicated through the Lagrange multipliers.

In order to use the LLM transient method, the fluid and structural equations need to be repre-

sented as energy functionals. Therefore, the fluid equations are originally derived in a displacement

formulation to assure that the fluid functional is in the form of energy. Thus, the Euler equation with

small compressibility effects for the pressure term and small displacement considerations is ultimately

used to derive the fluid formulation. However, to enforce the irrotational condition and reduce the

computational cost, the fluid displacement model is transformed into a displacement potential model

by a concept termed the gradient matrix. The gradient matrix is created based on the gradient of

the displacement potential. Park et al. [101] proposed a similar transformation concept, except that

the fluid model was transformed into a pressure based model. This appears a little more complicated

because of the creation of the transformation matrix based on the divergence of the displacement, the

need to include an irrotational constraint term, and an inversion of the overall matrix. In addition,

the formation of the gradient matrix is performed at an element level with finite element concepts,

which creates a straightforward process.

In addition to the fluid formulation, the LLM method has been extended for transient dynamic

analysis of an acoustic FSI by the following items. The assumption that the fluid is inviscid has

required the consideration of the normal component. The interface solver can become ill-conditioned

and requires a scaling methodology that not only provides a well conditioned system of equations but

also can account for the non-matching meshes at the interface. Energy difference across the interface is

provided to alert of any issues with the analysis. A method for including non-linear effects is provided.

Finally, a simple reduced order formulation is examined for the LLM transient method. A further

benefit of the LLM concept is the ease in which a vibration analysis can be conducted with the newly

derived fluid formulation based on the displacement potential.

4

There are several other transient concepts for an acoustic FSI. Monolithic approaches are abun-

dant [15, 29, 132], but are not efficient when solving large systems. Wandinger [121] used a Craig-

Bampton method for reduction of the system of equations for coupled fluid-structure systems. Walsh

et al. [120] also began with a monolithic set of equations and use standard domain decomposition

strategies for parallel computation. However, the problems in their study had long thin regions of alter-

nating fluid and solid domains, which decrease the convergence rate of the Finite Element Tearing and

Interconnection (FETI) methods that they used. Thus, they did not desire a partition based on the

heterogeneous systems. Felippa and Deruntz [45] provided a staggered partitioned integration method

termed Cavitating Acoustic Finite Elements to handle cavitation effects and partitioned the fluid and

the solid domains. This was enhanced by Sprague and Geers [114] with their Cavitating Acoustic

Spectral Element (CASE) formulation, which is also used in this study to compare and contrast the

LLM transient method. This procedure necessarily incorporates predictors and has to be carefully

designed to avoid stability degradation. A transient dynamic method that has similar characteristics

to the LLM transient method is the method proposed by Herry et al. [70]. However, this method uses

the Mortar method in conjunction with Schur’s dual formulation for the non-matching meshes and

has yet to be extended to FSI.

All of these methods can benefit from the LLM method in terms of dealing with non-matching

meshes. At the heart of the LLM concept is the ability to link displacements and related state variables

across the interface of distinct subdomains. Discrete, collocated Lagrange multipliers simplify the

numerical integration required to create the connection matrices. An intelligent, yet simple method

for discretizing the interface frame coined the zero-moment rule [108] satisfies an interface patch test

criterion for the LLM concept. Therefore, the LLM method for relating non-matching meshes can be

used to create the coupling matrices used in other transient methods. This is demonstrated with the

CASE method for the problems in this study as well.

Without a doubt, there are several concepts that have been proposed that are suitable for creat-

ing the coupling matrices for non-matching meshes that can be extended to fluid-structure interaction

problems. One of the more popular concepts are penalty like methods. The standard penalty method

either requires a very large penalty parameter, destroying the condition number of the resulting ma-

trix problem, or, in case the condition number is to be retained, is limited to first order energy-norm

accuracy [65]. Therefore, Hansbo and Hermansson [65] proposed a variation of the penalty concept

that incorporates Nitsche’s method. Here, the penalty parameter is chosen from a perspective of sta-

bility and the lower bound is recommended. The jump of the displacements at the interface in this

method is not enforced to be zero and thus can lead to a non-conservative system. However, in a

5

vibration analysis the modes that have large kinematic continuity discrepancies are in the upper part

of the spectrum. Thus, this can easily be used for a vibration analysis, but might suffer in a transient

analysis.

Another broad category for non-matching meshes is the master-slave concept. In the standard

master-slave technique, the nodes on the slave boundary are constrained to lie on the master boundary.

Dohrmann et al. [32] extend the master-slave concept by modifying the slave boundary to ensure

satisfaction of the patch test. This assured that a node on the master boundary would not penetrate

or pull away from the slave boundary, which is a common problem in the standard master-slave

concept. This did require a modification to the stiffness matrix on the slave boundary and has yet to

be extended to FSI problems.

Consistent interpolation [38] schemes are popular in aerodynamics. The original concept had

the draw back of not being strictly conservative. However, Farhat et al. [38] extended the concept

to satisfy that the interface is physically conservative. In this study, a tight relationship between

this concept, the Mortar method, and the LLM method is shown. The major concern with all of

these methods is assuring that the coupling matrices do not become singular or lose the conservative

property. In aeroelastic problems, this is not a concern because the fluid mesh is typically much finer

than the structure mesh. However, problems addressed in this study may have different mesh relations.

There are also weighted-residual methods that form conservative data transfer between different

meshes. Here, the residual of the state variables is minimized at the interface. The Common-refinement

Based method leads to set of integral equations that ultimately form a linear system of equations that

are solved at each iteration [69]. The linear system of equations is solved for the target state variables

given the source state variables. A difficult aspect in this method is the selection of the area to

integrated over to form the necessary matrices.

One of the fastest growing weighted-residual methods are the Lagrange multiplier methods. In

these methods, the weight functions used to enforce the residual of the state variables are the Lagrange

multipliers that have the physical meaning of tractions. One of the more prominent methods is the

Mortar method. Recently, the Mortar method is associated with the Lagrange multiplier enforcing the

residual of the displacements of the subdomains at the interface [4, 8, 16, 20, 38, 41, 72, 125]. In con-

trast, the LLM method uses two separate Lagrange multipliers to enforce the residual of the subdomain

displacements with an interface frame displacement; therefore, creating a three-field method.

One of the more difficult aspects of using any Lagrange multiplier method is the methodology for

doing the cumbersome integration of products of functions on unrelated meshes [66]. For instance, first

assume the boundary between two subdomains can be separated into disjoint segments and on each of

6

these segments the jump of the displacements across the interface is enforced by Lagrange multipliers.

For simplicity, assume the Lagrange multipliers inherit the approximation and discretization of the

trace of one of the two subdomain’s interface boundary. Also assume this does not match the trace

of the other subdomain. Eventually there will be a requirement of integrating piecewise polynomials

between the Lagrange mesh and the unrelated mesh.

Hansbo et al. [66] avoids this complicated issue of integrating piecewise polynomials on unrelated

meshes by incorporating global polynomial multipliers. This results in a global coupling of all the

variables on the interface boundary of the disjoint segment. In a contact problem, there is a small

zone of contact and the global coupling will not cause the problem to grow excessively in size [66].

However, this could be an issue for problems with large contact surfaces such as the fluid-structure

interaction in an earthquake analysis of a dam.

A simple method to avoid this arduous task of integrating piecewise polynomials is to choose

discrete multipliers where the shape function of the Lagrange multipliers is the Dirac delta function.

However, this requires an educated choice of the discretization of the interface frame to assure stability

in the transient analysis. By conducting a simple vibration analysis, kinematic continuity issues can

be assessed to determine when problems of this nature will arise. This is shown in this study for the

gravity dam example problem. Therefore, this study provides simple rules for the discretization of the

interface frame for the Mortar, the Localized Lagrange Multiplier, and the Consistent Interpolation

methods. The use of the LLM method in conjunction with the zero-moment rule for the discretization

of the interface frame prevails as the logical choice, for it does have the ability to pass an interface

patch test.

This thesis begins by developing the necessary components for FSI with the use of Localized

Lagrange Multipliers. Then, a partitioned transient algorithm is presented. In addition, the use of the

LLM for coupling fluid domain variables with structure domain variables is also provided. The use of

these concepts is then extended to problems in this study to verify the added contributions.

1.2 Main Thesis Contributions

The main contribution of this thesis is the extension of the LLM method to acoustic fluid-

structure interaction problems. This can provide an efficient, partitioned transient algorithm. A

method was originally proposed by Park et al. [101]. However, no problem was actually solved in this

publication. During the implementation stage of the concept, it was discovered that using a pressure-

based transformation for the fluid was more complicated than using a displacement potential based

transformation. Therefore, a methodology for using the displacement potential is created in this study

7

that produces a new fluid formulation. In addition, the scaling methodology for the interface solver is

modified to incorporate non-matching meshes. Furthermore, a method for including non-linear effects

in the LLM transient method is presented and tested by studying cavitation. In addition, normal

displacements are required in the constraint functional, because the fluid is assumed inviscid. Finally,

with the new fluid formulation, simple vibration analyses are conducted.

Given the above additions, the fundamental LLM concept for non-matching meshes is then used

to develop coupling matrices that can be used in other acoustic FSI codes. These relation matrices are

demonstrated in the CASE method and compared to similar matrices created by the Mortar and the

Consistent Interpolation methods. This also provided comparisons with the LLM transient method.

Finally, an analytical piston problem with a silent boundary has been developed that resembles

problems in this study. A piston problem is typically used in aerospace engineering for the simple

demonstration of aeroelastic codes. Replacing the closed end with a silent boundary condition as if

the piston went on infinitely, extended the main concept, and parameters for water were used instead

of air.

1.3 Manuscript Organization

This manuscript begins with the derivation of the LLM concept for an acoustic fluid-structure

interaction. During this derivation, the fluid displacement based model is transformed into a displace-

ment potential model. Chapter 2 concludes with a set of governing equations for the FSI system.

Chapter 3 then develops a transient method using the set of governing equations. Then, the use

of the LLM concept for coupling in other FSI codes is presented. Chapter 3 concludes with an er-

ror discussion and error measures used throughout this study. Chapter 4 is used to verify the new

fluid formulation. A rigid wall is moved against the fluid and computational determined pressure

is compared against analytical results. In addition, the plane-wave approximation for the radiating

boundary condition is evaluated and an implementation for surface waves is shown. Chapter 5 derives

the analytical solution for a piston problem with an infinite boundary. This is then used to evaluate

the transient partitioned method and the discretization method of the interface frame for 3-D brick

elements. Next, a vibration analysis is performed on a gravity dam problem in Chapter 6. A modal

analysis demonstrates kinematic continuity problems that can arise with the Mortar method and the

LLM method. A simple set of rules is provided when using either of these methods to assure kinematic

continuity across the interface. The chapter ends with a theoretical verification of the concept through

the use of frequency response functions. In Chapter 7, transient analyses are performed on the gravity

dam problem with matching and non-matching interfaces to verify the concept. Chapter 8 examines

8

the ability of the LLM transient method to include popular computational methods of reduced order

modeling and non-linear effects. Then, an operational cost of the LLM method is discussed. Chapter 9

is used to verify the coupling concept of the LLM method on curved surfaces by exploring the transient

analysis of an arch dam during seismic excitation. Finally, a summary and possible future work is

presented in Chapter 10.

Chapter 2

Localized Lagrange Multiplier Method for Acoustic FSI

2.1 Introduction

This chapter provides a new approach for the interface coupling of an acoustic fluid-structure

interaction (FSI) with the use of Localized Lagrange Multipliers (LLM). The LLM method was orig-

inally derived through a variational formulation [98], which entails partitioning the overall system

into non-overlapping subsystems. Then the interface (ΓB) between the two subsystems is treated by

an interface displacement frame and a localized version of the method of Lagrange multipliers, see

Fig. 2.1. The Lagrange multipliers (λk) enforce the state variables of the partition models to that

of the frame. Thus, the multiplier fields enforce directly the interface compatibility and equilibrium

conditions, without dissipation mechanisms [98]. The overall concept is depicted in Fig. 2.1, where

the benchmark dam problem of this study is partitioned into an acoustic fluid, a dam, and a rock

foundation the dam rests upon.

Therefore, the method begins by deriving and summing the energy expressions of the subsystems,

the structure and the fluid in this case. Then, an interface constraint is identified and constructed into

an interface energy functional. Instead of requiring the partitioned systems’ displacements to be the

same, the two partitioned systems’ displacements are constrained to be equal to the global reference

interface frame displacement [101]. Then the interface energy functional is summed with the energy

functional of the subsystems to obtain the total energy functional.

Πtotal = ΠF + ΠS + ΠB , (2.1)

where ΠF and ΠS are the governing space-time functionals for the isolated fluid and structure, and

ΠB is the interface constraint functional that generates the fluid-structure interface conditions as

stationary conditions.

The variation of the total energy functional yields the governing equations of the system. The

governing equations can be discretized spatially and temporally to ultimately approximate the state

10

ΩF

ΩS

ΩD

(a) System

ΩF

ΩS

ΩD

(b) Partitioned systems

λD2

λS2

λD1λF

λFλS1

ΓB , Interface frame

ΩF

ΩS

ΩD

(c) LLM Method

Figure 2.1: LLM Concept: (a)Given a system. (b) Divide the system into subdomains. (c) Insert aninterface displacement frame that is linked by the Local Lagrange Multipliers.

11

of the system throughout time. The Lagrange multiplier fields are represented by delta functions

collocated at nodes of the interacting systems. This has the benefit that multiplier node values become

simple interaction node forces and may be applied directly as such to the coupled model meshes. The

frame displacements are discretized by piecewise linear shape functions. The node locations of the

frame can be determined by the zero-moment rule to assure that a constant stress can be passed from

one subdomain to the other.

After the spatial discretization process, a simple transformation procedure is used to obtain the

governing equations of the fluid in terms of the displacement potential from the basic displacement

formulation. This has two distinct advantages. The first is the reduction of the degree of freedoms

at each node in a two-dimensional and three-dimensional analysis. The second is the enforcement of

the irrotational condition, which removes any circulation modes. The transformation procedure for

obtaining the governing equations for the fluid in terms of the displacement potential is based on the

fact that gradient of the displacement potential generates the displacement field.

2.2 System’s Energy Functionals

As previously mentioned, the LLM method begins by partitioning the overall system into sub-

systems and deriving the subsystems’ energy functionals. The fluid-structure system can naturally be

partitioned into a fluid system and a structure system.

By selecting the displacement of the structure system as the master variable, the energy func-

tional for the structure is the total potential energy functional (ΠTPE). This is written below in

d’Alembert form with the inertial and damping forces expressed as modified body forces.

ΠS = ΠTPE =1

2

ΩS

σijǫijdΩS −∫

ΩS

uSi(bSi − ρS uSi − dS uSi)dΩS

−∫

ΓS

uSiTSidΓS,

(2.2)

where a subscript S refers to the structure, σij and ǫij are the stress and strain tensors, ρ represents

the density, dS is the structural damping parameter, ui represents the displacements, bi represents the

body forces, Ti is the surface traction, Ω represents the domain, Γ represents the physical boundary,

and the superscript dots designate time differentiation. The summation convention is used until the

domain is discretized for the finite element method.

”A functional is an integral expression that implicitly contains the governing differential equa-

tions for a particular problem” [29]. In order to maintain consistency for the LLM method the fluid

functional must be represented by the work of the system. Thus, the fluid derivation will begin with

12

the fluid displacement as the primary variable in order to assure that work is represented for the func-

tional. In addition, the fluid is assumed to be a linear acoustic fluid that is initially at rest; thus, the

fluid is compressible, irrotational, inviscid, constant density, and adiabatic. Finally, it is also assumed

that the displacements are so small that the acceleration is given by∂2ui∂t2

instead of∂2ui∂t2

+ uj∂ui∂xj

[54]. These assumptions are valid for the examples in this work.

In order to assure that the fluid functional is in the correct terms, the derivation of the fluid

begins by examining the rate of change of the total energy. The rate of change of the total energy is

equal to the rate at which work is being done, plus the rate at which heat is being added. The total

energy can be represented by the internal energy plus the kinetic energy. Therefore, the rate of change

of the total energy is equivalent to the rate of change of the principle of conservation of energy, and

may be written as [31]

D

Dt

ΩF

(ρF e+1

2ρF uF iuF i)dΩF =

ΓF

uF iTF idΓF+

ΩF

uF ibF idΩF −∫

ΓF

qinidΓF ,

(2.3)

where a subscript F refers to the fluid, e is the internal energy per unit mass, qi is the conductive

heat flux leaving the control volume, and ni is the unit outward normal. Note that the fluid is

derived in a lagrangian coordinate system. This has the following advantages: the elements can be

incorporated with structural computer codes; the resulting global coefficient matrix is symmetric and

positive definite; and it is easier to implement the interaction with the structure [63]. In addition, the

use of the lagrangian coordinate system is employed because of the acoustic fluid used in this study.

In order to simplify the above rate of change of the conservation of energy equation, Eq. (2.3),

four steps are taken. First, it is assumed that a disturbance to a linear acoustic fluid travels at a

sufficiently fast speed such that the heat conduction term maybe neglected [31]. Second, the internal

energy term is represented by the energy equation [31] with the assumption that the fluid is inviscid,

constant density, and adiabatic.

D

Dt(ρF e) = ρF

∂e

∂t+ ρF uF i

∂e

∂xi= −p∂uF i

∂xi, (2.4)

where p is the fluid pressure. Third, using the constitutive equation that expresses the small com-

pressibility of a liquid, the pressure can be represented as [46]

p = −K∂uFk∂xk

= −ρF c2∂uFk∂xk

, (2.5)

where K represents the bulk modulus, and c is the fluid speed of sound. Finally, the kinetic energy

term with the assumption of small amplitude motions becomes

D

Dt(1

2ρF uF iuF i) =

1

2ρF uF iuF i +

1

2ρF uF iuF i = ρF uF iuF i. (2.6)

13

Inclusion of the above simplifications in the rate of change of the conservation of energy equation,

Eq. (2.3), produces

ΩF

ρF c2

(

∂uFk∂xk

)(

∂uF i∂xi

)

dΩF +

ΩF

ρF uF iuF idΩF =

ΓF

uF iTF idΓF +

ΩF

uF ibF idΩF .

(2.7)

Integrating the above rate of change equation with respect to time yields the energy functional for a

linear acoustic fluid.

ΠF =1

2ρF c

2

ΩF

(

∂uF i∂xi

)(

∂uFk∂xk

)

dΩF +

ΩF

ρFuF iuF idΩF

−∫

ΩF

uF ibF idΩF −∫

ΓF

uF iTF idΓF .

(2.8)

The final piece of the puzzle for the total system energy functional is to identify the interface

constraint and functional. Once this is obtained, then the total system energy functional can be

obtained by summing the interface functional, the fluid functional, and the structure functional. In

proceeding with the LLM method [98], an interface constraint is identified, and with it an interface

functional is formed that is in the form of energy, since the system’s functional is in terms of energy. To

form the interface constraint, the boundary between the two subsystems is separated and an interface

frame is inserted, see Fig. 2.1. In the original LLM method, compatibility of boundary displacements of

two connected frames is enforced by flux fields [98]. Alturi [2], Tong [116], and Felippa [43, 44] proposed

and studied variations of this functional concept for the construction of hybrid finite elements for which

the interior displacements and interface forces are eliminated at the element level [48].

However, at this point care must be taken because of the inviscid fluid system. In a typical

structure-structure interaction with the LLM method, the constraint would require that all displace-

ments of a subsystem at the boundary be equal to the frame’s boundary displacement (uBi). This

would also be the constraint for a viscous fluid [37]. However, due to the inviscid property of the fluid

in this study, the constraint is that the normal displacements of the systems must be equal to the

normal displacements at the boundary [122].

ΠB =

ΓF B

λF ini(uF ini − uBini)dΓFB +

ΓSB

λSini(uSini − uBini)dΓSB , (2.9)

where B refers to the global boundary, ΓFB is the boundary between the fluid and the frame, ΓSB is

the boundary between the structure and the frame, and ni is the normal vector at that boundary.

Finally, the total energy functional is the sum of the subsystems’ energy functionals plus the

14

constraint functional. Once again, this can be done because all functionals are in the form of energy.

ΠTotal =1

2

ΩS

σijǫijdΩS +1

2ρF c

2

ΩF

(

∂uF i∂xi

)(

∂uFk∂xk

)

dΩF

−∫

ΩS

uSi(bSi − ρS uSi − dS uSi)dΩS −∫

ΩF

uF i(bF i − ρF uF i)dΩF

−∫

ΓS

uSiTSidΓS −∫

ΓF

uF iTF idΓF

+

ΓF B

λF ini(uF ini − uBini)dΓFB +

ΓSB

λSini(uSini − uBini)dΓSB .

(2.10)

2.3 Finite Element Implementation

At this point, it is easier to continue if the domains and the boundaries are discretized into

elements in the standard finite element fashion. Thus, numerical techniques can be used to solve the

problems posed in this study. This section will begin with the discretization of the domains. Then

discretization and numerical integration methods will be discussed for the boundary and associated

terms. This section concludes with a matrix form of the total system’s energy functional.

2.3.1 Discretization of the Partitioned Domains

Both the geometry and the state variables of the elements in the partitioned domains are inter-

polated with standard shape functions, N (linear in this study), such as,

u(e)Si = NSu

eS , u

(e)F i = NFueF , (2.11)

where the superscript e represents an element in the discretized domain. This provides the following

domain element matrices:

KeS =

ΩeS

BTEB dΩeS ,

CeS = dS

ΩeS

NTSNS dΩ

eS ,

MeS = ρS

ΩeS

NTSNS dΩ

eS ,

f eS =

ΩeS

NTSbS dΩ

eS +

ΓeS

NTSTS dΓ

eS ,

KeF = ρF c

2

ΩeF

(∇ · NF )T (∇ ·NF ) dΩeF ,

MeF = ρF

ΩeF

NTFNF dΩ

eF ,

f eF =

ΩeF

NTFbF dΩ

eF +

ΓeF

NTFTF dΓ

eF .

(2.12)

15

Thus, it is assumed that the structure for now behaves linear-elastic with a constitutive matrix (E),

and a strain displacement matrix (B). These elemental expressions can be assembled into a global

system in the standard finite element format. The fluid stiffness matrix is associated only with the

volumetric deformation of the fluid [74], not the bending type deformation which does not cause

volume change. Proper numerical integration is needed to assure that only volumetric deformation

is accounted for in the fluid. Therefore, reduced integration is used. The removal of the zero energy

modes associated with the use of reduced integration is discussed in Chapter 6.

2.3.2 Lagrange Multiplier Boundary Discretization

The association of the Lagrange multiplier boundary appears to be the main difference between

many Lagrange multiplier methods and variations of the methods. In the LLM method, the Lagrange

multipliers are associated with the distinct partitioned subdomain’s boundary. The LLM method

requires that the partitioned boundary displacements be the same as the interface frame displacements

[101]; thus, the localized Lagrange multipliers link the sub-domains to the interface frame. Therefore,

the Lagrange multiplier is localized, which produces two Lagrange multipliers at the interface. In

the popular Mortar method and similar variations, the Lagrange multipliers are associated at the

boundary and directly connect the two subdomains; thus, there is only one Lagrange multiplier at the

interface. In the Mortar method the constraint is that the subdomains boundary displacements be

made equal (uSB = uFB). The difference is illustrated in Fig. 2.2. Further discussion of the Mortar

method and its merits are discussed in Chapter 6 and Appendix B.

Interface Frame

λF λD

ΩF ΩD

(a) LLM MethodInterface Frame

λ

ΩF ΩD

(b) Mortar Method

Figure 2.2: Comparison of the LLM method to the Mortar method. (a) Localized Lagrange multipliers.(b) Classical Lagrange multiplier.

With the Lagrange multipliers associated with the distinct boundary domains, it is only logical

and simple to discretize the Lagrange boundary the same as the distinct domain boundary [98, 104],

as illustrated in Fig. 2.3. The simplest choice for multiplier interpolation is node-force collocation as

mentioned by Park et al. [97, 98, 48]. This is accomplished by selecting the distinct domain boundary

16

discretization for the Lagrange multiplier discretization, as previously mentioned, and then relate the

shape functions of the Lagrange multiplier to a Dirac delta function [48].

Fluid Domain Structure Domain

λF location λS locationInterfaceFrame

Figure 2.3: Lagrange Multipliers are collocated with subdomains boundary nodes.

λ(e)Si = NλSλ

eS , λ

(e)F i = NλFλ

eF , (2.13)

λ(xm)D(x− xm) =

λ(xm) ifx = xm

0 otherwise(2.14)

Thus,

NλSi = D(x− xi). (2.15)

This greatly simplifies the integration because of the following property of the Dirac delta function

[80].∫ d

−dδ(t− a)f(t)dt = f(a); −∞ ≤ −d≪ a≪ d ≤ ∞. (2.16)

Therefore, the following pseudo-boolean matrices are created from the appropriate integral terms in

the total system’s energy functional, Eq. (2.10) due to the finite element approximations,

ΓSB

λSiniuSinidΓSB ⇒ λTS

[

ΓeSB

NTλSn

eneTNS dΓeSB

]

uS = λTS [BeS ]uS,

ΓF B

λF iniuF inidΓFB ⇒ λTF

[

ΓeF B

NTλFneneTNF dΓ

eFB

]

uF = λTF [BeF ]uF ,

(2.17)

where BeS and Be

F are the structure and fluid Boolean matrices that also include the normal component.

These should not be confused with the structures strain-displacement matrix. Here, the integration

is carried out along the surface of the element of the structure or fluid boundary denoted by ΓSB or

ΓFB, which is different than integrating along the interface frame element boundaries. These element

matrices are assembled into global matrices BS and BF that are only associated with the boundaries.

Therefore, the matrices BS and BF collocate the Lagrange multipliers as point (concentrated) forces

with the subdomain’s displacement nodes, but also take into account the normal component required

17

because of the inviscid fluid assumption. It should be noted that if all elements of a column of the above

matrices are zero, then that column should be removed from the matrix for the following derivation.

A column may become zero due to the fact that we are only concerned with the normal displacements

at the boundary, because of the inviscid fluid assumption.

2.3.3 Interface Frame Discretization

The interface frame is also discretized into elements, where the interface frame elements are

one degree less then the domain elements, due to the surface nature of the boundary. The node

location of the elements on the interface frame is currently a field under study. Felippa, Park, and

Rebel [48] recommend determining the node locations by maintaining a constant stress state through

a method termed the zero-moment rule. Herry, DiValentine, and Combescure [70], and Combescure

and Faucher [28] recommend placing nodes at corresponding node locations for both the fluid domain

and the structure domain. Several authors recommend the discretization to be either with the fluid

domain discretization on the interface or the structure discretization on the interface, and generally,

the more refined mesh is chosen as the discretization. An analysis for the appropriate interface frame

discretization is performed in Chapters 5 and 6. At this current point, the zero-moment rule is the

only known method for passing the patch test. However, it is only useful for the LLM method and

not other Lagrange multiplier methods, such has the Mortar method. Therefore, a brief description is

provided at the end of this section, for further detailed analysis please see the following work of Rebel,

Park, and Felippa [108, 102, 103, 48].

Assuming an adequate discretization of the interface frame, the interface frame boundary dis-

placements are interpolated with standard shape functions (linear in this study).

u(e)Bi = NBueB. (2.18)

However, the elements of the boundary frame are once again one order lower than the elements of

the domain (structure or fluid), because the frame is a surface on the domain elements. In this

study, if the subdomain consists of three dimensional brick elements, then the boundary elements

are two dimensional, four-node quadrilateral elements. In addition, if the subdomain consists of two-

dimensional quadrilateral elements, then the boundary element is a two-node, piecewise linear frame

element.

With the approximations of the interface boundary displacements and the Lagrange multipliers,

the following connection matrices are created from the remaining appropriate integral terms in the

18

system’s energy functional, Eq. (2.10),

ΓB

λSiniuBinidΓB ⇒ λTS

[

ΓeB

NTλSn

eneTNB dΓeB

]

uB = λTS [LeS ]uB ,

ΓB

λF iniuBinidΓB ⇒ λTF

[

ΓeB

NTλFneneTNB dΓ

eB

]

uF = λTF [LeF ]uB ,

(2.19)

where LeS and LeF are the connection matrices that include the normal component. Here, the integra-

tion is carried out along the surface of the interface frame boundary elements, and is noted by ΓB .

This greatly simplifies and speeds up the integration, because of the integration property of the Dirac

delta function, which is used for the Lagrange multipliers’ shape functions. These element matrices

are assembled into global matrices LS and LF that are only associated with the boundaries. The Lk

matrix, where the subscript k refers to a partitioned system, serves to relate the interface frame degree

of freedoms to the particular subsystem’s boundary degree of freedoms with, once again, the effect of

the normal component. An example of computing the Lk matrix without the normal component can

be found in work by Park, Felippa, and Rebel [103]. It is the Lk matrix that handles the non-matching

mesh effect. It should be noted that if all elements of a row or a column of the above matrices are

zero, then that row or column should be removed from the matrix for the following derivation. A row

or column may become zero due to the fact that we are only concerned with the normal displacements

at the boundary, because of the inviscid fluid assumption.

Next, the node locations of the interface frame are determined by the zero-moment rule.

2.3.3.1 Zero-moment Rule for Interface Node Placement

The zero-moment rule originally emerged from work on contact-impact problems studied by

Rebel, Park, and Felippa [108, 102, 103]. Though, the majority of the work in this section on the zero-

moment rule is taken from their paper, A simple algorithm for localized construction of non-matching

structural interfaces [48]. The main concept of the zero-moment rule is to assure that the interface

forces (the Lagrange multipliers) can satisfy a constant stress state along the interface frame. This can

be thought of as assuring the patch test [131] at the interface boundary or an interface patch test. As

pointed out by Bergan and Nygard [10], the patch test was generally used for a posteriori evaluation;

however, here it is used a priori, which was also done by Bergan and Nygard.

Therefore, to begin the procedure, a computation of the interface frame’s forces that are asso-

ciated with a constant stress state σc along the interface is needed. In order to obtain these frame

forces, the following procedure is provided [48]:

(1) Select a layer of elements along the interface of each partitioned subdomain.

19

(2) Given a typical element (e), obtain the strain-displacement relation Be and evaluate this at

the element centroid to get Be(0).

(3) The contribution of the element to the constant stress node forces is

fe = V [Be(0)]Tσc, (2.20)

where V denotes the volume, area or length of the element depending on its dimensionality.

(4) The interface forces at each partitioned subdomain k to be used in the interface patch test

can be obtained as

λk(σc) = LTfb, fb = AT

b f , f = [(f (1))T (f (2))T . . . (f (n))T]T, (2.21)

where L is a boolean extractor of the interface nodal degrees of freedom, Ab is the assembly

matrix that maps the elemental contributions into boundary node forces, and n is the number

of interface forces.

Given the constant stress state forces from the subdomains, the interface frame nodes can be

determined in order to maintain a constant state of stress from one domain to another. This begins by

first examining the variation of the boundary functional in regards to the variation of the boundary

displacement.

δΠB(uBi) = −∫

ΓB

δuBiλBidΓB , (2.22)

where λBi refers to the boundary forces, which is a sum of the partitioned subdomains’ interface forces.

Upon discretization of the boundary displacements and collocating the subdomains’ boundary forces

at their respective boundary nodes, the following resultant force nj, moment mj , and corresponding

frame displacements acting at a frame point xBj = (xj , yj , zj) will arise.

λBj =

nj

mj

, uBj =

uxj

uθj

, (2.23)

where the moment is taken into account, because the subdomains’ interface forces create internal

rotation on the frame. The resulting force and moment are determined by mapping the individual

subdomain interface forces onto the frame and evaluating the force and moment at the frame point,

as depicted in Fig. 2.4 for a 2-D case. Thus, if there are M frame nodes, the stationary condition

requires the following:

δΠB(uB) =

M∑

j=1

λTBδuBj =

M∑

j=1

nTj δuxj +

M∑

j=1

mTj δuθj = 0 (2.24)

20

Fluid Domain

Structure Domain

λF

λS

Frame

(a) Interface constant stress forces

λF

λS

Frame

(b) Mapped forces onto frame line

shaded portion used for

computing the resultant

force and moment at

frame point j.

λF

λS

nj

mj

(c) Resultant force and moment deter-mined at frame point j

Figure 2.4: Graphical representation of the determination of the interface nodes in 2-D.

21

When the subdomains’ forces are mapped onto the frame, considered as a free body, the frame must

be in self-equilibrium. In addition, it should not experience any deformational energy. Thus, the only

admissible frame displacements that would not cause any deformation on the frame are the frame’s

self-equilibrium modes, which are also the rigid-body modes of the frame. Therefore, Eq. (2.24),

becomes

δΠB(uB) =

M∑

j=1

nTj

δαx +

M∑

j=1

mTj

δαθ = 0, (2.25)

where αx and αθ are the translation and rotational rigid-body amplitudes of the frame.

It is important to note that for the problems in this study, the normal force is the only force

acting. Therefore, the force nj and moment mj acting at the frame point j of coordinates xj, can be

expressed as

nj

mj

=∑

s

Is

χTs

nks, (2.26)

where nks refers to the particular subdomains’ (k) boundary force mapped onto the frame at point s,

Is =

ixs 0 0

0 iys 0

0 0 izs

, ixs =

1 if xj − xks ≥ 0

0 if xj − xks < 0and similarly for other expressions, (2.27)

and

χs =

0 −(zj − zks) (yj − yks)

(zj − zks) 0 −(xj − xks)

−(yj − yks) (xj − xks) 0

if (xj ≥ xks, yj ≥ yks, zj ≥ zks)

and χ = 0 otherwise(2.28)

where (xks, yks, zks) are the locations where the subdomains interface forces are mapped onto the

frame. Also the localized Lagrange multipliers of each domain mapped onto the frame are restacked

so that they are ordered from min(xB) to max(xB). Thus, for the 2-D case, the force nj and moment

mj are readily obtained from the contributions of the shaded area to the left of j, as shown in Fig. 2.4.

Consequently, Eq. (2.25) with the above Eqs. (2.26, 2.27, and 2.28) becomes

δΠB(uB) =

(

Ntotal∑

s=1

nTksIs

)

δαx +

M∑

j=1

(

N∑

s=1

nTksχs

)

j

δαθ = 0, (2.29)

where N is the number of mapped nodes contributing to the frame point j. Since, δαx and δαθ are

independent, the resulting conditions are:

Translational force equilibrium:

Ntotal∑

s=1

nTks = 0,

Moment equilibrium:

M∑

j=1

(

N∑

s=1

nTksχs

)

j

= 0.

(2.30)

22

Therefore, the locations of the frame nodes are those points j that satisfy the moment equilibrium

condition, Eq. (2.30). Clearly, this can be computationally expensive and challenging if all nodes are

searched for simultaneously. Hence, it is advised to incrementally sweep the frame area and identify

one frame node at a time. For instance, in a 2-D problem, one would begin at an end of the frame

and progress to the other end, as shown in Fig. 2.4. Thus, the moment equilibrium equation is solved

at each frame node.(

N∑

s=1

nTksχs

)

= 0, (2.31)

In summary, the frame nodes are located at the roots of the moment equilibrium equation. For

curved surfaces in this study, the physical surface and physical subdomain forces from a constant

stress, Eq. (2.21), are mapped into a reference coordinate system (ξ, η), similar to the finite element

method. Then, the frame node locations are determined and mapped back to the physical surface. It

is also important to note that if a frame area has a change in its configuration then each configuration

needs to be analyzed in this methodology. For instance, in a two dimensional analysis, assume that the

intersection of two subdomains (Γij = ∂Ωi ∩ ∂Ωj) can be decomposed into a set of disjoint straight-

line pieces (γi). Thus, each straight-line piece is analyzed in this method. Finally, it should be

noted that the zero moment rule has been proven for the assumption that the Lagrange multipliers

are collocated with the subdomains’ nodes with shape functions composed of a Dirac delta function,

Eq. 2.15. Once again, the description in this section on the zero-moment rule was taken from Rebel,

Park, and Felippa’s work [48].

2.3.4 Lagrange Multiplier Shape Functions Revisited

Given the above analysis, we take a moment to revisit the shape functions of the Lagrange

multipliers (λ(e)Si = NλSλ

eS , λ

(e)F i = NλFλ

eF ). In Section 3.4, it is noted that there is the potential for

mathematical optimality if the Lagrange multiplier’s shape functions are linear interpolations among

the frame elements. In other words, the interface error induced by the method on the solution of the

coupled subdomains’ model problem is not worse than the local subdomains’ discretization errors [38].

The difficult aspect of obtaining this mathematical optimality is in the integration of the term

L:∫

ΓB

λSiniuBinidΓB ⇒ λTS

[

ΓeB

NTλSn

eneTNB dΓeB

]

uB = λTS [LeS]uB, (2.32)

and similar for the fluid equivalent term. In order to perform this integration with linear shape

functions for the Lagrange multipliers, one would first want to integrate over the area of the straight

23

line pieces, γi, of the interfaces rather than the boundary elements, see Fig. 2.5 for a 2-D example.

LS =

γi

NTλSnnTNB dγi, ∀i = 1, ..., Nγ , (2.33)

where Nγ is the number of straight line pieces that make up the interface of the subdomains. In

Fig. 2.5, the integration is depicted for linear shape functions and Dirac delta shape functions, for

a straight line piece γi in 2-D with four Lagrange multiplier nodes and five subdomain nodes. The

subdomain could be thought of as either the structure or the fluid. Also, the figure assumes that the

integration is performed over the normal components; thus,

λk(x) =[

Nλ1 Nλ2 Nλ3 Nλ4

]

λ1

λ2

λ3

λ4

, uB(x) =[

NB1 NB2 NB3 NB4 NB5

]

uB1

uB2

uB3

uB4

uB5

. (2.34)

Also, in Fig. 2.5, only the components associated with the Lagrange multiplier shape function Nλ2 are

depicted.

One can easily see that using standard shape functions for the Lagrange multipliers will mean

integrating products of piecewise polynomials on unrelated meshes. As noted by Hansbo et al [66],

this is not easily done in practice for problems in R3. The complication of this integration is avoided

in this study by using the Dirac delta shape functions. However, this does come with a sacrifice in the

order of the discretization error as discussed in Section 3.4.

2.3.5 Matrix Form of the Total System Functional

Finally, given the finite element discretization of the domains and the interface frame, Eq. (2.10)

can be expressed in matrix form as

ΠTotal = uTS (1

2KSuS + CSuS + MSuS − fS)+

uTF (1

2KFuF + MF uF − fF )+

λTS (BTSuS − LSuB) + λTF (BT

FuF − LFuB).

(2.35)

The next step in the LLM method would be to take the variation of the system functional to obtain

the governing equations. However, before proceeding, it is advantageous to take into account the

irrotational assumption of the fluid by use of the displacement potential.

24

1 2 3 4 5

1 2 3 4

γiNλ2NB1dγi

γiNλ2NB2dγi

γiNλ2NB3dγi

γiNλ2NB4dγi

γiNλ2NB5dγi

= 0

= 0

= 0= 0

= NB2(x)

= NB3(x)

Linear Nλ Dirac delta Nλ

Area for integration 6= 0

Boundary nodes

Subdomain nodes

Figure 2.5: Integration over the straight line piece γi.

25

2.4 Inclusion of Irrotational Assumption by the Displacement Potential

2.4.1 Circulation Modes

Up to this point, the fluid equations are in terms of fluid displacements. This choice, however,

brings on a serious computational difficulty: the appearance of spurious kinematic modes. Since an

acoustic fluid is irrotational and inviscid, its internal energy should respond only to volumetric changes.

As a result displacement-based elements can become highly rank deficient.

As previously mentioned, the fluid stiffness matrix uses reduced integration, only to account for

the volumetric deformation. Unfortunately, this leads to element stiffness matrices which are rank-

deficient and thus, produce extra zero-energy (spurious modes) [21]. It has been widely reported that

some of these spurious modes are due to circulation modes [11]. With the assumption that the fluid

is irrotational, one would want to remove these circulation modes.

Hamdi et al. [64] and Wilson et al. [124] proposed a penalty method in order to prescribe

irrotationality on the displacements to remove these circulation modes; this was successful for the

cases considered in those references. However, with the penalty method it reduces the zero frequency

spurious modes into higher frequency spurious modes with the hope that only the lower modes will be

excited. In addition, with a coarse mesh there may be spurious modes with nonzero frequencies due

to the fact that the mass matrix is fully integrated. To address these issues Chen et al. [21] and Kim

et al. [74] proposed the use of the penalty method and a projected mass matrix. This allowed for the

retention of low frequency modes while removing the spurious modes. In the cases studied by Kim

et al. [74], these low frequency modes were attributed to the effect of sloshing. Unfortunately, this

method is complex and requires iteration to determine an appropriate penalty value [124]. Bermudez

et al. [12] proposed the use of Raviart-Thomas elements to remove the circulation modes. Here,

the degrees of freedom for the fluid correspond to the mean fluxes of the displacements through the

element faces. This would require a modification to the standard structure finite element formulations,

if it were to be used with the LLM method.

Another method for removal of the circulation modes is to formulate the fluid governing equa-

tions in terms of the displacement potential as originally devised by Newton [90], [89], [91], [92]. As

demonstrated by Felippa and Deruntz [45], Zienkiewicz et al. [130], and Sprague and Geers [113], this

formulation enforces the irrotationality constraint, and can easily account for cavitation. By assuming

the fluid is irrotational, the fluid displacement can be described by the gradient of a scalar known as

the displacement potential.

uF i = ∇ψ. (2.36)

26

In the models developed by the above authors, the fluid equations were developed with the dis-

placement potential before the finite element discretization. However, during the formulation of the

governing equations with the use of the displacement potential, an equation was spatially integrated.

Thus, if one were to formulate a functional with these governing equations, it would not be in terms of

energy; hence, a problem with incorporating the LLM method in this energy variational formulation.

A method to include the LLM method with these formulations is included in Section 3.3.

2.4.2 Gradient Matrix

In order to utilize the benefits of both methods, where irrotational effects are constrained by

the use of the displacement potential, and the functional for the governing equations is in terms of

energy, the following method is performed. First, the fluid displacement, which is all ready discretized

at each node, is represented in terms of the displacement potential at each node by a gradient matrix,

D; thus,

uF i = ∇ψ =⇒ uF = Dψ. (2.37)

Here, the degree of freedoms are reduced at each node if the model is in two or three dimensions.

The gradient matrix, D, converts nodal displacement potentials into nodal displacement values. The

nodal displacement potentials correspond to the displacement potential values at the nodal locations

of the discretized fluid. Thus, the gradient matrix can be created by spatial numerical differentiation.

For instance, by knowing the displacement potential values at a node and the surrounding nodes,

various numerical differentiation techniques can be used to numerically differentiate the displacement

potentials in the corresponding directions to obtain the displacement values in those directions. Thus,

the gradient matrix can be formed.

The simplest method to create the gradient matrix is to take advantage of the Finite Element

method and previously existing software and matrices. This approach is best described by an example.

First, assume for simple clarification, that the fluid is in two dimensions and is discretized with bilinear

quadrilateral elements. A typical bilinear quadrilateral element would appear as depicted in Fig. 2.6.

At node 1, the nodal displacements, u1x and u1

y, are determined by using the nodal values of the

displacement potentials (ψ1, ψ2, ψ3, and ψ4); thus,

u1x =

∂ψ1

∂x, u1

y =∂ψ1

∂y. (2.38)

In order to solve this, we begin with the use of the chain rule with the element axes ξ and η,

∂ψ1

∂ξ=∂ψ1

∂x

∂x

∂ξ+∂ψ1

∂y

∂y

∂ξ,

∂ψ1

∂η=∂ψ1

∂x

∂x

∂η+∂ψ1

∂y

∂y

∂η. (2.39)

27

ψ1ψ2

ψ3

ψ4

x

yu1x

u1y

ξ

η

Figure 2.6: Bilinear quadrilateral element

This can be written in matrix notation as

∂ψ1

∂ξ

∂ψ1

∂η

=

∂x∂ξ

∂y∂ξ

∂x∂η

∂y∂η

∂ψ1

∂x

∂ψ1

∂y

= J

∂ψ1

∂x

∂ψ1

∂y

, (2.40)

where J is the jacobian. In order to solve for∂ψ1

∂xand

∂ψ1

∂y, use

∂ψ1

∂x

∂ψ1

∂y

= J−1

∂ψ1

∂ξ

∂ψ1

∂η

. (2.41)

A forward difference scheme is used for the differentiation of the unknown values on the right hand

side of the above equation; thus,

∂ψ1

∂ξ=ψ2 − ψ1

l12,

∂ψ1

∂η=ψ4 − ψ1

l14, (2.42)

where l12 is the distance between nodes 1 and 2, and l14 is the difference between nodes 1 and 4,

which in both cases will equal two in typical natural coordinates. This relation and the inverse of the

Jacobian provide the required partial derivatives, u1x =

∂ψ1

∂xand u1

y =∂ψ1

∂y.

To further fix the idea, consider once again the 4-node quadrilateral displacement-potential-

based fluid element pictured in Fig. 2.6. The geometry is defined in terms of the usual natural

coordinates ξ and η. The element has four degrees of freedom, which are the displacement potentials

ψi at the corners i = 1, 2, 3, 4. Coordinates x, y and the displacement potential ψ are interpolated

isoparametrically:

1

x

y

ψ

=

1 1 1 1

x1 x2 x3 x4

y1 y2 y3 y4

ψ1 ψ2 ψ3 ψ4

N1

N2

N3

N4

, (2.43)

28

with the bilinear shape functions N1 = 14(1− ξ)(1−η), N2 = 1

4 (1+ ξ)(1−η), N3 = 14(1+ ξ)(1+η) and

N4 = 14(1− ξ)(1+ η). Corner coordinate differences are abbreviated as xij = xi−xj and yij = yi− yj.

The Jacobian determinant is J = 14(A + A1ξ + A2η) with A = 1

2 (x31y42 − x42y31) is the area of the

quadrilateral, A1 = 12 (x34y12 − x12y34) and A2 = 1

2 (x23y14 − x14y23).

In terms of the node displacement 4-vector ψe = [ψ1 ψ2 ψ3 ψ4]T the Cartesian gradients ux =

∂ψ/∂x and uy = ∂ψ/∂y are given by

ux

uy

=1

8J

y24 y31 y42 y13

x42 x13 x24 x31

+

y43 y34 y12 y21

x34 x43 x21 x12

ξ +

y32 y14 y41 y23

x23 x41 x14 x32

η

ψe.

(2.44)

Evaluating (2.44) at the nodes yields

ueF =

ux1

uy1

ux2

uy2

ux3

uy3

ux4

uy4

=1

4

y24/J1 y41/J1 0 y12/J1

x42/J1 x14/J1 0 x21/J1

y23/J2 y31/J2 y12/J2 0

x32/J2 x13/J2 x21/J2 0

0 y34/J3 y42/J3 y23/J3

0 x43/J3 x24/J3 x32/J3

y34/J4 0 y41/J4 y13/J4

x43/J4 0 x14/J4 x31/J4

ψ1

ψ2

ψ3

ψ4

= DeF ψ

e, (2.45)

in which Ji are the corner Jacobians. These can be rapidly calculated from

4J1 = A−A1 −A2 = x14y21 − x21y14 = x42y14 − x14y42,

4J2 = A+A1 −A2 = x21y32 − x32y21 = x13y21 − x21y13,

4J3 = A+A1 +A2 = x32y43 − x43y32 = x24y32 − x32y24,

4J4 = A−A1 +A2 = x43y14 − x14y43 = x31y43 − x43y31.

(2.46)

The 8 × 4 gradient matrix DeF relates the Cartesian node displacement components to the node

displacement potentials. A similar transformation: ueF = DeF ψ

e can be constructed for any fluid

element based on the displacement potential element by appropriate differentiation and evaluation at

nodes.

Therefore, as the elements are assembled the global gradient matrix (D) can be created. How-

ever, instead of summing the contributions, an average must be taken. Once again, an example can

best explain. Assume the fluid is discretized with two bilinear quadrilateral elements as depicted in

Fig. 2.7. Notice for this example the element axes are aligned with the global axes.

29

ψ1 ψ2 ψ3

ψ4ψ5ψ6

ξ, x

η, y

Figure 2.7: Two connected bilinear quadrilateral elements

In this example the x-direction displacement (u2x) of node 2 will be determined. For the first

element on the left,

u2x =

ψ2 − ψ1

l12. (2.47)

The second element (on the right) contributes the following:

u2x =

ψ3 − ψ2

l23. (2.48)

In assembling, the average is taken therefore, the total contribution is

u2x =

ψ2 − ψ1

2l12+ψ3 − ψ2

2l23. (2.49)

If the lengths between the nodes are the same l12 = l23, then the central difference formula is used.

ψ2 − ψ1

2l12+ψ3 − ψ2

2l23=ψ3 − ψ1

2l12(2.50)

Therefore, the D matrix would have the following form, where uF = Dψ:

u1x

u1y

u2x

...

u6y

(12x1)

=

− 1l12

1l12

0 0 0 0

− 1l16

0 0 0 0 1l16

− 12l12

(

12l12

− 12l23

)

12l23

0 0 0

......

......

......

− 1l16

0 0 0 0 1l16

(12x6)

ψ1

ψ2

ψ3

ψ4

ψ5

ψ6

(6x1)

(2.51)

Here, it can be noted that generally the error will be of order (h) for the nodes on the edges and

of order (h2) for interior nodes. This is seen because a Forward Euler finite difference method is used

for the derivatives on the boundaries, and a Central Difference method is used for the interior nodes,

if the elements are of equal length.

30

This representation of the fluid displacement, uF = Dψ, is then inserted into the total functional

equation, Eq. (2.35), to produce:

ΠTotal = uTS (1

2KSuS + CSuS + MSuS − fS)+

ψT (1

2DTKFDψ + DTMFDψ − DT fF )+

λTS (BTSuS − LSuB) + λTF (BT

FDψ − LFuB).

(2.52)

By applying it in this manner, the rotational modes that were associated with the mass matrix are

removed, because it did not use reduced integration. This concern was addressed with a projected

mass matrix by Chen et al. [21] and Kim et al. [74]. To simplify further derivations, the following

matrices are defined:

Mfd = DTMFD, Kfd = DTKFD. (2.53)

2.5 The Variation of the Energy Functional

The variation of the functional in Eq. (2.52) leads to the governing equations of the system.

δΠ(uS ,ψ,λS ,λF,uB) =

δuTS (MS uS + CSuS + KSuS + BSλS − fS)+

δψT (Mfdψ + Kfdψ + DTBFλF − DT fF )+

δλTS (BTSuS − LSuB) + δλTF (BT

FDψ − LFuB)+

δuTB(−LTSλS − LTFλF )

(2.54)

The sum of all the terms that are multiplied by the same differential are equated to zero, due to the

stationary condition. This leads to the following equation set:

KMS 0 BS 0 0

0 KMfd 0 DTBF 0

BTS 0 0 0 −LS

0 BTFD 0 0 −LF

0 0 −LTS −LTF 0

uS

ψ

λS

λF

uB

=

fS

DT fF

0

0

0

, (2.55)

where,

KMS = KS + CSd

dt+ MS

d2

dt2, KMfd = Kfd + Mfd

d2

dt2.

As noted by Park et al [101], we can make the following observations. First, the Lagrange multipliers

handle the interface interaction. Second, the partition subsystems’ boundary displacements (uS ,

uF = Dψ) are equated to the interface frame displacements (uB) as seen in row three and four of

31

Eq. (2.55). Finally, the Lagrange multipliers at a node obey Newton’s third law as seen in row five of

Eq. (2.55).

2.6 Summary

This chapter developed a set of governing equations for an acoustic fluid-structure interaction

system. The set of governing equations uses Localized Lagrange Multipliers for weak enforcement of

kinematic variables. The process begins by deriving energy functionals for the fluid, the structure, and

the interface of the two domains. The total functional of the system is the sum of the three energy

functionals. Components of the total functional are spatially discretized with finite element methods.

Standard linear shape functions are used for the displacement variables. However, a standard linear

shape function for the Lagrange multiplier field would lead to cumbersome integration of products of

functions on unrelated meshes. A simplification consists of taking multipliers as the conjugate forces

of the node displacements of nodes located on the interaction surface. If so, the multiplier shape

functions become simply node collocated delta functions.

Given the spatial discretization of the fluid, a gradient matrix is formed that allows for the

governing equations of the fluid to be transformed in terms of a displacement potential. This gradient

matrix is formed at the element level and assembled into a global matrix. However, the average

contribution of an element must be used for a specific displacement.

Chapter 3

Transient Concepts Using Localized Lagrange Multipliers

3.1 Introduction

This chapter presents transient methods using the previous developed coupling concept with Lo-

calized Lagrange Multipliers (LLM). First, a transient analysis is formed from the governing equations

developed with the use of the LLM. Then, other established transient methods are explored with the

use of LLM. Discretization errors and error measures for the transient analysis conclude this chapter.

The first section in this chapter develops a transient analysis method given the semi-discrete

governing equations, Eq. (2.55), with the LLM. The semi-discrete governing equations are integrated

in time using a partitioned analysis procedure. Direct time integration is used in preference to a

modal response analysis to preserve the ability to do nonlinear analysis, as in modeling cavitation.

The governing equations for the fluid, structure, and interface partitions are processed separately. At

each time step, the interface equations receive vector information from the interacting systems and

are algebraically solved for frame displacements and Lagrange multipliers. The multiplier results are

broadcast to the fluid and structure partitions, and used to update displacements and displacement

potentials for the next step. Any convenient solver for the fluid and structure partitions may be used,

and need not be the same. The interface set of equations have the potential to be ill-conditioned due

to scaling issues. This is resolved by scaling the interface system of equations.

The coupling concept of the LLM method can be used in other partitioned transient methods.

This is demonstrated in the next section of this chapter. The coupling equations for the LLM method

are used to develop relations from the fluid pressure to the structure force, and from the structure

displacements to the fluid displacements.

The chapter concludes with a section on error. Only the discretization error at the boundary

interface is discussed. Then, methods for measuring the transient error are discussed, which include

Geers’ comprehensive error factor [114] and energy differences at the interface.

33

3.2 Partitioned Transient Analysis

This section begins by developing a transient analysis of the system given the semi-discrete

governing equations, Eq. (2.55). This ultimately leads to an interface equation that is used for the

transient analysis. This interface equation solves for the Lagrange multiplier values at a certain time

step. These values are then used to update the structure and fluid variables (uS and ψ).

A time discretization of Eq. (2.55) is performed to begin the transient analysis. Thus, a numerical

time-stepping method for integration of the structure and fluid differential equations is used. The time

stepping method chosen for the problems in this study is the popular Newmark method [25], which

is heavily used in earthquake engineering. The Newmark method approximates the displacement and

velocity at the next time step (n+ 1) by

un+1 = un + ∆tun +1

2∆t2

[

2βun+1 + (1 − 2β)un]

,

un+1 = un + ∆t[

γun+1 + (1 − γ)un]

,

(3.1)

where ∆t is the time step value, n is the current time step, n + 1 is the next time step, β and

γ are parameters that determine the stability and accuracy characteristics of the algorithm under

consideration, see Table 3.2 [71].

Table 3.1: Properties of well-known members of the Newmark method

Method Type β γ Stability condition Order of accuracy

Trapezoidal Rule Implicit 14

12 Unconditionally 2

Linear Acceleration Implicit 16

12 Conditional 2

Fox-Goodwin Implicit 112

12 Conditional 2

Central Difference Explicit 0 12 Unconditional 2

Inserting the above approximations into the first and second rows of Eq. (2.55), and moving all

terms except the values of the structure displacement and the fluid displacement potential to the right

hand side yields

un+1S = FS(gn+1

S − BSλn+1S ), (3.2)

ψn+1

= FF (gn+1F − DTBFλ

n+1F ), (3.3)

34

where

FS =[

MS + ∆tγCS + ∆t2βKS

]

−1,

gn+1S = fn+1

S − CSunS + ∆t(1 − γ)unS − KSunS + ∆tunS +∆t2

2(1 − 2β)unS,

FF =[

Mfd + ∆t2βKfd

]

−1,

gn+1F = DT fn+1

F − Kfdψn + ∆tψn

+∆t2

2(1 − 2β)ψ

n.

(3.4)

Note that the values in gn+1F and gn+1

S are known, assuming that the forcing values are known.

Next, the second time derivative of rows three and four of Eq. (2.55) are evaluated at time step

(n+ 1). Then Eq. (3.2) and Eq. (3.3) are inserted for the values of un+1S and ψ

n+1to yield

BTSFSBSλ

n+1S + LSu

n+1B = BT

SFSgn+1S , (3.5)

BTFDFFDTBFλ

n+1F + LF un+1

B = BTFDFFgn+1

F , (3.6)

with unknown variables on the left hand side.

Finally, using the two above equations (3.5 and 3.6) and the remaining row of Eq. (2.55) at time

step (n + 1), the following equation set is obtained.

BTSFSBS 0 LS

0 BTFDFFDTBF LF

LTS LTF 0

λn+1S

λn+1F

un+1B

=

BTSFSg

n+1S

BTFDFFgn+1

F

0

(3.7)

This is referred to as the Interface Equation. Note that this equation set is only as large as the degrees

of freedom’s of the three state variables at the boundary; thus, the benefit in reducing the degrees of

freedom at the interface.

In order to solve the transient analysis, the Lagrange multipliers are determined by solving the

Interface equation, Eq. (3.7). These known values are then inserted into Eq. (3.2) and Eq. (3.3).

Therefore, the next time step values of the structure acceleration and the fluid second time derivative

of the displacement potential can now be determined. Then Eq. (3.1) can be used to obtain the

displacement and the displacement potential. With these values, the process can be repeated and

continued until the time period of the evaluation is completed. Notice that in this staggered procedure

there is no prediction step, which can cause loss of accuracy and instability. However, care must be

taken in solving the Interface equation, Eq. (3.7), because it can be ill-conditioned. Therefore, the

Interface equation set is scaled in a manner similar to the method by Yue [127] and discussed in

Section 3.2.1 to resolve the ill-conditioning. Finally, it is pointed out that this transient procedure

allows for the use of different solvers for the fluid system and the structural system. A diagram

illustrating the time stepping procedure for the Trapezoidal Rule is depicted in Fig. 3.1.

35

InterfaceInterfaceInterface

StructureStructure

FluidFluid

etc.

etc.

t0 t1 t2time steptime steptime step

f1 f2 f3

ψ1

= FF (g1F − DTBFλ

1F ) ψ

2= FF (g2

F − DTBFλ2F )

u1S = FS(g1

S − BSλ1S) u2

S = FS(g2S − BSλ

2S)

ψ1, ψ

1, ψ1 ψ

2, ψ

2, ψ2

u1S , u1

S, u1S u2

S , u2S, u2

S

λ1F , f1

λ2F , f2

λ1S , f1 λ2

S , f2

Figure 3.1: Graphical representation of the time stepping procedure for the LLM method.

36

If the nonlinear effect of cavitation is to be included, then it is recommended to use the Central

Difference method. This is recommended because of the removal of the stiffness matrix in the matrix to

be inverted, to be discussed in Section 8.2. Unfortunately, care must be taken if the Central Difference

method is used, because of the stability concerned with the size of the time step value. In the Central

Difference Method, which is an explicit direct integration method, the time step must be small enough

that information does not propagate more than the distance between adjacent nodes during a single

time step. This is referred to as the CFL condition [29]. A convenient measure to assure this for fluids

is the Courant number [51].

The transient analysis provides some clear benefits of the LLM method. First, the partitioned

analysis can be solved without the normal need for a prediction step. This is advantageous because

typically the prediction step in staggered integration procedures can cause instability. Second, the

fluid, structure, and the interface equation can be solved by separate solvers tailored to the particular

system. Third, the fluid and the structure can be solved in parallel. Finally, the interface equation

will generally be a smaller system set to solve and should be faster. It should also be noted that

the transient analysis is solved for the second time derivatives of the state variables, because of the

advantage of determining cavitation, which is discussed in Section 8.2

3.2.1 Scaling the Interface Equation

As previously mentioned, care must be taken when evaluating the Interface Equation, Eq. (3.7),

because it can be an ill-conditoned equation. The ill-conditioned nature is of great concern in solving

the linear algebra system given by the Interface Equation, because a small perturbation in the right

hand side vector can cause a large change in the solution vector [3]. The ill-conditioned nature arises

because the values of the interface equation matrix vary greatly in size, due to the heterogeneous

physical systems and the magnitude of the connection matrices (L). Large loss of significance errors

will be introduced and the propagation of rounding errors will be worse, due to the considerable

difference in values of the interface equation matrix [3]. Therefore, to avoid the problem, the interface

equation matrix is scaled so that the values are near one. Dongarra et al. [33], Forsythe and Moler

[53], Atkinson [3], and Golub and Van Loan [60] among others, have addressed the scaling issue of

matrices. In this subsection, a method for scaling the interface equation is shown that extends the

work of Yue and Park [127].

Yue mentions two approaches for scaling the ill-conditioned matrices [127]. The first approach

is to scale the interface force (Lagrange multipliers) along the partitioned boundaries [62, 100]. This

37

begins by modifying the constraint functional with normalization matrices DS , DF , such has:

Πconstraint = λTS DS(BTSuS − LSuB) + λTF DF (BT

FDψ − LFuB). (3.8)

This ultimately leads to a normalization of the interface equation as shown by Gumaste et al. [62] and

can be illustrated by inserting the above constraint function into Eq. (2.52) and following the steps in

Section 2.5.

The second approach is the displacement scaling method. The basic idea is to scale the mag-

nitude of the diagonal terms of matrices (BTSFSBS and BT

FDFFDTBF ) to the order of one. This

causes a special need for modifying the third equation (Newton’s third law) of the interface equation,

Eq (3.7), and here is the extension to Yue’s work. In modifying the third equation, the objective is

to either average the majority of the terms to near one, or to take the maximum term and scale it to

one and scale all others by the scaling factor. The interface equation is shown below again for clarity.

BTSFSBS 0 LS

0 BTFDFFDTBF LF

LTS LTF 0

λn+1S

λn+1F

un+1B

=

BTSFSg

n+1S

BTFDFFgn+1

F

0

.

First, the scaling of the matrices (BTSFSBS and BT

FDFFDTBF ) in the interface equation matrix

is done in the following manner. Define the following for clarity:

FS = BTSFSBS ,

FF = BTFDFFDTBF ,

GS = BTSFSg

n+1S ,

GF = BTFDFFgn+1

F .

(3.9)

Scaling matrices for the terms (BTSFSBS and BT

FDFFDTBF ) can be created as

SS =[

(diag(FS))12

]

−1,

SF =[

(diag(FF ))12

]

−1,

(3.10)

where diag() refers to using only the diagonal components of the matrices. Therefore, the matrix

multiplication STS FSSS will result in a matrix with ones on the diagonal, and similarly for STF FFSF .

Hence, the first and second equation of the interface equation matrix would have the form:

STS FSSS(S−1S λ

n+1S ) + STSLS(un+1

B ) = STSGS ,

STF FFSF (S−1F λ

n+1F ) + STFLF (un+1

B ) = STF GF ,(3.11)

38

where the equations are pre-multiplied by the transpose of their respective scaling matrices and the

inverse of the scaling matrices are used so that unity on the diagonals is obtained. Note that taking

the inverse of the scaling matrices is rather simple, since the matrices only consist of diagonal terms.

In the above equation, Eq. (3.11), the term STSLS and STFLF have now become orders of magni-

tude away from unity. For example, if the structure is made from standard concrete, then the values

in the matrix STSLS are of the order of 105. Thus, it is desirable to scale these values to one as

well. However, this can become complicated because of the non-matching meshes and the use of the

zero-moment rule. The use of the zero-moment rule can create a boundary discretization that has

a different number of degree of freedoms than either the structure interface boundary or the fluid

interface boundary, which precludes the use of the method by Yue [127]. Once again, the objective

is to scale the numbers of these matrices to the order of one. This can be accomplished by using the

following scaling parameter:

s =1

2

(

ni∑

1

nj∑

1

|(STSLS)ij |nsz

+

ni∑

1

nj∑

1

|(STFLF )ij|nfz

)

, (3.12)

where ni and nj are the rows and columns of the respective matrices, and nsz and nfz are the number

of nonzero values of the respective matrices. A square scaling matrix for the concerned matrices STSLS

and STFLF can be created with this parameter as:

(SB)ii =1

s. (3.13)

Thus, it only has values on the diagonal and its size is the size of the column of LF or LS, which is the

number of degree of freedoms accounted for on the interface boundary with the normal component

taken into account.

Therefore, the three equations of the interface equation are scaled to the order of one, and are

written as follows:

STS FSSS(S−1S λ

n+1S ) + STSLSSB(S−1

B un+1B ) = STSGS ,

STF FFSF (S−1F λ

n+1F ) + STFLFSB(S−1

B un+1B ) = STF GF ,

STBLTSSS(S−1S λ

n+1S ) + STBLTFSS(S−1

S λn+1F ) = 0.

(3.14)

In matrix form:

STS FSSS 0 STSLSSB

0 STF FFSF STFLFSB

STBLTSSS STBLTFSS 0

S−1S λ

n+1S

S−1S λ

n+1F

S−1B un+1

B

=

STSGS

STF GF

0

. (3.15)

The original interface equation, Eq. (3.7), will be used and referenced in the remainder of this

study to avoid clutter. However, it should be noted that this matrix has been scaled with success for

39

the problems in this study in the above manner. Some simple comparisons showing the advantage of

this scaling procedure are discussed in Chapter 6.

3.3 Inclusion of Concept for other FSI Methods (i.e. CASE)

It is sufficient to say that many organizations already have an acoustic fluid code and a structure

code that they are content with using, but want to be able to handle non-matching meshes. This can

be accomplished by using the salient features of the LLM method to develop a relationship between

the structure’s boundary displacements and the fluid’s boundary displacements as well as the systems

boundary forces. In most typical FSI codes, the structure displacement is passed to the fluid and the

fluid passes the pressure to the structure; thus, two relations are provided. The first is the relation

between the fluid’s pressure and the structure’s force at the interface boundary. By relating these to

the Lagrange multipliers and using the relation between the Lagrange multipliers the interface patch

test is satisfied. The second relation is between the subsystem’s interface displacements and can be

found from the relation of the subsystem’s forces.

3.3.1 Relation Between Fluid Pressure and Structure Force

In order to assure that the interface patch test is satisfied a relationship between the fluid

pressure at the boundary and the Lagrange multiplier needs to be determined. Given the fact that the

interface constraint functional is in terms of energy, the normal value of the fluid’s Lagrange multiplier

can be thought of as being equal to the opposite of the fluid’s pressure determined at the boundary

pFB, which is in accordance with Newton’s third law.

λnF = −pFB, (3.16)

where the superscript n refers to the normal value in this section, and should not be confused with

any time-stepping value. In addition, the structure’s normal value of the Lagrange multiplier can

be thought of as the normal value of the structure force being applied at the structures interface

boundary. Therefore, it is only necessary to determine the normal value of the structure’s Lagrange

multiplier in terms of the fluid normal value of the Lagrange multiplier. Once the normal value of

the structure’s Lagrange multiplier is known then the force acting on the structures wetface can be

determined by calculating the average normal at the structure’s node and multiplying this vector to

the relevant normal structural Lagrange multiplier value.

Therefore, the interface constraint energy functional is used to develop a relation between the

partitioned FSI subsystems’ Lagrange multipliers. The interface constraint energy functional is rewrit-

40

ten below for convenience.

ΠB =

ΓF B

λF ini(uF ini − uBini)dΓFB +

ΓSB

λSini(uSini − uBini)dΓSB .

As noted for an acoustic fluid structure interaction, it is desirable to have the Lagrange multipliers in

terms of the normal components. Hence, only the normal values are needed, and the boundaries can

use the normal values as their primary variables to be discretized.

ΠB =

ΓF B

λnF i(unF i − unBi)dΓFB +

ΓSB

λnSi(unSi − unBi)dΓSB. (3.17)

The primary variables (Lagrange multipliers and displacements) can be discretized as before except

with the realization that the normal vector is not needed and there is only one-degree of freedom at

each node due to the normal value being the variable component.

ΠB = (λnF )T

([

ΓeF B

NTλFNFdΓFB

]

unF −[

ΓeB

NTλFNBdΓB

]

unB

)

+

(λnS)T

([

ΓeSB

NTλSNSdΓSB

]

unS −[

ΓeB

NTλSNBdΓB

]

unB

)

,

(3.18)

where it assumed that the integrations on the elements are assembled into global matrices, which

produces:

ΠB = (λnF )T(

(BnF )TunF − LnFunB

)

+ (λnS)T(

(BnS)TunS − LnSu

nB

)

. (3.19)

The variation of the interface constraint energy functional leads to the following set of equations at

the interface,

δΠB = δ(λnF )T(

(BnF )TunF − LnFunB

)

+

δ(λnS)T(

(BnS)TunS − LnSu

nB

)

+

δ(unB)T(

−(LnF )TλnF − (LnS)TλnS)

+

δ(unF )T (BnFλ

nF )+

δ(unS)T (BnSλ

nS) .

(3.20)

The third equation in Eq. (3.20) can be used to find the relation between the two normal

Lagrange multipliers.

λnS = −[

LnS(LnS)T]−1

LnS(LnF )TλnF (3.21)

This then provides the relation between the fluids pressure and the acting normal force at the interface

boundary of the structure, fSB;

fnSB =[

LnS(LnS)T]−1

LnS(LnF )TpFB. (3.22)

41

By using this relationship, a constant stress can be passed from the fluid subsystem to the structure

subsystem. However, there is the possibility that the matrix[

LnS(LnS)T]

is singular. This can happen

when the structure mesh is refined within each fluid element. There are two methods to get around

this issue. The first is to obviously remesh. The second is to use a relationship between the fluid

displacements and the structure displacements from Eq. 3.20, then with the use of the virtual work at

the interface, a relation between the Lagrange multipliers can be determined as discussed in the next

section. However, using the displacements values relation does not assure that the interface patch test

is passed.

3.3.2 Relation between Subsystems’ Boundary Displacements

In many acoustic FSI methods, there is a need to know the discretized fluid’s boundaries nor-

mal displacement in terms of the discretized structures normal displacement. For instance, in the

CASE method [114] (see Appendix A) the normal displacement at the boundary effects the boundary-

interaction vector.

be =

Γe

φ∇ψ · nidΓ =⇒∫

Γe

φuF i · nidΓ, (3.23)

where φ is a column vector composed of 1-D, N th-order-polynomial basis functions.

Therefore, the first and second equations of Eq. (3.20) can be used to develop a relationship

between the fluid’s normal boundary displacements and the structure’s normal boundary displacements

such as

unFB = LnF[

(LnS)TLnS]−1

(LnS)TunSB, (3.24)

where unFB and unSB refer to the normal displacements of the fluid and the structure respectively at

the boundary.

Therefore, given the structure’s normal displacements at the nodes of the structure boundary,

the fluid’s normal displacement at the nodes of the fluid boundary are determined. The structure’s

normal component can be determined by the known structure’s displacements of an element and

creating the normal components from these values. The normal value at a node is the average of the

normal values of the elements surrounding the node.

When maintaining the normal component in the integration of the connection matrices L, i.e.

for the structure,

LS =

ΓeB

NTλSn

eneTNB dΓeB (3.25)

a problem can arise in the inversion of the matrix[

(LS)TLS]

−1, because it has the potential to become

singular. Therefore, it is imperative to use the normal values of the displacements for the derivation.

42

Another option to determine the relation between the subsystems boundary forces is to use

the relation between the susbsystems’ boundary displacements and the principle of virtual work.

This results in using the transpose of the matrix created by the relation between the normal fluid

displacement and the normal structure displacement [38] for the relation among the forces. This can

be seen by examining the virtual work on the boundaries.

δWFB = (λnF )TunFB, (3.26)

where W is the work and FB refers to the fluid boundary. Inserting

unFB = LnF[

(LnS)TLnS]−1

(LnS)TunSB = RnFSu

nSB ,

where RnFS is the relation from the structure normal displacements to the fluid normal displacements,

into Eq. (3.26) provides

δWFB = (λnF )TRnFSu

nSB. (3.27)

The virtual work along the structure boundary is given by

δW SB = (λnS)TunSB. (3.28)

Therefore, by equating Eq. (3.27) and Eq. (3.28) it can be seen that

λnS = (RnFS)TλnF . (3.29)

This relationship can be used for the forces, but it does not guarantee that the patch test will be

satisfied. Also, there is the issue where the matrix[

(LnS)TLnS]

can become singular. It can become

singular in the opposite case of the previous one used for the Lagrange multiplier’s relation for it would

be come singular if the fluid mesh is refined within a structure element.

In conclusion, the Lagrange multiplier relation should be used for the force relation in order to

pass the interface patch test. Unfortunately, this can become an issue if the structure is more refined

then the fluid. However, there is a method to get around the issue, but can result in non-conservative

forces at the interface.

3.4 Error/ Accuracy

This section begins with a discussion on the order of the discretization error in strain associated

with LLM method previously derived. Here, we concentrate on the error associated with spatial

discretization at the interface boundary, because the method is capable of being implemented with

43

any time integration, which would dictate the error due to the time discretization. Then posteriori

error estimates are mentioned that are used in the remaining chapters.

Fortunately, the Mortar method, discussed in Appendix B.1, has resulted in significant studies

on the error associated with the method [8, 112, 125, 126]. The Mortar method achieves coupling

between two non-overlapping domains, Ωi, Ωj, by using a Lagrange multiplier to weakly enforce the

constraint ui − uj = 0 along the interface boundary [4].

Γλ(ui − uj)dΓ = 0. (3.30)

This method provides a discrete solution that satisfies optimal error estimates with respect to nat-

ural norms [8]. For instance, if the subdomains (Ωs) are discretized with linear finite elements on a

mesh with a characteristic size hs, then the global discretization error of the coupled problem grows

asymptotically as O(hs). This error was derived with the following assumptions.

• The domain Ω is partitioned into non-overlapping polygonal subdomains Ωs.

• The interface Γij of two subdomains intersection Ωi ∩ Ωj can be decomposed into a set of

disjoint straight line pieces γi.

• The triangulations over the different domains Ωs are assumed independent of each other, with

no compatibility enforced across interfaces.

• The mortar discretization is the trace of the triangulation from one of the domains Ωs, and it

uses the same piecewise polynomial interpolation over the element.

Given item four of the assumptions, where the discretization of the Lagrange space is equal to one of the

domains, it is determined that this is the same for the LLM method across the interface boundary to

each subdomain. Therefore, the global discretization error for the LLM method grows asymptotically

as O(hS) +O(hF ) +O(hB) with linear shape functions used for the Lagrange multipliers as described

in Section 2.3.4. In other words, the interface error induced by the LLM method on the solution is

not worse than the local fluid, structure, and boundary discretization errors.

However, the above error is associated with linear shape functions for the Lagrange multipliers

given the fluid and structure formulations in this study. The error for the formulation developed with

Dirac delta functions for the shape functions are assumed to be different. Farhat et al. [38] note

that using an interpolation method, where one subdomain’s interface Γi values are interpolated to the

interface of another Γj, gives a discretizatioin error of the coupled domains that grows asymptotically

as O(√hi). In essence, the LLM method with Dirac delta shape functions for the Lagrange multipliers,

44

ultimately interpolates the values of the subdomains to the interface frame, see Appendix B.2. Thus,

the discretization error of the LLM method with Dirac delta shape functions grows asymptotically as

O(√hF ) +O(

√hS) +O(

√hB) in strain.

The error is not mathematically optimal, but it will be shown to be sufficient for problems in

this study. In Appendix C, the discretization error for the two dimensional problems is shown to

be of order O(h2f + h2

s + h2b) in displacements, where the displacement variables at the interface are

interpolated with linear shape functions, and the intersection of the subdomains is composed of linear

disjoint segments. The remainder of the Chapter provides error measures used to evaluate the LLM

method in the problems of this study.

3.4.1 Geer’s C-Error Method

Geers’ comprehensive error factor [114] is used to quantify the error of a transient response

history relative to an accepted benchmark solution. The comprehensive error factor is given by C =√M2 + P 2, in which

M =√

ϑcc/ϑbb − 1, P =1

πarccos(ϑbc/

ϑbbϑcc),

ϑbb =1

t2 − t1

∫ t2

t1

b2(t)dt, ϑcc =1

t2 − t1

∫ t2

t1

c2(t)dt,

ϑbc =1

t2 − t1

∫ t2

t1

b(t)c(t)dt,

(3.31)

where c(t) is a candidate solution in the form of a response history, b(t) is the corresponding benchmark

history, and t1 ≤ t ≤ t2 is the time span of interest. The integrations are done numerically with the

composite Simpson’s rule [75], for this study. M is the magnitude error factor, which is insensitive to

phase discrepancies, and P is the phase error factor, which is insensitive to magnitude discrepancies.

3.4.2 Energy Error

Before moving to the numerical experiments, another type of error measurement that will be

used throughout the experiments should be discussed. This error is the energy or momentum difference

at the interface. If the interface is viewed as being a closed system, then the energy gained or released

on one side of the interface frame is equal to the energy lost or absorbed by the other side of the

interface. Thus, the difference in the energy on the two sides of the interface is used for an error

measurement. Two types of error can be examined for the LLM method. The first is the difference at

the interface frame.

W (n)interror = uTB(−LTSλS − LTFλF ), (3.32)

45

where n refers to the time step, and int is the for the interface frame. This will provide a graphical

representation of the error at the interface throughout the time span of interest. This difference

will be used to notice any error that is really introduced by the scaling of the interface equation,

see Section 3.2.1, for the Lagrange multipliers are specifically solved only at the interface equation.

According to the third equation of the Interface equation, Eq. 3.7, mathematically the system is

conservative at the boundary as the forces are equal and opposite in accordance with Newton’s third

law. However, when using the LLM transient method there is a minor error noticed with the above

equation that is not present when using the CASE method with the LLM concept for the non-matching

meshes. This indicates that the scaling is introducing a little error into the system. This is to be further

discussed in Chapter 6.

The next error is associated with the energy at the boundary of the different domains. In this

error, the energy at one subdomain’s boundary is compared with the energy at the other subdomain.

W (n)suberror = λTS (BTSuS) + λTF (BT

FDψ), (3.33)

where sub refers to the error associated with difference between energy on the different subdomains.

Once again, this is provided at each time step and will be used for visual purposes. This error is

associated more with the LLM method as it looks at the energy difference between the subdomains’

boundaries. Also, note that any error created by the gradient matrix, D, will be introduced into this

measurement. In addition, the plus sign is used because the Lagrange multipliers are theoretically

equal and opposite at the interface frame.

An energy difference can also be used with other FSI codes that incorporate the LLM method,

see Section 3.3. This is simply the same error as mentioned above for the energy difference at the

subdomains’ boundaries.

W (n)CASEerror = fTSB(uSB) − fTFB(uFB), (3.34)

where fSB is the force at the structures wet face produced from the fluid pressure converted to the

structure boundary DOF by some method, fFB is the force at the fluid dry face from the fluid pressures,

uSB is the structure displacements at the wet face, and uFB is the fluid displacements at the dry face.

In a staggered integration procedure the correct time step value of the uSB needs to be used, which is

the predicted value. This error measure is used to graphically compare different non-matching mesh

techniques.

46

3.5 Summary

In this chapter, the Localized Lagrange Multiplier (LLM) method is extended for use in an

acoustic fluid-structure interaction (FSI). The main feature was the transient analysis in the time

domain that was developed. Conveniently, this transient analysis allows for a partitioned method that

has the potential for an always stable method, if implicit transient analysis methods are chosen for

the time integration.

However, if a cavitation analysis is required, then it is more computationally efficient to use

the explicit time integration method known as the Central Difference method, to be discussed in

Chapter 8. Ultimately, this cavitation analysis leads to significant computational inefficiency when

compared to the Cavitating Acoustic Spectral Element (CASE) method [114], to be discussed in

Chapter 8. Therefore, the ability to couple non-matching meshes with the LLM method is extended

to the CASE method. However, this has limitations in that it may not always pass the interface patch

test.

Finally, it was noted that the LLM method has the potential for optimality in regards to the

discretization error. However, a simpler integration procedure was provided using Dirac delta shape

functions for the Lagrange multipliers that affected the discretization error. Nonetheless, this simpler

integration procedure is used throughout this study to help evaluate the method. The rest of the

chapters in this study are used for validation and comparison of the LLM method.

Chapter 4

Validation of Fluid Code: Pressure on Dam Face

4.1 Introduction

This chapter is intended to show that the fluid formulation derived for the LLM method with

the gradient matrix can provide accurate results. In the derivation of the fluid formula, Section 2.4, it

is shown that the discretization error of the fluid displacement is greatest at the edges of the domain

O(h), and the order of the discretization error is O(h2) for interior nodes. Therefore, the pressure on

the edge of the fluid is compared between an analytical model and the derived computational model.

In addition, this chapter provides an analysis with the use of silent boundaries. There is also a brief

section on the inclusion of surface waves.

4.2 Problem Description and Analytical Solution

The problem studied is a rigid dam excited by an earthquake, uq, as depicted in Fig. 4.1. This

problem was studied by Chopra [23], where an analytical solution was determined for the pressure on

the dam face. Chopra made the following assumptions during this study. The effect of the length of

the fluid is negligible for L/H > 3, where L is the length of the fluid and H is the depth of the fluid.

In addition, the effect of surface waves could be ignored with little loss of accuracy as long as the

depth is greater then 100 ft, for a typical ground motion period below three seconds.

Given the above assumptions, the following analytical form of the pressure was determined by

Chopra [23]:

p(0, y, t) =4ρc

π

∞∑

n=1

(−1)n−1

2n− 1cosλny

∫ t

0uq(τ)Joλnc(t− τ)dτ, (4.1)

where ρ is the fluid density, c is the velocity of sound in the fluid, λn = (2n − 1)π/(2H), y is the

vertical height on the dam face, H is the depth of the fluid, Jo is the Bessel function of the first kind

of order zero, t is time, and uq(τ) is the given earthquake acceleration. Parameters for these values are

shown in Table 4.1. The El Centro earthquake of May 18, 1940, was chosen as the excitation for this

48

Figure 4.1: Rigid dam problem for fluid code verification.

49

problem, and only the x-direction of the acceleration was used. This provided the analytical solution

shown in Fig. 4.2 for the first eight seconds of the earthquake.

ρ 1000 kg/m3

c 1500 m/sH 80.0 my 40.0 m

time duration 8.0 s

Table 4.1: Parameters for the pressure on the dam face problem.

4.3 Fluid Code Validation

Given the analytical solution, a comparison with computational models can be conducted. In

these models, the pressure is computed in elemental gauss points surrounding the point of interest,

then interpolated to the point. The point of interest is on the dam face located forty meters from

the bottom of the fluid. The pressure in the element is determined with the use of the displacement

volume strain matrix and the fluid displacements. In the current finite element discretization scheme

the pressure inside an element, and at a certain location in the element, can be determined by the

following relation that is typically used for the computation of stress in structures.

p = −ρF c2BvsueF , (4.2)

where ueF is found from the global displacements (uF = Dψ), and Bvs is the displacement volume

strain matrix at a particular location in an element [36].

Bvs = ∇ ·NF (4.3)

This can be saved during the initial calculation of the stiffness matrix of the fluid. However, reduced

integration is used and thus, the pressure at the center of the element will be determined. As noted

by Dow [34], a smoothed solution of the pressure would more closely approximate the actual pressure

solution. Therefore, the pressure is found at gauss points for full integration. Given these values, the

element pressure at the desired location is interpolated from the gauss points. If the desired location

is on an edge of element, then the average from the neighboring elements is used. This method is

used, because it requires the gradient matrix (uf = Dψ) which is one of the new components of the

fluid formulation and needs to be verified.

50

0 1 2 3 4 5 6 7 8−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

2.5x 10

5

Time (seconds)

Pressure

Figure 4.2: Analytical pressure on the rigid dam.

Figure 4.3: Fluid mesh with a characteristic length of 20 m.

51

The verification begins by examining three fluid meshes. The first fluid mesh is depicted in

Fig. 4.3, where the element length is twenty meters. The second mesh reduces the characteristic

length by half to ten meters, and the third mesh again reduces the length by half to five meters.

In order to evaluate the fluid formulation for the LLM method, three figures are provided for

visual comparison, please see Fig. 4.4, Fig. 4.5, and Fig. 4.6. In addition, on each figure is the C-error,

M-error, and P-error. These error measures are explained in Section 3.4. As can be seen, the error

is reduced for each refined mesh. The magnitude error is greater then the phase error, which could

be associated with the silent boundary and extending the period of analysis past three seconds as

recommended in Chopra’s study.

0 1 2 3 4 5 6 7 8−3

−2

−1

0

1

2

3

4x 10

5

Time (seconds)

Pre

ssur

e (P

a)

C−Error = 0.46; M−Error = −0.39; P−Error = 0.24

LLMAnalytical

Figure 4.4: Pressure on dam face with a fluid characteristic length of 20 m.

4.4 Incorporation of Silent Boundaries

An issue that needs to be considered for problems in this work is that of silent boundaries (non-

reflecting boundaries). Silent boundaries are needed to truncate the domain volume to a finite extent.

Thus, this section explains the incorporation of silent boundaries with the fluid formulation and the

structure formulation as depicted in Eq. (2.54). It is desirable for the boundary to perfectly radiate

outgoing waves (no spurious reflections). In addition, it is desired that the boundary be as close to

52

0 1 2 3 4 5 6 7 8−2.5

−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

2.5x 10

5

Time (seconds)

Pre

ssur

e (P

a)

C−Error = 0.31; M−Error = −0.26; P−Error = 0.18

LLMAnalytical

Figure 4.5: Pressure on dam face with a fluid characteristic length of 10 m.

0 1 2 3 4 5 6 7 8−2.5

−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

2.5x 10

5 C−Error = 0.26; M−Error = −0.20; P−Error = 0.17

Time (seconds)

Pre

ssur

e (P

a)

LLMAnalytical

Figure 4.6: Pressure on dam face with a fluid characteristic length of 5 m.

53

the finite structure as possible for computational efficiency.

The Perfectly Matched Layer (PML) method [9] would meet these requirements. This method

was developed in the frequency domain for electromagnetic wave absorption. It has been found that the

PML absorbing boundary condition provides much higher absorption than other previous absorbing

boundary conditions in finite-difference methods [82]. Hence, much work as been done to extend the

method to the finite difference time domain. Several authors, [22, 128, 82], have extended the method

for elastic waves with the finite difference method. Geers and Qi, [57] demonstrated the ability of the

PML to be used in acoustics with the finite difference method in the frequency and the time domain.

They also provided criteria for using the PML in acoustics. Extensions into the finite element method

have been performed for elastodynamics by several as well, [6, 7, 79, 81] in the frequency domain.

Basu et al., [7], were able to develop the PML method for finite elements in the time domain.

However, it has been found that the implementation burden of the PML would not be justified

in the problems of this study. Brahtz and Heilbron [17], and Calayir and Karaton [19], mention that

the effect of the radiating condition on the solution is generally negligible if the reservoir length is

taken three or more times then the depth of the reservoir. Since it is desired to capture the majority

of the cavitation region in the fluid computational domain, meeting the requirement for the length of

the reservoir is desirable. In regards to the structure, 5% Rayleigh damping is present in the domain

and this helps with the attenuation of the elastic wave.

However, a silent boundary is still provided for problems in this study. The plane wave approxi-

mation is provided for the fluid. The structure’s silent boundary uses the Viscous Damping Boundary

method (VDB) of Lysmer and Kuhlemeyer [83]. This was tested by comparing extended models that

would not effect the response in the time period of the two dimensional dam problems without fluid

to a truncated model with the VDB. Here it was found that a length of two times the height of the

dam provided negligible effects on the solution of displacements on the dam face. It has been noted

these absorbing boundary conditions are only approximated, and small reflections from the boundaries

should be expected. Komatitsch et al., [77] mentions that the amplitude of these reflections is of the

order of a few percent of the amplitude of the direct wave. Obviously, these conditions are optimal for

waves impinging normal to boundary and rapidly become less accurate when the angle of incidence

differs from the normal [26].

The simplest method to be incorporated into the above derived fluid formulations would be

the use of a plane wave approximation (PWA) for the silent boundary. It is valid for short acoustic

wavelengths (high frequencies) and is useful for modeling early stages of the transient interaction [58].

The PWA is chosen over the Doubly Asymptotic Approximation (DAA), [56, 59], because the waves

54

are assumed to behave as a plane wave and the extent of the fluid length previously mentioned. The

PWA relates the pressure to the velocity by the following:

p = ρF cun, (4.4)

where c is once again the fluid speed of sound, p is the pressure, and un is the normal component of

the velocity at the boundary [42]. In the energy functional of the fluid, there is a term that is related

to the surface traction. In this term, the PWA can be introduced through the relation between the

traction and the pressure

TF i = −pni. (4.5)

This ultimately can lead to a damping term that is added to the fluid equations as the following:

CeF =

ΓFnrb

cρFNTneneTNdΓFnrb, (4.6)

where ΓFnrbis the silent boundary, n is the normal vector at that boundary, and N is a standard

linear shape function. In order to use the displacement potential as the fluid state variable, the fluid

damping matrix is modified by the gradient matrix (D), as was done to the fluid stiffness and mass

matrix, see section 2.4.

Cfd = DTCFD (4.7)

This then fits into the variation of the systems energy functional, Eq. (2.54), as

δΠ(uS ,ψ,λS ,λF ,uB) =

δuTS (MSuS + CSuS + KSuS + BSλS − fS)+

δψT (Mfdψ + Cfdψ + Kfdψ + DTBFλF − DT fF )+

δλTS (BTSuS − LSuB) + δλTF (BT

FDψ − LFuB)+

δuTB(−LTSλS − LTFλF ).

(4.8)

A partitioned transient analysis can then be performed as discussed in section 3.2.

A silent boundary for the structure is included by the simple Viscous Damping Boundary method

(VDB) of Lysmer and Kuhlemeyer [83]. The VDB is similar to the plane wave approximation method

in that the VDB uses the concept of applying viscous dampers to the DOF on the boundary element.

However, for an elastic media, such as soil, there are primary waves and secondary waves that travel

through the media. Thus, the speed of these waves is used instead of the speed of sound in the fluid,

which was used for the PWA. Primary waves are compression waves that travel through solids, liquids,

and gases. Secondary waves are shear waves that only travel through solids.

55

In order to develop a boundary condition that ensures that all energy arriving at the boundary

is absorbed, Lysmer and Kuhlemeyer [83] found that the most promising expression for the boundary

condition is expressed by the following equations.

σzz = aρVpuz Primary wave

τzx = bρVsux Secondary wave

τyz = bρVsuy Secondary wave

(4.9)

These equations are formulated for an incident wave consisting of a primary and secondary wave that

act at an angle θ from the z-axis as shown in figure (4.7) [55]. In the above equations, ρ is the density

of the media, ux, uy, uz represent the velocities in the x,y,z-direction, Vp and Vs are the velocities of the

primary and secondary waves, and a and b are dimensionless parameters. For small incident angles

(θ < 30), the most effective value for the dimensionless parameters (a and b) is one [83].

Figure 4.7: Viscous Damping Boundary concept [55].

The primary and secondary velocities can be computed as

Vp =

(1 − ν)E

(1 − 2ν)(1 + ν)ρ,

Vs =

E

2(1 + ν)ρ,

(4.10)

where E is the modulus of elasticity, and ν is the Poisson’s ratio.

Implementation of the method is fairly simple. It amounts to modifying the damping matrix.

This is accomplished by adding the following additional terms to the diagonal damping terms.

Cii =

ΓaρVpdΓ Primary wave, z-direction,

Cii =

ΓbρVsdΓ Secondary wave, x or y direction,

(4.11)

where Γ is the surface area of the element at the nonreflecting boundary.

56

4.5 Silent Boundary Validation

In the above pressure analysis, the fluid code was tested with a silent boundary. In Section 4.4

it was noted that the effect of the radiation condition had minimal effect on the response of the dam,

if the length of the fluid was three times the height of the fluid. In this section, these statements are

tested and evaluated. However, it should be noted that the authors that made the assertion about

the minimal effect of the radiation condition were evaluating the displacement of the dam. Here, the

pressure is examined. Figure 4.8 depicts the pressure at the face of the dam at forty meters for the

case with no silent boundary (shown in blue), for the mesh with a characteristic length of ten meters.

Here it is noted that there is a difference between the analysis with no silent boundary and one with

a silent boundary (shown in a dashed green line). The C-Error is in fact nearly, the value for the fluid

mesh with a characteristic length of twenty meters; hence, an improvement with the use of the silent

boundary.

0 1 2 3 4 5 6 7 8−4

−3

−2

−1

0

1

2

3x 10

5

Time (seconds)

Pre

ssur

e (P

a)

No Silent BoundaryAnalyticalWith Silent Boundary

C−Error = 0.43; M−Error = −0.35; P−Error = 0.25

Figure 4.8: Pressure on dam face with a fluid characteristic length of 10 m with and without a silentboundary.

Finally, the effect of including numerical damping is evaluated. Numerical damping in the fluid

domain is added to suppress the spurious oscillations (frothing) that can accompany the cavitation

analysis, which will be discussed in Section 8.2. It is recommended to use a stiffness proportional

57

Rayleigh damping ratio of 0.1%. In Fig. 4.9 it is noted that adding damping actually improves the

response of the system. The C-error was improved from 0.26 to 0.15, for a stiffness proportional

Rayleigh damping of 0.1%.

0 1 2 3 4 5 6 7 8−2

−1.5

−1

−0.5

0

0.5

1

1.5

2

2.5x 10

5 C−Error = 0.15; M−Error = −0.07; P−Error = 0.13

Time (seconds)

Pre

ssur

e (P

a)

LLMAnalytical

Figure 4.9: Pressure on dam face with a fluid characteristic length of 5 m with damping added to thefluid.

This is believed to be due to the fact that the plane wave approximation (PWA) for the silent

boundary is not exact; thus, adding a very minor amount of damping to the fluid helps to attenuate

waves reflecting from the PWA. As can be noticed it appears that adding this damping tends to

improve the results in the last few second of the analysis, see Fig. 4.9. Obviously, adding too much

damping can drastically change the results. Though, once again it has been found that adding a small

portion of damping to suppress frothing has very little effect on the response of the dam [36, 96].

In Fig. 4.10, damping is increased until the C-error is equivalent to the C-error associated with no

damping for the fluid mesh with a characteristic length of 5 meters. It was found that adding damping

improved the results until the damping reached 1.36%.

4.6 Inclusion of Surface Waves

Another issue that needs to be addressed is when an example problem contains a free surface,

as does the problem in this chapter. Here it is beneficial to include a term for the surface waves. By

58

0 0.5 1 1.5 20

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Proportional Damping (%)

C-E

rror

Figure 4.10: Pressure on dam face with a fluid characteristic length of 5 m with damping added tothe fluid.

59

enforcing the irrotational condition of the fluid, several of the zero-energy modes of the fluid system

have been removed. Unfortunately, as pointed out by Wilson and Khalvati [124], there are still zero-

energy modes associated with the surface nodes. These can be converted to low frequency sloshing

modes by incorporating surface waves in the derivations. Typically, surface waves are included by

modifying the pressure at the surface, as the following:

psurface(x, t) = patm − ρgun(x, t), (4.12)

where patm is the atmospheric pressure, g is the gravitational constant, and un is the normal displace-

ment at the surface.

patm

un(t)still water

Figure 4.11: Free surface wave

In the current fluid derivation, the atmospheric pressure can be included by the surface traction

term of the fluid equations. Thus, the remaining term to be included is that of the surface waves.

This is added by including the following in the appropriate location of the Stiffness matrix.

Ks =

Γsurface

ρF gNTFnnTNF dΓsurface (4.13)

It should be noted that in some systems this effect does not need to be included.

4.7 Summary

This chapter used an analytical model to verify the fluid code. The pressure on the dam face is

used for the comparison between the analytical model and the computational model. This parameter

was chosen over the displacement because in the computational models it would have a greater error

value. As can be expected, as the characteristic length of the element decreases, the C-error also

decreased. Note that the error is more significant in the last four seconds than in the first four

seconds. Two possible explanations are that the silent boundary can not truly attenuate the pressure

waves, or that after a period of three seconds the surface waves need to be accounted for in Chopra’s

analysis [23]. It was also shown that the addition of a silent boundary did improve the results; though,

the majority of the improvement was in the last four seconds. It was also demonstrated that adding a

modest amount of damping to the fluid domain actually improved the results up to a certain damping

60

ratio. Thus, it is recommended that the stiffness proportional Rayleigh damping ratio begin around

0.1%, but not exceed 1.0%. Finally, it is realized that a characteristic length of ten meters for the

fluid model begins to provide accurate results.

Chapter 5

Validation of Concept: Infinite Piston Problem

5.1 Introduction

This chapter has a three-fold purpose. First, this chapter is to again validate the fluid formulation

and the transient method. Second, this chapter is used to illustrate one of the major benefits of the

LLM method for transient analysis. This is accomplished by examining a spring-mass infinite piston

problem with an analytical solution. Third, the chapter examines the benefit of the zero-moment rule.

A piston problem has been studied by several authors, [15, 110, 40, 117, 86], mainly for the use

in aeroelastic problems with attention towards the stability and accuracy of the transient method used

for the fluid-structure interaction. In this chapter, a piston problem is proposed that more accurately

reflects the problems in this study. Therefore, a silent boundary is added to the end of the fluid instead

of the typical shut piston problem, and the fluid parameters are those of water instead of air. This is

then used to evaluate the partitioned transient method described in Section 3.2.

The constant debate always arises between monolithic and partitioned solution procedures. The

biggest complaint about the partitioned solution procedure is the typical instability associated with

the method [86], which is generally caused by the time lag between the integration of the fluid and the

structure [15]. In the typical partitioned method, the fluid and the structure equations are integrated

in time, and the interface conditions are enforced asynchronously [47]. Much work has been done to

improve the stability of the partitioned method. Farhat et al. [39], developed a partitioned procedure

for aeroelastic problems where the stability limit was governed only by the critical time step of the

explicit fluid solver. Also, prediction techniques have improved the accuracy of the partitioned scheme

and lead to better stability [107, 106].

In this study, a partitioned method has been provided that can be used with an implicit time

integration method; thus, accuracy can impose the time step, rather than stability. This is examined

with the test problem described below. In addition, the partitioned method in this study does not

require the use of a predictor or a corrector. This is beneficial, because most predictor-corrector

62

schemes are custom made for a specific combination of time integration schemes and not applicable

to higher order schemes [117].

5.2 Problem Description

The problem to be analyzed in this chapter consists of a spring-mass system that is restrained to

only move in the x-direction and is in contact with an infinite body of water, see Fig. 5.1. The left end

of the spring mass system is excited by a prescribed displacement; for simplicity, the excitation that

is chosen is a sine function with a given amplitude (Xo) and a given frequency (ω). The parameters

for the problem are defined in table (5.1). Note that this simple model is a crude approximation of a

dam under seismic excitation.

us(t)

xprescribed = Xo sin(ωt)

k

m Fluid (ρ, c)

Figure 5.1: Infinite piston fluid-structure system.

ρ 1000 kg/m3

c 1500 m/sm 1.0 kgk 8000 N/mω 1 rad/sXo 0.01

Area 0.01 m2

Table 5.1: Parameters for the infinite piston fluid-structure system.

The objective is to analyze the system analytically, then computationally with already proven methods,

and finally to use the Localized Lagrange Multipliers (LLM) method previously derived in Ch.2.

5.3 Analytical Model for 1-D Problem

In order to analyze the fluid-structure system, the following assumptions are made. First, the

fluid is a linear acoustic fluid. The fluid is compressible, inviscid, and irrotational. Second, there

are no body forces acting on the fluid in this simple one-dimensional problem. Third, it is assumed

63

that the displacements of both bodies are so small that the acceleration is given by∂2ui∂t2

instead of

∂2ui∂t2

+ uj∂ui∂xj

.

With the above assumptions, the Eulerian equation for a fluid in one dimension becomes:

ρ∂2u

∂t2= −∂p

∂x. (5.1)

This equation can be rewritten with the fluid constitutive equation (p = −ρc2∇u), due to the small

compressibility of a liquid [46] to yield

ρ∂2u

∂t2= ρc2

∂2u

∂x2. (5.2)

The above equation is the classical wave equation for the fluid, which is used as the equation of motion

for the fluid in this one-dimensional problem. Therefore, a solution of the form u(x, t) is desired.

In order to solve this partial differential equation, some boundary conditions and initial con-

ditions need to be defined. The initial conditions for the system are simple. It is assumed that the

system is at rest; therefore,

u(x, 0) = u(x, 0) = 0. (5.3)

However, the boundary conditions are a little more complicated.

To determine the boundary condition on the left side of the system, the free-body-diagram

(FBD) of the mass, as shown in Fig. 5.2, is examined. Here the water applies a force that is equal to

the pressure multiplied by the area of the mass. Thus, the FBD provides the boundary equation at

mus

pAmk(xprescribed − us)

Figure 5.2: Free body diagram of a mass-spring system with water.

the far left end of the system:

mu(0, t) + pA+ k(u(0, t) − xp) = 0. (5.4)

Inserting the fluid constitutive equation and the prescribed displacement (xp = Xo sin(ωt)) into the

above equation yields

mu(0, t) − ρc2Au′(0, t) + ku(0, t) − kXo sin(ωt) = 0, (5.5)

where a prime (u′(x, t)) denotes the partial derivative with respect to x, a superposed dot (u(x, t))

denotes the partial derivative with respect to t, Xo is the amplitude of the excitation, and ω is the

frequency of the excitation.

64

In order to solve the wave equation, another boundary condition is required. To obtain this

boundary condition it is assumed that as the water goes to infinite, the displacement is zero.

limx→∞

u(x, t) = 0. (5.6)

This can be approximated with the plane wave approximation (PWA) [42] at a distance L.

p(L) = ρcu(L, t). (5.7)

Once again, the constitutive equation can be used for the pressure and after some simplification the

following boundary condition is found:

u(L, t) = −cu′(L, t). (5.8)

Therefore, the problem is to solve the partial differential equation,

∂2u

∂t2= c2

∂2u

∂x2, (5.9)

with the conditions

mu(0, t) − ρc2Au′(0, t) + ku(0, t) = kXo sin(ωt), (5.10)

u(L, t) = −cu′(L, t), (5.11)

u(x, 0) = 0, (5.12)

u(x, 0) = 0. (5.13)

To solve the equation, the Laplace transform is taken with respect to (t) of Eq. (5.9) [80].

L

∂2u(x, t)

∂t2

= s2L u(x, t) − su(x, 0) − u(x, 0) = c2L

∂2u(x, t)

∂x2

. (5.14)

Two terms drop out, due to the conditions (5.12) and (5.13). On the right, it is assumed that the

integration and differentiation can be interchanged as:

L

∂2u(x, t)

∂x2

=

0e−st

∂2u

∂x2dt =

∂2

∂x2

0e−stu(x, t)dt =

∂2

∂x2L u(x, t).

(5.15)

Writing U(x, s) = L u(x, t); the following is obtained

s2U(x, s) = c2∂2U(x, s)

∂x2,

or∂2U(x, s)

∂x2− s2

c2U(x, s) = 0. (5.16)

65

According to Graff [61], Eq. (5.16) has the solution

U(x, s) = (C1)cosh(s

cx)

+ (C2)sinh(s

cx)

. (5.17)

The transformed boundary conditions (5.10) and (5.11) become

ms2U(o, s) − ρc2A∂U(o, s)

∂x+ kU(o, s) =

kXoω

s2 + ω2, (5.18)

and

sU(L, s) = −c∂U(L, s)

∂x, (5.19)

where Eq. (5.12) and Eq. (5.13) along with the argument in Eq. (5.15) were used. To determine the

constants C1 and C2, the appropriate form of the solution, Eq. (5.17) is inserted into the transformed

boundary conditions. Then, the two unknowns (C1 and C2) can be determined. This yields the

following:

C1 = −C2 =kwXo

(k +Acρs+ms2)(s2 + ω2). (5.20)

The purpose of this exercise is to develop an analytical model that can be compared to a

computational model with the LLM method. Therefore, the only interest is in the displacement at

the mass (u(0, t) or U(0, s)). This has the solution:

U(0, s) = C1, (5.21)

since cosh(0) = 1, and sinh(0) = 0. Thus, the desired solution can be obtained by taking the inverse

Laplace transform of the constant C1. This produces the displacement of the mass as shown in Fig. 5.3,

for the parameters of the problem.

5.4 Current Computational Models

The displacement of the mass system was computed with two known computational models to

verify that the solution is correct. Both models discretize the fluid with finite elements. The first model

is the CAFE method (Appendix A), which uses the displacement potential for the formulation of the

equations [45]. The second model uses pressure for the formulation of the fluid equations and is derived

in Cook’s book, Concepts and Applications of Finite Element Analysis [29], where the functional for

the fluid portion of the problem becomes

Π =

ΩF

(

p2,x +p2,y +p2,z2

+1

c2pp

)

dΩF + ρ

Γs

unp dΓs +β

c

Γnrb

pp dΓnrb, (5.22)

where Γs is the surface in contact with the structure, un is the normal acceleration at the dry face,

and Γnrb is the surface in contact with the silent boundary.

66

0 5 10 15 20−6

−4

−2

0

2

4

6x 10

−3

Time (s)

Dis

pla

cem

ent

(m)

Figure 5.3: Displacement of the mass in the infinite piston fluid-structure system by an analyticalmethod.

67

Figure 5.4 illustrates that all methods produce the same results; although, this did require that

each method use a prediction step and a time step that met the CFL condition. The blue line in the

figure is the analytical method; the black, dashed colored line is produced by the fluid formulation with

displacement potentials (CAFE); and the red line is produced by the fluid formulation with pressure.

Therefore, this solution will be used to verify the LLM method. In addition, the LLM method will be

compared to the CAFE method.

0 2 4 6 8 10−6

−4

−2

0

2

4

6x 10

−3

AnalyticalCAFEPressure

Time (s)

Dis

pla

cem

ent

(m)

Figure 5.4: Displacement of the mass in the infinite piston fluid-structure system by existing compu-tational models.

5.5 LLM Method Results in 3-D

5.5.1 Matching Meshes Results

The model used for the LLM method is a three dimensional spring, mass, and fluid column as

depicted in Fig. 5.5. This produces the results as shown in Fig. 5.6. As can be seen, the LLM method

produces almost identical results; in fact, the C-error associated with this figure is 0.0007, for a time

step of 0.01 seconds.

5.5.2 Non-matching Results

Given that the matching interface produced such accurate results, it was decided to attempt

non-matching meshes on this problem as well. Therefore, the mass was divided into four elements and

the fluid face was divided into nine elements as depicted in Fig. 5.7. First, the analysis was carried

out with the interface frame discretized as the structure. This also produced no visible errors in the

68

0.1 m

0.1 m

0.1 m

Fluid Silent Boundaryk4 Mass

Figure 5.5: Model with 3-D elements for the LLM method.

0 2 4 6 8 10−6

−4

−2

0

2

4

6x 10

−3

Time (s)

Dis

pla

cem

ent

(m)

LLM MethodAnalytical

Figure 5.6: Comparison between analytical and LLM method with no visible error; C-error = 0.0007.

69

graph, and the C-error was minimal at 0.00062. Then, the same problem was run with the interface

discretized by the zero-moment rule, which produced the interface nodes as depicted in Fig. 5.8. Once

again, the solution produced no visible results and the C-error was negligible at 0.00046. However,

these results are being compared to an analytical solution that was actually solved numerically in

Matlab with the Laplace transform performed in Mathematica.

m

0.1 m

0.1 m

Fluid Silent Boundaryk5 Mass

1

30

Figure 5.7: Model with non-matching meshes.

Therefore, to compare non-matching meshes, it was decided to contrast different meshes with one

that had 100 elements on the faces with a characteristic length of 1100 meters. For a non-matching mesh,

the mass of the structure was created with four elements with a characteristic length of 120 meters. The

fluid was then varied from a characteristic length of 120 to 1

90 . In this analysis, the interface frame was

discretized in the same manner as the structure. This produced the C-error results shown in Fig. 5.9.

At first glance, this looks as if no concrete conclusions can be drawn. However, the characteristic

lengths that appear to have a reduced error actually have the interface nodes at the zero-moment

locations. Coincidentally for these meshes, the zero-moment nodes correspond to the structure nodes.

The other values with a larger error do not meet the zero-moment rule requirement. It is noted that

the zero-moment rule improves the results for this analysis.

In comparing the LLM method with the CAFE method, both produced the same results for this

particular problem. A difference will be noticed in the two dimensional analysis. However, the most

notable item is the time required to run the analysis. The CAFE method requires a much smaller time

step to remain stable. This stability requirement is caused by the prediction step in the method, the

required prediction at the silent boundary, and the system of equations are numerically integrated with

an explicit central difference scheme. By discretizing the fluid into thirty elements and truncating the

fluid at three meters, the time step for the CAFE method was ∆t = 1x10−6 seconds. In comparison,

by use of the LLM method with an implicit trapezoidal method for the numerical integration, a time

step of ∆t = 1x10−1 seconds was used to produce accurate results. Thus, much fewer time steps

were needed for the LLM method. This more than compensated for the increase in computational

70

Frame nodes

Fluid nodes

Structure nodes

2

45

1

18m

m

Figure 5.8: Interface frame for the non-matching mesh.

10−2

10−1

10−5

10−4

10−3

10−2

Characteristic Length (m)

C-e

rror

Zero Moment Matching

Figure 5.9: C-error for the infinite piston system with non-matching meshes.

71

cost per time step by the LLM method. For instance, by programming both methods in Matlab, the

CAFE method took a little under three hours. The LLM method took nine seconds. Obviously, this

is problem dependent; in some cases, the time step will naturally need to be small to accurately model

the physics.

5.5.3 Stability Verification

Finally, the main point of this chapter was to verify that the partitioned transient method

produced with the LLM method, see Section 3.2, does not have any stability issues if used with an

implicit integration method. Thus, accuracy is imposing the time step rather than stability. Therefore,

an analysis was run with varying time steps. The first time step was taken to be one second, which is

far above the CFL condition for the fluid. The error can visibly be noticed at a time step value of 0.4

seconds. Then, the analysis was run until convergence, which is around 0.01 seconds, see Fig. 5.10.

10−2

10−1

100

10−4

10−3

10−2

10−1

Time step (s)

C-e

rror

Figure 5.10: Time step values and the error associated with the time steps.

5.6 Summary

This chapter provided a problem with an analytical solution of a spring-mass infinite piston. This

was used to verify the ability of the LLM method for a simple fluid-structure interaction. Originally,

this was used to validate the concept of using the gradient matrix; thus, a one-dimensional problem

was studied. The problem was then used to demonstrate the ability of the LLM method with three-

dimensional elements. Given the use of three-dimensional elements, the effect of the zero-moment rule

was shown to be beneficial for non-matching meshes. Finally, the chapter concludes with the fact that

72

the partitioned transient method can be unconditionally stable, if two A-stable integration schemes

are chosen for the fluid and structure models.

Chapter 6

Gravity Dam Benchmark Study: Vibration Analysis

The Koyna dam in India is one of the most famous and studied cases of gravity dams under an

earthquake loading. The majority of the studies involve the non-linear response of the dam, because

during an earthquake in 1967 the dam experienced cracking on its down stream face. However, most

have realized the importance of including the fluid in the simulations. Some of the first work was done

by Zienkiewicz et al. [130] and Chopra et al. [24].

In this chapter, the Koyna dam is also used for FSI to compare different non-matching mesh

techniques. After a short introduction to the problem, this chapter begins with a modal analysis.

This modal analysis provides a good physical understanding of the problem. In addition, kinematic

continuity issues concerned with the discretization of the interface frame are discussed by examining

the mode shapes of the system. This provides some general rules for the discretization of the interface

frame in regards to the method chosen for managing the non-matching meshes. Then, a frequency

content of the acceleration is provided in order to know the dominant frequencies of the earthquake.

The linear vibration portion concludes by investigating a frequency response function of the system.

This illustrates that the LLM method theoretically works for non-matching meshes and gives another

perspective for the physical understanding of the system. The chapter finishes with an example of the

interface patch test to illustrate the benefit of the zero-moment rule, for it assures kinematic continuity

with the LLM method.

6.1 Problem Description

The problem studied in this chapter is a two dimensional dam under seismic excitation. The

dam modeled is the Koyna dam. In December of 1967 an earthquake shook the region and caused

cracking in the down stream direction, see Fig. 6.1. In this study, the dam and the rock foundation

are assumed linear, elastic, isotropic, and homogenous. The dimensions of the dam are shown in

Fig. 6.2 with a coarse mesh. The mesh is truncated at both ends and silent boundaries are applied at

74

those ends. The structures silent boundary was tested by moving the mesh until no visible error was

detected in the results for the dam crest displacement without the fluid. The fluid’s boundary was

discussed in Chapter 4. The parameters for the system are defined in Table 6.1. Several items have

been studied for this problem ranging from crack propagation [67] to the effect of reservoir bottom

absorption [129]. Although, all should be conducted in a true analysis of a gravity dam, in this study

the focus is on the FSI with non-matching meshes.

(a) View of the Koyna dam. (b) Leaks in downstream face.

Figure 6.1: Downstream views of the Koyna dam: (a)Present day. (b) After the earthquake ofDecemeber 1967.

The El Centro earthquake of May 18, 1940, was chosen as the excitation for this problem. The

El Centro earthquake is a low frequency earthquake with a predominant period of 0.55 seconds, see

Fig. 6.3 for the acceleration component in the horizontal direction.

6.2 Linear Vibration Analysis

Before the transient analysis is carried out, a vibration analysis is performed to help determine

the accuracy of the methods. This section begins with a method for deriving the eigenvalue analysis for

the use of Localized Lagrange Multipliers and one for Mortar like methods [8, 13]. The mode shapes

and corresponding frequency are then displayed with the frequency content of the seismic acceleration

so that the relevant modes are determined which will be useful for reduced order modeling. Frequency

Response Functions (FRF) are then examined to give a better physical understanding.

6.2.1 Modal Analysis

An eigenvalue analysis is carried out in a similar manner as that of Park, Felippa, and Ohayon

[99]. Here, it is assumed that the damping and forcing functions are zero for the analysis and that the

75

Dam

Fluid

InterfaceRock Foundation

50 m

0 m

131 m

157 m

Figure 6.2: Koyna dam-reservoir system.

Acc

eler

atio

n(G

)

Time (s)

0

0 5 10 15 20−0.4

−0.3

−0.2

−0.1

0.1

0.2

0.3

Figure 6.3: El Centro earthquake May 18, 1940, horizontal component.

76Econcrete 31.46x109 Pa

ν 0.2ρconcrete 2690 kg/m3

Erock 18x109 Paν 0.2

ρrock 1830 kg/m3

cfluid 1439 m/sρfluid 1019 kg/m3

Table 6.1: Parameters for Koyna dam model.

structural displacements and the fluid displacement potentials are a linear combination of the mode

shapes.

uS = ΦSq, ψ = Φfdϕ, (6.1)

where

[KS − ωSiMS ]ΦSi = 0, [Kfd − ωfdiMfd]Φfdi = 0. (6.2)

Therefore, by replacing the displacement and displacement potential with Eq. (6.1) in the vari-

ational equation set, Eq. (2.55), and pre-multiplying the first two equations by the transpose of the

eigenvector matrices the following is obtained:

ΛS − ω2I 0 ΦTSBS 0 0

0 Λfd − ω2I 0 ΦTfdD

TBF 0

BTSΦS 0 0 0 −LS

0 BTFDΦfd 0 0 −LF

0 0 −LTS −LTF 0

q

ϕ

λS

λF

uB

=

0

0

0

0

0

, (6.3)

where ΛS and Λfd are matrices that contain the particular system’s eigenvalues on the diagonals, and

I is the identity matrix.

Thus, it can be noticed that the combined system’s eigenvectors (Φc) and eigenvalues (Λc) can

be determined from solving:

AcΦc = BcΦcΛc, (6.4)

where

Ac =

ΛS 0 ΦTSBS 0 0

0 Λfd 0 ΦTfdD

TBF 0

BTSΦS 0 0 0 −LS

0 BTFDΦfd 0 0 −LF

0 0 −LTS −LTF 0

; Bc =

I 0 0 0 0

0 I 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

. (6.5)

77

The eigenvectors (Φc) are currently in modal coordinates and need to be in physical coordinates.

This can be accomplished by the following:

Φ1c = ΦSΦcS (displacement coordinates),

Φ2cd = ΦfdΦcfd (displacement potential coordinates),(6.6)

where ΦcS is the eigenvector portion that is associated with the structure and Φcfd is the eigenvector

portion associated with the fluid. In addition, the system’s eigenvectors that are associated with the

fluid need to be modified from displacement potential coordinates to displacement coordinates. This

can be accomplished by multiplying by the gradient matrix:

Φ2c = DΦ2cd (displacement coordinates). (6.7)

A similar procedure can be carried out to obtain the eigenvalues and eigenvectors if using the

Mortar method, Appendix B. Beginning with the previous derived mode shapes of the subdomains

(ΦS , Φfd), and using Eq. B.2, the following system equation is developed for the Mortar method:

ΛS − ω2I 0 ΦTS BS

0 Λfd − ω2I ΦTfdD

T BF

BTSΦS BT

FDΦfd 0

q

ϕ

λ

=

0

0

0

. (6.8)

Again, the combined system’s eigenvectors (Φc) and eigenvalues (Λc) can be determined from solving:

AcΦc = BcΦcΛc, (6.9)

where

Ac =

ΛS 0 ΦTS BS

0 Λfd ΦTfdD

T BF

BTSΦS BT

FDΦfd 0

; Bc =

I 0 0

0 I 0

0 0 0

. (6.10)

As previously mentioned, the eigenvectors are currently in modal coordinates and would need to be

changed into physical coordinates.

6.2.2 Kinematic Continuity Issues

With the above derivations, the first item to be discussed is kinematic continuity with the

discretization of the interface frame. In order to evaluate the possibilities of non-matching meshes, the

mesh depicted in Fig. 6.4 is examined with the Mortar method and the LLM method. The interface

frame is then discretized as the structure mesh, because this satisfies the zero-moment condition to

pass a consistent stress. A selection of the mode shapes and frequencies for the LLM method are shown

78

0

0

100

100

200

200 300 400 500

50

150

x

y

Figure 6.4: Koyna dam-reservoir system with non-matching meshes

in Fig. 6.5. All of the mode shapes for the LLM method satisfied kinematic continuity at the interface.

Recall that due to the fluid inviscid assumption, only the normal components of the displacements are

constrained at the interface.

On the other hand, the Mortar method with the interface discretized according to the zero-

moment rule exhibits mode shapes where the fluid moves into the structure as illustrated in Fig. 6.6.

It should be noted that this is only caused when the following conditions exist: the interface frame

is discretized with the coarse mesh, the refined mesh is further discretized inside an element of the

coarse mesh, and the Lagrange multipliers shape functions are Dirac delta shape functions that provide

discrete values of the Lagrange multipliers at the nodes.

While the Mortar method produces undesirable results, the LLM method with the Lagrange

multiplier shape functions is not without faults as well. For instance, if the interface frame is discretized

as the refined mesh, and the refined mesh is further discretized inside an element of the coarse mesh,

then a similar problem can be noted for the LLM method, see Fig. 6.7. As discussed in Appendix B,

this formulation of the LLM method produces the same relationship for the boundary displacements

as the Consistent method if the interface frame is discretized as the boundary of the structure and the

structure is more refined then the fluid. Thus, the relevance of using the zero-moment rule is apparent.

Given the above analysis, the following rules need to be applied with discrete Lagrange multipli-

ers collocated at the nodes of the Lagrange boundary. If using the Mortar method, the discretization

of the frame cannot be the coarse mesh. If using the Consistent method, the refined mesh should

always be interpolated to the coarse mesh. This is generally not a problem in aeroelastic problems, for

the fluid is usually much more refined. If using the LLM method, then the interface frame should not

be discretized as the refined mesh. This is actually a positive asset, as using the refined mesh causes

79

(a) 6th mode: 0.54 Hz (b) 20th mode: 1.36 Hz

(c) 21st mode: 2.53 Hz (d) 27th mode: 4.94 Hz

(e) 34th mode: 6.69 Hz (f) 40th mode: 8.13 Hz

(g) 46th mode: 10.1 Hz (h) 49th mode: 11.3 Hz

(i) 55th mode: 12.4 Hz (j) 65th mode: 13.8 Hz

Figure 6.5: Mode frequencies and mode shapes of the system

80

Figure 6.6: 5th Mode shape of the system for the Mortar method with the interface discretized as thecoarse mesh.

Figure 6.7: 30th Mode shape of the system for the LLM method with the interface discretized as therefined mesh.

81

more equations to be solved in the interface equation, Eq. 3.7.

6.2.3 Frequency Content of Acceleration

Given the mode shapes, the frequency content of the acceleration is examined to note which

mode shapes of the system are important. Using the Fast Fourier Transform (FFT), the frequency

contents of the acceleration for the El Centro earthquake are obtained as shown in Fig. 6.8. Thus,

for the horizontal acceleration, the eigenvectors with a frequency up to 8Hz need to be included, and

the vertical acceleration needs the eigenvectors with a frequency up to 15Hz to capture the dominant

shapes. A portion of the dominant mode shapes are shown in Fig. 6.5 for the LLM method.

replacemen

Frequency (Hz)

Pow

ersp

ectr

um

00 10 15 20 25

0.5

1

1.5

2

3

2.5

4

3.5

5

5

4.5

(a) Horizontal accelerationFrequency (Hz)

Pow

ersp

ectr

um

00 10 15 20 255

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

(b) Vertical acceleration

Figure 6.8: Frequency content of the seismic acceleration.

6.2.4 Frequency Response Function

Next, a selected frequency response function is examined to give a better understanding of the

system. The frequency response plots describe the dynamic response due to a persistent harmonic

excitation with various excitation frequencies [119]. By comparing the vibration amplitude for various

excitation frequencies, the critical vibration frequency for the system can be determined. Here, a

partitioned analysis and an assembled analysis are performed to clarify the dynamic contribution of

each subsystem to the global dynamics.

First, the system is partitioned into two subsystems as was previously done for the LLM method.

Frequency response functions are computed for both subsystems. Figure 6.9 illustrates the frequency

response amplitude of each subsystem in the horizontal (x-direction). The input and output locations

for the frequency response plots are depicted in the top portion of the figure. All response magnitudes

82

are normalized by the static displacement of a unit force at the input degree of freedom (DOF). It

can be noticed that the fluid experiences a zero energy mode when there is no interaction from the

structure. There is only one non-physical zero-energy mode for the fluid when it is partitioned from

the system, which represents the constant displacement potential mode. This is expected, because

this zero-energy mode corresponds to a rigid body mode for the fluid that is formulated in terms of a

displacement potential. This also demonstrates that the fluid formulation only has zero-energy modes

associated with rigid body modes. However, a free surface condition must be applied to remove any

accompanying zero-energy modes with the free surface as noted by Wilson et al. [124].

Input DOF

Output DOF

(a) Input/ output for structure

Input DOF

Output DOF

(b) Input/ output for fluid

Frequency (Hz)

Nor

mal

ized

vib

ration

mag

nitude 103

102

101

100

10−1

10−20 5 10 15 20

(c) Frequency response of the structure

Frequency (Hz)

Nor

mal

ized

vib

ration

mag

nitude 1012

1010

108

106

104

102

100

10−2

10−4

0 5 10 15 20

(d) Frequency response of the fluid

Figure 6.9: Frequency response of the partitioned subsystems.

A couple of items can be noted for the assembled system. First, it is noted that for low frequencies

the system acts as the structure alone, see Fig. 6.10. However, after 3Hz, it is noted that the system

responds to more frequencies than the structure alone, but still responds to the dominate structure

frequencies. Clearly, the assembled system’s response is different than the two systems alone. This

justifies the need for the fluid-structure interaction. Second item to be noted is if the output DOF

is the dam face at the top of the water, and the input DOF is associated with the fluid Lagrange

multiplier at the bottom, then this will produce the same result if the input DOF is associated with

the structure Lagrange multiplier at the same location, as shown in Fig. 6.11. This demonstrates that

the fluid formulation with the LLM method is working. However, if the input DOF stays the same,

such as the one associated with the dam Lagrange multiplier, but the output DOF changes from the

83

structure to the fluid then different results are produced, as shown in Fig. 6.12. This is expected,

because the displacement constraint is that of the normal displacements.

Frequency (Hz)

Nor

mal

ized

vib

ration

mag

nitude 103

102

101

10−1

10−2

100

0 5 10 15 20

Structure AloneAssembled System

(a) Frequency response of the system comparedto the structure.

Frequency (Hz)

Nor

mal

ized

vib

ration

mag

nitude 104

103

102

101

10−1

10−2

10−3

100

0 5 10 15 20

Fluid AloneAssembled System

(b) Frequency response of the system comparedto the fluid.

Figure 6.10: Frequency response of the assembled system.

6.3 Examples of Passing the Patch Test

In this section, an example of the zero-moment rule used to pass the patch test is presented.

The advantage of a simple method used to preserve a constant state of stress is extremely useful. This

has been a recent field of study in contact problems [30, 88]. Crisfield [30] proposed the determination

method of the interpolation function of the contact force field based on the contact patch test, and

demonstrated that an appropriate choice of the function leads to an improvement in accuracy. Here,

Felippa, Park, and Rebel’s [48] work is utilized to maintain an energy-preserving frame. It is on the

frame where the interface equilibrium condition is enforced with the use of the localized Lagrange

multipliers as described in Section 2.3.3.1.

For simplicity, attention is focused on the interface frame with the nodal forces from the subdo-

mains. An example is the interface frame with subdomain forces as shown in Fig. 6.13(b), which is a

characteristic portion of the system with the mesh shown in Fig. 6.13(a). In Fig. 6.13(b), the desired

solution (the bottom forces) is provided, if a uniform stress is applied to the top surface. The top sur-

face constant stress is proportioned to the top nodes by conventional methods, such as node-by-node

lumping.

In this example, six cases are evaluated. The first three consist of the LLM method with

the interface frame discretized as the coarse structure mesh, Fig. 6.14(a); the refined fluid mesh,

84

Frequency (Hz)

Nor

mal

ized

vib

ration

mag

nitude

103

102

101

10−1

10−2

100

0 5 10 15 20

Input DOF = Structure λ

Input DOF = Fluid λ

Figure 6.11: Frequency response of the assembled system with the dam face as the output DOF.

Frequency (Hz)

Nor

mal

ized

vib

ration

mag

nitude

103

102

101

10−1

10−2

10−3

100

0 5 10 15 20

Output DOF = Structure uS

Output DOF = Fluid uF

Figure 6.12: Frequency response of the assembled system with the dam Lagrange multiplier as theinput DOF.

(a) A non-matching mesh for the dam system problem.

0.50.5

1.01.01.0 1.01.0

1.51.5 1.5

0.750.75

f = 6.0

f = 6.0

8

33

(b) A characteristic portion of thenon-matching mesh.

Figure 6.13: An example of a constant state of stress across the interface frame.

85

Fig. 6.14(c); and the use of the zero-moment rule, Fig. 6.14(e). The only discretization that produces

a constant stress state on the bottom is the discretization with the zero-moment rule. Once again, the

LLM method with the coarse mesh, Fig. 6.14(a) is the same as the Consistent Interpolation method,

see Appendix B. Note that with the LLM method, regardless of the interface frame discretization,

the sum of the forces on one side is equal to the other. The next three cases consist of the Mortar

method with the interface frame discretized as the coarse structure mesh, Fig. 6.14(b); the refined

fluid mesh, Fig. 6.14(d); and the use of the zero-moment rule, Fig. 6.14(f). Only the interface with the

coarse mesh, Fig. 6.14(b), has a constant stress on the bottom nodes. However, this interface frame

discretized with the coarse mesh is not conservative, because the sum of the forces on one side is not

equal to the sum of the forces on the other.

These examples are only relevant for the Lagrange multipliers that are interpolated with the use

of the Dirac delta function. However, this provides a very simple, efficient method with the use of the

zero-moment rule that can pass the interface patch test for the LLM method. With the information

from the linear vibration analysis and the zero-moment rule, the next logical step is to begin the

transient analysis.

6.4 Summary

This chapter provided a theoretical understanding of the LLM concept through the exploration

of a frequency response function for the FSI. Also, the linear vibration analysis and the frequency

content of the acceleration provide a physical understanding of the system. It was discovered that the

structure is the dominant subsystem, and would generally require a more refined mesh for a better

representation of the system. If a nonlinear analysis of the structure were performed, then the mesh

would most definitely be more refined than the fluid.

The linear vibration study demonstrated that kinematic continuity problems could develop if

the interface frame was poorly discretized. This produced a set of rules for the discretization of the

interface frame. However, the use of the zero-moment rule for the LLM method will always assure

kinematic continuity at the interface. An example of how the LLM method and the zero-moment rule

are capable of propagating a consistent stress across two domains, an interface patch test, is shown at

the conclusion of this chapter.

86

0.50.5

1.01.01.0 1.01.0

0.830.83

1.33 1.331.66

f = 6.0

f = 6.0

(a) LLM: coarse mesh at interface.

0.50.5

1.01.01.0 1.01.0

0.50.5

1.01.0 1.0

f = 6.0

f = 4.0

(b) Mortar: coarse mesh at interface.

0.50.5

1.01.01.0 1.01.0

0.50.5

2.0 2.01.0

f = 6.0

f = 6.0

(c) LLM: refined mesh at interface.

0.50.5

1.01.01.0 1.01.0

0.830.83

1.33 1.331.66

f = 6.0

f = 6.0

(d) Mortar: refined mesh at interface.

0.50.5

1.01.01.0 1.01.0

1.51.5 1.5

0.750.75

8

33

f = 6.0

f = 6.0

(e) LLM: zero-moment mesh at inter-face.

0.50.5

1.01.01.0 1.01.0

8

3 3

0.610.61

1.77 1.771.22

f = 6.0

f = 6.0

(f) Mortar: zero-moment mesh at in-terface.

Figure 6.14: An example of a constant state of stress across the interface frame, where ”x” is thelocations of the interface nodes.

Chapter 7

Gravity Dam Benchmark Study: Transient Analysis, and Error

7.1 Introduction

This chapter is concerned with the transient response of the Koyna dam subject to the earthquake

loading of the El Centro earthquake in 1940. This chapter begins with a description of the load vector

for the CASE method and the LLM method. Then an analysis with matching meshes is conducted for

both the CASE and the LLM transient methods. Before exploring non-matching meshes, the scaling of

the interface equation for the LLM method is examined with energy errors. Next, an investigation with

non-matching meshes is conducted. For the non-matching meshes, the LLM transient method is first

study. Then the CASE transient method with the inclusion of the LLM for the non-matching meshes

is compared with the CASE transient method with the non-matching handled by the Mortar method.

Once again, it is noted that the LLM method is capable of including the consistent interpolation

method and in locations where it does, it is noted in this chapter.

In this chapter, the acceleration data of the earthquake is amplified by a factor of 1.5. The

seismic excitation is modeled in both the horizontal direction and the vertical direction. It is assumed

that the entire base is excited at the same excitation rate at all nodes at a particular period in time.

Thus, the acceleration, velocity, and displacement are known at the base of the rock foundation at

each given time.

7.2 Load Vector

In many earthquake engineering analyses the load vector is determined from a relative dis-

placement concept [68], given below. In the CASE method the fluid code must receive the total

displacement. In the CASE method this is not a problem, because the total displacement to be sent to

the fluid code can be determined by adding the relative displacement to the given ground displacement.

However, in the LLM method the Interface equation, Eq. (3.7) also needs the total displacement, total

velocity, and the total acceleration. This can also be accomplished by adding the relative displace-

88

ment to the ground displacement and similar for the velocity and the acceleration. However, to avoid

confusion and to create simplicity, it is best to use the total displacement, total velocity, and total

acceleration in the equations. This load vector is described following the relative displacement concept

load vector.

7.2.1 Relative Displacement’s Load Vector

To illustrate the concept, consider the typical dynamic structure shown in Fig. 7.1, where ki

represents the lateral stiffness of a member, mi is the mass of the member, and lateral viscosity is

ignored for clarity. Here, the structure is subject to an earthquake induced displacement, velocity, and

acceleration (ug, ug, ug). Summation of the forces produces the following:

m1u1 − k2(u2 − u1) = 0,

m2u2 + k2(u2 − u1) − k3(u3 − u2) = 0,

m3u3 + k3(u3 − u2) = 0.

(7.1)

k3

2

k3

2

k2

2

k2

2

u3

u2

u1 = ugug

x2

x3

m3m3m3

m2m2m2

m1 m1m1

k3(u3 − u2)

k2(u2 − u1)k3(u3 − u2)

k2(u2 − u1)

m3u3

m2u2

m1u1

FBDRigid Body Motion

Figure 7.1: Relative displacement concept.

A relative displacement is defined as follows given the earthquake displacement (ug) and accel-

eration (ug).

x1 = u1 − ug; thus, x1 = u1 − ug.

x2 = u2 − ug; thus, x2 = u2 − ug.

x3 = u3 − ug; thus, x3 = u3 − ug.

(7.2)

89

Inserting Eq. (7.2) in Eq. (7.1) provides the following equation set in matrix form.

m1 0 0

0 m2 0

0 0 m3

x1

x2

x3

+

k2 −k2 0

−k2 k2 + k3 −k3

0 −k3 k3

x1

x2

x3

= −

m1 0 0

0 m2 0

0 0 m3

ug

ug

ug

. (7.3)

Note the boundary condition x1 = u1 − ug = 0, provides the reduced system

m2 0

0 m3

x2

x3

+

−k2 + k3 −k3

−k3 k3

x2

x3

= −

m2 0

0 m3

ug

ug

. (7.4)

Once again, it is important to remember that the total displacement is the sum of the relative dis-

placement plus the ground displacement. This formulation is used for the CASE method problems.

7.2.2 Total Displacement’s Load Vector

The total displacement’s load vector maintains the equations of motion of the system in terms of

the total state variables. If the structure is redrawn in a more intuitive motion, as shown in Fig. 7.2,

the summation of the forces will still be the same; however, in the below equation, Eq. (7.5), the

displacements have been reversed to reflect the situation in Fig. 7.2.

m1u1 + k2(u1 − u2) = 0.

m2u2 − k2(u1 − u2) + k3(u2 − u3) = 0.

m3u3 − k3(u2 − u3) = 0.

(7.5)

k3

2

k3

2

k2

2

k2

2

u3

u2

u1 = ugug

m3m3 m3

m2m2m2

m1m1m1

k3(u2 − u3)

k2(u1 − u2)

k3(u2 − u3)

k2(u1 − u2)

m3u3

m2u2

m1u1

FBDRigid Body Motion

Figure 7.2: Total displacement concept.

90

This provides the matrix form:

m1 0 0

0 m2 0

0 0 m3

u1

u2

u3

+

k2 −k2 0

−k2 k2 + k3 −k3

0 −k3 k3

u1

u2

u3

=

0

0

0

. (7.6)

Once again, the boundary condition needs to be applied u1 = ug and u1 = ug; thus, a nonzero

prescribed boundary is used. A simple method is to augment the force vector (f) by multiplying

the full mass and stiffness matrices by their respective state variable values with only the prescribed

values:

f = −

m1 0 0

0 m2 0

0 0 m3

u1 = ug

0

0

k2 −k2 0

−k2 k2 + k3 −k3

0 −k3 k3

u1 = ug

0

0

. (7.7)

Then, the set of equations associated with the prescribed boundaries are removed to yield the following

system of equations:

m2 0

0 m3

u2

u3

+

−k2 + k3 −k3

−k3 k3

u2

u3

=

k2ug

0

. (7.8)

7.3 Matching Mesh Comparison for CASE/CAFE and LLM

Before evaluating the non-matching capabilities of the proposed method, a set of benchmark

data is established for the matching mesh case for the transient analysis with the CASE and the LLM

methods. In both transient methods, the mesh is refined until convergence is observed for the relative

displacement at the crest of dam. Both methods showed great convergence with the mesh shown

below in Fig. 7.3. The C-error for both methods with varying meshes is shown in Fig. 7.4. Here, the

benchmark mesh contains 3598 DOF for the system. The displacements for both methods are shown

in Fig. 7.5, where they are remarkably similar. Thus, each solution will be used for the benchmark

results for the given method when evaluating non-matching meshes.

7.4 Scaling the Interface Matrix Verification

Before the non-matching analysis is discussed, a brief moment is taken to verify the benefit of

scaling the interface equation, Eq. 3.7, as described in Section 3.2.1. The interface equation has a

potential to be extremely ill-conditioned. The ill-conditioned nature arises because the values of the

interface equation matrix may vary greatly in size, due to the heterogeneous physical systems and

the magnitude of the connection matrices (L). For example, if the standard parameters are used for

91

Figure 7.3: Portion of Koyna dam mesh used as benchmark mesh for C-error.

Total System DOF

C-e

rror

0.1

0.01100 1000 10000

(a) LLM Transient method

Total System DOF

C-e

rror

0.1

0.01

100 1000 10000

(b) CASE Transient method

Figure 7.4: C-error values with matching meshes.

0 1 2 3 4 5 6 7 8−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

Time (s)

Dis

pla

cem

ent

(m)

CASE

LLM

Figure 7.5: Crest Displacement of Koyna dam of converge results for the transient analysis by theLLM method and the CASE method.

92

the physical components of the system shown in Fig. 6.13(a), then the interface matrix will have a

condition number of 5.9845x106 where the condition number was determined from the spectral norm.

However, if the matrix is scaled as discussed in Section 3.2.1, then the matrix is well conditioned with

a condition number of 38.5592. Clearly, the benefit of scaling is realized.

However, this scaling does introduce a small numerical rounding error. This is noticed by

examining the difference in energy at the interface frame, as discussed in Section 3.4 and written

below for convenience.

W (n)interror = uTB(−LTSλS − LTFλF )

The only contributor to this error is the difference in the Lagrange multipliers at the boundary. This

can be caused by the connection matrices, LS and LF , or the scaling matrices, SF , SS , and SB , used

for the interface equation. If matching meshes are examined, then the only cause of error would be

from the scaling matrices, which is noticed in Fig. 7.6. Here, it is noticed that the difference in energy

is extremely minimal. The maximum absolute value is 3.15x10−7 Nm when the work on one side of

the interface is equal to 2.2x108 Nm at the maximum value. Thus, it appears that this is caused by

round-off error during the solutions of the systems with the scaling matrices. There is no interface

equation in matching meshes with the CASE transient method; however, the connection matrices LS

and LF are used. Thus, for the matching mesh case, there will be no round-off error and, as expected,

there is no energy difference at the interface. This was noticed for all of the matching meshes run

under the CASE method.

0 1 2 3 4 5 6 7 8−4

−3

−2

−1

0

1

2

3x 10

−7

Time (s)

Ener

gy(Nm

)

Maximum = −3.15x10−7Nm

Figure 7.6: Energy difference caused by scaling at the interface of matching meshes. Energy on oneside of the interface is of the magnitude of 106 Nm on average.

93

7.5 Non-Matching Mesh Comparisons for both CASE and LLM

In this section, the ability of the localized Lagrange multipliers for non-matching meshes of

different subdomains is evaluated with the transient method developed by the LLM method and

the CASE method. Some general trends will be discussed in regards to the different methods for

non-matching meshes. Before, discussing the pros and cons of each method an example problem is

provided.

As an example, the mesh shown in Fig. 6.13(a) is tested, and provides a discretization for a

portion of the interface as shown in Fig. 6.14. Here, the fluid length is characterized by the height

of the elements in contact with the dam (5.5 m in this problem). The structure length is the height

of elements at the same location (8.25 m in this problem). The interface is then discretized as either

matching the structure mesh, matching the fluid mesh, or by the zero-moment rule. In some other

simulations, the zero-moment rule is the same as the structure or the fluid.

The relative displacement at the crest of the dam is shown in Fig. 7.7 where the transient

method is provided by the LLM concept for this example problem. This figure depicts the three

different results for the different interface frame discretizations. As is typical for the majority of the

simulations, the difference in relative displacement is small. Here, it is noted that the zero-moment

rule does not provide the best results in terms of the C-error. However, it is very close.

0 1 2 3 4 5 6 7 8−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

Time (s)

Dis

pla

cem

ent

(m)

Zero moment rule (C-error = 0.105)

Refined match (C-error = 0.168)

Coarse match (C-error = 0.104)

Figure 7.7: Dam crest displacement values for different interface meshing with LLM for the transientanalysis and non-matching meshes.

It should be noted that there is a situation in other simulations with the potential to become

problematic. If the interface is discretized as the refined mesh, and the refined mesh is discretized

94

within the coarse mesh, then the forces from the refined mesh inside the coarse mesh are not transferred

to the coarse mesh, this is depicted in Fig. 7.8. This can become a serious issue if it is difficult to

determine which is the refined mesh or if the meshes change from being the refined to the coarse along

the boundary. Fortunately, this can be detected by examining the energy error across the interface

as explained in Section 3.4. If this situation arises, the system will become non-conservative in terms

of energy at the subdomains’ interface boundaries. However, this will be avoided if the zero-moment

rule is used, as it assures that with the LLM method all forces are transferred to the other domain.

In addition, by matching the coarse mesh the forces are correctly transferred.

Fluid Boundary

Interface frame

λF

λS

Structure Boundary

(a) Interface meshmatching the refinedmesh

Fluid Boundary

Interface frame

λF

λS

Structure Boundary

(b) Interface meshmatching the coarsemesh

Figure 7.8: Transfer of forces with the LLM method.

The difference in the energy across the interface frame discretized by the zero-moment rule for the

example problem is shown in Fig. 7.9. The average absolute value of the energy difference is 1.16x10−7

Nm, which is incredibly insignificant compared to the average energy on the structural boundary at

the interface (6.78x106 Nm). Thus, it is shown that using the localize Lagrange multipliers creates a

conservative system.

The same example problem is used with the CASE transient code with the use of the LLM

method for the non-matching meshes as described in Section 3.3. This provides the relative displace-

ment values as shown in Fig. 7.10. Once again, it is difficult to notice a great variance between the

different techniques for discretizing the interface frame. Also, there can be a serious issue if the frame

is discretized as the refined mesh and the refined mesh is significantly more refined than the coarse

mesh. This can result in a severely ill-conditioned matrix,[

LnS(LnS)T]

, which is to be inverted and

used for the relation between the subdomain’s forces. It is again pointed out that the zero-moment

rule will assure that this will not happen, or the coarse mesh should be used for the interface frame

discretization.

Now, the use of the Mortar method with the CASE transient analysis as explained in Ap-

95

0 1 2 3 4 5 6 7 8−8

−6

−4

−2

0

2

4

6

8

10x 10

−7

Time (s)

Ener

gy(Nm

)

Average Energy Difference = 1.16x10−7Nm

Figure 7.9: Energy difference across the interface frame. Average energy on structure side is6.78x106Nm.

0 1 2 3 4 5 6 7 8−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

Time (s)

Dis

pla

cem

ent

(m)

Zero moment rule (C-error = 0.181)

Refined match (C-error = 0.183)Coarse match (C-error = 0.178)

Figure 7.10: Dam crest displacement values for different interface meshing with CASE for the transientanalysis and LLM for the non-matching meshes.

96

pendix B.1 is reviewed. It is typical for the interface frame in the Mortar method to inherit the

discretization of one of the two sides of the subdomains. This provides the relative displacement val-

ues as shown in Fig. 7.11. As can be expected, there is not a great difference. However, like the LLM

method, the Mortar method can also create a severely ill-conditioned matrix used for the correlation

of the subdomain forces. In the Mortar method, a severely ill-condition matrix arises from using the

coarse mesh as the discretization for the interface frame. Here, it can be noted that there are times

when the zero moment rule will be the same as the coarse mesh; thus, the zero-moment rule cannot

be used for the Mortar method. In addition, it should be noted that the C-error for the Mortar

method with the refined mesh at the interface is the same as the LLM method with the CASE tran-

sient analysis. This is expected as described in Appendix B and would be the same as the Consistent

Interpolation method.

0 1 2 3 4 5 6 7 8−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

Time (s)

Dis

pla

cem

ent

(m)

Refined match (C-error = 0.178)Coarse match (C-error = 0.182)

Figure 7.11: Dam crest displacement values for different interface meshing with CASE for the transientanalysis and Mortar for the non-matching meshes.

Finally, to verify that the methods actually work, a variety of different non-matching meshes

are examined. Here it is expected that as the total number of degree of freedoms (DOF) increases,

the C-error should decrease. Therefore, four figures are provided in Fig. 7.12. Each figure shows the

C-error versus the total DOF. If the data is fit with a best-fit exponential curve, then it is evident

that refining the model does improve the C-error. It is expected that there will be discrepancies

around a total DOF value for refining the structure mesh has a greater influence than refining the

fluid mesh. This was envisioned in the linear vibration analysis where the frequency response function

of the system generally followed that of the structure.

97

0 500 1000 1500 2000 2500 3000 35000

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

C-e

rror

Total DOF

(a) CASE transient with LLM and Interface same asCoarse.

0 500 1000 1500 2000 2500 3000 35000

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

C-e

rror

Total DOF

(b) CASE transient with LLM and Interface same asZero-moment rule.

0 500 1000 1500 2000 2500 3000 35000

0.05

0.1

0.15

0.2

0.25

0.3

0.35

C-e

rror

Total DOF

(c) LLM transient with LLM and Interface same as Zero-moment rule.

0 500 1000 1500 2000 2500 3000 35000

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

C-e

rror

Total DOF

(d) CASE transient with Mortar and Interface same asRefined.

Figure 7.12: C-error for different Total DOF of non-matching meshes.

98

7.6 Summary

The main findings of this chapter pertain to the discretization of the interface frame regardless

of the method used for non-matching meshes. The primary discovery was that if the zero-moment

rule is not used with the LLM method, then there is potential for concerns either with highly ill-

conditioned relation matrices or incredibly non-conservative energy systems. This observation was

first noted during the linear vibration analysis where kinematic continuity issues were examined with

the use of the system’s mode shapes. This was also discussed during the transient analysis for non-

matching meshes. In the transient analysis, the use of the zero-moment rule performed nearly as

well as any other interface discretization; thus, establishing that the zero-moment rule and the LLM

method should be the preferred procedure for non-matching meshes.

During the transient analysis, it is noted that running the LLM method given the implementation

presented in this study with the interface frame discretized as the coarse mesh produces the same results

as the Consistent Interpolation method discussed in Appendix B. Finally, it was also observed that

for problems in this study the structure required a greater refinement then the fluid for convergence.

Therefore, a need for non-matching meshes at the interface is justified.

Chapter 8

Reduced Order Modeling and Cavitation with the Transient LLM method

Any multi-physics system contains several difficult aspects for implementation. This study has

focused on the interaction of the heterogeneous systems. Now, the direction of study takes a slight turn

and looks at reduced order formulations and the nonlinear effect of cavitation for the transient analysis

developed in conjunction with the LLM method. This chapter then concludes with a discussion of

the operational cost of using the LLM transient method versus other explicit, staggered, partitioned

methods.

8.1 Reduced Order Modeling

This section discusses the possibility of using reduced order formulations for the problems in

this study. The objective of any reduced order model (ROM) is to reduce the size of the system; thus,

reducing the number of equations to be solved. For instance, suppose the size of uS and ψ are large

and of order ns and nf respectively. Then, a reduction to the number of degrees of freedom can be

achieved through a set of ks and kf basis vectors:

uS(t) =

ks∑

j

qj(t)φj = ΦSq(t),

ψ(t) =

kf∑

j

ϕj(t)φj = Φfdϕ(t),

(8.1)

where ks and kf are smaller than ns and nf , and ΦS and Φfd are the set of basis vectors. Here it can

be noticed that the fluid has already been reduced from a displacement to a displacement potential,

where the gradient matrix, D, was composed of the basis vectors. Hence, by forcing the irrotational

condition the fluid system is reduced by half.

uF = Dψ.

Therefore, in a similar manner, but including the structural equation, the systems energy func-

tional, Eq. 2.54, is modified. Thus, pre-multiplying the energy functional of the structure by the

100

structure set of basis vectors and doing the equivalent for the fluid, produces the following equation

set:

δΠ(q,ϕ,λS,λF ,uB) =

δqT (ΦTSMSΦSq + ΦT

SCSΦSq + ΦTSKSΦSq + ΦT

SBSλS −ΦTS fS)+

δϕT (ΦTfdMfdΦfdϕ+ ΦT

fdCfdΦfdϕ+ ΦTfdKfdΦfdϕ+ ΦT

fdDTBFλF − ΦT

fdDT fF )+

δλTS (BTSΦSq − LSuB) + δλTF (BT

FDΦfdϕ− LFuB)+

δuTB(−LTSλS − LTFλF ).

(8.2)

Given this reduced equation set the transient analysis can be performed in the same manner as in

Section 3.2.

One of the greatest difficulties in any multi-physics problem is the selection of the basis to use for

the reduction. Allen [1] recommends the use of proper orthogonal decomposition (POD) for coupled

multiphysics problems. Ohayon [95] uses Ritz vectors for situations with FSI, where the Ritz vector

for the fluid is determined by considering the response of the fluid to a prescribed displacement at

the fluid-structure interface. Despite all warnings, the set of basis vectors chosen for this study is the

physical eigenmodes, see Section 6.2.

With the coupled system, it was noticed that there are several low frequency sloshing modes that

have little effect on the structure. Thus, if these were chosen has a basis for reduction, the structure

system equations could potentially become severely ill-conditioned. An alternate method to address

this concern would be to use a reduced basis set for the structure that consists of eigenvectors from the

coupled system that visibly show effect on the structure. However, this can be tedious to determine.

Therefore, each subsystem’s eigenmodes are used for the reduction. However, has Ohayon notes [95]

that if the dry structure modes are used, very slow convergence can be expected. Nonetheless, using

the subsystem’s eigenmodes can demonstrate the ability of the method to incorporate reduced order

modeling.

8.1.1 Computational Results for ROM

In order to evaluate the ability of the LLM method to incorporate reduced order modeling, the

refined non-matching mesh as shown in Fig. 8.1 is used. First, all the eigenvectors of the fluid are used

while the structure’s use of eigenvectors for the basis are being decreased until a substantial error is

noticed. Then, the same procedure is repeated for the fluid. The results are shown in Fig. 8.2, where

the frequency associated with each eigenvector is displayed. As can be seen, the fluid can be reduced

101

to a frequency much closer to the excitation frequency, see Fig. 6.9, where a maximum value of 15Hz

should be considered. The ability of the fluid to be reduced further is expected for a couple of reasons.

First, as already noted, the dam has a greater influence on the system than that of the fluid. Second,

the fluid can include many low frequency modes in the first few sets of modes. Third, using the dry

modes of the structure has demonstrated the need to include more mode shapes, but this does not

limit the ability for this particular problem. However, there is a concern about how quickly the fluid

selection can become inaccurate. Given the above, if the maximum excitation frequency is around

15Hz then the fluid eigenvectors that are associated with two times that amount should be included

and the structure system should include its eigenvectors associated with six times that amount. When

performed in this problem, each system’s equations are reduced to 20% their original value with the

response as shown in Fig. 8.3. Here the response of the ROM is compared to the refined system that

originally showed convergence. It can be noticed that the ROM performs exceptionally well given the

fact that the meshes are non-matching.

Figure 8.1: Characteristic portion of the non-matching mesh used for the ROM test.

There are methods for improvement of the ROM; however, none are present in this portion of

the study so that the potential of the method can be substantiated. It has been demonstrated here

that it can be rather easy to include a reduced formulation for the LLM method. Next, the effect of

a non-linear analysis is examined through the physical concept of cavitation.

8.2 Cavitation

The main benchmark problem of this study is a dam under seismic excitation. During severe

ground motion, it is possible that the absolute water pressure may drop below the water vapor pressure,

producing the phenomenon known as cavitation [5, 85, 113]. The pressure of the water may drop to the

102

0 5 10 15 20 25 30 350.06

0.08

0.1

0.12

0.14

0.16

0.18

0.2

Frequency (Hz)

C-e

rror

(a) Reducing fluids system only.

20 40 60 80 100 120 140 1600.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

Frequency (Hz)

C-e

rror

(b) Reducing structure system only.

Figure 8.2: Reduction of system up to the frequency of the eigenvector.

0 1 2 3 4 5 6 7 8−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

Time (s)

Dis

pla

cem

ent

(m)

ROM (C-error = 0.066)Refined matching mesh

Figure 8.3: Dam crest displacement values for a reduced order model using 20% of the eigenvectors ofthe structure and the fluid.

103

water vapor pressure during the period when the dam’s wetface is pulling away from the fluid, which

has been noted in numerous articles and in an experiment by Niwa and Clough [93]. This changes the

hydrodynamic forces on the dam and has the potential to effect the response of the dam. Clough and

Chu-han [27] have noted the possibility that the tensile stress can be increased by 20 − 40% during

the impact of the water against the dam face caused by the closure of the cavitation region. This is a

concern due to the fact that the increased tensile stress may cause crack growth.

When the dam face moves away from the fluid, the absolute water pressure falls below the

vapor pressure causing a micro bubble formulation known has cavitation [105]. In order to model

this phenomenon two models have been presented in the literature. Both models utilize a bilinear

relationship and note that during the process of cavitation the compressibility and the bulk modulus

of the water are reduced towards zero, see Fig. 8.4 [114], where p is the pressure, pv is the vapor

pressure, K is the bulk modulus of the fluid, ρ0 is the density of the saturated liquid, and ρ is

the instantaneous density. Thus, the first model utilizes the fact that the incipience of cavitation is

associated with a local sudden drop in the tangent water compressibility and thus modifies the fluid

stiffness [36, 96]. The second model utilizes the fact that the pressure of the cavitation region has

changed and makes appropriate corrections through the pressure term [45, 109, 114, 130].

p

(ρ − ρ0)/ρ00

Κ

−pv

Figure 8.4: Density pressure relation for a bilinear fluid [114].

The method of including cavitation by modifying the pressure is incredibly convenient, because

it simplifies the analysis. The absolute pressure is either the equilibrium pressure plus pressure de-

termined from the bulk modulus and the volume strain, or it is zero. If the analysis begins in static

equilibrium [130], then

p =

−Kδuiδxi

, if p+ ph > 0,

−ph, otherwise,(8.3)

where it has been assumed that the vapor pressure is negligible in comparison to the atmospheric

pressure (pv = 0), ph is the hydrostatic pressure, and K represents the bulk modulus. In order to

104

utilize this relation, a variable in the fluid equations needs to be related to the pressure term. The

second derivative of the displacement potential is related to the pressure by using Eq. (2.5), Eq. (2.36),

and the wave equation, all written below for convenience.

p = −K∂uFk∂xk

= −ρF c2∂uFk∂xk

.

uF i = ∇ψ.

ψ = c2∇2ψ. (8.4)

Therefore, the second time derivative of the displacement potential is equal to the negative pressure

via the fluid density:

ψ =−pρ. (8.5)

Conveniently, the transient analysis is carried out by solving for the second time derivative of the

state variables. However, it would be an error to adjust this value for the LLM method. In essence,

modifying this second time derivative of the displacement potential modifies the displacement of the

fluid. This then causes conservation of energy concerns at the interface, because the force displacement

relationship on the structure side is not the same as the force displacement relation on the fluid side.

The LLM transient method actually illustrates this problem. This is because the LLM method enforces

the subdomain boundary displacements to be the same and is conservative in an energy sense. If the

above analysis were to be used, then during the time step, the fluid boundary displacement would

be reassigned another value. This causes a greater value for the Lagrange multiplier of the structure,

as it would tend to try and glue the values back together. Ultimately, this could lead to drift of the

structure.

Therefore, the logical implementation of cavitation for the LLM method is to adjust the bulk

modulus of the fluid. By adjusting the bulk modulus, the displacement of the fluid boundary and

the structure boundary maintain their equilibrium values, yet the fluid’s rigidity drastically changes

during cavitation. This method is described in the remainder of this section. Here, the internal energy

can change within an element, but not the overall system energy.

8.2.1 Modification of the Fluid Element Stiffness for Cavitation

Intuitively, the region of cavitation will be characterized by approximately a uniform pressure

that is equal to the pressure of the gas. Mechanically, the region of cavitation experiences a bulk

modulus close to zero, and this condition persists until the pressure is increased again by the collapse

of the cavitated region [36]. A simple but satisfactory model to implement this phenomenon is a

105

bilinear one in which the bulk modulus is that of water, considered as an acoustic medium, when

pressure is superior to the vaporization pressure, and zero when pressure is less then or equal to the

vaporization pressure. The modification of the constitutive properties during cavitation is similar to

crack smearing approaches for tensile crack growth in concrete [19, 73].

To incorporate this bilinear behavior, the pressure must be calculated at each time step. Due to

the assumption of the small compressibility of a liquid, the pressure can be calculated from Eq. (2.5).

This leads to the pressure being determined by the volumetric strain displacement matrix as done in

Section 4.3. Therefore, the pressure is found at gauss points for full integration. Given these values,

the element pressure is determined at the center, as computed stresses are usually most accurate at

locations within an element rather than on its boundaries [29].

Another method for computing the pressure in the element is to compute the pressure at the

nodes, p, due to the following relation.

p = −ρψ, (8.6)

which was derived from above, Eq. (8.5). Given the pressure at the nodes, the element pressure can

be assumed to be the average of the node pressures. This appears to be more convenient, because the

displacement volume strain matrix does not need to be stored. In addition, this would remove the

error associated with determining the fluid displacement from the displacement potential (uF = Dψ).

Regardless of the method chosen for determining the pressure, caution must be taken, as this is the

dynamic pressure if the fluid structure system is initially in static equilibrium. The total pressure

would be the dynamic pressure plus the hydrostatic pressure.

When the total pressure in the element drops below the vapor pressure, then the stiffness matrix

is modified to suppress rigidity for the element. If the total pressure of the element is found to be

less than the vapor pressure, then the stiffness matrix can be modified by assembling the element

stiffness matrix calculated as before, but negative. However, to reduce computational cost during each

time step, it is best to determine the cavitation effect of each element on the matrix Kfd during the

preprocessing phase. Then, if the element is found to be in a cavitation region, the matrix Kfd can

be modified appropriately.

In addition, if cavitation analysis is to be performed it is best to use the Central Difference

form of the Newmark family methods (β = 0, γ = 12). Using the Central Difference is best for two

reasons. First, for an accurate account of cavitation, a small enough time step is needed to assure that

a pressure wave does not move through an element. Since the CFL condition is already being met, it

is better to use an accurate method for small time steps such as the Central Difference method. The

second advantage is due to the computational cost. Using the transient analysis from Section 3.2, the

106

matrix to be inverted is the following for the fluid:

FF =[

Mfd + ∆tγCfd + ∆t2βKfd

]

−1. (8.7)

This matrix is used several times for the determination of the Interface equation, Eq. (3.7), and during

the transient analysis. Thus, it is originally factored with an LU factorization [75], for computational

efficiency during reuse. Therefore, it would be beneficial if this matrix could be factored only once

and reused. Hence, by modifying the stiffness to account for cavitation, the Central Difference method

allows for the matrix FF to only be factored once, for β = 0.

An algorithm to be used during transient analysis is as follows:

(1) Given the forcing values (fn+1S , fn+1

F ), and the previous state variables of the domains (unS ,

unS , unS , ψn, ψ

n, ψ

n) solve the Interface equation, Eq. (3.7), for the Lagrange multipliers

(λn+1S , λn+1

S ).

BTSFSBS 0 LS

0 BTFDFFDTBF LF

LTS LTF 0

λn+1S

λn+1F

un+1B

=

BTSFSg

n+1S

BTFDFFgn+1

F

0

(2) Insert the Lagrange multipliers (λn+1S , λn+1

S ) into Eq. (3.2) and Eq. (3.3), and solve for un+1S

and ψn+1

un+1S = FS(gn+1

S − BSλn+1S ),

ψn+1

= FF (gn+1F − DTBFλ

n+1F ).

(3) Update the values of un+1S , un+1

S , ψn+1, and ψn+1

, with the Eq. (3.1).

(4) Determine the fluid element pressure, pe by either Eq. (4.2) or Eq. (8.6).

(5) Assuming the analysis begins in static equilibrium, check if the element is in the cavitation

region and adjust the matrix Kfd appropriately.

(6) Repeat for the next time step, given the modified matrix Kfd.

8.2.2 Frothing: Spurious Oscillations

After the cavitation region has formed it is possible that the earthquake direction may change.

This causes the micro bubbles region created by cavitation to collapse and and the cavitation region

closes up again. As a result of the sudden closure of the fluid cavities, an impact on the dam will occur

107

[96]. This impact phenomena results in a sharp increase in compressive pressure followed by spurious

oscillations (frothing) [36]. To eliminate these high frequency oscillations several authors [45, 109, 114]

have added a small amount of numerical damping. Others, [36, 96], have added a small amount of

damping to the water domain in their analysis. The small amount of damping required to remove the

frothing has been noted to have a small effect on the response of the system [96].

Given the fluid formulation for the LLM method, it would be easier to implement a modest

amount of damping into the fluid-domain by creating a fluid-damping matrix. This is also conve-

nient because a fluid damping matrix is created for the inclusion of the silent boundaries. El-Aidi

[36] recommended a 0.1 % stiffness-proportional damping in the fundamental pressure mode of the

fluid element. Oskouei [96] recommends a 5.0 % mass-proportional damping plus a variable stiffness

proportional damping. In this study, a classical damping matrix is proposed by the Rayleigh model:

CFd = αKKFd + αMMFd, (8.8)

where the coefficients are determined by the first nonzero frequency values. A damping ratio of 0%

is used for the Mass term and a value of 0.1% is the beginning point for the Stiffness term. The

Stiffness term’s damping ratio is adjusted as appropriate. The Stiffness proportional damping is used

to remove the high frequency oscillations caused by frothing. The mass-proportional term has been

omitted because it would provide some artificial numerical stability during the time-marching process

[73].

8.2.3 Bleich-Sandler Plate Problem

The Bleich Sander plate problem [14] is used to validate the cavitation method for the LLM

method. This benchmark problem has the only known analytical solution for a cavitation problem.

The study is made up of a semi-infinite column of fluid with a constant section that supports a

mass. Gravity and atmospheric pressure are included for this study. An exponential pressure wave

propagates thru the fluid, see Fig. 8.5. The reflection of this wave on the mass creates cavitation.

The fluid is described by 100 elements with the properties given in Table 8.1. Numerical damping

is necessary to suppress frothing with the coefficient of αK = 10−5 and αM = 0 in Eq. (8.8). The

results provided in the following section demonstrate the accuracy of the LLM method.

8.2.3.1 Numerical Solution

In Fig. 8.6, the numerical solution produced by the LLM method is compared to the analytical

solution of this problem given by Bleich and Sandler [14]. The small differences are due, according to

108

x

D

W

g

Plane-Wave Boundary

Exponential

Pressure Wave

patm

Pinc = P0e

t+(x−x0)/τ

t

H(

t+(x−x0)/τt

)

Figure 8.5: Schematic model of the Bleich-Sandler plate problem.

109ρF 989 kg/m3

c 1451 m/sPatm 0.101 MPag 9.81 m/s2

D 3.81 mW 144.7 kg/m2

P0 0.710 MPaτ 0.996 ms

Table 8.1: Parameters for the Bleich-Sandler plate problem

Sprague and Geers [113], to the use of the plane wave approximation which is not perfect for the whole

range of frequencies, and to the simplicity of the elements as they do not allow perfect propagation.

To solve this problem, Sprague and Geers propose the use of high order elements.

8.2.4 Cavitation in Koyna Dam Analysis

In this section, the reservoir cavitation effects caused by the El Centro earthquake response of

the Koyna dam are investigated. In order to truly consider the effects of cavitation, the El Centro

earthquake response is first amplified so that the maximum magnitude is of the order of 1 g and then

again to 1.5 g in the horizontal direction. The mesh of the system for this analysis is shown in Fig. 8.7,

because of its low C-error in the linear analysis for both the CASE transient analysis (C-error = 0.04)

and the LLM transient analysis (C-error = 0.06). In both methods, frothing is suppressed. The

CASE transient method uses an artificial damping coefficient, β = 0.25 and the LLM method uses a

stiffness proportional damping term with a Rayleigh damping ratio of 0.1%. The LLM method with

the zero-moment rule is used for the non-matching meshes in both transient analyses.

A comparison of the relative displacement of the crest of the dam for the LLM transient method

and the CASE transient method is shown in Fig. 8.8. The cavitation zone is shown for the LLM

transient method in Fig. 8.9. This is has expected with the cavitation being near the surface of the

water and primarily next to the dam’s face, which is on the right side of the figures. In these figures,

frothing has been suppressed. If the minor amount of damping is not added to the fluid domain then

spurious oscillations can be noticed away from the face of the dam as shown in Fig. 8.10.

An interesting observation is that including the effect of cavitation reduces the relative displace-

ment. This observation has been noted by several authors that have examined the effect of cavitation

[36, 27, 96, 49]. Zienkiewicz [130] notes a similar response for the majority of the cases studied in

that article. However, the Koyna dam example does show a spike in the relative displacement after

the major acceleration component of the El Centro earthquake has passed for a second. It is believed

110

Time (s)

Mas

sM

otio

n(m/s

) Bleich & SandlerLLM method

Cavitation Results

Without Cavitation

0

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016

-0.1

-0.2

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

(a) Velocity response for the mass

Time (s)

Xlo

cation

(m)

Bleich & SandlerLLM method

Closure Point

0.002 0.004 0.008 0.01 0.012 0.014

1

2

3

(b) Cavitation zone

Figure 8.6: Bleich-Sandler plate problem results

Figure 8.7: Koyna dam mesh used for cavitation analysis.

0 2 4 6 8−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

Time (s)

Dis

pla

cem

ent

(m)

Cavitation ResultsWithout Cavitation

(a) CASE transient method.

0 1 2 3 4 5 6 7 8−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

Time (s)

Dis

pla

cem

ent

(m)

Cavitation Results

Without Cavitation

(b) LLM transient method.

Figure 8.8: The effects of cavitation on the relative displacement of the dam crest.

111

(a) 1 second (b) 2 seconds

(c) 3 seconds (d) 4 seconds

(e) 5 seconds (f) 6 seconds

(g) 7 seconds (h) 8 seconds

Figure 8.9: Cavitation zone during selective time intervals with the Dam’s face on the right side.

(a) 5 seconds (b) 6 seconds

Figure 8.10: Cavitation zone during selective time intervals with the Dam’s face on the right sidewithout suppressing frothing.

112

that this spike is caused by the spurious oscillations [36], as Zienkiewicz did not try to preclude the

effects of frothing. Zienkiewicz does note that the effect of cavitation is not serious, as found by the

previous authors cited.

Niwa and Clough, [93] performed experimental results with an attempt to include cavitation.

The dam model was separated from the water by a plastic membrane. Since air could access the

dam-membrane interface, separation (i.e. cavitation) occurred whenever the absolute water pressure

was less then the atmospheric pressure. In their results, the authors concluded that in low intensity

accelerations (0.156g) cavitation would have little effect. However, they believe that at high intensity

accelerations (1.21g) cavitation can have a serious effect during the impact action. However, they did

not run tests attempting to suppress the cavitation for comparison. In all fairness, it would be very

difficult to truly model cavitation, and then turning it off would be nearly impossible. In addition, the

scaled model, about 1/150 the actual size, does not attempt to scale the hydrostatic pressure at the

locations of cavitation. Therefore, this study presents cavitation results for the relative displacement

with a maximum acceleration value of 1.5g shown in Fig. 8.11. It is possible that the effect of the

impact is not felt with the acceleration provided in the positive x-direction; therefore, the magnitude

is changed to the other direction and the relative displacement results are shown in Fig. 8.12. In both

cases, including the effect of cavitation actually appeared to reduce the relative displacements. Hence,

as concluded by many [36, 27, 96, 49, 130], the effect of cavitation is minimal and not a necessity for

problems of this nature.

0 2 4 6 8−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

0.5

Time (s)

Dis

pla

cem

ent

(m)

Cavitation ResultsWithout Cavitation

Figure 8.11: The effects of cavitation on the relative displacement of the dam crest with the maximumacceleration equal to 1.5g.

113

0 2 4 6 8−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

0.4

Time (s)

Dis

pla

cem

ent

(m)

Cavitation ResultsWithout Cavitation

Figure 8.12: The effects of cavitation on the relative displacement of the dam crest with the maximumacceleration equal to 1.5g, but with the acceleration in the opposite direction than previously.

114

8.3 Operational Count

Up to this point, it has been mentioned but not demonstrated that there is additional com-

putational cost of using the LLM transient method versus any other partitioned method with the

non-matching meshes handled by a basic relation matrix. In the LLM method the additional cost is

realized during the solution of the interface equation at each time step; whereas, other partitioned

methods would be using matrix multiplication for transferring loads from one subdomain’s mesh to

another.

It is assumed that the interface equation matrix, Eq. 3.7, has been factored with an LU fac-

torization. This upfront cost would be similar to the cost of creating the relation matrices for the

other methods. Also assume that the interface frame is discretized with the same amount of degree

of freedoms as the refined mesh boundary, say the structure ns. The fluid boundary has nf degrees

of freedom. According to Kincaid, [75], the long operations (multiplication) cost of solving a matrix

equation with an LU factorization for m time steps is

cost =

(

1

2+ml

)

(ne)2, (8.9)

where ne is the number of equations. Therefore, the cost for the interface equation is

costllm =

(

1

2+ml

)

(2ns + nf )2, (8.10)

where ml is the number of time steps for the LLM method. The long operations cost for using a

relation matrix during each time step would be

costmatrix = me(2nsnf ), (8.11)

where me is the number of time steps for this other method. Here there are two instances in which

matrix vector multiplication needs to happen, once for the displacement relation and once for the

force relation. If cavitation is not included, then the LLM transient method can use implicit time

integration schemes where the time step is dictated by accuracy. Assuming that other staggered

partitioned integration procedures use an explicit time integration method with me time steps, the

logical question is, when is it better to use the LLM method (i.e. when are there less operations)?

costllm ≤ costmatrix.

ml(4n2s + 4nsnf + n2

f ) +

(

2n2s + 2nsnf +

1

2n2f

)

≤ 2menfns.

ml ≤2menfns

4n2s + 4nsnf + n2

f

− 1

2.

(8.12)

115

As a worse case scenario, it can be assumed that the number of degrees of freedoms of the two

subdomains are the same, then the number of steps for the LLM transient method should be less than2

9the time steps of the explicit staggered partitioned procedure.

ml ≤2

9me − 1. (8.13)

8.4 Summary

The main point of this chapter is to examine the abilities of the LLM transient method and the

LLM method for non-matching meshes with popular concepts currently being explored for computa-

tional modeling. First, reduced ordering modeling (ROM) was demonstrated for the LLM transient

method with non-matching meshes and a basis chosen as the eigenvectors from the subsystems. Here,

it was noted that a substantial reduction could be accomplished. Next, a non-linear effect of cav-

itation was considered, where it was noted that the effect of including cavitation was small and a

conservative approach to dam analysis would not need to include this effect. Nonetheless, including

cavitation had no effect on the ability of the LLM method for non-matching meshes. However, using

the LLM transient method to include cavitation increased the computational cost as compared to the

CASE transient method, because of the need to use the explicit Central Difference time integration

method in the LLM transient code. Thus, a long operational cost analysis was conducted, and it was

determined that if the number of LLM transient time steps were less than2

9the number of time steps

of any partitioned method, then it was beneficial to use the LLM transient method.

Chapter 9

Partitioned 3-D Problem with Curved Surface

The main purpose of this chapter is to evaluate the LLM method for non-matching meshes on

a curved surface and the implementation of the zero-moment rule on such a surface. The effect of

mapping the interface nodes on the curved surface and the creation of the connection matrices are

explored. A benchmark problem that has a curved surface is the Morrow Point arch dam, which has

been extensively studied in the literature [35, 52, 87, 84, 94, 115, 123].

9.1 Problem Description

The problem studied in this chapter is the Morrow Point dam located on the Gunnison River

in southwest Colorado. The river is confined to a narrow channel cut deep in the rock-walled canyon.

The general geometry is shown in Fig. 9.1. The dam has a height of 142m and a crest length of 219m.

A detailed description of the dam is available from the U.S. Department of the Interior [76]. This arch

dam is generally studied in literature, because of its nearly symmetrical design. Thus, the majority of

the studies take advantage of the symmetry and only model half the dam.

Fok et al. [52] was one of the first to demonstrate the significance of including the effect of the

fluid, the rock foundation, and bottom absorption. Many studies have been conducted to determine

the influence of the surrounding rock at the dam [115, 52, 94, 35]. Tan et al. [115] illustrate that

modeling the rock foundation with the same stiffness as the dam has little effect compared to modeling

the rock foundation as a rigid foundation. However, decreasing the stiffness of the rock by one-forth

the amount of the stiffness of the dam appeared to have significant effects. In this study, the rock

foundation is rigid for comparison purposes with the literature, because the effect of the foundation

has provided varying results. This also provides a better evaluation of the interface condition.

Another controversial topic is the application of the ground motion. It is possible to define the

earthquake input as a rigid-body translation of the basement rock on which the finite element model

of the dam and foundation are supported, as was done for the gravity dam analysis in the previous two

117

(a) View of the Morrow Point dam.

142 m

219 m

(b) Morrow Point dam Geometry. [123]

Figure 9.1: Morrow Point dam representation.

118

chapters. However, there is little data about earthquake motion at a depth from the surface, because

most of the available strong motion records are from accelerometers located at the ground surface [52].

In addition, when truncating the rock foundation to a finite extent, some have applied the motion

to the side of the truncation where others have not. Tan et al. [115] assume the foundation rock is

massless for the dynamic analysis, and the earthquake input is specified as spatially-uniform motion of

the basement rock. Since there is no wave propagation mechanism in the massless foundation rock, the

specified base rock motion is transmitted without modification to the dam-foundation rock interface

[52]. Conveniently, this concern is alleviated with the assumption of a rigid foundation. Thus, the

ground motion is uniformly specified at the dam portions in contact with the rock foundation.

The excitation for this problem is the Taft earthquake of July 21, 1952, recorded at the Lincoln

School Tunnel. In this problem, the greatest acceleration data is applied in the downstream direction

and the other two in the vertical and cross stream directions. The downstream acceleration is shown

in Fig. 9.2 along with its frequency content.

0 5 10 15 20−1.5

−1

−0.5

0

0.5

1

1.5

2

Acc

eler

atio

n(m/s

2)

Time (s)

(a) Acceleration

0 5 10 15 20 250

0.5

1

1.5

2

2.5

Frequency (Hz)

Pow

ersp

ectr

um

(b) Frequency content

Figure 9.2: Ground motion and frequency content of the seismic acceleration recorded at the 1952Taft Lincoln School Tunnel, California earthquake.

A model of the dam and the rigid foundation used for this problem is shown in Fig. 9.3. The

parameters for this problem are defined in Table 9.1.

9.2 Creation of the Connection Matrices

One of the difficult aspects of the LLM method is creating the connection matrices, Lk, where k

is one of the subdomains, the structure or the fluid in this study. This is obviously compounded when

the interface surface is not a normal shape. An atypical shape is the curved surface of the arch dam.

119

Edam 2.76x1010 Paν 0.2

ρdam 2478 kg/m3

cfluid 1440 m/sρfluid 1000 kg/m3

Rayleigh damping 5 %

Table 9.1: Parameters for Morrow Point dam model.

Figure 9.3: Morrow Point dam model used for this study. The fluid is not shown and the foundation,shown in brown, is rigid.

120

This section provides a simple method for creating these matrices on such a surface. Recall that the

connection matrices are used to relate the interface displacements to the two subdomains’ interface

displacements.

ΠB = λTS (BTSuS − LSuB) + λTF (BT

FuF − LFuB).

The first task is to create the interface nodes. The simplest case is to use the trace of the

discretization of either the structure or fluid subdomain at the interface. However, the zero-moment

rule assures that a conservative system is created and no matrices will be ill-conditioned; thus, it

is discussed. If the surface is fairly regular and not disjointed, then the subdomains’ nodes can be

mapped to a surface element as shown in Fig. 9.4. With the subdomains’ nodes mapped to an interface

η

ξ

(a) Structure

η

ξ

(b) Fluid Domain

Figure 9.4: Mapping of the subdomains nodes to an interface element for use in the zero-moment rule.

element, then the zero-moment rule can be used to determine the discretization of the interface as

shown in Fig. 9.5. The zero-moment rule is used in one-dimension, along one of the mapped element’s

natural coordinates, to determine the node location along the axis. Then, the process is repeated

along the other mapped natural coordinate axis.

Given the following discretization on the mapped interface element, the connection matrices can

be easily created from simplifying the integrations on this mapped interface element. The simplification

can take place because of the use of Dirac delta shape functions for the Lagrange multipliers. For

instances, the connection matrix for the structure, LS, is created as follows:

(1) The process would be to loop through each created boundary frame element.

(2) The normal vector for each boundary frame element is found by mapping the boundary frame

element coordinates back to the physical coordinates of the system. An alternative is to

determine the normal vector of each structure node and use this value.

121

η

ξ

Figure 9.5: Interface nodes ”x” determined by the zero-moment rule.

122

(3) In each boundary frame element, the location of any structure nodes within the element is

determined.

(4) Given the mapped coordinates of the boundary frame’s nodes and the mapped coordinates of

the structure nodes, the natural coordinates of a structure node within the frame element is

deduced.

(5) Knowing the natural coordinates within a boundary frame element of a structure node, the

shape functions of the boundary frame element provide a relation between the nodes of the

frame element and the structure node.

(6) This relationship in conjunction with the normal vector provide the necessary portion of the

connection matrix for the boundary element nodes and the structure node.

This is possible, because of the use of collocated Lagrange multipliers with Dirac delta shape functions.

If linear shape functions were used instead, then the connection matrices would be difficult to form

given the need for piecewise polynomial integration.

9.3 Non-matching Mesh Results

Non-matching meshes for the fluid domain and the structure domain are used for the Morrow

Point arch dam to test the effect of creating the connection matrices as described in the previous

section. First, matching meshes were used until convergence was obtained with the mesh for the dam

as shown in Fig. 9.6. The relative displacement at the crest of the dam in the center of the arch is

used for comparison purposes. The results of the relative displacements are comparable with published

results. There are some minor differences, but these are attributed to the fact that the dam and the

fluid models are slightly different then published models.

Given the results for the converged model, the fluid is then changed so that there are non-

matching meshes at the interface. Figure 9.7 illustrates four different results, where the number

of fluid nodes at the interface is used to indicate the difference in the non-matching meshes. The

structure has 63 nodes at the interface, and the fluid ranges from 42 to 108 nodes at the interface.

The maximum difference between the matched simulation and the non-matched simulation for the

time span is presented in Fig. 9.7. In addition, the C-error for the non-matched mesh compared to

the matched mesh is provided. As can be seen, there is very little difference between the two results;

with the maximum difference, being noted when the meshes are further apart. It is also interesting

to note that the overall relative magnitude is less for the non-matching meshes with the fluid as the

coarse mesh than when the fluid is more refined than the structure.

123

Figure 9.6: Structure Mesh that showed Convergence.

0 5 10 15−0.06

−0.04

−0.02

0

0.02

0.04

0.06

Time (s)

Dis

pla

cem

ent

(m)

Non-matchingMatching

Maximum Difference = 0.005 m at 6.77 sec.

C-error = 0.079M-error = −0.073P-error = 0.029

(a) 42 Fluid Interface Nodes

0 5 10 15−0.06

−0.04

−0.02

0

0.02

0.04

0.06

Time (s)

Dis

pla

cem

ent

(m)

Non-matchingMatching

Maximum Difference = 0.005 m at 6.75 sec.

C-error = 0.079M-error = −0.076P-error = 0.022

(b) 48 Fluid Interface Nodes

0 5 10 15−0.06

−0.04

−0.02

0

0.02

0.04

0.06

Time (s)

Dis

pla

cem

ent

(m)

Non-matchingMatching

Maximum Difference = 0.004 m at 7.1 sec.

C-error = 0.048M-error = 0.036P-error = 0.031

(c) 88 Fluid Interface Nodes

0 5 10 15−0.06

−0.04

−0.02

0

0.02

0.04

0.06

Time (s)

Dis

pla

cem

ent

(m)

Non-matchingMatching

Maximum Difference = 0.010 m at 7.1 sec.

C-error = 0.083M-error = −0.027P-error = 0.078

(d) 108 Fluid Interface Nodes

Figure 9.7: Relative displacement comparison for non-matching mesh versus matched mesh with 63structure interface nodes.

124

Once again, a cavitation analysis is performed. Here, the acceleration is amplified by a factor of

five to increase the upstream acceleration to be near 1 g. The results are shown in Fig. 9.8 where the

effect of cavitation is minimal, and it tends to reduce the relative displacement.

0 5 10 15−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

Time (s)

Dis

pla

cem

ent

(m)

Cavitation ResultsWithout Cavitation

Figure 9.8: The effects of cavitation on the relative displacement of the dam crest with the maximumacceleration equal to 1 g.

9.4 Summary

This chapter illustrates the ease of implementing the LLM method for non-matching meshes

with a curved surface in a three dimensional analysis of the Morrow Point arch dam. This was

accomplished by mapping the trace of the subdomains’ discretization onto an interface element. Once

the node locations for the subdomains is known on the interface element, then the zero-moment rule

can be used to discretize the interface boundary. Then, the connection matrices Lk can easily be

determined on this mapped interface element. This was verified by comparing non-matching meshes

with a matching mesh converged result. Finally, the chapter concluded with the realization that

cavitation does not greatly effect the relative displacement in this problem.

Chapter 10

Conclusion

This thesis developed an acoustic fluid-structure interaction methodology with Localized La-

grange Multipliers (LLM) used for the coupling of the heterogeneous systems: the emphasis was on

the treatment of non-matching meshes. Beginning with a variational energy formulation, an efficient

partitioned transient analysis was developed, where all the benefits of a partitioned method are re-

alized. This transient analysis is advantageous, because it does not require a prediction step in the

algorithm, which can greatly effect the stability of the any transient method. With the coupling

of the systems performed by LLM, non-matching meshes at the interface are easily managed. The

salient feature of the coupling with LLM is also used to create coupling matrices that can be used in

other acoustic FSI codes. Even with the use of discrete Lagrange multipliers, convergence has been

demonstrated analytically and experimentally for non-matching meshes.

A new fluid formulation was needed in order to exploit the transient analysis that has been de-

veloped. The formulation is carried out by first employing a fluid displacement formulation to assure

an energy form. The displacement potential formulation is then obtained by means of a displacement

potential to displacement transformation. This created a gradient matrix that related fluid domain

displacement potentials to the fluid domain displacements, which was inserted into the derived gov-

erning equations of the FSI system. This formulation was verified with an analytical solution for the

movement of a rigid wall against a fluid domain. This test problem also provided validation of the

silent boundary used in this study and surface waves. Future work could include creating the gradient

matrix analytically with higher order elements.

An analytical piston problem was devised to examine the ability of the new transient method.

This demonstrated that the transient method can be unconditionally stable, if two A-stable integration

schemes are chosen for the fluid and structure models. This test problem was also used with three-

dimension models to illustrate the benefit of using the zero-moment rule to discretize the interface

frame.

126

A study was conducted on the seismic excitation of the Koyna Dam to demonstrate the benefits

of the new developed transient method and the use of the LLM for the coupling of the non-matching

meshes. One of the benefits of the new transient formulation is the ability to conduct a symmetric

vibration analysis to gain a theoretical and physical understanding of the system. This vibration anal-

ysis was used to study kinematic continuity issues with non-matching mesh techniques at the interface

of the fluid-structure system. Rules for implementation of the LLM, Mortar, and Consistent Interpo-

lation coupling methods were developed by observing conditions when continuity was not maintained.

In addition, a theoretical verification of the continuity constraints was demonstrated with a frequency

response analysis.

Following the vibration analysis, a transient analysis was executed with non-matching meshes.

The new LLM transient method was verified and different coupling procedures were explored with the

use of the cavitating acoustic spectral element (CASE) transient method [114]. This illustrated the

need to obey the rules developed for the different coupling methods to assure conservative systems

and to create well conditioned coupling matrices. During the transient analysis, it was observed that

for problems in this study, the structure required a greater refinement than the fluid for convergence.

Therefore, a need for non-matching meshes at the interface is justified. In conjunction with the

vibration analysis and frequency content of the excitation, a simple reduced order formulation is

found effective for the Koyna dam problem even with non-matching meshes.

The Koyna dam test problem was also used to research the ability of the new LLM transient

method for the non-linear effect of cavitation. A method was developed that resembled a crack-

smearing approach. Unfortunately, this concept did not perform as well as the CASE transient method

when evaluating cavitation. A simple operational cost analysis showed that the number of steps in

the new LLM transient analysis needs to be less than 29 the number in any comparative partitioned

method. This can generally be achieved because of the ability to use implicit time integrations and the

lack of a prediction step in the LLM method. Overall, this method is best used for acoustic problems

without cavitation; such as, a fluid in a container. It is possible that including the non-linear effect

of crack propagation in the LLM transient method could be comparative to other methods in that

regard, and this could be a starting point for future research.

Finally, the study concludes with the coupling ability of the LLM method on curved surfaces.

The problem studied is the Morrow Point Arch Dam. A three-dimension analysis is performed with

non-matching discretizations at the interface of the fluid and the dam. Here, it was noted that the

non-matching meshes had very little effect on the converged results. In addition, cavitation is shown

to have minimal effect on the overall response of the dam.

127

10.1 Future Work

Through the creation of a new fluid formulation and some basic assumptions a transient method

with a simple ability to couple non-matching meshes at an interface has been developed for an acoustic

fluid-structure interaction. The next logical step would be to try and create a similar concept for fluid

flow problems. This is obviously more complicated with the need for an Arbitrary Lagrangean Eulerian

formulation for the fluid field. However, the simplicity of the partitioned method without the required

prediction step has great appeal. A good starting point would be problems with low Reynolds numbers

such as the aeroelastic behavior of a bridge deck [18] or blood flowing through an artery [50, 118].

Another possible future extension to this work is incorporating an extensive model of the Morrow

Point Dam that would include the curvature of the river and an extensive portion of the mountains

surrounding the dam. An explosion of work in the 1990’s lead to the use of spectral elements for

the approximation solution of linear elastodynamic equations. This has been shown to provide an

efficient tool for simulating elastic wave propagation in realistic geological structures in two and three-

dimensions [78, 111]. In a method similar to Casadei [20], the structure elements that were believed to

behave nonlinear could be partitioned from the linear elements. In addition, the fluid would naturally

be partitioned from all structure systems. The transient method developed from the LLM concept

could be used for the analysis.

Bibliography

[1] Matthew Allen. Reliability-Based Design Optimization of Multiphysics, Aerospace Systems.PhD thesis, University of Colorado, Boulder, 2005.

[2] S.N. Alturi. Advances in Computer Methods for Partial Differential Equations. AICA, RutgersUniversity, New Brunswick, NJ, 1975. On ’hybrid’ finite-element models in solid mechanics.

[3] K.E. Atkinson. An Introduction to Numerical Analysis. John Wiley and Sons, New York, 2edition, 1989.

[4] F. Baaijens. A fictitious domain/mortar element method for fluid-structure interaction.International Journal for Numerical Methods in Fluids, 35:743–761, 2001.

[5] A.H. Barbat. Seismic soil-structure-fluid interaction analysis. In R.W. Lewis, P. Bettess, andE. Hinton, editors, Numerical Methods in Coupled Problems, pages 353–386, New York, 1984.John Wiley and Sons.

[6] U. Basu and A.K. Chopra. Perfectly matched layers for time-harmonic elastodynamics of un-bounded domains: theory and finite element implementation. Computer Methods in AppliedMechanics and Engineering, 192(11-12):1337–1375, 2003.

[7] U. Basu and A.K. Chopra. Perfectly matched layers for transient elastodynamics of unboundeddomains. International Journal for Numerical Methods in Engineering, 59(8):1039–1074, 2004.

[8] F. Belgacem. The mortar finite element method with lagrange multipliers. NumerischeMathematik, 84:173–197, 1999.

[9] J. P. Berenger. A perfectly matched layer for the absorption of electromagnetic waves. Journalof Computational Physics, pages 185–200, 1994.

[10] P.G. Bergan and M.K. Nygard. Finite elements with increased freedom in choosing shapefunctions. International Journal for Numerical Methods in Engineering, 20(4):643–663, 1984.

[11] A. Bermudez and R. Rodriquez. Finite element computation of the vibration modes of a fluid-solid system. Computer Methods in Applied Mechanics and Engineering, 119:355–370, 1994.

[12] A. Bermudez, R. Rodriquez, P. Gamallo, and M. Nogueiras. Approximation of a structuralacoustic vibration problem by hexahedral finite elements. IMA Journal of Numerical Analysis,2005. accepted for publication.

[13] C. Bernardi, Y. Maday, and AT. Patera. A new nonconforming approach to domain decompo-sition: the mortar element method. Technical report, Universite Pierre at Marie Curie, Paris,France, 1990.

129

[14] H. H. Bleich and I. S. Sandler. Interaction between structures and bilinear fluids. InternationalJournal of Solids and Structures, 6:617–639, 1970.

[15] F. Blom. A monolithical fluid-structure interaction algorithm applied to the piston problem.Computer Methods in Applied Mechanics and Engineering, 167:369–391, 1998.

[16] D. Braess, W. Dahmen, and C. Wieners. A multigrid algorithm for the mortar finite elementmethod. Journal on Numerical Analysis, 37:48–69, 2000. SIAM.

[17] H.A. Brahtz and C.H. Heilbron. Water pressures on dams during earthquakes. Transactions,98, 1933. ASCE.

[18] A. Braun and A. Awruch. Numerical simulation of the wind action on a long-span bridge deck.Journal of the Brazilian Society of Mechanics Science and Engineering, 25(4):352–363, 2003.

[19] Y. Calayir and M. Karaton. Seismic fracture analysis of concrete gravity dams including dam-reservoir interaction. Computers and Structures, 83:1595–1606, 2005.

[20] F. Casadei, E. Gabellini, G. Fotia, F. Maggio, and A. Quarteroni. A mortar spectral/finiteelement method for complex 2d and 3d elastodynamic problems. Computer Methods in AppliedMechanics and Engineering, 191:5119–5148, 2002.

[21] Harn Chen and Robert Taylor. Vibration analysis of fluid-solid systems using a finite elementdisplacement formulation. International Journal for Numerical Methods in Engineering, 29:683–698, 1990.

[22] W.C. Chew and Q.H. Liu. Perfectly matched layers for elastodynamics: A new absorbingboundary condition. Journal of Computational Acoustics, 4:341–359, 1996.

[23] A. K. Chopra. Hydrodynamic pressures on dams during earthquakes. Journal of the EngineeringMechanics Division, 93(EM6), December 1967. ASCE.

[24] A. K. Chopra and P. Chakrabarti. Earthquake analysis of concrete gravity dams including dam-water-foundation rock interaction. Earthquake Engineering and Structural Dynamics, 9:363–383,1981.

[25] Anil K. Chopra. Dynamics of Structures: Theory and Applications to Earthquake Engineering.Prentice Hall, Inc., New Jersey, 1995.

[26] R. Clayton and B. Engquist. Absorbing boundary conditions for acoustic and elastic waveequations. Bulletin of the Seismological Society of America, 67:1529–1540, 1977.

[27] R.W. Clough and C. Chu-Han. Seismic cavitation effects on gravity dams. In R.W. Lewis,P. Bettess, and E. Hinton, editors, Numerical Methods in Coupled Problems, pages 571–598,New York, 1984. John Wiley and Sons.

[28] A. Combescure and V. Faucher. A time and space mortar method for coupling linear modalsubdomains and non-linear subdomains in explicit structural dynamics. Computer Methods inApplied Mechanics and Engineering, 192:509–533, 2003.

[29] R.D. Cook, D.S. Malkus, M.E. Plesha, and R.J. Witt. Concepts and Applications of FiniteElement Analysis. John Wiley and Sons, New York, 4 edition, 2002.

[30] M.A. Crisfield. Re-visiting the contact patch test. International Journal for Numerical Methodsin Engineering, 48:435–449, 2000.

130

[31] I.G. Currie. Fundamental Mechanics of Fluids. Marcel Dekker, Inc., New York, 3 edition, 2003.

[32] C. Dohrmann, S. Key, and M. Heinstein. A method for connecting dissimilar finite ele-ment meshes in two dimensions. International Journal for Numerical Methods in Engineering,48(5):655–678, June 2000.

[33] J. Dongarra, J. Bunch, C. Moler, and G. Stewart. Linpack user’s guide. SIAM Publications,1979. Philadelphia.

[34] J. Dow. Finite Element Method and Error Analysis Procedures. Academic Press, London, 1999.

[35] Z. Duron and J. Hall. Experimental and finite element studies of the forced vibration responseof morrow point dam. Earthquake Engineering and Structural Dynamics, 16:1021–1039, 1988.

[36] Bahaa El-Aidi. Nonlinear Earthquake Response of Concrete Gravity Dam Systems. PhD thesis,California Institute of Technology, Pasadena, 1989.

[37] C. Farhat and M. Lesoinne. On the accuracy, stability, and performance of the solution ofthree-dimensional nonlinear transient aeroelastic problems by partitioned procedures. In AIAAStructures, Structural Dynamics, and Materials Conference and Exhibit, 37th, pages 629–641.AIAA, 1996. AIAA Meeting Papers, Paper 96-1388.

[38] C. Farhat, M. Lesoinne, and P. LeTallec. Load and motion transfer algorithms for fluid/structureinteraction problems with non-matching discrete interfaces: Momentum and energy conservation,optimal discretization and application to aeroelasticity. Compute Methods in Applied Mechanicsand Engineering, 157:95–114, 1998.

[39] C. Farhat, M. Lesoinne, and N. Maman. Mixed explicit/implicit time integration of coupledaeroelastic problems: three-field formulation, geometric conservation and distributed solution.International Journal for Numerical Methods in Fluids, 21:807–835, 1995.

[40] C. Farhat, S. Piperno, and B. Larrouturo. Partitioned procedures for the transient solution ofcoupled aeroelastic problems; part i: model problem, theory and two-dimensional application.Computer Methods in Applied Mechanics and Engineering, 124:79–112, 1995.

[41] V. Faucher and A. Combescure. A time and space mortar method for coupling linear modalsubdomains and non-linear subdomains in explicit structural dynamics. Compute Methods inApplied Mechanics and Engineering, 192:509–533, 2003.

[42] C.A. Felippa. A family of early-time approximations for fluid-structure interaction. Journal ofApplied Mechanics, 47(4):703–708, December 1980.

[43] C.A. Felippa. Parameterized multifield variational principles in elasticity: Ii. hybrid functionalsand the free formulation. Communications in Applied Numerical Methods, 5:89–98, 1989.

[44] C.A. Felippa. A survey of parameterized variational principles and applications to computationalmechanics. Computer Methods in Applied Mechanics and Engineering, 113:109–139, 1994.

[45] C.A. Felippa and J.A. Deruntz. Finite element analysis of shock-induced hull cavitiation.Computer Methods in Applied Mechanics and Engineering, 44:297–337, 1984.

[46] C.A. Felippa and R. Ohayon. Mixed variational formulation of finite element analysis of acous-toelastic/slosh fluid-structure interaction. Journal of Fluids and Structures, 4:35–57, 1990.

[47] C.A. Felippa, K.C. Park, and C. Farhat. Partitioned analysis of coupled mechanical systems.Comput. Methods Appl. Mech. Engrg., 190:3247–3270, 2001.

131

[48] C.A. Felippa, K.C. Park, and G. Rebel. A simple algorithm for localized constructioon ofnonmatching structural interfaces. International Journal for Numerical Methods in Engineering,53:2117–2142, 2002.

[49] G. Fenves and L. Vargas-Loli. Nonlinear dynamic analysis of fluid-structure systems. Journal ofEngineering Mechanics, 114(2), 1988.

[50] M. Fernandez and M. Moubachir. A newton method using exact jacobians for solving fluid-structure coupling. Computers and Structures, 83:127–142, 2005.

[51] J.H. Ferziger and M. Peric. Computational Methods for Fluid Dynamics. Springer, London,1996.

[52] K. Fok and A. Chopra. Earthquake analysis of arch dams including dam-water interaction,reservoir boundary absorption and foundation flexibility. Earthquake engineering and structuraldynamics, 14:155–184, 1986.

[53] G. Forsythe and C Moler. Computer Solution of Linear Algebraic Systems. Prentice-Hall,Englewood Cliffs, NJ, 1967.

[54] Y.C. Fung and Pin Tong. Classical and Computational Solid Mechanics. World Scientific,London, 2001.

[55] K. Garatani and H. Matsumoto. Large-scale parallel wave propagation analysis.http://geofem.tokyo.rist.or.jp.

[56] Thomas L. Geers and B.A. Lewis. Doubly asymptotic approximations for transient elastody-namics. International Journal of Solids and Structures, 34(11):1293–1305, 1997.

[57] Thomas L. Geers and Quan Qi. Evaluation of the perfectly matched layer for computationalacoustics. Journal of Computational Physics, 139:166–183, 1998.

[58] T.L. Geers. Transient response analysis of submerged structures. In T. Belytschko and T.L.Geers, editors, Finite Element Analysis of Transient Nonlinear Structural Behavior, volume 26,pages 83–94. ASME, New York, 1977.

[59] T.L. Geers. Doubly asymptotic approximations for transient motion of submerged structures.Journal of the Acoustical Society of America, 64:1500–1508, 1978.

[60] G. Golub and C. Van Loan. Matrix Computations. John Hopkins Press, Baltimore, 1983.

[61] K.F. Graff. Wave Motion in Elastic Solids. Oxford University Press, New York, 1975.

[62] U.A. Gumaste, K.C. Park, and K.F. Alvin. A family of implicit partitioned time integrationalgorithms for parallel analysis of heterogeneous structural systems. Computational Mechanics,24:463–475, 2000.

[63] F.H. Hamdan and P.J. Dowling. Fluid-structure interation: Application to structure in an acous-tic fluid medium, part 1: An introduction to numerical treatment. Engineering Computations,12:749–758, 1995.

[64] M.A. Hamdi and Y. Ousset. A displacement method for the analysis of vibrations of coupledfluid-structure systems. International Journal for Numerical Methods in Engineering, 13:139–150, 1978.

132

[65] P. Hansbo and J. Hermansson. Nitsche’s method for coupling non-matching meshes in fluid-structure vibration problems. Computational Mechanics, 32:134–139, 2003.

[66] P. Hansbo, C. Lovadina, I. Perugia, and G. Sangalli. A lagrange multiplier method for thefinite element solution of elliptical interface problems using non-matching meshes. NumerischeMathematik, 100:91–115, 2005.

[67] D.W. Harris, N. Snorteland, T. Dolen, and F. Travers. Shaking table 2-d models of a concretegravity dam. Earthquake Engineering and Structural Dynamics, 29:769–787, 2000.

[68] G.C. Hart and K. Wong. Structural Dynamics for Structural Engineers. John Wiley and Sons,Inc., New York, 2000.

[69] M. Heath and X. Jiao. Common-refinement based data transfer between nonmatching meshes inmultiphysics simulations. International Journal for Numerical Methods in Engineering, 61:2402–2427, 2004.

[70] B. Herry, L. Di Valentin, and A. Combescure. An approach to the connection between subdo-mains with non-matching meshes for transient mechanical analysis. International Journal forNumerical Methods in Engineering, 55:973–1003, 2002.

[71] Thomas J.R. Hughes. The Finite Element Method: Linear Static and Dynamic Finite ElementAnalysis. Dover Publications, Inc., New York, 2000.

[72] Lars Aaes Jakobsen. A finite element approach to analysis and sensitivity analysis of timedependent fluid-structure interaction systems. PhD thesis, Aalborg University, Denmark, 2002.

[73] N. Khalili, M. Yazdchi, and S. Valliappan. Non-linear seismic behavior of concrete gravity damsusing coupled finite element-boundary element technique. International Journal for NumericalMethods in Engineering, 44:101–130, 1999.

[74] Y. Kim and C Yun. A spurious free four-node displacement-based fluid element for fluid-structureinteraction analysis. Engineering Structures, 19(8):665–678, 1997.

[75] D. Kincaid and W. Cheney. Numerical Analysis: Mathematics of Scientific Computing. BrooksCole., United States, 2002.

[76] D. King. Hydraulic model studies for morrow point dam. Technical Report 37, United StatesDepartment of the Interior, Bureau of Reclamation, 1967. A water resources technical publica-tion.

[77] D. Komatitsch, C. Barnes, and J. Tromp. Wave propagation near a fluid-solid interface: aspectral-element approach. Geophysics, 65(2):623–631, 2000.

[78] D. Komatitsch, J. Vilotte, R. Vai, J. Castillo-Covarrubias, and F. Sanchez-Sesma. The spec-tral element method for elastic wave equations - application to 2-d and 3-d seismic problems.International Journal for Numerical Methods in Engineering, 45:1139–1164, 1999.

[79] Dimitri Komatitsch and Jeroen Tromp. A perfectly matched layer absorbing boundary conditionfor the second-order seismic wave equation. Geophysical Journal International, 154:146–153,2003.

[80] Erwin Kreyszig. Advanced Engineering Mathematics. John Wiley and Sons, Inc., New York,eighth edition, 1999.

133

[81] Q.H. Liu. Perfectly matched layers for elastic waves in cylindrical and spherical coordinates.Journal for Acoustical Society of America, 105(4):2075–2084, 1999.

[82] Q.H. Liu and J. tao. The perfectly matched layer for acoustic waves in absorptive media. Journalfor Acoustical Society of America, 102(4):2072–2082, 1997.

[83] John Lysmer and R.L. Kuhlemeyer. Finite dynamic model for infinite media. Journal of theEngineering Mechanics Division, 95(EM4):859–876, August 1969. Proc. ASCE.

[84] O. Maeso, J. Aznrez, and J. Dominguez. Effects of space distribution of excitation on seismicresponse of arch dams. Journal of Engineering Mechanics, pages 759–768, July 2002.

[85] K. Makinen. Cavitation models for structures excited by a plane shock wave. Journal of Fluidsand Structures, 12:85–101, 1998.

[86] C. Michler, S.J. Hulshoff, E.H. van Brummelen, and R. de Borst. A monolithic approach tofluid-structure interaction. Computers and Fluids, 33:839–848, 2004.

[87] H. Mirzabozorg and M. Ghaemian. Non-linear behavior of mass concrete in three-dimensionalproblems using a smeared crack approach. Earthquake Engineering and Structural Dynamics,34:247–269, 2005.

[88] Y. Miyazaki and K.C. Park. A formulation of conservinc impact system based on localized la-grange multipliers. International Journal for Numerical Methods in Engineering, 2006. acceptedfor publication.

[89] R.E. Newton. Effects of cavitation on underwater shock loading- axisymmetric geometry. Tech-nical Report NPS-69-78-017PR, Naval Postgraduate School, Monterey, CA, 1978.

[90] R.E. Newton. Effects of cavitation on underwater shock loading- part 1. Technical ReportNPS-69-78-013, Naval Postgraduate School, Monterey, CA, 1978.

[91] R.E. Newton. Finite element analysis of shock-induced cavitation. ASCE, Spring Convention,1980. Preprint 80-110.

[92] R.E. Newton. Effects of cavitation on underwater shock loading- plane problem. TechnicalReport NPS-69-81-001, Naval Postgraduate School, Monterey, CA, 1981.

[93] A. Niwa and R. Clough. Non-linear seismic response of arch dams. Eathquake Engineering andStructural Dynamics, 10:267–281, 1982.

[94] C. Noble. Seismic analysis of morrow point dam. Technical Report UCRL-JC-147993, LawrenceLivermore National Laboratory, April 2002.

[95] R. Ohayon. Reduced models for fluid-structure interaction problems. International Journal forNumerical Methods in Engineering, 60:139–152, 2004.

[96] A.V. Oskouei and A.A. Dumanoglu. Nonlinear dynamic response of concrete gravity dams:cavitation effect. Soil Dynamics and Eathquake Engineering, 21:99–112, 2001.

[97] K.C. Park and C.A. Felippa. A variational framework for solution method developments instructural mechanics. Journal of Applied Mechanics, 65:242–249, 1998.

[98] K.C. Park and C.A. Felippa. A variational principle for the formulation of partitioned structuralsystems. International Journal for Numerical Methods in Engineering, 47:395–418, 2000.

134

[99] K.C. Park and C.A. Felippa. Reduced-order partitioned modeling of coupled systems: formu-lation and computational algorithms. In NATO Advanced Research Workshop, Bled, Slovenia,pages 267–289. Faculty of Civil and Geodetic Engineering of University of Ljubljana, Solvenia,2004.

[100] K.C. Park, C.A. Felippa, and U.A. Gumaste. A localized version of the method of lagrangemultipliers and its applications. Computational Mechanics, 24:476–490, 2000.

[101] K.C. Park, C.A. Felippa, and R. Ohayon. Partitioned formulation of internal fluid-structureinteraction problems by localized lagrange multipliers. Comput. Methods Appl. Mech. Engrg.,190:2989–3007, 2001.

[102] K.C. Park, C.A. Felippa, and G. Rebel. Interfacing nonmatching fem meshes: the zero mo-ment rule. In W.A. Wall, K.U. Bletzinger, and K. Schweizer, editors, Trends in ComputationalStructural Mechanics, Barcelona, Spain, 2001. CIMNE.

[103] K.C. Park, C.A. Felippa, and G. Rebel. A contact formulation based on localized lagrangemultipliers: formulation and application to two-dimensional problems. International Journal forNumerical Methods in Engineering, 54:263–297, 2002.

[104] K.C. Park, M.R. Justino Jr., and C.A. Felippa. An algebraically partitioned feti method forparallel structural analysis: algoritm description. International Journal for Numerical Methodsin Engineering, 40:2717–2737, 1997.

[105] D.K. Paul, O.C. Zienkiewicz, and E. Hinton. Transient dynamic analysis of reservoir-dam in-teraction using staggered solution schemes. In Numerical Methods for Coupled Problems, pages321–334, Swansea U.K., 1981. Pineridge Press.

[106] S. Piperno. Explicit/implicit fluid/structure staggered procedures with a structural predictorand fluid subcycling for 2d inviscid aeroelastic simulations. International Journal for NumericalMethods in Fluids, 25:1207–1226, 1997.

[107] S. Piperno and C. Farhat. Partitioned procedures for the transient solution of coupled aeroelasticproblems-part ii energy transfer analysis and three-dimensional applications. Computer Methodsin Applied Mechanics and Engineering, 190:3147–3170, 2001.

[108] G. Rebel, K.C. Park, and C.A. Felippa. A contact-impact formulation based on localized lagrangemultipliers. Technical Report CU-CAS-00-18, University of Colorado, Boulder, CO, July 2000.

[109] G. Sandberg. A new finite element formulation of shock-induced hull cavitation. ComputerMethods in Applied Mechanics and Engineering, 120:33–44, 1995.

[110] G. Sandberg. A new strategy for solving fluid-structure problems. International Journal forNumerical Methods in Engineering, 38:357–370, 1995.

[111] G. Seriani and E. Priolo. Spectral element method for acoustic wave simulation in heterogeneousmedia. Finite Elements in Analysis and Design, 16:337–348, 1994.

[112] P. Seshaiyer and M. Suri. Uniform hp convergence results for the mortar finite element method.Mathematics of Computers, 69(230):521–546, 1999.

[113] M.A. Sprague and T.L. Geers. Spectral elements and field separation for an acoustic fluid subjectto cavitation. Journal of Computational Physics, 184:149–162, 2003.

135

[114] M.A. Sprague and T.L. Geers. A spectral-element method for modeling cavitation in tran-sient fluid-structure interaction. International Journal for Numerical Methods in Engineering,60:2467–2499, 2004.

[115] H. Tan and A. Chopra. Dam-foundation rock interaction effects in earthquake response of archdams. Journal of Structural Engineering, pages 528–538, May 1996.

[116] P. Tong. New displacement finite element method for solid continua. International Journal forNumerical Methods in Engineering, 2:73–83, 1970.

[117] A.H. van Zuijlen and H. Bijl. Implicit and explicit higher order time integration schemes forstructural dynamics and fluid-structure interaction computations. Computers and Structures,83:93–105, 2005.

[118] S. Canic and A. Mikelic. Effective equations modeling the flow of a viscous incompressible fluidthrough a long elastic tube arising in the study of blood flow through small arteries. Journal ofApplied Dynamical Systems, 2(3):431–463, 2003. SIAM.

[119] H.V. Vu and R.S. Esfandiari. Dynamic Systems: Modeling and Analysis. Irwin McGraw-Hill,New York, 1997.

[120] T. Walsh, G. Reese, K Pierson, H. Sumali, J. Dohner, and D. Day. Computational and exper-imental techniques for coupled acoustic/structure interactions. Technical Report SAND2003-4381, Sandia National Laboratories, January 2004.

[121] J. Wandinger. A symmetric craig-bampton method of coupled fluid-structure systems.Engineering Computations, 15(4):450–461, 1998.

[122] X. Wang and K. Bathe. Displacement/pressure based mixed finite element formulations foracoustic fluid-structure interaction problems. International Journal for Numerical Methods inEngineering, 40:2001–2017, 1997.

[123] B. Weber. Rational Transmitting Boundaries for Time-Domain Analysis of Dam-ReservoirInteraction. PhD thesis, Institut fur Baustatik und Konstruktion, Zurich, 1994.

[124] E.L. Wilson and M. Khalvati. Finite elements for the dynamic analysis of fluid-solid systems.International Journal for Numerical Methods in Engineering, 19:1657–1668, 1983.

[125] B. Wohlmuth. Hierarchical a posteriori error estimators for mortar finite element methods.Journal of Numerical Analysis, 36(5):1636–1658, 1999. SIAM.

[126] B. Wohlmuth. A mortar finite element method using dual spaces for the lagrange multipliers.Journal of Numerical Analysis, 38(3):989–1012, 2000. SIAM.

[127] Xue Yue. Development of joint elements and solution algorithms for dynamic analysis of jointedstructures. PhD thesis, University of Colorado at Boulder, 2002.

[128] Y.Q. Zeng, J.Q. He, and Q.H. Liu. The application of the perfectly matched layer in numericalmodeling of wave propagation in poroelastic media. Geophysics, 66(4):1258–1266, 2001.

[129] C. Zhao, T.P. Xu, and S Valliappan. Seismic response of concrete gravity dams including water-dam-sediment-foundation interaction. Computers and Structures, 54(4):705–715, 1995.

[130] O.C. Zienkiewicz, D.K. Paul, and E. Hinton. Cavitation in fluid-structure response (with par-ticular reference to dams under earthquake loading). Earthquake Engineering and StructuralDynamics, 11:463–481, 1983.

136

[131] O.C. Zienkiewicz and R.L. Taylor. The Finite Element Method: Basic Formulation and LinearProblems, volume 1. McGraw-Hill Book Company, London, fourth edition, 1989.

[132] O.C. Zienkiewicz and R.L. Taylor. The Finite Element Method: Solid and Fluid MechanicsDynamics and Non-linearity, volume 2. McGraw-Hill Book Company, London, fourth edition,1989.

Appendix A

CASE/CAFE Method

This appendix is used to provide a brief description of the Cavitating Acoustic Spectral Elements

(CASE) method [114], which is an improvement to the Cavitating Acoustic Finite Element method

[45]. The primary differences between this method and the LLM method is the modeling of the fluid,

modeling the silent boundary, and the time marching process. The modeling of the structure is the

same as in this study with traditional displacement based finite elements.

A.1 Fluid Equations

In solving the fluid equations the CASE method begins by approaching the problem as done

by Newton for treating a cavitating acoustic fluid. This necessitates the use of the displacement

potential, which is a scalar quantity, as the primary variable in the formulation of the finite element

matrix equations. However, instead of the traditional finite element approach the fluid volume is

discretized with spectral elements as performed by Sprague and Geers [114]. Finally, the CASE

method incorporates the fluid code with the structure code and the silent boundary in a time-marching

staggered solution procedure as performed by Felippa and Deruntz [45]. However, the use of the

spectral elements greatly speeds up the process for the mass like matrix, the capacitance, is diagonal.

In evaluating this current problem, some basic assumptions have been made about the fluid.

First, the fluid is inviscid and irrotational; though, the fluid is modeled as being compressible. In

addition, the displacements are small; thus, the density is constant. The fluid is modeled with a

bilinear constitutive relation to account for cavitation. Therefore, the fluid cannot transmit negative

pressure and cavitation is treated as a macroscopically homogeneous phenomenon.

A.1.1 Fluid Governing Equations

Assuming the fluid is modeled in a three-dimensional domain. Therefore, we begin by examining

an element of the fluid with an element volume (dV ). The mass of the element is ρ dV where ρ is the

138

fluid density. The force acting on the element is −∇pdV , where p is the total pressure. The fluid is

under a body force field bg (i.e. gravity), which is assumed to be the gradient of a time independent

potential (∇Bg). Appling Newton’s second law:

(∇Bg)dV − (∇p)dV = (ρd)dV

leads to the momentum equation:

∇Bg − ∇p = ρd (A.1)

where d is a vector of the total fluid displacement in the Cartesian directions, and a superposed dot

denotes temporal differentiation. Due to the assumption of fluid irrotational motion, the displacement

field can be expressed in terms of the gradient of a scalar function.

∇ψ = −ρd (A.2)

where ψ = ψ(d, t) is the displacement potential, and the factor ρ is introduced for notational conve-

nience. By inserting Eq. (A.2) into Eq. (A.1) and spatial integrating we arrive at the following

−ψ = Bg − p+ C

where C is a constant of integration. The spatial integration is the portion that makes it difficult

to form an energy functional with the governing equations that would be used in the LLM method.

At static equilibrium ψ = 0 and the fluid pressure is equal to the hydrostatic pressure (ph). If the

body force is considered only to be caused by gravity, then it is equal to the hydrostatic pressure (ph).

Therefore, the constant of integration (C) is equal to zero. This leads to the equation of motion of

the fluid in terms of the displacement potential.

ψ = p− ph (A.3)

The constitutive equation for a linear acoustic fluid is

p− ph = −K∇d, (A.4)

where K is the fluid bulk modulus. It is a property that characterizes compressibility. An increase in

the pressure leads to a decrease in the volume. The fluid bulk modulus (K) can also be represented

by the speed of sound (c) and the fluid density for the uncavitated acoustic fluid as

K = c2ρ. (A.5)

The densified relative condensation is defined as

s = −ρ∇d, (A.6)

139

then inserting Eq. (A.5) and Eq. (A.6) into Eq. (A.4), the constitutive equation is obtained in terms

of the densified relative condensation.

p− ph = c2s (A.7)

Comparing the constitutive equation, Eq. (A.7), with the equation of motion, Eq. (A.3), the following

relationship is evident.

ψ = c2s (A.8)

By applying the wave equation, ψ = c2∇2ψ, to the equation of motion, Eq. (A.3), and then inserting

this into the constitutive equation, Eq. (A.7), the govern equation for fluid system is obtained.

s−∇2ψ = 0 (A.9)

Finally, cavitation is included in the fluid equations. Cavitation is the spontaneous vaporization

of a fluid. It happens when the fluid pressure drops below the vapor pressure of the fluid. A simple

yet effective mathematical model to represent cavitation is to assume that the cavitating region is

macroscopically homogeneous and at zero total pressure [45]. The basic premise is that the fluid

cannot support negative pressures. Therefore, the constitutive equation of the fluid is written as

p =

ph + c2s, if s > −ph/c2

0, otherwise.(A.10)

The equation of motion becomes

ψ =

p− ph, if s > −ph/c2,−ph, otherwise.

A.1.2 Spectral Elements

In the following, we derive the discrete spectral-elements for the fluid volume as done by Sprague

and Geers [114]. The concept incorporates a subparametric discretization, where first-order (trilinear)

basis functions are used for geometry representation and higher-order basis functions are used for

field-variable representations. The volume occupied by the fluid is separated into hexagonal elements

defined by eight corner points. The discretized fluid volume is expressed as

X = NTX, Y = NTY, Z = NTZ,

where X, Y and Z are column vectors of element-corner-point locations in global coordinates, N is a

column vector of standard trilinear shape functions, and a superscript T denotes transposition.

140

The dependent field variables are derived in the Section A.1.1, and they are the displacement po-

tential (ψ) and the densified relative condensation (s). These dependent field variables are interpolated

within each element as

s(ξ, η, ζ, t) = φT (ξ, η, ζ)se(t),

ψ(ξ, η, ζ, t) = φT (ξ, η, ζ)ψe(t),(A.11)

where se and ψe are the column vectors of the node values of ψ and s, respectively. φ is a column

vector of 1-D, N th-order-polynomial basis functions φi(ξ)φi(η)φi(ζ); ξ, η, and ζ are element natural

coordinates. The essence of the spectral-element method lies in the choice of the field variable shape

functions (φ) and the associated quadrature rule [114]. In this study the choice of φ is defined by

Lagrangian interpolants given by

φi(ξ) = − (1 − ξ2)P ′

N (ξ)

N(N + 1)PN (ξi)(ξ − ξi),

where PN is the Legendre polynomial of degree N , the prime denotes differentiation with respect

to the argument, and ξi is the ith Gauss-Lobatto-Legendre (GLL) quadrature point defined by the

corresponding root of

(1 − ξ2)P ′

N (ξ) = 0. (A.12)

Element-node locations are coincident with the quadrature points, which are located at the zeros of

Eq. (A.12).

The governing equation, Eq. (A.9), is discretized with a standard Galerkin approach:∫

Ωe

φ(s−∇2ψ)dΩ = 0.

Application of the divergence theorem yields∫

Ωe

φsdΩ +

Ωe

∇φ · ∇ψdΩ =

Γe

φ∇ψ · ~ndΓ, (A.13)

where Ωe is the element domain, Γe is its surface, and ~n is the outward-normal vector to Γe. By

inserting the dependent field variables, Eq. (A.11), on the left side of Eq. (A.13) yields the element-

level algebraic equations

Qese + Heψe = be.

In this equation the capacitance (Q) and the reactance (H) are symmetric square matrices given

by

Qe =

Ωe

φφTdΩ, He =

Ωe

(∇φ)(∇φ)T dΩ, (A.14)

and the boundary-interaction vector (b) is defined by

be =

Γe

φ∇ψ · ~ndΓ. (A.15)

141

Note that in be, ∇φ is maintained in its continuum form because it is provided by the displacements

at the non-reflecting and structure boundaries. These element-level systems may be assembled into a

global system

Qs + Hψ = b. (A.16)

The element level capacitance and reactance are determined by evaluating the integrals in the

matrix Eq. (A.14) with GLL quadrature [114]. One of the advantages of using spectral elements is

that the capacitance matrix is diagonal, which greatly speeds up the solving of the linear system in the

explicit time integration scheme used in this method. The capacitance matrix is diagonal, because the

nodes and the quadrature points are coincidental. The boundary-interaction vector (b) is determined

by the displacements of the structure and the silent boundary of the fluid.

A.1.3 Silent Boundary of the Fluid

In order to have a finite domain for the fluid volume a silent boundary is implemented at the

water’s truncation edge. The PWA is a boundary fluid element, where the nodes of the element

coincide with the nodes of the fluid elements at the finite extent of the fluid. Thus, the PWA elements

lie on the brick faces of the fluid at the finite extent. The displacement of the PWA is given by

usb =1

ρcpsb,

where usb and psb are column vectors of the silent boundary displacements and the silent boundary

pressures, respectively. The circle above the pressure term denotes temporal integration.

A.1.4 Explicit Time Integration

The goal of the time integration of the fluid equations is to ultimately determine the pressure

that is applied to the structure and the silent boundary at each time step. This is accomplished by

solving for the relative condensation (s) in Eq. (A.16) at the particular time step, then to input this

time step value of the relative condensation (s) into the equation for the pressure in a bilinear fluid,

Eq. (A.10).

First, an artificial damping term is introduced into the fluid equation of motion, Eq. (A.3), to

reduce the phenomenon known as frothing.

ψ = p− ph + βtc2s.

The artificial damping term is βtc2s, in which β is a dimensionless damping coefficient that varies

from 0 to 1, and t is the fluid time step (tn+1 − tn). The value of s is estimated by a backward

142

difference formula as

s = (sn − sn−1)1

t .

The fluid displacement potential can be updated as follows:

ψn+1 = ψn + t(pn − ph + βtc2s),ψn+1 = ψn + t(ψn+1).

Inserting the above updated fluid displacement into Eq. (A.16) yields

Qsn+1 = bn+1 − Hψn+1,

where the linear system can be solved for the next relative condensation (sn+1). In solving the linear

system the advantage of the capacitance matrix (Q) being diagonal is utilized, and the boundary inter-

action vector (b) is determined from predicted structure displacements and predicted silent boundary

displacements. This updated value of the relative condensation (sn+1) can then be used to update the

fluid pressure with a test for cavitation as follows

pn+1 = max

ph + c2sn+1,0

.

A.1.5 Stability of Fluid Time Integration

In using trilinear shape functions for the geometry of the fluid elements, Felippa and Deruntz

[45] performed a Fourier Stability analysis and discovered and upper bound for the critical time as

tcr ≤2

c√

λmax(1 + 2β),

where β is the artificial-damping constant, and λmax is the maximum eigenvalue of the global gener-

alized eigenvalue problem

(H − λQ)z = 0,

where z is the eigenvector associated with the eigenvalue λ. Sprague and Geers [114] discovered that it

is more efficient to use Gerschgorin’s theorem to obtain an upper bound for the λmax of each element.

The greatest Gerschgorin λmax of the elements is then used to determine an upper bound for the time

increment of the fluid.

A.2 Structure, Fluid, and the Silent Boundary Coupling

With the dam and the soil modeled as a monolithic structure, there are now three semi-discrete

systems that interact along their interfaces: structure, fluid-volume, and the fluid silent boundary.

143

Displacements are outputted from the structure model and the fluid silent boundary to the fluid-

volume mesh, from which pressures are sent back.

The displacements of the structure and the fluid silent boundary are incorporated into the fluid

equations through the boundary-interaction vector (A.15), which becomes

be =

Γst

φxedΓ +

Γsb

φφTusbdΓ

where xe is the average normal displacement (positive going into the fluid) at the center of the wet-

structure element, and usb is the silent boundary displacements at the coinciding nodes.

The fluid pressures can be numerically integrated over the wet-structure element and applied to

the structure nodes that are on the wet-structure face. This forcing occurs as a pressure normal to

the dam face and a normal to the soil that is in contact with the water. The forcing created by the

wet-structure element is evaluated as

f ei = γei

Γwet

NpedΓ, iǫX,Y,Z,

where N is a column vector of standard bilinear shape functions, pe is the average pressure over the

element face, and γei is the cosine of the angle between the structure-element normal and the ith global

Cartesian direction. This element force can be proportioned to the wet-structure element nodes.

Finally, the last interaction is the silent boundary. As previously mentioned the fluid volume

sends pressure to the silent boundary. As discussed in Section A.1.3, the silent boundary then displaces

due to these pressures.

A.3 Staggered Integration

A staggered solution procedure is used to numerically integrate the structure, the fluid, and the

silent boundary. In the solution procedure, only displacements and pressures are passed back and forth

between the different modules. The process begins by assuming that all values are know at time n.

Then the fluid-volume values at time n+ 1 are determined by predicting the structure displacements

and the silent boundary displacements with a Euler scheme (i.e. upred = un +t(un)). This provides

the pressure at n+ 1. Given the fluid values, the structure and the silent boundary displacements are

updated to the n+ 1 time. Then the process is repeated.

Appendix B

Mortar Method and the Consistent Interpolation Based Method

In this appendix a brief moment is taken to discuss two of the more popular methods currently

used for non-matching meshes. These are the Mortar method and the Consistent interpolation method.

The relationship between the LLM method and these methods is considered in this appendix because

they are evaluated throughout this study. Here, it is discovered that there is a very strong relationship

among the Consistent Interpolation method, the Mortar method, and the LLM methods. In fact,

if the Lagrange multiplier’s shape functions use Dirac delta functions as the interpolation and the

interface space is discretized the same as the coarse boundary space, then the Consistency Interpolation

method and the LLM method will yield an equivalent relationship between the structure’s boundary

displacements and the fluid’s boundary displacements. The Mortar method will also be equal to the

Consistency Interpolation method, if the interface frame is discretized as the refined mesh and the

Consistent Interpolation method is interpolating the refined mesh on the coarse mesh. The LLM

method and the Mortar method both begin with an insertion of interface frame. However, the Mortar

method associates the Lagrange multipliers with this frame, whereas, the LLM method uses the frame

for a new variable (the boundary displacement) and associates domain specific Lagrange multipliers

with each domains boundary. Given the formulation discussed in this study, the LLM method with

the zero-moment rule is the only one of the three methods that can pass the interface patch test.

In addition, the formulation of the LLM method in terms of virtual work, produced a high fidelity,

efficient transient analysis.

B.1 Mortar Method

The discretization of the Lagrange multiplier boundary appears to be the main difference between

many Lagrange multiplier methods and variations of the methods. One of the most popular methods

is the method termed the ”mortar method” introduced by Bernardi [13]. As can be imagined, many

variations of this method have arisen. However, there appears to be one main common thread among

145

all mortar methods, which is that the Lagrange multipliers are located with the boundary frame and

share the same discretization as the boundary frame. During this study, this distinction will be referred

to as the Mortar method. This is a clear difference between the majority of the Lagrange multiplier

methods [20, 41] and the LLM method. Schur dual formulation can be used to more easily handle

integrating products of piecewise polynomials on unrelated meshes [70], see Section 2.3.4.

The Mortar method introduces a layer of mortar elements between two adjacent distinct domains

with a common boundary [72]. Applied to a FSI, coupling between the fluid and solid is achieved

by using a Lagrange multiplier to weakly enforce the constraint uF ini − uSini along the interface

boundary [4]. The energy functional for the interface constraint in the mortar method is commonly

written as the following:

ΠB =

ΓB

λBini(uF ini − uSini)dΓB , (B.1)

Once again, in the LLM method, the Lagrange multiplier’s boundaries are instead associated with

the particular system’s domain and discretized with the particular system’s domain. Therefore, the

Lagrange multiplier nodes are collocated with the system’s boundary nodes [103].

With a similar Finite Element implementation as discussed in Section 2.3, the total functional

for the FSI system derived with the Mortar method with the use of Eq. (B.1) for the interface portion

is as follows:

ΠTotal = uTS (1

2KSuS + CSuS + MSuS − fS)+

ψT (1

2Kfdψ + Cfdψ + Mfdψ − DT fF )+

λT (BTSuS − BT

FDψ).

Here in the mortar method, the Bk matrix not only serves as the Boolean matrix, but relates the

system’s boundary degree of freedoms to the degree of freedoms at the interface frame. The sum of

all the terms that are multiplied by the same differential are equated to zero, due to the stationary

condition. This leads to the following equation set for the Mortar method:

KMS 0 BS

0 KMfd −DT BF

BTS −BT

FD 0

uS

ψ

λ

=

fS

DT fF

0

. (B.2)

Given this formulation, a transient analysis can be developed in a similar manner as discussed in

Chapter 2. Herry et al. [70] carries out an analogous transient analysis with the mortar method,

but discretizes the interface frame with all nodes on both subdomains. As done by Jakobsen [72], a

relationship between the fluid boundary displacements uFB and the structure boundary displacements

146

uSB can be found as

uFB =[

BF BTF

]

−1BF BT

S . (B.3)

Given this relationship, the correlation between the subdomains’ forces can be determined by use of

the virtual work at the boundary of the two subdomains as described in Section 3.3.

B.2 Consistent Method

Another common method for FSI where the computational domains have non-matching discrete

interfaces is the Consistent Interpolation method developed by Farhat et al. [38]. Originally, the

method developed from a need to determine the force on a structure from the pressure of the fluid.

Therefore, the fluid pressure was determined at structural gauss points using the discretization method

of the fluid and the same shape functions of the fluid element; thus, the term Consistent Interpolation

method. Given the pressure at the structural gauss points, the force integral for the structure was

easily computed with a Quadrature rule that computes approximate nodal loads. This is a very logical

idea. Unfortunately, as mentioned by Farhat et al. [38], this is strictly speaking non-conservative, for

the method does not guarantee that the sum of the discrete loads on the wet surface of the structure

are exactly equal to the sum of the fluid loads on the fluid dry face. Although, in several numerical

examples this was not an issue.

In order to resolve the non-conservative property a larger number of gauss points, then used for

the element level stiffness matrix, can be used if a high pressure gradient is anticipated over a certain

area. However, to make the method more practical to implement Farhat et al. [38] proposed the

following modification to the method. In this study, this modification is referred to has the Consistent

Interpolation method.

First, each fluid grid point, j, on the fluid boundary (dry face) is paired with the closest structural

element on the boundary (wet face). Then the natural coordinates Xj, in the structure element Ω(e)S

of the fluid point (or its projection onto Ω(e)S ) are determined. Next, the fluid displacement, uF , are

interpolated inside Ω(e)S using the same shape functions, Ni used in the structural element.

uFj = uF (j) = uS(Xj) =∑

Ni(Xj)uSj . (B.4)

This provides a relationship between the fluid boundary displacements and the structure boundary

displacements. Given this relationship and the fact that energy is conserved at the interface, a rela-

tionship between the forces can be determined as done in Section 3.3.1.

147

B.3 Relationship between LLM, Mortar, and Consistent Interpolation

The LLM method, the Mortar method, and the Consistent Interpolation method, will produce

the exact same relationship between the subdomains’ boundary displacements given certain discretiza-

tions and as long as the Lagrange space uses shape functions of Dirac delta functions. For instance,

if the LLM method discretizes the interface boundary the same as the coarse mesh of the two subdo-

mains, then in essence the refined mesh displacements would be projected onto the coarse mesh and

interpolated among the values of the coarse mesh’s displacements. This is the same as the Consistent

Interpolation method discussed above if the fluid mesh is more refined then the structure mesh. Like-

wise, if the interface boundary is discretized as the refined mesh, then the Mortar method projects

the refined mesh displacements onto the coarse mesh and interpolates with the coarse mesh values of

displacement. A simple diagram of this process is shown in Fig. B.1, where the refined (fluid) mesh

is projected onto the coarse (structure) mesh shown with dashed lines. The solid lines illustrate how

the structure displacements provide the values of the fluid displacements.

148

111

1113

13

23

23

(a) LLM method

1 1

1111

13

13

23

23

(b) Mortar method

1113

13

23

23

(c) Consistent Interpolation method

Figure B.1: An example of how the LLM method, the Mortar method, and the Consistent Interpolationmethod can produce the same relation matrices.

Appendix C

Discretization Error Bound for Linear Interpolation of the Field Variables at the

Interface in the LLM Method

This appendix evaluates the discretization error in displacements for linear interpolations of the

displacement variables on the interface boundary. A similar procedure as presented here is provided

by Heath and Jiao [69].

First it is assumed that domain of the system is decomposed into two non-overlapping sub-

domains (ΩS and ΩF ). The intersection of the subdomains is assumed to be composed of a set of

disjoint straight line pieces Γi. Thus, the subdomains are in two dimensions and the intersection of

the domains is composed of one dimensional straight line segments. In the LLM method, an interface

frame is inserted in between the subdomains and the trace of the true solution is defined on the in-

terface. The jump on an interface segment is enforced with two separate lagrange multipliers to form

the interface functional:

ΠB =

Γi

λSi(uSi − uBi)dΓi +

Γi

λF i(uF i − uBi)dΓi. (C.1)

To begin the discretization evaluation, attention is focused on answering the following question, what

is the error of the approximation on the interface frame given an exact function of the solution on one

side of the interface (for the fluid uF (x))? Let a piecewise linear function uF (x) be an approximation

to the exact function uF (x), on the fluid subdomain interface boundary, then it is known [29]

‖uF (x) − uF (x)‖∞ ≤ 1

8h2f‖u′′F (x)‖∞, (C.2)

where hf denotes the mesh size of the fluid trace at the interface.

In this study, the Lagrange multiplier shape functions were that of Dirac delta functions. Thus,

if the frame is interpolated with piecewise linear shape functions then the approximation to the true

solution is obtained from

uB(x) =

nB∑

j

uB(tj)ϕj (C.3)

150

where uB(tj) = uF (tj) at the interface node locations, nB is the number of boundary elements on the

interface segment, and ϕi are linear basis functions (shape functions) [69]. This is depicted in Fig. C.1.

uF (x)

x

uF1 uF2 uF3 uF4

uB1 uB2 uB3

uB(x)

uF (x)

Figure C.1: Discretization error for the portion of the constraint functional∫

ΓiλF i(uF i − uBi)dΓi

Let [a, b] be a typical interval of an interface line segment Γi, and ∆(x) represent uF (x)− uB(x).

Considering ∆(x) over the interval [a, b], it is noted that ∆(x) reaches its maximum at either x = a,

x = b, or x = c where ∆′(c) = 0. For the first two cases, uB(a) = uF (a) and uB(b) = uF (b); hence,

the error is bounded by 18h

2f‖u′′F (x)‖∞ ≤ 1

8 (h2f + h2

b)‖u′′F (x)‖∞, where hb represents the mesh size of

the boundary elements. For the case x = c, suppose c − a ≤ hb2 ≤ b − c. Expanding ∆ in a Taylor

series about a,

∆(a) = ∆(c) + (c− a)∆′(c) +1

2(c− a)2∆′′(w), (C.4)

where a < w < c. Since ∆(a) ≤ 18h

2f‖u′′F (x)‖∞, ∆′(c) = 0, and ∆′′(x) = u′′F (x) − u′′B(x) = u′′F (x),

|∆(c)| ≤ 1

2(c− a)2∆′′(w) +

1

8h2f‖u′′F (x)‖∞ ≤ 1

8(h2f + h2

b)‖u′′F (x)‖∞. (C.5)

Therefore, ‖uF (x) − uB(x)‖∞ = O(h2f + h2

b).

By proceeding in a similar fashion, the structure side would give an equivalent error. Thus the

total discretization error would be of the order O(h2f + h2

s + h2b). This represents the discretization

error at the interface used to enforce the jump conditions.