integration of fluid sloshing models with complex vehicle

163
Integration of Fluid Sloshing Models with Complex Vehicle System Algorithms BY BRYNNE E. NICOLSEN B.Sc. in Bioengineering, University of Illinois at Chicago, 2015 THESIS Submitted as partial fulfillment of the requirements for the degree of Doctor of Philosophy in Mechanical Engineering in the Graduate College of the University of Illinois at Chicago, 2019 Chicago, Illinois Defense Committee: Professor Ahmed A. Shabana, Chair and Advisor, Mechanical and Industrial Engineering Michael Brown, Mechanical and Industrial Engineering Craig D. Foster, Civil and Materials Engineering Thomas J. Royston, Bioengineering James O’Shea, Exponent

Upload: others

Post on 30-Oct-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Integration of Fluid Sloshing Models with Complex Vehicle System Algorithms

BY

BRYNNE E. NICOLSEN

B.Sc. in Bioengineering, University of Illinois at Chicago, 2015

THESIS

Submitted as partial fulfillment of the requirements

for the degree of Doctor of Philosophy in Mechanical Engineering

in the Graduate College of the

University of Illinois at Chicago, 2019

Chicago, Illinois

Defense Committee:

Professor Ahmed A. Shabana, Chair and Advisor, Mechanical and Industrial Engineering

Michael Brown, Mechanical and Industrial Engineering

Craig D. Foster, Civil and Materials Engineering

Thomas J. Royston, Bioengineering

James O’Shea, Exponent

ii

This thesis is dedicated to my family.

iii

ACKNOWLEDGEMENTS

I would like to thank my advisor, Dr. Ahmed Shabana, for his continued guidance and

encouragement during my studies. I am grateful for the opportunities he has provided me, and for

pushing me to never settle for less than my best effort. I would also like to thank the members of

my thesis committee: Dr. Michael Brown, Dr. Craig Foster, Dr. Thomas Royston, and Dr. James

O’Shea, for reviewing my work and providing valuable feedback. I am also grateful to Dr. Antonio

Recuero and Dr. Liang Wang for their invaluable assistance and encouragement when I first joined

the Dynamic Simulation Laboratory and had much to learn.

I would like to acknowledge the financial support of the National Science Foundation,

without which focusing on my research would have been much more difficult.

I would also like to thank my friends and colleagues in the Dynamic Simulation Laboratory

with whom I worked over several years, including Mohil Patel, Shubhankar Kulkarni, and

Emanuele Grossi. Each has contributed to my growth and education, through direct assistance and

informative discussions. I am also grateful to my internship supervisors at Navistar, Inc., Stefano

Cassara and Dr. Brendan Chan, for providing me with the opportunity to gain hands-on experience

learning about vehicle dynamics and working in an industry setting. Finally, I would like to thank

my family for teaching me to view the world from a scientific perspective and for providing me

with an upbringing that allowed me to pursue my dreams, and my fiancé Brian Tinsley, whose

support and encouragement were invaluable during my most difficult times.

iv

CONTRIBUTIONS OF THE AUTHORS

Chapter 2 represents a published manuscript (Nicolsen et al., 2017). My research advisor, Dr.

Ahmed A. Shabana, contributed to the review of this manuscript and guidance of the work. My

colleague, Dr. Liang Wang, contributed to developing the fluid constitutive model, the fluid-tank

contact model, and guidance of the work. I contributed to developing the vehicle model, the

numerical simulations, the tire/ground and fluid/tank contact models, and writing the manuscript.

Chapter 3 represents a published manuscript (Shi et al., 2017). My research advisor, Dr. Ahmed

A. Shabana, contributed to the review of this manuscript and guidance of the work. My colleagues,

Dr. Huailong Shi and Dr. Liang Wang, contributed to developing the fluid constitutive model, the

fluid/tank contact model, the numerical simulations, and writing the manuscript. I contributed to

the numerical simulations, the fluid/tank contact model, and writing the manuscript. Chapter 4

represents work that is not yet published. My research advisor, Dr. Ahmed A. Shabana, contributed

to review of this manuscript and guidance of the work. I contributed to developing the models, the

numerical simulations, and writing the manuscript.

v

TABLE OF CONTENTS

1. INTRODUCTION................................................................................................................... 1

1.1 Fluid Sloshing Phenomenon in Freight Transport .............................................................. 2

1.2 Fluid Modeling Techniques ................................................................................................ 4

1.3 Electronically-Coupled Pneumatic (ECP) Braking ............................................................. 7

1.4 Geometrically-Accurate ANCF/FFR Finite Elements ........................................................ 8

1.5 Scope and Organization of the Thesis ................................................................................. 9

2. FLUID MODELING WITH HIGHWAY VEHICLE APPLICATIONS ........................ 15

2.1. Basic Force Concepts ....................................................................................................... 16

2.2. Continuum-Based Inertia Force Definitions .................................................................... 19

2.3. ANCF Description of the Fluid Geometry ....................................................................... 22

2.4. ANCF Fluid Constitutive Model ...................................................................................... 25

2.5. Fluid-Tank Interaction...................................................................................................... 29

2.6. Vehicle Model Components ............................................................................................. 31

2.7. Specified Motion Trajectories .......................................................................................... 38

2.8. Equations of Motion ......................................................................................................... 41

2.9. Numerical Results ............................................................................................................ 42

2.10. Concluding Remarks ....................................................................................................... 55

3. FLUID MODELING WITH RAILROAD VEHICLE APPLICATIONS ....................... 58

3.1. Basic Inertia Force Definitions ........................................................................................ 59

3.2. Integration of Geometry and Analysis for Railroad Sloshing .......................................... 63

3.3. Fluid/Tank Interaction Forces .......................................................................................... 70

3.4. ANCF Fluid Constitutive Equations ................................................................................ 74

3.5. Integration with MBS Railroad Vehicle Algorithms ....................................................... 78

3.6. Numerical Simulations ..................................................................................................... 83

3.7. Concluding Remarks ........................................................................................................ 97

4. GEOMETRICALLY ACCURATE REDUCED ORDER FLUID MODELS ................ 99

4.1. FE Mesh Geometry and Position Vector Gradients ....................................................... 100

4.2. Finite Element Formulations .......................................................................................... 103

4.3. Fluid Modeling Approaches ........................................................................................... 109

4.4. Fluid/Tank Contact ......................................................................................................... 111

vi

TABLE OF CONTENTS (continued)

4.5. Equations of Motion ....................................................................................................... 114

4.6. Numerical Examples ...................................................................................................... 115

4.7. Concluding Remarks ...................................................................................................... 125

5. SUMMARY AND CONCLUSIONS ................................................................................. 127

6. APPENDIX A ...................................................................................................................... 132

7. APPENDIX B ...................................................................................................................... 134

8. APPENDIX C ...................................................................................................................... 136

9. REFERENCES .................................................................................................................... 138

10. VITA..................................................................................................................................... 147

vii

LIST OF TABLES

TABLE 1.1. ECONOMIC CHARACTERSTICS OF THE TRANSPORTATION INDUSTRY

IN 2007 ............................................................................................................................................2

TABLE 1.2. FREIGHT TONNAGE IN 2007 .................................................................................3

TABLE 2.1. MBS MODEL INERTIAL PROPERTIES ...............................................................32

TABLE 2.2. SUSPENSION PARAMETERS ...............................................................................38

TABLE 2.3. INITIAL VELOCITIES ............................................................................................42

TABLE 3.1. TRACK GEOMETRY ..............................................................................................80

TABLE 4.1. SLOSHING BOX MODEL INFORMATION .......................................................116

TABLE 4.2. REFERENCE CONDITIONS ................................................................................117

TABLE 4.3. NORMALIZED VEHICLE MODEL CPU TIMES ...............................................124

viii

LIST OF FIGURES

Figure 2.1. Force diagrams of a vehicle during (a) straight-line motion and (b) curve negotiation

........................................................................................................................................................16

Figure 2.2. Change in tire contact force during curve negotiation: (a) theoretical values, (b)

simulation results ...........................................................................................................................17

Figure 2.3. Tank geometry .............................................................................................................22

Figure 2.4. Initially curved fluid geometry ....................................................................................24

Figure 2.5. ANCF fluid mesh ........................................................................................................25

Figure 2.6. Fluid configurations.....................................................................................................26

Figure 2.7. Fluid-tank interaction in the (a) radial and (b) longitudinal direction .........................30

Figure 2.8. Brush Tire model coordinate systems .........................................................................34

Figure 2.9. Ackermann steering mechanism ..................................................................................36

Figure 2.10. Steering mechanism geometry ..................................................................................37

Figure 2.11. Trajectory constraint coordinate systems ..................................................................39

Figure 2.12. Commercial medium-duty tanker truck model ..........................................................42

Figure 2.13. Velocity during braking .............................................................................................43

Figure 2.14. Fluid sloshing due to braking ....................................................................................44

Figure 2.15. Normal force on a front tire and a rear tire during braking .......................................44

Figure 2.16. Position of fluid center of mass relative to tank during braking ...............................45

Figure 2.17. Flat free surface at steady state after braking ............................................................46

Figure 2.18. Lane change trajectory ..............................................................................................46

Figure 2.19. Lateral sloshing due to lane change maneuver ..........................................................48

Figure 2.20. Normal force on (a) a left-hand tire and (b) a right-hand tire during a lane change .48

ix

LIST OF FIGURES (continued)

Figure 2.21. Position of fluid center of mass relative to tank during lane change.........................49

Figure 2.22. Lateral friction force on (a) a left-hand tire and (b) a right-hand tire during a lane

change ............................................................................................................................................49

Figure 2.23. Lateral slip velocity on a left-hand tire during a lane change ...................................50

Figure 2.24. Curve trajectory .........................................................................................................51

Figure 2.25. Normal force on an outer tire and an inner tire during curve negotiation .................52

Figure 2.26. Lateral friction force on (a) an outer tire and (b) an inner tire during curve negotiation

........................................................................................................................................................52

Figure 2.27. Outward inertia force on fluid during curve negotiation ...........................................54

Figure 2.28. Position of fluid center of mass relative to tank during curve negotiation ................54

Figure 2.29. Normalized velocity of the fluid center of mass in the (a) longitudinal and (b) lateral

and vertical directions ....................................................................................................................55

Figure 3.1. A planar flexible body negotiating a curve ..................................................................61

Figure 3.2. Wheel/rail contact ........................................................................................................64

Figure 3.3. Curved ANCF rail element ..........................................................................................66

Figure 3.4. Fluid and tank geometry ..............................................................................................66

Figure 3.5. Cross-section mesh of the fluid inside a cylindrical tank ............................................67

Figure 3.6. ANCF solid element in the (a) curved reference and (b) straight configurations .......68

Figure 3.7. Initially curved ANCF fluid mesh ...............................................................................39

Figure 3.8. Tank geometry and coordinate systems ......................................................................71

Figure 3.9. Railroad vehicle model ................................................................................................79

Figure 3.10. Flowchart of the numerical solution procedure .........................................................84

x

LIST OF FIGURES (continued)

Figure 3.11. Lateral component of gravity and outward inertia forces of the fluid .......................86

Figure 3.12. Position of the tank center with respect to the track in the lateral direction at (a)

40km/h, (b) 60 km/h, and (c) 90 km/h ...........................................................................................87

Figure 3.13. Tangential component of fluid gravity and inertia forces at 40 km/h .......................88

Figure 3.14. Liquid center of mass with respect to the tank in the longitudinal direction ............89

Figure 3.15. Tread normal contact force of truck lead wheelset (a) 40km/h, (b) 60 km/h, (c) 90

km/h ...............................................................................................................................................90

Figure 3.16. Flange normal contact force of truck lead wheelset (a) 40km/h, (b) 60 km/h, (c) 90

km/h ...............................................................................................................................................91

Figure 3.17. Tread lateral contact force of lead wheelset of lead truck (a) 40km/h, (b) 60 km/h, (c)

90 km/h ..........................................................................................................................................92

Figure 3.18. Flange lateral contact force of lead wheelset of lead truck (a) 40km/h, (b) 60 km/h,

(c) 90 km/h .....................................................................................................................................92

Figure 3.19. Average normal contact forces of lead and rear trucks in the traction case ...............93

Figure 3.20. Fluid center of mass longitudinal displacement with respect to the tank in the traction

case .................................................................................................................................................94

Figure 3.21. Coupler forces between two cars in the case of braking using (a) Conventional brake,

(b) ECP brake .................................................................................................................................94

Figure 3.22. Front car fluid center of mass displacement with respect to the tank during braking (a)

Longitudinal, (b) Lateral direction .................................................................................................96

Figure 3.23. Braking animation of two tank-cars filled with liquid in the (a) Parked state, (b)

Braking state ..................................................................................................................................96

xi

LIST OF FIGURES (continued)

Figure 4.1. Cylindrical vehicle tank .............................................................................................100

Figure 4.2. Fluid configurations...................................................................................................101

Figure 4.3. Floating Frame of Reference formulation .................................................................106

Figure 4.4. Box boundary conditions ...........................................................................................117

Figure 4.5. Maximum deformation of most refined (a) ANCF and (b) ANCF/FFR fluid meshes

......................................................................................................................................................118

Figure 4.6. Vertical corner node position of (a) ANCF meshes and (b) ANCF/FFR meshes .....119

Figure 4.7. Normalized CPU times for the sloshing box models ................................................120

Figure 4.8. Medium-duty tanker truck MBS model ....................................................................120

Figure 4.9. Lane change path .......................................................................................................121

Figure 4.10. Lateral position of fluid center of mass with respect to tank during lane change

maneuver ......................................................................................................................................122

Figure 4.11. (a) Normal force on (a) a left-hand tire and (b) a right-hand tire during a lane change

......................................................................................................................................................123

Figure 4.12. (a) Lateral friction force on (a) a left-hand tire and (b) a right-hand tire during a lane

change ..........................................................................................................................................124

xii

SUMMARY

A new continuum-based total-Lagrangian liquid sloshing approach that accounts for the effect of

complex fluid and tank geometry on highway and railroad vehicle dynamics is developed in this

thesis. A unified geometry/analysis mesh is used from the outset to examine the effect of liquid

sloshing on vehicle dynamics during curve negotiation and braking, including electronically

controlled pneumatic (ECP) brakes. ECP brakes produce braking forces uniformly and

simultaneously across all railroad cars and are employed in order to reduce stopping distances and

coupler forces. In these motion scenarios, the liquid experiences large displacements and

significant changes in shape that can be captured effectively using the finite element (FE) absolute

nodal coordinate formulation (ANCF). ANCF-FEs can describe complex mesh geometries and

the change in inertia due to the change in the fluid shape, allowing for accurately capturing the

effect of the sloshing forces during motion scenarios.

The liquid sloshing models are integrated with three-dimensional multibody system (MBS)

highway and railroad vehicle algorithms which account for the nonlinear tire/ground and

wheel/rail contact. The three-dimensional contact force formulations used in this thesis account

for the longitudinal, lateral, and spin forces that influence the vehicle stability. A continuum-based

fluid constitutive model is employed, and a penalty-based fluid-tank contact algorithm is

developed. In order to examine the effect of the liquid sloshing on the vehicle dynamics during

curve negotiation, a general and precise definition of the outward inertia force is defined, which

for flexible bodies does not take the simple form used in rigid body dynamics. Specified motion

trajectories are used to examine the vehicle dynamics in different scenarios including deceleration

during straight-line motion, rapid lane change, and curve negotiation. The balance speed and

xiii

SUMMARY (continued)

centrifugal effects in the case of a railroad tank-car partially filled with liquid are studied and

compared with an equivalent rigid body model during curve negotiation and braking scenarios. In

particular, the results obtained in the case of the ECP brake application of two railroad freight car

model are compared with the results obtained when using conventional braking. The highway

vehicle tire contact forces are investigated and the effects of fluid sloshing on the vehicle stability

are demonstrated.

Lastly, for the first time the newly developed absolute nodal coordinate

formulation/floating frame of reference (ANCF/FFR) solid FEs are integrated with a fully

nonlinear MBS algorithm. ANCF/FFR-FEs are able to capture initially curved structures such as

the fluid within a cylindrical tank while retaining the same number of degrees of freedom as

conventional elements and taking advantage of modal reduction techniques, resulting in faster

simulation times compared to the higher-order ANCF elements. The solid element is developed in

terms of constant geometric coefficients which are obtained using the matrix of position vector

gradients defined in the reference configuration. No geometry distortion occurs when computer-

aided design (CAD) models are converted to FE meshes using ANCF/FFR elements because such

meshes are developed using ANCF elements, which are related to B-splines and Non-Uniform

Rational B-Splines (NURBS) by a linear mapping. The fluid constitutive model, which is based

on the Navier-Stokes fluid model, is developed and the incompressibility conditions, which are

enforced using a penalty approach, are defined. A sloshing box model and a medium-duty tanker

truck model with a tank half filled with water are developed in order to investigate the ability of

the new ANCF/FFR elements to model the fluid sloshing in comparison to fluid meshed using

ANCF elements. The fluid/tank contact formulation, which is enforced using a penalty approach,

xiv

SUMMARY (continued)

is described. It is shown that while the sloshing amplitudes of the ANCF/FFR box meshes are

reduced compared to the converged ANCF meshes, the general sloshing behavior is still captured

at a significantly reduced CPU time, indicating that the ANCF/FFR elements can contribute to

significant improvement of the computational efficiency in applications in which capturing some

geometric changes due to the fluid displacement is not critical. This conclusion is confirmed by

the results of the highway vehicle lane change simulation – the sloshing amplitudes of the center

of mass predicted using the ANCF/FFR fluid mesh during the lane change are found to be in a

good agreement with what predicted by the ANCF mesh. Furthermore, the results of the overall

vehicle-dynamics, as measured by the tire contact forces predicted using the two different meshes,

are found to be in a good agreement. The results obtained demonstrate that if the goal is to

accurately capture the free-surface displacement of the fluid, then ANCF elements are better

candidates due to their high order and ability to capture complex shapes. However, if the goal is

to perform efficient simulations to obtain the overall vehicle motion, then using ANCF/FFR

elements are a better alternative.

1

CHAPTER 1

INTRODUCTION

Fluid sloshing is the motion of a liquid in a container subjected to forced oscillations and occurs

in a moving container which is not fully filled. One industry which is significantly affected by

fluid sloshing is freight transportation – an increase in demand for crude oil and other hazardous

materials (HAZMAT) has in turn increased the number of highway and railroad vehicles

transporting these materials. Fluid sloshing can have a significant effect on vehicle dynamics,

especially in curve negotiation and traction and braking scenarios. It is clear that thorough

understanding of this complex phenomenon is necessary in order to design safe and reliable

vehicles. While many approaches have been used to model fluid sloshing, such as discrete inertia,

discrete element, and computational fluid dynamics models, each has drawbacks which limit its

scope of application. Finite element analysis (FEA), however, addresses many of these

shortcomings by using a general, physics-based approach. One objective of this thesis is to

integrate high-fidelity finite element (FE) fluid sloshing models with complex multibody system

(MBS) highway and railroad vehicle algorithms in order to study the effect of fluid sloshing on

vehicle dynamics and stability. Additionally, the effect of fluid sloshing on the coupler forces of

railroad vehicles during braking scenarios will be investigated, including the efficacy of

electronically-controlled pneumatic brakes. The standard definition for the centrifugal force will

also be assessed and it will be shown that it is inadequate for describing the outward inertia forces

of flexible bodies. Finally, fluid sloshing models based on two different FE formulations will be

compared in order to assess their ability and efficiency in accurately capturing the fluid sloshing

phenomenon.

2

1.1. Fluid Sloshing Phenomenon in Freight Transport

The field of fluid dynamics has been extensively studied for decades using, for the most part,

Eulerian approaches. Another area of application that has recently seen significant advances is

vehicle dynamics, which is often examined using MBS algorithms based on a total Lagrangian

approach. Nonetheless, fluid-vehicle interaction impacts many areas of science and technology

including rail, highway, aerospace, and marine transportation. Although materials, including crude

oil and other HAZMAT, are transported using a variety of methods, including shipping vessels

and pipelines, transportation by highway vehicle dominates the industry, generating more revenue

and creating more jobs than the other modes of transportation combined, as shown by the data

presented in Table 1.1. Due to the extent of public roads in the US and the sheer volume of freight

vehicles, the tonnage of materials transported using highway vehicles far outweighs all other

methods. This is true for both non-hazardous and hazardous materials, as shown in Table 1.2 (U.S.

Department of Transportation, 2011).

Table 1.1. Economic characteristics of the transportation industry in 2007 (U.S. Department of

Transportation, 2011)

Mode Establishments Revenue (millions) Paid Employees

Highway 120,390 217,833 1,507,923

Railway* 563 49,400 169,891

Waterway 1,721 34,447 75,997

Pipeline 2,529 25,718 36,964

*Data for Railway are for 2009.

Rollovers are more common in tanker trucks than passenger vehicles because trucks have

a higher center of gravity. Rollovers can occur due to a variety of reasons, including vehicle and

road conditions, load size, and the most common, driver error, which accounts for up to 78% of

tanker truck rollovers (U.S. Department of Transportation, 2007). Hazardous materials are

3

regularly transported by tanker trucks, and accidents in which the tank is compromised and the

contents are released can lead to damage to the environment and the surrounding infrastructure,

fires and explosions, and civilian injuries and casualties (Shen et al., 2014; WSB-TV 2015). In

the last decade alone, highway transportation accidents comprised the majority of all HAZMAT

incidents, with 144,296 out of a total of 166,494 incidents; other incidents include air, railway, and

water transportation accidents. Highway accidents have also proven to be the most deadly and

costly, accounting for 100 out of 105 documented fatalities and 1,520 out of 2,129 injuries, at a

cost of $6.1 billion out of $8.2 billion in damages (U.S. Department of Transportation, PHMSA,

2015). Therefore, thorough testing and virtual prototyping are necessary to ensure better vehicle

design and stability. However, because physical prototyping is expensive, inefficient, and time-

consuming, it is necessary to develop accurate predictive models to investigate the effect of liquid

sloshing on the dynamics of highway vehicles subject to different loading conditions and motion

scenarios.

Table 1.2. Freight tonnage in 2007 (U.S. Department of Transportation, 2011)

Mode Hazardous Materials Non-Hazardous Materials Total Tons

(Thousands) Tons

(Thousands)

Percentage

of Mode

Tons

(Thousands)

Percentage

of Mode

Highway 1,202,825 14 7,575,888 86 8,778,713

Railway 129,743 7 1,731,564 93 1,861,307

Waterway 149,794 37 253,845 63 403,639

Pipeline 628,905 97 21,954 3 650,859

Air 362 10 3,256 90 3,618

As shown in Table 1.2, railroad vehicles are the second most common mode of

transportation of freight materials. The high demand of crude oil and other HAZMAT

transportation has resulted in many serious and environmentally-damaging highway and railroad

4

accidents (King and Trichur, 2015; Wronski, 2013). In railroad transportation, liquid sloshing can

have a significant effect on railroad vehicle dynamics, especially in curve negotiation and traction

and braking scenarios (Vera et al., 2005; Wang et al., 2014). Although statistics show that most of

the accidents were caused by misuse or careless driving by the operator, extensive mathematical

and empirical studies must be performed to examine the effect of liquid sloshing on vehicle

dynamics and stability.

1.2. Fluid Modeling Techniques

Although recent advances allow for modeling more accurate fluid behavior, most commonly used

models are insufficient in adequately capturing the dynamics of the fluid in complex motion

scenarios, particularly in the cases of three-dimensional finite rigid body rotations. Early sloshing

models represented the fluid as a series of planar pendulums or mass-spring systems (Graham,

1951; Graham and Rodriguez, 1952; Abramson, 1966; Zheng et al., 2012); spherical and

compound pendulums were later used to capture nonlinearities in the motion and damping was

added to include the effect of energy dissipation (Ranganathan et al., 1989). Discrete inertia models

have been used extensively in studying sloshing dynamics in the aerospace industry since the

1960s (Abramson, 1966; Cui et al., 2014; Nichkawde et al., 2004). Coefficients for these

equivalent mechanical models can be obtained from experimental results, or using potential flow

solutions (Dodge, 2000). However, while these discrete inertia models have been improved over

time, such models cannot be used to accurately capture the change in inertia due to a change in

fluid shape and the complex dynamics that results from the vehicle motion (Liu and Liu, 2010).

Furthermore, the discrete rigid body models do not allow for modeling the continuous free surface

of the fluid, and it has been found that while the solution of a pendulum system agreed well with

5

a computational fluid dynamics (CFD) model for tanks with low fill levels, increasing the fill level

resulted in underprediction of the sloshing amplitudes and forces. Tank models with low fill levels

are relevant to aerospace applications and contribute to better understanding of the vehicle

behavior during low-fuel scenarios; however, for highway and railroad vehicles it is desirable to

transport the cargo at maximum capacity in order to reduce the sloshing behavior and maximize

transportation efficiency, so discrete inertia models are not suitable for studying such systems.

The discrete element method (DEM), in which the fluid is modeled as a system of small

particles, has also been used to study fluid sloshing (Cundall and Strack, 1979). DEM has the

advantage of capturing the mixing of different fluids and is thus often used in multi-phase

simulations (Monaghan, 2012; Nishiura et al., 2014). DEM is also capable of capturing fluid

separation and is therefore an important tool in studying fluid-structure interaction (Boffi and

Gastaldi, 2016; Pingle et al., 2012). However, while torsional motion and fluid separation are

important for systems such as multi-phase flow and fluid-structure interaction, these phenomena

do not significantly affect the overall vehicle dynamics. Furthermore, DEM models often suffer

from very high problem dimensionality, particularly in the case of vehicle tanks with cargo in

excess of several thousand gallons – for example, 15,000-gallon cargo is common in the case of a

rail tank-car. A large tank volume requires millions of particles to accurately capture the sloshing

behavior, which significantly increases the computation cost. Thus, DEM models are also not

optimal for studying sloshing dynamics in vehicle applications.

The governing equations of fluid dynamics, including conservation of mass and

momentum, are in general highly nonlinear coupled differential equations, and a closed-form

analytical solution does not exist. Computational fluid dynamics (CFD) is a numerical approach

for solving the fluid equations, where the fluid volume is divided into many control volumes over

6

which the partial differential equations are converted to discrete equations which can be solved

iteratively (Griffiths and Boysan, 1996). There are several commercially-available CFD solvers

such as ANSYS Fluent (ANSYS, 2019). Eulerian-based CFD solvers are commonly used by the

fluid dynamics community due to the Eulerian nature of many current problems of interest.

However, many MBS algorithms are based on a Lagrangian formulation, and integration with

Eulerian CFD solvers has been shown to be problematic. Fluid sloshing is often studied in fixed

containers which are subjected to forced excitation, so Eulerian approaches are suitable. Practically,

however, fluid sloshing often occurs in dynamic systems, such as rockets and highway and railroad

vehicles, resulting in highly nonlinear centrifugal and Coriolis forces which are not captured using

existing CFD models. Thus, in order to use the CFD approach to study a complex mechanical

system, it would be necessary to combine an Eulerian fluid sloshing problem with a Lagrangian

MBS algorithm. Pape et al. (2016) attempted to integrate a CFD model with the commercial

vehicle-dynamics software TruckSim, but due to the incompatible nature of the two solvers,

integration was found to be impossible and a discrete inertia model was instead used. It is due to

this incompatibility that CFD models are not suitable for studying fluid sloshing in complex

mechanical systems. Furthermore, modeling the fluid free surface is also challenging using

Eulerian methods.

In the case of liquid sloshing problems, accurate definition of the geometry of the fluid and

container is necessary in order to develop a general computational framework that can be

effectively used to shed light on the effect of sloshing in complex motion scenarios. In order to

take advantage of the Lagrangian nature of existing MBS algorithms, a number of continuum-

based fluid models have been developed. Wang et al. (2015) developed a low-order fluid sloshing

model based on the FE/FFR formulation. The FE/FFR formulation uses a modal approach to

7

reduce significantly the problem dimensionality. The numerical results obtained using a railroad

vehicle model showed that while an increase in the fluid viscosity improves the stability at low

velocities due to the damping effect, at high velocities the increase in the fluid viscosity leads to

an increase in the vehicle hunting oscillations when compared to an equivalent rigid-body fluid

model. However, because the small-deformation FFR elements may not be capable of showing

severe sloshing behavior, Wei et al. (2015) developed a total-Lagrangian ANCF sloshing model.

This non-modal, non-incremental approach leads to a constant mass matrix and zero Coriolis and

centrifugal forces. It was shown that a single ANCF element can capture much more severe

deformation compared to a large number of conventional FFR elements. The FFR and ANCF

approaches are both fully Lagrangian and thus can be easily integrated with existing MBS

algorithms.

1.3. Electronically-Coupled Pneumatic (ECP) Braking

The coupler is a device which is used to connect two railcars. In the case of sudden braking,

excessive coupler forces can be generated between the railcars, which can be exacerbated by fluid

sloshing. These forces can cause damage to the couplers, which can shorten their lifespans and

require premature replacement. Failure of the coupler can also allow the railcars to separate, which

can lead to runaway cars, collisions, or derailment. The purpose of the newly introduced

electronically controlled pneumatic (ECP) braking system, recommended for long freight trains,

is to apply the braking forces uniformly and simultaneously on all railcars. This new technology

can improve both train safety and operations by reducing the coupler forces and decreasing

stopping distances. Studies have shown that the ECP braking system, as compared to conventional

8

braking, leads to a 40% reduction in the stopping distance and significant reduction in the coupler

forces (Aboubakr et al., 2016).

1.4. Geometrically-Accurate ANCF/FFR Finite Elements

Recently a new class of elements which can model initially curved structures and also allow for

using modal reduction techniques has been proposed (Shabana, 2017B). These elements, referred

to as ANCF/FFR elements, are based on the ANCF kinematic description and are thus able to

capture initially curved geometry in the reference configuration, such as fluid within a tank, tires,

and leaf springs. The position vector gradients which define the initially curved geometry of the

mesh are written in terms of finite rotations by use of a consistent rotation-based formulation

(CRBF) velocity transformation matrix, thus allowing for the development of reduced-order

models which can be used in both structural and MBS applications. Furthermore, because ANCF

elements are related to B-splines and Non-Uniform Rational B-Splines (NURBS) by a simple

linear mapping, conversion from CAD geometry to analysis meshes is a straightforward process

which avoids distortion of the mesh. Conventional finite elements are not related to CAD models

by a linear mapping, and consequently, conversion to FE meshes is a much more costly and error-

prone process, costing the U.S. automotive industry alone over $600m in 2011 (Mackenzie, 2012).

The convergence of ANCF/FFR element frequencies has been shown to agree well with

commercial FE software for simple geometries (Zhang et al., 2019; Tinsley and Shabana, 2019),

but the performance of these elements has not yet been evaluated in the case of full vehicle models

or compared to ANCF elements.

9

1.5. Scope and Organization of the Thesis

Chapter 2 was first published in the Journal of Sound and Vibration (Nicolsen et al., 2017) and is

reproduced in this thesis with permission, which is provided in Appendix A. In this chapter, a

formulation that correctly captures the geometry of the fluid and tank is used in order to accurately

represent the distributed inertia and elasticity of the fluid. In order to develop these new and unique

sloshing geometry models, ANCF elements that produce accurate geometry are used, eliminating

the need for using B-spline and NURBS representations for developing the complex fluid

geometry. The effect of the initially curved fluid geometry, which cannot be captured accurately

using existing FE formulations, is properly accounted for, leading to a systematic integration of

the geometry and analysis by adopting one fluid mesh from the outset. Such an important goal

cannot be achieved using other MBS formulations that employ modal representation for the fluid

displacements, as in the case of the FFR formulation (Wang et al., 2014).

The ANCF geometry/analysis mesh developed is used to formulate the inertia forces using

a non-modal continuum-based approach. Proper definition of the inertia forces is necessary in

order to be able to predict the effect of the sloshing on the vehicle dynamics and stability. In

particular, a continuum-based and general definition of the centrifugal forces in terms of the fluid

displacement is developed and used to shed light on the approximation made using the simple rigid

body dynamics formula 2

smV r . Accurate definition of the centrifugal forces is particularly

important in the definition of the vehicle balance speed that should not be exceeded during curve

negotiations. An ANCF fluid/tank car walls penalty contact formulation is developed and used to

determine the generalized contact forces associated with the ANCF nodal coordinates which

include absolute position and gradient vectors. The penalty contact formulation developed in this

10

chapter takes into account the fluid large displacement and complex geometry that result from the

sloshing effect.

It is shown in this chapter how general constitutive fluid models can be developed and

integrated with ANCF complex fluid geometry models, thereby opening the door for future

investigations that focus on adopting new and highly nonlinear constitutive fluid models as well

as experimenting with different tank designs that have different, complex, and unconventional

geometries. In so doing, the field of liquid sloshing can be significantly advanced to a new level.

The analysis presented in this chapter demonstrates for the first time how an ANCF liquid

sloshing model can be integrated with an MBS system computational algorithm that ensures that

the kinematic algebraic constraint equations are satisfied at the position, velocity, and acceleration

levels. Such new ANCF fluid/MBS algorithms will allow for investigating a large class of liquid

sloshing problems that cannot be solved using existing approaches. The purpose of this analysis is

to create a high-fidelity model which is capable of capturing more details than can be described by

existing modeling methods. It is important to note that simple models can still be valuable if real-

time simulations are required. In these cases, both simple vehicle and fluid models can be used to

significantly reduce the computer simulation time. High fidelity continuum-based models, on the

other hand, are necessary in order to account for the distributed inertia and viscoelasticity of the

fluid.

The use of the formulation and computational procedure developed in this chapter is

demonstrated using a fully nonlinear MBS model of a commercial medium-duty tanker truck

developed using the general-purpose MBS software SIGMA/SAMS (Systematic Integration of

Geometric Modeling and Analysis for the Simulation of Articulated Mechanical Systems). The

fluid in the tank is represented by an ANCF mesh which allows for capturing the change in inertia

11

due to the change in shape of the fluid, as well as visualizing the change in the fluid free surface

while correctly capturing the centrifugal forces. The interaction between the rigid tank walls and

the ANCF fluid is formulated using the penalty approach. The MBS model includes a suspension

system and Pacejka’s brush tire model is introduced to represent the ground-tire contact (Pacejka,

2006). Specified motion trajectories are used to examine three different working conditions –

deceleration under straight-line motion, rapid lane changing, and negotiating a curve. Reduced

integration is used to increase computational efficiency when the fluid viscosity forces are

calculated. The results show that sloshing has the effect of increasing contact forces on some

wheels and decreasing contact forces on other wheels. Severe sloshing behavior can cause vehicle

instability; in extreme cases, wheel lift and vehicle rollover may occur.

Chapter 3 was first published in the Journal of Multi-Body Dynamics (Shi et al., 2017)

and is reproduced in this thesis with permission, which is provided in Appendix B. In this chapter,

the computational continuum-based total Lagrangian approach is used to study the effect of liquid

sloshing on railroad vehicle dynamics during curve negotiation and braking scenarios. A unified

geometry/analysis mesh (Cottrell et al., 2007; Shabana 2017) is used from the outset to define the

tank-car and fluid configuration, demonstrating a successful integration of computer aided-design

and analysis (I-CAD-A) for an important and practical problem. The general definition of the

liquid outward inertia forces, which is fundamentally different from the case of rigid body

dynamics, is defined in this chapter using the FFR formulation, and it is shown that the

conventional centrifugal force definition used to define the vehicle balance speed during curve

negotiations is a special case of the more general expression. The geometry description of both the

tank and the fluid using ANCF elements is discussed, and it is shown how ANCF elements can be

used with cubic spline function representation to define the geometry of the rigid rails. The

12

formulation of the liquid/tank interaction forces and the search method used to define the fluid/tank

contact points are described and the constitutive fluid model used in the total Lagrangian and non-

incremental solution procedure adopted in this paper is briefly discussed. The integration of the

liquid sloshing model in computational MBS railroad vehicle algorithms, the track geometry, and

the three-dimensional wheel/rail contact force model are elaborated. The components of the MBS

vehicle model and the fluid model data used to examine the effect of liquid sloshing on the

performance of the newly introduced ECP brake system and the rail vehicle dynamics during curve

negotiations are also detailed. In this chapter, the results obtained using the ECP braking force

model are compared with the results obtained using the conventional air brake system. In order to

improve the efficiency of the simulation, integration techniques such as the Hilber-Hughes-Taylor

(HHT) method (Aboubakr et al., 2015) and reduced integration when calculating the fluid viscous

forces are used.

Chapter 4 presents a general procedure which can be used to develop geometrically

accurate spatial finite elements capable of capturing initially curved geometries. The ANCF/FFR

elements are developed in terms of constant geometric coefficients obtained using the matrix of

position vector gradients defined in the reference configuration. While the solid element is used in

this investigation, this procedure can also be used to develop spatial beam and plate elements. It is

shown how the ANCF gradient vector coordinates can be written in terms of the CRBF finite

rotations, which are then replaced by infinitesimal rotations. These elements, which have the same

number of degrees of freedom as conventional finite elements, can be used to develop efficient

reduced-order models with both structural and MBS applications. Due to increasing reliance on

virtual prototyping, ANCF/FFR elements have important implications for many industries

including the automotive, railroad, and aerospace industries.

13

A fluid constitutive model based on the Navier-Stokes fluid model is developed for the

ANCF/FFR elements. A similar approach as is used in this chapter can be applied to other

constitutive models including crude oil (Grossi and Shabana, 2017) and other HAZMAT, paving

the way for future investigations that can develop new, nonlinear fluid constitutive models which

can be integrated with complex MBS algorithms. The fluid/tank contact formulation, which is

based on the penalty approach, is also adopted in this chapter, and the generalized contact forces

associated with the ANCF and ANCF/FFR coordinates are developed. The approach used in this

chapter can be generalized to arbitrary tank geometries such as those featuring an oval cross-

section or hemispherical ends, which are also common in highway vehicles, or half-ellipsoid ends,

which are common on rail tank cars.

It is demonstrated for the first time how an ANCF/FFR fluid sloshing model can be

integrated with computational MBS algorithms. The algorithm used in this chapter again ensures

that the kinematic algebraic constraint equations are satisfied at the position, velocity, and

acceleration levels, and by using modal reduction techniques, parallel computation, and reduced

integration, it is possible to develop efficient fluid/vehicle models. The general-purpose MBS

software SIGMA/SAMS is again used to develop a fully nonlinear model of a tanker truck in order

to demonstrate the use of the ANCF/FFR elements in fluid modeling and to compare their behavior

with the higher-order ANCF elements. Three vehicle models are developed, in which the fluid is

represented using an ANCF mesh, an ANCF/FFR mesh, and a rigid body fixed to the tank; the

third model is used in order to isolate the effect of the fluid sloshing on the vehicle dynamics. The

contact between the flexible fluid and the tank walls, as well as the fluid incompressibility

conditions, are enforced using a penalty approach. In the vehicle model, which includes a

suspension system, the tire-ground contact is formulated using Pacejka’s brush tire model (Pacejka,

14

2006). The tanker truck model with a tank half-filled with water is tested in a rapid lane change

scenario in order to induce significant sloshing. It is concluded that the ANCF/FFR formulation

can be effective in modeling fluid sloshing problems when efficient simulations are desired. The

results obtained demonstrate that the overall vehicle motion of the low-fidelity ANCF/FFR model

is in good agreement with the high-fidelity ANCF fluid sloshing model. However, if capturing

accurately the deformation of the fluid free surface or the change in distributed inertia due to the

fluid sloshing is necessary, then ANCF elements are the better candidate – they require fewer

elements than conventional elements and fewer degrees of freedom compared to meshfree

approaches (Atif et al., 2019).

15

CHAPTER 2

FLUID MODELING WITH HIGHWAY VEHICLE APPLICATIONS

The objective of this chapter (published as Nicolsen et al., 2017) is to develop a new total

Lagrangian continuum-based liquid sloshing model that can be systematically integrated with

multibody system (MBS) algorithms to allow for studying complex motion scenarios. The new

approach allows for accurately capturing the effect of the sloshing forces during curve negotiation,

rapid lane change, and accelerating and braking scenarios. In these motion scenarios, the liquid

experiences large displacements and significant changes in shape that can be captured effectively

using the finite element (FE) absolute nodal coordinate formulation (ANCF). ANCF elements are

used in this investigation to describe complex mesh geometries, to capture the change in inertia

due to the change in the fluid shape, and to accurately calculate the centrifugal forces, which for

flexible bodies do not take the simple form used in rigid body dynamics. A penalty formulation is

used to define the contact between the rigid tank walls and the fluid. A fully nonlinear MBS truck

model that includes a suspension system and Pacejka’s brush tire model is developed. Specified

motion trajectories are used to examine the vehicle dynamics in three different scenarios –

deceleration during straight-line motion, rapid lane change, and curve negotiation. It is

demonstrated that the liquid sloshing changes the contact forces between the tires and the ground

– increasing the forces on certain wheels and decreasing the forces on other wheels. In cases of

extreme sloshing, this dynamic behavior can negatively impact the vehicle stability by increasing

the possibility of wheel lift and vehicle rollover.

16

2.1. Basic Force Concepts

In this section, a simplified planar vehicle model subjected to discrete forces is analyzed in order

to have an understanding of how the contact forces on the tires change as a tanker truck enters a

curve. A force diagram for this model during straight-line motion is presented in Fig. 2.1a, where

,w tF is the tank gravity force at the tank center of mass located at a vertical distance tz from the

ground, ,w fF is the fluid gravity force at the fluid center of mass located at a vertical distance

fz

from the ground, LN and RN are the normal forces on the left and right wheels, respectively,

located a distance ay from point O , and the motion is in the horizontal plane in the direction of

the dashed arrow as shown in the figure. During straight-line motion, the fluid is not displaced

laterally and there are no centrifugal forces exerted on the vehicle. By taking the moments of the

forces about point O , as expected, these steady-state normal forces are found to be

( )w, , 2L R t w fN N F F= = + ; that is, each wheel carries half of the total weight of the vehicle.

(a) (b)

Figure 2.1. Force diagrams of a vehicle during (a) straight-line motion and (b) curve negotiation

17

This is contrasted by the force diagram in Fig. 2.1b, where the vehicle has entered a

counter-clockwise constant-radius curve, indicated by the dashed arrow above the diagram.

Centrifugal forces ,C tF and

,C fF are exerted on the tank and fluid, respectively, lateral friction

forces fF are exerted on both tires in the opposite direction, and the center of mass of the fluid

has shifted laterally due to the centrifugal force, displacing the gravity force ,w fF by a distance

fy . Taking the moments due to these forces, one can obtain the equations for the left and right tire

contact forces in this case as , ,L w t w f RN F F N= + − and

( ), , , , , 2R c t t c f f w f f a w t w fN F z F z F y y F F = + + + +

, respectively. It can be shown that in the case

of straight-line motion, these equations reduce to the equations given previously because the

centrifugal forces C,tF and

,C fF and the lateral displacement of the fluid fy will be equal to 0.

Using this simple analysis, one can examine how the contact forces on the tires change

when a vehicle enters a curve. In Fig. 2.2a, the steady-state normal force equations are used to

calculate the contact forces for the first 0.7s, then the constant-radius curve contact force equations

are used for the following 9.3s. This represents a vehicle driving in a straight line initially before

entering a constant-radius curve at 0.7s, where it remains for the rest of the simulation. While the

results of this figure, obtained using the simple analysis and the simple force equations previously

presented in this section, do not capture the oscillations of the fluid because the lateral shift of the

fluid fy is assumed to remain constant for simplicity, it is evident from these results that the

contact force on the outer tire increases and that on the inner tire decreases. This change is due to

the lateral shift of the center of mass of the fluid, which is a result of the outward inertia force

acting on the fluid. The lateral shift of the fluid and thus the outward inertia forces act to increase

the roll moment and thus increase the contact force on the outer tire.

18

(a) (b)

Figure 2.2. Change in tire contact force during curve negotiation: (a) theoretical values, (b)

simulation results ( Left wheel, Right wheel)

These simplified results can be made more realistic by using simulation results in the

equations instead of constant theoretical values. By replacing the position of the center of mass of

the fluid fy and

fz and the centrifugal force on the fluid ,C fF with the simulation results that will

be presented in detail in Section 2.9.3, the resulting contact forces calculated by the previously

derived equations will capture the sloshing behavior. This effect is evident from the results

presented in Fig. 2.2b, where the contact forces oscillate with time due to the oscillatory motion

of the liquid. The discontinuity in the plot is due to the fact that the theoretical calculations assume

a sudden change from straight-line to constant-radius curve trajectories. In more realistic scenarios,

a spiral segment is used to connect the straight and curved sections in order to ensure a smooth

transition.

19

2.2. Continuum-Based Inertia Force Definitions

Inertial forces play an important role in the dynamics and stability of a vehicle negotiating a curve.

The centrifugal force is exerted on the vehicle in the outward normal direction of the curve. If the

bank angle of the curve is zero, the only opposing force is the inward lateral friction force due

to the contact between the tires and the ground. When the bank angle is different from zero, the

centrifugal force of a vehicle with mass m is also opposed by the component of the gravity force

which is parallel to the bank of the curve. If the rigid body assumptions are used and additionally

the vehicle forward velocity sV along the tangent to a curve of radius of curvature r is assumed

constant, one must have an upper limit on the velocity sV , called the balance speed, such that

2 sins fnmV r mg F= + , where g is the gravity constant and fnF is the component of the friction

force along the normal to the curve. Clearly, in deriving this force expression, the effect of other

forces such as suspension forces is not taken into account. It follows that the balance speed is

defined by ( )sins fnV r mg F m= + . Because the friction force cannot be predicted with high

degree of accuracy, a conservative estimate of the balance speed is normally defined in rigid body

dynamics as sinsV rg = ; this is the formula often used to develop operation guidelines for

vehicles negotiating curves. A vehicle negotiating a curve with radius of curvature r must not be

operated at a speed higher than the balance speed in order to avoid rollover. It is clear from the

equation sinsV rg = , in which the effect of friction is neglected and the assumption of rigidity

is used, that the balance speed does not depend on the mass of the vehicle, and therefore, the

guidelines specify a balance speed for a curve with specific geometry defined by the radius of

curvature and bank angle. It is clear that in the case of liquid sloshing, the simple expression of the

20

balance speed sinsV rg = cannot be in general used because the outward inertia force does not

take the simple form of 2

smV r .

When ANCF finite elements are used, the expression of the outward inertia force differs

significantly from the expression used in rigid body dynamics. For ANCF finite elements, the

vector of nodal coordinates can be written as the sum of two vectors as o d= +e e e , where oe is

the vector of nodal coordinates before displacement and de is the vector of displacements that

include large liquid reference displacements including finite rotations as well as the liquid

deformations. Therefore, the outward inertia force, as will be demonstrated in this section,

becomes function of the liquid motion and the simple expression 2

smV r is no longer applicable

for the calculation of the balance speed or for accurate force analysis during curve negotiations.

Furthermore, the vector oe can be used to systematically account for the initial curved geometry

of the liquid. As described in the literature, this can be accomplished by using the matrix of position

vector gradients oJ , where ( )o o= = J X x Se x , where x defines the element parameters in

the straight configuration, S is the shape function matrix, and o=X Se defines the reference

configuration before displacement.

In order for the vehicle to safely remain on the road, the outward inertia force must not

exceed the sum of the inward friction and gravity forces. Although the centrifugal force on a rigid

body negotiating a curve takes a simple form, as previously mentioned, the same expression does

not apply to curve negotiation of a flexible body, because such a force expression is function of

the deformation (Ibrahim et al., 2001). In general, the outward inertia force inF of a flexible body

or an ANCF finite element negotiating a curve is defined as T

oin o o

VF dV= r n , where o and oV

are, respectively, the mass density and volume of the flexible body in the reference curved

21

configuration, r is the absolute acceleration vector of an arbitrary point on the body, and n is the

outward unit normal vector to the curve. The volume in the curved reference configuration is

related to the volume in the straight configuration V before the liquid assumes the shape of the

container by the equation o odV J dV= , where o oJ = J . It is clear from the equation

T

oin o o

VF dV= r n that the component of the acceleration along the tangent to the curve will not

contribute to the outward force vector. When ANCF finite elements are used, the absolute

acceleration vector of an arbitrary point can be written as =r Se . If the flexible body is discretized

using en ANCF elements, the outward inertia force vector that must be used to define the vehicle

balance speed can be written as ( )1 1

Te e

j jo o

n nj j j T j j j j

in o o o oj jV VF dV dV

= == = r n n S e , where

superscript j refers to the element number. One can also write

( ) ( )1 1

e e

jo

n nT j j j j T j j

in o oj jVF dV

= =

= = n S e n S e , where

jo

j j j j

o oV

dV= S S . A standard FE

assembly procedure can be used by writing j j=e B e , where

jB is a Boolean matrix and e is the

vector of nodal coordinates of the body. It follows that ( )1

enT j j T

in jF

== =n S B e n Se , where

1

en j j

j==S S B is the constant assembled matrix of the constant element

jS matrices. Numerical

integration can be systematically used to evaluate the outward inertia force T

inF =n Se if analytical

integration of the element shape functions is to be avoided. In this case, one can create a mesh of

pn points on the flexible body and if an assumption is made that the mesh consists of only one

type of ANCF elements, then approximation of inF can be written as ( )1

pnT k k

in kF m

== n S e ,

where km is the lumped mass associated with the mesh point k ,

kS is the assembled matrix of

22

the element ( )k j jk j=S S x B matrices, and jk

x is the vector of the element spatial coordinates

T

x y z=x evaluated at the mesh point k that corresponds to element j .

Alternatively, one can use the moment of mass to write C o oV

m dV= r r , where m is the

total mass of the liquid, Cr is the global position vector of the liquid center of mass, and =r Se

when ANCF finite elements are used. It follows that ( ) ( )o o

C o o o oV V

dV m dV m = = r r S e ,

which can be simply written as ( )C m=r Se , and ( )C m=r Se . Therefore, the outward inertia

force vector can be written in an alternate form as ( )T

in CF m= n r . Because of the liquid

oscillations, Cr will not remain constant relative to the curve, and as a consequence, the outward

inertia force is not in general constant as in the case of a rigid body negotiating a curve.

2.3. ANCF Description of the Fluid Geometry

Figure 2.3. Tank geometry

23

In this section, the development of the initially curved ANCF geometry of the fluid that assumes

the shape of a rigid cylindrical tank is discussed. The tank used in this chapter has a cylindrical

geometry as shown in Fig. 2.3, and therefore, it is required for the ANCF fluid mesh to have the

same the shape of the container it fills and at the same time represent different levels of the free

surface. The use of the ANCF absolute positions and gradients as nodal coordinates allows for

efficient shape manipulation and for obtaining the accurate complex geometry without the need

for using the CAD B-spline and NURBS representations that have rigid recurrence structure

(Shabana, 2015; Patel et al., 2016). As previously mentioned, in the ANCF description, the

assumed displacement field can be written as ( , ) ( ) ( )t t=r x S x e , where r is the global position

vector, [ ]Tx y z=x is the vector of the element spatial coordinates, t is time, S is the time-

independent element shape function matrix, and e is the vector of the element nodal coordinates

that include absolute position and gradient coordinates (Shabana, 2017A). The superscript j that

refers to the element number is omitted here for notational simplicity. The vector of element nodal

coordinates e can be written as o d= +e e e , where oe is the vector of nodal coordinates in the

reference configuration and de is the vector of nodal displacements. The assumed displacement

field can then be written as ( )( , ) ( ) ( ) ( )o dt t t= +r x S x e e . Using the general continuum mechanics

description ( )( , ) ,t t= +r X X u X , where X is the absolute position vector of an arbitrary point in

the reference configuration and u is the displacement vector, one can write o=X Se and d=u Se .

By choosing the elements in the vector oe appropriately, initially curved structures can be defined

in a straightforward manner using ANCF elements (Shabana, 2015).

24

Figure 2.4. Initially curved fluid geometry

The fully-parameterized ANCF solid element (Olshevskiy et al., 2013; Wei et al., 2015),

based on an incomplete polynomial representation, is used in this chapter to represent the fluid by

applying the proper fluid constitutive model which will be discussed in Section 2.4. In this case,

the arbitrary fluid material point on element j can be written as

8,1 ,2 ,3 ,4

1

j k k k k jk j j

k

S S S S=

= = r I I I I e S e , where I is the 3 3 identity matrix; detailed shape

function and nodal coordinate expressions can be found in Appendix C of this thesis. For example,

consider the element j which has the initially curved structure shown in Fig. 2.4. The matrix of

position vector gradients at node k can be written as ( ) ( ) ( ) ( ) jk jk jk jk

x y zo o o o

= xr r r r . For the

specific element geometry shown in Fig. 2.4, by adjusting the magnitude of the gradient vector

( )1j

yo

r without changing the gradient vector orientation, the position vector gradients at node 5 will

be ( ) ( ) ( ) ( )5 5 5 5 j j j j

x y zo o o o

= xr r r r , where is the stretch factor used to represent the stretch

25

of the edge; the value of can be obtained by taking the ratio between the arc lengths of curves

5-8 and 1-4. Following this procedure, the complex geometry of the fluid structure can be created,

as shown in Fig. 2.5. The mesh used in this chapter consists of 48 ANCF solid elements and the

mesh has a total number of degrees of freedom of 1260.

Figure 2.5. ANCF fluid mesh

2.4. ANCF Fluid Constitutive Model

A general ANCF fluid constitutive model that can account for the initially curved configuration is

developed in this section. The proposed fluid model ensures the continuity of the displacement

gradients at the nodal points and allows for imposing a higher degree of continuity across the

element interface by applying algebraic constraint equations that can be used to eliminate

dependent variables and reduce the model dimensionality at the prepossessing stage (Wei et al.,

2015). In order to describe the fluid-structure interaction, the penalty approach, described in

Section 2.5, is used to evaluate the contact and friction forces between the fluid and the rigid tank.

By using the non-modal ANCF approach, the fluid elastic forces can be formulated without

imposing restrictions on the amount of deformation and rotation within the elements. Figure 2.6

26

shows the three configurations of the fluid; the straight, curved reference, and current

configurations. As previously mentioned, the volume of the fluid in the curved reference

configuration oV is related to the volume in the straight configuration V using the relationship

o odV J dV= , where o oJ = J is the determinant of the matrix of position vector gradients

( )o o= = J X x Se x . Therefore, integration with respect to the reference domain can be

converted to integration with respect to the straight element domain. This allows for using the

original element dimensions to carry out the integrations associated with the initially curved

configuration.

Figure 2.6. Fluid configurations

The matrix ( )o o= J Se x is constant and can be evaluated at the integration points using

the ANCF element shape function and the vector of nodal coordinates in the reference

configuration (Shabana, 2015). The matrix of position vector gradients X Y Z= =J r X r r r ,

which is used to determine the Green-Lagrangian strain tensor ( ) 2T= −ε J J I , can be written as

27

( ) ( ) ( )1

1

x y z x y z e oo oo

−− = = = =

r r xJ r r r r r r J J

X x X, where

( )e = = J r x Se x . The relationship between the volume defined in the current configuration

v and the volume in the curved reference configuration oV can be written as odv JdV= where

J = J . It follows that 1

o e o o edv JdV J dV J dV−= = =J J .

The linear fluid constitutive equations can be defined using the Cauchy stress tensor and

can be assumed as ( ) tr 2 vol devp = − + + = +σ D I D σ σ where the temperature effect is

neglected and the fluid is assumed to be incompressible, σ is the symmetric Cauchy stress tensor,

p is related to the hydrostatic pressure, and are Lame’s material constants, I is a 3 3

identity matrix, tr refers to the trace of a matrix, and D is the rate of deformation tensor (Spencer,

1980; Shabana, 2017A). If the incompressibility condition is imposed using a penalty method, the

first two terms will vanish and the constitutive model reduces to 2dev dev=σ D . It is convenient to

use the second Piola-Kirchoff stress tensor since it is associated with the Green-Lagrangian strain

tensor defined in the reference configuration. One has 1 1 1 1

2 2T T

P dev devJ J− − − −= =σ J σ J J D J , where

1 1T− −=D J εJ , ( )T T 2= +ε J J J J and = J r X . For an arbitrary element j in the fluid body, the

virtual work of the fluid stress forces can be written as

1

2: :j j

o

j j j j j j j j

s dev P o

v V

W dv dV −

= − = − σ J J σ ε (2.1)

where ( )j j j j = ε ε e e . The virtual work of the fluid viscous forces can then be written as

( )1 1

2 : 2T

j jo o

j j j j j j j j j j j j

s P o r r o v

V V

W dV J dV − −

= − = − = σ ε C ε C : ε Q e (2.2)

28

where Tj j j

r =C J J is the right Cauchy-Green deformation tensor, and upon using the identity

1 1T Tj j j j j j

e o e oJ − −= = =J J J J J , the vector of generalized viscosity forces j

vQ associated with the

ANCF nodal coordinates can be written as

( ) ( )

( )

1 1 1 1

1 1

02 2

2

j jo

j

j jj j j j j j j j j j j j

v r r o r rj j

V V

jj j j j j

e r r j

V

J dV J dV

J dV

− − − −

− −

= − = −

= −

ε εQ C ε C : C ε C : J

e e

εC ε C :

e

(2.3)

In this case, the integration over the current configuration domain is converted to integration over

the straight configuration domain.

The incompressibility condition is imposed using the penalty method. Figure 2.6 shows

that the volume relation between the reference and current configuration is j j j

odv J dV= , therefore,

1j jJ = =J and 0jJ = still hold for the initially curved fluid. By assuming the penalty energy

function ( )2

1 2j j j

IC ICU k J= − and the dissipation function ( )2

2j j j

TD TDU c J= , where j

ICk and j

TDc

are the two penalty coefficients, the generalized penalty forces associated with the ANCF nodal

coordinates that result from imposing the two penalty conditions can be defined as

( )( )

( )

1

=

T

T

j j j j j j j

IC IC IC

j j j j j j j

TD TD TD

U k J J

U c J J

= = −

=

Q e e

Q e e (2.4)

where ( )trj j jJ J= D and j j j jJ J = e e . Knowing that

( ) ( ) ( )j j j j j j j j j j

X Y Z Y Z X Z X YJ = = = r r r r r r r r r , j jJ e can be written more explicitly, by

differentiating any of the three expressions for jJ with respect to

je , as

( ) ( ) ( )T T T

T Tj j

j j j j j j j j j

X Y Z Y Z X Z X Yj j

J J = = + +

S r r S r r S r r

e e (2.5)

29

By defining the generalized forces associated with the fluid element coordinates j

e , the

generalized forces associated with the fluid body coordinates e can be obtained using a standard

FE assembly procedure.

2.5. Fluid-Tank Interaction

The fluid should remain inside the tank regardless of the severity of the sloshing and these

boundary conditions of the mesh can be defined in multiple ways. One can impose constraints on

the boundary nodes, using either Lagrange multipliers or elimination of dependent variables.

Because this method is often more computationally expensive, in this thesis, the penalty method

is used to formulate the interaction between the fluid body and the rigid tank walls. The tank

deformation is not considered in this analysis because the main focus of this thesis is on studying

the sloshing. Figure 2.7a shows the contact geometry in the radial direction; the radius of the tank

is tr , superscript t refers to the tank,

fr is an arbitrary point on the fluid body, and

tR is the

position vector of the tank reference point defined in the global coordinate system. The position of

an arbitrary point of the fluid defined in the tank local coordinate system can be written as

( ) 1 2 3

T T Tft t ft t f t ft ft ftu u u = = − = u A u A r R , where

tA is the 3 3 transformation matrix

which defines the tank orientation, and the bar notation means the vector or matrix is defined in

the body local coordinate system. The inequality ( ) ( ) ( )2 2 2

2 3

ft ft tu u r+ implies that the fluid point

is inside the tank and there is no need for applying a penalty force. On the other hand, the equation

( ) ( ) ( )2 2 2

2 3

ft ft tu u r+ implies that penalty forces must be applied in order to prevent the fluid from

penetrating the tank walls. In this case, the penetration can be evaluated as

( ) ( )2 2

2 3

ft ft tu u r = + − . The unit normal n at the fluid/tank contact point can be defined as

30

( ) ( )2 2

2 3 2 30T

t ft ft ft ftu u u u = +

n A . The location of the contact point on the tank wall with

respect to the tank coordinate system can be defined as 1 0 0T

t t t ftr u = + u n A . This equation

can be used to define the global position vector of the contact point on the tank as t t t= +r R u .

The relative velocity vector can be defined as ft f t

r = −v r r . The components of the relative velocity

between the fluid and the tank points along the normal vector and the tangent plane at the contact

point can be defined, respectively, as ft T ft

rn rv =n v and ( )ft ft T ft

rt r r= −v v n v n . If the magnitude of the

tangential relative velocity ft

rtv is larger than zero, one can define the unit tangent vector

ft ft

rt rt=t v v . The magnitude of the penalty normal contact force can be defined as

pn p pF k C = + , where pk and

pC are penalty stiffness and damping coefficients (Wei et al.,

2015). The penalty force vector can then be defined as p pn p pnF F= − −F n t , where

p is an

assumed friction coefficient between the fluid and the tank walls.

(a) (b)

Figure 2.7. Fluid-tank interaction in the (a) radial and (b) longitudinal direction

31

Knowing the ANCF element j on which the fluid/tank contact point lies, one can develop

an expression for the generalized penalty contact forces associated with the fluid element nodal

coordinates. The virtual work of the penalty force acting on the fluid and tank can be written as

T f T t

p p pW = −F r F r , which can be written as ( )T j j T t t t t

p p pW = − −F S e F R u G θ (Shabana,

2013), where j

S is the element shape function matrix evaluated at the contact point, t

u is the skew

symmetric matrix associated with the vector t

u , and t

G is the matrix that relates the absolute

angular velocity vector t

ω of the tank to the time derivatives of the tank orientation parameters

tθ , that is,

t t t=ω G θ . It follows that the generalized reaction forces exerted on the element j of

the ANCF fluid body can be written as Tj j

ep p=Q S F , while the generalized penalty forces exerted

on the tank and associated with the tank reference coordinates t

R and tθ are given, respectively,

as

,Tt i t t

R p p= − = −F F F G u F (2.6)

A similar procedure can be used to evaluate the interaction forces between the fluid and the rigid

tank in the longitudinal direction. This contact geometry is shown in Fig. 2.7b.

2.6. Vehicle Model Components

The MBS model used in this chapter consists of 21 bodies which have 147 absolute coordinates

because Euler parameters are used to describe the body orientations. These bodies are subjected to

115 constraint equations, leading to a model with 32 degrees of freedom. The 10 tires are modeled

using Pacejka’s brush tire model, which is discussed in Section 2.6.1 (Pacejka, 2006). The four-

bar Ackermann steer axle, which allows the truck to turn, is described in Section 2.6.2. Other

bodies include two rear axles, the cab, the tank, the frame rails, and a ground body. Inertial

32

properties for the individual bodies are listed in Table 2.1; the products of inertia are assumed to

be zero. The wheelbase of this vehicle is 4.064 m and the track width is 1.939 m. The front wheels

are attached to the steer axle using the steering knuckles with revolute joints; this allows the wheels

to rotate about the lateral axis to produce the desired forward motion, and about the vertical axis

to allow the vehicle to turn. For simplicity, the steering knuckles and tie rod arms are modeled as

a single body. The rear wheels are connected to the drive axles with revolute joints allowing

rotation about the lateral axis. The cab and tank are rigidly attached to the frame rails and these

three bodies are assumed in this model to represent the chassis (sprung mass). The capacity of the

tank is roughly 3,000 gallons, which is typical of a medium-duty commercial vehicle that services

residential areas. In order to induce the most extreme sloshing scenarios, the tank is assumed to be

half-filled with water (with viscosity of 0.001 kg/m.s (White, 2011)). The penalty coefficients used

in this chapter to enforce the incompressibility conditions 1jJ = and 0jJ = are 91 10 and

41 10 , respectively. The chassis and axles are connected by the suspension system which is

modeled using linear spring-damper elements as explained in Section 2.6.3. The stiffness and

damping coefficients used in this model are provided in Table 2.2.

Table 2.1. MBS model inertial properties

Component Mass (kg) Ixx (kg.m2) Iyy (kg.m2) Izz (kg.m2)

Wheels 56.7 4.25 7.77 4.25

Front Axle 313 797 85.1 797

Steering Knuckle

and Tie Rod Arm 86.8 3.63 3.43 5.41

Tie Rod 25.0 5.75 0.0077 5.75

Rear Axle 410 202 10.6 202

Cab 6804 3685 5265 6425

Tank 1301 567 2324 2299

Frame Rails 1579 151 3340 3340

Rigid Fluid 5464 1204 11877 12555

33

2.6.1. Brush Tire Model

In this thesis, the forces exerted on the tires by the ground are calculated using the brush tire model

(Pacejka, 2006). Four coordinate systems which define the orientations of the tire body i and the

ground body j are introduced. The ground coordinate system which describes the orientation of

the ground body is given by the matrix j j j j = A i j k , where the columns j

i , jj , and

jk

are unit vectors along the ground coordinate axes j j j

X Y Z . In this chapter, j

A is assumed to be

the identity matrix except in the case of uneven terrains, such as a hill, bumpy road, or inclined

ramp. The three other coordinate systems are used in the tire formulation and are depicted in Fig.

2.8. The tire coordinate system defined by the matrix i i i i = A i j k is rigidly attached to the

center of the tire and rotates with the tire. The matrix i i i i

o o o o = A i j k describes the

intermediate tire coordinate system (ITCS) which is also rigidly attached to the center of the tire,

but does not share the pitch rotation with the tire. The axes of i

oA are defined as

( )i i j i i j i

o =

A j k j j k j . Finally, the contact point coordinate system located at the

contact point between the ground and the tire is defined by the matrix

i i i i i i i j

c c c c o c c = = − A i j k i k i k . Using the ITCS transformation matrix i

oA , the contact

point between the tire and the ground can be defined as i i i i

c o oc= +r R A u , where i

R is the global

position of the ITCS origin and i

ocu is the position of the contact point with respect to the ITCS

origin. The velocity vector of the contact point can then be obtained by differentiating the position

vector with respect to time, and is defined as i i i i

c oc= + r R ω u , where i

R is the velocity of the

ITCS origin, i

ω is the absolute angular velocity vector of the tire defined in the global coordinate

system, and i i i

oc o oc=u A u .

34

Figure 2.8. Brush Tire model coordinate systems

The brush tire model accounts for the normal, lateral, longitudinal, and rotational friction

forces, as well as an aligning torque, also referred to as the spin moment in rail vehicle dynamics.

In this thesis, it is assumed that the material properties are the same in the lateral and longitudinal

directions, which is a simplifying assumption often made in the literature (Pacejka, 2006; Patel et

al., 2016; Pelc, 2007; Bolarinwa et al., 2012; Bruzelius et al., 2013). The normal force in the

contact point coordinate system is calculated simply as 1.5 | |z z z zF K C = − − where K is the

radial stiffness coefficient of the tire, z is the vertical penetration of the tire with the ground, C

is the radial damping coefficient, and z is the rate of change of the penetration. In order to

determine the lateral and longitudinal friction forces, the slip angle and slip ratio vector ξ are

needed. First, the slip velocity sv is defined as 0 0TT i i

s sx sx c rc c rcv v = = v i v j v ,

where rcv is the velocity of the tire with respect to the ground at the contact point. The slip ratio

35

vector ξ is 0 0T T

x y sx r sy rv v v v = = − − ξ , where rv is the tire forward velocity.

The slip angle is determined as ( )1tan y −= . The model parameter is defined as

22 / 3p zc a F = where pc is the tread element stiffness per unit length, a is half the contact

patch length, is the friction coefficient, and zF is the magnitude of the normal force. The

coefficients pc and are specified for both the lateral and longitudinal directions, so both x

and y can assume different values. In this thesis, it is assumed that

, ,p x p yc c= and x y = , so

x y = . The slip angle where the pure sliding starts, sl , is defined as ( )1tan 1sl −= .

The lateral and longitudinal friction forces on tire body i in the contact point coordinate

system are determined by ( ) ( )3sgn 1i

j sj j z jF v F = − − if | | sl and ( )sgni

j sj j zF v F= − if

| | sl for ,j x y= where 1j j j = − is a simplifying model parameter. The aligning torque

i

zM is also calculated in the contact point coordinate system depending on the slip angle , as

( ) ( ) 3sgn 1i

z y z y yM aF = − − if | | sl or 0i

zM = if | | sl in the case of pure sliding. To

determine the moment due to the rotational friction force ( )i

y rM in the intermediate tire coordinate

system, a sinusoidal function is used to smooth the forces near 0y = . This moment is defined as

( ) ( )sgni

y y r zrM F = − if | |y t and ( ) ( )sgn( ) sin | | 2

i

y y r z y trM F = − when | |y t ,

where r is the rotational friction coefficient, and t is assumed to be 101.0 10− . The forces and

moments obtained in this section for the tire can be defined in the appropriate coordinate system

for the inclusion in the Newton-Euler equations that govern the motion of the tire which is treated

in the brush model as a rigid body.

36

2.6.2. Ackermann Steering Mechanism

Figure 2.9. Ackermann steering mechanism

In order for a vehicle to be able to negotiate a turn with minimal tire scrub, the Ackerman steering

mechanism is often used. The Ackermann steering system is a four-bar mechanism which allows

the wheels to be oriented at different angles with respect to the forward direction. The four-bar

mechanism consists of the front axle as the ground link, two tie rod arms, and a tie rod, as shown

in Fig. 2.9. The tie rod is connected to the tie rod arms with spherical joints to avoid over-

constraining the mechanism. By specifying the geometry of the Ackerman mechanism, each of the

front wheel forward velocity vectors remains tangent to a circular arc whose origin is located at

the instantaneous center of rotation of a line element connecting the centers of the two wheels,

thus reducing tire scrub. The linkage geometry can be defined by two equations, cosh r = and

2 sins t r = − , where h is the distance between the axle and the tie rod, r is the length of the tie

rod arm, is the angle between the tie rod arm and the normal to the axle, s is the length of the

tie rod, and t is the length of the axle, as shown in Fig. 2.10. Because the length of the axle and

the wheelbase are known for the vehicle model used in this chapter to be 1.939 m and 4.064 m,

37

respectively, can be calculated as 13.42o. In the optimization study by De-Juan et al. (2012), it

was found that for an axle length of 1.5 m, the optimum tie rod length and tie rod arm length are,

respectively, 1.27 m and 0.3 m. To determine the remaining parameters of the steering mechanism,

it was assumed that the steering mechanism dimensions are proportional to the dimensions of the

optimal mechanism geometry obtained in the study by De-Juan et al. (2012). Using this assumption,

the geometry parameters obtained were r = 0.3878 m, s = 1.7590 m, and h = 0.3772 m.

Figure 2.10. Steering mechanism geometry

2.6.3. Suspension System Design

The suspension system of the truck is modeled using linear spring-damper elements, the

parameters of which are listed in Table 2.2. The spring-dampers are oriented to provide restoring

forces in the longitudinal, lateral, and vertical directions – the vertical springs are used to support

the weight of the chassis, cab, tank, and fluid, while the longitudinal and lateral springs are used

to prevent relative motion in the longitudinal and lateral directions, respectively. The spring-

dampers are located at each end of the three axles, resulting in 18 elements total. The vertical

springs on the rear axles are initially compressed to aid in supporting the weight of the tank and

38

fluid and to minimize oscillations at the beginning of the simulation when the vehicle reaches

equilibrium.

Table 2.2. Suspension parameters

X Direction Y Direction Z Direction

Spring

(N/m)

Damper

(N.s/m)

Spring

(N/m)

Damper

(N.s/m)

Spring

(N/m)

Damper

(N.s/m) 91.25 10 41 10 85 10 41 10 67.5 10 55 10

2.7. Specified Motion Trajectories

In numerical simulations, two methods can be used to produce the MBS motion; the first is to

apply forces on the system components, while the second is to specify motion trajectories using

algebraic constraint equations. The latter approach of using constraint equations is more

appropriate when it is required to precisely follow certain trajectories and correctly capture their

geometry, as is the case in this chapter. In order for the tanker truck model to follow different

specified paths necessary to create the motion scenarios to be investigated in this chapter, trajectory

coordinate constraints must be imposed. Three coordinate systems are used to define the type of

the trajectory coordinate constraint used in this thesis: the global coordinate system XYZ , the

trajectory coordinate system ti ti tiX Y Z , and the body coordinate system

ir ir irX Y Z , as shown in

Fig. 2.11 (Shabana et al., 2008). Six trajectory coordinates i

p are used to specify the motion of a

body, where T

i i ir ir ir ir irs y z = p , is is the arc length along the user-specified

trajectory, iry and

irz are the lateral and vertical displacements of the body with respect to the

specified trajectory, and ir ,

ir , and ir are the three Euler angles describing the relative

rotations of the body coordinate system with respect to the trajectory coordinate system. Because

39

a curve can be completely defined using one parameter, Frenet frame geometry is employed to

write the matrix ti

A that defines the orientation of the trajectory coordinate system in terms of

three Euler angles ti , ti , and ti which can be expressed in terms of the arc length parameter

is as ( )ti ti is = , ( )ti ti is = , and ( )ti ti is = (Shabana et al., 2008). The transformation

matrix ir

A that defines the orientation of the body coordinate system with respect to the trajectory

coordinate system is developed using the three time-dependent Euler angles ( )ir t , ( )ir t , and

( )ir t . Using this description, the global position vector of an arbitrary point on the body can be

written as i i i i

p p= +r R A u , where i

R is the global position of the origin of the body coordinate

system, i ti ir=A A A is the transformation matrix which defines the orientation of the body

coordinate system in the global coordinate system, and i

ppu is the position vector of the arbitrary

point, defined in the body coordinate system.

Figure 2.11. Trajectory constraint coordinate systems

40

While trajectory coordinate constraints can be applied to the translation and/or orientation

of a body, only translational coordinate constraints are needed in this thesis to specify the vehicle

forward motion during straight line motion, rapid lane change, and curve negotiation. A

translational trajectory coordinate constraint on body i can be written as R, ( ) 0i

k kC p f t= − = for

1,2,3k = , where kC is the constraint function corresponding to the trajectory coordinate R,

i

kp ,

R,

i

kp is the thk component of

Ti i ir ir

R s y z = p , and ( )f t is the time-varying function

defining the values of the trajectory coordinate ,

i

R kp . For example, to constrain the vehicle to move

along a specified path with a constant forward velocity sV , the constraint applied to the body

coordinate system of the front axle can be written as 0i i

o sC s s V t= − − = , where i

os is the initial

arc length coordinate. Because in this thesis, the equations of motion are developed using the

absolute Cartesian coordinate formulation, it is necessary to define the trajectory coordinates in

terms of the absolute Cartesian coordinates using the relationship

( ), ,i ir ir i ti ti irs y z = − − =g R R A u 0 , where ir

R is the global position of the body coordinate

system with respect to the trajectory coordinate system and ir

u is the position vector of the center

of mass of the body in the trajectory coordinate system. For a given set of absolute Cartesian

coordinates, this set of nonlinear equations can be solved iteratively to determine the arc length

parameter is as well as the coordinates

iry and irz . The constraint Jacobian matrix associated

with the absolute Cartesian coordinates can also be systematically developed and used to enforce

the constraints at the position, velocity, and acceleration levels. The driving constraint forces that

produce the desired motion can be determined using Lagrange multipliers and the trajectory

coordinate constraint Jacobian matrix (Shabana et al., 2008).

41

2.8. Equations of Motion

The vectors and matrices defined in the previous sections enter into the general formulation of the

equations of motion for the MBS vehicle model, which may include rigid and flexible bodies. The

equations of motion can be written as (Shabana et al., 2008):

r

r

T

r rr

Tee

T

d

=

q

e

s

q e s

M 0 0 C Qq

Qe0 M 0 C

0s0 0 0 C

QλC C C 0

(2.7)

where rM and eM are the mass matrices associated with the rigid and flexible ANCF body

coordinates, respectively, rqC , eC , and sC are the constraint Jacobian matrices associated with

the rigid, elastic, and non-generalized trajectory coordinates, respectively, rq , e , and s are the

accelerations of the rigid reference, elastic, and non-generalized trajectory coordinates,

respectively, λ is the vector of Lagrange multipliers associated with the constraints, rQ and eQ

are the vectors of generalized forces associated with the rigid and elastic coordinates, respectively,

and dQ is a vector resulting from the second time derivative of the vector of constraint equations.

The equations of motion are solved numerically using the Adams-Bashforth integration technique,

and a solution algorithm that ensures that the constraint equations are satisfied at the position,

velocity, and acceleration levels. Because the equations of motion are second order differential

equations, two sets of initial conditions (coordinates and velocities) are required to obtain a unique

solution. The initial coordinates and velocities of the rigid bodies, 0r

q and 0r

q , the initial

coordinates and velocities of the nodes of the flexible ANCF bodies, 0e and 0e , and the initial

non-generalized coordinates and velocities 0s and 0s are user-specified and known at the

beginning of the simulation. The initial velocities of all bodies and nodes vary depending on the

42

maneuver being considered in the numerical simulation. These maneuver scenarios are presented

in Table 2.3.

Table 2.3. Initial velocities

Maneuver Braking Lane Change Curve Negotiation

Initial Velocity

(mph) 55 55 25

2.9. Numerical Results

Figure 2.12. Commercial medium-duty tanker truck model

The dynamic behavior of the tanker truck model, created in the MBS dynamics software

SIGMA/SAMS and shown in Fig. 2.12 was examined using three different motion scenarios, each

of which produces different fluid motion. In the first scenario, the truck decelerates under straight-

line motion, such that the fluid primarily exerts longitudinal forces on the tank. In the second

scenario, the truck performs a lane change, which causes the fluid to exert alternating lateral forces

on either side of the tank. In contrast, in the third scenario the truck is assumed to negotiate a wide

43

curve, such that the lateral motion of the fluid is continuous and exerted on one side of the tank

only. Steady state is achieved before the simulation results are reported in order to eliminate the

transient effects. Another model in which the fluid is represented by a rigid body with equivalent

inertial properties was also created; the rigid fluid body is rigidly attached to the tank so that the

sloshing motion is prevented but the fluid inertia is correctly accounted for. This model was also

examined using the same three scenarios so that the effect of the fluid sloshing on vehicle dynamics

can be isolated and evaluated.

2.9.1. Straight Line Deceleration Scenario

Figure 2.13. Velocity during braking

For this scenario, the truck begins at a highway speed of 55 mph and brakes to 20 mph as seen in

the velocity - position plot in Fig. 2.13. This scenario could occur if braking is suddenly applied

in an attempt to avoid a rear-end collision. As seen in Fig. 2.14, the sloshing phenomenon is clearly

evident as the fluid moves in the longitudinal direction towards the front end of the tank as a result

44

of the sudden braking. It is important to note that the section on the top of the tank is an exterior

section that the fluid cannot enter (refer to Fig. 2.3); therefore, the fluid fills the front portion of

the tank during braking.

Figure 2.14. Fluid sloshing due to braking

Figure 2.15. Normal force on a front tire and a rear tire during braking

( Rigid model front tire, ANCF model front tire,

Rigid model rear tire, ANCF model rear tire)

45

Figure 2.16. Position of fluid center of mass relative to tank during braking

( Longitudinal direction, Lateral direction, Vertical direction)

The normal forces increase on the front tires and decrease on the rear tires, as evident in

Fig. 2.15, because the center of mass of the fluid moves towards the front of the tank, as seen in

the plot of the fluid center of mass presented in Fig. 2.16. A second peak occurs in the contact

forces on the front tires near 7s due to the rebounding motion of the fluid after it impacts the rear

of the tank and again sloshes longitudinally towards the front of the tank. Although the normal

forces also increase on the front tires and decrease on the rear tires in the equivalent rigid body

model due to the shift in inertia of the chassis on the suspension system, also seen in Fig. 2.15, the

magnitudes are much less significant because the center of mass of the rigid fluid body remains

constant with respect to the tank and the relative displacement of the chassis on the suspension

system is small compared to the longitudinal displacement of the flexible fluid. The fluid free

surface of the ANCF model returns to a flat shape once the truck reaches the lower speed, as seen

46

in Fig. 2.17, and the normal forces return to approximately equilibrium in both models, as evident

by the results presented in Fig. 2.15.

Figure 2.17. Flat free surface at steady state after braking

2.9.2. Lane Change Scenario

Figure 2.18. Lane change trajectory

47

In the second motion scenario, the truck performs a lane change over a standard-width US highway

lane of 3.7m, as seen in the plot of lateral - longitudinal position in Fig. 2.18. The lane change is

completed in a relatively short time of 4s so that the fluid sloshing readily occurs, as seen in the

series of images in Fig. 2.19 depicting the change in the free surface of the fluid mesh. The shift

in the center of mass of the flexible fluid mesh as the truck negotiates the turns of the lane change

causes the normal forces exerted on the tires on the outer edge of the curve to be greater than those

exerted on the inner tires, as seen in Fig. 2.20. For the first half of the lane change, the left tire (Fig.

2.20a) is the inner tire, and for the second half of the lane change, the right tire (Fig. 2.20b) is the

inner tire. This effect is also evident in the rigid body model due to the outward inertia forces,

however the peak forces exerted are greater for the ANCF model than for the rigid body model

due to the sloshing behavior, which is evident in the plot of the fluid center of mass presented in

Fig. 2.21. Furthermore, it can be seen that the peaks in the normal forces on the outer tires of the

rigid body model are the same after both the first and second halves of the lane change – 8 kN on

the inner tire and 20.5 kN on the outer tire in both cases. However, this is not the case for the

ANCF model tires – the changes in the normal forces from equilibrium are greater after the second

half of the lane change (5.5 kN and 22.5 kN) as compared to the first (6.5 kN and 21 kN). This is

because the forces of the tank walls on the fluid during the second half of the lane change act in

the same direction as the motion of the rebounding fluid, causing the lateral shift of the fluid to be

larger than would have occurred due to free vibration only. After the lane change is completed, the

normal forces on the tires of the ANCF model oscillate about a nominal value due to the lateral

sloshing of the fluid, whereas the normal forces remain at steady-state after the rigid fluid model

negotiates the lane change because the lateral position of the fluid is fixed with respect to the tank.

48

Figure 2.19. Lateral sloshing due to lane change maneuver

Figure 2.20. Normal force on (a) a left-hand tire and (b) a right-hand tire during a lane change

( Rigid model, ANCF model)

49

Figure 2.21. Position of fluid center of mass relative to tank during lane change

( Longitudinal direction, Lateral direction, Vertical direction)

Figure 2.22. Lateral friction force on (a) a left-hand tire and (b) a right-hand tire during a lane

change

( Rigid model, ANCF model)

50

These effects are also apparent in the lateral friction force results presented in Fig. 2.22.

The peak lateral friction force on the inner tire is comparable between the ANCF model and the

rigid body model because although the normal force is less for the ANCF model tire due to the

outward shift of the fluid, the lateral velocity of the tire with respect to the ground is greater for

the ANCF model tire, which negates the effect of the decreased normal force. This is clear from

the results presented in Fig. 2.23, where the lateral slip velocity of a left tire is greater for the

ANCF model during the lane change.

Figure 2.23. Lateral slip velocity on a left-hand tire during a lane change

( Rigid model, ANCF model)

2.9.3. Curve Scenario

For the third scenario, the truck negotiates a wide curve as seen in Fig. 2.24, similar to an onramp

or exit ramp of a highway, except that the bank angle is assumed zero for simplicity. While driving

along a road of constant curvature and in the case of zero bank angle, the outward centrifugal force

on a rigid vehicle due to the curve of the road is counteracted by the lateral friction force exerted

51

on the tires; that is, 2

smV r mg= where m is the mass of the vehicle, sV is the forward velocity,

r is the radius of curvature of the road, is the coefficient of friction between the tires and the

road, and g is the gravitational constant.

Figure 2.24. Curve trajectory

As previously discussed in this paper, the maximum speed at which the vehicle can traverse

a curve without sliding can then be calculated as sV gr= . For example, for a radius of curvature

of 115 ft and a coefficient of friction of 0.7, the maximum calculated speed is 33.9 mph. However,

due to the high center of gravity of the truck and increased chance of rollover, in practice vehicles

cannot traverse a curve at the theoretical maximum speed. According to the National Highway

Traffic Safety Administration (NHTSA), the maximum speed at which an average fully-loaded

tractor-trailer can negotiate a curve of 150 ft is 30mph; at greater speeds, the chance of rollover is

greatly increased (U.S. Department of Transportation, NHTSA, 2015). Therefore, for this analysis,

the constant forward speed is chosen to be 25 mph and the radius of curvature of the track is set to

52

150 ft. As previously mentioned, the road is assumed to be flat with no super-elevation; this

assumption is consistent with methods used in the literature (Pape et al., 2016).

Figure 2.25. Normal force on an outer tire and an inner tire during curve negotiation

( Rigid model outer tire, ANCF model outer tire,

Rigid model inner tire, ANCF model inner tire)

Figure 2.26. Lateral friction force on (a) an outer tire and (b) an inner tire during curve

negotiation

( Rigid model, ANCF model)

53

The normal and lateral forces exerted on the tires of the rigid body model are larger on the

outer tires and smaller on the inner tires, as seen in Figs. 2.25 and 2.26, respectively. This is due

to the roll moment that is exerted from the centrifugal force on the vehicle. Because the radius of

curvature is constant and the center of mass of the fluid in the rigid body model is not able to move,

the centrifugal force is constant once the truck enters the curve, and thus the normal and lateral

forces are constant as well. The contact forces on the tires of the ANCF model, however, overshoot

the constant value exerted on the rigid body model tires, and oscillate due to the sloshing motion

and the change in the center of mass of the fluid, as seen in Figs. 2.25 and 2.26. Furthermore, while

the centrifugal force on the rigid fluid is 2

smV r as discussed in Section 2.2, it has been

demonstrated that this is not the case for flexible bodies (Shi et al., 2017). This is evident in Fig.

2.27, where the outward inertia force on the flexible fluid mesh oscillates with a maximum

amplitude that exceeds the nominal rigid body model value by nearly 16%. This oscillation is due

to the changing location of the center of mass of the fluid – while it is constant relative to the

vehicle for the rigid body model, the sloshing phenomenon occurring in the flexible model results

in oscillation of the center of mass, and thus the effective radius of curvature changes as the vehicle

negotiates the curve. The sloshing amplitudes and thus the inertia and contact forces decrease with

time for the ANCF model due to the fluid viscosity and the friction forces between the fluid and

tank walls. In order to better understand the results presented in Fig. 2.27, Fig. 2.28 shows the

position of the center of mass of the liquid with respect to the tank. This figure shows that because

of the liquid oscillations, the simple equation 2

smV r used to calculate the centrifugal force in

rigid body dynamics is no longer applicable in the case of liquid sloshing. Figure 2.29 shows the

components of the normalized velocity of the liquid center of mass obtained by dividing by the

54

vehicle forward velocity. Figure 2.29a shows the dimensionless velocity component tangent to the

curve, while Fig. 2.29b shows the other two components.

Figure 2.27. Outward inertia force on fluid during curve negotiation

( Rigid model, ANCF model)

Figure 2.28. Position of fluid center of mass relative to tank during curve negotiation

( Longitudinal direction, Lateral direction, Vertical direction)

55

Figure 2.29. Normalized velocity of the fluid center of mass in the (a) longitudinal and (b) lateral

and vertical directions

( Longitudinal direction, Lateral direction, Vertical direction)

The magnitudes of the contact forces are not identical to those predicted by the analytical

model in Section 2.1 because only two wheels were included in that analysis, and the weight of

the tank and fluid is actually distributed over 10 wheels. However, the orders of magnitude of the

contact forces are the same and the relative changes in the forces were well predicted, and thus the

analytical model verifies the simulation results.

2.10. Concluding Remarks

A total Lagrangian ANCF fluid formulation that can be systematically integrated with fully

nonlinear MBS vehicle algorithms is proposed in this chapter. The new approach can capture the

fluid distributed inertia and viscosity, can accurately predict the change in inertia due to the change

in shape of the fluid, and can visualize the change in the fluid free surface, unlike other discrete

inertia models which do not capture these significant details. The outward forces on the fluid

56

during curve negotiation are derived and it is shown that these forces do not take the same simple

form as the case of a rigid body negotiating a curve. As discussed in the paper, accurate modeling

of the fluid geometry using ANCF elements can be achieved without the need for using

computational geometry methods such as B-spline and NURBS representations which have a rigid

recurrence structure unsuitable for MBS analysis. By using the approach proposed in this chapter,

one geometry/analysis mesh is used from the outset. The fluid constitutive law and the fluid/tank

interaction forces are developed. The penalty method is used to ensure that the fluid remains within

the boundaries defined by the tank geometry. Both normal and tangential penalty contact forces

are considered in this chapter. The MBS vehicle model components are described and the

dynamics of the vehicle is examined using three contrasting motion scenarios in order to study the

effect of sloshing on vehicle dynamics. The braking scenario examines the case of longitudinal

sloshing of the fluid, a rapid lane change produces alternating lateral fluid forces on the tank, and

curve negotiation sheds light on the case of steady-state outward forces due to the centrifugal effect.

The results presented in this chapter demonstrate that depending on the scenario, the

sloshing phenomenon can increase the contact forces on some wheels while decreasing contact

forces on other wheels, and this can lead to vehicle instability. In the case of brake applications,

the fluid in the partially-filled tank surges forward, causing uneven wheel loading; in cases of

severe braking, wheel lift may occur on the rear wheels of the vehicle. This can lead to difficulty

controlling the vehicle and increased stopping distances due to lessened road contact, and the

decrease in stability may result in jack-knifing for tractor-trailer vehicles. When entering a curve

or performing a lane change, the outward centrifugal forces cause lateral displacement of the fluid,

which also causes uneven wheel loading. In extreme cases of high speed or small radius of

curvature, these changes could be significant enough to induce wheel lift on the inner wheels and

57

increase the possibility of rollover compared to an equivalent truck carrying rigid materials.

Furthermore, in cases where tire friction forces are decreased, such as on wet or icy roads, the

possibility for vehicle instability increases even further and drivers must exercise extra caution.

Future analysis can result in defining general rules for speed reduction of a tanker truck entering a

curve in order to maximize vehicle stability and driver safety, proposing modifications to the tank

geometry to reduce sloshing amplitudes and forces, studying the effect of viscosity and

incompressibility on the fluid dynamic behavior, and comparison with conventional finite

elements as well as the smoothed particle hydrodynamics (SPH) method (Wasfy et al., 2014).

58

CHAPTER 3

FLUID MODELING WITH RAILROAD VEHICLE APPLICATIONS

A new continuum-based liquid sloshing approach that accounts for the effect of complex fluid and

tank-car geometry on railroad vehicle dynamics is developed in this chapter (published as Shi et

al., 2017). A unified geometry/analysis mesh is used from the outset to examine the effect of liquid

sloshing on railroad vehicle dynamics during curve negotiation and during the application of

electronically controlled pneumatic (ECP) brakes that produce braking forces uniformly and

simultaneously across all cars. Using a non-modal approach, the geometry of the tank-car and fluid

is accurately defined, a continuum-based fluid constitutive model is employed, and a fluid-tank

contact algorithm is developed. The liquid sloshing model is integrated with a three-dimensional

multibody system (MBS) railroad vehicle algorithm which accounts for the nonlinear wheel/rail

contact. The three-dimensional wheel/rail contact force formulation used in this study accounts for

the longitudinal, lateral, and spin creep forces that influence the vehicle stability. In order to

examine the effect of the liquid sloshing on the railroad vehicle dynamics during curve negotiation,

a general and precise definition of the outward inertia force is defined, and in order to correctly

capture the fluid and tank-car geometry, the absolute nodal coordinate formulation (ANCF) is

used. The balance speed and centrifugal effects in the case of tank-car partially filled with liquid

are studied and compared with the equivalent rigid body model in curve negotiation and braking

scenarios. In particular, the results obtained in the case of the ECP brake application of two freight

car model are compared with the results obtained when using conventional braking. The traction

analysis shows that liquid sloshing has a significant effect on the load distribution between the

59

front and rear trucks. A larger coupler force develops when using conventional braking compared

with ECP braking, and the liquid sloshing contributes to amplifying the coupler force in the ECP

braking case compared to the equivalent rigid body model which does not capture the fluid

nonlinear inertia effects. Furthermore, the results obtained in this chapter show that liquid sloshing

can exacerbate the unbalance effects when the rail vehicle negotiates a curve at a velocity higher

than the balance speed.

3.1. Basic Inertia Force Definitions

A rail vehicle can safely negotiate a curve if the outward inertia force does not exceed the sum of

the lateral gravity force component and the inward friction force. However, as mentioned

previously, the centrifugal force of a flexible body does not take the simple form of 2mV R , where

m is the mass of the vehicle, V is the forward velocity, and R is the radius of curvature of the

curve (Shi et al., 2017). As presented in Chapter 2, a straightforward method to determine the

outward inertia force in the case of flexible body dynamics is to use the projection of the inertia

force vector on the outward normal to the curve, which has the form i

i i i

VdV r n in the case of

a flexible body i , where i and

iV are, respectively, the mass density and volume of the flexible

body, i

r is the acceleration vector, and n is the outward unit normal to the curve. This inertia

force expression is general and includes the effect of other deformation-dependent forces such as

gyroscopic moments and Coriolis forces.

60

3.1.1. FFR Inertia Forces

The form of the inertia forces depends on the method used to formulate the kinematic and dynamic

equations. When ANCF elements are used, the inertia forces take a simple form and the mass

matrix becomes constant. While ANCF elements will be used in this chapter, another widely used

formulation, the floating frame of reference (FFR), is used in this section to shed light on the form

of the inertia forces in the case of curve negotiation and to show that the rigid body assumption

leads to the definition of the centrifugal forces used in rigid body dynamics. To this end, a simple

planar example is used in this section (Shi et al., 2017).

Unlike the ANCF description, in the FFR formulation, a flexible body coordinate system

is introduced and the motion of a planar body i in the system is defined using two coupled sets of

coordinates, the reference coordinates T T

i i i

r =

q R and the elastic coordinates i

fq , where i

R

describes the body reference translation, i defines the reference orientation, and i

fq defines the

body deformation with respect to its reference. In the FFR formulation, there is no separation

between the rigid body motion and the elastic deformation, and therefore, the FFR description does

not imply any simplifying assumptions. The generalized coordinates for a planar deformable body

i can then be written as T T T

i i i i

f =

q R q . Using these generalized coordinates, the global

position vector of an arbitrary point on the deformable body can be written as

( )i i i i i i i i i

o f= + = + +r R A u R A u S q (3.1)

where iu is the local position vector defined in the body coordinate system, and

iA is the

transformation matrix that defines the body orientation and is expressed in terms of the angle i .

The local position vector iu can be written as i i

o f+u u , in which i

ou is local position vector of the

61

arbitrary point in the undeformed state and i

fu is the deformation vector which can be written

using the technique of the separation of variables as i i

fS q in which i

S is a space-dependent shape

function matrix. The acceleration vector can be derived by differentiating the position vector twice

with respect to time as

( )2

2i i i i i i i i i i i i i i i

f f = + + − +r R A u A S q A u A S q (3.2)

Substituting this equation into the inertia force expression, one obtains

( ) ( ) ( )2

1 1 2i

i i i i i i i i i i i i i i i i i i i i i i

f f f fV

dV m = + + + − + + r R A I S q A S q A I S q A S q (3.3)

where 1 i

i i i i

oV

dV= I u , i

i i i i

VdV= S S , and in the case of planar motion, the transformation

matrix and its partial derivative are given, respectively, as

cos sin sin cos,

sin cos cos sin

i i i i

i i

i i i i

− − −= =

− A A (3.4)

Figure 3.1. A planar flexible body negotiating a curve

If the flexible body negotiates a circular curve with a constant forward velocity, as shown

in Fig. 3.1, the motion constraints are defined as T

sin cosi i iR − − = R 0 , where R is the

62

radius of curvature. In this special planar case, the unit outward normal to the curve takes the form

T

sin cosi i = − n . If the arc length traveled by the reference point is defined as is , then the

constraints at the acceleration level are written as

( )2

cos sin

sin cos

ii i

i i

i i

ss

R

−= +

R (3.5)

in which the identities i is R = and i is R = are used. Using the preceding equations with the

outward inertia force i

i i i i

VF dV= r n , and assuming a constant forward velocity (that is, 0is = ),

one obtains

( )( )

2 2T T T

1

0 0 12

1 1 0

i i ii i i i i i i i i

f f f

s s sF m

R R R

− = − + − + +

− − S q I S q S q (3.6)

In the case of steady state motion, where i

f =q 0 and i

f =q 0 , the preceding equation

reduces to ( ) ( ) ( )22 T

10 1i i i i i i i

fF m s R = − + +I S q , which shows that, even when the time

derivatives of the elastic coordinates are zeros, the outward inertia force of a deformable body

depends on the deformation and differs from ( )2

i im s R used in rigid body dynamics. In the case

of a rigid body with a centroidal body coordinate system, 1

iI and i

fq vanish, and iF reduces to

( )2

i i iF m s R= − , which demonstrates clearly that the centrifugal force in the case of rigid body

dynamics is a special case of the more general expression used in flexible body dynamics. The

FFR analysis presented in this section sheds light on the fundamental differences between the

inertia force definitions used in rigid and flexible body dynamics. These fundamental differences

must be considered in the case of liquid sloshing in railroad vehicles which experience large

displacements.

63

3.1.2. ANCF Inertia Forces

In this chapter, three-dimensional ANCF elements are used in the analysis of liquid sloshing, and

therefore, the general expression of the outward inertia force T

i

i i i

VdV r n will be used, as

discussed in the previous chapter. The displacement field of an ANCF element j is defined in the

global coordinate system as ij ij ij=r S e , where

ijS is the element shape function matrix and

ije is

the vector of the ANCF element nodal coordinates. Because in the ANCF kinematic description,

a body (structure) coordinate system is not used, direct comparison with rigid body dynamics

cannot be easily made as in the case of the FFR formulation. Nonetheless, one can show the

equivalence of the ANCF and FFR kinematic description. One can also show that

T T

i

i i i i i

cV

dV m = r n r n , in which im is the total mass of the ANCF flexible body and i

cr is the

acceleration vector of the body center of mass. The constant mass matrix of element j of the

ANCF flexible body i is defined as T

ij

ij ij ij ij ij

VdV= m S S , where

ijS is space-dependent shape

function matrix, ij is the element density, and ijV is the element volume (Shabana, 2017A). The

position vector of the center of mass can be written as ( )1

eni ij ij i

c jm

== r S e , where

1

eni ij

jm m

== ,

ijm is the mass of element j , ij

ij ij ij ij

VdV= S S , and en is the total number of elements. It

follows that ( )1

eni ij ij i

c jm

== r S e . In order to define the outward inertia force for the liquid body,

the unit outward normal i

n to the curve should also be defined.

3.2. Integration of Geometry and Analysis for Railroad Sloshing

In railroad vehicle system applications, accurate definition of the liquid/tank geometry and

wheel/rail geometry, shown in Fig. 3.2, is necessary for thorough investigation of the sloshing

64

effect. In this chapter, ANCF elements are used to describe both the track and liquid/tank geometry.

The track geometry is described using ANCF beam elements, while the liquid and tank are

modeled using ANCF solid elements. The wheel is modeled as a surface of revolution, and

therefore, no FE discretization is required. The procedure described in this section allows for the

use of a unified geometry/analysis mesh from the outset for the study of the liquid sloshing as well

as the wheel/rail contact.

Figure 3.2. Wheel/rail contact

3.2.1. ANCF Track Geometry

The wheel/rail contact forces that define the vehicle stability depend on the geometry of the wheel

and rail profiles. The wheels and rails can be modeled as rigid or flexible bodies depending on the

focus of the investigations. The track can be tangent (straight line) or curved; curved tracks are

formed using constant radius and spiral segments. The spiral sections are designed to have a

curvature that varies linearly along the spiral arc length, thereby allowing smoothly joining a

tangent track segment with a circular segment. Figure 3.3 shows a three-dimensional fully

65

parameterized ANCF beam element used in this investigation to describe the geometry of a curved

track segment. The geometry of the rail segment is defined by the geometry of the space curve and

the profile geometry. The three-dimensional fully parameterized ANCF beam element used in this

study has at each node 12 coordinates that contain positions and position vector gradients; that is,

for a node k , the vector of coordinates is defined as T T T T T

, 1, 2ijk ijk ijk ijk ijk

x y z k = =

e r r r r ,

where , , ,ijk ijk x y z = =r r , ijk

r is the global position vector at node k , and ,x y , and z are

the element spatial coordinates (Shabana, 2017A). In the case of a fully parameterized beam

element, the parameters ,x y , and z are independent and can be used to define the three

independent position gradient vectors , , ,ijk ijk x y z = =r r . In railroad vehicle dynamics, the

profile of the rail is measured using a device called a mini-prof that produces cubic spline data

which define the profile geometry. Therefore, the surface of the rail can be described using the

parametric expression ( )y f z= . If the profile geometry changes along the rail space curve, the

more general parametric equation ( ),y f x z= can be used. The profile geometry defined by the

parametric equation ( ),y f x z= can be integrated systematically with the fully parameterized

ANCF beam element to define the rail surface geometry at the contact points. The rail surface

geometry is used in the numerical solution algorithm to define the location of the wheel/rail contact

points, the velocity creepages, and the creep contact forces. The definition of these kinematic and

force variables requires the definition of the tangent plane and the normal vector to this plane. If

rs defines the rail arc length and y defines the lateral rail parameter, one can define the

longitudinal and lateral tangent vectors at an arbitrary point on the rail surface using the ANCF

kinematic equations ( )( )r

ij ij r

sx x s= r r , and

66

( ) ( )( ) ( )( )ij ij ij ij

y y x x y z z y= + + r r r r , respectively. The unit normal vector to the

rail surface that corresponds to element j can be defined as ( )r r

ij ij ij ij ij

y ys s= n r r r r . If the

rail is assumed rigid, the nodal coordinates of the element are constants and assume their initial

values. If the rail is assumed flexible, the nodal coordinates will change with time in response to

the wheel/rail contact forces. Therefore, the ANCF geometry description presented in this section

can be applied to both rigid and flexible rails. However, because the focus of this chapter is on

railroad liquid sloshing, the rail is assumed to be rigid.

Figure 3.3. Curved ANCF rail element

3.2.2. Liquid/Tank Geometry

Figure 3.4. Fluid and tank geometry

67

In this section, an initially curved ANCF fluid, shaped according to the rail tank-car geometry, is

modeled using fully parameterized ANCF solid elements. The tank is assumed to consist of a

cylinder with half-ellipsoid ends, as shown in Fig. 3.4. The tank-car and the fluid geometries enter

into the definition of the fluid/tank contact forces formulated in this chapter using a penalty method

in which both the normal and friction forces are considered. Because the liquid has relatively larger

deformation than the tank, the tank is assumed to be rigid.

Figure 3.5. Cross-section mesh of the fluid inside a cylindrical tank

In order to define an initially curved fluid geometry/analysis mesh consistent with the

geometry of the railroad tank which consists of a cylinder and two half-ellipsoid ends, it is required

that the fluid mesh at the boundary has the same curved shape as the tank. Wei et al. (2015)

demonstrated that fewer ANCF fluid elements can describe the fluid motion compared with the

FFR formulation. The solid element used in this investigation is a fully parameterized ANCF

68

element with 8 nodes; each node has 12 coordinates, T T T T T

, 1, 2, ,8ijk ijk ijk ijk ijk

x y z k = =

e r r r r

[16, 26]. The initially curved ANCF solid elements are used to model the fluid inside the tank with

a cross-section geometry defined by eight elements, as shown in Fig. 3.5, where in this figure, r

is the radius of the cylindrical tank, h is a measure of the height of the liquid free surface, and the

angular parameters , , and are used to determine nodal positions and gradients. In Fig.

3.6, the nodes and element numbers are labeled such that the nodes with solid circles represent the

element master nodes used to define the element dimensions; examples of master nodes are shown

where node 1 in Fig. 3.6a is the master node for the straight element and node 2 in Fig. 3.6b is the

master node for an initially curved element. The element dimensions in the reference configuration

are assumed a , b , and c as shown in Fig. 3.6b.

Figure 3.6. ANCF solid element in the (a) curved reference and (b) straight configurations

69

The ANCF gradient vectors can be conveniently used for efficient shape manipulation in

order to accurately define the fluid geometry; for example, if there is no stretch or change of shape

at a node of the fluid, the gradients will assume values that correspond to the straight configuration,

that is, T

1 0 0 for the first gradient vector ijk

xr , T

0 1 0 for the second gradient vector ijk

yr ,

and T

0 0 1 for the third gradient vector ijk

zr . In the case of an element that has a reference

configuration different from the straight configuration, as in Fig. 3.b, the gradients can be adjusted

to properly define the desired geometry. For example, referring to the geometry of the fluid mesh

in Figs. 3.5 and 3.6, the gradients in the reference configuration can be obtained as

( )4

670 1 0z l c= r , ( )T5

160 sin cosy l b = − − r , and

( )T8

160 cos siny l b = − − r , where the superscript ij is dropped for simplicity, the number

superscript refers to the node number, and the angles and and the arc lengths 67l and 16l can

be determined according to the free surface height h and radius of the cylinder r . Figure 3.7

shows the complete mesh of the fluid inside a tank; the mesh has 75 nodes, 32 elements, and a

total of 900 degrees of freedom.

Figure 3.7. Initially curved ANCF fluid mesh

70

3.3. Fluid/Tank Interaction Forces

The penalty method is also used in this chapter to formulate the fluid/tank interaction forces that

produce the sloshing oscillations. The boundary surfaces of the fluid mesh are regarded as the

potential contact surfaces and points on these surfaces are monitored throughout the simulation in

order to determine the contact points. It will be explained later in this section how the contact

points are identified in the case of the cylindrical tank and also in the case of the tank ellipsoidal

ends whose geometry is important, particularly in the case of sudden braking application.

3.3.1. Normal Contact Force

The penalty forces, which include normal and tangential friction forces, are applied on both the

fluid and the tank bodies only when interpenetration occurs. A contact frame at the contact point

is introduced in order to define the normal and tangential forces along the axes of this contact

frame. Knowing the relative penetration , and its time rate , between the fluid ANCF element

and the tank at the contact point, the normal contact force can be evaluated using the formula

1.5

nf K C = − − , where K and C are the penalty coefficients associated with the penetration

and the penetration rate, respectively, and represents the absolute value. In the expression used

in this investigation for the normal force, the exponent on the penetration in the stiffness term was

chosen to be 1.5 to increase smoothness near zero penetration. Other force models, including a

linear relationship, can also be used. The absolute value term is included in the damping term to

ensure that the normal force is equal to 0 when there is no penetration. It follows that the tangential

friction force can be written as t nf f= , where is the coefficient of friction between the fluid

and tank at the interface. Determining the friction coefficient between a fluid and solid surface is

not a trivial matter, and is not the focus of this work. It is a function of the texture of the solid

71

surface as well as the viscosity of the fluid, and is highly sensitive to changes in the liquid-solid

interface (Petravic, 2007; Pit et al., 1999). A relatively large value of 0.5 = was chosen to reduce

relative motion between the fluid and solid surfaces and approximate the no-slip condition

characteristic of viscous Newtonian fluids (White, 2011).

3.3.2. Relative Position

Figure 3.8. Tank geometry and coordinate systems

The position vector of a potential contact point P on ANCF solid element j of the fluid body f

can be written as fj fj fj

P P=r S e , where fj

PS is the shape function matrix evaluated at point P , and fj

e

is the vector of nodal coordinates of the ANCF element j . If the global position vector of the

origin of the coordinate system of the rigid tank body t is defined as t

R , the relative position and

velocity vectors of the potential contact point on the fluid with respect to the tank can be written

as ft f t

P P= −u r R and ft f t

P P= −u r R , respectively. In order to define the penetration and the

penetration rate , the relative position vector ft f t

P P= −u r R is defined in the local tank body

72

coordinate system as ( )T Tft t ft t f t

P P P= = −u A u A r R , where tA is the transformation matrix that

defines the orientation of the tank coordinate system in the global coordinate system. Similarly,

the relative velocity between the contact points on the fluid and tank bodies can be written as

( )Tft t ft t ft

Pr P P= −v A u ω u , where t

ω is the skew-symmetric matrix that defines the tank absolute

angular velocity vector t

ω . The relative position and velocity vectors at the potential contact point

can be used to define the penetration and its rate . The origin of the body coordinate system

of the tank is chosen to be at the tank geometric center, as shown in Fig. 3.8. Using the symmetry

of the tank, the tank can be divided into two geometry sections, the cylindrical and ellipsoidal

sections. The cylindrical section has length L and radius r , while the three axes of the half-

ellipsoid are defined as a , b , and c , and satisfy the relationship b c r= = . Two local coordinate

systems, t t t

c c cx y z and t t t

e e ex y z , are introduced for the cylindrical and ellipsoidal sections, respectively,

for the convenience of defining the normal and tangential contact forces at the contact point.

3.3.3. Cylindrical Region

In the case that the contact occurs in the cylindrical section of the tank, the normal vector at the

contact point t

Pn is simply directed to the tank center and can be defined as

T

2 30t ft ft

P P Pu u = − − n , where , 1,2,3ft

Plu l = , are the three components of the vector ft

Pu defined

in the tank cylindrical section coordinate system t t t

c c cx y z as shown in Fig. 3.8. The unit normal

vector at the contact point Tˆ t t t t

P P P P=n n n n can be used to define the tangential relative velocity

vector as ( ) ( )ˆ ˆft ft ft t t

Pr Pr Pr P Pt= − v v v n n . A unit vector along the tangential relative velocity can be

defined as ( ) ( )t ft ft

P Pr Prt t

=t v v . Using these definitions, the penetration and penetration rate can

73

be defined, respectively, as Tt t

P P r = −n n and ˆft t

Pr P = v n . If 0 , the normal contact and

friction forces at the contact point can be evaluated, respectively, as 3/2

nf K C = − − and

t nf f= , respectively. Therefore, the penalty force vector can be written as ˆ t t

P n P t Pf f= −f n t . This

penalty force vector can be defined in the global coordinate system as t

P P=F A f . The generalized

contact forces exerted on element j of the ANCF fluid body can be defined using the virtual work

and can be written as T Tfj fj fj t

P P P P P= =Q S F S A f . In this equation, fj

PQ is the vector of generalized

forces associated with the ANCF nodal coordinates of the fluid element j . The resultant contact

forces on the rigid tank are equal in magnitude but opposite in direction to the forces exerted on

the fluid. The generalized contact forces associated with the generalized coordinates of the rigid

tank is T T T Tt t t t

P P P P = − − Q F F A u G , where

tG is the transformation matrix which relates the

angular velocity vector to the time derivatives of the orientation parameters, t t t=ω G θ , t

Pu is the

skew matrix associated with the vector t

Pu which defines the contact point on the tank in the tank

coordinate system and can be written as T

1ˆ 0 0t t ft

P P Pr u = + u n , t

ω is the absolute angular

velocity vector of the tank reference defined in the tank coordinate system, and tθ is the set of

parameters used to define the orientation of the tank coordinate system. In this thesis, Euler

parameters are used as the orientation coordinates.

3.3.4. Ellipsoidal Region

In the case that the contact occurs in the ellipsoidal sections of the tank, which can be determined

by evaluating 1 2ft

Pu L , the relative position and velocity vectors of the contact point can be

defined with respect to the coordinate system t t t

e e ex y z which is located at the ellipsoid center as

74

shown in Fig. 3.8. By introducing the analytical expression of an ellipsoid, one can check if the

condition ( ) ( ) ( )2 2 2

1 2 3 1ft ft ft

P P Pu a u b u c+ + is satisfied to determine if a contact occurs between

the fluid and the half-ellipsoids. Assuming that the position vector of the contact point on the tank

wall t

Pu with respect to the origin of the tank coordinate system is parallel to the vector ft

Pu which

defines the location of the contact point on the fluid in the same coordinate system, one can

calculate t

Pu by using the ellipsoid geometry function. The normal vector at the potential contact

point can be written as T

2 2 2

1 2 32t ft ft ft

P P P Pu a u b u c = − n . Having determined the normal

vector, a procedure similar to the one used for the tank cylindrical section can be used to determine

the normal and tangential velocity components as well as the penetration and its rate . Using

this information, the normal and tangential friction forces can be calculated and used to determine

the generalized forces associated with the generalized coordinates of the ANCF fluid and rigid

tank bodies.

3.4. ANCF Fluid Constitutive Equations

In order to demonstrate the use of the general procedure proposed in this chapter, an

incompressible Newtonian fluid model is used, where the viscous forces as well as the

incompressibility conditions of the fluid can be formulated systematically based on the Navier-

Stokes equations, as was presented in Chapter 2. The resultant stresses are used to define the

generalized viscous forces of the ANCF fluid element. Using higher-order ANCF solid elements,

fewer elements are needed to model the liquid sloshing compared to the conventional FE method

and the FE/FFR approach (Wang et al., 2015). These ANCF elements can also accurately capture

75

the initial shape as well as the complex shapes that result from the liquid sloshing as previously

explained in this thesis.

In order to consider the initially curved configuration of a fluid element that interacts with

a curved tank surface, the relationships between the volumes in various configurations will be

reviewed. Let V , oV , and v be the volumes in the straight, curved reference, and current deformed

configurations, respectively, and x , X , and r are the corresponding position vectors of an

arbitrary fluid point in these three configurations. The position vectors in the reference and current

configurations are written, respectively, as o=X Se and =r Se , in which S is the shape function

matrix and oe and e are the nodal position vectors defined in the reference and current

configurations, respectively. The relation between the volume in the initially curved reference

configuration and the volume in the straight configuration can be defined as o odV dV= J , where

refers to the determinant of a matrix and oJ is the constant Jacobian matrix of the position

vector gradients and is defined as ( ) :o o o= = = xJ X x Se x S e , in which = xS S x is a

third-order tensor that defines the derivatives of the shape function matrix with respect to the

straight configuration parameters x . The relationship between the volume defined in the current

configuration and the volume in the curved reference configuration can be written as odv dV= J ,

where J is the Jacobian matrix of position vector gradients defined as

( )( ) 1

e o

−= = =J r X r x x X J J , in which ( ) :e = = = xJ r x Se x S e . It follows that

1

o e o o edv dV dV dV−= = =J J J J J J . Therefore, integrations carried out over the initially

curved reference configuration domain can be systematically converted to integrations over the

straight configuration domain, allowing for using the element dimensions defined in the initially

straight configuration throughout the entire simulation regardless of the amount of the fluid

76

displacements. Using the principle of conservation of mass, the density defined in the initial

straight configuration can be used.

3.4.1. Viscosity and Penalty Forces

The penalty method is also used in this chapter to impose the incompressibility condition of the

fluid elements. For an incompressible fluid element j , the determinant of the matrix of position

vector gradients must be equal to one, that is 1j jJ = =J and its first derivative 0jJ = . In this

case, the Navier-Stokes stress relationship reduces to 2j j

dev f=σ D , where j

D is the rate of

deformation tensor, j

devσ is the deviatoric portion of the symmetric Cauchy stress tensor, and f

is the coefficient of shear viscosity (Spencer, 1980; Shabana, 2017A). In this thesis, the mass

density remains constant because of the incompressibility condition, and the effect of temperature

is neglected. In general, the virtual work of the fluid viscous forces can be written in terms of the

second Piola-Kirchoff stress tensor 2

j

Pσ and Green-Lagrangian Strain tensor j

ε since they are

defined in the reference configuration as 1

2: :j j

o

j j j j j j j j

v dev P ov V

W dv dV −

= − = − σ J J σ ε in which

( )T

2j j j= −ε J J I and ( )1 1 T

2

j j j j j

P devJ− −

=σ J σ J . In order to define the fluid viscous forces, the

constitutive model 2j j

dev f dev=σ D and the kinematic relationship ( )1 1T

j j j j

dev

− −

=D J ε J are used

leading to

( ) ( )

( )

1 1 1 1

1 1

T T

2 :

2 :

jo

jo

j j j j j j j j j

v f oV

j j j j j j

f r r oV

W J dV

J dV

− − − −

− −

= −

= −

J J ε J J ε

C ε C ε

(3.7)

77

where Tj j j

r =C J J is the right Cauchy-Green deformation tensor. Using the virtual work of the

preceding equation, the viscous forces can be defined as

( )1 1

2 :j

o

jj j j j j j

v f r r ojVJ dV

− − = −

ε

Q C ε Ce

(3.8)

Since this integral is defined in the curved reference configuration, the volume relationship defined

in the preceding section can be used to change the domain of integration to the straight

configuration.

In order to impose the incompressibility condition, the penalty method is applied at both

the position and velocity levels, 1jJ = and 0jJ = , respectively. The strain energy and dissipation

penalty functions can be written as in Chapter 2 as ( ) ( )2

1 2 1j j

IC ICU k J= − and

( ) ( )2

1 2j j

TD TDU c J= , respectively, where ICk and TDc are the penalty coefficients. The associated

penalty forces can be derived as ( )( )1j j j j j j

IC IC ICU k J J= = − Q e e and

( )j j j j j j

TD TD TDU c J J= = Q e e , where

( ) ( ) ( )T T Tj j j j j j j j j j j j j

x y z y z x z x yJ J = = + + e e S r r S r r S r r (3.9)

and ( )trj j jJ J= D , in which ( )tr refers to the trace of a matrix, , , ,j x y z =S , refers to the

partial derivative of the shape function matrix with respect to the coordinate defined in the

straight configuration, j j j

=r S e for , ,x y z = , and j

e and j

e are the element nodal coordinate

and velocity vectors, respectively, defined in the current configuration. Consequently, the viscosity

and penalty forces of the fluid element can be written as j j j j

s v IC TD= + +Q Q Q Q .

78

3.4.2. Fluid Element Equations of Motion

The virtual work of the inertia forces of the fluid element j is written as

j

j j j j j j j j

i vv

W dv = = r r m e e , where T

i

j j j j j

VV

dV= m S S , and j

v and j

V are the

densities defined in the current and straight configuration, respectively, and they are related by the

equation j j j

V e v = J . This demonstrates that the mass matrix is a constant and symmetric matrix

regardless of the amount of the fluid displacement. The virtual work of the externally applied

forces can also be written as T

j

j j j j j j

e e ev

W dv = = f r Q e , in which T

j

j j j j j

e e eV

dV= Q S f J is

defined as the body force applied on the fluid element; this force expression is obtained by using

the relationship between the volumes in the current and straight configurations. Applying the

principle of virtual work for the fluid element j , one obtains the element equations of motion as

j j j j

e s= −m e Q Q , where j

eQ is the vector of the body forces and j

sQ is the vector of the elastic

forces.

3.5. Integration with MBS Railroad Vehicle Algorithms

The fluid model proposed in this thesis is implemented in an MBS railroad vehicle algorithm in

this chapter in order to develop new liquid sloshing models with significant details. In this section,

the detailed railroad vehicle model used in this chapter is introduced, the track and wheel/rail

profile geometries are discussed, and the three-dimensional elastic wheel/rail contact formulation

which allows for wheel/rail separation is briefly explained.

3.5.1. MBS Vehicle Model

The railroad MBS vehicle model considered in this chapter is shown in Fig. 3.9 and consists of

two trucks and one tank car, where each truck consists of two wheelsets, two equalizer bars, one

79

stub sill, one frame, and one bolster. The MBS vehicle model is thus assumed to have 16 bodies

including 15 rigid bodies and one flexible body representing the fluid. The equalizer bars are

connected to the wheelsets by journal bearings, and the frames are connected to the equalizer bars

using spring-damper elements that represent the primary suspension. The bolster is connected to

the frame using a revolute joint, and the tank is assumed to be rigidly connected to the two stub

sills which are connected to the lead and rear bolsters by spring-damper elements representing the

secondary suspension. The dimensions and inertia properties of the trucks are the same as that

presented in the literature (Shabana et al., 2005). The forward velocity of the vehicle is defined

using a trajectory coordinate constraint function that allows the vehicle to negotiate both tangent

and curved tracks. An elastic contact formulation that allows for wheel/rail separation is used to

define the wheel/rail dynamic interaction in the MBS vehicle algorithm, that is, the wheel is

assumed to have six degrees of freedom with respect to the rail (Shabana et al., 2008). The railroad

vehicle model with a rigid tank-car has 67 degrees of freedom, while in the case of the fluid tank-

car, the model has 900 additional degrees of freedom used to describe the liquid motion inside the

tank.

Figure 3.9. Railroad vehicle model

80

3.5.2. Track Geometry and Wheel/Rail Profiles

In order to examine the effect of the liquid sloshing on the wheel/rail contact and vehicle response,

different simulation scenarios are considered using different track geometries. A curved track is

used to understand the effect of liquid sloshing on the vehicle dynamics when the vehicle

negotiates a curve; Table 3.1 shows the data of the curved track used in this chapter. A tangent

track is also used in the braking scenario to analyze the effect of liquid sloshing on the coupler

forces. It is important to note that in the case of a flexible fluid model, the centrifugal forces do

not take the simple form as in the case of the rigid tank-car, as previously mentioned in this thesis.

For this reason, it is important to perform simulations to determine if the liquid sloshing will affect

the balance speed and the vehicle safety. The track is modeled as a rigid body with zero degrees

of freedom for all of the simulation scenarios considered.

Table 3.1. Track geometry

Segment

Points

No.

Distance

(ft)

Curvature

(deg)

Super-

elevation

(in)

Grade

(%)

Right rail

cant angle

(rad)

Left rail

cant angle

(rad)

A 0 0 0 0 0.025 -0.025

B 100 0 0 0 0.025 -0.025

C 300 3 3 0 0.025 -0.025

D 600 3 3 0 0.025 -0.025

E 800 0 0 0 0.025 -0.025

F 1000 0 0 0 0.025 -0.025

G 1200 -3 -3 0 0.025 -0.025

H 1500 -3 -3 0 0.025 -0.025

I 1700 0 0 0 0.025 -0.025

J 2800 0 0 0 0.025 -0.025

81

The wheel and rail profiles used in this investigation are the same as the profiles used by

Sanborn et al. (2008), the AAR-1B-WF which has a 1:20 taper in the tread section of the wheel

and is commonly used with freight cars in North America, and UIC 60 rail profile. The diameter

of the wheel is 914 mm, and both the wheel and rail profiles are assumed to be in unworn

conditions. The tank used in this chapter has the same dimension used by Wang et al. (2016) with

a length of 11.9 m and radius of roughly 1.5 m for the cylindrical part, and has a maximum capacity

of 15,000 gallons.

3.5.3. Wheel/Rail Contact

A three-dimensional elastic contact formulation is used to define the wheel/rail interaction forces.

This formulation allows for wheel/rail separation, and therefore, the wheel has six degrees of

freedom with respect to the rail. The geometries of the wheel and rail contact surfaces are described

using surface parameters, as shown in Fig. 3.2. The wheel surface parameters are referred to as

T

1 2

w w ws s = s , where 1

ws is the wheel profile surface parameter and 2

ws is the wheel radial

surface parameter (Shabana, 2008). The rail surface parameters are referred to as T

1 2

r r rs s = s ,

where 1

rs is the longitudinal surface parameter that describes the distance traveled and 2

rs is the

rail profile surface parameter. The assumptions of non-conformal wheel/rail contact are used. In

order to determine the contact point, the following four algebraic equations are formulated:

( )T

T T T T

1 2 1 2,w r wr r wr r r w r w = = C s s r t r t n t n t 0 . In this equation, , 1,2; ,k k k

l ls l k w r= = =t r

are the tangent vectors to the surface at the contact point, r

n is the normal to the rail surface, and

wr w r= −r r r is the relative position between the potential contact points on the wheel and rail. The

four nonlinear algebraic equations ( ),w r =C s s 0 can be solved for the four surface parameters.

82

These four surface parameters are used to define the potential contact points on the wheel and rail.

The wheel/rail penetration and the relative velocity along the normal to the rail are defined,

respectively, as Twr r = r n and

Twr r = r n . If there is a penetration between the wheel and rail, an

elastic force model is used to define the normal force (Shabana, 2008). The normal force, the

creepages, the dimensions of the contact ellipse, and the wheel and rail material properties are used

to define the tangential creep force and spin moment using Kalker’s USETAB subroutine (Kalker,

1990). The dimensions of the contact ellipse are determined using Hertz’s contact theory which

requires the evaluation of the principal curvatures. It is also important to point out that the

nonlinear algebraic equations ( ),w r =C s s 0 are used only to determine the geometric surface

parameters and there are no constraint forces associated with these algebraic equation since

wheel/rail separation is allowed. More details on the elastic contact formulation used in this study

can be found in the literature (Shabana, 2008).

3.5.4. MBS Equations of Motion

The general equations of motion for an MBS system that consists of rigid and flexible bodies,

including bodies modeled using the ANCF formulation, can be written in terms of the rigid body

reference coordinates rq , ANCF nodal coordinates e , and the non-generalized surface parameters

s used in the contact formulations, as (Shabana, 2005; Shabana, 2013)

r

r

r rr

ee

d

=

q

e

s

q e s

M 0 0 C Qq

Qe0 M 0 C

0s0 0 0 C

QλC C C 0

(3.10)

83

where rq , e , and s are, respectively, the second time derivatives of the reference, ANCF, and

non-generalized coordinates; rM and eM are, respectively, the rigid body mass matrix and ANCF

constant mass matrix; rqC , eC , and sC are the Jacobian matrices of the constraint equations

associated, respectively, with the reference, ANCF, and non-generalized coordinates; λ is the

vector of Lagrange multipliers; rQ and eQ are, respectively, the vectors of generalized forces

associated with the rigid and elastic coordinates, and dQ is the quadratic velocity vector that

results from the differentiation of the nonlinear algebraic constraint equations twice with respect

to time. The numerical procedure used in this chapter ensures that the nonlinear algebraic

constraint equations are satisfied at the position, velocity, and acceleration levels. A flowchart

depicting the numerical solution procedure is shown in Fig. 3.10.

3.6. Numerical Simulations

In the numerical investigation presented in this section, the effect of liquid sloshing on the vehicle

dynamics is examined. Simulations of a vehicle negotiating a curved track are performed in order

to evaluate the wheel/rail contact forces and the movement of the center of mass of the tank and

fluid, and to have a better understanding of the effect of liquid sloshing on vehicle dynamics when

the vehicle forward velocity is varied. The traction and braking scenarios on a tangent track are

also considered in order to examine the load transfer between the two trucks and the coupler forces

between the two vehicles. These braking scenarios are used to evaluate the effect of liquid sloshing

on the performance of the ECP brake system. The numerical results obtained in this chapter show

that the liquid sloshing does not have a pronounced effect on the vehicle critical speed, but it does

affect the change of the wheel load when the vehicle negotiates a curve at a velocity different from

the balance speed. In the case of vehicle traction and braking, there is significant fluid motion due

84

Figure 3.10. Flowchart of the numerical solution procedure

85

to the acceleration and deceleration of the vehicle. The numerical results are obtained in this

section using tanks partially filled with water.

3.6.1. Curve Negotiation and Balance Speed Analysis

In this section, the curved track described in Table 3.1 is used to investigate the effect of liquid

sloshing on vehicle dynamics by evaluating the wheel/rail contact forces and the movement of the

center of mass of the tank and fluid when the vehicle forward velocity is varied. The track has a

radius of curvature of roughly 582 m which results in a balance speed of roughly 63 km/h defined

in the case of rigid body dynamics as tV gRh G= , where g is the gravity constant, R is the

radius of curvature of the curve, th is the super-elevation, and G is the track gauge. In order to

examine the impact of liquid sloshing on curve negotiation, forward velocities of 40 km/h, 60 km/h,

and 90 km/h are considered in this chapter.

3.6.1.1. Centrifugal Forces

The effect of the centrifugal force when a vehicle is negotiating a curve is to push the vehicle

outward along the direction normal to the curve. The inertia forces of the fluid can be expressed

as the product of the acceleration of the center of mass and the total mass. In order to obtain the

centrifugal force of the fluid and analyze its longitudinal motion, the normal and tangent vectors

can be determined by using the transformation matrix of the track frame relative to the tank body

(Shabana, 2008).

86

Figure 3.11. Lateral component of gravity and outward inertia forces of the fluid

( Gravity force, 40km/h, 60km/h, 90 km/h)

Figure 3.11 shows the outward inertia force and the lateral component of gravity force of

the fluid, which is due to the super-elevation of the track. In order to plot the curves at various

velocities in one figure, the traveling time along the curve is normalized to a dimensionless

parameter which represents the curve length. The results presented in this figure show that the

centrifugal force exerted on the fluid is smaller than the lateral component of the gravity force

when the vehicle travels below the balance speed; however, a larger centrifugal force is exerted

when the vehicle speed is above the balance speed. The resultant force can affect the tank motion

as shown in Fig. 3.12 in which the lateral displacement of the geometric center of the tank with

respect to the track is plotted for both the fluid and rigid body models at various velocities. This

figure illustrates that the liquid sloshing can exacerbate the unbalanced effects when the vehicle

negotiates a curve at a velocity away from the balance speed. It can also be found that when the

vehicle travels near the balance speed, there are no significant differences between the fluid and

87

rigid body model since the liquid exerts the same magnitude of centrifugal force as the lateral

component of gravity, as seen in Fig. 3.11. Traveling at a speed greater than the balance speed

causes instability, which is evident by the results presented in Fig. 3.12c, where the oscillations of

the rigid body model increase after the first curve. However, due to the damping effect of the liquid

motion, the fluid model experiences increased stability compared to the rigid body model.

(a) (b)

(c)

Figure 3.12. Position of the tank center with respect to the track in the lateral direction at (a)

40km/h, (b) 60 km/h, and (c) 90 km/h ( Rigid, Fluid)

88

Figure 3.13 depicts the tangential component of the inertia and gravity forces at a velocity

of 40 km/h, which is used to investigate the longitudinal motion of the fluid inside the tank when

the vehicle is negotiating a curve. It was found that the flexible fluid experiences more than three

times the tangential forces than the rigid fluid due to the liquid sloshing. Numerical simulations

also show that increasing the vehicle forward velocity can increase the tangential forces applied

on the fluid, which will cause larger movement in the longitudinal direction compared to the case

of a lower speed, as shown in Fig. 3.14 in which the position of the center of mass of the liquid

with respect to the tank in the longitudinal direction is plotted in various velocity cases. The results

presented in this figure also show that the relative displacement increases with vehicle velocity.

Figure 3.13. Tangential component of fluid gravity and inertia forces at 40 km/h

( Gravity force, Inertia force)

89

Figure 3.14. Liquid center of mass with respect to the tank in the longitudinal direction

( 40km/h, 60km/h, 90 km/h)

3.6.1.2. Wheel/Rail Contact Forces

The wheel/rail contact forces are also examined in the case of the vehicle negotiating a curve and

are used to examine the impact of the liquid sloshing on wheel/rail interaction. Figures 3.15 and

3.16 depict the normal contact forces on the tread and flange, respectively, on both left and right

wheels of the lead wheelset of the lead truck. The normal forces on the tread show that the liquid

sloshing can intensify the load variance on both the right and left wheels, and tends to increase the

amount of the load variation when the vehicle does not travel at the balance speed during the curve

negotiation case. However, the normal forces on the flange increase with the vehicle forward

velocity and there are no noticeable differences between the fluid and rigid body models.

90

(a) (b)

(c)

Figure 3.15. Tread normal contact force of truck lead wheelset (a) 40km/h, (b) 60 km/h, (c) 90

km/h. ( Rigid-Right, Fluid-Right, Rigid-Left, Fluid-Left)

(a) (b)

91

(c)

Figure 3.16. Flange normal contact force of truck lead wheelset (a) 40km/h, (b) 60 km/h, (c) 90

km/h. ( Rigid-Right, Fluid-Right, Rigid-Left, Fluid-Left)

Figures 3.17 and 3.18 depict the lateral contact forces on the tread and flange, respectively,

which exhibit similar patterns compared to the normal forces. It is clearly shown that the lateral

forces on the flange increase with the vehicle forward velocity since more lateral forces are needed

to balance the centrifugal forces, which also increase with the vehicle velocity.

(a) (b)

92

(c)

Figure 3.17. Tread lateral contact force of lead wheelset of lead truck (a) 40km/h, (b) 60 km/h,

(c) 90 km/h. ( Rigid-Right, Fluid-Right, Rigid-Left, Fluid-Left)

(a) (b)

(c)

Figure 3.18. Flange lateral contact force of lead wheelset of lead truck (a) 40km/h, (b) 60 km/h,

(c) 90 km/h. ( Rigid-Right, Fluid-Right, Rigid-Left, Fluid-Left)

93

3.6.2. Traction and Braking Analysis

In this section, motion scenarios are used to examine the impact of liquid sloshing on the vehicle

nonlinear dynamics in the traction and braking cases.

3.6.2.1. Vehicle Traction Analysis

Figure 3.19. Average normal contact forces of lead and rear trucks in the traction case

( Rigid-Lead, Fluid-Lead, Rigid-Rear, Fluid-Rear)

In the traction scenario, a single vehicle model is used and a trajectory velocity constraint is applied

on the lead stub sill to represent the vehicle traction scenario. The trajectory velocity constraint

causes the vehicle to accelerate according to user-specified trajectory and velocity relationships

(Shabana, 2008). A constant acceleration of 0.3 m/s2 is used to accelerate the vehicle to 20 km/h

in 15 s and then maintain this constant velocity. The average contact forces of the four wheels of

the lead and rear trucks are plotted in Fig. 3.19. It is evident that the liquid sloshing has a significant

effect on the load distribution during the vehicle traction; there is an approximately 13% difference

in the normal load compared with that of the rigid body model. After traction, the liquid continues

94

to experience sloshing towards a steady state which can be clearly seen in Fig. 3.20, which shows

the longitudinal displacement of the fluid center of mass with respect to the tank during the traction.

It is apparent that there is a maximum motion of roughly 0.7 m of the center of mass, which will

certainly affect the wheel load as shown in Fig. 3.19.

Figure 3.20. Fluid center of mass longitudinal displacement with respect to the tank in the

traction case

3.6.2.2. Vehicle Braking Analysis

(a) (b)

Figure 3.21. Coupler forces between two cars in the case of braking using (a) Conventional

brake, (b) ECP brake ( Rigid, Fluid)

95

In order to consider the coupler force between cars, a two tank-car model is developed based on

the single MBS vehicle model to simulate the braking scenarios, in which the coupler is

represented as a linear spring-damper force element. The nonlinear braking torque associated with

the vehicle loads and forward velocities is applied on the wheelsets to perform braking in this

analysis. The case where only the lead car brakes as well as the case where both cars brake are

simulated in order to consider the usual brake situations. These two scenarios are used to examine

the effects of liquid sloshing on the dynamic response of the train during braking when the

conventional and ECP brake systems are used. In this chapter, the train is decelerated from 40

km/h to 5 km/h in 40s with a nonlinear braking force. The coupler used in this model has a stiffness

coefficient 300 MN m and a damping coefficient of 200 kN s m . The coupler forces are plotted

in Fig. 3.21 in the case where only the lead car brakes as well as the case of both cars braking, for

both the fluid and rigid body models. By comparing the results in Figs. 3.21a and 3.21b, it can be

seen that a larger coupler force is exerted when only the lead car brakes regardless of whether the

flexible fluid or the equivalent rigid fluid model is used, which shows the significance of ECP

application in railroad vehicles. In the case of brake application on the lead car only, the flexible

fluid model has essentially the same coupler force as that of the rigid fluid model initially, while

it experiences a larger value as the vehicle velocity decreases. However, in the case of uniform

and simultaneous brake application on both cars (ECP), the flexible fluid model shows

significantly larger coupler forces than the rigid body model for the entire scenario due to the

increased relative motion of the fluid inside the tank as shown in Fig. 3.22, which depicts the

longitudinal and lateral displacement of the center of mass of the fluid with respect to the tank. It

is clearly shown that in the case of ECP braking, the fluid moves more significantly because the

vehicle system experiences a more severe acceleration resulting from increased braking torques

96

that are applied as compared to the conventional brake scenario. Figure 3.23 shows the

configuration of these two cars partially filled with water in braking, and it can be clearly seen that

there is significant liquid motion due to the deceleration of the vehicle.

(a) (b)

Figure 3.22. Front car fluid center of mass displacement with respect to the tank during braking

(a) Longitudinal, (b) Lateral direction ( Conventional brake, ECP brake)

(a)

(b)

Figure 3.23. Braking animation of two tank-cars filled with liquid in the (a) Parked state, (b)

Braking state

97

3.7. Concluding Remarks

In this chapter, a new approach is proposed for the integration of a continuum-based sloshing

model with a fully nonlinear MBS rail vehicle model. The contributions of this chapter are as

follows: (1) A unified geometry/analysis mesh is used from the outset in order to accurately capture

complex fluid and tank geometries as well as the nonlinear dynamic behavior of the fluid and

vehicle. The approach developed in this chapter is used to examine the effects of liquid sloshing

on railroad vehicle dynamics when negotiating a curve and during traction or braking; (2) The

method of the tank and fluid geometry description is described and it is shown how a unified

geometry/analysis mesh can be developed for both the rigid rail and continuum fluid bodies. The

search method used to define the fluid/tank contact points is outlined and the penalty force model

used to describe the fluid/tank interaction forces is formulated; (3) The fluid constitutive equations

that account for the viscosity and incompressibility effects are presented. The liquid sloshing

model developed in this study is integrated with the MBS railroad vehicle model which takes into

consideration the nonlinear three-dimensional wheel/rail contact forces and the wheel and rail

profile geometries; and (4) In order to systematically examine the effect of the motion of the

flexible fluid on railroad vehicle dynamics when the vehicle is negotiating a curve, a general

definition of the outward inertia force of a flexible body using both FFR and ANCF descriptions

is investigated. The analysis presented in this chapter shows that this force depends strongly on

the motion of the continuum and does not take the simple form used in the case of rigid body

dynamics.

Comparative simulations are performed to examine the liquid sloshing effects by using

flexible and rigid body fluid models. It is shown that the liquid sloshing can exacerbate the

unbalanced effects when the vehicle travels at a velocity away from the balance speed, but this

98

effect decreases when the forward velocity is close to the balance speed because the liquid

experiences the same centrifugal force as the rigid fluid body in this case. The results in the traction

analysis show that the liquid motion can significantly affect the load distribution between the front

and rear trucks. Comparing with the ECP braking case, there is a larger coupler force when the

conventional braking is used for both the flexible and rigid body fluid models. Nonetheless, the

results obtained for the model considered in this chapter demonstrate that the liquid sloshing

amplifies the coupler force greatly in the ECP braking case compared to the equivalent rigid body

model because the latter model cannot capture the fluid nonlinear inertia effects.

99

CHAPTER 4

GEOMETRICALLY ACCURATE REDUCED ORDER FLUID MODELS

The objective of this chapter is to integrate for the first time the newly developed absolute nodal

coordinate formulation/floating frame of reference (ANCF/FFR) solid finite elements (FE) with a

fully nonlinear multibody system (MBS) algorithm. ANCF/FFR elements are able to capture

initially curved structures such as the fluid within a cylindrical tank while retaining the same

number of degrees of freedom as conventional elements and taking advantage of modal reduction

techniques, resulting in faster simulation times compared to the higher-order ANCF elements. In

this chapter, the solid element is developed in terms of constant geometric coefficients which are

obtained using the matrix of position vector gradients defined in the reference configuration. No

geometry distortion occurs when computer-aided design (CAD) models are converted to FE

meshes using ANCF/FFR elements because such meshes are developed using ANCF elements,

which are related to B-splines and Non-Uniform Rational B-Splines (NURBS) by a linear mapping.

The fluid constitutive model, which is based on the Navier-Stokes fluid model, is developed and

the incompressibility conditions, which are enforced using a penalty approach, are defined. A

sloshing box model and a medium-duty tanker truck model with a tank half filled with water are

developed in order to investigate the ability of the new ANCF/FFR elements to model the fluid

sloshing in comparison to fluid meshed using ANCF elements. The fluid/tank contact formulation,

which is enforced using a penalty approach, is described. It is shown that while the sloshing

amplitudes of the ANCF/FFR box meshes are reduced compared to the converged ANCF meshes,

the general sloshing behavior is still captured at a significantly reduced CPU time, indicating that

100

the ANCF/FFR elements can contribute to significant improvement of the computational

efficiency in applications in which capturing some geometric changes due to the fluid displacement

is not critical. This conclusion was confirmed by the results of the highway vehicle lane change

simulation – the sloshing amplitudes of the center of mass predicted using the ANCF/FFR fluid

mesh during the lane change are found to be in a good agreement with what predicted by the ANCF

mesh. Furthermore, the results of the overall vehicle-dynamics, as measured by the tire contact

forces predicted using the two different meshes, are found to be in a good agreement. The results

obtained in this investigation demonstrate that if the goal is to accurately capture the free-surface

displacement of the fluid, then ANCF elements are better candidates due to their high order and

ability to capture complex shapes. However, if the goal is to perform efficient simulations to obtain

the overall vehicle motion, then using ANCF/FFR elements are a better alternative.

4.1. FE Mesh Geometry and Position Vector Gradients

Figure 4.1. Cylindrical vehicle tank

The tanker truck considered in this investigation features a tank with cylindrical geometry, as

shown in Fig. 4.1. It is necessary for the FE mesh to have an accurate representation of the initially

101

curved geometry in order to develop a realistic virtual model in which the fluid fills the tank. In

this section, the method used to produce the initially curved reference-configuration geometry of

the ANCF and ANCF/FFR meshes used to model the fluid inside the tank is described. The fluid

body can be described using three different configurations, as depicted in Fig. 4.2. The curved

reference configuration is the initial configuration of the fluid inside the tank at the start of the

simulation. Its volume is denoted as oV with position vector X . The initially curved reference

configuration is achieved by shaping a straight configuration, with volume V and position vector

x . During the simulation, the fluid may experience large deformations due to the forces exerted

by the tank. These deformed shapes are referred to as the current configuration, with volume v

and position vector r . The volumes of the curved reference and current configurations are related

by the determinant of the matrix of position vector gradients = J r X according to odv JdV= ,

where J = J and denotes the determinant of a matrix. The curved reference configuration is

related to the straight reference configuration by o odV J dV= .

Figure 4.2. Fluid configurations

102

Furthermore, the mesh in the current configuration can be mapped to the straight reference

configuration; by rewriting the matrix of position vector gradients as

( )( ) 1

e o

−= = =J r X r x x X J J , where eJ relates the volumes in the current and straight

reference configurations, one can write 1

o o e o o edv JdV JJ dV J dV J dV−= = = =J J . Using these

identities, integrations carried out over the curved configurations can be systematically converted

to integrations over the straight reference configuration, regardless of the amount of deformation

that occurs within the fluid mesh.

The ANCF/FFR solid element is developed using the ANCF solid element as a basis, which

features eight nodes with three gradient vectors and three absolute position coordinates, for a total

of 12 coordinates per node and 96 coordinates per element (Olshevskiy et al., 2013). The cubic

shape functions allow for the creation of complex geometries, such as a half-cylindrical fluid mesh

which can be coincident with the tank walls. The shape functions ( )1,2,3,4k

is i = at node

( )1,2, ,8k k = for this element are as follows:

( ) ( )( )( )( )( ) ( )( ) ( )( )( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

1

1

2 1 11

2

1 2 11

3

1 1 21

4

1 1 1 1

1 1 2 1 2 1 2

1 1 1 1

1 1 1 1

1 1 1 1

k k k

k k k k kk k k

k k k k kk k k

k k k k kk k k

k

k k k

k k k

k

k

k

s

s a

s b

s c

+ + +

+ − − −+

+ − − −+

+ − − −+

= − + − + − + −

+ + − + − − + − −

= − − − −

= − − − −

= − − − −

(4.1)

where , , and are the dimensionless parameters along the three coordinate axes, and k ,

k , and k are the parameters evaluated at node k . The mesh in the straight reference

configuration can be shaped into a half-cylinder by adjusting the gradient vectors at each node.

More information about using the gradients to develop initially curved meshes can be found in the

literature (Shi et al., 2017).

103

The ANCF/FFR solid element also features eight nodes, however the nodal coordinate

vector includes infinitesimal rotations and a local position vector, instead of the gradient vectors

and absolute position vector used by ANCF elements (Shabana, 2018; Tinsley and Shabana, 2019).

The transformation from gradient vectors to infinitesimal rotations and the derivation of the new

shape function matrix will be described in the following section.

4.2. Finite Element Formulations

In this section, the ANCF formulation, the conversion of nodal gradient vectors to infinitesimal

rotations, and the FFR formulation are described.

4.2.1. Absolute Nodal Coordinate Formulation

The position of an arbitrary point on an ANCF body can be written as

( ) ( ) ( ) ( ) ( )( ), o dt t t= = +r x S x e S x e e , where ( )S x is the space-dependent shape function matrix,

e is the time-dependent vector of nodal coordinates defined in the current configuration, oe is the

vector of nodal coordinates in the curved reference configuration, and de is the vector of

deformation coordinates, where o d+e = e e and the nodal coordinates are defined in the global

coordinate system. Using general continuum mechanics, the position vector can be rewritten as

o d= +r r r , where or is the position of the point in the curved reference configuration and dr is

the displacement vector, and thus o o=r Se and d d=r Se , demonstrating that ANCF elements are

isoparametric. The matrix of position vector gradients which relates the current and curved

reference configurations can be written as 1

e o= −J J J , where o o= = xJ X x S : e ,

e r= = xJ x S : e , xS is a third-order tensor defining the derivatives of the shape functions with

104

respect to the coordinates in the straight reference configuration, and : denotes the tensor double

contraction. The velocity and acceleration equations are obtained by differentiating the position

equation once and twice, respectively, with respect to time and can be written as =r Se and

r = Se . Using the velocity equation, the kinetic energy can be defined as

( ) ( ) ( ) ( )1 2 1 2 1 2T T T T

v vT dV dV = = = r r e S S e e Me , where

T

vM dV= S S is the constant,

symmetric mass matrix. The ANCF generalized external forces can be obtained using the principal

of virtual work. The virtual work of the external forces can be written as

T T T

e fW = = =F r F S e Q e , where T

f =Q S F is the vector of generalized forces associated

with the ANCF elastic coordinates.

4.2.2. Conversion of Position Vector Gradients to Infinitesimal Rotations

The matrix of position vector gradients for the ANCF/FFR elements can be written as

( )1 2 3, ,e o o = =J JJ A J , where A is an orthogonal rotation matrix which is a function of the

three rotation parameters 1 , 2 , and 3 and can be written as

( ) ( ) ( )1 1 2 3 2 1 2 3 3 1 2 3, , , , , ,= = A J a a a where the columns ia ( )1,2,3i = are

orthogonal unit vectors. Using the assumption of infinitesimal rotations, the rotation matrix A can

be written as = +A I θ , where θ is a skew-symmetric matrix associated with the vector of rotation

coordinates 1 2 3

T =θ . Substituting this equation into the equation for the matrix of

position vector gradients eJ yields ( )e o o o o= = + = +J AJ I θ J J θJ . This equation can be

rewritten as

1 2 3 1 2 3 1 2 3e o o o o o o o o o o o o = + = + = + − − − J J θ J J J J θJ θJ θJ J J θ J θ J θ where

105

( )1,2,3oi i =J are the skew-symmetric matrices associated with the column vectors ( )1,2,3oi i =J

of oJ . Thus, the vector of element nodal coordinates which includes position vectors and position

vector gradients can be written as

( ) ( ) 1 1 1

2 2 2

3 3 3

o

x o o d

o d o cd

x o o

x o o

t t

− = + = = + = + −

r r I 0

r J 0 J re e e e Be

r J 0 J θ

r J 0 J

(4.2)

where 0

k k

cd=e Be . At an arbitrary node k , the nodal coordinates can be written as

( )k k k k k

o d o cdt = + = +e e e e Be , where

( ) ( )1

2

3

11

22

33

; ; ;

k k

ok kk kx ok k k ko d

o cdk kk kx oo

kkkoox

t t

− = = = = − −

r I 0r

r 0 JJ re e B e

r 0 JJ θ

0 JJr

(4.3)

Recalling from Sec. 4.2.1 that the position of an arbitrary point on the ANCF body can be

written as ( ) ( ) ( ) ( )( )o d= t t= +r S x e S x e e , one can substitute Eqn. (4.1) into this position

equation to yield ( ) ( ) ( ) ( )( )o cd o d cd= t t= + = +r S x e S x e Be Se S e where d =S SB is the new

displacement shape function matrix. This demonstrates that the new ANCF/FFR-FEs are non-

isoparametric elements, because different shape functions are used to interpolate the geometry and

displacements; however, the effect of the initial curved geometry is still accounted for in the shape

functions dS and the position equation can be used for both straight and curved geometries. It can

also be shown that the displacement shape function matrix has a proper set of rigid body modes at

the velocity level (Shabana, 2017B), which is important because the velocity equation is used to

define the element inertia, which will be shown in the following section. The displacement

106

shape function matrix for node k ( )1,2,...,8k = can be written as

4 3 4 2 4 1 4

1 2 3( )k k k k k k k k k k

d o o oS S S S− − − = = − + + S S B I J J J and can be assembled at the element level

as 1 2 8

d d d d = S S S S .

4.2.3. Floating Frame of Reference Formulation

The position vector i

Pr to an arbitrary point P on a flexible body i , as shown in Fig. 4.3, can be

written in the FFR formulation as i i i

P P= +r R u , where i

R is the global position vector to the body

coordinate system and i

Pu is the global position of point P with respect to the body coordinate

system. i

Pu can also be written i i i

P P=u A u , where i

A is the orthogonal transformation matrix

which describes the orientation of the body reference system, and i

Pu is the local position vector

to point P in the body coordinate system. In the FFR formulation, the local position vector i

Pu

can be decomposed as i i i

P O f= +u u u , where i

Ou is the position vector to point P in the reference

configuration, and i

fu is the local deformation vector. i

fu can be written as i i i

f f=u S q , where i

S

is the shape function matrix of body i , and i

fq is the vector of the FFR elastic coordinates. Thus,

the global position vector of point P is ( )i i i i i i

P O f= + +r R A u S q .

Figure 4.3. Floating Frame of Reference formulation

107

The velocity equation can be obtained by differentiating the position equation with respect

to time, and by knowing that i

Ou and i

S are constant with respect to time. The velocity of point P

can therefore be written as i i i i i i i

P P f= + +r R A u A S q . In the preceding equation, the first term is the

velocity of the body reference and the third term is the velocity of point P due to local deformation.

The second term results from differentiation of the rotation matrix with respect to time and is equal

to zero in the case of rigid body motion. The second term can be written as

( ) ( )i i i i i i i i i i i i i i i i i

P P P P P= = − = − = − =A u A ω u A u ω Au ω Au G θ Bθ , where i

G is the velocity

transformation matrix which relates the angular velocities i

ω and the time derivatives of the

coordinates iθ , and i i i i

P= −B A u G . This partitioning allows for writing the velocity equation in

matrix form as

i

i i i i i

p

i

f

=

R

r I B A S θ

q

(4.4)

where I is a 3x3 identity matrix. Defining the left-hand matrix as i

L , the velocity equation can be

written in simple form as i i i

P =r Lq , where i i i i = L I B A S and T T T T

i i i i

f =

q R θ q is the

vector of generalized velocities.

To obtain the acceleration vector, the matrix form of the velocity equation is differentiated

with respect to time as i i i i i

P +=r Lq Lq . In the preceding equation, the first term includes quantities

which are quadratic in the velocities, such as the Coriolis terms, and the second term includes

quantities linear in the accelerations. It can also be shown that this equation may be written as

( ) ( )2i i i i i i i i i i i i i i i i i

P P P P P O N T C D= + + + + = + + + +r R ω ω u α u ω Au A u a a a a a , where i

Oa is the

acceleration of the body reference, i

Na is the normal component of the acceleration, i

Ta is the

108

tangential component, i

Ca is the Coriolis component, and i

Da is the acceleration due to the local

deformation. It is clear that in the case of rigid body motion, the last two terms are equal to zero

because i

Pu is constant.

Using the velocity equation that was previously defined, the kinetic energy can be written

as ( ) ( )1 2 1 2T T T

i i

i i i i i i i i i i i

V VT dV dV = = r r q L Lq . Because the velocities are not a function of

the volume, this equation can be written as ( ) ( )1 2 1 2T T T

i

i i i i i i i i i i

VT dV= =q L L q q M q where

T

i

i i i i i

VdV= M L L is the symmetric positive definite mass matrix of body i . The mass matrix

can be partitioned as

.

i i i

RR R Rf

i i i

f

i

ffsym

=

m m m

M m m

m

(4.5)

where

T

T T

i i i i i i i

RR RV V

i i i i i i i i i

RfV V

i i i i i i i i i i i

f ffV V

dV dV

dV dV

dV dV

= =

= =

= =

m I m B

m S m B B

m B A S m S S

(4.6)

It is clear that i

RRm and i

ffm are constant with respect to time; in general, however, the FFR mass

matrix is highly nonlinear and varies with time.

The virtual work of the generalized external forces acting on body i , including the gravity

force and the tank contact, can be written as T Ti i i i

e e fW = =F r Q q , where i

F is the external force

vector, ir is the virtual change in the position vector, and

T T T Ti i i i

e R f =

Q Q Q Q is the vector

of generalized external forces; i i

R =Q F and ( )T

i i i i i

P = −Q A u G F are the vectors of generalized

109

forces associated with the reference translation and rotation, respectively, and ( )T

i i i i

f =Q A S F is

the vector of generalized forces associated with the deformation.

4.3. Fluid Modeling Approaches

In this thesis, the fluid in the vehicle tank is assumed to be water. The constitutive model and the

method for enforcing the fluid incompressibility are discussed in this section.

4.3.1. Fluid Constitutive Model

The constitutive model used in this chapter is based on the Navier-Stokes equations and is

applicable to Newtonian fluids; in this thesis, water is the fluid considered. This constitutive model

has been validated against experimental and numerical techniques and found to be in good

agreement (Grossi and Shabana, 2017). The fluid equations can be defined by the symmetric

Cauchy stress tensor as ( )tr 2 vol devp = − + + = + σ D I D σ σ , where p is related to the

hydrostatic pressure, ( )tr refers to the trace of a matrix, D is the rate of deformation tensor, I is

a 3x3 identity matrix, and and are the Lamé constants.

Because the penalty approach is used to enforce the incompressibility in this investigation,

as will be discussed in the following section, the first two terms of the stress tensor vanish, reducing

to 2dev dev=σ D . The second Piola-Kirchoff stress tensor is defined in the reference configuration

and is therefore often used in Lagrangian analysis. The second Piola-Kirchoff stress tensor can be

obtained from the Cauchy stress tensor as 1

2

T

P devJ − −=σ J σ J , where X Y X= =J r X r r r is

the matrix of position vector gradients, X is the vector of coordinates in the curved reference

configuration, and J=J , as defined previously. The virtual work of the fluid viscous forces for

110

an element j can be written as ( )0

1

2 0: :jj

j j j j j j j j

s dev Pv V

W dv dV −

= − = − σ J J σ ε where

( )( )1 2Tj j j= −ε J J I is the Green-Lagrange strain tensor,

jv is the volume of element j in the

current configuration, 0

jV is the volume of element j in the curved reference configuration, and

: indicates the double contraction. Using the expression for 2Pσ and the identity

( ) ( )1T

j j j− −

=D J ε J , the virtual work of the viscous forces becomes

( ) ( )0

1 1

02 :T

j

j j j j j j

s r r vV

W J dV − −= − = C εC ε e Q e , where

( ) ( )0

1 1

02 :j

j j j j j

v r rV

J dV − −= − Q C εC ε e is the vector of generalized viscous forces and j

e are

the nodal coordinates of element j . By using the identity 1 1

0 0

j j j j j j

e eJ− −

= = =J J J J J , the vector

of generalized viscous forces can be rewritten as

( ) ( ) ( ) ( )0

1 1 1 1

02 : 2 :j j

i j j j j j j j j j j

v r r r rV V

J dV J dV − − − −= − = − Q C ε C ε e C ε C ε e (4.7)

This allows for carrying out integration in the straight reference configuration, which is constant

in time.

4.3.2. Fluid Incompressibility

In the case of incompressible fluids such as water, the fluid incompressibility can be enforced

using one of several methods. The continuity condition ( )( ) ( )( ), t t +r r reduces to

0 =r in the case of incompressible materials, where is the density of the fluid, because for

incompressible fluids, the density is constant over both space and time. This constraint equation

can be applied at each integration point to ensure the fluid remains incompressible. However, this

method introduces a significant number of constraint equations, thus reducing the computational

111

efficiency of the model. Therefore, in this chapter, the penalty approach is instead used again to

enforce the incompressibility. The incompressibility condition is enforced at both the position

( )1jJ = and velocity ( )0jJ = levels using the energy function ( ) ( )2

1 2 1j j j

C CU k J= − and the

dissipation function ( ) ( )2

1 2j j j

D DU c J= , respectively. The penalty coefficients are chosen such

that jJ remains within a specified tolerance of 1. The generalized forces associated with

incompressibility of element j can be defined as ( ) ( )( )1Tj j j j j j j

C C CU k J J= = − Q e e at the

position level and ( ) ( )Tj j j j j j j

C D DU c J J= = Q e e at the velocity level, where

j j j jJ J = e e and ( )trj j jJ J= D . By using the identity

( ) ( ) ( )j j j j j j j j j j

X Y Z Y Z X Z X YJ = = = r r r r r r r r r , any of the three expressions can be used to

calculate j jJ e as ( ) ( ) ( )j j j j j j j j j j j

X Y Z Y Z X Z X YJ = + + e S r r S r r S r r . The generalized

forces on body j , j j j j

v C D= + +Q Q Q Q , can be assembled using a standard finite element assembly

procedure to determine the generalized forces acting on the fluid body.

4.4. Fluid/Tank Contact

In this section, the contact detection algorithm between the fluid and tank will be developed. The

contact is checked in two directions – radially in the cylindrical section and longitudinally for the

flat tank ends. The contact forces, based on the penalty approach, will also be discussed for the

case of an ANCF/FFR or ANCF fluid body.

112

4.4.1. Contact Detection

Consider a point P on the fluid body, which is located at position Pr in the global frame. In order

to determine if contact between fluid point P and the tank is occurring, the global position of

point P with respect to the tank coordinate system is obtained by

1 2 3

Tt t t t t

P P P P Pu u u = = − u r R , where t

R is the position vector to the tank coordinate system.

The position vector of the point P relative to the tank coordinate system is transformed to the

local coordinate system using the matrix t

A which defines the orientation of the tank coordinate

system as Tt t

P P=u A u . Contact is detected separately in the radial direction in the cylindrical

portion of the tank, and in the longitudinal direction at the tank ends.

4.4.1.1. Radial Direction

The longitudinal direction of the tank is along the X axis of the tank coordinate system, and thus

radial contact is checked in the local Y Z− plane. In order to detect contact in this plane, the

projection matrix P is defined as ˆ ˆ= − P I a a , where a is a unit vector along the tank

longitudinal axis. The projection of the local position vector t

Pu is thus calculated as ( )t t

P Pr=u Pu ,

which is the position of the fluid point in the radial plane. Given that the radius of the tank is tr ,

contact is occurring between point P and the tank if the length of ( )t

Pr

u exceeds the tank radius,

or if ( ) ( ) ( )t t t t

P P Pr r r

r= u u u , and the amount of penetration can be calculated as

( )t t

Pr

r r = −u . The unit normal vector along which the penalty force acts is defined as

( ) ( )t t

r P Pr r

= −n u u .

113

4.4.1.2. Longitudinal Direction

The vehicle tank features flat ends, and thus contact is detected by comparing the position of the

fluid point relative to the tank’s length. The local position vector of point P in the tank coordinate

system is projected on a unit vector along the tank’s length, that is ( ) ˆt t

P Pl= u u a . If the magnitude

of this vector is longer than half the length of the tank; that is, if ( ) ( ) ( ) 2t t t t

P P Pl l l

L= u u u ,

where tL then point P is in contact with the tank end. The unit normal vector ln along which the

penalty force acts is opposite to the unit vector along the length of the tank, ˆ−a .

4.4.2. Penalty Forces

In the case that contact occurs between a fluid point and the tank, the penalty forces are exerted on

the fluid and the tank at the point of contact. The contact force in the tank coordinate system is

defined as n tf f= +f n t , where nf and tf are the magnitudes of the normal and tangential contact

forces, and n and t are unit vectors along the normal and tangential directions, respectively. In

this investigation, nf is defined as 1.5

n nf K Cv = + , where K and C are penalty coefficients

at the position and velocity level, respectively, the  term is used to ensure the contact force is

0 when there is no penetration, the exponent on the penetration in the stiffness term was chosen to

be 1.5 to increase smoothness near zero penetration, and nv is the normal velocity of the contact

point which is defined as n Pv = r n , where Pr is the absolute velocity of the contact point

(Shabana et al., 2008). The tangential force due to friction between the fluid and the tank walls is

calculated using a smoothing function as sin2

nt n

s

vf f

v

=

if

n sv v and t nf f= if n sv v ,

114

where sv is a specified limit to the velocity smoothing function. The unit vector in the tangential

direction is calculated as ( ) ( )P Pt t=t r r , where ( )P t

r is the tangential velocity, which is

calculated as ( )P P ntv= −r r n . The local force vector can be transformed to the global coordinate

system as t=F A f , to be included in the system equations of motion which will be discussed in

Sec. 4.5.

4.4.3. Generalized Forces

Using the penalty force vector defined in the previous section, the generalized external forces

associated with the translation, rotation, and elastic coordinates are calculated. The generalized

external forces can be obtained using the principle of virtual work as T

e eW = =F r Q q , where

F is the external force vector including the penalty forces, q are the virtual changes in the

coordinates, and eQ is the vector of generalized external forces. Using the procedure outlined in

Sec. 3, the generalized force vector can be calculated as T

T T T

e R f = Q Q Q Q , where R = −Q F ,

T t

P =Q G u F , and ( )T

f =Q AS F in the case of an ANCF/FFR fluid mesh and T

f =Q S F in the

case of an ANCF fluid mesh. These generalized external forces are included in the system

equations of motion, which will be discussed in the following section.

4.5. Equations of Motion

The equations of motion for the vehicle system considered in this study can be expressed in the

augmented Lagrangian form as:

115

r

f

r f

T

r r r c

T

f f f

d

+

=

q

q

q q

M 0 C q Q Q

0 M C q Q

λ QC C 0

(4.8)

In this equation, the subscript r refers to the rigid body coordinates and the subscript f refers to

the elastic coordinates in the case of a flexible body. rq is the vector of accelerations of the rigid

body coordinates, fq are the accelerations of the elastic coordinates, and λ is the vector of the

Lagrange multipliers associated with the constraints. rM and fM are the mass matrices and

rqC

and fqC are the constraint Jacobian matrices. rQ and

fQ are the vectors of external forces, cQ

is a vector including Coriolis and centrifugal terms, and dQ is a vector of terms quadratic in the

velocities that arises due to the second time differentiation of the constraint equations. The

equations of motion are solved using the Adams-Bashforth numerical method, ensuring that the

constraints are satisfied at the position, velocity and acceleration levels, and parallel computing is

used to increase the computational efficiency.

4.6. Numerical Examples

A sloshing box problem is first considered in order to compare the fluid behavior between the

ANCF and ANCF/FFR elements in the case of an initially straight mesh. The vehicle model and

motion scenario considered in this investigation are then detailed, which are developed in order to

demonstrate the integration of an ANCF/FFR fluid sloshing model with initially curved geometry

with a complex nonlinear MBS algorithm. The models are developed using the general-purpose

MBD software SIGMA/SAMS.

116

4.6.1. Sloshing Box Model

A simple sloshing box model is first simulated in order to compare the fluid sloshing behavior

between ANCF and ANCF/FFR finite elements in the case of a mesh with initially straight sides.

Cube-shaped meshes with sides equal to 1 m are created with varying numbers of elements such

that the effect of mesh refinement on the sloshing behavior can be studied. Information about the

meshes tested is listed in Table 4.1.

Table 4.1. Sloshing box model information

ANCF/FFR ANCF

Number of

Elements 27 64 216 512 1000 1 8 27 64

Number of

Degrees of

Freedom

384 750 2058 4374 7986 96 324 768 1500

4.6.1.1. Boundary Conditions

Figure 4.4 depicts the four different types of nodes in the fluid mesh – corner (black), edge (red),

face (blue), and free (green) nodes. For the ANCF/FFR element meshes, the fluid remains inside

the box by choosing the appropriate FFR reference conditions. At the preprocessing stage, the

degrees of freedom of each type of node with respect to the body coordinate system are removed

in order to ensure that the fluid remains inside the box and to remove the redundant rigid body

motion from the FFR shape functions. These reference conditions are summarized in Table 4.2

and the colors of each node refer to Fig. 4.4. More information about the FFR reference conditions

can be found in the literature (Shabana et al., 2018; Shabana and Wang, 2018).

117

Figure 4.4. Box boundary conditions

Table 4.2. Reference conditions

Type of Node Node Color Degrees of Freedom

Corner Black None

Edge Red One (along edge axis)

Face Blue Two (within face plane)

Free Green All

The ANCF reference node (ANCF-RN) is used to enforce the boundary conditions of the

ANCF meshes. The ANCF-RN is a node that is associated with the ANCF mesh but does not

belong to any element. Continuity conditions between the reference node and the other nodes of

the mesh are defined at the preprocessing stage in order to remove the degrees of freedom of each

node with respect to the reference node such that the fluid remains inside the box. The continuity

conditions are the same as those listed in Table 4.2, except that the degrees of freedom are defined

with respect to the ANCF-RN rather than the body coordinate system, as is the case for ANCF/FFR

meshes. More information about the ANCF-RN can be found in the literature (Shabana, 2015).

118

4.6.1.2. Results

(a) (b)

Figure 4.5. Maximum deformation of converged (a) ANCF and (b) ANCF/FFR fluid meshes

The fluid boxes are subjected to a sinusoidal forcing function ( ) ( )0.1sin 8f t t= in order to

generate significant sloshing motion. Figure 4.5 shows the maximum deformation of the

converged ANCF and ANCF/FFR meshes. The vertical position of one of the corner nodes of each

mesh is plotted in Fig. 4.6. In the case of ANCF elements, convergence is achieved using 27

elements, while 512 elements are required to achieve convergence for the ANCF/FFR meshes.

Although the deformation of the ANCF/FFR meshes is smaller than that of the ANCF meshes, the

sloshing behavior is evident, which indicates that the reduced-order ANCF/FFR models can be

useful in situations where modeling the overall fluid motion is necessary and fast simulation time

is desirable, but accurately capturing the fluid free surface is not required, such as in real-time

vehicle modeling which is common in the automotive industry.

119

(a) (b)

Figure 4.6. Vertical corner node position of (a) ANCF meshes and (b) ANCF/FFR meshes

(ANCF meshes: 1, 8, 27, 64 elements;

ANCF/FFR meshes: 27, 125, 216, 512, 1000 elements)

Figure 4.7 plots the CPU times for each mesh, normalized with respect to the 27-element

ANCF/FFR mesh CPU time, versus the number of elements in the mesh, N. The results in Fig. 4.7

demonstrate that although an increased number of ANCF/FFR elements are required to achieve

convergence compared to the ANCF elements, the CPU time for the converged ANCF/FFR mesh

is three times faster than for the converged ANCF mesh due to the use of modal reduction

techniques. In the next section, the effect of the fluid sloshing on the vehicle dynamics will be

investigated using meshes of initially curved ANCF and ANCF/FFR elements.

120

Figure 4.7. Normalized CPU times for the sloshing box models

( ANCF, ANCF/FFR)

4.6.2. Tanker Truck Model

In this section, a tanker truck model is considered in order to highlight the differences between

ANCF and ANCF/FFR finite elements for the analysis of fluid sloshing effects on overall vehicle

motion. The tank is assumed to be half-filled with water in order to induce the most severe sloshing.

Figure 4.8. Medium-duty tanker truck MBS model

121

The vehicle considered in this investigation is the medium-duty tanker truck, pictured in

Fig. 4.8, which was described in Chapter 2. The model includes 20 rigid bodies with 32 degrees

of freedom, in addition to the fluid body. The rigid bodies include 10 tires, a four-bar Ackermann

steering mechanism, two rear axles, the frame, a cab, a tank, and a ground body. The tires, which

are modeled using Pacejka’s brush tire model (Pacejka, 2006), are connected to the axles using

revolute joints, and the cab, tank, and frame are rigidly attached. The sprung mass is supported by

a suspension system of 18 spring-damper elements which provide restoring forces in the

longitudinal, lateral, and vertical directions. The wheelbase is 4.064m, the track width is 1.939m,

and the total vehicle weight is 111.67kN. The volume of the tank is roughly 3,000 gallons, typical

of medium-duty trucks which service residential areas.

Figure 4.9. Lane change path

The tanker truck model performs a lane change over a standard-width U.S. highway lane

of 3.7m, as seen in Fig. 4.9. The lane change is negotiated in a short time of 4s in order to induce

122

noticeable lateral sloshing motion. The ANCF fluid is meshed using 48 elements with 1260

degrees of freedom, and the ANCF/FFR fluid is meshed using 112 elements with 1350 degrees of

freedom. Further mesh refinement was shown not to have a significant effect on the solution. The

solution is also obtained for a model in which the fluid is represented by a rigid body fixed to the

inside of the tank in order to isolate the effect of the fluid sloshing on the vehicle motion.

Figure 4.10. Lateral position of fluid center of mass with respect to tank during lane change

maneuver

( ANCF, ANCF/FFR)

Figure 4.10 shows the lateral position of the center of mass of the fluid with respect to the

tank center for the two flexible fluid models. While the first peak agrees well between the two

models, the later peaks of the ANCF/FFR model are elevated compared to the ANCF model. This

is due to the reduced deformation of the fluid, demonstrated in the sloshing box example. Because

the deformation of the ANCF/FFR is less than the ANCF fluid, the magnitude of the viscous forces

is smaller, resulting in reduced energy dissipation and increased sloshing amplitudes. However,

123

while the motion is not identical between the ANCF and ANCF/FFR meshes, the peak

displacements achieved during the lane change maneuver are similar, leading to similar effects on

the vehicle dynamics. This is evident in Fig. 4.11, which shows the normal forces, and Fig. 4.12,

which shows the lateral friction forces on a left-hand and right-hand tire of the three vehicle models.

The sloshing forces exerted by the flexible fluid on the tank propagate through the suspension to

the tires, resulting in oscillations in the contact forces. The peak magnitudes of the forces exerted

on the tires during the lane change maneuver match very closely. The rigid fluid model, however,

underpredicts the peak contact forces due to the absence of the sloshing, and no oscillation in the

forces occurs after the lane change is complete because the fluid is fixed to the tank.

(a) (b)

Figure 4.11. (a) Normal force on (a) a left-hand tire and (b) a right-hand tire during a lane change

( ANCF, ANCF/FFR, Rigid)

124

Figure 4.12. Lateral friction force on (a) a left-hand tire and (b) a right-hand tire during lane

change ( ANCF, ANCF/FFR, Rigid)

Table 4.3 lists the CPU times normalized with respect to the rigid fluid simulation CPU

time. The ANCF model simulation required over 5.5 times as long to run compared to the

ANCF/FFR model, highlighting the efficiency of the reduced-order ANCF/FFR fluid model

compared to the higher order ANCF model. It is clear from these results that although the reduced-

order ANCF/FFR fluid model does not capture the fluid deformation as accurately as the high-

order ANCF fluid mesh, the overall vehicle motion closely resembles that of the vehicle model

with ANCF fluid but at a reduced computational expense, demonstrating that the new ANCF/FFR

elements can effectively be used to model fluid sloshing in vehicle applications.

Table 4.3. Normalized vehicle model CPU times

Fluid Model Type Normalized CPU Time

Rigid 1

ANCF/FFR 785.9

ANCF 4,325.5

125

4.7. Concluding Remarks

Reduced-order models are often used in MBS and structural applications, however finite element

meshes using conventional elements with infinitesimal rotations as nodal coordinates do not

correctly capture the geometry in the reference configuration. For this reason, the geometrically-

exact ANCF/FFR elements were recently proposed in order to address this fundamental modeling

problem. In this chapter, the ANCF/FFR solid elements are integrated for the first time with a fully

nonlinear MBS vehicle algorithm. ANCF/FFR elements are able to capture initially curved

structures, such as the fluid within a cylindrical tank, while retaining the same number of degrees

of freedom as conventional elements and taking advantage of modal reduction techniques,

resulting in faster simulation times compared to the higher-order ANCF elements. In this chapter,

the solid element is developed in terms of constant geometric coefficients which are obtained using

the matrix of position vector gradients defined in the reference configuration. ANCF elements are

used as the basis for developing the ANCF/FFR elements, which are therefore related to B-splines

and NURBS by a linear mapping, and thus no geometry distortion occurs when CAD models are

converted to FE meshes. The procedure for writing the ANCF gradient vector coordinates in terms

of the FFR nodal rotations is described, which allows for effectively separating the initial geometry

from the displacement coordinates. Using this procedure, the new displacement shape function

matrix, which accounts for the initially curved structure, can be developed. The fluid constitutive

model, which is based on the Navier-Stokes fluid model, is detailed and the incompressibility

conditions, which are enforced using a penalty approach, are defined. The boundary conditions

between the fluid and tank, also enforced using a penalty approach, are detailed. Both normal and

tangential contact forces are included in this chapter. A model of a box filled with fluid which is

subjected to sinusoidal forcing is considered and the components of the MBS vehicle model

126

considered in this chapter, a medium-duty tanker truck with tank half filled with water, are

described. The fluid in both cases is meshed using ANCF and ANCF/FFR elements, and an

equivalent vehicle model in which the fluid is represented by a rigid body is also developed in

order to shed light on the effect of fluid sloshing on the vehicle dynamics when the different finite

element formulations are used.

The sloshing box model demonstrates that while the sloshing amplitudes of the ANCF/FFR

meshes are reduced compared to the converged ANCF fluid meshes, the general sloshing behavior

is still captured and at a significantly reduced CPU time, indicating that the elements may be useful

in certain applications such as vehicle dynamics where the capturing the exact shape of the free

surface is not critical and rapid simulation time is advantageous. This conclusion was confirmed

by the results of the highway vehicle lane change simulation – the sloshing amplitudes of the center

of mass predicted using the ANCF/FFR fluid mesh are found to be in a good agreement with what

predicted by the ANCF mesh. Furthermore, the results of the overall vehicle-dynamics, as

measured by the tire contact forces predicted using the two different meshes, are found to be in a

good agreement. Therefore, it is shown that if the goal is to accurately capture the free-surface

displacement of the fluid, then ANCF elements are better candidates due to their high order and

ability to capture complex shapes. However, if the goal is to perform efficient simulations in order

to obtain the overall vehicle motion, then using ANCF/FFR elements are a better alternative.

127

CHAPTER 5

SUMMARY AND CONCLUSIONS

In Chapter 2, a total Lagrangian ANCF fluid formulation that can be systematically integrated with

fully nonlinear MBS vehicle algorithms is proposed. The new approach can capture the fluid

distributed inertia and viscosity, can accurately predict the change in inertia due to the change in

shape of the fluid, and can visualize the change in the fluid free surface, unlike other discrete inertia

models which do not capture these significant details. The outward forces on the fluid during curve

negotiation are derived and it is shown that these forces do not take the same simple form as the

case of a rigid body negotiating a curve. As discussed in this chapter, accurate modeling of the

fluid geometry using ANCF elements can be achieved without the need for using computational

geometry methods such as B-spline and NURBS representations which have a rigid recurrence

structure unsuitable for MBS analysis. By using the approach proposed in this chapter, one

geometry/analysis mesh is used from the outset. The fluid constitutive law and the fluid/tank

interaction forces are developed. The penalty method is used to ensure that the fluid remains within

the boundaries defined by the tank geometry. Both normal and tangential penalty contact forces

are considered in this study. The MBS vehicle model components are described and the dynamics

of the vehicle is examined using three contrasting motion scenarios in order to study the effect of

sloshing on vehicle dynamics. The braking scenario examines the case of longitudinal sloshing of

the fluid, a rapid lane change produces alternating lateral fluid forces on the tank, and curve

negotiation sheds light on the case of steady-state outward forces due to the centrifugal effect.

The results presented in this chapter demonstrate that depending on the scenario, the

sloshing phenomenon can increase the contact forces on some wheels while decreasing contact

128

forces on other wheels, and this can lead to vehicle instability. In the case of brake applications,

the fluid in the partially-filled tank surges forward, causing uneven wheel loading; in cases of

severe braking, wheel lift may occur on the rear wheels of the vehicle. This can lead to difficulty

controlling the vehicle and increased stopping distances due to lessened road contact, and the

decrease in stability may result in jack-knifing for tractor-trailer vehicles. When entering a curve

or performing a lane change, the outward centrifugal forces cause lateral displacement of the fluid,

which also causes uneven wheel loading. In extreme cases of high speed or small radius of

curvature, these changes could be significant enough to induce wheel lift on the inner wheels and

increase the possibility of rollover compared to an equivalent truck carrying rigid materials.

Furthermore, in cases where tire friction forces are decreased, such as on wet or icy roads, the

possibility for vehicle instability increases even further and drivers must exercise extra caution.

Future analysis can result in defining general rules for speed reduction of a tanker truck entering a

curve in order to maximize vehicle stability and driver safety, proposing modifications to the tank

geometry to reduce sloshing amplitudes and forces, studying the effect of viscosity and

incompressibility on the fluid dynamic behavior, and comparison with conventional finite

elements as well as the smoothed particle hydrodynamics (SPH) method (Wasfy et al., 2014).

In Chapter 3, the newly proposed approach for the integration of a continuum-based

sloshing model with vehicle models is presented using a fully nonlinear MBS rail vehicle model.

The contributions of this chapter are as follows: (1) A unified geometry/analysis mesh is used from

the outset in order to accurately capture complex fluid and rail tank-car geometries as well as the

nonlinear dynamic behavior of the fluid and vehicle. The approach developed in this chapter is

used to examine the effects of liquid sloshing on railroad vehicle dynamics when negotiating a

curve and during traction or braking; (2) The method of the tank-car and fluid geometry description

129

is introduced and it is shown how a unified geometry/analysis mesh can be developed for both the

rigid rail and continuum fluid bodies. The search method used to define the fluid/tank contact

points is outlined and the penalty force model used to describe the fluid/tank interaction forces is

formulated; (3) The fluid constitutive equations that account for the viscosity and incompressibility

effects are presented. The liquid sloshing model developed in this chapter is integrated with the

MBS railroad vehicle model which takes into consideration the nonlinear three-dimensional

wheel/rail contact forces and the wheel and rail profile geometries; and (4) In order to

systematically examine the effect of the motion of the flexible fluid on vehicle dynamics when the

vehicle is negotiating a curve, a general definition of the outward inertia force of a flexible body

using both FFR and ANCF descriptions is investigated. The analysis presented in this chapter

shows that this force depends strongly on the motion of the continuum and does not take the simple

form used in the case of rigid body dynamics.

Comparative simulations were performed to examine the liquid sloshing effects by using

flexible and rigid body fluid models. It is shown that the liquid sloshing can exacerbate the

unbalanced effects when the vehicle travels at a velocity away from the balance speed, but this

effect decreases when the forward velocity is close to the balance speed because the liquid

experiences the same centrifugal force as the rigid fluid body in this case. The results in the traction

analysis show that the liquid motion can significantly affect the load distribution between the front

and rear trucks. Comparing with the ECP braking case, there is a larger coupler force when the

conventional braking is used for both the flexible and rigid body fluid models. Nonetheless, the

results obtained for the model considered in this chapter demonstrate that the liquid sloshing

amplifies the coupler force greatly in the ECP braking case compared to the equivalent rigid body

model because the latter model cannot capture the fluid nonlinear inertia effects.

130

In Chapter 4, the ANCF/FFR elements are integrated for the first time with a fully nonlinear

MBS vehicle algorithm. ANCF/FFR elements are able to capture initially curved structures, such

as the fluid within a cylindrical tank, while retaining the same number of degrees of freedom as

conventional elements and taking advantage of modal reduction techniques, resulting in faster

simulation times compared to the higher-order ANCF elements. In this paper, the solid element is

developed in terms of constant geometric coefficients which are obtained using the matrix of

position vector gradients defined in the reference configuration. ANCF elements are used as the

basis for developing the ANCF/FFR elements, which are therefore related to B-splines and

NURBS by a linear mapping, and thus no geometry distortion occurs when CAD models are

converted to FE meshes. The procedure for writing the ANCF gradient vector coordinates in terms

of the FFR nodal rotations is described, which allows for effectively separating the initial geometry

from the displacement coordinates. Using this procedure, the new displacement shape function

matrix, which accounts for the initially curved structure, can be developed. The fluid constitutive

model, which is based on the Navier-Stokes fluid model, is detailed and the incompressibility

conditions, which are enforced using a penalty approach, are defined. The boundary conditions

between the fluid and tank, also enforced using a penalty approach, are detailed. Both normal and

tangential contact forces are included in this study. A model of a box filled with fluid which is

subjected to sinusoidal forcing is considered and the components of the MBS vehicle model

considered in this investigation, a medium-duty tanker truck with tank half filled with water, are

described. The fluid in both cases is meshed using ANCF and ANCF/FFR elements, and an

equivalent vehicle model in which the fluid is represented by a rigid body is also developed in

order to shed light on the effect of fluid sloshing on the vehicle dynamics when the different finite

element formulations are used.

131

The sloshing box model demonstrates that while the sloshing amplitudes of the ANCF/FFR

meshes are reduced compared to the converged ANCF fluid meshes, the general sloshing behavior

is still captured and at a significantly reduced CPU time, indicating that the elements may be useful

in certain applications such as vehicle dynamics where the capturing the exact shape of the free

surface is not critical and rapid simulation time is advantageous. This conclusion was confirmed

by the results of the highway vehicle lane change simulation – the sloshing amplitudes of the center

of mass predicted using the ANCF/FFR fluid mesh are found to be in a good agreement with what

predicted by the ANCF mesh. Furthermore, the results of the overall vehicle-dynamics, as

measured by the tire contact forces predicted using the two different meshes, are found to be in a

good agreement. Therefore, it is shown that if the goal is to accurately capture the free-surface

displacement of the fluid, then ANCF elements are better candidates due to their high order and

ability to capture complex shapes. However, if the goal is to perform efficient simulations in order

to obtain the overall vehicle motion, then using ANCF/FFR elements are a better alternative.

132

APPENDIX A

Permission for use of author’s previously published Journal of Sound and Vibration article in this

thesis.

133

134

APPENDIX B

Permission for use of author’s previously published Journal of Multi-Body Dynamics article in

this thesis.

135

136

APPENDIX C

The three-dimensional ANCF solid element, with an incomplete polynomial representation, used

in this investigation is an 8-node element. The nodal coordinates jk

e at the node k of the finite

element j can be defined as

T T T T T

1, ,8jk jk jk jk jk

x y z k = =

e r r r r (C. 1)

where jk

r is the absolute position vector at the node k of the element j , and jk

xr , jk

yr and jk

zr are

the position vector gradients obtained by differentiation with respect to the spatial coordinates ,x y

and z , respectively. The displacement field of each coordinate of the solid fluid element can be

defined using an incomplete polynomial with 32 coefficients as

( ) 2 2 2

1 2 3 4 5 6 7 8 9 10

3

3 3 3 2 2 2 2 2 2

11 12 13 14 15 16 17 18 19

3 3 3 3 3 3 2

20 21 22 23 24 25 26 27

2 2

28 29 30 31

3

, ,x y z x y z x y z xy yz xz

x y z x y x z y z xy xz yz

xyz x y x z xy y z x

x

z yz x yz

xy z xy yzz xy

= + + + + + + + + +

+ + + + + + + + +

+ + + + + + + +

+ ++ + 3

32z xyz+

(C. 2)

In this equation, , 1, 2, ,32k k = , are the polynomial coefficients. Using this polynomial

description, the shape functions of the ANCF brick fluid element can be derived as follows:

( ) ( )( )( )

( )( ) ( )( ) ( )( )( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

( ) ( ) ( ) ( )

1,1

2 1 11,2

1 2 1,3

1 1 21,4

1

1 1 1 1

1 1 2 1 2 1 2

1 1 1 1

1 1 1 1

1 1 1 1

k

k

k

k k k

k k k k kk k

k k k k kk k

k k k k kk k

k

k k k

k k k

k

k

k

S

S a

S b

S c

+ + +

+ − −

+

−+

+ − − −

+ − − −+

= − + − + − + −

+

− − + − − + − −

= − − − −

= − − − −

= − − − −

1,2, ,8k

=

(C. 3)

where ,a b , and c are, respectively, the dimensions of the element along the axes ,x y , and z

directions, / , / , /x a y b z c = = = , , , 0,1 , and , ,k k k are the dimensionless

137

nodal locations for node k . The position vector of an arbitrary material point on element j can be

written as

8,1 ,2 ,3 ,4

1

j k k k k jk j j

k

S S S S=

= = r I I I I e S e (C. 4)

where I is the 3 3 identity matrix, j

S and j

e are, respectively, the element shape function

matrix and the vector of nodal coordinates which can be written as

T T T T T T T T

1,1 1,2 1,3 1,4 8,1 8,2 8,3 8,4

1 2 3 4 5 6 7 8

j

Tj j j j j j j j j

S S S S S S S S =

=

S I I I I I I I I

e e e e e e e e e (C. 5)

138

REFERENCES

1. Aboubakr, A.K., Volpi, M., and Shabana, A.A., 2015, “Efficient and robust implementation of

the TLISMNI method”. Journal of Sound and Vibration, 353, pp. 220-242.

2. Aboubakr, A.K., Volpi, M., Shabana, A.A., Cheli, F., and Melzi, S., 2016, “Implementation of

electronically controlled pneumatic brake formulation in longitudinal train dynamics

algorithms”. Proceedings of the Institution of Mechanical Engineers, Part K: Journal of Multi-

body Dynamics, online first, DOI: 1464419316628764

3. Abramson, H.N., 1966, “The dynamic behaviour of liquids in moving containers”. NASA SP-

106.

4. Aliabadi, S., Johnson, A., and Abedi, J., 2003, “Comparison of finite element and pendulum

models for simulation of sloshing”. Computers and Fluids, 32, pp. 535-545.

5. ANSYS, 2019, “ANSYS Fluent”. https://www.ansys.com/products/fluids/ansys-fluent

(accessed 6/11/2019).

6. Ashtianti, I.H., Rakheja, S., and Ahmed, A.K.W., 2015, “Hunting and derailment analysis of

a partly-filled railway tank car”, International Heavy Haul Association Conference, Perth,

Western Australia.

7. Atif, M.M., Chi, S.W., Grossi, E., and Shabana, A.A., 2019, “Evaluation of breaking wave

effects in liquid sloshing problems: ANCF/SPH comparative study”, Nonliner Dyn DOI:

10.1007/s11071-019-04927-5.

8. Bolarinwa, E.O., Mahadevaiah, E., Marzougui, D., and Opiela, K.S., 2012, “The development

of an enhanced finite element tire model for roadside safety hardware assessment”. Journal of

Multibody Dynamics, 226, pp. 206-219.

139

9. Boffi, D. and Gastaldi, L., 2016, “Discrete models for fluid-structure interactions: the finite

element immersed boundary method”, Discrete and Continuous Dynamics Systems: Series S

9(1), pp. 89-107, DOI: 10.3934/dcdss.2016.9.89.

10. Bruzelius, F., Hjort, M., and Svendenius, J., 2013, “Validation of a basic combined-slip tyre

for use in friction estimation applications”. Journal of Automobile Engineering, 228, pp. 1622-

1629.

11. Cottrell, J.A., Hughes, T.J.R., and Reali, A., 2007, “Studies of Refinement and Continuity in

the Isogeometric Analysis”, Comput. Methods in Appl. Mech. Engrg., 196, pp. 4160-4183.

12. Chen, G.F., and Wu, C.H., 2011, “Effects of excitation angle and coupled heave–surge–sway

motion on fluid sloshing in a three-dimensional tank”, J. Mar. Sci. Technol 16, pp. 22-50, DOI:

10.1007/s00773-010-0111-0.

13. Cui, D., Yan, S., Guo, X., and Gao, R., 2014, “Parametric resonance of liquid sloshing in

partially filled spacecraft tanks during the powered-flight phase of rocket”. Aerospace Science

and Technology, 35, pp. 93-105.

14. Cundall, P.A. and Strack, O.D.L., 1979, “A discrete numerical model for granular assemblies”,

Geotechnique 29(1), pp. 47-65.

15. De-Juan, A., Sancibrian, R., and Viadero, F., 2012, “Optimal synthesis of function generation

in steering linkages”. International Journal of Automotive Technology, 13, pp. 1033-146.

16. Di Gialleonardo, E., Premoli, A., Gallazi, S., and Bruni S., 2013, “Sloshing effects and running

safety in railway freight vehicles”, Vehicle System Dynamics 51(10), pp. 1640-1654, DOI:

10.1080/00423114.2013.814797.

17. Dodge, 2000, “The new ‘dynamic behavior of liquids in moving containers’”, Southwest

Research Institute, San Antonio, Texas.

140

18. Gialleonardo, E.D., Premoli, A., Gallazi, S., and Bruni S., 2013, “Sloshing effects and running

safety in railway freight vehicles”, Vehicle System Dynamics 51(10), pp. 1640-1654, DOI:

10.1080/00423114.2013.814797.

19. Godderidge, B., Turnock, S., and Mingyi, T., 2011, “A rapid method for the simulation of

sloshing using a mathematical model based on the pendulum equation”, Computers & Fluids

57, pp. 163-171.

20. Graham, E.W., 1951, “The forces produced by fuel oscillations in a rectangular tank”. Douglas

Aircraft Cooperation, SM-13748.

21. Graham, E.W. and Rodríguez, A.M., 1952, “The characteristics of fuel motion which affect

airplane dynamics”. Journal of Applied Mechanics, 74, pp. 381-388.

22. Griffiths, W.D. and Boysan, F., 1996, “Computational Fluid dynamics (CFD) and empirical

modelling of the performance of a number of cyclone samplers”, J. Aerosol Sci. 27(2), pp. 281-

304.

23. Grossi, E. and Shabana, A.A., 2017, “Validation of a total lagrangian ANCF solution

procedure for fluid-structure interaction problems”, J. Verif. Valid. Uncertain. Quantif. 2(4),

041001-1-041001-13.

24. Grossi, E. and Shabana, A.A., 2018, “ANCF analysis of the crude oil sloshing in railroad

vehicle systems”, Journal of Sound and Vibration 433, pp. 493-516.

25. Ibrahim, R.A., 2005, Liquid Sloshing Dynamics: Theory and Applications, Cambridge

University Press.

26. Ibrahim, R.A., Pilipchuk, V.N., and Ikeda, T., 2001, “Recent advances in liquid sloshing

dynamics”. Applied Mechanics Reviews, 54, pp. 133-199.

141

27. Kalker, J.J., 1990, Three-Dimensional Elastic Bodies in Rolling Contact, Kluwer Academic

Publishers, Dordrecht, The Netherlands.

28. King, C. and Trichur, R., 2015, “Train carrying crude oil derails in Ontario”. WSJ,

http://www.wsj.com/articles/train-carrying-crude-oil-derails-in-ontario-1424015961.

29. Liu, M.B. and Liu, G.R., 2010, “Smoothed particle hydrodynamics (SPH): an overview and

recent developments”. Archives of Computational Methods in Engineering, 17, pp. 25-76.

30. Mackenzie, D., 2012, “Curing ill surfaces”, SIAM News, 6 April 2011.

31. Monaghan, J.J., 2012, “Smoothed particle hydrodynamics and its diverse applications”, Annu.

Rev. Fluid Mech. 44, pp. 323-46, DOI: 10.1146/annurev-fluid-120710-101220.

32. Nichkawde, C., Harish, P.M, and Ananthkrishnan, N., 2004, “Stability analysis of a multibody

system model for coupled slosh-vehicle dynamics”. Journal of Sound and Vibration, 275, pp.

1069-1083.

33. Nicolsen, B., Wang, L., and Shabana, A.A., 2017, “Nonlinear finite element analysis of liquid

sloshing in complex vehicle motion scenarios”. Journal of Sound and Vibration, 405, pp. 208-

233.

34. Nishiura, D. and Sakagushi, H., 2014, “Wet dispersion mechanism of fine aggregates in

multiphase flow with solid beads under simple shear”, AIChE Journal 60(12), pp. 4076-4085.

35. Olshevskiy, A., Dmitrochenko, O., and Kim, C.W., 2013, “Three-dimensional solid brick

element using slopes in the absolute nodal coordinate formulation”. ASME Journal of

Computational and Nonlinear Dynamics, 9, doi:10.1115/1.4024910

36. Pacejka, H.B.,2006, Tyre and Vehicle Dynamics, Elsevier, Oxford, England, Second Edition.

142

37. Pape, D., Thornton, B., and Yugulis, K., 2016, “Slosh characteristics of aggregated

intermediate bulk containers on single-unit trucks”. Federal Motor Carrier Safety

Administration, U.S. Department of Transportation, Washington, D.C.

38. Patel, M., Orzechowski, G., Tian, Q., and Shabana, A.A., 2016, “A new multibody system

approach for tire modeling using ANCF finite elements”. Proceedings of the Institution of

Mechanical Engineers, Part K: Journal of Multi-body Dynamics, 230, pp. 69-84.

39. Pelc, J., 2007, “Towards realistic simulation of deformations and stresses in pneumatic tyres”.

Applied Mathematical Modelling, 31, pp. 530-540.

40. Petravic, J., 2007, “On the equilibrium calculation of the friction coefficient for liquid slip

against a wall,” Journal of Chemical Physics, Vol. 127.

41. Piegl, L. and Tiller, W., 1997, The NURBS Book, Second Edition, Springer, New York.

42. Pit, R., Hervet H., and Leger, L., 1999, “Friction and slip of a simple liquid at a solid surface,”

Tribology Letters, pp. 147-152.

43. Ranganathan, R., Rakheja, S., and Sankar, S., 1989, “Steady turning stability of partially filled

tank vehicles with arbitrary tank geometry”. Journal of Dynamic Systems, Measurement, and

Control, 111, pp. 481–489.

44. Sanborn, G., Tobaa, M. and Shabana, A.A., 2008, “Coupling between structural deformations

and wheel-rail contact geometry in railroad vehicle dynamics”. Proceedings of the Institution

of Mechanical Engineers, Part K: Journal of Multi-body Dynamics, 222(4), p.381-392.

45. Shabana, A.A., 2013, Dynamics of Multibody Systems, Cambridge University Press, Oxford,

England, Fourth Edition.

46. Shabana, A.A., 2015, “ANCF tire assembly model for multibody system applications”.

Journal of Computational and Nonlinear Dynamics, 10, ID 024504.

143

47. Shabana, A.A., 2017A, Computational Continuum Mechanics, Cambridge University,

Cambridge, UK, Third edition.

48. Shabana, A.A., 2017B, “Geometrically accurate floating frame of reference finite elements for

the small deformation problem”, Proc. Inst. Mech. Eng. Part K: J. Multibody Dyn. 232(2), pp.

286–292.

49. Shabana, A.A., 2018, “Geometrically accurate infinitesimal-rotation spatial finite elements”,

Proc. Inst. Mech. Eng. Part K: J. Multibody Dyn. 233(1) pp. 182-187, DOI:

10.1177/1464419318774948.

50. Shabana, A.A., Tobaa, M., Sugiyama, H. and Zaazaa, K., 2005. “On the Computer

Formulations of the Wheel/Rail Contact Problem”. Nonlinear Dynamics, 40(2), p.169-193.

51. Shabana, A.A., Wang, G., and Kulkarni, S., 2018, “Further investigation on the coupling

between the reference and elastic displacements in flexible body dynamics”, Journal of Sound

and Vibration 427, pp. 159–777, DOI: 10.1016/j.jsv.2018.02.054.

52. Shabana, A.A., Zaazaa, K.E., and Sugiyama, H., 2008, Railroad Vehicle Dynamics: A

Computational Approach, Taylor & Francis/CRC, Boca Raton, FL.

53. Shen, X., Yan, Y., Li, X., Xie, C., Wang, L., 2014, “Analysis on tank truck accidents involved

in road hazardous materials transportation in China”. Traffic Injury Prevention, 15, pp. 762-

768.

54. Shi, H., Wang, L., Nicolsen, B., Shabana, A.A., 2017, “Integration of geometry and analysis

for the study of liquid sloshing in railroad vehicle dynamics”. Proceedings of the Institution of

Mechanical Engineers, Part K: Journal of Multi-body Dynamics, 231(4), pp. 608-629.

55. Spencer, A.J.M., 1980, Continuum Mechanics, Longman, London, England.

144

56. Sugiyama, H., and Shabana, A., A., 2007, “Trajectory Coordinate Constraints in Multibody

Railroad Vehicle Systems”, Journal of System Design and Dynamics 1(3), DOI:

10.1299/jsdd.1.481.

57. Tinsley, B. and Shabana, A.A., 2019, “Eigenvalue convergence of the geometrically accurate

spatial beam, plate, and solid elements”, Technical Report # MBS2019-1-UIC.

58. U.S. Department of Transportation, Federal Motor Carrier Safety Administration, 2007,

“Cargo tank rollover myths and truths”. https://www.fmcsa.dot.gov/sites/fmcsa.dot.gov/files/

docs/FINAL_Cargo_Tank_Rollover_Fact_Sheet_508.pdf (accessed 2.16.17).

59. U.S. Department of Transportation, Federal Highway Administration, 2011, “Freight facts and

figures”. http://www.ops.fhwa.dot.gov/freight/freight_analysis/nat_freight_stats/docs/

11factsfigures/pdfs/fff2011_highres.pdf (accessed 2.16.17).

60. U.S. Department of Transportation, Pipeline and Hazardous Materials Safety Administration,

2011, “Hazardous Materials Incident Report”, Incident ID E-2011080424. Accessed 5/31/2019

(https://portal.phmsa.dot.gov/xmlpserver/loading.jsp?_xdo=%2FEIA+Forms%2

F5800Incident.xdo&_id=1410e6fe-d686-4593-b73c-7176221c2c83&_genstatus=).

61. U.S. Department of Transportation, National Highway and Traffic Safety Administration,

2015, “Electronic stability control systems for heavy vehicles”. https://www.transportation.

gov/sites/dot.gov/files/docs/FMVSS-136-Final-Rule-05182015-final.pdf (accessed 2/16/17).

62. U.S. Department of Transportation, Pipeline and Hazardous Materials Safety Administration,

2015, “10 year incident summary reports”. https://hip.phmsa.dot.gov/analyticsSOAP/saw.dll?

Dashboard&NQUser=HazmatWebsiteUser1&NQPassword=HazmatWebsiteUser1&PortalPa

th=/shared/Public%20Website%20Pages/_portal/10%20Year%20Incident%20Summary%20

Reports (accessed 2.16.17).

145

63. Vera, C., Paulin, J., Suarez, B., and Gutiérrez, M., 2005, “Simulation of freight trains equipped

with partially filled tank containers and related resonance phenomenon”. Proceedings of the

Institution of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit, 219(4), pp.

245-259.

64. Wang, L., Jimenez-Octavio, J.R., Wei, C., and Shabana, A.A., 2014, “Low Order Continuum-

Based Liquid Sloshing Formulation for Vehicle System Dynamics”, Journal of Computational

and Nonlinear Dynamics, 10(2), ID 021022.

65. Wang, L., Jimenez-Octavio, J.R., Wei, C., and Shabana, A.A., 2015, “Low order continuum-

based liquid sloshing formulation for vehicle system dynamics”. Journal of Computational

and Nonlinear Dynamics, 10, ID 021022.

66. Wang, L., Shi, H. and Shabana, A.A., 2016, “Effect of the tank-car thickness on the nonlinear

dynamics of railroad vehicles”. Proceedings of the Institution of Mechanical Engineers, Part

K: Journal of Multi-body Dynamics, DOI: 10.1177/1464419316642930.

67. Wasfy, T.M., Wasfy, H.M., and Peters, J.M., 2014, “Coupled multibody dynamics and

smoothed particle hydrodynamics for predicting liquid sloshing for tanker trucks”, ASME

DETC2014-35191, Proceedings of the ASME 2014 International Design Engineering

Technical Conferences & Computers and Information in Engineering Conference, Buffalo,

New York.

68. Wei, C., Wang, L., and Shabana, A.A., 2015, “A total Lagrangian ANCF liquid sloshing

approach for multibody system applications”. Journal of Computational and Nonlinear

Dynamics, 10(5), ID 051014.

69. White, F.M., Fluid Mechanics, 2011, The McGraw-Hill Companies, Inc., New York, Seventh

edition.

146

70. Wronski, R., 2013, “As more oil flows by rail, concerns grow about safety of the tank-cars”.

tribunedigital-chicagotribune, http://articles.chicagotribune.com/2013-11-29/news/ct-

dangerous-tank-cars-met-1129-20131129_1_tank-cars-oil-production-more-oil-flows.

71. WSB-TV Atlanta, Driver dies in fiery tanker crash on I-985 in Hall County,

http://www.wsbtv.com/news/local/tanker-truck-fire-shuts-down-i-985-hall-county/17241607,

2015 (accessed 2.16.17).

72. Xue-Lian, Z., Xian-Sheng, L., and Yuan-Yuan, R., 2012, “Equivalent mechanical model for

lateral liquid sloshing in partially filled tank vehicles”, Mathematical Problems in Engineering,

Article ID: 162825, DOI: 10.1155/2012/162825.

73. Yang, S.H., Lee, S.B., Park, J.H., Han, S.Y., Choi, Y.M., DO, J., Kwon, S.H., and Molin, B.,

2016, “Experimental study on piston- and sloshing-mode moonpool resonances”, J. Mar. Sci.

Technol. 21, pp. 715-728, DOI: 10.1007/s00773-016-0386-x.

74. Zhang, Z., Wang, T., and Shabana, A.A., 2019, “Development and implementation of

geometrically accurate reduced-order models: Convergence properties of planar beams”,

Journal of Sound and Vibration 439, pp. 457–478, DOI: 10.1016/j.jsv.2018.06.005.

75. Zheng, X., Li, X., and Ren, Y., 2012, “Equivalent mechanical model for lateral liquid sloshing

in partially filled tank vehicles”. Mathematical Problems in Engineering, ID 162825.

147

VITA

Brynne Elizabeth Nicolsen

EDUCATION

University of Illinois at Chicago (UIC)

• Ph.D. in Mechanical Engineering; Advisor: Prof. Ahmed A. Shabana || Aug. 2015 – July 2019

GPA: 4.0/4.0

• B.S. in Bioengineering, Minor in Mechanical Engineering || Aug. 2011 – May 2015

GPA: 4.0/4.0, Summa Cum Laude

RELEVANT COURSEWORK

Finite Element Analysis I and II – Nonlinear Finite Element Analysis – Dynamics of Mechanical Systems –

Computer-Aided Analysis of Multibody Systems – Vibration of Discrete and Continuous Systems –

Continuum Mechanics – Theory of Elasticity – Plasticity – Numerical Analysis – Mathematical Theory of Finite

Elements – Mathematical Methods for Engineers

SKILLS

Software: Experienced: SIGMA/SAMS, ANSYS, PTC Creo, Matlab, Fortran || Familiar: MSC Adams,

HyperMesh, NX, TeamCenter

Communication: Exemplary written and oral communication

Interpersonal: Strong leadership qualities, skilled at instruction, collaborates well in groups

RESEARCH EXPERIENCE

Graduate

Dynamic Simulation Laboratory || UIC || June 2015 – July 2019

Comparison of Small- and Large-Deformation Finite Elements for the Study of Fluid Sloshing || Jan. 2016 –

July 2019

• Developed general contact algorithm applicable to rigid or flexible bodies of different finite element types

• Compared fluid sloshing behavior using different finite element formulations in highway vehicle system

applications

Fluid Sloshing Phenomena using Large-Deformation Finite Elements || June 2015 – Dec. 2016

• Developed medium-duty tanker truck model with fluid represented by ANCF continuum finite elements

• Ran model under different scenarios to investigate the effects of fluid-tank interaction on vehicle

dynamics and stability

Undergraduate

Dynamic Simulation Laboratory || UIC || Oct. 2014 – May 2015

• Implemented DEM (discrete element method) module in the general-purpose flexible MBD software

SIGMA/SAMS

• Modeled tire-soil interaction using the DEM approach and ANCF finite elements

• Developed Fortran subroutines including interparticle contact detection and Verlet numerical integration

148

PUBLICATIONS

• Nicolsen, B., Shi, H., Wang, L., Shabana, A., "Integration of Geometry and Analysis for the Study of Liquid Sloshing in Vehicle System Dynamics," Proceedings of the 2017 ASME International Mechanical Engineering Congress and Exposition, November 3-9, 2017, Tampa, FL, USA.

• Nicolsen, B., Wang, L., and Shabana, A., “Nonlinear Finite Element Analysis of Liquid Sloshing in Complex Vehicle Motion Scenarios,” Journal of Sound and Vibration, Vol. 405, May 2017.

• Shi, H., Wang, L., Nicolsen, B., and Shabana, A.A., “Integration of Geometry and Analysis for the Study

of Liquid Sloshing in Railroad Vehicle Dynamics”, Proc IMechE Part K: J Multibody Dynamics, Online

March 2017.

• Contreras, U., Nicolsen, B., Tian, Q., Recuero, A., Shabana, A., “Integration of ANCF and Discrete

Element Method for Multibody Vehicle Applications,” Proceedings of the ASME 2015 International Design

Engineering Technical Conference & Computers and Information in Engineering Conference, August 2-

5, 2015, Boston, MA, USA.

INTERNSHIP EXPERIENCE

Graduate

Navistar, Inc.: Vehicle Dynamics Engineer || Lisle, IL || May 2016 – Aug. 2016

• Developed 20 steer axle and 34 leaf spring model libraries using newly-enhanced modeling techniques

to increase speed of model development

• Assisted in building state-of-the-art hardware-in-the-loop setup to simulate electronic stability control of

class A tractors in support of FMVSS136 vehicle stability regulations

• Characterized test data of 13 tires from a variety of manufacturers in order to develop more accurate

virtual tire models

• Processed and analyzed vehicle test results for use in model correlation and validation

Undergraduate

Hospira: Systems Test Engineer || Lake Forest, IL || May 2016 – Aug. 2016

• Worked with engineers to write, test, and edit protocols for the testing of infusion pumps for re-

manufacture

TEACHING ASSISTANTSHIPS

Graduate

Introduction to Engineering Design and Graphics || UIC || Jan. 2019 – May 2019

• Co-managed a group of 17 undergraduate TAs in developing coursework material and assisting during

lectures

• Developed new methodologies to improve TA productivity and efficiency

• Led discussions on fair TA compensation and redesigning course content

Professional Development Seminar || UIC || Jan. 2016 – May 2017

• Met with students weekly to aid in developing resumes and cover letters and in securing full-time jobs

• Collaborated with three other TAs to maintain course website, record presentations, and ensure lectures

ran smoothly

149

Introductory Vibration Theory || UIC || Aug. 2015 – May 2016

• Guided students in learning to formulate and solve the differential equations of motion for vibrational

systems

• Assisted professor in preparing and grading assignments, projects, and exams, and maintaining course

website

Undergraduate

Engineering Orientation || UIC || Aug. 2013 – Dec. 2013

• Educated a group of freshman students about the field of bioengineering and the curriculum at UIC

• Assisted students in developing their first resumes and applying to summer internships

PRESENTATIONS

• U.S. National Congress for Theoretical and Applied Mechanics, June 4-9, 2018, Chicago IL

Integration of Geometry and Analysis for the Study of Liquid Sloshing in Complex Vehicle Systems

• ASME International Mechanical Engineering Congress and Exposition, November 3-9, 2017, Tampa, FL

Integration of Geometry and Analysis for the Study of Liquid Sloshing in Vehicle System Dynamics

• Workshop on Computational and Nonlinear Dynamics for Industry, May 27, 2016, Chicago, IL

Liquid Sloshing and the Contact Problem

PROFESSIONAL SERVICE

Journal Reviewer:

• Journal of Multi-Body Dynamics, Feb. 2017 || Vehicle System Dynamics, Sept. 2017 || Journal of Rail

and Rapid Transit, July 2018 || Journal of Sound and Vibration, July 2018 || Journal of Vibration and

Acoustics, Sept. 2018

Conference Reviewer:

• ASME International Gas Turbine Institute Exposition, June 2019 (2 papers)

EXTRACURRICULAR ACTIVITIES

Pi Tau Sigma: Mechanical Engineering Honors Society || Member Oct. 2018 – Present

American Society for Engineering Education || Member Aug. 2018 – Present

Tau Beta Pi: Engineering Honors Society || Recording Secretary Apr. 2014 – May 2015; Member Nov. 2012

– Present

UIC Bioengineering Student Journal: Editor Jan. 2013 – May 2016, Exec. Editor Jan 2014 – May 2014,

Author May 2014