You are on page 1of 116

GEOMETRICALLY NONLINEAR MULTI-PATCH

ISOGEOMETRIC ANALYSIS OF BEAMS: TIMOSHENKO


AND EULER-BERNOULLI BEAM THEORIES

BY

MR. DUY VO

A DISSERTATION SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF
PHILOSOPHY (ENGINEERING AND TECHNOLOGY)

SIRINDHORN INTERNATIONAL INSTITUTE OF TECHNOLOGY

THAMMASAT UNIVERSITY

ACADEMIC YEAR 2020

COPYRIGHT OF THAMMASAT UNIVERSITY

Ref. code: 25636122300053IAQ


GEOMETRICALLY NONLINEAR MULTI-PATCH
ISOGEOMETRIC ANALYSIS OF BEAMS: TIMOSHENKO
AND EULER-BERNOULLI BEAM THEORIES

BY

MR. DUY VO

A DISSERTATION SUBMITTED IN PARTIAL FULFILLMENT OF


THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF
PHILOSOPHY (ENGINEERING AND TECHNOLOGY)

SIRINDHORN INTERNATIONAL INSTITUTE OF TECHNOLOGY

THAMMASAT UNIVERSITY

ACADEMIC YEAR 2020

COPYRIGHT OF THAMMASAT UNIVERSITY

Ref. code: 25636122300053IAQ


THAMMASAT UNIVERSITY
SlRINDI JORN INTERNATlONAL INSTITUTE OF TECHNOLOGY

DISSERTATION

BY

MR.DUY VO

E:--JTITLED

GEOMETRICALLY NONLINEAR MULTI-PATCH lSOGEOMETRIC ANALYSIS


OF BEAMS: TIMOSHENKO AND EULER-BERNOULLI BEAM THEORIES

was approved as partial fulfillment of the requirements for the degree of Doctor of
Philosophy (Engineering and Technology)

on November 17, 2020

Chairperson
(Associate Professor Jaroon Rungamomrat, Ph.D.)

Member and Advisor �- A -


(Professor Pruettha Nanakom, D.Eng.)
/ 1/d:::,-_
Member and Co-Advisor ----------
fl--/
(Associate Professor Tinh Quoc Bui, D.techn.)

Member 'f4-,
(Associate Profess.9r Nakhorn Po?varodom, Ph.D.)
--
Member )l� �.
(Research Assistant �ofc�e)Pakawat Sancharoen, Ph.D.)

Member n<-r-, (uU


�-
(Assistant Professor Ganchai Tanapornraweekit, Ph.D.)

Director
(Professor Pruettha Nanakom, D.Eng.)
(1)

Thesis Title GEOMETRICALLY NONLINEAR


MULTI-PATCH ISOGEOMETRIC
ANALYSIS OF BEAMS:
TIMOSHENKO AND EULER-
BERNOULLI BEAM THEORIES

Author Mr. Duy Vo

Degree Doctor of Philosophy (Engineering and


Technology)

Faculty/ University Sirindhorn International Institute of


Technology/ Thammasat University

Thesis Advisor Professor Pruettha Nanakorn, D.Eng.

Thesis Co-Advisor Associate Professor Tinh Quoc Bui,


D.techn.

Academic Years 2020

ABSTRACT

Isogeometric analysis (IGA) is a recently developed numerical technique that aims


at unifying geometric design and computational finite element analysis into one model.
The technique is conducted by using the same basis functions for both computer-aided
design and finite element analysis. In this study, two new efficient Timoshenko and Euler-
Bernoulli beam formulations are presented in the context of IGA for geometrically
nonlinear analysis of spatial beam structures.

A novel Timoshenko beam formulation is proposed by means of the total


Lagrangian description using the Green-Lagrange strain tensor and the second Piola-

Ref. code: 25636122300053IAQ


(2)

Kirchhoff stress tensor as the energy conjugate pair. This approach facilitates consideration
of hyperelastic materials in analysis of highly flexible beam structures. Although in the
context of the conventional finite element approach, many beam formulations have been
developed by using the Green-Lagrange strain and second Piola-Kirchhoff stress tensors,
there virtually exist no isogeometric Timoshenko beam formulations that are derived by
using this conjugate pair. Three-dimensional beam configurations are reduced into one-
dimensional structures using the beam axis and director vectors of the cross-sections. The
displacements of the beam axis and the total cross-sectional rotation are considered as
unknown kinematics. The cross-sectional rotation is represented by an orthogonal tensor,
which is parameterized by a vector-like parameter. Updating the cross-sectional rotations
is performed purely through natural exponentiation and superposition of relevant rotational
quantities. This enables the proposed Timoshenko beam formulation to tackle beam
structures undergoing large displacements and rotations without any restriction in
magnitude.

In this thesis, a novel Euler-Bernoulli beam formulation is also proposed. In the


developed Euler-Bernoulli beam formulation, a geometric description that is similar to the
one employed in the Timoshenko beam formulation is utilized. However, kinematics are
described by the displacements of the beam axis and the axial cross-sectional rotation, not
the total one, showing some advantages which will be discussed later in the thesis. The
orthogonality between the cross-sections and the beam axis is satisfied by using the
smallest rotation mapping for the description of finite cross-sectional rotations. The use of
the smallest rotation mapping reduces the nonlinearity of the employed strain
measurements with respect to the unknown kinematics and offers linearization that is much
more efficient than those of existing isogeometric Euler-Bernoulli beam formulations.
Furthermore, a penalty-free approach is introduced to deal with rigid connections in multi-
patch beam structures in the context of geometrically nonlinear analysis. A novel nonlinear
transformation between the total cross-sectional rotation and the unknown kinematics is
derived, which facilitates the use of the total cross-sectional rotations at the ends of patches
as discrete unknowns. This approach also allows straightforward enforcement of rotational
boundary conditions.

Ref. code: 25636122300053IAQ


(3)

To show the accuracy and efficiency of the proposed beam formulations, some
benchmark and well-established numerical examples with various types of beams, i.e.,
straight, curved, pre-twisted beams, and lattice-like beam structures, are analyzed. The
obtained results are compared with those in the literature, obtained from both analytical
and numerical methods.

Keywords: Isogeometric analysis (IGA), Non-uniform rational B-spline


(NURBS), Geometrically nonlinear analysis, Timoshenko and Euler-Bernoulli beam
formulations, Multi-patch beam structures, Pre-twisted beams.

Ref. code: 25636122300053IAQ


(4)

ACKNOWLEDGEMENTS

“Every journey of thousand miles begins with a single small step”

My study life at Sirindhorn International Institute of Technology (SIIT),


Thammasat University finally comes to the end. Words could not be enough to express my
deepest appreciation to my advisor, Prof. Pruettha Nanakorn, for continuously supporting
me for the last five years, during my master’s and doctoral degree studies at SIIT. I also
would like to extend my deepest gratitude to my co-advisor, Assoc. Prof. Tinh Quoc Bui,
for giving me opportunities to work with his research group. As advisors, their guidance is
crucial for the accomplishment of this research. During this research, we together share
enjoyable stories, experiences, and lessons.

Furthermore, I would like express my sincere thanks to Prof. Jaroon Rungamornrat,


Assoc. Prof. Nakhorn Poovarodom, Research Asst. Prof. Pakawat Sancharoen, and Asst.
Prof. Ganchai Tanapornraweekit for serving as committee members. Their time and helpful
comments are extremely appreciated.

It would be a shortcoming to not recognize the assistance of colleagues in


Computational Engineering Lab at SIIT: Dr. Pana Suttakul, Dr. Alin Shakya, Mr. Duong
Huu Nghi, and Mr. Chorn Vithearin. I am fortunate to be a part of the team.

During this research, I spent eight months at Tokyo Institute of Technology, Tokyo,
Japan. I am very thankful to friends and colleagues in HIROSE & BUI Lab for making
memorable moments in Japan.

Scholarships from the ASEAN University Network/ Southeast Asia Engineering


Education Development Network (AUN/SEED-Net) and SIIT are greatly acknowledged.

Last but not least, special thanks are dedicated to my beloved family and my dear
girlfriend, Giang, for their patience and encouragement at all times.

Pathumthani, November 2020

Duy Vo

Ref. code: 25636122300053IAQ


(5)

TABLE OF CONTENTS

Page

ABSTRACT (1)

ACKNOWLEDGEMENTS (4)

TABLE OF CONTENTS (5)

LIST OF TABLES (9)

LIST OF FIGURES (10)

CHAPTER 1 INTRODUCTION 1

1.1 General 1

1.2 Statement of the problem 2

1.3 Objectives of the study 3

1.4 Scope of the study 3

CHAPTER 2 REVIEW OF LITERATURE 5

CHAPTER 3 RATIONAL B-SPLINE BASIS FUNCTIONS 9

3.1 B-splines 9

3.2 Rational B-splines 11

Ref. code: 25636122300053IAQ


(6)

CHAPTER 4 ISOGEOMETRIC TIMOSHENKO BEAM FORMULATION 13

4.1 Kinematic description of spatial Timoshenko beams 13

4.1.1 Beam axis and director vectors 13

4.1.2 Beam configurations 17

4.1.3 Curvilinear coordinate systems 18

4.1.4 Strain measurements and cross-sectional resultant forces 19

4.2 Variational formulations 23

4.2.1 Variations of unknown kinematics 23

4.2.2 Virtual work principle 25

4.3 NURBS-based discretization 26

4.3.1 Discretization of the kinematics 26

4.3.2 Beam axis parameterizations 28

4.3.3 Discretization of the strain measurements 28

4.3.4 Linearization of the system equations 30

4.3.5 Kinematic update procedure 34

4.4 Numerical examples and discussions 35

4.4.1 Cantilever quadrant subjected to an end moment 36

4.4.2 Pre-twisted circular arch 39

4.4.3 Rolled-up straight cantilever beam 42

4.4.4 Lateral buckling of a cantilever beam 47

4.4.5 Twisting of a circular ring 48

CHAPTER 5 ISOGEOMETRIC EULER-BERNOULLI BEAM FORMULATION 52

5.1 Kinematic description of spatial Euler-Bernoulli beams 52

Ref. code: 25636122300053IAQ


(7)

5.1.1 Smallest rotation (SR) mapping and director vectors of Euler-Bernoulli beams
52

5.1.2 Variations of kinematics 54

5.1.3 Virtual work principle 55

5.2 NURBS-based Euler-Bernoulli beams 57

5.2.1 NURBS discretization of the kinematics 57

5.2.2 Parameterizations of the beam axis 58

5.2.3 Nonlinear system equations 58

5.2.4 Linearization of the system equations 60

5.3 Multi-patch Euler-Bernoulli beams 61

5.3.1 Axial rotation and total cross-sectional rotations 62

5.3.2 Transformations of degrees of freedom 62

5.4 Numerical examples and discussions 64

5.4.1 Square frame subjected to opposite compression forces 65

5.4.2 Lee’s frame 68

5.4.3 A pre-twisted circular arch 70

5.4.4 Twisting of a circular ring 74

5.4.5 Compression of a lattice tower 77

CHAPTER 6 CONCLUSIONS 81

REFERENCES 83

APPENDICES 90

APPENDIX A 91

Ref. code: 25636122300053IAQ


(8)

APPENDIX B 93

APPENDIX C 95

APPENDIX D 98

BIOGRAPHY 100

Ref. code: 25636122300053IAQ


(9)

LIST OF TABLES

Tables Page

4.1 Rolled-up straight cantilever beam – Single circle test: comparison of the number of
degrees of freedom. 44

5.1 Twisting of a circular ring. Comparison of the number of degrees of freedom for a
half of the ring. 75

Ref. code: 25636122300053IAQ


(10)

LIST OF FIGURES

Figures Page

3.1 (a) Cubic B-spline basis functions with the knot vector 𝛷 =
0, 0, 0, 0, 0.25, 0.5, 0. ,75, 0.75, 1, 1, 1, 1. (b) An associated cubic B-spline curve with
knots and control points. 10

3.2 (a) NURBS and B-spline curves-the associated weights of the NURBS curve are
given in the parentheses, and the unit weights are dropped. (b) Associated ration B-spline
and B-spline basis functions. 12

4.1 Reference continuum beam configurations. 14

4.2 Beam axis and director vectors in the reference and current configurations. 16

4.3 Continuum beam configurations. 17

4.4 Cantilever quadrant subjected to an end moment. 36

4.5 Cantilever quadrant subjected to an end moment: (a) Strain energy obtained by
various NURBS curves of different degrees. (b) Contribution of the membrane and shear
strain energy. 37

4.6 Pre-twisted circular arch. 39

4.7 Pre-twisted circular arch – Convergence tests for the deformation of the arch axis. 40

4.8 Pre-twisted circular arch: (a) Displacement 𝑢𝑋. (b) Displacement 𝑢𝑌. (c)
Displacement 𝑢𝑍. (d) Reference and deformed configurations. 41

4.9 Rolled-up straight cantilever beam. 42

4.10 Rolled-up straight cantilever beam – Single circle test: (a) Convergence tests for the
position of the free end. (b) Normalized load-displacement curves of the free end. (c)
Reference and final deformed configurations. 43

4.11 Rolled-up straight cantilever beam – Double circle test: (a)Convergence tests of the
position of the free end. (b) Normalized load-displacement curves of the free end. (c)
Reference and deformed configurations. 45

Ref. code: 25636122300053IAQ


(11)

4.12 Lateral buckling of a cantilever beam. 46

4.13 Lateral buckling of a cantilever beam: (a) Convergence tests for the deformation of
the beam axis. (b) Load-displacement curves of the free end. (c) Reference and deformed
configurations. 47

4.14 Twisting of a circular ring. 48

4.15 Twisting of a circular ring – Load-displacement curves of the point 𝐵. 49

4.16 Twisting of a circular ring – Deformed configurations. 50

5.1 Director vectors in the current configuration of Euler-Bernoulli beams. 53

5.2 Square frame subjected to opposite compression forces: (a) Problem set up. (b)
Equivalent model. 65

5.3 Square frame: (a) Normalized load-displacement curves. (b) Deformed


configurations. 66

5.4 Lee’s frame: (a) Problem set up. (b) Normalized load-displacement curves of the
point 𝐷. (c) Deformed frame axes. 67

5.5 A pre-twisted circular arch – Problem set up. 68

5.6 A pre-twisted circular arch: (a) Top view. (b) NURBS representations of the arch and
control points. 69

5.7 A pre-twisted circular arch – A simulation in Abaqus. 71

5.8 A pre-twisted circular arch – Convergence test for the relative error in the
deformation of the arch axis. 71

5.9 A pre-twisted circular arch: (a) Displacement 𝑢𝑋. (b) Displacement 𝑢𝑌. (c)
Displacement 𝑢𝑍. (d) Reference and deformed configurations (with a scale factor of 1000
for displacements). 72

5.10 Twisting of a circular ring: (a) Problem set up. (b) Moment-rotation curve of the
point 𝐷. 74

5.11 Twisting of a circular ring – Deformed configurations. 76

Ref. code: 25636122300053IAQ


(12)

5.12 Compression of a lattice tower: (a) Problem set up. (b) Target surface and embedded
curves. 77

5.13 Compression of a lattice tower: (a) NURBS representation of the generatrix of the
target surface with the axis of revolution. (b) Parent domains of the embedded curves. 78

5.14 Compression of a lattice tower – Comparison of the displacement at the joints 𝐾 and
𝑁. 79

5.15 Compression of a lattice tower – Deformed configurations: (a) Beam formulation


developed by Vo, Nanakorn and Bui (2020). (b) The proposed formulation. 80

Ref. code: 25636122300053IAQ


1

CHAPTER 1

INTRODUCTION

1.1 General

Among various beam theories, both the Euler-Bernoulli and the Timoshenko beam
theories are popular. Both theories assume that the beam cross-sections are rigid. The major
difference between them lies in the orientations of beam cross-sections. While the former
theory assumes that rigid beam cross-sections are always perpendicular to the beam axis
during deformation, the latter theory releases this assumption by including the shear effect.
In practice, the Euler-Bernoulli beam theory is suitable for analysis of thin beams, whereas
the Timoshenko beam theory has a wider range of applications since its ability to deal with
both thin and thick beams. Nevertheless, the performance of finite element formulations
based on the Timoshenko beam theory suffers from the well-known shear locking
phenomenon when accounting for thin beams. Special treatments are thus required to
eliminate the shear locking issue.

Recently, isogeometric analysis (IGA) (Hughes, Cottrell & Bazilevs, 2005) has
become a powerful numerical approach in the field of computational mechanics. The
primary concept of IGA is to fill in the gap between geometric design and computational
analysis by using the same basis functions used in computer-aided design (CAD), e.g.,
rational B-spline basis functions, for both geometric descriptions of domains and
approximation of unknown fields. With this approach, exact representations of complex
reference geometry, generated by CAD, can be used directly in the analysis without any
modifications. In addition, the use of CAD basis functions with high order of continuity
for the approximation of unknown fields results in better accuracy per degree of freedom.
These advantages of IGA have been verified for analysis of structural members, i.e., plates
and shells (Dornisch & Klinkel, 2014; Dornisch, Klinkel & Simeon, 2013; Dornisch,
Müller & Klinkel, 2016; Kiendl, Bazilevs, Hsu, Wüchner & Bletzinger, 2010; Kiendl,
Bletzinger, Linhard & Wüchner, 2009), beams and rods (Fang, Yu, Van Lich & Bui, 2019;

Ref. code: 25636122300053IAQ


2

Vo & Nanakorn, 2020; Yu, Hu, Zhang & Bui, 2019; Yu, Zhang, Hu & Bui, 2019), and
other engineering applications (Lai et al., 2017; Yu, Lai & Bui, 2019).

For beams subjected to large displacements, the deformed configurations are


naturally complicated. In standard finite element analysis (FEA), the number of elements
required for a smooth and accurate geometric representation of a deformed configuration
can be greater than that required for the aimed accuracy of displacements at only some
specific locations. This kind of occurrence is even more pronounced in analysis of
structures that are sensitive to geometric imperfection. However, in IGA associated with
NURBS, complicated deformed configurations of beams can be captured better, and
consequently obtained results are more accurate with a smaller number of elements. This
is one of the major motivations to develop advance approaches using IGA for analysis of
beam structures.

1.2 Statement of the problem

In literature, for shear deformable beams, the geometrically exact beam theory
(Simo, 1985) is intensively considered. However, in the geometrically exact beam theory,
the deformation gradient tensor and first Piola-Kirchhoff stress tensor are used as the
energy conjugate pair in the derivation of finite element equations. This approach is not
convenient when considering complicated material constitutive laws, e.g. hyperelastic
materials or functionally graded materials, where the Green-Lagrange strain tensor is
widely used. In these cases, beam formulations using the Green-Lagrange strain tensor and
second Piola-Kirchhoff stress tensor in the derivation of finite element equations are
preferable.

As the use of Timoshenko beam formulations for analysis of thin beams suffers
from the shear locking effects, the development of Euler-Bernoulli beam formulations is
necessary to get rid of special treatments of the shear locking issue. In fact, the efficiency
and accuracy of Euler-Bernoulli beams with the IGA approach have been shown by many
previous studies. However, due to the lack of rotational degrees of freedom in Euler-
Bernoulli beam formulations with the IGA approach, analyzing multi-patch beams is still

Ref. code: 25636122300053IAQ


3

challenging. In this study, this limitation is handled by using the end rotations of patches
as degrees of freedom. Numerical results show the accuracy and efficiency of the
developed beam formulation in modelling beam structures consisting of multi-patches.

1.3 Objectives of the study

This study aims to develop effective beam formulations in association with the IGA
approach using kinematic assumptions of the Timoshenko beam theory and the Euler-
Bernoulli beam theory. The main objectives of this thesis are as follows:

- To construct a novel isogeometric Timoshenko beam formulation using the


variational approach where the Green-Lagrange strain tensor and the second Piola-
Kirchhoff stress tensor are used to facilitate consideration of complicated material
constitutive laws.

- To develop a novel isogeometric Euler-Bernoulli beam formulation using the


variational approach with the Green-Lagrange strain tensor and the second Piola-
Kirchhoff stress tensor, and to remove, from the proposed Euler-Bernoulli beam
formulation, the intrinsic limitation of isogeometric Euler-Bernoulli beam
formulations for analyzing multi-patch beams by considering the end cross-
sectional rotations of patches as degrees of freedom.

- To verify the accuracy and performance of developed beam formulations, some


benchmark and well-established problems whose exact solutions or numerical
results are available in the literature are analyzed.

1.4 Scope of the study

The scope of this study is as follows:

- Linear elasticity is assumed.

- The kinematic assumptions of the Timoshenko beam theory and the Euler-
Bernoulli beam theory are considered.

Ref. code: 25636122300053IAQ


4

- Conservative loads are considered.

Ref. code: 25636122300053IAQ


5

CHAPTER 2

REVIEW OF LITERATURE

Spatial flexible beam structures subjected to large displacements and rotations are
often encountered in a wide range of engineering applications, perhaps most notably in
space, marine, and offshore structures. In order to gain insight into mechanical behavior of
these structures, many spatial beam formulations have been proposed (Bathe & Bolourchi,
1979; Cardona & Geradin, 1988; Crisfield, 1990; Ibrahimbegovic, 1997; Ibrahimbegović,
1995; Ibrahimbegović, Frey & Kožar, 1995; Reissner, 1973; Simo, 1985; Simo & Vu-
Quoc, 1986). A thorough discussion on the historical development of spatial beam
formulations is beyond the scope of this work, and curious readers may refer to some
compact reviews, e.g., (Greco & Cuomo, 2015; Meier, Popp & Wall, 2019).

Recently, isogeometric analysis (IGA) (Hughes, Cottrell & Bazilevs, 2005) has
become a powerful numerical approach in the field of computational mechanics. The
primary concept of IGA is to fill in the gap between geometric design and computational
analysis by using the same basis functions used in computer-aided design (CAD), e.g.,
rational B-spline basis functions, for both geometric descriptions of domains and
approximation of unknown fields. With this approach, exact representations of complex
reference geometry, generated by CAD, can be used directly in analysis without any
modifications. In addition, the use of CAD basis functions with high order of continuity
for the approximation of unknown fields results in better accuracy per degree of freedom.
These advantages of IGA have been verified for analysis of structural members, i.e., plates
and shells (Dornisch & Klinkel, 2014; Dornisch, Klinkel & Simeon, 2013; Dornisch,
Müller & Klinkel, 2016; Kiendl, Bazilevs, Hsu, Wüchner & Bletzinger, 2010; Kiendl,
Bletzinger, Linhard & Wüchner, 2009), beams and rods (Fang, Yu, Van Lich & Bui, 2019;
Vo & Nanakorn, 2020; Yu, Hu, Zhang & Bui, 2019; Yu, Zhang, Hu & Bui, 2019), and
other engineering applications (Lai et al., 2017; Yu, Lai & Bui, 2019).

Regarding spatial Timoshenko beams, some formulations based on the


isogeometric collocation method (IGC) have been considered. Linear analysis of beams

Ref. code: 25636122300053IAQ


6

using IGC is performed by Auricchio, Beirão da Veiga, Kiendl, Lovadina and Reali (2013).
The governing equations are given in two forms, i.e., the displacement-based form and the
mixed one, where both displacements and forces are primary unknowns. The locking
effects in the displacement-based formulation can be avoided by using rational B-spline
basis functions of high degree, whereas the mixed formulation is completely locking-free
regardless of the degrees of the basis functions used. The IGC approach is applied to the
geometrically exact beam model (Simo, 1985) for geometrically nonlinear analysis of
beams by Weeger, Yeung and Dunn (2017), and shortly later also by Marino (2016, 2017).
Different parameterizations of cross-sectional rotations are used, namely quaternions in the
former work and vector-like parameters in the latter works. Later, the IGC approach with
the geometrically exact beam model has been further applied to several engineering
problems, i.e., frictionless contacts (Weeger, Narayanan, De Lorenzis, Kiendl & Dunn,
2017), shape optimization (Weeger, Narayanan & Dunn, 2019), and nonlinear dynamic
analysis of beam structures (Marino, Kiendl & De Lorenzis, 2019, 2019). Through all the
works discussed above, the IGC approach has been shown to be effective for nonlinear
analysis of beams where the numerical integration is avoided, and the need for the higher
continuity of displacements required by the governing equations is easily handled by the
smoothness of rational B-spline basis functions. Therefore, the IGC approach permits good
accuracy to be obtained with reduced computational costs. However, for problems with
complex mechanical phenomena, where the governing equations are not available, the IGC
approach cannot be used. In this case, the variational approach can be considered as an
effective alternative. One should also note that variational Timoshenko beam formulations
with the IGA approach have not been widely considered. Most recently, the variational
formulation for the geometrically exact beam model (Simo, 1985) is used with the IGA
approach by Choi and Cho (2019) for design sensitivity analysis of beam structures.

Recently, several isogeometric shell formulations have been proposed with the
energy conjugate pair of the Green-Lagrange strain tensor and the second Piola-Kirchhoff
stress tensor (Dornisch & Klinkel, 2014; Dornisch, Klinkel & Simeon, 2013; Dornisch,
Müller & Klinkel, 2016; Kiendl, Bletzinger, Linhard & Wüchner, 2009). This approach
facilitates analysis of shell structures made of hyperelastic materials (Chen et al., 2014;
Kiendl, Hsu, Wu & Reali, 2015; Tepole, Kabaria, Bletzinger & Kuhl, 2015), whose strain

Ref. code: 25636122300053IAQ


7

energy density functions are usually described in terms of the Green-Lagrange strain
tensor. In the conventional finite element approach, many beam formulations have been
developed using the Green-Lagrange strain tensor and the second Piola-Kirchhoff stress
tensor. However, there virtually exist no isogeometric Timoshenko beam formulations
derived with this energy conjugate pair of the strain and stress tensors.

Regarding implementation, beam formulations based on the Timoshenko beam


theory are preferable to the Euler-Bernoulli beam theory. In the Timoshenko beam theory,
the cross-sectional rotation is not a function of only the beam axis. Because of that, the
beam axis and the cross-sectional rotation can be separately approximated. When the Euler-
Bernoulli beam theory is employed, the orthogonality between each cross-section and the
beam axis allows the development of beam formulations in which the cross-sectional
rotation is not one of the unknowns. These beam formulations are referred to as rotation-
free beam formulations. Beams with high slenderness, whose shear deformations are
negligible, are common in practice and they can be modelled accurately by the Euler-
Bernoulli beam theory. In fact, if a Timoshenko beam formulation, in which the axis
displacements and cross-sectional rotation are separately approximated, is used to model a
slender beam, it is likely that the accuracy of the analysis suffers from the well-known
effect of shear locking. To lessen the shear locking effect in Timoshenko beam
formulations, special manipulations must be devised. On the contrary, analysis of highly
slender beams using rotation-free beam formulations based on the Euler-Bernoulli beam
theory does not suffer from the shear locking effect. Moreover, since the cross-sectional
rotation is not one of the unknowns in rotation-free beam formulations, there is no degree
of freedom related to the cross-sectional rotation. Consequently, these formulations require
less degrees of freedom to yield the same accuracy.

Only a few IGA Euler-Bernoulli beam formulations have been proposed (Bauer,
Breitenberger, Philipp, Wüchner & Bletzinger, 2016; Bauer, Wüchner & Bletzinger, 2019;
Greco & Cuomo, 2013, 2014; Radenković & Borković, 2018; Raknes et al., 2013). By
using only the displacements of the beam axis as unknown kinematics, nonlinear dynamic
analysis of beams is performed by Raknes et al. (2013). Since only displacements are used
as the unknowns, torsional behavior cannot be considered. For analysis of beams with
torsional effects and pre-twisted configurations, the axial rotation angle around the beam

Ref. code: 25636122300053IAQ


8

axis is included in the formulations developed by Greco and Cuomo (2013) and
Radenković and Borković (2018) for linear analysis, and by Bauer, Breitenberger, Philipp,
Wüchner and Bletzinger (2016) for geometrically nonlinear analysis. The formulation
developed by Greco and Cuomo (2013) is later extended by Greco and Cuomo (2014) for
assemblies of multi-patch beam structures by using the cross-sectional rotations at the ends
of patches as discrete unknowns. Although the formulations proposed by Greco and Cuomo
(2013, 2014) are derived with finite deformation kinematics, only examples in the linear
regime are shown. Geometrically nonlinear analysis of multi-patch beam structures is
performed by Bauer, Wüchner and Bletzinger (2019). Rigid connections between patches
is considered by using some additional constraints, and the penalty approach is used to
incorporate these constraints into the virtual work principle. In this approach, the accuracy
is affected by the choices of problem-dependent coefficients. Additionally, for beam
structures with many patches, the inclusion of these additional constraints is not practically
convenient.

Although very promising results have been reported (Bauer, Breitenberger, Philipp,
Wüchner & Bletzinger, 2016; Bauer, Wüchner & Bletzinger, 2019; Greco & Cuomo, 2013,
2014), it is noticed that strain measurements are highly nonlinear in terms of the unknown
kinematics since the Euler-Rodriguez formula is used for the description of finite cross-
sectional rotations. Consequently, the linearization of the system equations is
computationally expensive. In the context of geometrically nonlinear analysis, the
enforcement of rigid connections in multi-patch beams still requires an approach that is
more convenient for practical applications.

Ref. code: 25636122300053IAQ


9

CHAPTER 3

RATIONAL B-SPLINE BASIS FUNCTIONS

In IGA, rational B-spline basis functions are used both for the approximations of
the geometry and unknown variable fields. In this section, rational B-spline basis functions
and their properties are briefly introduced. More details about these functions can be found
in the textbook of Piegl and Tiller (1997).

3.1 B-splines

A B-spline curve 𝐂(𝜂) ∈ ℝ𝑑 is a piecewise parametric curve written as a linear


combination of B-spline basis functions, i.e.,

𝐂(𝜂) = ∑ 𝑁𝑖,𝑝 (𝜂)𝐏𝑖 (3.1)


𝑖=1

where 𝜂 is the parametric variable and 𝑁𝑖,𝑝 (𝜂) is a B-spline basis function of degree 𝑝. In
addition, 𝐏𝑖 ∈ ℝ𝑑 is a control point while 𝑛 is the number of the control points.

Define a knot vector 𝛷 as

𝛷 = [𝜂1 , 𝜂2 , … , 𝜂𝑛+𝑝+1 ] (3.2)

where 𝜂𝑖 ≤ 𝜂𝑖+1 are knot values. A knot vector is called uniform if its elements are equally
spaced. In addition, a knot vector is called open if the multiplicities of knot values at the
ends are both equal to 𝑝 + 1. Hereafter, knot vectors in this paper refer to nonuniform,
open knot vectors.

B-spline basis functions are defined by the Cox-de Boor relation as

1 if 𝜂 ∈ [𝜂𝑖 , 𝜂𝑖+1 )
𝑁𝑖,0 (𝜂) = { (3.3)
0 otherwise

Ref. code: 25636122300053IAQ


10

Figure 3.1 (a) Cubic B-spline basis functions with the knot vector 𝛷 =
[0, 0, 0, 0, 0.25, 0.5, 0. ,75, 0.75, 1, 1, 1, 1]. (b) An associated cubic B-spline curve with
knots and control points.

𝜂 − 𝜂𝑖 𝜂𝑖+𝑝+1 − 𝜂
𝑁𝑖,𝑝 (𝜂) = 𝑁𝑖,𝑝−1 (𝜂) + 𝑁 (𝜂).
𝜂𝑖+𝑝 − 𝜂𝑖 𝜂𝑖+𝑝+1 − 𝜂𝑖+1 𝑖+1,𝑝−1 (3.4)

B-spline basis functions have the property of the partition of unity, namely
∑𝑛𝑖=1 𝑁𝑖,𝑝 (𝜂) = 1. They are non-negative and locally supported on [𝜂1 , 𝜂𝑛+𝑝+1 ]. At any
point on [𝜂1 , 𝜂𝑛+𝑝+1 ], there are at most 𝑝 + 1 basis functions that are non-zero. In addition,
at an internal knot value, namely a knot value that is not the first or the last one, B-spline
basis functions are 𝐶 𝑝−𝑘 continuous, where 𝑘 is the multiplicity of the knot value and 1 ≤
𝑘 ≤ 𝑝 (Piegl & Tiller, 1997). The continuity of a B-spline curve follows the continuity of
its B-spline basis functions. In general, a B-spline curve does not pass through all its control
points. However, if an internal knot value of a B-spline curve has the multiplicity of 𝑝, the
B-spline curve interpolates the control point at that knot value. In addition, a B-spline curve

Ref. code: 25636122300053IAQ


11

having an open knot vector interpolates its first and last control points. A so-called control
polygon is obtained by connecting consecutive control points with linear functions.

As an example, cubic B-spline basis functions, i.e., 𝑝 = 3, and an associated planar


cubic B-spline curve are shown, respectively, in Figure 3.1a-b. In this study, conforming
to a common practice, the parametric variable 𝜂 varies from 0 to 1. In IGA, the curve
segment over a distinct knot span [𝜂𝑙 , 𝜂𝑙+1 ] is referred to as an element while the whole
curve is referred to as a single patch.

3.2 Rational B-splines

B-spline basis functions 𝑁𝑖,𝑝 (𝜂), defined by the Cox-de Boor relation, are in fact
polynomials, written in different forms. Consequently, B-spline curves are polynomial
curves and are not capable of exactly representing Conic sections, such as circles and
ellipses. Another class of curves, called non-uniform rational B-spline (NURBS) curves,
does not have this drawback. NURBS curves generally have non-uniform knot vectors and
are constructed from so-called rational B-spline basis functions. A NURBS curve is defined
by

𝑛
∑𝑛𝑖=1 𝑁𝑖,𝑝 (𝜂)𝑤𝑖 𝐏𝑖
𝐂(𝜂) = = ∑ 𝑅𝑖,𝑝 (𝜂)𝐏𝑖 . (3.5)
∑𝑛𝑖=1 𝑁𝑖,𝑝 (𝜂)𝑤𝑖
𝑖=1

Here, 𝑤𝑖 is a scalar weight coefficient or simply a weight, and 𝑅𝑖,𝑝 (𝜂) is a rational B-spline
basis function defined by

𝑁𝑖,𝑝 (𝜂)𝑤𝑖
𝑅𝑖,𝑝 (𝜂) = . (3.6)
∑𝑛𝑗=1 𝑁𝑗,𝑝 (𝜂)𝑤𝑗

In this study, all 𝑤𝑖 s are restricted to be positive. As a result, the denominators in Eqs. (3.5)
and (3.6) are always positive, and the division by zero is explicitly avoided. It should be
noted that, if all 𝑤𝑖 s in a rational B-spline basis function 𝑅𝑖,𝑝 (𝜂) are identical, the rational
B-spline basis function 𝑅𝑖,𝑝 (𝜂) simply becomes a B-spline basis function 𝑁𝑖,𝑝 (𝜂).
Therefore, B-spline curves are special cases of NURBS curves.

Ref. code: 25636122300053IAQ


12

Figure 3.2 (a) NURBS and B-spline curves-the associated weights of the NURBS
curve are given in the parentheses, and the unit weights are dropped. (b) Associated
ration B-spline and B-spline basis functions.

As an example comparison, a NURBS curve and a B-spline curve having the same
control points are shown in Figure 3.2a. The non-unit weights used in the NURBS curve
are given in the parentheses. The associated rational B-spline and B-spline basis functions
are shown in Figure 3.2b. It can be noticed that the NURBS curve exactly represents a
circle.

Ref. code: 25636122300053IAQ


13

CHAPTER 4

ISOGEOMETRIC TIMOSHENKO BEAM FORMULATION

4.1 Kinematic description of spatial Timoshenko beams

The kinematic description of a spatial beam based on the Timoshenko beam theory
is presented here. In this study, the unstressed configuration of the beam is treated as the
reference configuration. From now onwards, as a rule for notations, upper-case and lower-
case letters are used to denote the kinematic quantities in the reference and current
configurations, respectively. The convective coordinates of the beam are denoted by
𝜉 (𝑖) , 𝑖 = 1,2,3, where 𝜉 (1) is the convective coordinate along the beam axis. In addition,
𝜉 (2) and 𝜉 (3) denote the convective coordinates along the principal axes of the cross-
sections. The ranges of Roman and Greek indices are from 1 to 3 and 2 to 3, respectively,
unless otherwise stated. In this study, the arc-length of the reference beam axis is used as
𝜉 (1) . For convenience, 𝜉 (1) is interchangeably written as 𝜉.

4.1.1 Beam axis and director vectors

Any configuration of a spatial beam under the Timoshenko beam theory can be
represented by the beam axis and the cross-sections. Here, director vectors are the unit
vectors that are aligned with the principal axes of the cross-sections. The determination of
the director vectors is given below.

4.1.1.1 Reference beam axis

The reference beam axis is generally assumed to be a non-degenerate curve, which


means that it has no zero curvature at any point. Each cross-section of the beam is initially
planar and perpendicular to the reference beam axis. Example reference beam
configurations are shown in Figure 4.1.

Ref. code: 25636122300053IAQ


14

Figure 4.1 Reference continuum beam configurations.

Consider a cross-section C intersecting the reference beam axis at a point 𝑄0 in


Figure 4.2. The position of 𝑄0 is expressed in a global Cartesian coordinate system, defined
by the base vectors 𝐞𝑖 , as

𝐑 0 (𝜉) = 𝑋𝑖0 (𝜉)𝐞𝑖 . (4.1)

The unit tangent vector of the reference beam axis at 𝑄0 can be identified as

d𝐑 0 (𝜉)
𝐀1 = = 𝐑′0 (𝜉). (4.2)
d𝜉

Hereafter, the prime notations represent the derivatives with respect to 𝜉.

Ref. code: 25636122300053IAQ


15

For a non-degenerate reference beam axis, it is always possible to determine a set


of three vectors, i.e., the unit tangent vector 𝐀1 shown above, the unit normal vector 𝐍, and
the unit binormal vector 𝐁, expressed as

𝐀′1
𝐍= 𝐁 = 𝐀1 × 𝐍. (4.3)
‖𝐀′1 ‖

The three equations in Eqs. (4.2) and (4.3) are the so-called Frenet-Serret formulas. For
later use, the derivatives of 𝐍 and 𝐁 with respect to 𝜉 are given by

𝐀′′1 (𝐀′′1 ∙ 𝐀′1 )𝐀′1


𝐍′ = − 𝐁 ′ = 𝐀1 × 𝐍 ′ . (4.4)
‖𝐀′1 ‖ (‖𝐀′1 ‖)3

One can see that ‖𝐀′𝟏 ‖ = 0 for a straight beam. Consequently, 𝐍 and 𝐁 are not defined.
Later in subsection 4.4.3, an alternative method is introduced for the determination of 𝐍
and 𝐁 of a straight beam.

Denote the director vectors of the cross-section C by 𝐀 2 and 𝐀 3 . In some cases, 𝐀 2


and 𝐀 3 are identical with 𝐍 and 𝐁 (see Figure 4.1a). However, they can be generally
determined from 𝐍 and 𝐁 by rotating these vectors around 𝐀1 with an angle 𝜙 (see Figure
4.1b), i.e.,

𝐀2 cos 𝜙 sin 𝜙 𝐍
[ ]=[ ] [ ]. (4.5)
𝐀3 − sin 𝜙 cos 𝜙 𝐁

Their derivatives with respect to 𝜉 are given by

𝐀′2 0 1 𝐀2 ′ cos 𝜙 sin 𝜙 𝐍′


[ ′]=[ ][ ]𝜙 +[ ] [ ]. (4.6)
𝐀3 −1 0 𝐀 3 − sin 𝜙 cos 𝜙 𝐁 ′

4.1.1.2 Current beam axis

Due to the application of the external loads, the cross-section C moves to the cross-

section c in the current configuration, and the point 𝑄0 on the reference beam axis moves

Ref. code: 25636122300053IAQ


16

Figure 4.2 Beam axis and director vectors in the reference and current configurations.

to the point 𝑞0 on the current beam axis as shown in Figure 4.2. The position of 𝑞0 is given
by

𝐫0 (𝜉) = 𝐑 0 (𝜉) + 𝐮0 (𝜉) (4.7)

where 𝐮0 (𝜉) is the translational displacement vector of the beam axis. The tangent vector
of the current beam axis at 𝑞0 can be written as

d𝐫0 (𝜉)
𝐚1 = = 𝐫0′ (𝜉). (4.8)
d𝜉

In the Timoshenko beam theory, the cross-section c is still planar but generally no
longer perpendicular to the current beam axis. As a result, the director vectors 𝐚2 and 𝐚3
of the cross-section c cannot be determined from the current beam axis. In this study, they
are determined from 𝐀 2 and 𝐀 3 as

𝐚𝛼 = 𝚲𝐀 𝛼 (4.9)

where 𝚲 is an orthogonal or rotation tensor belongs to a special orthogonal group 𝑆𝑂(3)


defined as

Ref. code: 25636122300053IAQ


17

Figure 4.3 Continuum beam configurations.

𝑆𝑂(3) = {𝚲: ℝ3 → ℝ3 |𝚲𝚲T = 𝐈 ∧ det 𝚲 = 1}. (4.10)

Here, 𝐈 is an identity tensor or matrix of size 3 × 3. The rotation tensor 𝚲 is used to


describe finite rotations in a three-dimensional space. In this study, 𝚲 is parameterized by
a vector-like parameter by following Rodrigues’ rotation formula (see Appendix A). From
Eq. (4.9), it follows that

𝐚′𝛼 = 𝚲′ 𝐀 𝛼 + 𝚲𝐀′𝛼 . (4.11)

4.1.2 Beam configurations

Let 𝑄 be an arbitrary material point on C as shown in Figure 4.3. The position vector
of 𝑄 can be determined as

Ref. code: 25636122300053IAQ


18

𝐑(𝜉 (1) , 𝜉 (2) , 𝜉 (3) ) = 𝐑 0 (𝜉) + 𝜉 (𝛼) 𝐀 𝛼 . (4.12)

The point 𝑄 moves to the point 𝑞 on c in the current configuration. The position
vector of 𝑞 is expressed as

𝐫(𝜉 (1) , 𝜉 (2) , 𝜉 (3) ) = 𝐫0 (𝜉) + 𝜉 (𝛼) 𝐚𝛼 . (4.13)

It can be seen from Eqs. (4.12) and (4.13) that the position of any point of the beam
can be determined by the beam axis and the director vectors of the cross-section containing
the point. However, the beam axis 𝐫0 (𝜉) and the director vectors 𝐚𝛼 of the cross-section in
the current configuration are unknown. Note that 𝐚𝛼 is related to 𝐀 𝛼 through the rotation
tensor 𝚲 as shown in Eq. (4.9). For the proposed Timoshenko beam formulation, the current
beam axis 𝐫0 (𝜉) and the rotation tensor 𝚲(𝜉) are considered as the unknown kinematics.

4.1.3 Curvilinear coordinate systems

Covariant base vectors are defined, respectively, for a point in the reference and
current configurations as

𝜕𝐑(𝜉 (1) , 𝜉 (2) , 𝜉 (3) ) 𝜕𝐫(𝜉 (1) , 𝜉 (2) , 𝜉 (3) )


𝐆𝑖 = 𝐠𝑖 = . (4.14)
𝜕𝜉 (𝑖) 𝜕𝜉 (𝑖)

Substituting Eqs. (4.12) and (4.13) into Eq. (4.14) yields

𝐆1 = 𝐀1 + 𝜉 (𝛼) 𝐀′𝛼 𝐠1 = 𝐚1 + 𝜉 (𝛼) 𝐚′𝛼 (4.15)

𝐆𝛼 = 𝐀 𝛼 𝐠 𝛼 = 𝐚𝛼 . (4.16)

The reciprocal contravariant vectors 𝐆𝑖 are obtained from the following relation as

1 𝑖=𝑗
𝐆𝑖 ∙ 𝐆𝑗 = { (4.17)
0 𝑖 ≠ 𝑗.

A metric tensor 𝐆 is defined in the contravariant curvilinear coordinate system 𝐆𝑖 ⊗ 𝐆 𝑗 as

Ref. code: 25636122300053IAQ


19

𝐆 = 𝐺𝑖𝑗 𝐆𝑖 ⊗ 𝐆 𝑗 (4.18)

where the coefficients 𝐺𝑖𝑗 are the covariant metric components in the reference
configuration, defined as

𝐺𝑖𝑗 = 𝐆𝑖 ∙ 𝐆𝑗 . (4.19)

Similarly, the covariant metric components in the current configuration are defined as

𝑔𝑖𝑗 = 𝐠 𝑖 ∙ 𝐠 𝑗 . (4.20)

4.1.4 Strain measurements and cross-sectional resultant forces

4.1.4.1 Strain measurements

The deformation gradient tensor 𝐅 is given by

𝐅 = 𝐠 𝑖 ⊗ 𝐆𝑖 . (4.21)

The Green-Lagrange strain tensor 𝐄 can thus be expressed as

1 T 1 1
𝐄= (𝐅 𝐅 − 𝐆) = (𝐠 𝑖 ∙ 𝐠 𝑗 − 𝐆𝑖 ∙ 𝐆𝑗 )𝐆𝑖 ⊗ 𝐆 𝑗 = (𝑔𝑖𝑗 − 𝐺𝑖𝑗 )𝐆𝑖 ⊗ 𝐆 𝑗 . (4.22)
2 2 2

The components of the Green-Lagrange strain tensor 𝐄 in the contravariant curvilinear


coordinate system 𝐆𝑖 ⊗ 𝐆 𝑗 are expressed as

1
𝐸𝑖𝑗 = (𝑔 − 𝐺𝑖𝑗 ). (4.23)
2 𝑖𝑗

To obtain 𝐸𝑖𝑗 , 𝐺𝑖𝑗 and 𝑔𝑖𝑗 are determined, with the help of Eqs. (4.15) and (4.16) as

𝐺11 = 𝐀1 ∙ 𝐀1 + 2𝜉 (2) 𝐀1 ⋅ 𝐀′2 𝑔11 = 𝐚1 ∙ 𝐚1 + 2𝜉 (2) 𝐚1 ⋅ 𝐚′2


(4.24)
+ 2𝜉 (3) 𝐀1 ⋅ 𝐀′3 + 2𝜉 (3) 𝐚1 ⋅ 𝐚′3

Ref. code: 25636122300053IAQ


20

𝐺12 = 𝐀1 ∙ 𝐀 2 + 𝜉 (3) 𝐀 2 ⋅ 𝐀′3 𝑔12 = 𝐚1 ∙ 𝐚2 + 𝜉 (3) 𝐚2 ⋅ 𝐚′3 (4.25)

𝐺13 = 𝐀1 ∙ 𝐀 3 + 𝜉 (2) 𝐀 3 ⋅ 𝐀′2 𝑔13 = 𝐚1 ∙ 𝐚3 + 𝜉 (2) 𝐚3 ⋅ 𝐚′2 (4.26)

1 if 𝛼 = 𝛽 1 if 𝛼 = 𝛽
𝐺𝛼𝛽 = { 𝑔𝛼𝛽 = { (4.27)
0 if 𝛼 ≠ 𝛽 0 if 𝛼 ≠ 𝛽.

2
In the above expressions of 𝐺𝑖𝑗 and 𝑔𝑖𝑗 , all the quadratic terms, i.e., 𝛰 [(𝜉 (2) ) ],
2
𝛰 [(𝜉 (3) ) ], and 𝛰(𝜉 (2) 𝜉 (3) ), are ignored. By making use of Eqs. (4.24)-(4.27), the non-

zero components of the Green-Lagrange strain tensor 𝐄 are calculated as

𝛤11
12 𝐾 13 𝐾

1 (2) ⏞ (3) ⏞
𝐸11 = (𝐚1 ∙ 𝐚1 − 𝐀1 ∙ 𝐀1 ) + 𝜉 (𝐚1 ∙ 𝐚2 − 𝐀1 ∙ 𝐀 2 ) + 𝜉 (𝐚1 ∙ 𝐚3 − 𝐀1 ∙ 𝐀′3 )
′ ′ ′
2

(4.28)
𝛤12 𝐾23
1 ⏞ (3) ⏞
𝐸12 = [(𝐚1 ∙ 𝐚2 − 𝐀1 ∙ 𝐀 2 ) + 𝜉 (𝐚2 ∙ 𝐚′3 − 𝐀 2 ∙ 𝐀′3 )] (4.29)
2

𝛤13 𝐾32
1 ⏞ (2) ⏞
𝐸13 = [(𝐚1 ∙ 𝐚3 − 𝐀1 ∙ 𝐀 3 ) + 𝜉 (𝐚3 ∙ 𝐚′2 − 𝐀 3 ∙ 𝐀′2 )] (4.30)
2

or in compact expressions as

𝐸11 = 𝛤11 + 𝜉 (2) 𝐾12 + 𝜉 (3) 𝐾13 (4.31)

1 1
𝐸12 = (𝛤12 + 𝜉 (3) 𝐾23) 𝐸13 = (𝛤13 + 𝜉 (2) 𝐾32). (4.32)
2 2

The strain component 𝐸11 consists of one constant term and two linear terms with
respect to 𝜉 (2) and 𝜉 (3) , which are the convective coordinates along the principal axes of
the cross-sections. The constant term describes the membrane action while the linear terms
represent the bending actions. The term 𝛤11 in Eq. (4.31) is referred to as the membrane
strain, while the terms 𝜉 (2) 𝐾12 and 𝜉 (3) 𝐾13 are the bending strains. In fact, the terms 𝐾12
and 𝐾13 are terms that are similar to the bending curvatures in the directions of the principal

Ref. code: 25636122300053IAQ


21

axes of the cross-sections in the Euler-Bernoulli beam theory. However, because of the
cross-sectional rotations due to the shear deformation in the Timoshenko beam theory, 𝐾12
and 𝐾13 are not exactly curvatures.

Each of the other two strain components, 𝐸12 and 𝐸13 , has one constant term and
one linear term. The constant terms describe the shear actions whereas the linear terms
represent the torsional actions. The terms 𝛤12 /2 and 𝛤13 /2 represent the direct shear strains,
while 𝜉 (3) 𝐾23 /2 and 𝜉 (2) 𝐾32 /2 are the torsional shear strains. The terms 𝐾23 and 𝐾32 are
in fact the torsional curvatures and are related through the following relations, i.e.,

(𝐀 2 ∙ 𝐀 3 )′ = 𝐀 2 ∙ 𝐀′3 + 𝐀′2 ∙ 𝐀 3 = 0 (𝐚2 ∙ 𝐚3 )′ = 𝐚2 ∙ 𝐚′3 + 𝐚′2 ∙ 𝐚3 = 0. (4.33)

It follows from Eqs. (4.29), (4.30) and (4.33) that

𝐾23 = −𝐾32 . (4.34)

In other words, the two torsional curvatures have the same absolute value.

4.1.4.2 Constitutive relations and cross-sectional resultant forces

The second Piola-Kirchhoff stress tensor 𝐒 is the energy conjugate of the Green-
Lagrange strain tensor 𝐄. The second Piola-Kirchhoff stress tensor 𝐒 is expressed as

𝐒 = 𝑆 𝑖𝑗 𝐆𝑖 ⊗ 𝐆𝑗 . (4.35)

As the Green-Lagrange strain tensor 𝐄 has only three non-zero components, it follows that

𝑆 11 = 𝐸̅ 𝐸11 𝑆 12 = 2𝐺̅ 𝐸12 𝑆 13 = 2𝐺̅ 𝐸13 (4.36)

where 𝐸̅ , and 𝐺̅ are Young’s modulus and the shear modulus measured in the curvilinear
coordinate system 𝐆𝑖 ⊗ 𝐆𝑗 (see Appendix B).

The cross-sectional resultant forces are defined in this study as follows:

- Normal force:

Ref. code: 25636122300053IAQ


22

𝑁11 = ∫ 𝑆 11 d𝐴 = 𝐸̅ 𝐴𝛤11 (4.37)


𝐴

where 𝐴 is the cross-sectional area.

- Shear forces:

𝑁12 = ∫ 𝑆 12 d𝐴 = 𝐺̅ 𝐴𝛤12 (4.38)


𝐴

𝑁13 = ∫ 𝑆 13 d𝐴 = 𝐺̅ 𝐴𝛤13 . (4.39)


𝐴

- Bending moments:

𝑀12 = ∫ 𝜉 (2) 𝑆 11 d𝐴 = 𝐸̅ 𝐼2 𝐾12 + 𝐸̅ 𝐼23 𝐾13 (4.40)


𝐴

𝑀13 = ∫ 𝜉 (3) 𝑆 11 d𝐴 = 𝐸̅ 𝐼23 𝐾12 + 𝐸̅ 𝐼3 𝐾13 (4.41)


𝐴

where the cross-sectional moments of inertia 𝐼2 , 𝐼3 , and 𝐼23 are defined as

2 2
𝐼2 = ∫ (𝜉 (2) ) d𝐴 𝐼3 = ∫ (𝜉 (3) ) d𝐴 𝐼23 = ∫ 𝜉 (2) 𝜉 (3) d𝐴. (4.42)
𝐴 𝐴 𝐴

- Torsional moment

𝑀23 = ∫ (𝜉 (3) 𝑆 12 − 𝜉 (2) 𝑆 13 ) d𝐴 = 𝐺̅ 𝐽𝐾23 (4.43)


𝐴

where the polar moment of inertia 𝐽 is defined as

Ref. code: 25636122300053IAQ


23

2 2
𝐽 = ∫ [(𝜉 (2) ) + (𝜉 (3) ) ] d𝐴. (4.44)
𝐴

4.2 Variational formulations

4.2.1 Variations of unknown kinematics

Recall that the beam axis 𝐫0 (𝜉) and the rotation tensor 𝚲(𝜉) are considered as the
unknown kinematics in this study. The beam axis 𝐫0 (𝜉) is a member of the linear Euclidean
space while the rotation tensor 𝚲(𝜉) belongs to the orthogonal group 𝑆𝑂(3).
Mathematically, a beam configuration is a member of a nonlinear manifold Ω, defined as

Ω = {(𝐫0 (𝜉), 𝚲(𝜉))|𝐫0 (𝜉) ∈ ℝ3 , 𝚲(𝜉) ∈ 𝑆𝑂(3)}. (4.45)

Let 𝚲1 and 𝚲2 be two distinct rotation tensors. The summation 𝚲1 + 𝚲2 does not always
result in an orthogonal tensor. Hence, Ω is not a linear space but a nonlinear manifold. Due
to the mathematical complexity intrinsic to the space of beam configurations, the
admissible variations of the kinematics used in the derivation of the system equations are
explicitly defined.

As derived by Argyris (1982) and as shown in Appendix A, the rotation tensor 𝚲 is


̃, i.e.,
the exponential map of a skew-symmetric tensor 𝛉

̃).
𝚲 = exp(𝛉 (4.46)

In practice, it is more convenient to work with the axial vector of a skew-symmetric tensor
̃ is defined as
instead of the tensor itself. The axial vector 𝛉 of 𝛉

̃𝐚 = 𝛉 × 𝐚,
𝛉 ∀𝐚 ∈ ℝ3 . (4.47)

̃ is used to denote a skew-symmetric tensor whose axial vector


Hereafter, the notation (∗)
is (∗).

Ref. code: 25636122300053IAQ


24

By considering a virtual displacement vector 𝛿𝐮0 (𝜉) and a virtual axial vector
𝛿𝐰(𝜉), a varied beam axis 𝐫0𝜀 (𝜉) and a varied rotation tensor 𝚲𝜀 (𝜉) are defined as

𝐫0𝜀 (𝜉) = 𝐫0 (𝜉) + 𝛿𝐮0 (𝜉) = 𝐫0 (𝜉) + 𝜀𝛈(𝜉) (4.48)

𝚲𝜀 (𝜉) = exp(𝛉 ̃ (𝜉)) = exp(𝛉


̃(𝜉) + 𝛿𝐰 ̃(𝜉) + 𝜀𝐪
̃(𝜉)). (4.49)

Since 𝐫0𝜀 (𝜉) ∈ ℝ3 and 𝚲𝜀 (𝜉) ∈ 𝑆𝑂(3), it follows that a varied beam configuration
(𝐫0𝜀 (𝜉), 𝚲𝜀 (𝜉)) is always a member of Ω, i.e., (𝐫0𝜀 (𝜉), 𝚲𝜀 (𝜉)) ∈ Ω. Therefore, the varied
beam configuration (𝐫0𝜀 (𝜉), 𝚲𝜀 (𝜉)) is admissible. Straightforwardly, the admissible
variation of the unknown beam axis 𝐫0 (𝜉) is as explicitly given in Eq. (4.48), i.e.,

𝛿𝐫0 (𝜉) = 𝛿𝐮0 (𝜉) = 𝜀𝛈(𝜉). (4.50)

The admissible variation of the unknown rotation tensor is computed as

d𝚲𝜀
𝛿𝚲(𝜉) = 𝜀 | = 𝜀𝐪 ̃ (𝜉)𝚲(𝜉).
̃(𝜉)𝚲(𝜉) = 𝛿𝐰 (4.51)
d𝜀 𝜀=0

̃ (𝜉), by omitting the higher order


To better understand the physical meaning of 𝛿𝐰
variations, the varied rotation tensor 𝚲𝜀 (𝜉) can be written as

̃ 𝚲 = (𝐈 + 𝛿𝐰
𝚲𝜀 = 𝚲 + 𝛿𝚲 = 𝚲 + 𝛿𝐰 ̃ )𝚲. (4.52)

It can be seen from Eqs. (4.48) and (4.52) that the vector 𝛿𝐮0 (𝜉) is an infinitesimal virtual
displacement vector superposed onto the current beam axis, and the skew-symmetric tensor
̃ (𝜉) adds an infinitesimal virtual rotation to the current cross-sectional rotation.
𝛿𝐰
However, as aforementioned, it is more convenient to work with the virtual axial vector
̃ (𝜉). Thus, the vectors 𝛿𝐮0 (𝜉) and 𝛿𝐰(𝜉) are considered as the admissible
𝛿𝐰(𝜉) than 𝛿𝐰
variations of the unknown kinematics. In fact, 𝛿𝐰(𝜉) can be interpreted as an infinitesimal
virtual rotation superposed onto the current cross-sectional rotation (Simo & Vu-Quoc,
1986).

Once the variation of 𝚲 has been clearly defined, the variations of the director
vectors 𝐚2 and 𝐚3 are determined as

Ref. code: 25636122300053IAQ


25

𝛿𝐚𝛼 = 𝛿𝚲𝐀 𝛼 = 𝛿𝚲𝚲T 𝐚𝛼 . (4.53)

By using the expression of 𝛿𝚲 in Eq. (4.51), 𝛿𝐚𝛼 it follows that

𝛿𝐚𝛼 = 𝛿𝐰 × 𝐚𝛼 𝛿𝐚′𝛼 = 𝛿𝐰 ′ × 𝐚𝛼 + 𝛿𝐰 × 𝐚′𝛼 . (4.54)

4.2.2 Virtual work principle

The weak form of the equilibrium is expressed using the virtual work principle as

𝛿Π = 𝛿Π𝑖𝑛𝑡 − 𝛿Π𝑒𝑥𝑡 = 0 (4.55)

where 𝛿Π is the total virtual work. In addition, 𝛿Π𝑖𝑛𝑡 is the internal virtual work, and 𝛿Π𝑒𝑥𝑡
is the external virtual work.

In this study, the internal virtual work 𝛿Π𝑖𝑛𝑡 is expressed by means of the Green-
Lagrange strain tensor 𝐄 and the second Piola-Kirchhoff stress tensor 𝐒 as

𝛿Π𝑖𝑛𝑡 = ∫ 𝛿𝐄: 𝐒 d𝑉 = ∫ (𝛿𝐸11 𝑆 11 + 2𝛿𝐸12 𝑆 12 + 2𝛿𝐸13 𝑆 13 ) d𝑉 (4.56)


𝑉 𝑉

where 𝑉 is the initial volume of the beam. By using the expressions of the strain
components in Eqs. (4.31) and (4.32) and the definitions of the cross-sectional resultant
forces in Eqs. (4.37)-(4.44), the internal virtual work 𝛿Π𝑖𝑛𝑡 can be alternatively expressed
as

𝛿Π𝑖𝑛𝑡 = ∫(𝛿𝛤11 𝑁 11 + 𝛿𝛤12 𝑁12 + 𝛿𝛤13 𝑁13 + 𝛿𝐾12 𝑀12 + 𝛿𝐾13 𝑀13 + 𝛿𝐾23 𝑀23 ) d𝜉
0

(4.57)

where 𝐿 is the length of the reference beam axis. The following generalized strain and
stress vectors are introduced as

Ref. code: 25636122300053IAQ


26

𝚪 = [𝛤11 𝛤12 𝛤13 𝐾12 𝐾13 𝐾23 ]T (4.58)

𝐌 = [𝑁11 𝑁12 𝑁13 𝑀12 𝑀13 𝑀23 ]T . (4.59)

The generalized strain vector 𝚪 and generalized stress vector 𝐌 are related as follows:

𝐸̅ 𝐴 0 0 0 0 0
0 𝐺̅ 𝐴 0 0 0 0
0 0 𝐺̅ 𝐴 0 0 0
𝐌= 0 0 0 𝐸̅ 𝐼2 𝐸̅ 𝐼23 0 𝚪 = 𝐃𝚪. (4.60)
0 0 0 𝐸̅ 𝐼23 𝐸̅ 𝐼3 0
[0 0 0 0 0 𝐺̅ 𝐽]

Alternatively, the internal virtual work 𝛿Π𝑖𝑛𝑡 is compactly expressed as

𝛿Π𝑖𝑛𝑡 = ∫ 𝛿𝚪 T 𝐌 d𝜉. (4.61)


0

Hereafter, all the equations are expressed in matrix form for computational convenience.

The external virtual work 𝛿Π𝑒𝑥𝑡 can be expressed as

𝛿Π𝑒𝑥𝑡 = ∫(𝛿𝐮T0 𝐅 + 𝛿𝐰 T 𝐐) d𝜉 + 𝛿𝐮T0 (0)𝐟(0) + 𝛿𝐮T0 (𝐿)𝐟(𝐿) + 𝛿𝐰 T (0)𝐪(0) + 𝛿𝐰 T (𝐿)𝐪(𝐿)


0

(4.62)

where 𝐅 and 𝐐 are the distributed force and distributed moment vectors applied on the
beam axis. Furthermore, 𝐟(0), 𝐟(𝐿), 𝐪(0), and 𝐪(𝐿) are the concentrated force and
concentrated moment vectors applied at 𝜉 = 0 and 𝜉 = 𝐿, respectively.

4.3 NURBS-based discretization

4.3.1 Discretization of the kinematics

In this study, rational B-spline basis functions are used for both the geometric
approximation and the discretization of the unknown kinematics. A very brief introduction

Ref. code: 25636122300053IAQ


27

to rational B-spline basis functions and their properties has been presented in Chapter 3.
More details can be found in the textbook of Piegl and Tiller (1997).

The reference and current beam axes are represented as NURBS curves as

𝑛 𝑛

𝐑 0 (𝜂) = ∑ 𝑅𝑖 (𝜂) 𝐑 𝑖 𝐫0 (𝜂) = ∑ 𝑅𝑖 (𝜂) 𝐫𝑖 (4.63)


𝑖=1 𝑖=1

One should recall that 𝑅𝑖,𝑝 (𝜂) is a rational B-spline basis function of degree 𝑝, 𝜂 is the
parameter referred to as the knot parameter, and 𝑛 is the number of basis functions. In
addition, 𝐑 𝑖 and 𝐫𝑖 are the position vectors of the 𝑖th control point in the reference and
current configurations. Hereafter, the degree 𝑝 of basis functions is omitted for notational
simplicity. The displacement vector of the beam axis is then obtained as

𝑛 𝑛

𝐮0 (𝜂) = 𝐫0 (𝜂) − 𝐑 0 (𝜂) = ∑ 𝑅𝑖 (𝜂)(𝐫𝑖 − 𝐑 𝑖 ) = ∑ 𝑅𝑖 (𝜂)𝐩𝑖 (4.64)


𝑖=1 𝑖=1

where 𝐩𝑖 is the displacement vector of the 𝑖-th control point.

The tangent vectors of the beam axis are determined as

𝑛 𝑛

𝐀1 = ∑ 𝑅𝑖′ (𝜂)𝐑 𝑖 𝐚1 = ∑ 𝑅𝑖′ (𝜂)𝐫𝑖 . (4.65)


𝑖=1 𝑖=1

The variation of the displacement 𝛿𝐮0 (𝜉) and the virtual axial vector 𝛿𝐰 are
approximated as

𝑛 𝑛

𝛿𝐮0 (𝜂) = ∑ 𝑅𝑖 (𝜂)𝛿𝐩𝑖 𝛿𝐰 = ∑ 𝑅𝑖 (𝜂)𝛿𝐰𝑖 . (4.66)


𝑖=1 𝑖=1

The variations of the tangent vector 𝐚1 and the director vectors 𝐚2 and 𝐚3 are
expressed, respectively, as

Ref. code: 25636122300053IAQ


28

𝑛 𝑛

𝛿𝐚1 = ∑ 𝑅𝑖′ (𝜂)𝛿𝐫𝑖 = ∑ 𝑅𝑖′ (𝜂)𝛿𝐩𝑖 (4.67)


𝑖=1 𝑖=1

𝑛 𝑛
𝛿𝐚𝛼 = − ∑ 𝑅𝑖 (𝜂)𝐚̃𝛿𝐰
𝛼 𝑖 𝛿𝐚′𝛼 = − ∑[𝑅𝑖′ (𝜂)𝐚̃𝛿𝐰
𝛼
̃′
𝑖 + 𝑅𝑖 (𝜂)𝐚𝛼 𝛿𝐰𝑖 ].
𝑖=1 𝑖=1

(4.68)

4.3.2 Beam axis parameterizations

In this study, two different parameterizations of the beam axis, i.e., arc-length
parameterization 𝜉 and the NURBS parameterization 𝜂, are employed. Here, the relation
between the two parameterizations is established.

The arc-length in the reference configuration can be determined as

d𝐑 0 (𝜂)
d𝜉 = ‖ ‖ d𝜂 = ‖𝐑̇ 0 (𝜂)‖d𝜂 = 𝐼d𝜂. (4.69)
d𝜂

Hereafter, the dot notations indicate the derivatives with respect to 𝜂. It follows that

d𝑅𝑖 (𝜂) d𝜂 1
𝑅𝑖′ (𝜂) = = 𝑅̇ (𝜂). (4.70)
d𝜂 d𝜉 𝐼 𝑖

The determination of 𝑅̇𝑖 (𝜂) is available in the textbook of Piegl and Tiller (1997).

4.3.3 Discretization of the strain measurements

The variation of the generalized strain vector 𝚪 in Eq. (4.58) is determined as

Ref. code: 25636122300053IAQ


29

𝛿𝛤11 𝐚1T 𝛿𝐚1


𝛿𝛤12 𝐚T𝟐 𝛿𝐚1 + 𝐚1T 𝛿𝐚2
𝛿𝛤13 𝐚T𝟑 𝛿𝐚1 + 𝐚1T 𝛿𝐚3
𝛿𝚪 = = . (4.71)
𝛿𝐾12 (𝐚′2 )T 𝛿𝐚1 + 𝐚1T 𝛿𝐚′2
𝛿𝐾13 (𝐚′3 )T 𝛿𝐚1 + 𝐚1T 𝛿𝐚′3
[𝛿𝐾23 ]
[(𝐚′3 )T 𝛿𝐚2 + 𝐚T2 𝛿𝐚′3 ]

To express 𝛿𝚪 in a compact form, the following matrices are introduced, i.e.,

𝑅𝑖′ (𝜂)𝐚1T 𝟎T
𝑅𝑖′ (𝜂)𝐚T2 −𝑅𝑖 (𝜂)𝐚1T 𝐚̃2
𝑅𝑖′ (𝜂)𝐚T3 T
−𝑅𝑖 (𝜂)𝐚1 𝐚 ̃3
𝐇𝑖 = 𝐇 = [𝐇1 𝐇2 … 𝐇𝑛 ] (4.72)
𝑅𝑖′ (𝜂)(𝐚′2 )T ′
−𝑅𝑖 (𝜂)𝐚1T 𝐚 ̃′
̃2 − 𝑅𝑖 (𝜂)𝐚1T 𝐚2
𝑅𝑖′ (𝜂)(𝐚′3 )T ′ T
−𝑅 (𝜂)𝐚1 𝐚 T ̃
̃3 − 𝑅𝑖 (𝜂)𝐚1 𝐚′
𝑖 3
[ 𝟎T −𝑅𝑖′ (𝜂)𝐚T2 𝐚
̃3 ]

𝛿𝐩𝑖
𝛿𝐯𝑖 = [ ] 𝛿𝐯 = [𝛿𝐯1T 𝛿𝐯2T … 𝛿𝐯𝑛T ]T . (4.73)
𝛿𝐰𝑖

It follows that

𝛿𝚪 = ∑ 𝐇𝑖 𝛿𝐯𝑖 = 𝐇𝛿𝐯. (4.74)


𝑖=1

By using Eq. (4.74) in Eq. (4.61), the internal virtual work 𝛿Π𝑖𝑛𝑡 is rewritten as

𝛿Π𝑖𝑛𝑡 = 𝛿𝐯 ∫ 𝐇 T 𝐌 d𝜉.
T
(4.75)
0

To obtain the external virtual work 𝛿Π𝑒𝑥𝑡 in a compact form, the distributed load
vector 𝐝, point load vector 𝐏, and the matrix 𝐋 are defined as

𝐝 = [𝐅 T 𝐐T ]T (4.76)

𝐏 = [𝐟 T (0) 𝐪T (0) … 𝐟 T (𝐿) 𝐪T (𝐿)]T (4.77)

Ref. code: 25636122300053IAQ


30

𝐋 = [𝑅1 (𝜂)𝐈6 … 𝑅𝑛 (𝜂)𝐈6 ] (4.78)

where 𝐈6 is an identity matrix of size 6 × 6. The external virtual work 𝛿Π𝑒𝑥𝑡 now can be
expressed compactly as

𝛿Π𝑒𝑥𝑡 = 𝛿𝐯 ∫ 𝐋T 𝐝 d𝜉 + 𝛿𝐯 T 𝐏.
T
(4.79)
0

By substituting Eqs. (4.75) and (4.79) into Eq. (4.55), the system equations are
obtained as

𝐿 𝐿

∫ 𝐇 T 𝐌 d𝜉 = ∫ 𝐋T 𝐝 d𝜉 + 𝐏. (4.80)
0 0

4.3.4 Linearization of the system equations

The system equations in Eq. (4.80) are nonlinear in terms of the unknown
kinematics. As a result, the linearization of the system equations is required in order that
the Newton-Raphson algorithm can be used to solve them. To linearize the system
equations, first the virtual work equation in Eq. (4.55) is linearized. After that, the
linearized system equations are obtained by discretizing the linearized virtual work
equation.

The incremental forms of Eq. (4.66) are written as

𝑛 𝑛

Δ𝐮0 (𝜂) = ∑ 𝑅𝑖 (𝜂)Δ𝐩𝑖 Δ𝐰 = ∑ 𝑅𝑖 (𝜂)Δ𝐰𝑖 . (4.81)


𝑖=1 𝑖=1

The vectors Δ𝐩𝑖 s and Δ𝐰𝑖 s are considered as the degrees of freedom. They are collected in
a vector Δ𝐯 as

Δ𝐩𝑖
Δ𝐯𝑖 = [ ] Δ𝐯 = [Δ𝐯1T Δ𝐯2T … Δ𝐯𝑛T ]T . (4.82)
Δ𝐰𝑖

Ref. code: 25636122300053IAQ


31

The virtual work equation in Eq. (4.80) is linearized as (Dornisch, Müller &
Klinkel, 2016)

L[𝛿Π(𝐯, 𝛿𝐯)] = 𝛿Π + D[𝛿Π]Δ𝐯 = 0 (4.83)

where D[𝛿Π]Δ𝐯 is the linear increment of the total virtual work 𝛿Π with respect to Δ𝐯 given
by

𝐿 𝐿

D[𝛿Π]Δ𝐯 = D[𝛿Π𝑖𝑛𝑡 ]Δ𝐯 = ∫ 𝛿𝚪 T Δ𝐌 d𝜉 + ∫ Δ𝛿𝚪 T 𝐌 d𝜉. (4.84)


0 0

Here, it is assumed that the external loads are deformation independent.

The increment of the generalized stress vector 𝐌 can be obtained as

Δ𝐌 = 𝐃Δ𝚪 = 𝐃𝐇Δ𝐯. (4.85)

By using the above equation, the first integral term in Eq. (4.84) becomes

𝐿 𝐿 𝑛 𝑛 𝐿
T T
∫ 𝛿𝚪 Δ𝐌 d𝜉 = ∫ 𝛿𝚪 𝐃Δ𝚪 d𝜉 = ∑ ∑ 𝛿𝐯𝑖T (∫ 𝐇𝑖T 𝐃𝐇𝑗 d𝜉) Δ𝐯𝑗 . (4.86)
0 0 𝑖=1 𝑗=1 0

The increment of 𝛿𝚪 is obtained as

𝛿𝐚1T Δ𝐚1
𝛿𝐚1T Δ𝐚2 + 𝛿𝐚T2 Δ𝐚1 + 𝐚1T Δ𝛿𝐚2
𝛿𝐚1T Δ𝐚3 + 𝛿𝐚T𝟑 Δ𝐚1 + 𝐚1T Δ𝛿𝐚3
Δ𝛿𝚪 = . (4.87)
𝛿𝐚1T Δ𝐚′2 + (𝛿𝐚′2 )T Δ𝐚1 + 𝐚1T Δ𝛿𝐚′2
𝛿𝐚1T Δ𝐚′𝟑 + (𝛿𝐚′3 )T Δ𝐚1 + 𝐚1T Δ𝛿𝐚′𝟑
[𝛿𝐚T2 Δ𝐚′3 + (𝐚′3 )T Δ𝛿𝐚2 + (𝛿𝐚′3 )T Δ𝐚2 + 𝐚T2 Δ𝛿𝐚′3 ]

The increments of the kinematics in the above equation can be obtained by replacing, in
Eqs. (4.67) and (4.68), the symbol 𝛿 with the symbol Δ, i.e.,

Ref. code: 25636122300053IAQ


32

Δ𝐚1 = ∑ 𝑅𝑖′ (𝜂)Δ𝐩𝑖 (4.88)


𝑖=1

𝑛 𝑛
Δ𝐚𝛼 = ∑ −𝑅𝑖 (𝜂)𝐚̃Δ𝐰
𝛼 𝑖 Δ𝐚′𝛼 = ∑[−𝑅𝑖′ (𝜂)𝐚̃Δ𝐰
𝛼
̃′
𝑖 − 𝑅𝑖 (𝜂)𝐚𝛼 Δ𝐰𝑖 ].
𝑖=1 𝑖=1

(4.89)

Moreover, the increments of 𝛿𝐚2 and 𝛿𝐚3 can be written as

̃ Δ𝐰
Δ𝛿𝐚𝛼 = 𝛿𝐰 ̃ 𝐚𝛼 ̃′ )𝛿𝐰
Δ𝛿𝐚′𝛼 = (𝛿𝐰 ̃ 𝐚𝛼 + 𝛿𝐰 ̃′ )𝐚𝛼 + 𝛿𝐰
̃ (Δ𝐰 ̃ Δ𝐰
̃ 𝐚′𝛼 .

(4.90)

By substituting Eqs. (4.67)-(4.68), Eqs. (4.88)-(4.90) into Eq. (4.87), the second
integral term in Eq. (4.84) is then written as

𝐿 𝑛 𝑛 𝐿
T
∫ Δ𝛿𝚪 𝐌 d𝜉 = ∑ ∑ 𝛿𝐯𝑖T (∫ 𝐆𝑖𝑗 d𝜉) Δ𝐯𝑗 , (4.91)
0 𝑖=1 𝑗=1 0

where the matrix 𝐆𝑖𝑗 is defined as

𝐆𝑖𝑗,𝑢𝑢 𝐆𝑖𝑗,𝑢𝑤
𝐆𝑖𝑗 = [ ] (4.92)
𝐆𝑖𝑗,𝑤𝑢 𝐆𝑖𝑗,𝑤𝑤

and the components of 𝐆𝑖𝑗 are given by

𝐆𝑖𝑗,𝑢𝑢 = 𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)𝐈𝑁11 (4.93)

𝐆𝑖𝑗,𝑢𝑤 = [𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚


̃) T
2 ]𝑁
12
+ [𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚
̃)T
3 ]𝑁
13

′ T
+ [𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)(𝐚
̃)
2
T
+ 𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚̃
2) ]𝑀
12
(4.94)
T
+ [𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)(𝐚
̃)
3
T
+ 𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚̃

3) ]𝑀
13

Ref. code: 25636122300053IAQ


33

𝐆𝑖𝑗,𝑤𝑢 = [𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)𝐚


̃]𝑁
2
12
+ [𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)𝐚
̃]𝑁
3
13

+ [𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)𝐚 ̃′ ]𝑀12


̃2 + 𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)𝐚 2 (4.95)
+ [𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)𝐚 ̃′ ]𝑀13
̃3 + 𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)𝐚 2

𝐆𝑖𝑗,𝑤𝑤 = [𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚2 𝐚1T ) − 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚1T 𝐚2 )𝐈]𝑁12

+[𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚3𝐚1T ) − 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚1T 𝐚3 )𝐈]𝑁13

+[𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚2 𝐚1T ) − 𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚1T 𝐚2 )𝐈 + 𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)(𝐚2 𝐚1T )

−𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)(𝐚1T 𝐚2 )𝐈 + 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚′2 𝐚1T )

− 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚1T 𝐚′2 )𝐈]𝑀12 (4.96)

+[𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚3 𝐚1T ) − 𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚1T 𝐚3 )𝐈 + 𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)(𝐚3 𝐚1T )

−𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)(𝐚1T 𝐚3 )𝐈 + 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚′3 𝐚1T )


− 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)(𝐚1T 𝐚′3 )𝐈]𝑀13

+[𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)(𝐚


̃𝐚2̃ ̃𝐚
3−𝐚 3 ̃)]𝑀
2
23
.

Eventually, the linear increment of 𝛿Π with respect to Δ𝐯 is compactly written as

𝑛 𝑛

𝐷[𝛿Π]Δ𝐯 = ∑ ∑ 𝛿𝐯𝑖T 𝐊 𝑖𝑗 Δ𝐯𝑗 (4.97)


𝑖=1 𝑗=1

where

𝐊 𝑖𝑗 = ∫(𝐇𝑖T 𝐃𝐇𝑗 + 𝐆𝑖𝑗 ) d𝜉. (4.98)


0

The tangent stiffness matrix 𝐊 is defined as

𝑛 𝑛

𝛿𝐯 T 𝐊Δ𝐯 = ∑ ∑ 𝛿𝐯𝑖T 𝐊 𝑖𝑗 Δ𝐯𝑗 . (4.99)


𝑖=1 𝑗=1

Ref. code: 25636122300053IAQ


34

The linearized virtual work equation in Eq. (4.83) now yields the linearized system
equations, i.e.,

𝐿 𝐿

𝐊Δ𝐯 = ∫ 𝐋 𝐟 d𝜉 + 𝐏 − ∫ 𝐇 T 𝐌 d𝜉.
T
(4.100)
0 0

4.3.5 Kinematic update procedure

In the Newton-Raphson algorithm, the beam configuration must be updated


continuously. Assume that the configuration at the 𝑘-th iteration is known. The beam
(𝑘)
configuration parameters at the 𝑘-th iteration include the control points 𝐫𝑖 where 𝑖 =
(𝑘)
1,2, … , 𝑛, and the rotation tensors 𝚲𝑗 and its derivatives 𝚲𝑗′(𝑘) where 𝑗 = 1,2, … , 𝑛𝐺 . Here,
𝑛𝐺 denotes the number of Gauss points used in the Gauss integration, which is employed
in this study.
(𝑘) (𝑘)
The unknowns Δ𝐩𝑖 and Δ𝐰𝑖 are obtained as the solutions of Eq. (4.100). Since
the beam axis 𝐫0 (𝜂) is a member of ℝ3 , it is updated in a conventional additive manner by
using the following two relations, i.e.,

𝑛
(𝑘+1) (𝑘) (𝑘) (𝑘+1) (𝑘+1)
𝐫𝑖 = 𝐫𝑖 + Δ𝐩𝑖 𝐫0 (𝜂) = ∑ 𝑅𝑖 (𝜂)𝐫𝑖 . (4.101)
𝑖=1

The rotation tensor 𝚲 is a member of 𝑆𝑂(3). As a result, updating this quantity


requires an additional effort. The incremental axial vector Δ𝐰 and its derivative can be
obtained from Δ𝐰𝑖 as

𝑛 𝑛
(𝑘) (𝑘) (𝑘)
Δ𝐰 = ∑ 𝑅𝑖 (𝜂)Δ𝐰𝑖 Δ𝐰 ′(𝑘)
= ∑ 𝑅𝑖′ (𝜂)Δ𝐰𝑖 . (4.102)
𝑖=1 𝑖=1

The rotation tensor and its derivative at the 𝑗-th Gauss point are then updated as

Ref. code: 25636122300053IAQ


35

(𝑘+1) ̃ (𝑘) (𝜂(𝑗) )] 𝚲𝑗(𝑘)


𝚲𝑗 = exp[Δ𝐰 (4.103)

′ (𝑘)
𝚲𝑗′(𝑘+1) = exp[Δ𝐰
̃ (𝑘) (𝜂(𝑗) )] 𝚲𝑗 ̃ (𝑘) (𝜂(𝑗) )] 𝚲𝑗′(𝑘)
+ exp[Δ𝐰 (4.104)

where 𝜂(𝑗) is the 𝑗-th Gauss point. The determination of the derivative of the exponential
map of a skew-symmetric tensor is detailed in Appendix A. Once the rotation tensors at
the Gauss points are updated, the director vectors of the cross-sections at the Gauss points
are obtained by Eq. (4.9) and their derivatives are determined by Eq. (4.11).

4.4 Numerical examples and discussions

To show the accuracy and efficiency of the proposed Timoshenko beam


formulation, five numerical examples are solved. All the examples, except the last one, are
solved by using one NURBS patch. For simplicity, rectangular cross-sections are
considered in all the examples, and the moments of inertia are determined by Eq. (4.42) as
follows:

𝑏ℎ3 ℎ𝑏 3
𝐼2 = 𝐼3 = (4.105)
12 12

where ℎ and 𝑏 are, respectively, the thickness and width of the cross-section. Note that, for
a rectangular cross-section, 𝐼23 is zero.

In the proposed formulation, the effect of torsional warping of non-circular cross-


sections is not explicitly considered. To indirectly incorporate the effect of torsional
warping on the beam torsional stiffness, the polar moment of inertia, expressed in Eq.
(4.44), is not used for the considered rectangular cross-sections. Instead, the effective polar
moment of inertia is employed (Young, Budynas & Sadegh, 2012), i.e.,

1 3
𝑏 𝑏4
𝐽 = ℎ𝑏 [ − 0.21 (1 − )] for ℎ ≥ 𝑏. (4.106)
3 ℎ 12ℎ4

Ref. code: 25636122300053IAQ


36

Figure 4.4 Cantilever quadrant subjected to an end moment.

In case 𝑏 > ℎ, the roles of ℎ and 𝑏 in the above expression are simply interchanged. Note
that, in the Gauss integration, it is found that good results can be obtained with 𝑝 + 1 Gauss
points for each distinct knot span.

4.4.1 Cantilever quadrant subjected to an end moment

The first example involves the examination of the locking effects, i.e., membrane
locking and shear locking, on the numerical results obtained by the present formulation. A
cantilever quadrant in 𝑂𝑋𝑌 plane, shown in Figure 4.4, is considered. A concentrated
moment is applied at the free end in the positive direction of the 𝑍 axis with the magnitude
of

4𝐸𝐼2
𝑀=√ (4.107)
𝜋𝑅

Ref. code: 25636122300053IAQ


37

Figure 4.5 Cantilever quadrant subjected to an end moment: (a) Strain energy obtained
by various NURBS curves of different degrees. (b) Contribution of the membrane and
shear strain energy.

where 𝐸 is Young’s modulus. It is obvious that, for the given loading and boundary
conditions, there exists only the bending deformation. Therefore, this example is ideal for
the examination of the locking effects. In this example, only linear analysis is performed
by fully applying the moment 𝑀 in one step with one iteration.

The membrane, shear, and bending strains of the arch can be, respectively,
computed as (Adam, Bouabdallah, Zarroug & Maitournam, 2014)

d𝑢𝐀1 (𝜉) 𝑢𝐀𝟐 (𝜉)


𝜀𝑚 (𝜉) = − (4.108)
d𝜉 𝑅

𝑢𝐀1 (𝜉) d𝑢𝐀𝟐 (𝜉)


𝛾𝑠 (𝜉) = + − 𝜃𝐀𝟑 (𝜉) (4.109)
𝑅 d𝜉

d𝜃𝐀𝟑 (𝜉)
𝜒𝑏 (𝜉) = (4.110)
d𝜉

Ref. code: 25636122300053IAQ


38

where 𝑢𝐀1 (𝜉) and 𝑢𝐀2 (𝜉) are the displacements of the arch axis in the direction of the unit
tangent vector 𝐀1 and the director vector 𝐀 2 , respectively. Moreover, 𝜃𝐀3 (𝜉) is the cross-
sectional rotation around the director vector 𝐀 3 .

In this example, the locking effects are studied by considering the strain energy of
the arch. The numerical strain energy 𝐸𝑛𝑢𝑚 of the arch can be computed as

𝜋 𝜋 𝜋
𝑅 𝑅 𝑅
2 2 2
1 1 1
𝐸𝑛𝑢𝑚 = ∫ 𝜀𝑚 𝐸𝐴𝜀𝑚 d𝜉 + ∫ 𝛾𝑠 𝐺𝐴𝛾𝑠 d𝜉 + ∫ 𝜒𝑏 𝐸𝐼𝜒𝑏 d𝜉 = 𝐸𝑚 + 𝐸𝑠 + 𝐸𝑏
2 2 2
0 0 0

(4.111)

where 𝐸𝑚 , 𝐸𝑠 , and 𝐸𝑏 are the membrane, shear, and bending strain energy, respectively. In
addition, 𝐺 is the shear modulus. The exact solution for the strain energy 𝐸𝑒𝑥𝑎𝑐𝑡 is given
by

𝑀2 𝜋𝑅
𝐸𝑒𝑥𝑎𝑐𝑡 = = 1. (4.112)
4𝐸𝐼2

The arch is analyzed with various values of the ratio 𝑅/ℎ ranging from 1𝑒1
(relatively thick beams) to 1𝑒5 (very thin beams). The beam is also represented using
NURBS curves of different degrees 𝑝 = 2, 3, 4, 5, 6. The number of control points is fixed
at 15 to distinguish between the errors due to discretization and locking. The computed
values of the strain energy 𝐸𝑛𝑢𝑚 are plotted in Figure 4.5a together with the exact solution.
It is observed that, when 𝑅/ℎ = 1𝑒1, the present formulation offers good results for all the
values of 𝑝. When 𝑅/ℎ = 1𝑒5, the quadratic and cubic NURBS curves provide inaccurate
results.

The locking effects can be further shown by examining the ratio of the membrane
and shear strain energy to the total strain energy 𝐸𝑛𝑢𝑚 . This ratio is defined as

𝐸𝑚 + 𝐸𝑠
𝐶𝑚𝑠 = . (4.113)
𝐸𝑛𝑢𝑚

Ref. code: 25636122300053IAQ


39

Figure 4.6 Pre-twisted circular arch.

The ratio 𝐶𝑚𝑠 is computed and plotted in Figure 4.5b. Since there is only bending strain in
the beam, it is expected that 𝐶𝑚𝑠 is zero. It can be seen that, when 𝑅/ℎ is in the range of
thin beams, the majority of strain energy is due to the membrane and shear deformations
with the use of the quadratic and cubic NURBS curves. In contrast, for all values of 𝑅/ℎ,
the contribution of the membrane and shear deformations is entirely negligible with the use
of the NURBS curves of higher degrees. These results clearly reveal that the locking effects
happen when the quadratic and cubic NURBS curves are used.

In conclusion, the obtained numerical results show that the present formulation is
not locking-free, but the effects can be effectively reduced by using NURBS curves of high
degree for the approximations of the geometry and unknown kinematics.

4.4.2 Pre-twisted circular arch

The second example deals with the analysis of a quarter circular arch with a pre-
twisted configuration shown in Figure 4.6. In the same figure, the information of the cross-
sectional dimensions and material properties are also given. In the initial configuration, the

Ref. code: 25636122300053IAQ


40

Figure 4.7 Pre-twisted circular arch – Convergence tests for the deformation of the
arch axis.

arch axis lies in the 𝑂𝑋𝑌 plane while its cross-section rotates around the arch axis. The
initial twisted configuration is described by the angle 𝜙, defined in Figure 4.2, as

𝜋
𝜙 = (1 − 𝜂) . (4.114)
2

All the translational displacements and the axial rotations at both ends of the arch are fixed.
The arch is subjected to a uniformly distributed force 𝑞𝑍 = 1 kN/m in the negative direction
of the 𝑍 axis. This example is initially introduced by Greco and Cuomo (2013) and recently
considered by Radenković and Borković (2018). Once again, only linear analysis is
performed by fully applying the load in one step with one iteration. The excellent
convergence property of the IGA approach is one of its most prominent features. Here, this
property in the present formulation is verified thoroughly. As indicated by Bazilevs, Beirão
Da Veiga, Cottrell, Hughes and Sangalli (2006) and Greco and Cuomo (2013), when
NURBS curves of degree 𝑝 are used for the discretization of the geometry and unknown
kinematics, the minimum expected order of convergence is 𝑝 + 1 for displacements. As

Ref. code: 25636122300053IAQ


41

Figure 4.8 Pre-twisted circular arch: (a) Displacement 𝑢𝑋 . (b) Displacement 𝑢𝑌 . (c)
Displacement 𝑢𝑍 . (d) Reference and deformed configurations.

similar to Greco and Cuomo (2013), in this study, the order of convergence with respect to
the number of control points is considered. To test the convergence, the following relative
error is used, i.e.,

𝐿
∫0 ‖𝐫𝑛𝑢𝑚 − 𝐫𝑟𝑒𝑓 ‖𝑑𝜉
Relative error = 𝐿 (4.115)
∫0 ‖𝐫𝑟𝑒𝑓 ‖𝑑𝜉

where 𝐫𝑛𝑢𝑚 is the deformed arch axis obtained by using NURBS curves of various degrees
to approximate the geometry and the unknown kinematics, and 𝐫𝑟𝑒𝑓 is the deformed arch
axis obtained by using sextic NURBS curves with 100 control points. Figure 4.7 shows the
results of the convergence tests. The horizontal axis indicates the number of control points
while the vertical axis indicates the relative error. Each curve in the plot is associated with

Ref. code: 25636122300053IAQ


42

Figure 4.9 Rolled-up straight cantilever beam.

a fixed NURBS degree 𝑝. It can be observed that the minimum expected orders of
convergence of 𝑝 + 1 are achieved by all the considered NURBS degrees.

Next, the accuracy of the present formulation is verified by comparing the


displacements of the arch axis obtained by the present formulation with the results by those
formulations of Greco and Cuomo (2013) and Radenković and Borković (2018) . The arch
is also simulated by Radenković and Borković (2018) using Abaqus, a commercial finite
element package, with 314 B33 beam elements. In this study, the arch is represented using
sextic NURBS curves with 20 control points. Figure 4.8a-c show, respectively, the
displacement components 𝑢𝑋 , 𝑢𝑌 , and 𝑢𝑍 of the arch axis, obtained in this study and in the
works mentioned above. Note that the results in the work of Greco and Cuomo (2013) are
provide by Radenković and Borković (2018) through a private communication. The present
results match well with those from the literature. The reference and deformed
configurations obtained in this study are also shown in Figure 4.8d.

4.4.3 Rolled-up straight cantilever beam

This example studies a straight cantilever beam subjected to an end moment shown
in Figure 4.9. The problem is considered by many studies as a benchmark problem for

Ref. code: 25636122300053IAQ


43

Figure 4.10 Rolled-up straight cantilever beam – Single circle test: (a) Convergence
tests for the position of the free end. (b) Normalized load-displacement curves of the
free end. (c) Reference and final deformed configurations.

testing geometrically nonlinear beam formulations. This is because the beam exhibits very
large displacements and rotation. In addition, only the bending deformation occurs in the
beam and, because of that, the exact solutions can be easily derived. Since the beam axis
is straight and not a non-degenerate curve, the unit normal vector 𝐍 and the unit binormal
vector 𝐁 by the Frenet-Serret formulas are not defined. Subsequently, the director vectors
𝐀 2 and 𝐀 3 cannot be determined from Eq. (4.5). Here, an alternative method is used to
define 𝐍 and 𝐁. There always exists a rotation tensor 𝚲0 , such that (Crisfield, 2000)

𝐀1 = 𝚲0 𝐞1 . (4.116)

Note that 𝐀1 is the unit tangent vector of the reference beam axis. The vectors 𝐍 and 𝐁 are
then determined as

𝐍 = 𝚲 0 𝐞2 𝐁 = 𝚲 0 𝐞3 . (4.117)

Ref. code: 25636122300053IAQ


44

The director vectors 𝐀 2 and 𝐀 3 can then be determined from 𝐍 and 𝐁 by rotating them
around 𝐀1 with an angle 𝜙 using Eq. (4.5). In this study, for a straight beam, 𝐞1 is always
set along the initial beam axis and used as 𝐀1 , as shown in Figure 4.9.

The exact solutions for the displacements in the directions of the 𝑋 and 𝑌 axes of
the free end can be obtained, respectively, as

𝑢𝑋 sin 𝜔 𝑢𝑌 1 − cos 𝜔 𝑀𝐿
= −1 = 𝜔= . (4.118)
𝐿 𝜔 𝐿 𝜔 𝐸𝐼2

In this study, the beam is loaded up to two values of 𝑀, i.e., 2𝜋𝐸𝐼2 /𝐿 and 4𝜋𝐸𝐼2 /𝐿. The
first test is referred to as the single circle test while the second test is referred to as the
double circle test. In the single circle test, the beam is perfectly bent into a circle while in
the double circle test, the beam is wound around itself twice.

Table 4.1 Rolled-up straight cantilever beam – Single circle test: comparison of the
number of degrees of freedom.

Number of elements Number of


(Finite element degrees of
analysis)/distinct knot freedom
spans (IGA)

Bathe and Bolourchi (1979) 20 126

Ibrahimbegović, Frey and Kožar (1995) 10 66

Raknes et al. (2013) 21 75

Bauer, Breitenberger, Philipp, Wüchner 10 64


and Bletzinger (2016)

Present 1 48

4.4.3.1 Single circle test

Ref. code: 25636122300053IAQ


45

Figure 4.11 Rolled-up straight cantilever beam – Double circle test: (a)Convergence
tests of the position of the free end. (b) Normalized load-displacement curves of the
free end. (c) Reference and deformed configurations.

The convergence tests for the Euclidean norm ‖𝐫𝑒𝑛𝑑 ‖ of the free end at 𝑀 =
2𝜋𝐸𝐼2 /𝐿 are performed, and their results are plotted in Figure 4.10a. The orders of
convergence by all the used NURBS degrees are better than the minimum expected orders
of convergence of 𝑝 + 1.

Next, the beam is represented by septic NURBS curves with 8 control points. This
corresponds to 48 degrees of freedom. The displacements of the free end obtained by the
present formulation are plotted in Figure 4.10b together with the exact solutions. The
present results match well with the exact solutions. The deformed configuration of the
beam at the final load obtained in this study is shown in Figure 4.10c. It can be seen that

Ref. code: 25636122300053IAQ


46

Figure 4.12 Lateral buckling of a cantilever beam.

the deformed configuration, which is a circle, is perfectly represented by the present


formulation. Additionally, a comparison of the numbers of employed degrees of freedom
between the present formulation and some other beam formulations is shown in Table 4.1.
The present work employs the smallest number of degrees of freedom.

4.4.3.2 Double circle test

The results of the convergence tests for the double circle test are depicted in Figure
4.11a. The orders of convergence by all the employed NURBS degrees are found to be
better than the minimum expected orders of convergence of 𝑝 + 1.

Next, the beam is represented by sextic NURBS curves with 15 control points. This
corresponds to 90 degrees of freedom. The obtained displacements of the free end are
plotted in Figure 4.11b in comparison with the exact solutions. Good agreement between
the present results and the exact solutions is found. In addition, the deformed configurations
at some load steps, from this study, are visualized in Figure 4.11c. Very smooth
representations of these configurations are observed.

Ref. code: 25636122300053IAQ


47

Figure 4.13 Lateral buckling of a cantilever beam: (a) Convergence tests for the
deformation of the beam axis. (b) Load-displacement curves of the free end. (c)
Reference and deformed configurations.

4.4.4 Lateral buckling of a cantilever beam

The accuracy of the present formulation for analysis of beams under very large
displacements is demonstrated in the previous example. However, the previous problem is
a plane problem and involves only a two-dimensional displacement field. In this example,
the accuracy of the present formulation is further illustrated by using the buckling analysis
of a cantilever beam in Figure 4.12, which involves a three-dimensional displacement field.
The beam is subjected to a concentrated force 𝑃 at the free end. The extreme slenderness

Ref. code: 25636122300053IAQ


48

Figure 4.14 Twisting of a circular ring.

of the beam cross-section, i.e., ℎ/𝑏 = 20, must be noted. The lateral buckling of the beam
is induced by a perturbation lateral force of 0.001𝑃 as shown in Figure 4.12.

The convergence tests for the relative error defined in Eq. (4.115) at the final load
of 𝑃 = 2000 are performed. In this example, 𝐫𝑟𝑒𝑓 is the deformed beam axis obtained by
using sextic NURBS curves with 100 control points. The results of the convergence tests
are plotted in Figure 4.13a. It can be observed that the orders of convergence that are better
than the minimum expected orders of convergence of 𝑝 + 1 are obtained by all the used
NURBS degrees.

This beam is previously analyzed by Smoleński (1999), where it is split into ten
two-noded beam elements. To compare with the results obtained by Smoleński (1999), the
beam is represented in this study by sextic NURBS curves with 20 control points. The free-
end displacements obtained by this study and the work of Smoleński (1999) are compared
in Figure 4.13b. The comparison shows good agreement between the two solutions. The
deformations of the beam at some load steps, from this study, are also shown in Figure
4.13c.

4.4.5 Twisting of a circular ring

Ref. code: 25636122300053IAQ


49

Figure 4.15 Twisting of a circular ring – Load-displacement curves of the point 𝐵.

The examples considered so far involve the beam structures which can be
represented by only one NURBS patch. In fact, the presence of the rotational degrees of
freedom in the present formulation allows analysis of multi-patch beam structures to be
performed easily. This capability is shown in this example where twisting of a circular ring,
shown in Figure 4.14, is considered. The ring is fixed at the point 𝐵 and subjected a twisting
moment 𝑀 at the point 𝐶. The relation between the angle 𝜃 around the 𝑋 axis and the
moment 𝑀 at 𝐶 is investigated. The ring is first introduced by Goto, Watanabe, Kasugai
and Obata (1992). It is numerically solved by Pai and Palazotto (1996) using the multiple
shooting method and by Smoleński (1999) using the finite element method, among others.

In this study, the whole ring is analyzed. The ring is split into four NURBS patches,
and each patch is represented by septic NURBS curves. The total number of control points
used in this study for the whole ring is 28, which is equivalent to 15 control points if only

Ref. code: 25636122300053IAQ


50

Figure 4.16 Twisting of a circular ring – Deformed configurations.

half of the ring is analyzed. Note that 15 control points correspond to 90 degrees of
freedom. Goto, Watanabe, Kasugai and Obata (1992) simulate half of the ring with two
hundred beam elements with 1206 degrees of freedom, and Smoleński (1999) model half
of the ring using forty eight beam elements with 294 degrees of freedom. Figure 4.15 shows
the numerical results for the twisting angle 𝜃 at 𝐶, obtained by the present formulation and
the works of Goto, Watanabe, Kasugai and Obata (1992) and Smoleński (1999). The figure
also shows the numerical results obtained by Pai and Palazotto (1996) as reference results.
The numerical results from this study and Smoleński (1999) are virtually identical with the
reference results. Although many beam elements are used by Goto, Watanabe, Kasugai and
Obata (1992), the difference with the reference results is still remarkably noticeable. This
is probably due to the fact that the initial curvature of the ring is not well represented by
the geometry of the beam elements.

The deformed configurations of the ring are visualized in Figure 4.16 at every 90
degrees of the twisting angle 𝜃. It can be observed that the ring undergoes very large three-
dimensional deformations. Indeed, at the final equilibrium state considered in this study,
the ring is transformed into a smaller ring with a diameter of one third of its original size.

Ref. code: 25636122300053IAQ


51

With this example, the excellent efficiency and accuracy of the present formulation for
analysis of multi-patch beam structures subjected to very large three-dimensional
deformations are validated.

Ref. code: 25636122300053IAQ


52

CHAPTER 5

ISOGEOMETRIC EULER-BERNOULLI BEAM FORMULATION

5.1 Kinematic description of spatial Euler-Bernoulli beams

It is well-known that shear effects are neglected in the Euler-Bernoulli beam theory,
and as a result, beam cross-sections are always orthogonal to the beam axis during
deformation. To derive a new Euler-Bernoulli beam formulation, some modifications are
performed on the kinematic descriptions of Timoshenko beams in Chapter 4 to impose the
orthogonality between the cross-sections and the beam axis.

Mathematically, the orthogonality between the cross-sections and the beam axis
can be expressed as

𝐀 𝛼 ⋅ 𝐀1 = 0 𝐚𝛼 ⋅ 𝐚1 = 0 (5.1)

and consequently,

𝐀′𝛼 ⋅ 𝐀1 = −𝐀 𝛼 ⋅ 𝐀′1 𝐚′𝛼 ⋅ 𝐚1 = −𝐚𝛼 ⋅ 𝐚1′ . (5.2)

In Eqs. (5.1) and (5.2), 𝛼 is not 1.

By definition, the orthogonality between the cross-sections and the beam axis in the
reference configuration is automatically met. In the current configuration, an alternative
parametrization of the rotation tensor is presented to fulfil this orthogonality requirement.

5.1.1 Smallest rotation (SR) mapping and director vectors of Euler-Bernoulli beams

In the current configuration, the unit tangent vector of the beam axis is computed
as

𝐚1 𝐚1
𝐭= = . (5.3)
‖𝐚1 ‖ 𝑎1

Ref. code: 25636122300053IAQ


53

Figure 5.1 Director vectors in the current configuration of Euler-Bernoulli beams.

A set of three orthonormal vectors is considered, i.e., ̅̅


𝐀̅̅1 , ̅̅̅̅
𝐀 2 , and ̅̅̅̅
𝐀 3 . Choices of this set
of vectors will be discussed shortly. The three vectors are simultaneously rotated with the
same rotational angle and direction in such a way that the vector ̅̅
𝐀̅̅1 is rotated into the
vector 𝐭 by the smallest angle possible as described by Crisfield (2000). This mapping
scheme is called the smallest rotation (SR) mapping (Crisfield, 2000). The resulting vectors
of the SR mapping on the vectors ̅̅̅̅
𝐀 2 and ̅̅̅̅
𝐀 3 are given, respectively, by

̅̅̅̅
𝐀2 ⋅ 𝐭 ̅̅̅̅
𝐀3 ⋅ 𝐭
𝐧 = ̅̅̅̅
𝐀2 − (𝐭 + ̅̅
𝐀̅̅1 ) 𝐛 = ̅̅̅̅
𝐀3 − (𝐭 + ̅̅
𝐀̅̅1 ). (5.4)
1+𝐀 ̅̅̅̅1 ⋅ 𝐭 1+𝐀 ̅̅̅̅1 ⋅ 𝐭

Since ̅̅
𝐀̅̅1 , ̅̅̅̅
𝐀 2 , and ̅̅̅̅
𝐀 3 are rotated with the same rotational angle and direction, the vectors
𝐭, 𝐧, and 𝐛 in the above equations are also orthonormal. In analogous to the expressions of
the director vectors 𝐀 2 and 𝐀 3 in the reference configuration, the director vectors 𝐚2 and
𝐚3 are determined by a quasi-2D rotation of the vectors 𝐧 and 𝐛 around the vector 𝐭 with a
correction angle 𝜃. The resulting vectors of this rotation are given by

𝐚2 = cos 𝜃 𝐧 + sin 𝜃 𝐛 𝐚3 = − sin 𝜃 𝐧 + cos 𝜃 𝐛 (5.5)

𝐚′2 = 𝐚3 𝜃 ′ + cos 𝜃 𝐧′ + sin 𝜃 𝐛′ 𝐚′3 = −𝐚2 𝜃 ′ − sin 𝜃 𝐧′ + cos 𝜃 𝐛′ . (5.6)

Ref. code: 25636122300053IAQ


54

By denoting the rotation tensor associated with the mapping expressed via Eqs. (5.3)-(5.5)
as 𝚲𝐸𝐵 , the following compact relations can be written as

𝐭 = 𝚲𝐸𝐵 ̅̅̅̅
𝐀1 𝐚𝛼 = 𝚲𝐸𝐵 ̅̅̅̅
𝐀𝛼 . (5.7)

The determination of the vectors 𝐚2 and 𝐚3 is visualized in Figure 5.1.

It can be easily verified that the constraint for the current configuration in Eq. (5.1)
̅̅̅̅1 , ̅̅̅̅
is satisfied by 𝚲𝐸𝐵 . Since the set of vectors {𝐀 𝐀 2 , ̅̅̅̅
𝐀 3 } is given in advance, 𝚲𝐸𝐵 is entirely
identified by the vector 𝐭 and the axial rotation angle 𝜃. Recall that the vector 𝐭 can be
determined by the current beam axis 𝐫0 (𝜉) via Eq. (5.3). Therefore, for the proposed Euler-
Bernoulli beam formulation, the current beam axis 𝐫0 (𝜉) and the axial rotation angle 𝜃 are
considered as the primary kinematics.

̅̅̅̅1 , ̅̅̅̅
The method for obtaining suitable choices of {𝐀 𝐀 2 , ̅̅̅̅
𝐀 3 }, proposed by Meier,
Popp and Wall (2014), is adopted. In this study, the undeformed configuration is
considered as the reference configuration. Generally, at a point on the current beam axis
with the convective coordinate 𝜉, the unit tangent vector 𝐀1 and the director vectors 𝐀 2
and 𝐀 3 , determined at the same 𝜉 on the reference beam axis, can be chosen as the reference
̅̅̅̅1 , ̅̅̅̅
vectors {𝐀 𝐀 2 , ̅̅̅̅
𝐀 3 }. However, in some cases when the vectors 𝐀1 and 𝐭 are antiparallel,
̅̅̅̅1 , ̅̅̅̅
this choice of {𝐀 𝐀 2 , ̅̅̅̅
𝐀 3 } results in a singularity because 1 + 𝐀1 ⋅ 𝐭 = 0 and the relations
in Eq. (5.4) are thus undefined. This situation may happen when significantly large
deformations occur between the reference and current configurations. For these cases, the
unit tangent vector 𝐚1 and the director vectors 𝐚2 and 𝐚3 of a known configuration at a few
̅̅̅̅1 , ̅̅̅̅
loading steps prior to the current one can be used as {𝐀 𝐀 2 , ̅̅̅̅
𝐀 3 }, instead of {𝐀1 , 𝐀 2 , 𝐀 3 }.

5.1.2 Variations of kinematics

The variations of the beam axis 𝐫0 (𝜉) and the rotation tensor 𝚲𝐸𝐵 are determined
as

𝛿𝐫0 (𝜉) = 𝛿𝐮0 (𝜉) ̃ 𝚲𝐸𝐵 .


𝛿𝚲𝐸𝐵 = 𝛿𝐰 (5.8)

Ref. code: 25636122300053IAQ


55

Detailed derivations can be found in Chapter 4. Here, 𝛿𝐮0 (𝜉) is interpreted as an


infinitesimal virtual displacement vector superposed onto the current beam axis while
̃ = 𝑠𝑘𝑒𝑤(𝛿𝐰) adds an infinitesimal virtual rotation to the current rotation. The
𝛿𝐰
following relations are obtained, i.e.,

𝛿𝐭 = 𝛿𝐰 × 𝐭 (5.9)

𝛿𝐚𝛼 = 𝛿𝐰 × 𝐚𝛼 𝛿𝐚′𝛼 = 𝛿𝐰 ′ × 𝐚𝛼 + 𝛿𝐰 × 𝐚′𝛼 . (5.10)

For later use, a relation between the vector 𝛿𝐰 and the variations of the primary kinematics
is given by (Meier, Popp & Wall, 2014)

𝐭 ⊗ ̅̅
𝐀̅̅1 𝐭 × 𝛿𝐚1
𝛿𝐰 = 𝐭𝛿𝜃 + (𝐈 − ) . (5.11)
1 + ̅̅
𝐀̅̅1 ⋅ 𝐭 𝑎1

Additionally, the variations of 𝐧 and 𝐛 are derived as

(𝐭 + ̅̅̅̅
𝐀1 ) ⊗ ̅̅̅̅
𝐀2 + (𝐀 ̅̅̅̅2 ⋅ 𝐭)𝐈 (𝐀
̅̅̅̅2 ⋅ 𝐭)(𝐭 + ̅̅̅̅
𝐀1 ) ⊗ ̅̅̅̅
𝐀1
𝛿𝐧 = [− + ̅̅̅̅2 , 𝐭)𝛿𝐭
] 𝛿𝐭 = 𝐎(𝐀 (5.12)
̅̅̅̅
1 + 𝐀1 ⋅ 𝐭 ̅̅̅̅
(1 + 𝐀1 ⋅ 𝐭) 2

(𝐭 + ̅̅̅̅
𝐀1 ) ⊗ ̅̅̅̅
𝐀3 + (𝐀 ̅̅̅̅3 ⋅ 𝐭)𝐈 (𝐀
̅̅̅̅3 ⋅ 𝐭)(𝐭 + ̅̅̅̅
𝐀1 ) ⊗ ̅̅̅̅
𝐀1
𝛿𝐛 = [− + ̅̅̅̅3 , 𝐭)𝛿𝐭.
] 𝛿𝐭 = 𝐎(𝐀 (5.13)
̅̅̅̅
1 + 𝐀1 ⋅ 𝐭 ̅̅̅̅
(1 + 𝐀1 ⋅ 𝐭) 2

5.1.3 Virtual work principle

For computational convenience, from now onwards, equations are expressed in


matrix forms. Because of the orthogonality in Eq. (5.1), the shear strains 𝛤12 and 𝛤13 in Eqs.
(4.29) and (4.30) vanish. In addition, since there is no direct shear deformation, 𝐾12 and
𝐾13 in Eq. (4.28) become the bending curvatures. By using Eq. (5.2), alternative
expressions of 𝐾12 and 𝐾13 are obtained as

𝐾12 = 𝐀T2 𝐀′1 − 𝐚T2 𝐚1′ 𝐾13 = 𝐀T3 𝐀′1 − 𝐚T3 𝐚1′ . (5.14)

Ref. code: 25636122300053IAQ


56

Since the direct shear strains 𝛤12 and 𝛤13 vanish, the shear forces 𝑁12 and 𝑁13 also vanish.
A generalized strain vector 𝚪 and a generalized stress vector 𝐌 are defined for Euler-
Bernoulli beams as

𝚪 = [𝛤11 𝐾12 𝐾13 𝐾23 ]T 𝐌 = [𝑁11 𝑀12 𝑀13 𝑀23 ]T . (5.15)

Their relation can be obtained by using Eqs. (4.37), (4.40), (4.41), and (4.43) as

𝐸̅ 𝐴 0 0 0
0 𝐸̅ 𝐼2 𝐸̅ 𝐼23 0
𝐌 = 𝐃𝚪 𝐃= . (5.16)
0 𝐸̅ 𝐼23 𝐸̅ 𝐼3 0
[0 0 0 𝐺̅ 𝐽]

In similar to the Timoshenko beams in Chapter 4, the equilibrium is enforced by


using the virtual work principle as

𝛿Π = 𝛿Π𝑖𝑛𝑡 − 𝛿Π𝑒𝑥𝑡 = 0 (5.17)

where 𝛿Π𝑖𝑛𝑡 and 𝛿Π𝑒𝑥𝑡 are the internal and external virtual works, respectively. The
internal virtual work 𝛿Π𝑖𝑛𝑡 can be expressed as

𝛿Π𝑖𝑛𝑡 = ∫ 𝛿𝐄: 𝐒 d𝑉 = ∫ 𝛿𝚪 T 𝐌 d𝜉 (5.18)


𝑉 0

where 𝑉 and 𝐿 are the volume and length of the reference beam configuration, respectively.

The external virtual work 𝛿Π𝑒𝑥𝑡 is written as

𝛿Π𝑒𝑥𝑡 = ∫(𝛿𝐮T0 𝐅 + 𝛿𝜃𝑀) d𝜉 + 𝛿𝐮T0 (0)𝐟(0) + 𝛿𝐮T0 (𝐿)𝐟(𝐿) + 𝛿𝜃(0)𝑚(0) + 𝛿𝜃(𝐿)𝑚(𝐿)


0

(5.19)

where 𝐅 and 𝑀 are the distributed force and torsional moment. In addition, 𝐟(0) and 𝐟(𝐿)
are the concentrated forces applied at 𝜉 = 0 and 𝜉 = 𝐿, respectively. The same
interpretation holds for the concentrated torsional moments 𝑚(0) and 𝑚(𝐿).

Ref. code: 25636122300053IAQ


57

5.2 NURBS-based Euler-Bernoulli beams

5.2.1 NURBS discretization of the kinematics

The approximations of the beam axis in the reference and current configurations
are expressed as

𝑛 𝑛

𝐑 0 (𝜂) = ∑ 𝑅𝑖 (𝜂)𝐑 𝑖 𝐫0 (𝜂) = ∑ 𝑅𝑖 (𝜂)𝐫𝑖 . (5.20)


𝑖=1 𝑖=1

The displacement vector of the beam axis is determined as

𝑛 𝑛

𝐮0 (𝜂) = 𝐫0 (𝜂) − 𝐑 0 (𝜂) = ∑ 𝑅𝑖 (𝜂)(𝐫𝑖 − 𝐑 𝑖 ) = ∑ 𝑅𝑖 (𝜂)𝐩𝑖 (5.21)


𝑖=1 𝑖=1

where 𝐮𝑖 is the displacement vector of the 𝑖-th control point.

The tangent vectors are computed as

𝑛 𝑛

𝐀1 = ∑ 𝑅𝑖′ (𝜂)𝐑 𝑖 𝐚1 = ∑ 𝑅𝑖′ (𝜂)𝐫𝑖 . (5.22)


𝑖=1 𝑖=1

The axial rotation angle 𝜃 is approximated in the same manner as the beam axis as

𝑛 𝑛

𝜃 = ∑ 𝑅𝑖 (𝜂)𝜃𝑖 𝜃 ′ = ∑ 𝑅𝑖′ (𝜂)𝜃𝑖 . (5.23)


𝑖=1 𝑖=1

Eventually, there are four control variables at each control point, i.e., the displacement
vector 𝐩𝑖 having three components and the axial rotation angle 𝜃𝑖 . Vectors of control
variables are defined as

𝐯𝑖 = [𝐩T𝑖 𝜃𝑖 ]T 𝐯 = [𝐯1T 𝐯2T … 𝐯𝑛T ]T . (5.24)

Eqs. (5.22) and (5.23) subsequently yields

Ref. code: 25636122300053IAQ


58

𝑛 𝑛

𝛿𝐚1 = ∑ 𝑅𝑖′ (𝜂)𝛿𝐩𝑖 𝛿𝐚1′ = ∑ 𝑅𝑖′′ (𝜂)𝛿𝐩𝑖 (5.25)


𝑖=1 𝑖=1

𝑛 𝑛

𝛿𝜃 = ∑ 𝑅𝑖 (𝜂)𝛿𝜃𝑖 𝛿𝜃 = ∑ 𝑅𝑖′ (𝜂)𝛿𝜃𝑖



(5.26)
𝑖=1 𝑖=1

5.2.2 Parameterizations of the beam axis

As discussed in Chapter 4, the two parameterizations of the beam axis, i.e., arc-
length parameter 𝜉 and knot parameter 𝜂, can be related by using Eq. (4.69) as

d𝐑 0 (𝜂)
d𝜉 = ‖ ‖ d𝜂 = ‖𝐑̇ 0 (𝜂)‖d𝜂 = 𝐼d𝜂. (5.27)
d𝜂

It follows that

d𝑅𝑖 (𝜉) d𝜂 1 1 𝐑̇T0 (𝜂)𝐑̈ 0 (𝜂)


𝑅𝑖′ (𝜂) = = 𝑅̇ (𝜂) 𝑅𝑖′′ (𝜂) ̈
= 2 𝑅𝑖 (𝜂) − 𝑅̇𝑖 (𝜂)
d𝜂 d𝜉 𝐼 𝑖 𝐼 𝐼4

(5.28)

5.2.3 Nonlinear system equations

The variation of the generalized strain vector 𝚪 in Eq. (5.15) is obtained as

𝛿𝛤11 𝐚1T 𝛿𝐚1T 𝐚1T 𝛿𝐚1


𝛿𝐾 −(𝐚1′ )T 𝛿𝐚2 − 𝐚T2 𝛿𝐚1′ (𝐚1′ )T 𝐚
̃𝛿𝐰 − 𝐚T2 𝛿𝐚1′
𝛿𝚪 = [ 12 ] = = 2
. (5.29)
𝛿𝐾13 −(𝐚1′ )T 𝛿𝐚3 − 𝐚T3 𝛿𝐚1′ (𝐚1′ )T 𝐚
̃𝛿𝐰
3 − 𝐚T3 𝛿𝐚1′
𝛿𝐾23 [ (𝐚′ )T 𝛿𝐚 + 𝐚T 𝛿𝐚′ ] [ −𝐭 T 𝛿𝐰 ′ ]
3 2 2 3

By using Eq. (5.11), the vector 𝛿𝐰 can be determined from the primary kinematics and
their variations. Substituting Eqs. (5.11), (5.25), and (5.26) into Eq. (5.29) yields

Ref. code: 25636122300053IAQ


59

𝛿𝚪 = ∑ 𝐇𝑖 𝛿𝐯𝑖 = 𝐇𝛿𝐯 (5.30)


𝑖=1

where the matrices 𝐇𝑖 and 𝐇 are defined as

𝑅𝑖′ (𝜂)𝐚1T 0
′ ′ T
𝑅𝑖 (𝜂)(𝐚1 ) 𝐚 ̃𝐗(𝐀
2
̅̅̅̅1 , 𝐭) − 𝑅𝑖 (𝜂)𝐚2
′′ T
𝑅𝑖 (𝜂)(𝐚1′ )T 𝐚̃𝐭
2
𝐇𝑖 = (5.31)
′ ′ T
𝑅𝑖 (𝜂)(𝐚1 ) 𝐚 ̃𝐗(𝐀
3
̅̅̅̅
1 , 𝐭) − 𝑅 ′′
𝑖 (𝜂)𝐚 T
3 𝑅𝑖 (𝜂)(𝐚 ′
1 ) T
̃𝐭
𝐚3

[−𝑅𝑖 (𝜂)𝐭 ̅̅
T ′ (𝐀 ̅̅
𝐗 1 , 𝐭) − 𝑅𝑖 ′′
(𝜂)𝐭 T ̅̅̅̅
𝐗(𝐀1 , 𝐭) −𝑅𝑖′
(𝜂) ]

𝐇 = [𝐇1 𝐇2 … 𝐇𝑛 ]. (5.32)

̅̅̅̅1 , 𝐭) and its derivative are given in Appendix C.


The matrix 𝐗(𝐀

The internal virtual work 𝛿Π𝑖𝑛𝑡 can now be rewritten as

𝛿Π𝑖𝑛𝑡 = 𝛿𝐯 T ∫ 𝐇 T 𝐌 d𝜉. (5.33)


0

Regarding the external virtual work 𝛿Π𝑒𝑥𝑡 , the following matrices are introduced
as

𝐝 = [𝐅 T 𝑀]T (5.34)

𝐏 = [𝐟 T (0) 𝑚(0) … 𝐟 T (𝐿) 𝑚(𝐿)]T (5.35)

𝐂 = [𝑅1 (𝜂)𝐈4 𝑅2 (𝜂)𝐈4 … 𝑅𝑛 (𝜂)𝐈4 ] (5.36)

where 𝐈4 is an identity matrix of size 4 × 4. A compact expression of the external virtual


work 𝛿Π𝑒𝑥𝑡 is obtained as

𝛿Π𝑒𝑥𝑡 = 𝛿𝐯 T ∫ 𝐂 T 𝐝 d𝜉 + 𝛿𝐯 T 𝐏. (5.37)
0

Substituting Eqs. (5.33) and (5.37) into Eq. (5.17) yields the system equations as

Ref. code: 25636122300053IAQ


60

𝐿 𝐿

∫ 𝐇 𝐌 d𝜉 = ∫ 𝐂 T 𝐝 d𝜉 + 𝐏.
T
(5.38)
0 0

The system equations are nonlinear in terms of the control variables, and they are solved
using the Newton-Raphson algorithm. Their linearization is given below.

5.2.4 Linearization of the system equations

Recall that the external loads are assumed to be deformation independent in this
study, the linearization of the total virtual work 𝛿Π is expressed as

L[𝛿Π] = 𝛿Π + D[𝛿Π]Δ𝐯 = 0 (5.39)

where D[𝛿Π]Δ𝐯 is the linear increment of the total virtual work 𝛿Π with respect to Δ𝐯. This
increment can be determined as

𝐿 𝐿

D[𝛿Π]Δ𝐯 = D[𝛿Π𝑖𝑛𝑡 ]Δ𝐯 = ∫ 𝛿𝚪 T Δ𝐌 d𝜉 + ∫ Δ𝛿𝚪 T 𝐌 d𝜉. (5.40)


0 0

In the first term of the right-hand side of Eq. (5.40), Δ𝐌 can be obtained as

Δ𝐌 = 𝐃Δ𝚪 = 𝐃𝐇Δ𝐯. (5.41)

It follows that

𝐿 𝐿

∫ 𝛿𝚪 Δ𝐌 d𝜉 = 𝛿𝐯 (∫ 𝐇 T 𝐃𝐇 d𝜉) Δ𝐯.
T T
(5.42)
0 0

In the second term of the right-hand side of Eq. (5.40), Δ𝛿𝚪 can be expressed as

Ref. code: 25636122300053IAQ


61

Δ𝛿Γ11 𝛿𝐚1T Δ𝐚1


′ T̃
Δ𝛿𝐾12 −𝛿𝐰 T 𝐚
̃Δ𝐚
2

̃Δ𝐰
1 − 𝛿𝐰 𝐚1 𝐚 2 + (𝐚1′ )T 𝐚̃Δ𝛿𝐰
2 + (𝛿𝐚1′ )T 𝐚
̃Δ𝐰
2
Δ𝛿𝚪 = [ ]=
Δ𝛿𝐾13 T
−𝛿𝐰 𝐚 ̃Δ𝐚
3
′ T̃′
̃Δ𝐰
1 − 𝛿𝐰 𝐚1 𝐚 3

+ (𝐚1 ) 𝐚T
̃Δ𝛿𝒘
3
′ T
+ (𝛿𝐚1 ) 𝐚 ̃Δ𝐰
3
Δ𝛿𝐾23 [ ′ T
−(𝛿𝐰 ) Δ𝐭 − 𝐭 Δ𝛿𝐰 T ′ ]

(5.43)

in which the components of Δ𝛿𝚪 can be compactly expressed as

Δ𝛿𝛤11 = 𝛿𝐯 T 𝐋11 Δ𝐯 Δ𝛿𝐾12 = 𝛿𝐯 T 𝐋12 Δ𝐯 (5.44)

Δ𝛿𝐾13 = 𝛿𝐯 T 𝐋13 Δ𝐯 Δ𝛿𝐾23 = 𝛿𝐯 T 𝐋23 Δ𝐯. (5.45)

The matrices 𝐋11, 𝐋12, 𝐋13, and 𝐋23 are given in Appendix D. Using these matrices, the
second term in the right-hand side of Eq. (5.40) can be written as

𝐿 𝐿

∫ Δ𝛿𝚪 T 𝐌 d𝜉 = 𝛿𝐯 T [∫(𝐋11𝑁11 + 𝐋12 𝑀12 + 𝐋13 𝑀13 + 𝐋23 𝑀23 ) d𝜉] Δ𝐯. (5.46)
0 0

Substituting Eqs. (5.42) and (5.46) into Eq. (5.40) yields an alternative expression
of the linear increment D[𝛿Π]Δ𝐯 as

D[𝛿Π]Δ𝐯 = 𝛿𝐯 [∫(𝐇 T 𝐃𝐇 + 𝐋11 𝑁11 + 𝐋12 𝑀12 + 𝐋13 𝑀13 + 𝐋23 𝑀23 ) d𝜉] Δ𝐯 = 𝛿𝐯 T 𝐊Δ𝐯
T

(5.47)

where 𝐊 is referred to as the tangent stiffness matrix. Substituting Eqs. (5.33), (5.37), and
(5.47) into Eq. (5.39) finally gives

𝐿 𝐿

𝐊Δ𝐯 = ∫ 𝐂 T 𝐝 d𝜉 + 𝐏 − ∫ 𝐇 T 𝐌 d𝜉. (5.48)


0 0

5.3 Multi-patch Euler-Bernoulli beams

Ref. code: 25636122300053IAQ


62

5.3.1 Axial rotation and total cross-sectional rotations

In some scenarios, e.g., multi-patch beam structures or beams with rotational


boundary conditions, it is not convenient to use the Euler-Bernoulli beam formulation
presented in the last section. This is because the correction axial rotation angle 𝜃 is locally
defined with respect to each patch in a beam structure and may not in general be continuous
across multiple patches. These drawbacks can be removed by using the total cross-sectional
rotations at the ends of each patch as degrees of freedom. Here, a relation between the axial
rotation and the total cross-sectional rotation is derived.

The following relations can be obtained from Eq. (5.5), i.e.,

cos 𝜃 = 𝐧T 𝐚2 sin 𝜃 = 𝐛T 𝐚2 . (5.49)

Applying the operator 𝛿 to the above expressions yields

− sin 𝜃 𝛿𝜃 = 𝐚T2 𝛿𝐧 + 𝐧T 𝛿𝐚2 cos 𝜃 𝛿𝜃 = 𝐚T2 𝛿𝐛 + 𝐛T 𝛿𝐚2 . (5.50)

With the help of Eqs. (5.9), (5.10), (5.12), and (5.13), Eqs. (5.49) and (5.50) yield

sin2 𝜃 𝛿𝜃 = 𝐛T 𝐚2 [𝐧T 𝐚 ̅̅̅̅2 , 𝐭)𝐭̃]𝛿𝐰


̃2 + 𝐚T2 𝐎(𝐀 (5.51)

cos 2 𝜃 𝛿𝜃 = −𝐧T 𝐚2 [𝐛T 𝐚 ̅̅̅̅


̃2 + 𝐚T2 𝐎(𝐀 3 , 𝐭)𝐭̃]𝛿𝐰. (5.52)

Subsequently, the following relation is obtained, i.e.,

̅̅̅̅2 , 𝐭)𝐭̃] − 𝐧T 𝐚2 [𝐛T 𝐚


̃2 + 𝐚T2 𝐎(𝐀
𝛿𝜃 = {𝐛T 𝐚2 [𝐧T 𝐚 ̅̅̅̅3 , 𝐭)𝐭̃]}𝛿𝐰 = 𝐕𝛿𝐰.
̃2 + 𝐚T2 𝐎(𝐀

(5.53)

It will be shown shortly that this relation allows the consideration of the total cross-
sectional rotations at the ends of patches as discrete unknowns.

5.3.2 Transformations of degrees of freedom

Consider the second and second-last control points, i.e., 𝐫2 and 𝐫𝑛−1. Their position
vectors can be determined as

Ref. code: 25636122300053IAQ


63

𝐫2 = 𝐫1 + 𝑙1 𝐭1 𝐫𝑛−1 = 𝐫𝑛 − 𝑙𝑛−1 𝐭 𝑛 . (5.54)

Here, 𝐭1 and 𝐭 𝑛 are, respectively, the unit tangent vectors at the first and last ends of the
patch. In addition, the lengths of the first and last segments of the control polygon, 𝑙1 and
𝑙𝑛−1, are given by

𝑙1 = ‖𝐫2 − 𝐫1 ‖ 𝑙𝑛−1 = ‖𝐫𝑛 − 𝐫𝑛−1 ‖. (5.55)

Applying the operator 𝛿 to Eq. (5.54) yields

𝛿𝐩2 = 𝛿𝐩1 + 𝐭1 𝛿𝑙1 − 𝑙1 𝐭̃1 𝛿𝐰1 𝛿𝐩𝑛−1 = 𝛿𝐩𝑛 − 𝐭 𝑛 𝛿𝑙𝑛−1 + 𝑙𝑛−1 𝐭̃𝑛 𝛿𝐰𝑛 .

(5.56)

By using Eqs. (5.53) and (5.56), the following transformations are obtained, i.e.,

𝛿𝐩1 𝐈 𝟎3 𝟎 𝟎 𝛿𝐩1
𝛿𝜃 𝟎T 𝐕(0) 0 0 𝛿𝐰1
[ 1] = [ ] (5.57)
𝛿𝐩2 𝐈 −𝑙1 𝐭̃1 𝐭1 0 𝛿𝑙1
𝛿𝜃2 [𝟎T 𝟎T 0 1] 𝛿𝜃2

𝛿𝐩𝑛−1 𝟎 −𝐭 𝑛 𝐈 𝑙𝑛−1 𝐭̃𝑛 𝛿𝜃𝑛−1


𝛿𝜃 1 0 𝟎T 𝟎T 𝛿𝑙
[ 𝑛−1 ] = [ 𝑛−1 ] (5.58)
𝛿𝐩𝑛 𝟎 𝟎 𝐈 𝟎3 𝛿𝐩𝑛
𝛿𝜃𝑛 [0 0 𝟎T 𝐕(𝐿) ] 𝛿𝐰 𝑛

where 𝟎3 is a zero matrix of size of 3 × 3, and 𝟎 is a zero vector of size of 3 × 1.


Additionally, 𝐕(0) and 𝐕(𝐿) are evaluated using the relevant kinematic quantities at 𝜉 = 0
and 𝜉 = 𝐿, respectively.

Define a new vector of degrees of freedom 𝐱, via its variation, and a new point load
vector 𝐏𝐱 as

𝛿𝐱
= [𝛿𝐩1T 𝛿𝐰1T 𝛿𝑙1 𝛿𝜃2 𝛿𝐩T3 𝛿𝜃3 … 𝛿𝐩𝑛−2 𝛿𝜃𝑛−2 𝛿𝜃𝑛−1 𝛿𝑙𝑛−1 𝛿𝐩T𝑛 𝛿𝐰𝑛T ]T

(5.59)

Ref. code: 25636122300053IAQ


64

𝐏𝐱 = [𝐟 T (0) 𝐦T (0) 0 0 0 0 … 0 0 0 0 𝐟 T (𝐿) 𝐦T (𝐿)]T (5.60)

where 𝐦(0) and 𝐦(𝐿) are the concentrated moments applied at 𝜉 = 0 and 𝜉 = 𝐿,
respectively. With the help of Eqs. (5.57) and (5.58), the following transformations
between the control variables and the degrees of freedom can be obtained as

𝛿𝐯 = 𝐓𝛿𝐱 Δ𝐯 = 𝐓Δ𝐱. (5.61)

Here, the matrix 𝐓 is the transformation matrix.

The internal and external virtual works, i.e., 𝛿Π𝑖𝑛𝑡 and 𝛿Π𝑒𝑥𝑡 , can be alternatively
expressed in terms of the degrees of freedom as

𝐿 𝐿

𝛿Π𝑖𝑛𝑡 = 𝛿𝐯 T ∫ 𝐇 T 𝐌 d𝜉 = 𝛿𝐱 T 𝐓 T ∫ 𝐇 T 𝐌 d𝜉 (5.62)
0 0

𝐿 𝐿

𝛿Π𝑒𝑥𝑡 = 𝛿𝐯 T ∫ 𝐂 T 𝐝 d𝜉 + 𝛿𝐯 T 𝐏 = 𝛿𝐱 T 𝐓 T ∫ 𝐂 T 𝐝 d𝜉 + 𝛿𝐱 T 𝐏𝐱 . (5.63)
0 0

Subsequently, the system equations and their linearization can be written as

𝐿 𝐿

𝐓 ∫ 𝐇 𝐌 d𝜉 = 𝐓 ∫ 𝐂 T 𝐝 d𝜉 + 𝐏𝐱
T T T
(5.64)
0 0

𝐿 𝐿

𝐓 T 𝐊𝐓Δ𝐱 = 𝐓 T ∫ 𝐂 T 𝐝 d𝜉 + 𝐏𝐱 − 𝐓 T ∫ 𝐇 T 𝐌 d𝜉. (5.65)


0 0

5.4 Numerical examples and discussions

Considering the total cross-sectional rotations at patch ends as degrees of freedom


allows straightforward enforcement of rigid connections and rotational boundary
conditions. In this section, several numerical examples are shown to clarify this statement.
Except the last example, all examples consider beams with rectangular cross-sections,

Ref. code: 25636122300053IAQ


65

Figure 5.2 Square frame subjected to opposite compression forces: (a) Problem set up.
(b) Equivalent model.

whose moments of inertia and polar moment of inertia are, respectively, determined by
using Eqs. (4.105) and (4.106).

All the examples in this section are analyzed using NURBS curves with full inner-
element continuity.

5.4.1 Square frame subjected to opposite compression forces

In the first example, the convenience of using the proposed formulation to deal with
rotational boundary conditions is illustrated. Figure 5.2a shows a square frame subjected
to a pair of opposite compression forces. Due to the symmetry, only a quarter of the frame
is analyzed and the equivalent model used is visualized in Figure 5.2b. In the equivalent
model, two sextic NURBS curves with 13 control points, i.e., 𝐑1 to 𝐑13 , are used to
represent the two members.

Here, the enforcement of the supports at 𝐵 and 𝐶 without rotational degrees of


freedom is discussed. Let 𝑢𝑋𝑖 and 𝑢𝑌𝑖 denote the horizontal and vertical displacements of

Ref. code: 25636122300053IAQ


66

Figure 5.3 Square frame: (a) Normalized load-displacement curves. (b) Deformed
configurations.

the 𝑖th control point. At 𝐵 and 𝐶, only the horizontal displacement 𝑢𝑋1 and the vertical
displacement 𝑢𝑌13 are not restrained. Since the displacements of 𝐑1 and 𝐑13 are considered
as degrees of freedom, the restrained displacements at 𝐵 and 𝐶 can be straightforwardly
enforced as

𝑢𝑌1 = 0 𝑢𝑋13 = 0. (5.66)

On the contrary, the enforcement of the rotational boundary conditions at 𝐵 and 𝐶 is more
involved. In fact, due to the lack of rotational degrees of freedom, the following relations
should be used, i.e.,

𝑢𝑋1 = 𝑢𝑋2 𝑢𝑌12 = 𝑢𝑌13 . (5.67)

These constraints are determined based on the conditions that the tangent vectors at 𝐵 and
𝐶 must retain their original directions. The Lagrange multiplier approach can be used to
implement these constraints. However, it should be noticed that the above relations are
determined based on the initial geometry of the frame. In other words, the constraints used
to prescribe the rotational boundary conditions are unfortunately problem-dependent.

With the proposed formulation, the boundary conditions at 𝐵 and 𝐶 can be


conveniently enforced since rotational degrees of freedom are included. The accuracy of

Ref. code: 25636122300053IAQ


67

Figure 5.4 Lee’s frame: (a) Problem set up. (b) Normalized load-displacement curves
of the point 𝐷. (c) Deformed frame axes.

the proposed formulation is verified by comparing the unrestrained displacements at 𝐵 and


𝐶 obtained by the proposed formulation with those obtained by means of elliptical integrals
(Mattiasson, 1981). The result comparisons are visualized in Figure 5.3a where very good
agreement is observed. Furthermore, deformations of the quarter of the frame at several
loading steps are plotted in Figure 5.3b. It can be observed that the tangent vectors at 𝐵
and 𝐶 are always parallel to their directions in the reference configuration. Thus, the
enforcement of the rotational boundary conditions at these ends is confirmed.

Ref. code: 25636122300053IAQ


68

Figure 5.5 A pre-twisted circular arch – Problem set up.

5.4.2 Lee’s frame

In this example, the accuracy and robustness of the proposed formulation is further
examined by analyzing Lee's frame shown in Figure 5.4a. This example is widely
considered as a benchmark problem for geometrically nonlinear analysis of frames, and the
exact solutions are derived by Lee, Manuel and Rossow (1968). The frame is pinned at 𝐵
and 𝐹, and a concentrated force is prescribed at point 𝐷. The horizontal and downward
displacements at the point 𝐷, i.e., 𝑢 and 𝑣, are considered for result comparisons.

The frame consists of three segments, i.e., 𝐵𝐶, 𝐶𝐷, and 𝐷𝐹. In order to maintain
the right angle between the segments 𝐵𝐶 and 𝐶𝐷 and thereby enforce the rigid connection
at 𝐶, some special treatments must be used for isogeometric Euler-Bernoulli beam
formulations that do not include rotational degrees of freedom, e.g., adding a bending strip

at 𝐶 (Huang, He, Jiang, Qiao & Wang, 2016). With the proposed formulation, no special
treatments are required. The frame is analyzed by Simo and Vu-Quoc (1986) using ten
quadratic beam elements with 63 in-plane degrees of freedom. Most recently, the frame is
modeled by Yuan, Wang and Kardomateas (2019) using two ten-noded co-rotational
quadrature beam elements with 57 degrees of freedom.

Ref. code: 25636122300053IAQ


69

Figure 5.6 A pre-twisted circular arch: (a) Top view. (b) NURBS representations of
the arch and control points.

In this study, each segment is represented by a septic NURBS curve with 22 control
points or 42 in-plane degrees of freedom in total. A comparison between the results
obtained by the proposed formulation and the exact solutions is given in Figure 5.4b. It can
be observed that the normalized load-displacement curves exhibit severe snap-though and
snap-back behaviors. The obtained excellent agreement between the present results and the
exact solutions validates the accuracy and robustness of the proposed formulation for
stability analysis of beam structures. Deformed configurations of the frame at some loading
steps are also visualized in Figure 5.4c, showing the tangent vectors at 𝐶 and 𝐷 that are
determined separately from each segment. Even with very large displacements, the rigid
connections of patches are perfectly enforced.

Ref. code: 25636122300053IAQ


70

5.4.3 A pre-twisted circular arch

The capability of the proposed formulation for geometrically nonlinear analysis of


planar beam structures has been verified. Here and in the following examples, applications
of the proposed formulation in analysis of three-dimensional beam structures are explored.
A pre-twisted circular arch of radius 𝑅 = 1 shown in Figure 5.5 is analyzed. The most
interesting feature of this arch is the twisted configuration of the arch, whose axis is
restricted in the 𝑋𝑌 plane. With this example, many strong advantages of the proposed
formulation can be illustrated.

First, the exact geometric representation of this complex arch is straightforwardly


defined. The arch is split into two halves, i.e., 𝐵𝐶 and 𝐶𝐷. The top view of the arch in
Figure 5.6a shows the geometry of the axis of each half. It can be observed that the two
axes are parts of a circle of unit radius. Thus, NURBS representations of the axes can be
easily constructed. As for the twisted continuum configuration, as described in section 4.1,
the correction angles of the halves 𝐵𝐶 and 𝐶𝐷 are, respectively, given by

𝛼 𝛼 − 130
𝜙=− × 90, (0 ≤ 𝛼 ≤ 65) 𝜙= × 90, (65 ≤ 𝛼 ≤ 130) (5.68)
65 65

where 𝛼 is the angular coordinate of the arch (see Figure 5.6a). Here, the correction angles
are computed in degrees. By using NURBS representations for the axes and the above
correction angles, the exact geometric representation of the arch is easily obtained. On the
contrary, the simulation of this arch is not a trivial task when the conventional finite
element approach is used. In fact, this arch can be simulated by using many straight beam
elements, each with a constant cross-section. To reduce the errors of the geometric
approximation and to ensure the result accuracy, a dense mesh of finite elements should be
used. Figure 5.7 shows a screenshot of the arch in Abaqus with 600 B33 beam elements.
However, specifying the correct orientations of a large number of elements is highly time-
consuming. Additionally, a smooth geometric representation is not obtained, which can be
clearly seen in the magnified figure in Figure 5.7.

Along its whole axis, the arch is subjected to a uniformly distributed force in the
negative direction of the 𝑍-axis. The arch is also fixed at both ends, i.e., at 𝐵 and 𝐷. As

Ref. code: 25636122300053IAQ


71

Figure 5.7 A pre-twisted circular arch – A simulation in Abaqus.

Figure 5.8 A pre-twisted circular arch – Convergence test for the relative error in the
deformation of the arch axis.

discussed in the first example, the boundary conditions relating the displacements at the
arch ends can be straightforwardly prescribed. However, without rotational degrees of
freedom, some additional constraints are required to enforce the rotational boundary
conditions at 𝐵 and 𝐷. Figure 5.6b shows a NURBS representation of the arch axis with

Ref. code: 25636122300053IAQ


72

Figure 5.9 A pre-twisted circular arch: (a) Displacement 𝑢𝑋 . (b) Displacement 𝑢𝑌 . (c)
Displacement 𝑢𝑍 . (d) Reference and deformed configurations (with a scale factor of
1000 for displacements).

11 control points, i.e., 𝐑1 to 𝐑11 . The displacements along the coordinate axes of the 𝑖th
control point are denoted by 𝑢𝑋𝑖 , 𝑢𝑌𝑖 , and 𝑢𝑍𝑖 , respectively. If rotational degrees of freedom
are not used, the additional constraints for the arch in this example are given by

𝑢𝑌2 = −𝑢𝑋2 tan 650 𝑢𝑍2 = 0 (5.69)

𝑢𝑌10 = 𝑢𝑋10 tan 650 𝑢𝑍10 = 0. (5.70)

Once again, the constraints above are problem-dependent since they are determined based
on the initial geometry of the arch. For spatial beam structures with several complex

Ref. code: 25636122300053IAQ


73

rotational boundary conditions, determination of all additional constraints is tedious. With


the proposed formulation, the need of additional constraints is removed.

Higher accuracy per degree-of-freedom is one of the most prominent advantages of


using NURBS for analysis. As studied by Bazilevs, Beirão Da Veiga, Cottrell, Hughes and
Sangalli (2006) and Greco and Cuomo (2013), with the use of NURBS curves of degree 𝑝
for the discretization of unknown kinematics, a minimum convergence order of 𝑝 + 1 is
expected. In this example, to test this convergence property, the arch is analyzed by using
NURBS curves with different degrees. A linear analysis is performed with one full loading
step and one iteration. The relative error in the displacement of the arch axis is defined as

130
𝜋 𝑟𝑒𝑓
∫0180 ‖𝐫0𝑛𝑢𝑚 − 𝐫0 ‖ d𝜉
𝑟𝑒𝑙𝑎𝑡𝑖𝑣𝑒𝐸𝑟𝑟 = 130 . (5.71)
𝜋 𝑟𝑒𝑓
∫0180 ‖𝐫0 ‖ d𝜉

𝑟𝑒𝑓
Here, 𝐫0 denotes the deformed axis, obtained by using two septic NURBS curves with
100 control points, and 𝐫0𝑛𝑢𝑚 denotes the deformed axis obtained by the NURBS curve
being tested. In this test, the relative error is considered to be dependent on the number of
control points. The resulting relative errors are plotted in Figure 5.8. In the figure,
quadratic, cubic, and quartic NURBS curves, i.e., 𝑝 = 2, 3, and 4, are examined. It is
observed that the minimum expected order of 𝑝 + 1 is obtained by using cubic and quartic
NURBS curves.

Lastly, the accuracy of the proposed formulation is verified. Two septic NURBS
curves with 15 control points or 58 degrees of freedom are used for the analysis. Since no
reference results are found in the literature, the arch is analyzed with the beam formulation
presented by Vo, Nanakorn and Bui (2020) using two septic NURBS curves and 100
control points. Furthermore, the arch is also simulated in Abaqus with 600 B33 beam
elements. The displacements of the arch axis, i.e., 𝑢𝑋 , 𝑢𝑌 , and 𝑢𝑍 , are plotted in Figure
5.9a-c. Very good agreement is observed between the results obtained by the proposed
formulation and those using the formulation proposed by Vo, Nanakorn and Bui (2020)
and Abaqus. Additionally, the reference and deformed configurations of the arch are

Ref. code: 25636122300053IAQ


74

Figure 5.10 Twisting of a circular ring: (a) Problem set up. (b) Moment-rotation curve
of the point 𝐷.

visualized in Figure 5.9d. Both the reference and deformed configurations are represented
with very smooth geometric representations.

5.4.4 Twisting of a circular ring

This example is devoted to illustrating the accuracy and efficiency of the proposed
formulation for analysis of beam structures subjected to very large three-dimensional
displacements and rotations. A circular ring shown in Figure 5.10a is considered. The ring
is fixed at point 𝐵 and subjected to a concentrated moment 𝑀 around the 𝑋-axis at point
𝐷. A solution for the relation between the moment 𝑀 and the rotation angle 𝛼 is derived
by Pai and Palazotto (1996) using shooting methods, and this solution is used as the
reference. For an easy comparison, the same geometrical and material properties with those
in the work of Pai and Palazotto (1996) are used in this study, as provided in Figure 5.10a.
Additionally, for a better discussion, the term “connection” is used to denote a connection
between patches in IGA and between elements in the conventional finite element approach.

The ring has been analyzed in several published works, e.g., (Goto, Watanabe,
Kasugai & Obata, 1992; Meier, Popp & Wall, 2014; Smoleński, 1999; Vo, Nanakorn &

Ref. code: 25636122300053IAQ


75

Bui, 2020) using different sets of discrete unknowns. The formulations proposed by Goto,
Watanabe, Kasugai and Obata (1992), Smoleński (1999), and Vo, Nanakorn and Bui
(2020) are Timoshenko beam formulations in which the displacements of the beam axis
and the total cross-sectional rotation are considered as unknown kinematics. Therefore,
rigid connections and rotational boundary conditions can be easily considered. Meier, Popp
and Wall (2014) use similar unknown kinematics to those considered in the present study,
i.e., the translational displacements of the beam axis and the axial rotation. However, an
alternative approach is used to handle rigid connections. It can be seen from Figure 5.10a
that the unit tangent vectors at the point 𝐶, determined from the segments 𝐵𝐶 and 𝐶𝐷, are
initially identical. In the work of Meier, Popp and Wall (2014), the connection at this point
is handled by considering the tangent vectors as additional discrete unknowns. Clearly, this
approach cannot be directly used to consider connections of beams with kinks. In addition,
applications of concentrated moments require additional treatments. In the work of Meier,
Popp and Wall (2014), an accurate geometric representation of the beam axis is constructed
with a problem-dependent parameter that requires more computational efforts.

Table 5.1 Twisting of a circular ring. Comparison of the number of degrees of freedom
for a half of the ring.

Authors Number of degrees of freedom

Goto, Watanabe, Kasugai and Obata 1206


(1992)

Smoleński (1999) 294

Meier, Popp and Wall (2014) 135

Vo, Nanakorn and Bui (2020) 90

Present 58

Ref. code: 25636122300053IAQ


76

Figure 5.11 Twisting of a circular ring – Deformed configurations.

In this study, the ring is split into four segments, i.e., 𝐵𝐶, 𝐶𝐷, 𝐷𝐹, and 𝐹𝐵. Exact
geometric representations of the segments are easily obtained using four septic NURBS
curves. No additional treatments are required for the supports at 𝐵, and the connections of
patches at 𝐶, 𝐷 and 𝐹. Figure 5.10a shows the results of the rotation angle 𝛼 obtained by
the proposed formulation and those in literature (Goto, Watanabe, Kasugai & Obata, 1992;
Pai & Palazotto, 1996; Smoleński, 1999; Vo, Nanakorn & Bui, 2020). The results obtained
by Smoleński (1999) and Vo, Nanakorn and Bui (2020) are identical with the reference
result (Pai & Palazotto, 1996). A comparison of the numbers of degrees of freedom used
in the literature is reported in Table 5.1. One can observe excellent agreement between the
present result, where the least number of degrees of freedom is employed, and reference
result. With this obtained result, the accuracy and efficiency of the proposed formulation
are verified. Moreover, the ring exhibits very large deformations, and the use of Euler-
Bernoulli beam formulations, e.g., presented by Meier, Popp and Wall (2014) and the
proposed formulation, provides high accuracy with fewer degrees of freedom in
comparison with Timoshenko beam formulations, e.g., (Goto, Watanabe, Kasugai &

Ref. code: 25636122300053IAQ


77

Figure 5.12 Compression of a lattice tower: (a) Problem set up. (b) Target surface and
embedded curves.

Obata, 1992; Smoleński, 1999; Vo, Nanakorn & Bui, 2020). This feature proves the
efficiency of Euler-Bernoulli beam formulations in analysis of highly flexible beam
structures.

The deformations of the ring at every increment of 90 degrees of the rotation angle
𝛼 are plotted in Figure 5.11a-d. It is apparent that the ring undergoes very large three-
dimensional displacements and rotations. Indeed, at the final equilibrium state, the ring is
deformed into a smaller one with the radius of one-third of the original radius. With this
example, the accuracy and efficiency of the proposed formulation for analysis of spatial
multi-patch beam structures undergoing large displacements and rotations are clearly
illustrated.

5.4.5 Compression of a lattice tower

Ref. code: 25636122300053IAQ


78

Figure 5.13 Compression of a lattice tower: (a) NURBS representation of the


generatrix of the target surface with the axis of revolution. (b) Parent domains of the
embedded curves.

This example is concerned with analysis of a tower-like lattice structure shown in


Figure 5.12a. The lattice tower is constructed using 36 beam segments connected at the
joints. The tower is fixed at the basement joints, i.e., joints 𝐵, 𝐶, 𝐷 and 𝐹. A distributed
force 𝑞𝑍 is applied on the top of the tower. The tower is analyzed with the proposed
formulation using a quartic NURBS curve to represent a beam segment. The downward
displacements at the joints 𝐾 and 𝑁, i.e., 𝑣𝐾 and 𝑣𝑁 , are used for result comparisons. It can
be observed that several beams with different orientations are connected at each joint. Thus,
this example is ideal to illustrate the capability of the present formulation for analysis of
complex structures with arbitrarily oriented beams.

A four-step procedure described by Choi and Cho (2018) is used to construct the
tower. The necessary quantities for the construction of the tower are given here. The beam
segments of the tower are in fact curves embedded on a target surface. Figure 5.12b shows
the target surface and the embedded curves, i.e., red solid lines, used for this example. The
target surface is a revolution surface of a given generatrix. The generatrix and the axis of
revolution are shown in Figure 5.13a. The embedded curves are mapped onto the target

Ref. code: 25636122300053IAQ


79

Figure 5.14 Compression of a lattice tower – Comparison of the displacement at the


joints 𝐾 and 𝑁.

surface from lines in a parent domain shown in Figure 5.13b. With the given target surface
and lines in the parent domain, the mapping procedure presented by Choi and Cho (2018)
is performed to construct the tower considered in this example.

Since there exist no reference solutions for the downward displacements at the
joints 𝐾 and 𝑁, the beam formulation presented by Vo, Nanakorn and Bui (2020) is used
to analyze the same tower, and the obtained results are considered as the reference. The
result comparisons are shown in Figure 5.14. Good agreement between the results validates
the accuracy of the proposed formulation. Furthermore, to show the enforcement of the
rigid joints, the deformed configurations obtained by using the beam formulation
developed by Vo, Nanakorn and Bui (2020) and the proposed one are visualized in Figure
5.15a-b, respectively. Identical deformed configurations are observed.

Ref. code: 25636122300053IAQ


80

Figure 5.15 Compression of a lattice tower – Deformed configurations: (a) Beam


formulation developed by Vo, Nanakorn and Bui (2020). (b) The proposed
formulation.

Ref. code: 25636122300053IAQ


81

CHAPTER 6

CONCLUSIONS

In this contribution, two new efficient Timoshenko and Euler-Bernoulli beam


formulations are developed for geometrically nonlinear analysis of spatial beam structures
using the IGA approach. In both beam formulations, continuum beam configurations are
described by the beam axis and director vectors of cross-sections. The Green-Lagrange
strain and the second Piola-Kirchhoff stress tensors are used as the energy conjugate pair.

The first part of this study is devoted to the Timoshenko beam formulation. This
study is the first effort to derive such a formulation with the aforementioned energy
conjugate pair in the context of isogeometric analysis. This approach facilitates
consideration of hyperelastic materials in analysis of highly flexible beam structures. The
employed unknown kinematics include the displacements of the beam axis and the cross-
sectional rotation. The cross-sectional rotation is represented in terms of an orthogonal
tensor, which is parameterized by a vector-like parameter with the aid of Rodrigues’
rotation formula. Natural exponentiation and superposition are employed for updating the
cross-sectional rotations. This methodology enables the developed Timoshenko beam
formulation to simulate beams subjected to arbitrary large displacements and rotations.

The second part focuses on the Euler-Bernoulli beam formulation for analysis of
multi-patch beam structures. The displacements of the beam axis and the axial cross-
sectional rotation are considered as the unknown kinematics. Finite cross-sectional
rotations are described by using the smallest rotation mapping. This essentially reduces the
nonlinearity of the strain measurements in terms of the unknown kinematics, and
consequently lessens the computational efforts for the required linearization. To deal with
rigid connections in multi-patch beam structures, a novel nonlinear transformation between
the axial cross-sectional rotation and the total cross-sectional rotation is derived. This
transformation allows the use of the total cross-sectional rotations at the ends of beam
patches as discrete unknowns. The accuracy of this approach in the nonlinear regime is

Ref. code: 25636122300053IAQ


82

investigated for the first time. Additionally, with the end total cross-sectional rotations, the
difficulties found in enforcing rotational boundary conditions are removed.

The accuracy and efficiency of the proposed beam formulations are assessed via
several numerical examples. Regarding the locking phenomena of the Timoshenko beam
formulation, i.e., membrane and shear locking, the obtained numerical results indicate that
the locking effects can be significantly reduced by using NURBS curves of high degree for
the approximations of the geometry and unknown kinematics. However, the membrane
locking effect in the Euler-Bernoulli beam formulation is not yet elaborated in this study.
Beams with complex initial geometry, e.g., pre-twisted configurations, are conveniently
handled with the proposed method with the use of NURBS beam axes and director vectors.
The excellent convergence property of the present formulations is clearly observed in the
considered examples, where better orders of convergence than the theoretical expected
orders of convergence of 𝑝 + 1 are achieved. On the other hand, the capability of the
present Euler-Bernoulli beam formulation for analysis of complex multi-patch beam
structures undergoing large three-dimensional deformations is also validated. The obtained
results also show good agreement with those available in the literature.

In all the numerical examples, sextic and septic NURBS curves, i.e., 𝑝 = 6, 7, are
mainly used for the analysis to avoid the locking effects. Although acceptable results can
be obtained with the use of either quartic or quintic NURBS curves, i.e., 𝑝 = 4, 5, more
distinct knot spans are required to obtain the same accuracy. As a result, the total number
of Gauss points required for the numerical integration can in fact be increased when quartic
or quintic NURBS curves are used. This is obviously not computationally effective.
Therefore, sextic and septic NURBS curves are employed in this work to balance the
accuracy and computational efficiency. For future works, locking-free beam formulations
for geometrically nonlinear analysis are worth being developed.

Ref. code: 25636122300053IAQ


83

REFERENCES

Adam, C., Bouabdallah, S., Zarroug, M., & Maitournam, H. (2014). Improved numerical
integration for locking treatment in isogeometric structural elements, part i: Beams.
Computer Methods in Applied Mechanics and Engineering, 279, 1-28.
10.1016/j.cma.2014.06.023
Argyris, J. (1982). An excursion into large rotations. Computer Methods in Applied
Mechanics and Engineering, 32(1), 85-155. https://doi.org/10.1016/0045-
7825(82)90069-X
Auricchio, F., Beirão da Veiga, L., Kiendl, J., Lovadina, C., & Reali, A. (2013). Locking-
free isogeometric collocation methods for spatial Timoshenko rods. Computer
Methods in Applied Mechanics and Engineering, 263, 113-126.
https://doi.org/10.1016/j.cma.2013.03.009
Bathe, K.-J., & Bolourchi, S. (1979). Large displacement analysis of three-dimensional
beam structures. International Journal for Numerical Methods in Engineering,
14(7), 961-986. 10.1002/nme.1620140703
Bauer, A. M., Breitenberger, M., Philipp, B., Wüchner, R., & Bletzinger, K. U. (2016).
Nonlinear isogeometric spatial Bernoulli beam. Computer Methods in Applied
Mechanics and Engineering, 303, 101-127.
https://doi.org/10.1016/j.cma.2015.12.027
Bauer, A. M., Wüchner, R., & Bletzinger, K. U. (2019). Weak coupling of nonlinear
isogeometric spatial Bernoulli beams. Computer Methods in Applied Mechanics
and Engineering, 112747. https://doi.org/10.1016/j.cma.2019.112747
Bazilevs, Y., Beirão Da Veiga, L., Cottrell, J. A., Hughes, T. J. R., & Sangalli, G. (2006).
Isogeometric analysis: Approximation, stability and error estimates for h-refined
meshes. Mathematical Models and Methods in Applied Sciences, 16(07), 1031-
1090. 10.1142/S0218202506001455
Cardona, A., & Geradin, M. (1988). A beam finite element non-linear theory with finite
rotations. International Journal for Numerical Methods in Engineering, 26(11),
2403-2438. 10.1002/nme.1620261105

Ref. code: 25636122300053IAQ


84

Chen, L., Nguyen-Thanh, N., Nguyen-Xuan, H., Rabczuk, T., Bordas, S. P. A., & Limbert,
G. (2014). Explicit finite deformation analysis of isogeometric membranes.
Computer Methods in Applied Mechanics and Engineering, 277, 104-130.
https://doi.org/10.1016/j.cma.2014.04.015
Choi, M.-J., & Cho, S. (2018). Constrained isogeometric design optimization of lattice
structures on curved surfaces: Computation of design velocity field. Structural and
Multidisciplinary Optimization, 58(1), 17-34. 10.1007/s00158-018-2000-9
Choi, M.-J., & Cho, S. (2019). Isogeometric configuration design sensitivity analysis of
geometrically exact shear-deformable beam structures. Computer Methods in
Applied Mechanics and Engineering, 351, 153-183.
https://doi.org/10.1016/j.cma.2019.03.032
Crisfield, M. A. (1990). A consistent co-rotational formulation for non-linear, three-
dimensional, beam-elements. Computer Methods in Applied Mechanics and
Engineering, 81(2), 131-150. https://doi.org/10.1016/0045-7825(90)90106-V
Crisfield, M. A. (2000). Non-linear finite element analysis of solids and structures. Wiley.
Dornisch, W., & Klinkel, S. (2014). Treatment of Reissner–Mindlin shells with kinks
without the need for drilling rotation stabilization in an isogeometric framework.
Computer Methods in Applied Mechanics and Engineering, 276, 35-66.
http://dx.doi.org/10.1016/j.cma.2014.03.017
Dornisch, W., Klinkel, S., & Simeon, B. (2013). Isogeometric Reissner–Mindlin shell
analysis with exactly calculated director vectors. Computer Methods in Applied
Mechanics and Engineering, 253, 491-504.
http://dx.doi.org/10.1016/j.cma.2012.09.010
Dornisch, W., Müller, R., & Klinkel, S. (2016). An efficient and robust rotational
formulation for isogeometric Reissner–Mindlin shell elements. Computer Methods
in Applied Mechanics and Engineering, 303, 1-34.
http://dx.doi.org/10.1016/j.cma.2016.01.018
Fang, W., Yu, T., Van Lich, L., & Bui, T. Q. (2019). Analysis of thick porous beams by a
quasi-3D theory and isogeometric analysis. Composite Structures, 221, 110890.
https://doi.org/10.1016/j.compstruct.2019.04.062

Ref. code: 25636122300053IAQ


85

Goto, Y., Watanabe, Y., Kasugai, T., & Obata, M. (1992). Elastic buckling phenomenon
applicable to deployable rings. International Journal of Solids and Structures,
29(7), 893-909. https://doi.org/10.1016/0020-7683(92)90024-N
Greco, L., & Cuomo, M. (2013). B-spline interpolation of Kirchhoff-Love space rods.
Computer Methods in Applied Mechanics and Engineering, 256, 251-269.
http://dx.doi.org/10.1016/j.cma.2012.11.017
Greco, L., & Cuomo, M. (2014). An implicit g1 multi patch b-spline interpolation for
Kirchhoff–Love space rod. Computer Methods in Applied Mechanics and
Engineering, 269, 173-197. https://doi.org/10.1016/j.cma.2013.09.018
Greco, L., & Cuomo, M. (2015). Consistent tangent operator for an exact Kirchhoff rod
model. Continuum Mechanics and Thermodynamics, 27(4), 861-877.
10.1007/s00161-014-0361-x
Huang, Z., He, Z., Jiang, W., Qiao, H., & Wang, H. (2016). Isogeometric analysis of the
nonlinear deformation of planar flexible beams with snap-back. Acta Mechanica
Solida Sinica, 29(4), 379-390. 10.1016/s0894-9166(16)30241-5
Hughes, T. J. R., Cottrell, J. A., & Bazilevs, Y. (2005). Isogeometric analysis: CAD, finite
elements, NURBS, exact geometry and mesh refinement. Computer Methods in
Applied Mechanics and Engineering, 194(39–41), 4135-4195.
http://dx.doi.org/10.1016/j.cma.2004.10.008
Ibrahimbegovic, A. (1997). On the choice of finite rotation parameters. Computer Methods
in Applied Mechanics and Engineering, 149(1), 49-71.
https://doi.org/10.1016/S0045-7825(97)00059-5
Ibrahimbegović, A. (1995). On finite element implementation of geometrically nonlinear
reissner's beam theory: Three-dimensional curved beam elements. Computer
Methods in Applied Mechanics and Engineering, 122(1), 11-26.
https://doi.org/10.1016/0045-7825(95)00724-F
Ibrahimbegović, A., Frey, F., & Kožar, I. (1995). Computational aspects of vector-like
parametrization of three-dimensional finite rotations. International Journal for
Numerical Methods in Engineering, 38(21), 3653-3673. 10.1002/nme.1620382107
Kiendl, J., Bazilevs, Y., Hsu, M. C., Wüchner, R., & Bletzinger, K. U. (2010). The bending
strip method for isogeometric analysis of Kirchhoff–Love shell structures

Ref. code: 25636122300053IAQ


86

comprised of multiple patches. Computer Methods in Applied Mechanics and


Engineering, 199(37–40), 2403-2416.
http://dx.doi.org/10.1016/j.cma.2010.03.029
Kiendl, J., Bletzinger, K. U., Linhard, J., & Wüchner, R. (2009). Isogeometric shell
analysis with Kirchhoff–Love elements. Computer Methods in Applied Mechanics
and Engineering, 198(49–52), 3902-3914.
http://dx.doi.org/10.1016/j.cma.2009.08.013
Kiendl, J., Hsu, M.-C., Wu, M. C. H., & Reali, A. (2015). Isogeometric Kirchhoff–Love
shell formulations for general hyperelastic materials. Computer Methods in Applied
Mechanics and Engineering, 291, 280-303.
https://doi.org/10.1016/j.cma.2015.03.010
Klinkel, S., & Govindjee, S. (2002). Using finite strain 3D‐material models in beam and
shell elements. Engineering Computations, 19(3), 254-271.
10.1108/02644400210423918
Lai, W., Yu, T., Bui, T. Q., Wang, Z., Curiel-Sosa, J. L., Das, R., & Hirose, S. (2017). 3-d
elasto-plastic large deformations: Iga simulation by bézier extraction of NURBS.
Advances in Engineering Software, 108, 68-82.
https://doi.org/10.1016/j.advengsoft.2017.02.011
Lee, S. L., Manuel, F. S., & Rossow, E. C. (1968). Large deflections and stability of elastic
frame. Journal of the Engineering Mechanics Division, 94(2), 521-548.
Marino, E. (2016). Isogeometric collocation for three-dimensional geometrically exact
shear-deformable beams. Computer Methods in Applied Mechanics and
Engineering, 307, 383-410. https://doi.org/10.1016/j.cma.2016.04.016
Marino, E. (2017). Locking-free isogeometric collocation formulation for three-
dimensional geometrically exact shear-deformable beams with arbitrary initial
curvature. Computer Methods in Applied Mechanics and Engineering, 324, 546-
572. https://doi.org/10.1016/j.cma.2017.06.031
Marino, E., Kiendl, J., & De Lorenzis, L. (2019). Explicit isogeometric collocation for the
dynamics of three-dimensional beams undergoing finite motions. Computer
Methods in Applied Mechanics and Engineering, 343, 530-549.
https://doi.org/10.1016/j.cma.2018.09.005

Ref. code: 25636122300053IAQ


87

Marino, E., Kiendl, J., & De Lorenzis, L. (2019). Isogeometric collocation for implicit
dynamics of three-dimensional beams undergoing finite motions. Computer
Methods in Applied Mechanics and Engineering, 356, 548-570.
https://doi.org/10.1016/j.cma.2019.07.013
Mattiasson, K. (1981). Numerical results from large deflection beam and frame problems
analysed by means of elliptic integrals. International Journal for Numerical
Methods in Engineering, 17(1), 145-153. 10.1002/nme.1620170113
Meier, C., Popp, A., & Wall, W. A. (2014). An objective 3D large deformation finite
element formulation for geometrically exact curved Kirchhoff rods. Computer
Methods in Applied Mechanics and Engineering, 278, 445-478.
https://doi.org/10.1016/j.cma.2014.05.017
Meier, C., Popp, A., & Wall, W. A. (2019). Geometrically exact finite element
formulations for slender beams: Kirchhoff–Love theory versus Simo–Reissner
theory. Archives of Computational Methods in Engineering, 26(1), 163-243.
10.1007/s11831-017-9232-5
Pai, P. F., & Palazotto, A. N. (1996). Large-deformation analysis of flexible beams.
International Journal of Solids and Structures, 33(9), 1335-1353.
https://doi.org/10.1016/0020-7683(95)00090-9
Piegl, L., & Tiller, W. (1997). The NURBS book. New York, United States: Springer.
Radenković, G., & Borković, A. (2018). Linear static isogeometric analysis of an
arbitrarily curved spatial Bernoulli–Euler beam. Computer Methods in Applied
Mechanics and Engineering, 341, 360-396.
https://doi.org/10.1016/j.cma.2018.07.010
Raknes, S. B., Deng, X., Bazilevs, Y., Benson, D. J., Mathisen, K. M., & Kvamsdal, T.
(2013). Isogeometric rotation-free bending-stabilized cables: Statics, dynamics,
bending strips and coupling with shells. Computer Methods in Applied Mechanics
and Engineering, 263, 127-143. https://doi.org/10.1016/j.cma.2013.05.005
Reissner, E. (1973). On one-dimensional large-displacement finite-strain beam theory.
Studies in Applied Mathematics, 52(2), 87-95. 10.1002/sapm197352287

Ref. code: 25636122300053IAQ


88

Simo, J. C. (1985). A finite strain beam formulation. The three-dimensional dynamic


problem. Part i. Computer Methods in Applied Mechanics and Engineering, 49(1),
55-70. https://doi.org/10.1016/0045-7825(85)90050-7
Simo, J. C., & Vu-Quoc, L. (1986). A three-dimensional finite-strain rod model. Part ii:
Computational aspects. Computer Methods in Applied Mechanics and Engineering,
58(1), 79-116. https://doi.org/10.1016/0045-7825(86)90079-4
Smoleński, W. M. (1999). Statically and kinematically exact nonlinear theory of rods and
its numerical verification. Computer Methods in Applied Mechanics and
Engineering, 178(1), 89-113. https://doi.org/10.1016/S0045-7825(99)00006-7
Tepole, A. B., Kabaria, H., Bletzinger, K.-U., & Kuhl, E. (2015). Isogeometric Kirchhoff–
Love shell formulations for biological membranes. Computer Methods in Applied
Mechanics and Engineering, 293, 328-347.
https://doi.org/10.1016/j.cma.2015.05.006
Vo, D., & Nanakorn, P. (2020). A total Lagrangian Timoshenko beam formulation for
geometrically nonlinear isogeometric analysis of planar curved beams. Acta
Mechanica, 231(7), 2827-2847. 10.1007/s00707-020-02675-x
Vo, D., Nanakorn, P., & Bui, T. Q. (2020). A total Lagrangian Timoshenko beam
formulation for geometrically nonlinear isogeometric analysis of spatial beam
structures. Acta Mechanica, 10.1007/s00707-020-02723-6
Weeger, O., Narayanan, B., De Lorenzis, L., Kiendl, J., & Dunn, Martin L. (2017). An
isogeometric collocation method for frictionless contact of cosserat rods. Computer
Methods in Applied Mechanics and Engineering, 321, 361-382.
https://doi.org/10.1016/j.cma.2017.04.014
Weeger, O., Narayanan, B., & Dunn, M. L. (2019). Isogeometric shape optimization of
nonlinear, curved 3D beams and beam structures. Computer Methods in Applied
Mechanics and Engineering, 345, 26-51.
https://doi.org/10.1016/j.cma.2018.10.038
Weeger, O., Yeung, S.-K., & Dunn, M. L. (2017). Isogeometric collocation methods for
cosserat rods and rod structures. Computer Methods in Applied Mechanics and
Engineering, 316, 100-122. https://doi.org/10.1016/j.cma.2016.05.009

Ref. code: 25636122300053IAQ


89

Young, W. C., Budynas, R. G., & Sadegh, A. M. (2012). Roak's fomulas for stress and
strain. New York, United States: McGraw-Hill.
Yu, T., Hu, H., Zhang, J., & Bui, T. Q. (2019). Isogeometric analysis of size-dependent
effects for functionally graded microbeams by a non-classical quasi-3D theory.
Thin-Walled Structures, 138, 1-14. https://doi.org/10.1016/j.tws.2018.12.006
Yu, T., Lai, W., & Bui, T. Q. (2019). Three-dimensional elastoplastic solids simulation by
an effective iga based on bézier extraction of NURBS. International Journal of
Mechanics and Materials in Design, 15(1), 175-197. 10.1007/s10999-018-9405-x
Yu, T., Zhang, J., Hu, H., & Bui, T. Q. (2019). A novel size-dependent quasi-3D
isogeometric beam model for two-directional FG microbeams analysis. Composite
Structures, 211, 76-88. https://doi.org/10.1016/j.compstruct.2018.12.014
Yuan, Z., Wang, X., & Kardomateas, G. A. (2019). A co-rotational weak-form quadrature
planar beam element for geometric nonlinear static and dynamic analysis.
International Journal for Numerical Methods in Engineering, 0(0),
10.1002/nme.6183

Ref. code: 25636122300053IAQ


90

APPENDICES

Ref. code: 25636122300053IAQ


91

APPENDIX A

PARAMETERIZATION OF THE ROTATION TENSOR 𝚲

There are several choices for the parameterization of the rotation tensor 𝚲, e.g.,
Euler angles, quaternion parameters, and Rodrigues’ rotation formula (Crisfield, 2000). In
this study, Rodrigues’ rotation formula is used.

Any rotation can be described by an axis of rotation and a rotation angle about that
axis. Let 𝛉 = 𝜃𝑖 𝐞𝑖 be a vector-like parameter whose direction is the axis of rotation and the
magnitude 𝜃 = ‖𝛉‖ is the rotation angle about 𝛉. Unless the rotation angle 𝜃 is small, 𝜃𝑖
̃
should not be interpreted as the component rotation about 𝐞𝑖 . A skew-symmetric tensor 𝛉
is defined using the components of 𝛉 as

0 −𝜃3 𝜃2
̃ = [ 𝜃3
𝛉 0 −𝜃1 ]. (A.1)
−𝜃2 𝜃1 0

̃ (Argyris, 1982), i.e.,


The rotation tensor 𝚲 is the exponential map of 𝛉

sin 𝜃 1 − cos 𝜃
̃) = 𝐈 +
𝚲 = exp(𝛉 ̃+
𝛉 ̃𝛉
𝛉 ̃. (A.2)
𝜃 𝜃2

This formulation is known as Rodrigues’ rotation formula, which is the most robust
formulation for finite rotations in three-dimensional spaces. For a relatively small 𝜃, the
following properties should be used for stability, i.e.,

sin 𝜃 1 − cos 𝜃 1
lim ( )=1 lim ( ) = . (A.3)
𝜃→0 𝜃 𝜃→0 𝜃2 2

The derivative of 𝚲 with respect to 𝜉 is determined from

sin 𝜃 ′ sin 𝜃 ′ 1 − cos 𝜃 ′ 1 − cos 𝜃 ′


′ ̃ )′ = (
𝚲 = exp(𝛉 ̃
) 𝛉+ ̃
𝛉 +( ) 𝛉̃𝛉
̃+ (𝛉̃ 𝛉
̃+𝛉
̃𝛉̃′ )
𝜃 𝜃 𝜃 2 𝜃2

(A.4)

Ref. code: 25636122300053IAQ


92

𝛉′ = 𝜃𝑖′ 𝐞𝑖 (A.5)

𝛉′ ⋅ 𝛉
𝜃′ = (A.6)
𝜃

sin 𝜃 ′ 𝜃 cos 𝜃 − sin 𝜃 ′


( ) = 𝜃 (A.7)
𝜃 𝜃2

1 − cos 𝜃 ′ 𝜃 sin 𝜃 − 2(1 − cos 𝜃) ′


( ) = 𝜃 (A.8)
𝜃2 𝜃3

0 −𝜃3′ 𝜃2′
̃
𝛉 = [ 𝜃3′
′ 0 −𝜃1′ ]. (A.9)
−𝜃2′ 𝜃1′ 0

For a relatively small 𝜃, the following properties should be used for stability, i.e.,

sin 𝜃 ′ 1 − cos 𝜃 ′
lim (𝜃 ′ ) = lim [( ) ] = lim [( ) ] = 0. (A.10)
𝜃→0 𝜃→0 𝜃 𝜃→0 𝜃2

Ref. code: 25636122300053IAQ


93

APPENDIX B

YOUNG’S MODULUS AND THE SHEAR MODULUS IN THE


CURVILINEAR COORDINATE 𝐆𝒊 ⊗ 𝐆𝒋

The Saint Venant-Kirchhoff material model is used in this study. This model is
considered as an extension of the isotropic linear elastic material model to analysis of
problems with large strains. The relation between the second Piola-Kirchhoff stress tensor
𝐒 and the Green-Lagrange strain tensor 𝐄 can be expressed as

𝑆 𝑖𝑗 = 𝐶 𝑖𝑗𝑘𝑙 𝐸𝑘𝑙 (B.1)

where 𝐶 𝑖𝑗𝑘𝑙 are the material moduli expressed in the curvilinear coordinate system 𝐆𝑖 ⊗
𝐆𝑗 as

𝐶 𝑖𝑗𝑘𝑙 = 𝜆𝐺 𝑖𝑗 𝐺 𝑘𝑙 + 𝜇(𝐺 𝑖𝑘 𝐺 𝑗𝑙 + 𝐺 𝑖𝑙 𝐺 𝑗𝑘 ). (B.2)

In Eq. (B.2), 𝜆 and 𝜇 are the Lamé parameters determined as

𝜈𝐸 𝐸
𝜆= 𝜇 = 𝑘𝑠 =𝐺 (B.3)
(1 + 𝜈)(1 − 2𝜈) 2(1 + 𝜈)

where 𝐸 is Young’s modulus and 𝜈 is Poisson’s ratio. In addition, 𝐺 is the shear modulus
and 𝑘𝑠 is the shear correction factor which varies with the shape of the cross-section. For
rectangular cross-sections, 𝑘𝑠 = 5/6 can be used.

For the Timoshenko beam theory, the stress components 𝑆 22 , 𝑆 33 , and 𝑆 23 are
eliminated by static condensation (Klinkel & Govindjee, 2002). It then follows that

𝑆 11 = 𝐸̅ 𝐸11 𝑆 12 = 2𝐺̅ 𝐸12 𝑆 13 = 2𝐺̅ 𝐸13 (B.4)

where 𝐸̅ , and 𝐺̅ are defined as

𝐸 𝐺
𝐸̅ = 𝐺̅ = . (B.5)
(𝐺11 )2 𝐺11

Ref. code: 25636122300053IAQ


94

Furthermore, the following relation is assumed for relatively thick beams without causing
significant errors to the numerical results, i.e.,

𝐺11 ≈ 𝐀1 . 𝐀1 = 1. (B.6)

With the above assumption, 𝐸̅ and 𝐺̅ are constant over the cross-section. The assumption
simplifies the computation of the cross-sectional resultant forces.

Ref. code: 25636122300053IAQ


95

APPENDIX C

EXPLICIT EXPRESSIONS FOR THE VARIATIONS OF


KINEMATICS OF THE EULER-BERNOULLI BEAM
FORMULATION

The variation of the unit tangent vector 𝐭 and its derivative are given by

𝛿𝐭 = 𝐙𝛿𝐚1 𝛿𝐭 ′ = 𝐙′ 𝛿𝐚1 + 𝐙𝛿𝐚1′ (C.1)

where the matrix 𝐙 and its derivative are determined as

𝐈 − 𝐭𝐭 T 𝐭 ′ 𝐭 T + 𝐭(𝐭 ′ )T (𝐈 − 𝐭𝐭 T )(𝐚1T 𝐚1′ )


𝐙= 𝐙′ = − − .
𝑎1 𝑎1 (𝑎1 )3

(C.2)

The virtual rotation vector 𝛿𝐰 and its derivative are expressed as

𝐭 ⊗ ̅̅
𝐀̅̅1 𝐭 × 𝛿𝐚1
𝛿𝐰 = 𝐭𝛿𝜃 + (𝐈 − ) ̅̅̅̅1 , 𝐭)𝛿𝐚1
= 𝐭𝛿𝜃 + 𝐗(𝐀 (C.3)
1 + ̅̅
𝐀̅̅1 ⋅ 𝐭 𝑎1

̅̅̅̅1 , 𝐭)𝛿𝐚1 + 𝐗(𝐀


𝛿𝐰 ′ = 𝐭 ′ 𝛿𝜃 + 𝐭𝛿𝜃 ′ + 𝐗 ′ (𝐀 ̅̅̅̅1 , 𝐭)𝛿𝐚1′ (C.4)

̅̅̅̅1 , 𝐭) and its derivative are given by


where the matrix 𝐗(𝐀

̅̅̅̅1 T
𝐭𝐀 𝐭̃
̅̅̅̅1 , 𝐭) = (𝐈 −
𝐗(𝐀 ) (C.5)
T
̅̅̅̅1 𝐭 𝑎1
1+𝐀

T
T
𝐭 ′ ̅̅̅̅ ̅̅̅̅1 )
𝐀1 + 𝐭(𝐀
′ T ̅̅̅̅1 T [(𝐀
𝐭𝐀 ̅̅̅̅1 ′ ) 𝐭 + ̅̅̅̅T
𝐀1 𝐭 ′ ] 𝐭̃
̅̅̅̅1 , 𝐭) = {−
𝐗 ′ (𝐀 + }
T 2 𝑎1
1 + ̅̅
𝐀̅̅1 𝐭 (1 + ̅̅
𝐀̅̅1 𝐭)
T

(C.6)
T
̅̅̅̅1
𝐭𝐀 𝐭̃′ (𝐚1T 𝐚1′ )𝐭̃
+ (𝐈 − T
)[ − ].
1 + ̅̅
𝐀̅̅1 𝐭 𝑎1 (𝑎1 )3

Ref. code: 25636122300053IAQ


96

In the increment of 𝛿𝚪, the vector 𝛿𝐰 always appears as 𝐀T Δ𝛿𝐰 and 𝐀T Δ𝛿𝐰 ′ with
an arbitrary vector 𝐀. The lengthy derivation is skipped, and the following expressions are
provided as

𝐀T Δ𝛿𝐰 = 𝛿𝜃𝐀T 𝐙Δ𝐚1 + 𝛿𝐚1 𝐘(𝐀)Δ𝐚1 (C.7)

𝐀T Δ𝛿𝐰 ′ = 𝛿𝜃 ′ 𝐀T 𝐙Δ𝐚1 + 𝛿𝜃𝐀T 𝐙′ Δ𝐚1 + 𝛿𝜃𝐀T 𝐙Δ𝐚1′ + 𝛿𝐚1′ 𝐘(𝐀)Δ𝐚1


(C.8)
+ 𝛿𝐚1 𝐘 ′ (𝐀)Δ𝐚1 + 𝛿𝐚1 𝐘(𝐀)Δ𝐚1′

where the matrices 𝐘(𝐀) and 𝐘 ′ (𝐀) are given by

1 𝐀 ̃
̅̅̅̅1 𝐭𝐀T 𝐙 ̃
1 ̅̅
𝐀 ̅̅1 𝐭𝐀T 𝐭𝐀̅̅̅̅1 T 𝐙 𝐀 ̃𝐙 𝐀̃ 𝐭𝐚1T ̃
̅̅
1 𝐀T 𝐭𝐀 ̅̅1 𝐙
𝐘(𝐀) = − T + 2 + − − T
𝑎1 1 + ̅̅̅̅
𝐀1 𝐭 𝑎1 (1 + 𝐀 ̅̅̅̅1 T 𝐭) 𝑎1 (𝑎1 )3 𝑎1 1 + ̅̅𝐀̅̅1 𝐭
(C.9)
̃ 𝐭𝐀T 𝐭𝐚T
1 ̅̅̅̅
𝐀 1 1
+ T
(𝑎1 )3 1 + ̅̅
𝐀̅̅ 𝐭 1

′ (𝐀)
̃
̅̅
𝐚1T 𝐚1′ 𝐀 ̅̅1 𝐭𝐀T 𝐙 𝐀̃
1 ̅̅̅̅1 ′ 𝐭𝐀T 𝐙 + ̅̅̅̅
̃ 𝐭 ′ 𝐀T 𝐙 + ̅̅
𝐀 1
̃
𝐀 ̅̅1 𝐭𝐀T 𝐙′
𝐘 ={ T − T
(𝑎1 )3 1 + ̅̅ 𝐀̅̅ 𝐭 𝑎1 1 + ̅̅
𝐀̅̅ 𝐭
1 1

T
1 ̅̅̅̅
𝐀 1
̅̅̅̅1 ′ ) 𝐭 + ̅̅̅̅
̃ 𝐭𝐀T 𝐙 [(𝐀 T
𝐀1 𝐭 ′ ]
+ }
𝑎1 T 2
̅̅
(1 + 𝐀1 𝐭) ̅̅
(C.10)
̃ T
𝐚1T 𝐚1′ ̅̅
𝐀 ̅̅1 𝐭𝐀T 𝐭𝐀
̅̅̅̅1 𝐙
+ {−
(𝑎1 )3 T 2
(1 + ̅̅̅̅
𝐀1 𝐭)

𝐀̃1 𝐭𝐀T 𝐭𝐀

1 ̅̅̅̅ ̅̅̅̅1 T 𝐙 + ̅̅
̃
𝐀 ̅̅1 𝐭 ′ 𝐀T 𝐭𝐀̅̅̅̅1 T 𝐙 + ̅̅
̃
𝐀 ̅̅1 𝐭𝐀T 𝐭 ′ ̅̅̅̅
𝐀1 𝐙
T
+
𝑎1 T 2
(1 + ̅̅ 𝐀̅̅1 𝐭)

Ref. code: 25636122300053IAQ


97

T
̃
1 ̅̅
𝐀 ̅̅̅̅1 ′ ) 𝐙 + ̅̅
̅̅1 𝐭𝐀T 𝐭(𝐀 ̃
𝐀 ̅̅̅̅1 T 𝐙′
̅̅1 𝐭𝐀T 𝐭𝐀
+
𝑎1 T 2
(1 + ̅̅ 𝐀̅̅1 𝐭)
T
̃
̅̅
1 2𝐀 ̅̅̅̅1 T 𝐙 [(𝐀
̅̅1 𝐭𝐀T 𝐭𝐀 ̅̅̅̅1 ′ ) 𝐭 + ̅̅̅̅ T
𝐀1 𝐭 ′ ]
− }
𝑎1 T 3
̅̅
(1 + 𝐀1 𝐭) ̅̅

̃𝐙 𝐀
𝐚1T 𝐚1′ 𝐀 ̃𝐙′ ̃ 𝐭𝐚1T 𝐀
3𝐚1T 𝐚1′ 𝐀 ̃ 𝐭 ′ 𝐚1T + 𝐀
̃ 𝐭(𝐚1′ )T
+ {− + }+{ − }
(𝑎1 )3 𝑎1 (𝑎1)5 (𝑎1)3
̃𝐙
̅̅̅̅
𝐚1T 𝐚1′ 𝐀T 𝐭𝐀 ̃
1 𝐀T 𝐭 ′ ̅̅
𝐀 ̅̅1 𝐙
1
+{ T − T
(𝑎1 )3 1 + ̅̅
𝐀̅̅ 𝐭 𝑎1 1 + ̅̅ 𝐀̅̅ 𝐭
1 1

T
̅̅̃
1 𝐀T 𝐭𝐀 ̅̅1 ′ 𝐙 + 𝐀T 𝐭𝐀 ̃
̅̅
̃ 𝐙′ 1 𝐀T 𝐭𝐀
̅̅̅̅ ̅̅̅̅1′ ) 𝐭 + ̅̅
̅̅1 𝐙 [(𝐀 T
𝐀̅̅1 𝐭 ′ ]
1
− + }
𝑎1 T 𝑎1 T 2
1 + ̅̅𝐀̅̅1 𝐭 ̅̅
(1 + 𝐀 𝐭) ̅̅ 1

̃
𝐚1T 𝐚1′ ̅̅
𝐀 ̅̅1 𝐭𝐀T 𝐭𝐚1T
+ {− T
(𝑎1 )5 1 + ̅̅ 𝐀̅̅ 𝐭
1

𝐀̃1 𝐭𝐀T 𝐭𝐚1T + ̅̅


′ ̃ ̃ ̃
1 ̅̅̅̅ 𝐀 ̅̅1 𝐭 ′ 𝐀T 𝐭𝐚1T + ̅̅
𝐀 ̅̅1 𝐭𝐀T 𝐭 ′ 𝐚1T + ̅̅
𝐀 ̅̅1 𝐭𝐀T 𝐭(𝐚1′ )T
+ T
(𝑎1 )3 1 + ̅̅
𝐀̅̅1 𝐭
T
𝐀 1 1
̅̅̅̅1 ′ ) 𝐭 + ̅̅
̃ 𝐭𝐀T 𝐭𝐚T [(𝐀
1 ̅̅̅̅
T
𝐀̅̅1 𝐭 ′ ]
− }
(𝑎1 )3 T 2
̅̅
(1 + 𝐀1 𝐭)̅̅

Ref. code: 25636122300053IAQ


98

APPENDIX D

INCREMENT OF THE VARIATION OF THE GENERALIZED


STRAIN VECTOR 𝚪 OF THE EULER-BERNOULLI BEAM
FORMULATION

Before expressing the increment of the variation of 𝚪, two matrices are introduced
as

𝐐(𝐀, 𝑖, 𝑗) = −𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)[𝐗 T (𝐀 ̃′ 𝐀


̅̅̅̅1 , 𝐭)𝐚 ̃ ̅̅̅̅ ̃ ′
1 𝐗(𝐀1 , 𝐭) + 𝐘(𝐀𝐚1 )]
(D.1)
̅̅̅̅1 , 𝐭)𝐀
− 𝑅𝑖′ (𝜂)𝑅𝑗′′ (𝜂)𝐗 T (𝐀 ̃ + 𝑅𝑖′′ (𝜂)𝑅𝑗′ (𝜂)𝐀
̃ 𝐗(𝐀
̅̅̅̅1 , 𝐭)

𝐖(𝐀, 𝑖, 𝑗) = −𝑅𝑖 (𝜂)𝑅𝑗′′ (𝜂)𝐭 T 𝐀 ̃′ 𝐀


̃ − 𝑅𝑖 (𝜂)𝑅𝑗′ (𝜂)[𝐭 T 𝐚 ̃ ̅̅̅̅ ′ T̃
1 𝐗(𝐀1 , 𝐭) − (𝐚1 ) 𝐀𝐙] (D.2)

where 𝐀 ∈ ℝ3 is an arbitrary vector.

The increments of the variations of the components of 𝚪 are compactly expressed


as

𝑛 𝑛
𝑖𝑗
Δ𝛿𝛤11 = 𝛿𝐯 𝐋11 Δ𝐯 = ∑ ∑ 𝛿𝐯𝑖T 𝐋11 Δ𝐯𝑗
T
(D.3)
𝑖=1 𝑗=1

𝑛 𝑛
𝑖𝑗
Δ𝛿𝐾12 = 𝛿𝐯 𝐋12 Δ𝐯 = ∑ ∑ 𝛿𝐯𝑖T 𝐋12 Δ𝐯𝑗
T
(D.4)
𝑖=1 𝑗=1

𝑛 𝑛
𝑖𝑗
Δ𝛿𝐾13 = 𝛿𝐯 𝐋13 Δ𝐯 = ∑ ∑ 𝛿𝐯𝑖T 𝐋13 Δ𝐯𝑗
T
(D.5)
𝑖=1 𝑗=1

𝑛 𝑛
𝑖𝑗
Δ𝛿𝐾23 = 𝛿𝐯 𝐋23 Δ𝐯 = ∑ ∑ 𝛿𝐯𝑖T 𝐋23 Δ𝐯𝑗
T
(D.6)
𝑖=1 𝑗=1

𝑖𝑗 𝑖𝑗 𝑖𝑗 𝑖𝑗
where the matrices 𝐋11, 𝐋12, 𝐋13, and 𝐋23 are defined as

Ref. code: 25636122300053IAQ


99

𝑖𝑗 𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂)𝐈 𝟎


𝐋11 = [ ] (D.7)
𝟎T 0

−𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)𝐗 T (𝐀 ̃′ 𝐚


̅̅̅̅1 , 𝐭)𝐚 ′′
2 + 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)𝐚
𝑖𝑗 𝐐(𝐚2 , 𝑖, 𝑗) 1 ̃𝐭 ̃𝐭2
𝐋12 =[ ] (D.8)
𝐖(𝐚2 , 𝑖, 𝑗) −𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)𝐭 𝐚1 𝐚 T ̃′
̃𝐭
2

−𝑅𝑖′ (𝜂)𝑅𝑗 (𝜂)𝐗 T (𝐀 ̃′ 𝐚


̅̅̅̅1 , 𝐭)𝐚 ′′
3 + 𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)𝐚
𝑖𝑗 𝐐(𝐚3 , 𝑖, 𝑗) 1 ̃𝐭 ̃𝐭3
𝐋13 =[ ] (D.9)
𝐖(𝐚3 , 𝑖, 𝑗) −𝑅𝑖 (𝜂)𝑅𝑗 (𝜂)𝐭 𝐚1 𝐚 T ̃′
̃𝐭
3

𝑖𝑗
𝐋23
̅̅̅̅1, 𝐭)]T 𝐙 + 𝐘 ′ (𝐭)} − 𝑅𝑖′′ (𝜂)𝑅𝑗′(𝜂)[𝐗T (𝐀
−𝑅𝑖′ (𝜂)𝑅𝑗′ (𝜂){[𝐗 ′ (𝐀 ̅̅̅̅1 , 𝐭)𝐙 + 𝐘(𝐭)] − 𝑅𝑖′ (𝜂)𝑅𝑗′′ (𝜂)𝐘(𝐭) 𝟎
=[ ]
−𝑅𝑖 (𝜉)𝑅𝑗′ (𝜉)[(𝐭′ )T 𝐙 + 𝐭T 𝐙 ′ ] − 2𝑅𝑖′ (𝜂)𝑅𝑗′(𝜂)𝐭T 𝐙 − 𝑅𝑖 (𝜂)𝑅𝑗′′ (𝜂)𝐭T 𝐙 0

(D.10)

Ref. code: 25636122300053IAQ


100

BIOGRAPHY

Name Mr. Duy Vo

Date of Birth May 08, 1992

Education 2015: Bachelor of Engineering (Civil


Engineering), Faculty of Civil
Engineering, Ho Chi Minh City University
of Technology, Vietnam National
University

2017: Master of Science (Engineering and


Technology), School of Civil Engineering
and Technology, Sirindhorn International
Institute of Technology, Thammasat
University

Publications

International journals

Vo, D., & Nanakorn, P. (2020). Geometrically nonlinear multi-patch isogeometric analysis
of planar curved Euler–Bernoulli beams. Computer Methods in Applied
Mechanics and Engineering, 366, 113078.
https://doi.org/10.1016/j.cma.2020.113078.

Vo, D., & Nanakorn, P. (2020). A total Lagrangian Timoshenko beam formulation for
geometrically nonlinear isogeometric analysis of planar curved beams. Acta
Mechanica, 231(7), 2827-2847. https://doi.org/10.1007/s00707-020-02675-x.

Vo, D., Nanakorn, P., & Bui, T. Q. (2020). A total Lagrangian Timoshenko beam
formulation for geometrically nonlinear isogeometric analysis of spatial beam
structures. Acta Mechanica, 231(9), 3673-3701. https://doi.org/10.1007/s00707-
020-02723-6.

Ref. code: 25636122300053IAQ


101

Vo, D., Borković, A., Nanakorn, P., & Bui, T. Q. (2020). Dynamic multi-patch
isogeometric analysis of planar Euler-Bernoulli beams. Computer Methods in
Applied Mechanics and Engineering, 372, 113435,
https://doi.org/10.1016/j.cma.2020.113435

Suttakul, P., Nanakorn, P., & Vo, D. (2019). Effective out-of-plane rigidities of 2D lattices
with different unit cell topologies. Archive of Applied Mechanics, 89(9), 1837-
1860. https://doi.org/10.1007/s00419-019-01547-8.

International conferences

Duong, N. H., Vo, D., Nanakorn, P. (2019). Design of two-dimensional trusses by an


improved firefly algorithm. Proceeding of the 12th Pacific Structural Conference
(PSSC2019). Tokyo, Japan.

Vo, D., Nanakorn, P. (2017) . Large displacement analysis of 2D beams using field-
consistent beam elements with a higher-order axial-displacement interpolation.
Proceeding of 15th East Asia-Pacific Conference on Structural Engineering and
Construction (EASEC-15) (pp.1672-1678). Xi’an, P.R. China.

Book chapters

Vo, D., Nanakorn, P., & Bui, T. Q. (2020). Multi-patch geometrically nonlinear
isogeometric analysis of spatial beams with additive rotation updates. Reddy, J.
N., Wang, C. M., Luong, V. H., & Le, A. T., Proceeding of ICSCEA 2019 (pp.
1129-1136). Singapore: Springer Singapore.

Ref. code: 25636122300053IAQ

You might also like