You are on page 1of 17

Thin-Walled Structures 170 (2022) 108539

Contents lists available at ScienceDirect

Thin-Walled Structures
journal homepage: www.elsevier.com/locate/tws

Full length article

Geometrically exact static isogeometric analysis of arbitrarily curved plane


Bernoulli–Euler beam
A. Borković a,b ,∗, B. Marussig a , G. Radenković b,c
a
Institute of Applied Mechanics, Graz University of Technology, Technikerstraße 4/II, 8010 Graz, Austria
b
University of Banja Luka, Faculty of Architecture, Civil Engineering and Geodesy, Department of Mechanics and Theory of Structures, 78000 Banja Luka, Bosnia
and Herzegovina
c
Faculty of Civil Engineering, University of Belgrade, Bulevar kralja Aleksandra 73, 11000 Belgrade, Serbia

ARTICLE INFO ABSTRACT


Keywords: We present a geometrically exact nonlinear analysis of elastic in-plane beams in the context of finite but
Bernoulli–Euler beam small strain theory. The formulation utilizes the full beam metric and obtains the complete analytic elastic
Strongly curved beams constitutive model by employing the exact relation between the reference and equidistant strains. Thus, we
Geometrically exact analysis
account for the nonlinear strain distribution over the thickness of a beam. In addition to the full analytical
Analytical constitutive relation
constitutive model, four simplified ones are presented. Their comparison provides a thorough examination of
the influence of a beam’s metric on the structural response. As a benchmark result, an analytical solution for
a pure bending of a strongly curved cantilever beam is derived. We show that the appropriate formulation
depends on the curviness of a beam at all configurations. Furthermore, the nonlinear distribution of strain
along the thickness of strongly curved beams must be considered to obtain a complete and accurate response.

1. Introduction Several works consider nonlinear shear-deformable spatial beams


in the framework of IGA, [14–20]. In the context of the BE theory, a
Curved beams are fundamental structural components and their torsion- and rotation-free spatial cable formulation is developed in [21]
analysis is a classic topic of mechanics [1]. Although some analytical with attractive nonlinear dynamic applications. A spatial BE beam
solutions exist, see, e.g., [2], numerical methods are inevitable for deal- modeled as a ribbon with four degrees of freedom (DOF) is introduced
ing with the complex nonlinear behavior of beams. The finite element in [9], while a consistent tangent operator is derived in [22]. Based
method (FEM) is the most versatile numerical procedure for solving on [9], the authors in [23] discuss nonlinear applications of a spatial
partial differential equations. The first research on the geometrically- curved BE beam. Planar nonlinear BE beams are analyzed in [24–26].
exact FEM analysis of curved spatial beams was presented in a seminal Different approaches to multi-patch modeling are considered in [24,
paper by Reissner [3]. This theory has been enhanced in a series 25], while the paper [26] focuses on the weighted residuals and the
of papers [4–8]. In [4], Simo introduced the term geometrically exact
collocation of strong form.
beam theory to designate that the governing equations of motion are
When the beam is curved, the strain distribution over the cross
valid regardless of the magnitude of kinematic quantities. The papers
section becomes nonlinear. This leads to the coupling of axial and
referred to are mainly focused on the shear-deformable beam model,
bending actions, which is readily ignored in the mentioned references
and it is not until recently that the Bernoulli–Euler (BE) spatial model
in which simple decoupled constitutive relations are utilized. Nonlinear
was scrutinized [9–11]. One reason for this is that the equations of
distribution of strain is evident from the initial curvature correction term
the Simo–Reissner beam require only 𝐶 0 interelement continuity. On
the other hand, the BE model requires 𝐶 1 interelement continuity — a 𝑔0 = 1 − 𝜂𝐾, see [27], which relates lengths of the basis vectors at
requirement not fulfilled by standard Lagrange polynomials commonly the centroid and an arbitrary point of cross section to each other.
used in FEM. Herein, we resolve this issue by using isogeometric This effect becomes more prominent as the curvature of the beam axis
analysis (IGA), [12,13]. This approach utilizes (smooth) splines as basis (𝐾) and height of the cross section (ℎ) increase. The product of these
functions for the spatial discretization of the weak form of the equations two quantities 𝐾ℎ is a parameter referred to as the curviness of beam
of motion, and hence, it allows for an arbitrarily high continuity which changes along the beam length [28]. In this paper, we perform
between elements. a nonlinear analysis of arbitrarily curved uniform BE beams. The term

∗ Corresponding author at: University of Banja Luka, Faculty of Architecture, Civil Engineering and Geodesy, Department of Mechanics and Theory of Structures,
78000 Banja Luka, Bosnia and Herzegovina.
E-mail addresses: aborkovic@tugraz.at, aleksandar.borkovic@aggf.unibl.org (A. Borković).

https://doi.org/10.1016/j.tws.2021.108539
Received 9 April 2021; Received in revised form 6 August 2021; Accepted 7 October 2021
Available online 25 October 2021
0263-8231/© 2021 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

arbitrarily curved does not refer just to the variable curvature of the 2.1. Metric of the beam axis
beam axis, but more importantly, to its curviness 𝐾ℎ. We consider
geometries with large curviness such as 𝐾ℎ = 0.5. In [29], curved Let us revise some basic expressions of the metric of beam axis.
beams are grouped into small-, medium- and big-curvature beams,
A beam axis is a curve that passes through the centroids of all cross
depending on their curviness. A beam is said to have a small curvature
sections. It can be defined with either the arc-length coordinate 𝑠 or
if its curviness is infinitesimal, 𝐾ℎ ≪ 1, while it has a medium
some parametric coordinate 𝜉. The position vector of the beam axis in
curvature if approximately, (𝐾ℎ)2 ≪ 1. All the others belong to the
Cartesian coordinates is 𝐫 = {𝑥 = 𝑥1 , 𝑦 = 𝑥2 } and it is here defined as a
category of big-curvature, also known as strongly curved beams. In the
linear combination:
following, we will label a beam strongly curved when 𝐾ℎ > 0.1.
Although the basic theory of strongly curved beams has been known ∑
𝑁 ∑
𝑁

for a long time, see e.g. [30], it is only recently that modern numerical 𝐫 = 𝑥 𝛼 𝐢𝛼 = 𝑅𝐼 (𝜉)𝐫𝐼 , 𝑥𝛼 = 𝑅𝐼 (𝜉)𝑥𝛼𝐼 , (1)
𝐼=1 𝐼=1
techniques have been applied for the analysis of these beams. Linear
analysis of plane beams is given in [28,30,31] within the framework where 𝑅𝐼 are univariate NURBS basis functions, 𝐫𝐼 = {𝑥𝐼 , 𝑦𝐼 } are the
of IGA, while the spatial beams are analyzed in [32]. To the best of position vectors of the control points 𝐼, and 𝑁 is the total number of
our knowledge, this is the first paper that deals with the nonlinear control points [12]. Furthermore, 𝐢𝛼 = 𝐢𝛼 are the base vectors of the
analysis of strongly curved beams. The paper is based on our previous Cartesian coordinate system, see Fig. 1. The tangent base vector of a
works [28,31–33]. The exact metric of the plane BE beam is utilized beam axis is:
for the derivation of the weak form of equilibrium which is solved by
𝐠1 = 𝐫,1 = 𝑥𝛼,1 𝐢𝛼 . (2)
Newton–Raphson and arc-length methods. The geometric stiffness ma-
trix is derived by the variation of strain and external load with respect The other base vector, 𝐠2 , is perpendicular to 𝐠1 and has unit length.
to the metric. The strict derivation of constitutive relation allows us to Hence, we can choose the normal of the beam axis, as in [28], or we
derive reduced models and to compare them through numerical experi- can rotate 𝐠1 by 90 degrees and scale it. The latter approach is utilized
ments. Comparison with existing results demonstrates that the obtained
in present formulation and the base vector 𝐠1 is rotated anti-clockwise:
formulation is reliable for the finite rotation analysis of arbitrarily
curved beams. Moreover, the approach introduces an additional level
𝑥2,1 𝑥1,1 [ ]
of accuracy when dealing with strongly curved beams. Regarding the 1 0 −1
𝐠2 = √ 𝛬𝐠1 = − √ 𝐢1 + √ 𝐢2 = 𝑥𝛼,2 𝐢𝛼 , 𝛬= . (3)
benchmark results, the analytical solutions for the nonlinear behavior 𝑔 𝑔 𝑔 1 0
of strongly curved beams are rare in literature. Therefore, we derived
the analytical solution for the nonlinear problem of pure bending of Here, 𝑔 is a component of the metric tensor and also its determinant:
a strongly curved beam and compared the results with the numerical [ ] ( )
𝑔 0
ones. The present formulation is geometrically exact in the sense that it 𝑔𝛼𝛽 = 11 , 𝑔11 = 𝐠1 ⋅ 𝐠1 = det 𝑔𝛼𝛽 = 𝑔 . (4)
0 1
strictly defines a relation between work conjugate pairs, which allows
analysis of arbitrarily large rotations and displacements [4]. Since both This quantity relates differentials of arc-length and parametric coordi-

geometry and displacements are interpolated with the same spline nate as d𝑠 = 𝑔 d𝜉. Therefore, it equals the square of Jacobian of the
functions, the formulation exactly describes rigid-body modes and thus coordinate transformation from convective to arc-length coordinate.
it is frame invariant [34,35]. The metric of the beam axis is completely defined by the introduc-
The paper is structured as follows: The next section presents the tion of the Christoffel symbols. They follow from the differentiation of
fundamental relations of the beam metric. Then, the description of the base vectors with respect to the curvilinear coordinates [38]:
BE beam kinematics is provided. The finite element formulation is given
in Section 4 and numerical examples are presented in Section 5. The 𝐠𝛼,𝛽 = 𝑥𝛾,𝛼𝛽 𝐢𝛾 = 𝛤𝛼𝛽
𝛾 𝛾
𝐠𝛾 ⇒ 𝛤𝛼𝛽 = 𝑥𝛿,𝛼𝛽 𝑥,𝛾
𝛿
, (5)
conclusions are delivered in the last section. 𝛾
where 𝛤𝛼𝛽 are the Christoffel symbols of the second kind. The reciprocal
base vectors of the beam axis are:
2. Metric of the beam continuum
1
𝐠1 = 𝑥𝛼,1 𝐢𝛼 = 𝐠 and 𝐠2 = 𝑥𝛼,2 𝐢𝛼 = 𝐠2 , (6)
𝑔 1
A detailed and rigorous definition of the beam metric is presented
in this section. Following the classical BE assumption, a cross section and the reciprocal metric tensor is:
is rigid and remains perpendicular to the beam axis in the deformed [ 11 ]
𝑔 0 ( ) 1
configuration. This assumption leads to a degeneration of a 3D contin- 𝑔 𝛼𝛽 = , 𝑔 11 = 𝐠1 ⋅ 𝐠1 = det 𝑔 𝛼𝛽 = . (7)
0 1 𝑔
uum beam model into an arbitrarily shaped line. The present analysis is
performed with respect to the convective frame of reference while the The expression for the derivatives of base vectors in matrix form is:
complete beam kinematics is defined by the translation of the beam [ ] [ ][ ] [ 1 ][ ]
𝐠1,1 𝛤1 2
𝛤11 𝐠1 𝛤 𝐾̃ 𝐠1
axis. = 11 1 2 = 11 , (8)
𝐠2,1 𝛤21 𝛤21 𝐠2 −𝐾 0 𝐠2
In the notation, lowercase and uppercase boldface letters are used
for vectors and tensors or matrices, respectively. An overbar designates where 𝐾 is so-called signed curvature:
quantities at the equidistant line of the beam, and the asterisk sign
denotes the deformed configuration. Finally, the hat symbol specifies 𝑥1,1 𝑥2,11 − 𝑥2,1 𝑥1,11
1
𝐾 = −𝛤21 = −𝐠2,1 ⋅ 𝐠1 = , (9)
the local component of a vector, with respect to the curvilinear co- 𝑔 3∕2
ordinates. The direct and index notations are applied simultaneously,
depending on the context, and the standard summation convention and 𝐾̃ = 𝑔𝐾 is the signed curvature of beam axis with respect to the
is adopted. Greek index letters take values of 1 and 2. Partial and convective coordinate frame.
covariant derivatives with respect to the convective coordinates are It should be noted that, since we are dealing with both the signed
designated with (),𝛼 and ()|𝛼 , respectively. curvature and the modulus of curvature, the curviness 𝐾ℎ itself can be
The elaboration on the NURBS-based IGA modeling of curves is defined with either of them. In the remainder of the paper, 𝐾ℎ usually
excluded for brevity since it is readily available in the literature. refers to the curviness obtained with the modulus of curvature. It is
For a detailed discussion on IGA and NURBS, Refs. [12,36,37] are only in Section 5.3 that the so-called signed curviness is used, merely
recommended. for graphical representation.

2
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 1. Degeneration of a (planar) 3D continuum into an in-plane beam. Base vectors at the centroid and at an equidistant line are depicted.

2.2. Metric of an equidistant line where 𝐮 is the displacement vector of the beam axis. It is discretized
with NURBS, in the same way as the geometry, Eq. (1):
The degeneration from a 3D to a 2D beam model is trivial since ∑
𝑁
we are dealing with plane beams, and all quantities are constant along 𝐮= 𝑅𝐼 (𝜉)𝐮𝐼 . (15)
the 𝜁 coordinate, as illustrated in Fig. 1. In order to reduce the beam 𝐼=1

continuum from 2D to 1D, the metric of the complete beam must be 𝐮𝐼 is the vector of displacement components of a control point 𝐼 with
defined by some reference quantities. It is common to use the metric of respect to the Cartesian system. For the sake of a seamless transition
the beam axis as the reference. Therefore, let us define an equidistant to the discrete equation of motion which follows in Section 4, let us
line which is a set of points with 𝜂 = 𝑐𝑜𝑛𝑠𝑡. Its position and tangent base introduce the matrix of basis functions 𝐍 such that Eq. (15) can be
vectors are: written as:
𝐫̄ = 𝐫 + 𝜂𝐠2 , 𝐮 = 𝐍𝐪, (16)
(10)
𝐠̄ 1 = 𝐫̄ ,1 = 𝐠1 − 𝜂𝐾𝐠1 = 𝑔0 𝐠1 , with 𝑔0 = 1 − 𝜂𝐾.
where:
[ ] [ ]
Evidently, the base vector of this line is parallel to the base vector of 𝐪𝖳 = 𝐮𝖳1 𝐮𝖳2 ... 𝐮𝖳𝐼 ... 𝐮𝖳𝑁 , 𝐮𝖳𝐼 = 𝑢1𝐼 𝑢2𝐼 ,
beam axis. Their lengths differ for the initial curvature correction term [ ] [ ] (17)
𝐍𝖳 = 𝐑𝖳 𝐑𝖳 , 𝐑 = 𝑅1 𝑅2 ... 𝑅𝐼 ... 𝑅𝑁 .
𝑔0 [27]. The other base vector of the equidistant line is the same as the
one of beam axis 𝐠̄ 2 = 𝐠2 . The metric tensor of an equidistant line is: The material time derivative of the displacement field yields the veloc-
[ 2 ] ity field, i.e.:
𝑔 𝑔 0 ( )
𝑔̄𝛼𝛽 = 0 , det 𝑔̄𝛼𝛽 = 𝑔02 𝑔 = 𝑔̄ . (11) 𝐯 = 𝐫̇ ∗ = 𝐮̇ = 𝑢̇ 𝛼 𝐢𝛼 = 𝑣𝛼 𝐢𝛼 = 𝑥̇ ∗𝛼 𝐢𝛼 = 𝑢̂̇ 𝛼 𝐠∗𝛼 + 𝑢̂ 𝛼 𝐠̇ ∗𝛼 = 𝑣̂ 𝛼 𝐠∗𝛼 + 𝑢̂ 𝛼 𝐠̇ ∗𝛼 , (18)
0 1
Note that the initial curvature correction term 𝑔0 becomes zero for where 𝑢̂ 𝛼 and 𝑣̂ 𝛼 are the components of the displacement and velocity
[ ]
𝜂𝐾 = 1 which is physically prohibited. Since 𝜂 ∈ −ℎ∕2, ℎ∕2 , it follows with respect to the local curvilinear coordinates, respectively. Since the
that the upper bound of curviness is 2. convective coordinate deforms with the body, we must differentiate
both the local displacement components and the base vectors with
respect to time. However, it is more convenient to work with the global
3. Bernoulli–Euler beam theory
components of displacement and velocity since the base vectors of
the Cartesian coordinate frame are time-invariant. In that case, the
After the metric of the beam continuum is defined, the next step is components of velocity are simple partial derivatives of the global
to introduce a strain measure. For the convective coordinate frame, the displacement components with respect to time. Spatially discretized
Lagrange strain equals the difference between the current and reference form of the velocity follows from Eqs. (17) and (18): 𝐯 = 𝐮̇ = 𝐍𝐪. ̇
metrics: The work conjugate pair for the Cauchy stress tensor is the strain
( )
1 ∗ rate tensor. It is the symmetric part of the velocity gradient and its
𝜖𝛼𝛽 = 𝑔 − 𝑔𝛼𝛽 . (12)
2 𝛼𝛽 components equal the material time derivative of (12). The covariant
Due to the BE hypothesis, the shear strain vanishes and the position components of the strain rate tensor, with respect to the Cartesian
of the cross section is completely determined by the components of coordinates, can be written as:
translation of the beam axis. In other words, these components are the ( ) ( ) ( )
1 1 ∗𝛿 1 ∗
only generalized coordinates of the BE beam. This fact gives rise to the 𝑑𝛼𝛽 = 𝜖̇ 𝛼𝛽 = 𝑣̂ 𝛼|𝛽 + 𝑣̂ 𝛽|𝛼 = 𝑥,𝛼 𝑣𝛿,𝛽 + 𝑥∗𝛿
,𝛽 𝑣𝛿,𝛼 = 𝐠𝛼 · 𝐯,𝛽 + 𝐠∗𝛽 · 𝐯,𝛼 .
2 2 2
so-called rotation-free beam theories [28]. (19)

For BE beams, the only non-zero component of the strain rate is 𝑑11 :
3.1. Kinematics
𝑑11 = 𝑥∗𝛼 𝑣 = 𝐠∗1 · 𝐯,1 .
,1 𝛼,1
(20)
In the deformed configuration, the position vector of an equidistant
In order to define the kinematics of a beam continuum, the velocity
line is:
of an equidistant line must be introduced. It is the material derivative
𝐫̄ ∗ = 𝐫 ∗ (𝜂) = 𝐫 ∗ + 𝜂𝐠∗2 , (13) of (13):

where 𝐫 ∗ denotes the position vector of the deformed beam axis and the 𝐫̄̇ ∗ = 𝐯̄ = 𝐯 + 𝜂𝐯,2 , (21)
base vector 𝐠∗2 is calculated similarly as in Eq. (3). The position vector
where 𝐯,2 follows from the definition of shear strain:
of the deformed beam axis is:
( ) ( )
1 ∗ 1
𝐫 ∗ = 𝐫 + 𝐮, (14) 𝑑12 = 𝐠1 ⋅ 𝐯,2 + 𝐠∗2 ⋅ 𝐯,1 = 0 ⇒ 𝐯,2 = − ∗ 𝐠∗2 ⋅ 𝐯,1 𝐠∗1
2 𝑔

3
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

1 ( )
= − ∗ 𝐠∗1 ⊗ 𝐠∗2 𝐯,1 . (22) surface loads are reduced to the beam axis, it can be written for plane
𝑔
BE beams as:
For the purpose of further derivation, let us express the components of √ √
this gradient by index notation, [28]: 𝛿𝑃 = 𝜎̄ 11 𝛿 𝑑̄11 d𝑉 − 𝑝̂𝛼 𝛿 𝑣̂ 𝛼 𝑔 d𝜉 = 𝝈 ∶ 𝛿𝐝 d𝑉 − 𝐩⋅𝛿𝐯 𝑔 d𝜉 = 0,
∫𝑉 ∫𝜉 ∫𝑉 ∫𝜉
1 ∗𝛽 ∗
𝑣𝛼,2 = −𝐵𝛼∗𝛽 𝑣𝛽,1 , with 𝐵𝛼∗𝛽 = 𝑥 𝑥 . (23) (32)
𝑔 ∗ ,2 𝛼,1
where 𝝈 is the Cauchy stress tensor, 𝐝 is the strain rate tensor, and
3.2. Strain rate at an equidistant line 𝐩 is the vector of external line loads. All these quantities are measured
with respect to the current, unknown, configuration. For problems with
The strain rate at an equidistant line is, analogously to Eq. (20): deformation-independent loads, the only requirement is to linearize
the Cauchy stress. At the current (𝑛 + 1) configuration this stress is
𝑑̄11 = 𝐠̄ ∗1 ⋅ 𝐯̄ ,1 . (24) approximated by:
For the evaluation of this expression, let us first define the material d𝑡 (𝑛) 𝝈
(𝑛+1)
𝝈 ≈ (𝑛) 𝝈+ 𝛥𝑡(𝑛+1) , (33)
derivative of mixed velocity gradient, using Eq. (8): (𝑛+1) d𝑡
(𝑛)
𝐯,21 = 𝐠̇ ∗2,1 = −𝐾̇ ∗ 𝐠∗1 − 𝐾 ∗ 𝐯,1 . (25) where is the stress from the previous (𝑛) configuration expressed
(𝑛+1)
𝝈
with respect to the metric of the current (𝑛 + 1) configuration.
With this relation and Eqs. (10), (21) and (24), the equidistant strain
rate reduces to: Remark. The additive decomposition of stress, as in Eq. (33), is only
( ) ( ) ( ) ( )
valid if all of the terms are expressed with respect to the same metric.
𝑑̄11 = 𝑔0∗ 𝐠∗1 ⋅ 𝐯,1 + 𝜂𝐯,21 = 𝑔0∗ 𝐠∗1 ⋅ 𝐯,1 − 𝜂 𝐾̇ ∗ 𝐠̄ ∗1 − 𝜂𝐾 ∗ 𝐯,1
( ) (26) Therefore, the stress from the previous configuration is here adjusted
= 𝑔0∗ 𝑔0∗ 𝑑11 − 𝜂 𝐾̇ ∗ 𝑔 ∗ . to the current metric by an adjustment term. This term is approximately
At this point, we will introduce a curvature change of beam axis with equal to the ratio of determinants of the metric tensors in two configu-
respect to the convective coordinate: rations, c.f. Appendix A. The exact integration of the adjusted stress
is impractical because the adjustment term consists of the ratio of
𝜅 = 𝐾̃ ∗ − 𝐾̃ = 𝑔 ∗ 𝐾 ∗ − 𝑔𝐾, (27) the initial curvature correction terms at two configurations. For linear
elastic material adopted here, this is not a problem, since the stress
and its material derivative, the rate of curvature change:
can be calculated from the total strain. However, for the nonlinear
𝜅̇ = 𝑔̇ ∗ 𝐾 ∗ + 𝑔 ∗ 𝐾̇ ∗ = 2𝑑11 𝐾 ∗ + 𝑔 ∗ 𝐾̇ ∗ ⇒ 𝑔 ∗ 𝐾̇ ∗ = 𝜅̇ − 2𝑑11 𝐾 ∗ . (28) material behavior, it is necessary to use an incremental stress-updating
approach. For the future considerations, two incremental approaches
By inserting the last expression into Eq. (26), the final form of the are considered here. They are discussed in detail in Appendix A and
equidistant strain rate is obtained: compared in Section 5.1.4.
( ) [( ) ]
𝑑̄11 = 𝑔0∗ 𝑔0∗ 𝑑11 − 𝜂 𝜅̇ + 2𝜂𝑑11 𝐾 ∗ = 𝑔0∗ 1 + 𝜂𝐾 ∗ 𝑑11 − 𝜂 𝜅̇ . (29) An appropriate time derivative in Eq. (33) is designated with d𝑡 ()∕ d 𝑡
while 𝛥𝑡(𝑛+1) = 𝑡(𝑛+1) − 𝑡(𝑛) is the time increment. The rate form
This expression is analogous to the one derived in [28] for the equidis- of stress–strain relations requires an objective time derivative. Note
tant axial strain. It is valid for finite, but small, strain analysis, which that the material derivative is not objective, while the corotational
falls into the scope of the geometrical exact beam theory [6]. (Jaumann) and convective time derivatives fulfill this property [40,41].
Finally, representing the rate of curvature change as a function of The convective derivative of stress tensor results in the stress rate
velocity gradients of the beam axis is a requirement. It follows from tensor. An important fact is that the components of the stress rate tensor
Eqs. (8), (22), and (27): are equal to the material derivatives of the components of the stress
( )
𝜅̇ = 𝐾̃̇ ∗ = 𝛤̇ 11
∗2
= 𝐠∗1,1 ⋅ 𝐯,2 + 𝐠∗2 ⋅ 𝐯,11 = 𝐠∗2 ⋅ 𝐯,11 − 𝛤11
∗1
𝐯,1 . (30) tensor, [42]. Having this in mind, and after the insertion of Eq. (33)
into Eq. (32), the linearized form of the principle of virtual power at
the current configuration is obtained:
4. Finite element formulation

𝜎̄̇ 11 𝛿 𝑑̄11 d𝑉 𝛥𝑡 + 𝜎̄ 11 𝛿 𝑑̄11 d𝑉 = 𝑝̂𝛼 𝛿 𝑣̂ 𝛼 𝑔 d𝜉 . (34)
∫𝑉 ∫𝑉 ∫𝜉
In line with the previous derivation, we will formulate the isogeo-
metric BE element using the principle of virtual power. The generalized In the remainder of this paper, we neglect the time indices and aster-
coordinates are the components of the velocities of the control points. isks for the sake of readability. Note that this simplification does not
We start from the generalized Hooke law for linear elastic mate- introduce any notational ambiguity since (i) the stress and strain rates
rial, also known as the Saint Venant–Kirchhoff material model. This are instantaneous quantities, while the known stress is calculated at
material model is well-suited for the small strain and large rotation the previous configuration, and (ii) all integrations are performed with
analysis [39]. The only non-zero components of the stress and strain respect to the metric of the current configuration, in accordance with
rates are related as: the updated Lagrangian procedure [43].
( )2 In order to reduce the dimension from 3D to 1D, it is necessary to
𝜎̄̇ 11 = 𝐸 𝑔̄ 11 𝑑̄11 , (31) integrate the left-hand side of Eq. (34) over the area of the cross section.
Thus, integrals along the length of the beam axis are obtained:
where 𝐸 is the Young’s modulus of elasticity. It is a well-known fact
( )√
that this axial state of stress is an erroneous consequence of the BE 𝜎̄̇ 11 𝛿 𝑑̄11 d𝑉 𝛥𝑡 + 𝜎̄ 11 𝛿 𝑑̄11 d𝑉 = 𝑁̃̇ 𝛿𝑑11 + 𝑀
̃̇ 𝛿 𝜅̇ 𝑔 d𝜉 𝛥𝑡
assumptions. Nevertheless, it readily used due to its simplicity. ∫𝑉 ∫𝑉 ∫𝜉
(35)
( )√
+ 𝑁̃ 𝛿𝑑11 + 𝑀 ̃ 𝛿 𝜅̇ 𝑔 d𝜉 ,
∫𝜉
4.1. Principle of virtual power
where 𝑁̃ and 𝑀 ̃ are the stress resultant and the stress couple, which
The principle of virtual power represents a weak form of the equi- are energetically conjugated with the reference strain rates of the beam
librium. It states that at any instance of time, the total power of the axis, 𝑑11 , and 𝜅̇ , while 𝑁̃̇ and 𝑀
̃̇ are their respective rates:
external, internal and inertial forces is zero for any admissible virtual
state of motion. If the inertial effects are neglected, and body and 𝑁̃̇ = (𝑔0 )2 𝜎̄̇ 11 (1 + 𝜂𝐾) d𝜂 d𝜁 and ̃̇ = −
𝑀 𝜂(𝑔0 )2 𝜎̄̇ 11 d𝜂 d𝜁 . (36)
∫𝐴 ∫𝐴

4
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

These expressions are analogous to those obtained in [28]. By introduc- where the 4th moment of area can be easily calculated for rectangular
ing the vectors of generalized section forces, strain rates of the beam and circular cross sections.
axis, and external line loads:
[ ] [ ] [ ] 4.3. Variation of strains
𝐟 𝖳 = 𝑁̃ 𝑀̃ , 𝐞𝖳 = 𝑑11 𝜅̇ , 𝐩𝖳 = 𝑝1 𝑝2 , (37)
Eq. (34) can be written in compact matrix form as: Since the strain rate is a function of the generalized coordinates as
√ √ √ well as the metric, we must vary it with respect to both arguments. By
𝐟̇ 𝖳 𝛿𝐞 𝑔 d𝜉 𝛥𝑡 + 𝐟 𝖳 𝛿𝐞 𝑔 d𝜉 = 𝐩𝖳 𝛿𝐯 𝑔 d𝜉 . (38) noting that the variation of the tangent vectors can be expressed as:
∫𝜉 ∫𝜉 ∫𝜉
This equation is nonlinear and it will be linearized in Section 4.4. 𝛿𝐠𝛼 = 𝛿𝐯,𝛼 𝛥𝑡, (45)

4.2. Relation between energetically conjugated pairs the variations of reference strains are given by:
( )
The geometrically exact relations (35) and (36) are crucial for 𝛿𝑑11 = 𝛿 𝐠1 ⋅ 𝐯,1 = 𝛿𝐯,1 ⋅ 𝐯,1 𝛥𝑡 + 𝐠1 ⋅ 𝛿𝐯,1 ,
the accurate formulation of structural beam theories. In particular, [ ( )]
1
𝛿 𝜅̇ = 𝛿 𝐠2 ⋅ 𝐯,11 − 𝛤11 𝐯,1
they allow a rigorous definition of energetically conjugated pairs, and ( ) ( ) ( )
guarantee that the appropriate constitutive matrix is symmetric. By the 1
= 𝛿𝐯,2 · 𝐯,11 − 𝛤11 1
𝐯,1 𝛥𝑡 − 𝛿𝛤11 1
𝐠2 · 𝐯,1 + 𝐠2 · 𝛿𝐯,11 − 𝛤11 𝛿𝐯,1 .
introduction of Eqs. (29) and (31) into Eq. (36), we obtain the exact
relation between energetically conjugated pairs of stress and strain (46)
rates. The resulting symmetric constitutive matrix 𝐃 is derived in [28]:
Most of the terms in the previous expression are easily computed
[ ] since the variation is explicitly performed with respect to the unknown
𝐸 𝐴 −𝐼̄ variables. However, the first two addends of 𝜅̇ are exceptions since they
𝐟̇ = 𝐃𝐞, 𝐃= , (39)
𝑔2 −𝐼̄ 𝐼 require the variations of the velocity of the normal and the Christoffel
where: symbol. These variations must be represented via the variations of
generalized coordinates [33].
(1 + 𝜂𝐾)2 𝜂 (1 + 𝜂𝐾) 𝜂2
𝐴= d𝜂 d𝜁 , 𝐼̄ = d𝜂 d𝜁 , 𝐼= d𝜂 d𝜁 . The variation of the velocity of the normal follows directly from
∫𝐴 𝑔0 ∫𝐴 𝑔0 ∫𝐴 𝑔 0
Eq. (22):
(40) ( )
1
𝛿𝐯,2 = − 𝐠2 ⋅ 𝛿𝐯,1 𝐠1 . (47)
Importantly, these integrals can be analytically determined for the 𝑔
symmetrical solid cross section shapes [28]. In the following, we refer For the variation of the Christoffel symbols, we start from Eq. (5) and,
to the exact constitutive model given in Eq. (39) by 𝐷𝑎 . after some straightforward calculation, obtain:
We introduce four reduced models to examine the various influences ( ) [ ]
1 1 1(̃ )
of the exact constitutive relation. The first and the simplest of these is 𝛿𝛤11 = 𝛿 𝐠1,1 ⋅ 𝐠1 = 𝐠1 𝛿𝐯,11 + 1
𝐾 𝐠2 − 𝛤11 𝐠1 ⋅ 𝛿𝐯,1 𝛥𝑡. (48)
𝑔 𝑔
designated with 𝐷0 . It is often employed for thin beams, e.g. [19,25],
since it decouples axial and bending actions: By inserting Eqs. (47) and (48) into Eq. (46), the final expression for
[ ] the variation of the rate of curvature change is found:
𝐴0 0 ( )
𝐃0 = , 𝐴0 = d𝜂 d𝜁 , 𝐼0 = 𝜂 2 d𝜂 d𝜁 . (41) 1
0 𝐼0 ∫𝐴 ∫𝐴 𝛿 𝜅̇ = 𝐠2 · 𝛿𝐯,11 − 𝛤11 𝛿𝐯,1
{( ) ( )
The second reduced model, 𝐷1 , is based on the approximation: 𝑔0 → 1 1
− 𝐠2 ⋅ 𝛿𝐯,1 𝐠1 · 𝐯,11 − 𝛤11 𝐯,1 (49)
1. It is readily utilized for the analysis of curved beams with small 𝑔
[ ( ) ]( )}
curvature [9,32]. The model is specified by: 1
[ ] + 𝐠1 𝛿𝐯,11 + 𝐾̃ 𝐠2 − 𝛤11 𝐠1 ⋅ 𝛿𝐯,1 𝐠2 · 𝐯,1 𝛥𝑡.
𝐴1 −𝐼̄1
𝐃1 = , 𝐴1 = (1 + 𝜂𝐾)2 d𝜂 d𝜁 ≈ d𝜂 d𝜁 = 𝐴0 ,
−𝐼̄1 𝐼1 ∫𝐴 ∫𝐴 4.4. Discrete equation of motion
(42)
𝐼̄1 = 𝜂 (1 + 𝜂𝐾) d𝜂 d𝜁 = 𝐾𝐼0 , 𝐼1 = 𝜂 2 d𝜂 d𝜁 = 𝐼0 .
∫𝐴 ∫𝐴 In this subsection, the virtual power is spatially discretized and
where the quadratic term in the integrand of property 𝐴1 is disregarded. linearized. We start by introducing the matrix 𝐁𝐿 , which relates the
Finally, the models 𝐷2 and 𝐷3 are based on the Taylor approxima- reference strain rates of the beam axis with the velocities of control
tion of the exact expressions (40): points:
[ ]
𝐴2 −𝐼̄2 ̇
𝐞 = 𝐁𝐿 𝐪. (50)
𝐃2 = ̄ ,
−𝐼2 𝐼2
Using Eqs. (20) and (30), the vector of the reference strain rates can be
𝐴2 = d𝜂 d𝜁 = 𝐴0 , represented as:
∫𝐴
(43) [ ] [ ]
( ) ∑
2 [ ] 𝑥𝛼,1 0 𝑣𝛼,1
𝐼̄2 = 2𝐾𝜂 2 d𝜂 d𝜁 = 2𝐾𝐼0 , 𝐞 = 𝐇𝐰 = 𝐇𝛼 𝐰𝛼 , 𝐇 = 𝐇1 𝐇2 , 𝐇𝛼 = , 𝐰𝛼 = ,
∫𝐴 𝛼=1
1
−𝛤11 𝑥𝛼,2 𝑥𝛼,2 𝑣𝛼,11

𝐼2 = 𝜂 2 d𝜂 d𝜁 = 𝐼0 , (51)
∫𝐴
[ ] where the vector 𝐰 is defined by:
𝐴3 −𝐼̄3 [ 1] [ ]
𝐃3 = , 𝐰 𝐁11 ... 𝐁1𝐼 ... 𝐁1𝑁
−𝐼̄3 𝐼3 ̇
𝐰 = 𝐁𝐪, 𝐰= , 𝐁 = . (52)
2
( ) 𝐰 𝐁21 ... 𝐁2𝐼 ... 𝐁2𝑁
𝐴3 = 1 + 4𝜂 2 𝐾 2 d𝜂 d𝜁 = 𝐴0 + 4𝐾 2 𝐼0 ,
∫𝐴 The submatrices 𝐁𝛼𝐼 for an arbitrary control point 𝐼 consist of the
(44)
( ) derivatives of the basis functions, i.e.:
𝐼̄3 = 2𝐾𝜂 2 + 2𝜂 4 𝐾 3 d𝜂 d𝜁 = 2𝐾𝐼0 + 2𝐾 3 𝜂 4 d𝜂 d𝜁 ,
∫𝐴 ∫𝐴 [ ] [ ]
( 2 ) 1
𝑅𝐼,1 0 2
0 𝑅𝐼,1
𝐼3 = 𝜂 + 𝜂 4 𝐾 2 d𝜂 d𝜁 = 𝐼0 + 𝐾 2 𝜂 4 d𝜂 d𝜁 . 𝐁𝐼 = , 𝐁𝐼 = . (53)
∫𝐴 ∫𝐴 𝑅𝐼,11 0 0 𝑅𝐼,11

5
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

The matrix 𝐁𝐿 now follows as: particular, the full BE beam metric is incorporated in Eq. (55) and
the symmetric nature of the geometric stiffness term is stressed by the
𝐞 = 𝐇𝐰 = 𝐇𝐁𝐪̇ = 𝐁𝐿 𝐪,
̇ 𝐁𝐿 = 𝐇𝐁. (54) elegant and compact form of Eqs. (56) and (59). The latter confirms that
Next, we need to specify the explicit matrix form of the virtual power the formulation is valid and that the adopted force and strain quantities
generated by the known stress and the variation of strain rate with are energetically conjugated.
respect to the metric, Eqs. (38) and (46):
5. Numerical examples
√ ∑
2 ∑
2

𝖳 𝛼 𝖳
𝐟 𝛿𝐁𝐿 𝐪̇ 𝑔 d𝜉 𝛥𝑡 = (𝐰 ) 𝐆𝛽𝛼 𝛿𝐰𝛽 𝑔 d𝜉 𝛥𝑡 The aim of the subsequent numerical experiments is to validate
∫𝜉 ∫𝜉
𝛼=1 𝛽=1
the proposed approach and to examine the influence of curviness on

2 ∑
2 ( ) structural response.
(𝐰𝛼 )𝖳 𝐆𝛽𝛼 𝛿𝐰𝛽 = 𝑁̃ 𝑣𝛼,1 𝛿𝑣𝛼,1
The boundary conditions are imposed strongly and the rotations
𝛼=1 𝛽=1
( ) are treated with special care, see [28], since they are not utilized as
̃ 𝑣𝛼 𝑌 𝛽 𝛿𝑣 − 𝑣𝛼 𝐵 𝛽 𝛿𝑣 − 𝑣𝛼 𝐵̃ 𝛽 𝛿𝑣
+𝑀 ,
,1 𝛼 𝛽,1 ,11 𝛼 𝛽,1 ,1 𝛼 𝛽,11 DOFs. Locking issues are not considered, but the presence of membrane
locking is already detected for linear analyzes in [28], and it is present
(55)
in nonlinear analysis as well. However, its influence is alleviated with
where the matrix 𝐆𝛽𝛼 is: the usage of higher order basis functions [44]. It is interesting that high
[ ] interelement continuity increases the locking effect due to the presence
̃ 𝑌𝛼𝛽
𝑁̃ 𝛿𝛼𝛽 + 𝑀 ̃ 𝐵̃ 𝛼𝛽
−𝑀
𝐆𝛽𝛼 = , (56) of more constraints, which must be satisfied. This issue can be dealt
−𝑀 𝐵𝛼𝛽
̃ 0 with an appropriate numerical integration scheme [44]. The quest for
with: optimal quadrature rules in IGA is an ongoing topic, see e.g. [45], but
( ) the standard Gauss quadrature with 𝑝 + 1 integration points are used
1 ̃ 𝛽 1 𝛽 1
1 𝛽
𝑌𝛼𝛽 = 𝛤11 𝐵𝛼 − 𝐾 𝑥,2 − 𝛤11 𝑥,1 𝑥𝛼,2 , 𝐵̃ 𝛼𝛽 = 𝑥𝛽,1 𝑥𝛼,2 = 𝐵𝛽𝛼 . (57) here.
𝑔 𝑔
Since no rotational DOFs are required for BE beams, the implemen-
By introducing the total matrix of the generalized section forces:
[ ] tation of multi-patch structures receives much attention [25,46–49]. In
𝐆11 𝐆21 contrast to the implicit method suggested in [47], the applied approach
𝐆= , (58) is explicit and uses a constraint equation [36]. Concretely, the rotation
𝐆12 𝐆22
at the end of NURBS curve depends on the displacements of the last two
expression (55) can be written as: control points [37]. Therefore, the rotation of two NURBS curves at a
√ √ joint is a function of six displacement components. In order to prescribe
𝐟 𝖳 𝛿𝐁𝐿 𝐪̇ 𝑔 d𝜉 𝛥𝑡 = 𝐰𝖳 𝐆𝛿𝐰 𝑔 d𝜉 𝛥𝑡. (59)
∫𝜉 ∫𝜉 rigid connection between two patches, one displacement component is
expressed as a function of the other five. This function is then used
A careful inspection of Eqs. (56) and (58) reveals that the matrix of
to constrain the equations of equilibrium (63). Present approach is
generalized section forces, 𝐆, is symmetric [33]. Now, the integrands
straightforward and does not require the introduction of bending strips
in the equation of the virtual power (38) reduce to:
or end rotational DOFs [24,25,48], although it lacks the modeling
𝐟̇ 𝖳 𝛿𝐞 ≈ 𝐪̇ 𝖳 𝐁𝐿 𝖳 𝐃𝐁𝐿 𝛿 𝐪,
̇ flexibility of the mentioned references.
( ) (60)
𝐟 𝛿𝐞 = 𝐟 𝛿𝐁𝐿 𝐪̇ + 𝐁𝐿 𝛿 𝐪̇ = 𝐪̇ 𝖳 𝐁𝖳 𝐆𝐁𝛿 𝐪𝛥𝑡
𝖳 𝖳 ̇ + 𝐟 𝖳 𝐁𝐿 𝛿 𝐪,
̇ All the results are related to the load proportionality factor (𝐿𝑃 𝐹 ),
rather than the load intensity itself. Although the implemented code
where the first term is linearized by neglecting the variation of strain
allows the usage of large load increments for some examples, the results
rate with respect to the metric:
are mostly calculated with small increments in order to enable a clear
̇
𝛿𝐞 ≈ 𝐠1 · 𝛿𝐯,1 = 𝐁𝐿 𝛿 𝐪. (61) comparison of the obtained results via smooth equilibrium paths. For
problems that do not exhibit a snap behavior, it is possible to obtain
Finally, the equation of equilibrium reduces to: the deformed configurations for the prescribed load increments. Thus,
( 𝖳 )√ the relative differences between different models can be compared
𝐪̇ 𝖳 𝐁𝐿 𝐃𝐁𝐿 + 𝐁𝖳 𝐆𝐁 ̇
𝑔 d𝜉 𝛿 𝐪𝛥𝑡
∫𝜉 along the equilibrium path, c.f. Section 5.1. These graphs give a clearer
√ √ insight into discrepancies of the models than a simple comparison of
= 𝐩𝖳 𝐍 𝑔 d𝜉 𝛿 𝐪̇ − 𝐟 𝖳 𝐁𝐿 𝑔 d𝜉 𝛿 𝐪,
̇ (62) equilibrium paths.
∫𝜉 ∫𝜉
Since the time is a fictitious quantity in the present static analysis,
which is readily written in the standard form:
strain and stress rates are equal to strains and stresses, respectively.
( )
𝐊𝑇 𝛥𝐪 = 𝐐 − 𝐅, ̇
𝛥𝐪 = 𝐪𝛥𝑡 , (63)
5.1. Cantilever beam subjected to an end concentrated moment
where:
√ √ This example is a classic benchmark test for the validation of
𝐊𝑇 = 𝐁𝐿 𝖳 𝐃𝐁𝐿 𝑔 d𝜉 + 𝐁𝖳 𝐆𝐁 𝑔 d𝜉 , (64)
∫𝜉 ∫𝜉 nonlinear beam formulations [23–25]. The cantilever beam is loaded
is the tangent stiffness matrix and: at its free end with a moment 2𝑛𝐸𝐼𝜋∕𝐿, where 𝑛 defines the num-
ber of circles into which the beam rolls-up, Fig. 2. A square cross
√ √
𝐐= 𝐍𝖳 𝐩 𝑔 d𝜉 , 𝐅= 𝐁𝐿 𝖳 𝐟 𝑔 d𝜉 , (65) section is considered and its dimension varies from the set 𝑏 = ℎ ∈
∫𝜉 ∫𝜉
{0.05, 0.1, 0.2, 0.4, 0.8}. If the beam is rolled up into one circle (𝑛 = 1)
are the vectors of the external and internal forces, respectively. The the curviness becomes 𝐾ℎ ≈ {0.03, 0.06, 0.13, 0.25, 0.50} for 𝐿𝑃 𝐹 = 1. If
vector 𝛥𝐪 in Eq. (63) contains increments of the displacements. Due to the beam is rolled up into two circles (𝑛 = 2), the value of curviness at
the approximation introduced in Eq. (61), the solution of Eq. (63) does the final configuration is also doubled.
not satisfy the principle of virtual power directly, but has to be enforced A moment load can be applied in various ways for rotation-free
by additional numerical schemes. Here, both the Newton–Raphson and models. For example, an additional knot can be inserted to define a
arc-length methods are employed. couple [23], or the linear stress distribution can be imposed, [20].
The present derivation of the geometric stiffness matrix differs from Here, no additional knots are added and the force couple is defined by
the conventional procedure for nonlinear beam formulations [25]. In the external virtual power, analogous to [28]. Although the external

6
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

model slightly deviates. On the other hand, the results obtained with
the 𝐷𝑎 and 𝐷3 models are virtually indistinguishable from analytical
predictions for strongly curved beam, while different from those of the
thin beam. The 𝐷2 model deviates from the exact predictions, but with
a different sign than the 𝐷0 and 𝐷1 models. These results show that
the present approach can return accurate results for problems with
large displacements and rotations. For beams with small curvature,
the results are aligned with the classic predictions. As the curviness
increase, the difference between constitutive models becomes clearly
visible.
In order to present these findings more concisely, the relative differ-
ences of the reduced models with respect to the 𝐷𝑎 model are presented
Fig. 2. Cantilever beam. Geometry and applied load. in Fig. 7 for four different values of curviness. These results confirm
that the reduced model 𝐷3 returns reasonably accurate results, even
for strongly curved beams. It is interesting to note the relatively large
moment has constant direction, the forces change direction throughout differences of the 𝐷2 model, which suggest that the inclusion of higher
the whole deformation. Therefore, the external load vector must be order terms, as in the reduced constitutive model 𝐷3 , is required if
updated at each iteration to define the applied moment accurately. accurate results for strongly curved beams are sought, see Eqs. (43)
Additionally, this change of direction results with an important contri- and (44).
bution to the stiffness matrix, c.f. Appendix C. This matrix significantly Another important conclusion follows from this specific example.
improves convergence properties of the nonlinear solver. Since the curviness is constant along the beam, the introduced exact
Since the deformed beam axis has constant curvature at each con- metric has more impact on the beam response than it is the case
figuration, this problem can be solved analytically. Based on the con- with examples where the maximum curviness is local. We can observe,
siderations given in [50], an analytical solution for a strongly curved e.g., that if the curviness along the whole beam is less than 0.15,
beam is derived in Appendix B. the error of the simple decoupled 𝐷0 model is less than 0.3%. As the
curviness increases over 0.25, the error increase over 1.1%, etc.
5.1.1. Convergence analysis
We examine the convergence properties of the developed formu- 5.1.3. Specific aspects of strongly curved beams
lation for two constitutive models, i.e., 𝐷1 and 𝐷𝑎 , using 𝑛 = 1 and This example is sometimes misinterpreted due to the pure bending
ℎ = 0.2 → 𝐾ℎ ≈ 0.13. Since these two models return different structural conditions. It is assumed that there is no change of the length of beam
responses, different reference solutions must be utilized. For the 𝐷1 axis and the beam rolls-up into a perfect circle with the circumference
model, reference analytical solution for thin beam is employed [23]. equal to the length of the beam [19]. However, when we deal with a
For the 𝐷𝑎 model, the analytical solution is derived in Appendix B. The geometrically exact analysis that considers the effect of curviness, the
𝐿2 -norm of the relative error of displacement for 𝐿𝑃 𝐹 = 1 is considered strain is distributed nonlinearly across the height of the cross section
in Fig. 3 for three different polynomial orders with highest available with a non-zero value at the centroid. This strain is extensional and
interelement continuity. The expected order of convergence for linear its value is negligible for very thin beams, but becomes significant for
analysis is 𝑝 + 1, [9], and the obtained values are in line with these strongly curved beams. Consequently, the axis of a cantilever loaded
theoretical estimates. with tip moment will reach full circle for 𝑛 < 1 and its circumference
Fig. 4 illustrates the convergence of the normal force and the will be somewhat larger than the original length. In order to depict this,
bending moment using the 𝐷𝑎 model. The analytical solution is used the four configurations of the beam with ℎ = 0.2 are visualized in Fig. 8.
as reference. Since the second derivatives influence both the normal Note how the initially thin and slender beam becomes strongly curved
force and the bending moment, the expected order of convergence during the deformation. As a consequence, the metric, axial strain, and
is 𝑝 − 1. However, obtained orders are closer to 𝑝. The exception to stress across the height of the cross section are distributed nonlinearly.
this observation is the mesh with cubic splines. A reduced order of The distribution of stress is shown next to the deformed configurations
convergence for cubics is also observed and discussed in [14]. Accuracy and compared with the classic linear distribution. Evidently, the intra-
of normal force is poor due to the oscillations which are present for dos stress is greater, and the extrados stress is lesser than the equivalent
coarse meshes. Similar behavior is observed in the linear analysis [28]. linear stress. Due to the small extension of beam axis, the ends of the
Finally, we have considered the error of strain energy for 𝐷𝑎 model. beam overlap for 𝐿𝑃 𝐹 = 0.5 and 𝐿𝑃 𝐹 = 1. This overlap for 𝐿𝑃 𝐹 = 0.5
In order to examine the influence of the level of nonlinearity, two levels is zoomed in Fig. 8b.
of load are considered, 𝐿𝑃 𝐹 = 1 and 𝐿𝑃 𝐹 = 0.1. The reference solution Furthermore, due to the axial strain of the beam axis, it is necessary
is calculated analytically and the results are shown in Fig. 5. Theoretical to include the influence of flexural strain when calculating the normal
values of convergence rates, 𝑝, are obtained for quadratic and quartic section force [28]. To examine this effect, 40 elements are used for
splines, while the rate is reduced for cubic spline. The rates are not a model with 𝑛 = 1 and 𝐾ℎ = 0.25, and the influence of both
affected by the level of nonlinearity in the model, but the accuracy per axial and flexural strains is observed. All five constitutive models are
DOF is. In the rest of this example, quartic splines with 𝐶 3 interelement employed and the results are presented in Fig. 9. Clearly, the normal
continuity are exclusively used. force oscillates, but with negligible amplitude. We see that the axial
strain gives tension while flexural strain results in compression force.
5.1.2. Comparison of constitutive models For the 𝐷𝑎 model, these contributions have nearly equal absolute values
A discretization with 24 elements is employed to calculate the and the normal force oscillates around zero. The result is similar for
equilibrium path of the tip for 𝑛 = 2 and two different values of the 𝐷3 model, the normal force oscillates about the value close to 0.1.
curviness. The results are compared with the analytical ones and shown For the 𝐷2 model, the error becomes pronounced since the normal
in Fig. 6. For a beam with the curviness 𝐾ℎ ≈ 0.06, the results of force oscillates around approximately 9.3. Due to the fact that the
all models are in almost full agreement, Fig. 6a. However, when the influence of flexural strain on normal section force is disregarded in
cross section dimensions are increased such that the curviness becomes the 𝐷1 model, the only contribution to the normal force comes from the
𝐾ℎ ≈ 0.50, the discrepancies become significant, Fig. 6b. The 𝐷1 model axial strain, which oscillates around zero for this model. This result is
is fully aligned with the analytical solution for thin beams, while the 𝐷0 considered accurate, but it stems from the fact that the two errors have

7
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 3. Cantilever beam. Convergence of displacement for LPF=1 using two constitutive models: (a) 𝐷1 , (b) 𝐷𝑎 .

Fig. 4. Cantilever beam. Convergence of: (a) normal force, and (b) bending moment, for LPF=1.

Fig. 5. Cantilever beam. Convergence of strain energy for: (a) 𝐿𝑃 𝐹 = 1, and (b) 𝐿𝑃 𝐹 = 0.1.

canceled each other out. I.e. the dilatation is erroneously obtained as 5.1.4. Test of the algorithms for the update of internal forces
near zero, and the influence of the flexural strain is disregarded. These The simplicity of this example is ultimately utilized in testing the
two approximations gave the correct value of normal force. Finally, algorithms for the update of internal stress. Three approaches are
the 𝐷0 model gives a completely inaccurate result. For this model, the considered: (i) the current stress is calculated from the total strain, (ii)
axial strain is negative and there is no contribution from the flexural the current stress is calculated by Eq. (33) with the first order approx-
strain. imation of adjustment term, and (iii) the current stress is calculated

8
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 6. Cantilever beam. Comparison of the displacement components of the tip obtained by different constitutive models and analytically: (a) 𝐾ℎ ≈ 0.06, (b) 𝐾ℎ ≈ 0.50.

Fig. 7. Cantilever beam. Relative differences of the displacement of the tip for four different values of curviness. The results by the reduced models are compared with those from
the exact one.

Fig. 8. Cantilever beam. Reference and deformed configurations, and stress distribution across the height of cross section for four levels of load: (a) 𝐿𝑃 𝐹 = 0.25, (b) 𝐿𝑃 𝐹 = 0.50,
(c) 𝐿𝑃 𝐹 = 0.75, (d) 𝐿𝑃 𝐹 = 1.00.

by Eq. (33) with the zeroth order approximation of adjustment term. For the first load increment, which is 0.1 here, the internal forces
These approaches are designated as 𝐹 1, 𝐹 2 and 𝐹 3, respectively, and are calculated in a same way for all algorithms. The 𝐹 1 approach,
they are elaborated in Appendix A. as expected, gives the most accurate results and this is also approach
The exact value of the axial strain of beam axis for 𝑛 = 1 is utilized in all present examples. The 𝐹 2 algorithm can be successfully
calculated analytically, c.f. Appendix B. These values are compared employed for those beams with moderate curvature. Inevitably, error of
with the results obtained with three different numerical approaches 𝐹 2 algorithm rises as the curviness and strain increase. The algorithm
for the update of internal forces and the 𝐿2 error is shown in Fig. 10. 𝐹 3 is not recommended since it gives large errors, even for the small

9
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 9. Cantilever beam. Comparison of normal force and its contributions by axial and flexural strain for different constitutive models. The resulting normal force is zoomed on
all the graphs.

Fig. 10. Cantilever beam. Comparison of relative errors of axial strain of beam axis for three algorithms for the update of internal forces. Four values of curviness are considered:
(a) 𝐾ℎ ≈ 0.06, (b) 𝐾ℎ ≈ 0.13, (c) 𝐾ℎ ≈ 0.25, (d) 𝐾ℎ ≈ 0.50.

values of curviness. Interestingly, its error is practically invariant to the discretized with five quartic 𝐶 1 elements. The results obtained with the
value of curviness. presented models are given in Fig. 11b and they are in full agreement
with those from literature [25]. This is expected because the maximum
5.2. Lee’s frame curviness in this example is local and less than 0.1. In order to examine
the influence of curviness on this structure, the height of the cross
This example is also well-established in the literature in the context section is multiplied by 2.5 (ℎ1 = 0.05) and 5 (ℎ2 = 0.1), and
of nonlinear analysis of in-plane beams [24,51,52]. The frame consists the obtained examples are designated as 𝐿𝐹 1 and 𝐿𝐹 2, respectively.
of two rigidly connected beams and the displacement components at The example with the original height, ℎ = 0.02, is labeled 𝐿𝐹 . The
the point of force application are observed, Fig. 11a. Each beam is

10
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 11. Lee’s frame. (a) Geometry and applied load. (b) Comparison of the equilibrium paths for two characteristic displacement components using developed models and [25].

Fig. 12. Lee’s frame. Equilibrium paths for the frame 𝐿𝐹 1 using five different constitutive models: (a) complete equilibrium path, (b) zoomed parts of the equilibrium path.

equilibrium paths for the examples with increased height are shown in at outer fiber in Fig. 16. It is clear that, the exact stress distribution
Figs. 12 and 13, while the curviness at three characteristic sections is should be utilized in a post-processing phase even for structures with a
given in Fig. 14. maximum local curviness less than 0.1, such as the original example
It is evident that the curviness at these sections follows similar path 𝐿𝐹 . The difference of extreme stresses between the exact and the
for both observed examples. The difference in the value of curviness linear stress distributions is close to 4.6% for the 𝐿𝐹 example. As the
practically equals the difference in cross section heights. The maximum curviness increases, this error becomes more pronounced.
curviness for 𝐿𝐹 1 is 𝐾ℎ ≈ 0.23, while for 𝐿𝐹 2 it is 𝐾ℎ ≈ 0.46.
In the context of the structural response of these frames, all consti- 5.3. Multi-snap behavior of a parabolic arch
tutive relations return similar results in the case of 𝐿𝐹 1. Inspecting the
zoomed equilibrium path for the component 𝑣𝐴 allows us to estimate The final example deals with the nonlinear response of a parabolic
the difference between 𝐷𝑎 and 𝐷0 models which is close to 0.2% for arch depicted in Fig. 17a. The arch is simply supported and loaded
𝐿𝑃 𝐹 = 1, see Fig. 12b. For the 𝐿𝐹 2 example, the discrepancies are with the vertical force at the apex. For shallow arches, the snap-
more noticeable. The largest difference between 𝐷0 and 𝐷𝑎 models is through phenomenon occurs after which the structure reaches a stable
near 0.8% at the final configuration while the results returned by the equilibrium and shows stiffening behavior. However, for deeper arches,
𝐷2 , 𝐷3 , and 𝐷𝑎 models are virtually indistinguishable. a multiple snap behavior can be expected [33,53–56]. The aim of this
Although the maximum curviness for the examples with increased example is to show that the present formulation is capable of describing
cross section height is large, the difference between displacements highly-complex responses of arbitrarily curved structures. Furthermore,
obtained by different constitutive models is significantly lower in com- we discuss a physical reasoning behind the observed behavior [53].
parison with the previous example, Fig. 7. This is mainly due to the We consider four parabolic arches labeled 𝑅1, 𝑅2, 𝑅3, and 𝑅4. They
fact that the maximum curviness of Lee’s frame is local, while in the differ in rise 𝑓 and the magnitude of the applied load 𝑃 , as detailed in
previous example it was constant along the whole beam and, therefore, Fig. 17a. Due to the symmetry, one half of the arch is modeled with 16
more impactful on the structural response. This local character of the cubic 𝐶 2 elements. As a guiding solution, we use an Abaqus simulation,
maximum curviness is underlined in Fig. 15 where the deformed con- which employs about 30 straight cubic B23 elements [57]. It is worth
figurations for 𝐿𝐹 1 and 𝐿𝐹 2 are displayed for three 𝐿𝑃 𝐹 s. Moreover, noting that Abaqus could not converge with denser meshes.
the rotations of adjacent sections at the joint of patches match, which Fig. 17b shows the signed curviness at the apex for the beam with
confirms the correct application of the constraint equation. the largest curvature, i.e., R4. The graph is interesting due to the
Finally, the stress at the section 𝐵 is considered in Fig. 16. The multiple limit points. The important fact to note is that the maximum
constitutive model 𝐷𝑎 is utilized for the calculation of displacements local curviness of this beam is lower than 0.1.
and reference strains. Then, two expressions for the distribution of For the 𝑅1 case, a relatively simple structural response is obtained,
stress are utilized: (i) the exact one, based on Eqs. (29) and (31), and and the equilibrium path for 𝑣𝑠 is shown in Fig. 18a. As the rise
(ii) the linear one, based on the same equations but with 𝐾 = 0. increases to 0.5 m, one snap-back occurs, Fig. 18b. Since these two
These distributions are shown below the equilibrium paths for the stress arches have curviness less than 0.025, only the results obtained with the

11
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 13. Lee’s frame. Equilibrium paths for the frame 𝐿𝐹 2 using five different constitutive models: (a) complete equilibrium path, (b) zoomed parts of the equilibrium path.

Fig. 14. Lee’s frame. Curviness at three characteristic sections for the frames with increased height of cross section: (a) case 𝐿𝐹 1, (b) case 𝐿𝐹 2.

Fig. 15. Lee’s frame. Deformed configurations of structures with increased height for three values of 𝐿𝑃 𝐹 : (a) case 𝐿𝐹 1, (b) case 𝐿𝐹 2.

simplest, 𝐷0 , and the most elaborated, 𝐷𝑎 , models are displayed. They from these results in the vicinity of limit points, especially for the 𝑅4
agree well for almost the complete equilibrium path while negligible case. It is noteworthy that all models return the same final deformed
discrepancies can be observed at the load limit points. Also, the results configuration.
for the 𝑅1 case are practically indistinguishable from those obtained
In order to analyze this multi-snap phenomenon more thoroughly,
by Abaqus. Regarding the 𝑅2 case, small differences occur at the load
the eight different deformed configurations of the 𝑅4 case at 𝐿𝑃 𝐹 = 0
limit points. As the rise increases further, the arch behavior becomes
are given in Fig. 20. It can be seen that these configurations are similar
more complex and multiple snaps occur, see Fig. 19. Since the results
to the buckling/vibration eigenmodes of a simple straight beam. In
for the different constitutive models are in good agreement, it can
particular, the configurations designated with 𝑎, 𝑏, 𝑐, and 𝑑 resemble
be concluded that the influence of the curviness is not significant for
the 3rd, 5th, 7th, and 9th beam eigenshape, respectively. Each of these
these arches. However, an inspection of the equilibrium paths near
configurations is stiffer than the previous one and the absolute value
the load limit points reveals that a difference between the present
of load limit points increases as well. After configuration 𝑑, which
approach and the B23 element discretization exists, which increases
is seemingly straight, the arch cannot generate any more half-waves
with the curviness and complexity of the equilibrium path. The results
and starts to release accumulated strain energy. Hence, the point 𝑑 in
obtained by the 𝐷3 and 𝐷2 models are practically identical to those
Fig. 20 approximately represents an inflection point of the equilibrium
of the 𝐷𝑎 model. Furthermore, the models 𝐷0 and 𝐷1 deviate slightly
path. After this the load limit points decrease and the arch passes

12
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 16. Lee’s frame. Stress at the outer fiber at the section 𝐵 vs. 𝐿𝑃 𝐹 (above), and the distribution of stress across the cross-section height (below) for designated points on
equilibrium path: (a) case 𝐿𝐹 , (b) case 𝐿𝐹 1, (c) case 𝐿𝐹 2.

Fig. 17. Parabolic arch. (a) Geometry and applied load; (b) Signed curviness vs. 𝐿𝑃 𝐹 for 𝑅4 case.

Fig. 18. Parabolic arch. Equilibrium paths for: (a) 𝑅1 case, and (b) 𝑅2 case.

through all the previous configurations in reverse order. Finally, a loads of a simply supported, axially loaded beam with length 𝐿. Addi-
stable equilibrium branch is reached after the last load limit point. tionally, an estimate of the maximum possible axial strain of the beam
axis is taken into consideration by calculating the initial length of the
It should be noted that this multi-snap behavior depends heavily on
parabolic arch 𝐿∗ and by finding the Green–Lagrange strain:
the imposed conditions of symmetry. In reality some small perturba-
tions would probably exist in either the force position or the geometry,
𝐿∗2
𝑛 − 10
2
which would result in a simpler response [54]. 𝑒𝑅𝑛 = . (66)
2 ⋅ 102
The detected behavior is further investigated by observing the nor-
mal force at the apex. The applied forces for all cases are scaled in This is the strain of the parabola which deforms into a straight line.
relation to the largest one. Fig. 21 illustrates the resulting equilibrium Multiplication of this strain with the axial stiffness 𝐸𝐴0 gives an
paths and marks the values of 3rd, 5th, 7th, and 9th critical buckling estimate of a maximum normal compression force that can be generated

13
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

Fig. 19. Parabolic arch. Equilibrium paths for: (a) 𝑅3 case, and (b) 𝑅4 case.

Fig. 20. Parabolic arch. Deformed configurations of 𝑅4 case for 𝐿𝑃 𝐹 = 0.

in these arches due to the considered applied load. These forces are also the normal force that is large enough to snap the beam into this mode.
designated in Fig. 21. The analogous conclusion is valid for all observed cases. When there
The number of half-wavelength forms that an arch can attain during are no more new stable forms to reach, the curvature of equilibrium
this symmetric snapping is limited. This limit is directly influenced path changes sign and the arch goes through all these configurations in
by the length of the arch and the critical buckling forces of a simple reverse, until it finds the stable form. A similar discussion takes place
straight beam with length 𝐿. To summarize, the arch 𝑅4 cannot make in [53], where the rise, thrust and flexural stiffness are identified as
the 11th eigenshape, since its maximum axial strain cannot generate the causes for multi-snap behavior of circular arches. However, the

14
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

( )
Fig. 21. Parabolic arch. Normal force at the apex for all considered cases. Critical buckling forces and estimated maximum normal compression forces EA0 eRn are marked by
arrows.

graphical description given in Fig. 21 provides additional insight into A phenomenon of multi-snap behavior of a parabolic arch is revised
the nature of the phenomenon. and elaborated. The arch accumulates strain energy and deforms into
the configurations that resemble the eigenmodes of a straight beam.
6. Conclusions After the limit eigenshape is reached, the arch releases the accumulated
strain energy and finds a stable equilibrium branch.
The previously introduced Bernoulli–Euler (BE) beam model is here
An interesting direction for future research is the application of the
implemented in the geometrically nonlinear setting. A rigorous metric
proposed formulation to the statics and dynamics of spatial beams.
of the plane BE beam is utilized consistently for the derivation of the
weak form of the equilibrium. The spatial discretization of the virtual
power is performed by the isogeometric approach. The introduction CRediT authorship contribution statement
of the full beam metric provides a higher-order accurate BE beam
formulation. Four simplified models are derived in addition to the exact A. Borković: Conceptualization, Methodology, Software, Writing,
constitutive relation. The comparison of these models via numerical Validation, Funding acquisition. B. Marussig: Writing, Formal analysis,
examples allows a detailed analysis of the influence the formulation Visualization, Funding acquisition. G. Radenković: Conceptualization,
components have and of the most importance, the beam’s curviness. Methodology, Software, Writing.
The present formulation is geometrically exact, rotation-free, and fully
capable of dealing with complex responses of multi-patch structures.
The variation of the unknown metric in the external virtual power Declaration of competing interest
term due to the concentrated moment resulted with one important
contribution to the stiffness matrix. Although the external moment does The authors declare that they have no known competing finan-
not change direction, its implementation in the rotation-free theory cial interests or personal relationships that could have appeared to
requires use of the follower force couple. influence the work reported in this paper.
Three approaches for the calculation of the section force and mo-
ment are derived and discussed. The results obtained with the incre-
mental approach are satisfactory but further enhancements should be Acknowledgments
sought for the material nonlinear analysis.
Besides the results from the literature, the present numerical model During this work, our beloved colleague and friend, Professor Gligor
is also verified by the comparison with the novel analytical solution for Radenković (1956–2019), passed away. The first author acknowledges
the strongly curved cantilever beam. The results show that, in order to that his unprecedented enthusiasm and love for mechanics were crucial
correctly determine the axial strain at the centroid of a curved beam, for much of his previous, present, and future research.
a rigorous computational model must be employed. For beams with We acknowledge the support of the Austrian Science Fund (FWF):
curviness 𝐾ℎ < 0.1, the simple decoupled equations return reasonably M 2806-N.
accurate results for the displacement field. As the curviness increases,
its influence becomes noticeable and as a result a more involved and
Appendix A. Update of internal forces
complex model is required. An important factor is the domain where
the strong curviness exists. If this is relatively local, its influence on
the global response will not be as significant as when large portions of Due to the effect of curviness, the update of internal forces requires
the structure are strongly curved. Regardless of the constitutive model, special attention. In order to add the increment of stress as in Eq. (33),
utilizing the exact expressions for the equidistant strains and stresses the cumulative stress from the previous configuration must be adjusted
is recommended, since these have a nonlinear distribution even for to the metric of the current configuration, as discussed in Section 4.1.
beams with a small curvature. The inclusion of these expressions in a After the addition, the stress is integrated and the stress resultant
post-processing phase is a simple method for improving accuracy. and stress couple are obtained. First, let us find the relation between

15
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

previously calculated stress with respect to two configurations. The the curvature of current configuration 𝜒 = 𝐾 ∗ . Furthermore, this
magnitude of traction can be expressed as: quantity is related to the change of curvature with respect to the
| | parametric convective coordinate as 𝜅 = 𝜒𝑔 ∗ .
|( ) (𝑛) 𝒈̄ ||
|(𝑛) ̄ | || (𝑛) 11 (𝑛) The physical normal force and bending moment, derived in [28],
| 𝒕 | = | (𝑛) 𝜎̄ 𝒈̄ 1 ⊗ (𝑛) 𝒈̄ 1 √ 1 ||
| | | (𝑛) 𝑔̄ |
are:
| |
| | 1
| | 𝑁 = 𝐸𝐴0 𝜖(11) − 𝐸 𝐼̄ 𝜒𝑠 ,
|( ) (𝑛+1) 𝒈̄ || 2 (B.2)
| (𝑛)
|
= | (𝑛+1) 𝜎̄ 11 (𝑛+1)
𝒈̄ 1 ⊗ (𝑛+1)
𝒈̄ 1 √ 1 |
(A.1) 𝑀 = 𝐸𝐼𝜒𝑠 .
|,
| (𝑛+1) 𝑔̄ |
| | If we impose the conditions of equilibrium for this example, 𝑁 = 0
| |
and 𝑀 = 𝑀𝑒𝑥𝑡 , the previous equations reduce to the system of two
which allows us to write the required relation as:
equations with two unknowns. Since the 𝑎𝑟𝑡𝑎𝑛ℎ function appears in the
(𝑛) 11 (𝑛)
((𝑛) 𝑔0 )2 (𝑛) 𝑔 expressions for the geometrical properties of rectangular cross-section,
(𝑛)
𝜎̄ 𝑔̄ = (𝑛)
(𝑛+1)
𝜎̄ 11 (𝑛+1) 𝑔̄ ⟹ (𝑛)
(𝑛+1)
𝜎̄ 11 = (𝑛)
(𝑛+1)
𝜎̄ 11 . (A.2)
((𝑛+1) 𝑔 )2 (𝑛+1) 𝑔 these equations are transcendental and they must be solved numeri-
0
cally. Therefore, the solution to the problem is not truly analytical.
The adjustment term next to (𝑛) (𝑛+1)
𝜎̄ 11 consists of the ratio of the initial After the reference strains of the beam axis are calculated, the
curvature correction factors at two configurations which makes it inap- functions of displacements for a specified level of external load can be
propriate for exact integration in nonlinear beam analysis. Therefore, calculated as:
some approximation is required. The first order Taylor approximation 1 √
of this ratio yields: 𝑢(𝜉) = ∗ sin (𝐾 ∗ 𝑔 ∗ 𝜉) − 𝐿𝜉,
𝐾
( (𝑛) )2 ( )2 [ √ ] . (B.3)
𝑔0 1 − 𝜂 (𝑛) 𝐾 1
𝑣(𝜉) = ∗ cos (𝐾 ∗ 𝑔 ∗ 𝜉) − 1 ,
= ( ) ≈ 1 + 2𝜂 (𝑛+1) 𝜒 , (A.3) 𝐾
(𝑛+1) 𝑔 1 − 𝜂 (𝑛) 𝐾 + (𝑛+1) 𝜒
0
Appendix C. Stiffness matrix due to the external moment
where (𝑛+1) 𝜒 is the change of curvature in the current increment
with respect to the convective arc-length coordinate, as discussed in
Since the present formulation is rotation-free, the application of
Appendix B. With this approximated adjustment term, we can perform
external moments requires special attention. An external moment is
the integration, but the higher order terms with respect to 𝜂 must
modeled as a force couple which depends on the current metric of a
be disregarded. By this means the energetically conjugated section
beam. If this metric is varied, a contribution to the stiffness matrix is
forces from the previous configuration are adjusted to the metric of
revealed. The external virtual power due to the concentrated moment
the current configuration using Eqs. (36), (A.2), and (A.3):
load is:
(𝑛)
(𝑛+1)
𝑁̃ ≈ (𝑛)
(𝑛)
𝑁̃ − 2𝜒 (𝑛)
(𝑛)
̃,
𝑀
(A.4) 𝛿𝑃𝑒𝑥𝑡 = −𝑚 𝛿𝜔, (C.1)
(𝑛)
(𝑛+1)
𝑀 ̃ ≈ (𝑛)
(𝑛)
̃.
𝑀 √
where the angular velocity is 𝜔 = 𝐠2 ⋅ 𝐯,1 ∕ 𝑔. By varying the angular
Evidently, the energetically conjugate bending moment does not re- velocity with respect to the kinematics, the vector of external load is
quire adjustment, while the normal force does. obtained [28]. Let us vary the angular velocity with respect to the
This approach for the update of internal forces is the rational one, metric:
since the integration across the cross section should be avoided at ( ) ( )
each increment. Another method is to calculate the current internal 1 1 1
𝛿𝜔 = 𝛿 √ 𝐠2 ⋅ 𝐯,1 = 𝛿 √ 𝐠2 ⋅ 𝐯,1 + √ 𝛿𝐯,2 ⋅ 𝐯,1 . (C.2)
forces from the total strain and the current constitutive matrix. This 𝑔 𝑔 𝑔
approach is suitable for the present linear elastic analysis, but it is If we note that:
not applicable for materially nonlinear behavior where stress must ( )
1 1 ( )
be updated incrementally. As noted previously, the latter approach is 𝛿 √ =− 𝐯 ⋅ 𝐠1 ⊗ 𝐠2 𝛿𝐯,1 , (C.3)
designated as 𝐹 1 and former one as 𝐹 2. Additionally, an incremental 𝑔 𝑔 3∕2 ,1
approach for the update of internal forces without any adjustment of then the insertion of Eqs. (47) and (C.3) into Eq. (C.1) yields:
previously calculated forces is implemented and marked as 𝐹 3. When 𝑚 ( )
observing the Eq. (A.3), it can in fact be considered as the zeroth order 𝛿𝑃𝑒𝑥𝑡 = 𝐯 ⋅ 𝐠1 ⊗ 𝐠2 + 𝐠2 ⊗ 𝐠1 𝛿𝐯,1 . (C.4)
𝑔 3∕2 ,1
approximation. The results of these three approaches are compared in
Evidently, we can define the following matrix:
Section 5.1.4.
( )
𝐆̄1 = 𝑚 𝐠1 ⊗ 𝐠2 + 𝐠2 ⊗ 𝐠1 . (C.5)
1
𝑔 3∕2
Appendix B. Analytical solution for the roll-up of cantilever beam
In order to include the contribution of concentrated external moment
When a cantilever is loaded with tip moment, the curvature and to the tangent stiffness, this matrix must be added to the submatrix 𝐆11
̄ 1 is symmetric.
in Eq. (58). It is important to note that matrix 𝐆
the axial strain of beam axis are constant along the beam’s length. This 1
fact allows us to find the analytical solution for a problem involving
strongly curved BE beam. Let us introduce two measures of curvature References
change with respect to the arc-length coordinate, as derived in [50].
[1] E.H. Dill, Kirchhoff’s theory of rods, Arch. Hist. Exact Sci. 44 (1) (1992) 1–23.
The first is 𝜒, the change of curvature with respect to the convective
[2] A. Kimiaeifar, G. Domairry, S.R. Mohebpour, A.R. Sohouli, A.G. Davodi, Ana-
arc-length coordinate. The second is 𝜒𝑠 , the change of curvature with lytical solution for large deflections of a cantilever beam under nonconservative
respect to the arc-length coordinate, derived by the linearization of the load based on homotopy analysis method, Numer. Methods Partial Differential
spatial configuration. The relation between these two quantities is: Equations 27 (3) (2011) 541–553.
[3] E. Reissner, On finite deformations of space-curved beams, Z. Angew. Math. Phys.
𝜒𝑠 = 𝜒 − 2𝐾 ∗ 𝜖(11) , (B.1) ZAMP 32 (6) (1981) 734–744.
[4] J.C. Simo, A finite strain beam formulation. The three-dimensional dynamic
where 𝜖(11) = 𝜖11 ∕𝑔 ∗
is the physical axial strain of beam axis, while problem. Part I, Comput. Methods Appl. Mech. Engrg. 49 (1) (1985) 55–70.
( ) [5] A. Ibrahimbegović, On finite element implementation of geometrically nonlinear
𝑔 ∗ = 𝑔∕ 1 − 2𝜖(11) . Since the beam is initially straight, the change of Reissner’s beam theory: Three-dimensional curved beam elements, Comput.
curvature with respect to the convective arc-length coordinate equals Methods Appl. Mech. Engrg. 122 (1) (1995) 11–26.

16
A. Borković, B. Marussig and G. Radenković Thin-Walled Structures 170 (2022) 108539

[6] M.A. Crisfield, G. Jelenić, Objectivity of strain measures in the geometrically [31] A. Borković, S. Kovačević, G. Radenković, S. Milovanović, D. Majstorović,
exact three-dimensional beam theory and its finite-element implementation, Proc. Rotation-free isogeometric dynamic analysis of an arbitrarily curved plane
R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 455 (1983) (1999) 1125–1147. Bernoulli-Euler beam, Eng. Struct. 181 (2019) 192–215.
[7] S.N. Atluri, M. Iura, S. Vasudevan, A consistent theory of finite stretches and [32] G. Radenković, A. Borković, Linear static isogeometric analysis of an arbitrarily
finite rotations, in space-curved beams of arbitrary cross-section, Comput. Mech. curved spatial Bernoulli–Euler beam, Comput. Methods Appl. Mech. Engrg. 341
27 (4) (2001) 271–281. (2018) 360–396.
[8] C. Meier, A. Popp, W.A. Wall, Geometrically exact finite element formulations [33] G. Radenković, Finite Rotation and Finite Strain Isogeometric Structural Analysis,
for slender beams: Kirchhoff–Love theory versus Simo–Reissner theory, Arch. Faculty of Architecture Belgrad, 2017, (in Serbian).
Comput. Methods Eng. 26 (1) (2019) 163–243. [34] F. Armero, J. Valverde, Invariant Hermitian finite elements for thin Kirchhoff
[9] L. Greco, M. Cuomo, B-spline interpolation of Kirchhoff-Love space rods, Comput. rods. I: The linear plane case, Comput. Methods Appl. Mech. Engrg. 213–216
Methods Appl. Mech. Engrg. 256 (2013) 251–269. (2012) 427–457.
[10] C. Meier, A. Popp, W.A. Wall, An objective 3D large deformation finite element [35] Y.B. Yang, Y. Liu, Y.T. Wu, Invariant isogeometric formulations for three-
formulation for geometrically exact curved Kirchhoff rods, Comput. Methods dimensional Kirchhoff rods, Comput. Methods Appl. Mech. Engrg. 365 (2020)
Appl. Mech. Engrg. 278 (2014) 445–478. 112996.
[11] C. Meier, A. Popp, W.A. Wall, A locking-free finite element formulation and [36] J. Kiendl, K.U. Bletzinger, J. Linhard, R. Wüchner, Isogeometric shell analysis
reduced models for geometrically exact Kirchhoff rods, Comput. Methods Appl. with Kirchhoff–Love elements, Comput. Methods Appl. Mech. Engrg. 198 (49)
Mech. Engrg. 290 (2015) 314–341. (2009) 3902–3914.
[12] T.J.R. Hughes, J.A. Cottrell, Y. Bazilevs, Isogeometric analysis: CAD, finite [37] L. Piegl, W. Tiller, The NURBS Book, in: Monographs in Visual Communication,
elements, NURBS, exact geometry and mesh refinement, Comput. Methods Appl. Springer-Verlag, Berlin Heidelberg, 1995.
Mech. Engrg. 194 (39) (2005) 4135–4195. [38] P. Naghdi, The theory of shells and plates, in: C. Truesdell (Ed.), Linear
[13] Isogeometric analysis: progress and challenges, Comput. Methods Appl. Mech. Theories of Elasticity and Thermoelasticity, Springer, Berlin, Heidelberg, 1973,
Engrg. 316 (2017) 1. pp. 425–640.
[14] E. Marino, Isogeometric collocation for three-dimensional geometrically exact [39] M. Bischoff, K.U. Bletzinger, W.A. Wall, E. Ramm, Models and finite elements for
shear-deformable beams, Comput. Methods Appl. Mech. Engrg. 307 (2016) thin-walled structures, in: Encyclopedia of Computational Mechanics, American
383–410. Cancer Society, 2004.
[15] E. Marino, Locking-free isogeometric collocation formulation for three- [40] P. Wriggers, Nonlinear Finite Element Methods, Springer-Verlag, Berlin
dimensional geometrically exact shear-deformable beams with arbitrary initial Heidelberg, 2008.
curvature, Comput. Methods Appl. Mech. Engrg. 324 (2017) 546–572. [41] G.C. Johnson, D.J. Bammann, A discussion of stress rates in finite deformation
[16] O. Weeger, S.-K. Yeung, M.L. Dunn, Isogeometric collocation methods for problems, Int. J. Solids Struct. 20 (8) (1984) 725–737.
Cosserat rods and rod structures, Comput. Methods Appl. Mech. Engrg. 316 [42] G. Radenković, A. Borković, B. Marussig, Nonlinear static isogeometric analysis
(2017) 100–122. of arbitrarily curved Kirchhoff-Love shells, Int. J. Mech. Sci. 192 (2021) 106143.
[17] E. Marino, J. Kiendl, L. De Lorenzis, Isogeometric collocation for implicit dy- [43] K.J. Bathe, Finite Element Procedures, Klaus-Jurgen Bathe, Boston, Mass., 2007.
namics of three-dimensional beams undergoing finite motions, Comput. Methods [44] C. Adam, S. Bouabdallah, M. Zarroug, H. Maitournam, Improved numerical
Appl. Mech. Engrg. 356 (2019) 548–570. integration for locking treatment in isogeometric structural elements, Part I:
[18] D. Vo, P. Nanakorn, T.Q. Bui, A total Lagrangian Timoshenko beam formulation Beams, Comput. Methods Appl. Mech. Engrg. 279 (2014) 1–28.
for geometrically nonlinear isogeometric analysis of spatial beam structures, Acta [45] F. Auricchio, F. Calabrò, T.J.R. Hughes, A. Reali, G. Sangalli, A simple algorithm
Mech. 231 (9) (2020) 3673–3701. for obtaining nearly optimal quadrature rules for NURBS-based isogeometric
[19] A. Tasora, S. Benatti, D. Mangoni, R. Garziera, A geometrically exact isogeometric analysis, Comput. Methods Appl. Mech. Engrg. 249–252 (2012) 15–27.
beam for large displacements and contacts, Comput. Methods Appl. Mech. Engrg. [46] D. Vo, A. Borković, P. Nanakorn, T.Q. Bui, Dynamic multi-patch isogeometric
358 (2020) 112635. analysis of planar Euler–Bernoulli beams, Comput. Methods Appl. Mech. Engrg.
[20] M.-J. Choi, R.A. Sauer, S. Klinkel, An isogeometric finite element formulation 372 (2020) 113435.
for geometrically exact timoshenko beams with extensible directors, Comput. [47] L. Greco, M. Cuomo, An implicit G1 multi patch B-spline interpolation for
Methods Appl. Mech. Engrg. 385 (2021) 113993. Kirchhoff–Love space rod, Comput. Methods Appl. Mech. Engrg. 269 (2014)
[21] S.B. Raknes, X. Deng, Y. Bazilevs, D.J. Benson, K.M. Mathisen, T. Kvamsdal, 173–197.
Isogeometric rotation-free bending-stabilized cables: statics, dynamics, bending [48] L. Greco, M. Cuomo, An isogeometric implicit G1 mixed finite element for
strips and coupling with shells, Comput. Methods Appl. Mech. Engrg. 263 (2013) Kirchhoff space rods, Comput. Methods Appl. Mech. Engrg. 298 (2016) 325–349.
127–143. [49] G. Marchiori, A. Gay Neto, G. Marchiori, A. Gay Neto, Isogeometric analysis
[22] L. Greco, M. Cuomo, Consistent tangent operator for an exact Kirchhoff rod applied to 2D Bernoulli-Euler beam model: Imposition of constraints by Lagrange
model, Contin. Mech. Thermodyn. 27 (4) (2015) 861–877. and penalty methods, Lat. Am. J. Solids Struct. 17 (1) (2020).
[23] A.M. Bauer, M. Breitenberger, B. Philipp, R. Wüchner, K.U. Bletzinger, Nonlinear [50] G. Radenković, A. Borković, On the analytical approach to the linear analysis
isogeometric spatial Bernoulli beam, Comput. Methods Appl. Mech. Engrg. 303 of an arbitrarily curved spatial Bernoulli–Euler beam, Appl. Math. Model. 77
(2016) 101–127. (2020) 1603–1624.
[24] Z. Huang, Z. He, W. Jiang, H. Qiao, H. Wang, Isogeometric analysis of the [51] K.H. Schweizerhof, P. Wriggers, Consistent linearization for path following
nonlinear deformation of planar flexible beams with snap-back, Acta Mech. methods in nonlinear Fe analysis, Comput. Methods Appl. Mech. Engrg. 59 (3)
Solida Sin. 29 (4) (2016) 379–390. (1986) 261–279.
[25] D. Vo, P. Nanakorn, Geometrically nonlinear multi-patch isogeometric analysis [52] M. Rezaiee-Pajand, R. Naserian, Geometrical nonlinear analysis based on
of planar curved Euler–Bernoulli beams, Comput. Methods Appl. Mech. Engrg. optimization technique, Appl. Math. Model. 53 (2018) 32–48.
366 (2020) 113078. [53] A.B. Sabir, A.C. Lock, Large deflexion, geometrically non-linear finite element
[26] F. Maurin, F. Greco, S. Dedoncker, W. Desmet, Isogeometric analysis for analysis of circular arches, Int. J. Mech. Sci. 15 (1) (1973) 37–47.
nonlinear planar Kirchhoff rods: weighted residual formulation and collocation [54] H. Harrison, Post-buckling behaviour of elastic circular arches., Proc. Inst. Civ.
of the strong form, Comput. Methods Appl. Mech. Engrg. 340 (2018) 1023–1043. Eng. 65 (2) (1978) 283–298.
[27] R.K. Kapania, J. Li, On a geometrically exact curved/twisted beam theory under [55] M.J. Clarke, G.J. Hancock, A study of incremental-iterative strategies for
rigid cross-section assumption, Comput. Mech. 30 (5) (2003) 428–443. non-linear analyses, Internat. J. Numer. Methods Engrg. 29 (7) (1990)
[28] A. Borković, S. Kovačević, G. Radenković, S. Milovanović, M. Guzijan-Dilber, 1365–1391.
Rotation-free isogeometric analysis of an arbitrarily curved plane Bernoulli–Euler [56] Y.-B. Yang, M.-S. Shieh, Solution method for nonlinear problems with multiple
beam, Comput. Methods Appl. Mech. Engrg. 334 (2018) 238–267. critical points, AIAA J. 28 (12) (1990) 2110–2116.
[29] V. Slivker, Mechanics of Structural Elements: Theory and Applications, Softcover [57] M. Smith, ABAQUS/Standard User’s Manual, Version 6.9, 2009.
reprint of hardcover 1st ed. 2007 edition, Springer, 2010.
[30] A. Cazzani, M. Malagù, E. Turco, Isogeometric analysis of plane-curved beams,
Math. Mech. Solids 21 (5) (2016) 562–577.

17

You might also like