Accepted Manuscript A consistent Timoshenko hysteretic beam finite element model M. Amir, K.G. Papakonstantinou, G.P. Warn
PII: DOI: Reference:
S0020-7462(19)30083-6 https://doi.org/10.1016/j.ijnonlinmec.2019.07.003 NLM 3218
To appear in:
International Journal of Non-Linear Mechanics
Received date : 4 February 2019 Revised date : 3 June 2019 Accepted date : 7 July 2019 Please cite this article as: M. Amir, K.G. Papakonstantinou and G.P. Warn, A consistent Timoshenko hysteretic beam finite element model, International Journal of Non-Linear Mechanics (2019), https://doi.org/10.1016/j.ijnonlinmec.2019.07.003 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
A consistent Timoshenko hysteretic beam finite element model M. Amira,∗, K.G. Papakonstantinoub , G.P. Warnc a Graduate
Student Researcher, Dept. of Civil and Environmental Engineering, The Pennsylvania State University, University Park, PA 16802, USA. b Assistant Professor, Dept. of Civil and Environmental Engineering, The Pennsylvania State University, University Park, PA 16802, USA. c Associate Professor, Dept. of Civil and Environmental Engineering, The Pennsylvania State University, University Park, PA 16802, USA.
Abstract A parametrized Timoshenko hysteretic beam finite element model is developed, using consistent two-node elements, to efficiently simulate the nonlinear behavior of structures. New displacement interpolation functions are derived that satisfy both the exact equilibrium and kinematic conditions. Therefore, strain field continuity along the beam element is achieved, resulting in a consistent formulation without shear locking effects. Multiaxial interactions are considered through yield/capacity functions and distributed plasticity is accounted for by appropriate hysteretic interpolation functions. The suggested interpolation functions yield constant element matrices that do not require updating throughout the analysis, and the nonlinearities are captured in terms of hysteretic curvatures, axial and shear deformations, which are set to evolve through ordinary differential equations. A computationally efficient solution scheme is also suggested, without requiring any linearization, ∗ Corresponding
author Email addresses:
[email protected] (M. Amir),
[email protected] (K.G. Papakonstantinou),
[email protected] (G.P. Warn)
Preprint submitted to International Journal of Non-Linear Mechanics
July 9, 2019
to straightforwardly solve the resulting system of equations for quasi-static problems. The consistency, efficiency, versatility and validity of the suggested model is explained in detail and demonstrated through several numerical examples and comparisons with experimental data from available tests. Keywords: Hysteretic finite element, multiaxial interactions, distributed plasticity, ODEs/DAEs. 1
1. Introduction
2
In this paper, a consistent and efficient Timoshenko hysteretic beam finite element
3
model is developed. The presented approach deviates from traditional nonlinear
4
structural modeling techniques [1], including discrete plasticity uniaxial nonlinear
5
spring models [2], fiber-based approaches [3, 4], and solid finite elements mod-
6
eling, and suggests an alternative hysteretic framework, conveniently based on
7
spatially distributed, multiaxial, interacting stress resultants. The formulation is
8
largely inspired by the notable works of Triantafyllou and Koumousis [5, 6], yet this
9
paper suggests a newly developed formulation, mostly related to the consistency
10
and efficiency of the element.
11
The phenomenon of hysteresis appears in many scientific areas and a large
12
number of theoretical and mathematical models can be found in the literature,
13
e.g. [7–9]. Among the various available hysteretic models for structural systems,
14
e.g. [10–12], a considerably extended multiaxial variant of the smooth Bouc-Wen
15
model [13, 14] is emphasized and utilized herein, although the element formulation
16
idea and representation are general and alternative hysteretic models could also be
17
used. As shown in [15–17] however, the Bouc-Wen model expressions have the
18
significant attributes that they can be also derived following the physics of classical 2
19
multiaxial plasticity [18], and very compactly incorporate and satisfy the loading
20
rate, the yield/capacity criterion and the flow rule.
21
The original hysteretic Bouc-Wen model was initially developed by Bouc [13]
22
and later modified by Wen [14], and is described by first order nonlinear ordinary
23
differential equations (ODEs), often referred to as evolution equations. Over the
24
years, several extensions, modifications and variations have been suggested, e.g.
25
[19–25], and the model and its many variants have been extensively used in various
26
applications to simulate the response of diverse structural components/devices, as
27
for example described in [26]. Further extensions are also possible, for example,
28
relating hysteretic damping to viscoelasticity as suggested by [27–29]. Although
29
the original Bouc-Wen model and the vast majority of works in the literature only
30
consider a uniaxial formulation, in Casciati [15] and Sivaselvan and Reinhorn [16]
31
the model is elegantly formulated in a multiaxial plasticity context.
32
More recently, Triantafyllou and Koumousis [5, 6] suggested further multiaxial
33
plasticity enhancements and interactions, and developed an integrated hysteretic
34
beam-column element formulation, able to characterize both Timoshenko and
35
Euler-Bernoulli beam elements. In the considered beam element formulation,
36
hysteretic degrees of freedom are introduced, in terms of hysteretic curvatures, ax-
37
ial and shear deformations, and are set to evolve according to multiaxial Bouc-Wen
38
type evolution equations. The described element in Triantafyllou and Koumousis
39
[6] however, only satisfies weak equilibrium conditions for the Timoshenko case
40
and thus mesh convergence and universal/general performance for both Euler-
41
Bernoulli and Timoshenko beam types are importantly negatively affected. In this
42
paper, the exact equilibrium and kinematic conditions are simultaneously satisfied
43
due to the introduction of newly derived displacement field interpolation functions, 3
44
for a nonlinear Timoshenko beam element subjected to nodal forces. Therefore,
45
displacement and strain field continuity along the beam element are achieved,
46
resulting in a consistent formulation without any shear locking effects, regardless
47
of the element discretization, boundary conditions and beam thickness [30]. Dis-
48
tributed loads along the length of the element can be treated using the standard
49
finite element approach of equivalent nodal loads [31]. In its most comprehensive
50
form, of an inelastic Timoshenko beam formulation, the herein suggested element
51
can now also seamlessly provide compatible and accordant results to relevant
52
slender Euler-Bernoulli cases, and thus this most extensive Timoshenko format is
53
mainly described in this paper. Nonetheless, due to its parametrized nature, the
54
developed element is also directly transformable, by simple parametric changes, to
55
elastic Timoshenko and Euler-Bernoulli beam elements, exhibiting either elastic
56
or inelastic material behavior, and therefore can also be considered as a general
57
beam element model. Additionally, the advantages of the described element in
58
Triantafyllou and Koumousis [6] are also preserved in our suggested approach,
59
full interaction between axial and shear forces and moments can be considered, if
60
needed, through yield/capacity functions, and distributed plasticity is accounted
61
for by appropriate hysteretic interpolation functions.
62
One of the major advantages of the formulation in [6], and in this paper, is its
63
computational efficiency, since the time-varying element tangent stiffness matrix
64
is replaced by a constant elastic stiffness matrix and a hysteretic matrix, which both
65
need to be evaluated only once at the beginning of the analysis and do not need to be
66
updated, given that inelasticity is treated through the localized hysteretic evolution
67
equations. The whole formulation can be thus conveniently presented in state-
68
space form for dynamic cases, as shown in [5, 6], and be efficiently solved by any 4
69
generic ODE solver, avoiding any type of linearizations. In quasi-static problems,
70
however, the solution process is not as straightforward due to the presence of
71
algebraic equations now, in addition to the ODEs, resulting in a set of complex
72
and hard to solve differential algebraic equations (DAEs). Yet, a new and efficient
73
solution scheme for such problems is presented in this paper, where instead of
74
DAEs only a set of ODEs are again required to be solved, similar to the dynamic
75
cases. Through the suggested approach, the number of unknowns to be updated
76
at each pseudo time step of the resulting ODEs are also significantly reduced,
77
in comparison to both DAE systems and the typically used Newton’s method,
78
reducing computational complexity and increasing efficiency even further.
79
To summarize, the presented formulation is able to simulate the nonlinear
80
behavior of beam elements, using distributed plasticity and interaction effects, with
81
sufficient accuracy and efficient computational effort. As previously mentioned,
82
accuracy is achieved through the formulation of exact interpolation fields for the
83
nonlinear element, while efficiency is achieved through the direct use of stress
84
resultants at the section level, constant matrices and the newly devised solution
85
procedure. Thus, the suggested modeling scheme can be importantly employed
86
for the nonlinear static/dynamic analysis of large-scale frame structures, as well as
87
for system identification techniques, among many other applications dictating and
88
demanding computational efficiency.
89
In the remainder of this paper, the generality, consistency, efficiency, versa-
90
tility and validity of the suggested model is explained in detail and demonstrated
91
through several numerical examples and comparisons with experimental data from
92
available tests.
5
93
2. Hysteretic model
94
2.1. Uniaxial hysteretic model
95
Among the various hysteretic modeling approaches, a Bouc-Wen type model, based
96
on stress resultants, is adopted in this work, as already explained and justified. In
97
its simplest form, the hysteretic behavior is modeled through a set of two parallel
98
components; a linear one, contributing to the post-elastic kinematic hardening,
99
and a nonlinear, describing the hysteretic behavior according to a first order ODE,
100
otherwise known as a hysteretic evolution equation. As such, a uniaxial force
101
(P)-displacement (u) relation or any other work conjugate pair can be expressed
102
as:
n z Û uÛ P = αku + (1 − α)k z ; zÛ = 1 − h (β + γsgn(zu)) z
(1)
c
103
where k is the elastic loading stiffness, zch is the hysteretic displacement capacity,
104
α is the ratio of post-elastic to elastic stiffness, z is a hysteretic variable, β, γ,
105
n are model parameters, sgn(.) is the signum function and overdot indicates
106
differentiation with respect to time. Parameter n controls the smoothness of the
107
transition from elastic to inelastic regimes, whereas parameters β and γ control the
108
shape of the hysteretic loop. For the specific case in which the elastic unloading
109
stiffness is equal to the elastic loading stiffness, it can be proven that both β and γ
110
are equal to 0.5. Furthermore, to be thermodynamically admissible, the following
111
constraints should be imposed on the parameters β and γ [32]: −γ ≤ β ≤ γ
;
β+γ =1
(2)
112
In this paper, a generalized law is adopted that equivalently to Eq. (1) maps
113
generalized strains to their corresponding actions, in particular, centerline axial
114
strain (εu ) to axial force (N), curvature (εφ ) to moment (M), and shear strain (εγ ) 6
115
to shear force (Q) [6]. Hence, at any distance x along a beam element of length L,
116
the local M-εφ relation is given by: M(x) = αφ E Iεφ(x) + (1 − αφ )E I zφ(x) h n M h zÛφ(x) = 1 − h β + γsgn(M εÛφ ) εÛφ(x) M
(3)
c
117
where
118
moment, Mch = (1 − αφ )Mp is the hysteretic bending capacity, Mp is an estimated
Mh
= (1 − αφ )E I zφ represents the hysteretic component of the bending
119
plastic moment capacity of the section, E is the elastic modulus, I is the moment
120
of inertia, zφ is the hysteretic bending deformation, and αφ is the post-elastic to
121
elastic stiffness ratio for bending. Similar equations are also available for N-εu ,
122
and Q-εγ , which are omitted here for brevity. Combining the expressions for axial
123
force, bending moment and shear force, results in the following: e h P(x) = P(x) + P(x) = De ε(x) + HD z(x)
124
(4)
P = {N Q M }T ; ε = {εu εγ εφ }T ; z = {zu zγ zφ }T D=
"
EA
G As
EI
#
;
α=
"
αu
αγ αφ
#
;
De = αD
(5)
HD = (I − α)D
125
where I is the identity matrix, De is the elastic rigidity matrix, HD is the hysteretic
126
rigidity matrix, zu and zγ are the hysteretic axial and shear deformations, respec-
127
tively, A is the cross-section area, As is the effective shear area, G is the shear
128
modulus, αu and αγ are the axial and shear hardening parameters, respectively.
129
2.2. Multiaxial model with interaction
130
In the previous subsection, the axial force, moment and shear force have been
131
assumed to evolve following independent uniaxial plasticity laws, such that the 7
132
individual capacities are not constrained by a multiaxial yield/capacity surface.
133
If needed, the evolution laws can be modified to incorporate axial-moment-shear
134
interactions as follows [6]:
n zÛ (x) = (I − H1 H2 R) εÛ (x) ; H1 = Φ(Ph ) + 1 Ph = {N h Q h M h }T ; H2 = β + γsgn (Ph )T εÛ " T # −1 " T # ∂Φ ∂Φ ∂Φ ∂Φ D D R= ∂Ph ∂Ph ∂Ph ∂Ph
(6)
135
where H1 is a smooth function ranging in [0,1] corresponding to the elastic and
136
plastic regions respectively, H2 is a Heaviside function, Ph is the hysteretic com-
137
ponent of the axial force, shear force, bending moment, and Φ(Ph ) is the specified
138
yield/capacity function. The expression for Φ(Ph ) can be specified from pre-
139
defined yield/capacity functions, depending on the material and cross-sectional
140
properties of the beam element, e.g. [33] among many others. It should be also
141
noted, however, that an expression for Φ(Ph ) can also be numerically derived using
142
fiber analysis, [34] for example, and the function H1 and the matrix R are then
143
evaluated accordingly. Similarly, axial-moment interaction can be incorporated by
144
modifying the evolution equations as follows: εÛu
zÛu
= (I2×2 − H1 H2 R) εÛφ zÛφ (x) (x) h n Q h zÛγ(x) = 1 − h β + γsgn(Q εÛγ ) εÛγ(x) Qc
(7)
145
where functions H1 , H2 and the interaction matrix R are dependent on N h and M h
146
now, whereas Q h evolves independently based on the hysteretic shear capacity Q ch .
8
147
3. Suggested hysteretic finite element formulation
148
Friedman and Kosmatka [35] derived transverse displacement and rotation inter-
149
polation functions for linear Timoshenko beam elements that exactly satisfy the
150
differential equations associated with Timoshenko’s beam theory. These interpo-
151
lation functions, also used in Triantafyllou and Koumousis [6], are not entirely
152
compatible with the hysteretic formulations due to the additional hysteretic defor-
153
mation components z. Therefore in the present work, new consistent interpolation
154
functions are explicitly derived for a two-noded beam element, that comply with
155
the equilibrium equations of homogeneous nonlinear Timoshenko beam elements.
156
Six additional DOF are considered to account for hysteretic axial, bending and
157
shear deformations, both at the start (Node 1) and the end nodes (Node 2) of the
158
beam element. Hence, the nodal displacement and hysteretic DOF vectors can be
159
expressed as: d = {u1 w1 θ 1 u2 w2 θ 2 }T ; z = {zu1 zγ1 zφ1 zu2 zγ2 zφ2 }T
(8)
160
where d is the nodal displacement vector in local coordinates, consisting of longi-
161
tudinal displacement u, transverse displacement w, and rotation θ for both nodes,
162
and z represents the hysteretic deformation vector or else the six additional DOF,
163
each corresponding to their respective strains at the two nodes. To account for
164
distributed plasticity over the length of the beam element, compatible interpolation
165
functions are also suggested for z. A detailed description is presented in the sub-
166
sequent subsections, starting from the derivation of the interpolation functions for
167
displacement and hysteretic deformation, to the final formulation of the element
168
matrices using the variational principle of virtual work.
9
169
3.1. Beam kinematics and equilibrium equations
170
Timoshenko beam theory assumes that in the deformed configuration the cross
171
sections of a beam element remain plane but not necessarily perpendicular to the
172
longitudinal axis. Hence, for a 2D homogeneous beam element of length L in a
173
x − y plane, as in Fig. 1, the kinematic relations are expressed as: εu(x) =
∂u(x) ∂w(x) ∂θ (x) ; εγ(x) = − θ (x) ; εφ(x) = ∂x ∂x ∂x
(9)
174
In this formulation, only nodal loading is considered for the beam element, resulting
175
in the following equilibrium equations: ∂N(x) =0 ∂x
176
177
where N(x) = αu E Aεu(x) + (1 − αu )E Azu(x)
∂Q(x) = 0 where Q(x) = αγ G As εγ(x) + (1 − αγ )G As zγ(x) ∂x ∂ M(x) + Q(x) = 0 where M(x) = αφ E Iεφ(x) + (1 − αφ )E I zφ(x) ∂x
(10) (11) (12)
178
Eqs. (9) to (12) represent the kinematic conditions and the fundamental equilibrium
179
equations of the hysteretic Timoshenko beam element, without any nonlinear
180
geometric effects.
181
3.2. Consistent interpolation functions with distributed nonlinearity
182
In order to satisfy the equilibrium conditions in Eqs. (10) to (12), εu(x) , εγ(x) , zu(x)
183
and zγ(x) are assumed to be constant, whereas εφ(x) and zφ(x) are considered to
184
vary linearly along the length of the element. Therefore, based on the hysteretic
185
boundary conditions in Eq. (8), the following hysteretic axial, shear and bending
10
∂w ∂x εγ θ
y
y
∂w ∂x
w Q,w
M,θ N,u
x
x
(a)
(b)
Fig. and (b) (b) deformed deformed configuration. configuration. Fig.1.1. Beam Beamelement: element: (a) (a) undeformed undeformed configuration, configuration, and 104 186
187
188
3.2. Consistentareinterpolation functions with distributed nonlinearity deformations obtained at any distance x along the element: In order to satisfy the equilibrium conditions 1 1in Eqs. (10) to (12), εu(x) , εγ(x) , zu(x) zu(x) = zu1 + zu2 2 whereas 2 εφ(x) and zφ(x) are considered to and zγ(x) are assumed to be constant, 1 1 (13) zγ1 + zγ2 zγ(x) of = the element. vary linearly along the length Therefore, based on the hysteretic 2 2 following x boundary conditions in Eq. (8), the hysteretic axial, shear and bending x zφ(x) = 1 − zφ1 + zφ2 L x alongLthe element: deformations are obtained at any distance By substituting Eqs. (9) and (13) inEqs. (10) to (12), the following fundamental 1 1 zu(x)hysteretic = zu1Timoshenko + zu2 beam element are obtained: differential equations for the 2 2 ! ! 1∂ 2 w 1 (13) ∂ 2 u(x) ∂θ (x) zγ1 + zγ2(x) zγ(x) = = 0 ; α G A − = 0 αu E A 2 2 γ s ∂x ∂ x2 ∂ x 2 x x + z zφ2 φ1 − ∂ 2 θ (x)zφ(x) = 1 − L zzφ2 φ1 (14) L + αφ E I + (1 − αφ )E I 2 L ∂x ∂w(x) zγ2 + zγ1 αγ G As − θ (x) + (1 − αγ )G As =0 ∂x 2
11
10
189
Based on the typical beam element assumptions, the displacement fields along the
190
element are assumed as: u(x) = B0 + B1 x w(x) = C0 + C1 x + C2 x 2 + C3 x 3
(15)
θ (x) = D0 + D1 x + D2 x 2 191
where the coefficients Bi (i ∈ {0, 1}), C j ( j ∈ {0, 1, 2, 3}) and D k (k ∈ {0, 1, 2})
192
are derived using appropriate displacement boundary conditions in Eq. (8) and
193
satisfying Eq. (14). Therefore, the overall displacement and rotational fields can
194
be expressed as: N1 0 0 N2 0 u 0 = 0 N3 N4 0 N5 N6 d w θ 0 N7 N8 0 N9 N10 (x) 0 0 0 0 0 0 + 0 N11 N12 0 N13 N14 z 0 N15 N16 0 N17 N18 Nu HN u = Nw d + HN w z Nθ HN θ
12
(16)
195
196
where the derived displacement interpolation functions, Ni (i ∈ {1, 2, ... , 18} ), are provided in Eq. (17):
N1 = 1 − Lx ; N2 = Lx 3α 12α λ 2α N3 = L 3γ x 3 − L 2γ x 2 − Lφ x µ0 + 1 α 2(α +3α λ) N4 = Lγ2 x 3 − γ L φ x 2 + (αγ + 6αφ λ)x µ0 3α 12α λ 2α N5 = − L 3γ x 3 + L 2γ x 2 + Lφ x µ0 α (α −6α λ) N6 = Lγ2 x 3 − γ L φ x 2 − 6αφ λx µ0 6 2 6 N7 = L 3 x − L 2 x αγ µ0 3α 4(α +3α λ) N8 = L 2γ x 2 − γ L φ x µ0 + 1 6 2 6 N9 = − L 3 x + L 2 x αγ µ0 3α 2(α −6α λ) N10 = L 2γ x 2 − γ L φ x µ0 3 2 x − 12 x (1 − αγ )µ0 N11 = N13 = − L12 x 3 + 2L N12 = L2 x 3 − 3x 2 + L x (1 − αφ )µ0λ N14 = − L2 x 3 + 3x 2 − L x (1 − αφ )µ0λ 3 2 3 N15 = N17 = − L 2 x + L x (1 − αγ )µ0 N16 = L6 x 2 − 6x (1 − αφ )µ0λ 6 2 N18 = − L x + 6x (1 − αφ )µ0λ
and µ0 = 197
1 αγ +12αφ λ
; λ=
(17)
EI G As L 2
Similar to the linear Timoshenko beam element case, functions N1 − N10 map the
198
conventional nodal displacement and rotational DOF to the displacement fields.
199
To satisfy the equilibrium conditions and obtain consistent beam elements, the
200
additional functions, N11 − N18 , are also needed, connecting the displacement
201
202
fields to the introduced nodal hysteretic DOF as well.
Although the presented interpolation functions in Eq. (17) are particularly 13
203
derived for a fully nonlinear Timoshenko element, they are sufficiently general,
204
having the notable property of being entirely compatible to all other simpler cases
205
by an appropriate and simple choice of parameter values. For example, for the case
206
of linear beam elements, when αu = αγ = αφ = 1, the developed expressions for
207
N1 − N10 revert to the interpolation functions suggested by Friedman and Kosmatka
208
[35], and the hysteretic components of the interpolation functions, N11 − N18 ,
209
become equal to zero, nullifying the hysteretic DOF contributions. Similarly,
210
for the case of inelastic Euler-Bernoulli beam element, by specifying λ = 0 and
211
αγ = 1, the suggested interpolation functions coincide with the typically used ones
212
for these elements, as also proposed by Triantafyllou and Koumousis [5].
213
214
Similar to the displacement fields, the hysteretic deformation fields from Eq. (13), can be expressed in matrix form as: z(x)
1 0 z 2 u = zγ = 0 12 zφ (x) 0 0
1 0 2 0 0 0 0 12 0 z = Nz(x) z 1 − Lx 0 0 Lx
(18)
215
where Nz is the hysteretic interpolation matrix, with each element representing a
216
hysteretic interpolation function. From the derived displacement and hysteretic
217
interpolation functions, as shown in Eqs. (16) to (18), generalized strain fields
218
can be obtained at any distance x along the beam element, using the kinematic
219
relations of Eq. (9). Substituting Eq. (16) to Eq. (9), results in the following
14
220
strain-displacement relations:
ε(x)
221
222
where B(x)
∂u ε u ∂x ∂w = εγ = ∂x − θ εφ ∂θ (x) ∂ x ∂HNu ∂Nu ∂N ∂ x ∂H ∂ x Nw w = ∂ x − Nθ d + − H N θ z = B(x) d + HB(x) z ∂ x ∂HNθ ∂Nθ ∂x ∂x
(19)
T − 1 0 0 L 6αγ µ0 (L−2x) 12αφ λµ0 0 − L − 3 L 0 −6α λµ0 6αγ µ0 x − 4µ0(αγ +3λαφ ) φ 2 L L B(x) = 1 (20) L 0 0 6αγ µ0 (L−2x) 0 12αφ λµ0 L L 30 0 0 −6αφ λµ0 6αγ µ2 x − 2µ (αγL−6λαφ ) L 0 T 0 0 (1−αγ )µ0 3µ0 (1−αγ )(L−2x) 0 − 2 L2 0 0 (1 − α )λµ0 L − 6λµ (1−αφ )(L−2x) φ L (21) HB(x) = 0 0 0 (1−α )µ0 3µ0 (1−αγ )(L−2x) 0 − 2γ 2 L 6λµ0 (1−αφ )(L−2x) 0 0 −(1 − αφ )λµ L L and HB(x) are the linear and hysteretic strain matrices, respectively.
223
Similar to the discussion about the displacement interpolation fields, the strain
224
fields are also generally applicable. For example, the second row elements of
225
matrices B(x) and HB(x) in Eqs. (20) and (21), representing shear strain, become
226
zero for an Euler-Bernoulli beam case (αγ = 1 and λ = 0).
227
Substituting the strain and hysteretic deformation fields expressions from
228
Eqs. (18) and (19) to Eq. (4), the internal force vector of the beam element
15
229
can be reformulated as: P(x) = De ε(x) + HD z(x) = De (B(x) d + HB(x) z) + HD (Nz(x) z) ∴ P(x) = (De B(x) )d + (De HB(x) + HD Nz(x) )z
(22)
230
3.3. Derivation of element matrices
231
The stiffness and hysteretic matrices of the beam element are formulated in this
232
section by means of the virtual work principle [31], which can be written as: ∫ L T δd F = N(x) δεu(x) + Q(x) δεγ(x) + M(x) δεφ(x) dx (23) 0
233
where δd is the virtual nodal displacement vector, F is external nodal load, and
234
(δεu(x), δεγ(x), δεφ(x) ) are the relevant internal virtual generalized strains at any
235
distance x. By substituting the kinematic conditions of Eq. (9) to Eq. (23), the
236
237
following relations are obtained: ∫ L ∂w ∂u T + Q(x) δ −θ δd F = N(x) δ ∂x ∂x 0 ∂θ + M(x) δ dx ∂x ∫ L ∂(δw) ∂(δu) T δd F = + Q(x) N(x) ∂x ∂x 0 ∂(δθ) − Q(x) δθ dx + M(x) ∂x Integration by parts, results in: ∫ L L ∂N(x) T δd F = N(x) (δu) 0 − δu dx ∂x 0 ∫ L L ∂Q(x) + Q(x) (δw) 0 − δw dx ∂x 0 ∫ L ∫ L L ∂ M(x) + M(x) (δθ) 0 − δθ dx − Q(x) δθ dx ∂x 0 0 16
(24)
(25)
238
By imposing the integration limits at nodes 1 and 2, and considering Eqs. (10)
239
to (12), Eq. (25) becomes: δdT F = −δu1 N1 − δw1 Q1 − δθ 1 M1 + δu2 N2 + δw2 Q2 + δθ 2 M2
(26)
240
Hence, equilibrium is satisfied in a strong sense, such that the internal nodal forces
241
are exactly equal to the external ones, since the present formulation is based on
242
the exact equilibrium equations of the nonlinear beam element. Eq. (26) can be
243
further expressed as: δd F = δd T
T
−P1 P2
where
P1 = {N1 Q1 M1 }T P2 = {N2 Q2 M2 }T
(27)
244
Substituting the internal force vector expression obtained in Eq. (22) to Eq. (27)
245
and considering the nodal boundary conditions, results in: −P1
−De B(x=0) d+ = F= De B(x=L) P2 = Kd + Hz
−De HB (x=0) − HD Nz(x=0) z De HB (x=L) + HD Nz(x=L)
(28)
246
Eq. (28) represents the final concise force-displacement expression in local co-
247
ordinates for the hysteretic Timoshenko beam formulation. Consistent element
248
stiffness matrix K and hysteretic matrix H are obtained by substituting in Eq. (28)
249
the rigidity matrices De and HD from Eq. (5), the strain matrices B and HB from
250
Eqs. (20) and (21), and the hysteretic interpolation matrix Nz from Eq. (18),
251
resulting in:
17
αu AE 0 0 − αuLAE 0 0 L 6ψ1 6ψ1 12ψ1 1 0 0 − 12ψ L3 L2 L3 L2 0 6ψ1 4ψ2 2ψ3 1 0 − 6ψ 2 L L L2 K = α AE L αu AE − uL 0 0 0 0 L 0 − 12ψ1 − 6ψ1 0 12ψ1 6ψ1 − 3 2 3 2 L L L L 2ψ 6ψ 4ψ 6ψ 3 1 2 1 0 0 − L2 2 L L L 0 where ψ1 = αφ αγ µ E I ;
(29)
ψ2 = αφ (αγ + 3αφ λ)µ0 E I ; ψ3 = αφ (αγ − 6αφ λ)µ0 E I −h 0 0 u h φ αγ 6hγ 0 − L2 − L 0 − 3hγ −h (α + 6α λ) φ γ φ L H= hu 0 0 h φ αγ 0 6hγ L L2 0 − 3hLγ 6hφ αφ λ where hu = (1−α2u )AE ;
−hu
0 6hγ L2 3hγ − L
0 −
0
hu 0 0
0
6hγ L2 3h − Lγ
h φ αγ L −6hφ αφ λ 0 h φ αγ − L hφ (αγ + 6αφ λ) 0
(30)
hφ = (1 − αφ )µ0 E I ; hγ = (1 − αγ )µ0 E Iαφ
252
3.4. Generality of the beam element model
253
In relation to the previous discussion about the generality of the suggested interpo-
254
lation functions, the derived stiffness and hysteretic element matrices in Eqs. (29)
255
and (30) are applicable for both Euler-Bernoulli and Timoshenko beam element
256
formulations, and for both linear and nonlinear behaviors. For a pure elastic case
257
(αu = αγ = αφ = 1) for example, the hysteretic matrix H in Eq. (30) becomes
258
0, similar to what was mentioned earlier for the hysteretic interpolation functions,
259
and the elastic stiffness matrix K in Eq. (29) reverts to the typical stiffness matrix
260
of an elastic Timoshenko beam [35]. In addition, for λ = 0 and αγ = 1, the beam
261
matrices take the proper form for a nonlinear Euler-Bernoulli beam element [5], 18
262
and by combining the two cases, i.e. for λ = 0 and αu = αγ = αφ = 1, the conven-
263
tional stiffness matrix of a linear Euler-Bernoulli beam element is obtained [31].
264
The resulting stiffness and hysteretic matrices for each of these common cases are
265
shown in the Appendix. Evidently, other cases, combining linear and nonlinear
266
components, are immediately and readily possible with this parametrized form,
267
by merely changing the parameter values, and without altering anything in the
268
modeling of the problem or its solution process.
269
3.5. Numerical solution process
270
Eq. (28) can be interpreted as the conventional stiffness based formulation, aside
271
from the additional hysteretic DOF vector, z. Matrices K and H, and the vectors
272
d and F in Eq. (28) correspond to local element coordinates. Applying coordinate
273
transformation, the global force vector, Fg , can be obtained as: Fg = ΛT KΛdg + ΛT Hz
(Since d = Λdg ; Fg = ΛT F)
(31)
274
where Λ is the typical transformation matrix from global to local coordinate
275
system, and dg corresponds to the element nodal displacement DOF vector in
276
global coordinates [31]. Eq. (31) is thus expressed in terms of global matrices as: Fg = Kg dg + Hg z ; Kg = ΛT KΛ ;
Hg = ΛT H
(32)
277
where Kg and Hg are the element stiffness and hysteretic matrices in global
278
coordinates, with the former comprising the symmetric and positive definite system
279
stiffness matrix, Ks , by the typical direct stiffness matrix assembly techniques [31],
280
and the latter forming the system hysteretic matrix, Hs , by consistently appending
281
the Hg matrices for all the elements. The overall system force-displacement
282
expression can be then expressed as: 19
283
284
285
z(el=1) . . Fs = Ks ds + Hs zs ; zs = (33) . z (el=Nel ) where Fs and ds are the system nodal force and displacement DOF vectors respectively, z(el= j) is the hysteretic DOF vector of size 6 × 1 for the j th element
( j = {1, 2, · · · , Nel }), where Nel is the total number of elements in the system, and
286
zs is a vector consisting of the hysteretic DOF for all the elements. For example, in
287
a 2-dimensional problem of a cantilever beam, subjected to a concentrated external
288
force at its free end and discretized into 2-elements (Nel = 2), the size of known
289
external force vector, Fs , is (6 × 1), while the unknown displacement, ds , and
290
hysteretic DOF, zs , vectors are of size (6 × 1) and (12 × 1) respectively. Based
291
on Eq. (6), the evolution equation for the j th beam element, of length L j , can be
292
expressed as: zÛ (el= j) = Ω(el= j) εÛ (el= j) where, Ω(el= j)
(I − H1 H2 R)(x=0) 0 = (I − H1 H2 R)(x=L j ) 0 (el= j)
(34)
z(el= j) = {zu1 zγ1 zφ1 zu2 zγ2 zφ1 }T(el= j)
ε (el= j) = {εu1 εγ1 εφ1 εu2 εγ2 εφ2 }T(el= j)
293
and 0 is a square null matrix of size 3, function H1 and matrix R are shown in
294
Eq. (6). For numerical implementation purposes in quasi-static problems, function
295
H2 is now given as H2 ≈ β + γsgn((Ph )T (ε − ε0 )), which is slightly different from
296
Eq. (6), where ε0 is the initial strain at the beginning of each solution step. Eqs. (33)
297
and (34) have to be solved simultaneously to obtain the solutions, which results in 20
298
a system of DAEs.
299
The formulated DAEs can be solved using a Newton solution scheme, as
300
explained in more detail subsequently. However, in this paper a new, alternative
301
and more efficient approach is suggested for solving the resulting DAEs. The main
302
idea is to substitute the displacement expression from Eq. (33) to the evolution
303
equation of Eq. (34), so that the resulting expressions will only and simply be an
304
explicit first order ODE system with respect to the vector zs . Therefore, Eq. (33)
305
is reformulated as: −1 ds = K−1 s Fs − Ks Hs zs
(35)
306
and based on Eq. (19) the vector of strains for all the elements of the system, εs ,
307
referred to herein as system strain vector, is expressed in terms of the system DOF
308
as follows: (Bd + HB z)(el=1) ε(el=1) . .. .. εs = = . ε (el=Nel ) (Bd + HB z)(el=Nel ) (HB Lz )(el=1) zs (BΛLd )(el=1) ds + . .. = (BΛL ) d (el=Nel ) ds + (HB Lz )(el=Nel ) zs
(36)
= Bs ds + HBs zs
309
where Ld and Lz are the incidence (or connectivity) matrices for transforming
310
dg and z vectors, for each element, to ds and zs vectors respectively, resulting in
311
d = Λdg = ΛLd ds and z = Lz zs . The elements in the incidence matrices are 1,
312
corresponding to the same DOF in both row and column, and 0 otherwise [31].
313
Substituting now Eq. (35) to Eq. (36), the system strain vector is obtained as: −1 εs = (Bs K−1 s )Fs + (HBs − Bs Ks Hs )zs
21
(37)
314
where all the resulting matrices in the parentheses are constant. As seen in Eq. (37),
315
the strain is now a function of the unknown zs vector and therefore H2 , in Eq. (34),
316
also becomes a function of hysteretic DOF, similar to H1 and R. Consequently,
317
the whole matrix Ω for each element is now a function of hysteretic DOF. Based
318
on Eqs. (34) and (36), the overall evolution equation for the system is expressed as
319
follows:
320
Ω(el=1) Û ε(el=1) . . .. . zÛ s = . Ω(el=Nel ) εÛ (el=Nel ) −1 Û = Ωs εÛ s = Ωs (Bs K−1 zs s )Fs + Ωs (HBs − Bs Ks Hs )Û
(38)
Further simplification of Eq. (38) results in the following ODEs:
−1 zÛ s = (I − Ωs A)−1 Ωs EFÛ s ; A = HBs − Bs K−1 s Hs ; E = Bs Ks
(39)
321
where A and E are constant system matrices and Ωs is a function of zs , as already
322
explained. The ODEs in Eq. (39) are now the only equations needed to solve
323
the problem, having also a reduced number of unknowns, since the displacement
324
DOF have been eliminated from the evolution equations. Eq. (39) can thus be
325
straightforwardly solved now to obtain the hysteretic DOF, using any customary
326
numerical ODE solver, e.g. based on Runge-Kutta methods. As a post-processing
327
step, the resulting zs , at each or from all time steps, can be substituted to Eq. (35),
328
where Ks and Hs are constant, to evaluate the nodal displacement DOF.
329
Alternatively, the suggested DAEs in Eqs. (33) and (34) can be solved using
330
a Newton scheme, by approximating instead the ODE in Eq. (34) in an algebraic
331
form, resulting in the following expressions at the i th step, based also on Eq. (19):
332
Fs (i) = Ks ds (i) + Hs zs (i) z(el= j) (i) = z(el= j) (i − 1) + Ω(el= j) ∆ε(el= j) ; 22
(40) (41)
333
where ∆ε(el= j) = (B(d(i) − d(i − 1)) + HB (z(i) − z(i − 1)))(el= j)
(42)
334
Therefore, by solving Eqs. (40) and (41) by a typical iterative Newton scheme,
335
the unknown nodal displacements, ds , and hysteretic DOF, zs , at step (i) can be
336
obtained.
337
In the present work, both Newton and Runge-Kutta approaches have been
338
implemented, with both methods yielding the same results. The suggested ODEs
339
approach of Eq. (39) using a Runge-Kutta method has been however significantly
340
more computationally efficient in all the examined problems. Finally, for dynamic
341
cases, as also mentioned earlier, the whole formulation can be favorably and
342
straightforwardly described in a state-space ODE form, as in [6], and be again
343
efficiently solved by any generic ODE solver.
344
4. Numerical studies
345
The generality, consistency, versatility and efficiency of the proposed formulation is
346
demonstrated through numerical examples of a cantilever beam model subjected
347
to a concentrated external force at its free end, as shown in Fig. 2. For all
348
the examples, the beam is a W30 × 148 steel beam with yield strength of 380
349
MPa, elastic modulus of 200 GPa and shear modulus of 77 GPa. Based on the
350
cross section properties and yield strength, the moment and shear capacities are
351
calculated to be 3,100 kNm (Mp ) and 2,800 kN (Q p ), respectively. In addition,
352
other considered model parameters are αγ = αφ = 0.01, β = γ = 0.5, and n = 2,
353
for the Timoshenko case. The Euler-Bernoulli formulation is simply achieved by
354
setting the model parameters λ and αγ to 0 and 1, respectively, without any other
355
changes. The formulation is implemented in MATLAB and the solution schemes 23
described in Section 3.5 are employed. L x
W30x148
F
Fig. 2. Cantilever beam subjected to tip force. 356
357
4.1. Shear locking elimination
358
To assess the performance of the suggested formulation in relation to shear locking,
359
the responses of the cantilever beam over a broad range of scenarios are obtained
360
and compared. Since the shear deformation of a long slender beam is negligible
361
as compared to its flexural deformation, the solution obtained from a Timoshenko
362
formulation should converge to that obtained from an Euler-Bernoulli formula-
363
tion. However, this is not often achieved, mostly due to inconsistent interpolation
364
functions, providing reasonable results for short deep beams but overly stiff results
365
for long slender beams, a phenomenon known as shear locking. The suggested
366
formulation provides consistent interpolation functions and results naturally con-
367
verge toward the expected behavior of an Euler-Bernoulli formulation, even with
368
one element and without the need to resort to any approximate methods.
369
To demonstrate this point, the response of the cantilever beam shown in Fig. 2
370
is compared using both Timoshenko and Euler-Bernoulli formulations for different
371
length-to-depth (L/d) ratios, achieved by varying the length of the cantilever beam,
372
ranging from 0.4 m to 10.0 m. For all scenarios, the tip displacement response is
373
obtained when the beam is subjected to a static force of 100 kN at the free end. The 24
374
results are shown in Fig. 3, where y-axis shows the normalized tip displacement
375
with respect to the corresponding Euler-Bernoulli deflections, and x-axis reports
376
the L/d beam ratio. As seen in the figure, for small L/d ratios, the displacements
377
obtained from the Timoshenko formulation are substantially larger than the Euler-
378
Bernoulli case, suggesting significant shear deformations. However, for large L/d
379
values, the Timoshenko response gradually converges to the Euler-Bernoulli one. 12 Euler Beam Timoshenko Beam
10
w/wEuler
8 6 4 2 0 0
5
10
15
L/d Fig. 3. Difference between Euler-Bernoulli and Timoshenko beam formulation.
380
4.2. Timoshenko vs Euler-Bernoulli response investigation
381
In this subsection, the nonlinear global response is evaluated when the free end
382
of the cantilever beam is subjected to a monotonically increasing force, for three
383
different cases: (a) shear dominant beam, (b) intermediate beam, and (c) flexure
384
dominant beam. For these cases, different beam lengths and external forces
385
are considered, as shown in Table 1. Both Timoshenko and Euler-Bernoulli 25
386
formulations are tested and the results are compared, as shown in Fig. 4, where
387
x-axis represents the beam rotation, obtained by normalizing the displacement
388
of the free end with respect to the beam length, and y-axis reports the fixed end
389
bending moment of the beam.
390
Fig. 4(a) shows the results for the shear dominant beam with length 0.7 m. The
391
beam is subjected to a maximum tip force of 3,000 kN, which exceeds the plastic
392
shear capacity, while the corresponding maximum bending moment of 2,100 kNm
393
is within the elastic range. As expected, the Euler-Bernoulli formulation is not able
394
to capture the global inelastic behavior of the beam element, since the maximum
395
bending moment is below the plastic capacity. In contrast, the nonlinear behavior
396
is evident in the Timoshenko beam formulation, with an accurate transition from
397
the elastic to the inelastic regimes, corresponding to the shear capacity of the
398
section.
399
Fig. 4(b) shows the moment-rotation response for the intermediate beam with
400
length 1.1 m subjected to a maximum force of 3,000 kN, which corresponds
401
to a moment of 3,300 kNm. From these results, it is evident that the beam is
402
yielding both in shear and bending, hence the term intermediate beam. However,
403
the significant shear deformation of the beam correctly results in reduced stiff-
404
ness and increased rotation for the Timoshenko formulation, as compared to the
405
corresponding Euler-Bernoulli case.
406
Lastly, Fig. 4(c) shows the moment-rotation response of a 5 m long beam
407
subjected to a maximum tip force of 700 kN, which is less than the shear capacity
408
of the section but produces a moment of 3,500 kNm, which exceeds the flexural
409
capacity. Due to this beam geometry, there is negligible shear deformation and
410
the Euler-Bernoulli formulation agrees sufficiently well with the more general 26
Table 1. Different beam elements with Q p =2,800kN and Mp =3,100kNm Type
411
Beam length (L)
Max applied force (F)
Max moment (F L)
Shear dominant beam
0.7 m
3,000 kN (>Q p )
2,100 kNm (
Intermediate beam
1.1 m
3,000 kN (>Q p )
3,300 kNm (>M p )
Flexure dominant beam
5.0 m
700 kN (
3,500 kNm (>M p )
Timoshenko one, that proved reliable in all cases. (a) 0.7 m long beam
3000 2500
Moment (kNm)
Moment (kNm)
(b) 1.1 m long beam
Timoshenko formulation Euler formulation
2000 1500
(c) 5 m long beam
3500
3500
3000
3000
2500
2500
Moment (kNm)
3500
2000 1500
2000 1500
1000
1000
1000
500
500
500
0
0 0
0.005
0.01
0.015
0.02
Tip rotation
0 0
0.01
0.02
0.03
Tip rotation
0.04
0
0.04
0.08
0.12
0.16
Tip rotation
Fig. 4. Comparison of Timoshenko and Euler-Bernoulli beam formulations for: (a) shear dominant, (b) intermediate, and (c) flexure dominant cantilever beams.
412
4.3. Element consistency
413
The Timoshenko formulation is here employed for the same shear dominant,
414
intermediate and flexure dominant beams, now discretized into five elements, to
415
analyze the local beam behavior including shear strain and curvature distributions
416
within the beam. All the relevant example values are shown in Table 1, as before.
417
Figs. 5(a-c) give the response of the shear dominant beam. Fig. 5(a) shows
418
the global moment-rotation response, where points A and B represent indicative
419
points in the elastic and plastic domains, respectively. The local responses of the 27
420
beam element are presented in Figs. 5(b) and (c). The curvature for both points is
421
shown in Fig. 5(b) and the shear strain is reported in Fig. 5(c). Figs. 5(b-c) show a
422
slight curvature change but a large shear strain variation between points A and B,
423
verifying the shear dominant behavior at the local level.
424
Similarly, Fig. 5(d) shows the global moment-rotation response of the interme-
425
diate beam, whereas Figs. 5(e) and (f) show the local curvature and shear strain
426
plots, respectively, for points C and D. Significant changes in both curvature and
427
shear strain responses are observed from C to D, as expected, since the bending mo-
428
ment and shear force are both exceeding their yield capacities for the intermediate
429
beam.
430
Lastly, Figs. 5(g-i) correspond to the flexure dominant beam, such that Fig. 5(g)
431
displays the global response and Figs. 5(h-i) the local responses at points E and F.
432
As expected, these figures indicate slight shear strain change and large curvature
433
variation from E to F, correctly suggesting a flexure dominant response.
434
From all the curvature and shear strain plots, namely Figs. 5(b-c), (e-f) and
435
(h-i), model consistency in terms of nodal strain field continuity is observed for
436
both linear and nonlinear regimes. Thus, the suggested formulation is compatible
437
with all theoretical assumptions and produces consistent results at both global and
438
local levels for the entire range of beam behaviors.
439
4.4. Discretization and convergence
440
In this subsection, to further investigate the beam response both at the global and
441
local levels, a convergence study for the Timoshenko formulation is conducted
442
by discretizing the beam into different number of elements. In this illustrative
443
example, only the intermediate beam is considered and the beam is discretized
444
into 1, 5, 10 and 30 elements. Fig. 6(a) shows the global response in terms of 28
4000
0.08
(a-c): 0.7 m long beam response (b)
0.06 B
2000
1000
0 (c)
A B
A B
-0.01
Shear strain
3000
Curvature
Moment (kNm)
(a)
0.04
0.02
-0.02 -0.03 -0.04
A 0
0 0
0.005
0.01
0.015
0.02
0.025
-0.05 1
2
3
4
5
1
2
Nodes
Tip rotation 4000
0.08
(e)
1000
5
(f)
6
C D
-0.01
Shear strain
Curvature
0.06
2000
4
0
C D
D 3000
3
Nodes
(d-f): 1.1 m long beam response
(d)
Moment (kNm)
6
0.04
0.02
-0.02 -0.03 -0.04
C 0
0 0
0.01
0.02
0.03
-0.05 1
2
4000
5
6
1
2
0.06
2000
1000
4
5
6
0 (i)
E F
E F
-0.01
Shear strain
3000
3
Nodes
(h)
F
Curvature
Moment (kNm)
4
(g-i): 5 m long beam response
0.08 (g)
3
Nodes
Tip rotation
0.04
0.02
-0.02 -0.03 -0.04
E 0
0 0
0.02
0.04
Tip rotation
0.06
-0.05 1
2
3
4
Nodes
5
6
1
2
3
4
5
6
Nodes
Fig. 5. Strain continuity at the nodes considering 5 elements for: (a-c) shear dominant, (d-f) intermediate, and (g-i) flexure dominant cantilever beams.
29
445
support moment and beam rotation, and Fig. 6(b) represents the local curvature
446
response of the beam element corresponding to the peak moment of 3,300 kNm.
447
For both the global and local responses, convergence is nearly achieved with
448
10 elements. The local curvature response also indicates that by increasing the
449
number of elements a proper continuous and smooth curvature distribution along
450
the length of the beam element is obtained. As can be seen in Fig. 6(b), the
451
estimated plastic hinge length, i.e. the beam length with curvature values greater
452
than the yield curvature, has converged with 10 elements. For this example, the
453
yield curvature is given by My /E I, with My = 2, 700 kNm. Hence, the developed
454
formulation is able to provide consistent, continuous, converged and smooth strain
455
fields. (a)
3500
(b)
0.05 Nel=1 Nel=5 Nel=10 Nel=30 Yield
3000 0.04
Curvature (1/m)
Moment (kNm)
2500 2000 1500 1000
Nel=1 Nel=5 Nel=10 Nel=30
500
0.03
0.02
0.01
0
0 0
0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04
0
Tip rotation
0.2 L
0.4 L
0.6 L
0.8 L
L
Beam length
Fig. 6. Element discretization and convergence analysis for intermediate beam.
456
4.5. Combined axial-flexural loading
457
The effect of interaction is shown in this subsection by subjecting the beam element
458
to combined axial and transverse forces. As seen in the previous subsection, 30
459
convergence is achieved for 10 elements; therefore, only the converged case of the
460
intermediate beam with the Timoshenko formulation is presented here, with the
461
same monotonically increasing transverse force, as in the previous subsections,
462
and an additional constant axial force, applied at the free end of the beam element.
463
Four cases are considered, based on the magnitude of the applied axial force, i.e.
464
np = 0, 0.2, 0.3, 0.4, where np is the ratio of the applied axial force, N, to the
465
axial capacity of the cross-section, Np = 10, 600 kN. Axial-moment interaction is
466
considered using the Orbison criterion [33] and the results are presented in Fig. 7.
467
Fig. 7(a) presents the global moment-rotation response, whereas Fig. 7(b) shows
468
the local moment-curvature behavior at the fixed end. A reduction in only flexural
469
capacity is observed in these figures with the increasing magnitude of axial force
470
due to the axial-moment interaction, while the shear capacity remains unchanged for all the cases. (a) Global response
3000
3000
2500
2500
2000 1500
np= 0 1000
2000 1500 1000
np= 0.2 np= 0.3
500
(b) Local response
3500
Moment (kNm)
Moment (kNm)
3500
500
np= 0.4
0
0 0
0.01
0.02
0.03
0.04
0.05
0.06
0
Tip rotation
0.05
0.1
0.15
0.2
Curvature (1/m)
Fig. 7. Effect of axial-moment interaction for intermediate beam with Timoshenko formulation. 471
472
To further discuss and connect all the presented analyses in this section to31
473
gether, the L/d ratios of the analyzed shear dominant, intermediate and flexure
474
dominant beams in Fig. 3 are obtained as 0.9, 1.4 and 6.4, respectively. As ex-
475
pected, Fig. 3 shows a large variation between Euler-Bernoulli and Timoshenko
476
formulation responses for the L/d ratio of 0.9, deviation decreases for the L/d
477
ratio of 1.4, whereas identical responses are observed for the 6.4 L/d ratio. In
478
Fig. 4, only these three representative beam cases are considered with monotoni-
479
cally increasing load, in contrast to Fig. 3, to compare the overall global response
480
of Timoshenko and Euler-Bernoulli formulations for both elastic and inelastic
481
regimes. The global behavior of Figs. 4(a-c) correspond to Fig. 5(a), Fig. 5(d)
482
and Fig. 5(g), respectively, and due to different element discretizations, global
483
response variations are observed between the two figures. Note that for the shear
484
dominant beam, the global responses in Figs. 4 and 5 are alike, the deviation is
485
increasing for the intermediate beam, and a large variation is observed for the
486
flexure dominant beam. This is due to the fact that the constant shear strain field in
487
this case is equally captured by either one or five elements, however, as the bending
488
contribution increases, the curvature response is more accurately computed with
489
larger number of elements. To further analyze this, in Fig. 6, the intermediate beam
490
is discretized with different number of elements and the responses are compared
491
for both the global moment-rotation and the local curvature responses. Lastly, to
492
emphasize the effect of interaction, the converged case of Fig. 6 is employed and
493
presented in Fig. 7, with combined axial-flexural loading.
494
5. Experimental validation
495
To validate the suggested formulation, the simulated responses are compared
496
with the reported experimental beam test results from Berman and Bruneau [36]. 32
497
Berman and Bruneau [36] carried out extensive testing on twelve beam specimens,
498
consisting of four different lengths with three cross sections, observing varied
499
responses, ranging from shear to flexure dominant behaviors. In the present study,
500
two such beams are simulated for validation purposes: a shear dominant beam
501
with the shortest length, where the total beam rotation is largely governed by the
502
shear strain, and a flexure dominant beam with the longest length, such that the
503
beam deformation is predominantly due to flexure. The two ends of the beams are
504
restrained in the experimental setup, except for the transverse displacement at one
505
end, where the beam is subjected to a typical cyclic loading protocol with increasing
506
amplitude. To account for some observed flexibility at the end restraints of the
507
beams, linear rotational springs are considered at the beam ends in the analysis.
508
Further details about the experimental setup, cross sectional geometry, loading
509
protocol, etc. can be seen in [36].
510
5.1. Shear dominant beam
511
The shear dominant beam response is simulated using both the suggested Timo-
512
shenko and Euler-Bernoulli formulations. Specimen X3L1.2 in [36] is considered
513
with a length of 578 mm, web yield strength of 430 MPa, flange yield strength
514
of 370 MPa, elastic modulus of 200 GPa and shear modulus of 77 GPa. From
515
the beam mechanical properties, the moment and shear capacities are specified to
516
be as 310 kNm (Mp ) and 1,000 kN (Q p ), respectively. Other considered model
517
parameters, without any inverse method utilization, are αφ = 0.01, αγ = 0.01,
518
β = γ = 0.5, and n = 1, for the Timoshenko case. Fig. 8(a) presents the global
519
force-rotation response, where rotation is the ratio of transverse displacement at
520
the free end to the effective length of the member. Figs. 8(b-c) show the local
521
moment-curvature and shear force-shear strain responses at the fixed end, respec33
522
tively. Experimental curvature is calculated by taking the ratio of the difference
523
between the top flange and bottom flange strains over the total depth of the section,
524
while experimental shear strains are obtained from the strain gauges installed at
525
the mid-section and spanning at 45 degrees from the longitudinal axis. Similarly,
526
Figs. 8(d-f) present the Euler-Bernoulli response compared to the same experi-
527
mental data, where Fig. 8(d) shows the global behavior and Figs. 8(e-f) the local
528
cross section responses.
529
As seen in Figs. 8(a-c), the response of the beam is well represented by the
530
Timoshenko formulation. In contrast, using the same parameters, the obtained
531
responses from the Euler-Bernoulli formulation show significant differences with
532
the experimental data in Figs. 8(d-f). It is however interesting to note that the global
533
behavior in Fig. 8(d) approximates the response to some level of accuracy, but due
534
to the limitation of the Euler-Bernoulli formulation to capture shear deformations,
535
large curvature deviation is obtained, as compared to the experimental values.
536
This example, hence, demonstrates the suggested Timoshenko formulation
537
ability to effectively simulate the beam behavior both at the global and local levels
538
with increased accuracy. In the next validation, the same process is repeated for
539
the flexure dominant beam to demonstrate the overall model robustness.
540
5.2. Flexure dominant beam
541
In this example, the X1L3.0 flexure dominant beam specimen in [36] is simulated,
542
with a length of 1,441 mm, web yield strength of 480 MPa, flange yield strength
543
of 371 MPa, elastic modulus of 200 GPa and shear modulus of 77 GPa. Similar
544
to the previous example, model parameters for bending and shear capacities are
545
calculated as Mp = 302 kNm and Q p = 936 kN, respectively. Other used param-
546
eters, without any inverse method utilization again, are αφ = 0.02, αγ = 0.02, 34
Experiment
500
Moment (kNm)
Force (kN)
750
0
-750
-1500 -0.1
250
0
-250
0
0.05
0.1
-0.2
(d)Euler:Global
-750
250
0
-500 0.05
-750
-0.05
0.1
Total rotation (rad)
-0.7
0
0
0.05
0.1
(f)Euler:Local
1500
-250
0
0
Shear strain
Shear force (kN)
0
-0.05
750
-1500 -0.1
0.2
(e)Euler:Local
500
Moment (kNm)
Force (kN)
0
(c)Timoshenko:Local
Curvature (1/m)
750
-1500 -0.1
1500
-500 -0.05
Total rotation (rad)
1500
Hysteretic finite element
(b)Timoshenko:Local
Shear force (kN)
1500
(a)Timoshenko:Global
0.7
750
0
-750
-1500 -0.1
-0.05
Curvature (1/m)
0
0.05
0.1
Shear strain
Fig. 8. Comparison of experimental and simulated response for shear dominant beam.
547
β = γ = 0.5, n = 1, for the Timoshenko formulation. For the Euler-Bernoulli
548
formulation all the parameters again remain the same, except for λ and αγ , which
549
are set to 0 and 1, respectively.
550
In Fig. 9, the Euler-Bernoulli and Timoshenko formulation responses agree
551
well, both at the global and local levels. It is also interesting to note that the
552
response obtained from the suggested Timoshenko formulation is able to capture
553
the small linear shear strain in the element without any shear locking effects. In
554
this scenario, the Timoshenko formulation is again capable to capture the response,
555
however, the Euler-Bernoulli formulation is also sufficient here.
35
Experiment
(c)Timoshenko:Local
600
600
300
300
300
0
-600 -0.06
Shear force (kN)
600
-300
0
-300
-600 -0.03
0
0.03
0.06
-0.5
Total rotation (rad)
0
-300
-600 -0.25
0
0.25
0.5
-0.01
(e)Euler:Local
300
300
-600
Shear force (kN)
300
Moment (kNm)
600
-300
0
-300
-600 -0.03
0
0.03
Total rotation (rad)
0.06
-0.5
0.005
0.01
(f)Euler:Local
600
0
0
Shear strain
600
-0.06
-0.005
Curvature (1/m)
(d)Euler:Global
Force (kN)
Hysteretic finite element
(b)Timoshenko:Local
Moment (kNm)
Force (kN)
(a)Timoshenko:Global
0
-300
-600 -0.25
0
0.25
Curvature (1/m)
0.5
-0.01
-0.005
0
0.005
0.01
Shear strain
Fig. 9. Comparison of experimental and simulated response for flexure dominant beam.
556
6. Conclusions
557
In this paper, a consistent and computationally efficient hysteretic beam finite ele-
558
ment model is presented. The hysteretic modeling component, expressed in terms
559
of stress resultants and featuring distributed multiaxial plasticity with interactions,
560
is fully integrated into the element formulation, enabling numerous important
561
features. Nodal displacement and strain fields continuity along the element is
562
achieved, due to development of new displacement field interpolation functions.
563
These functions take into account the introduced hysteretic degrees of freedom
564
and satisfy both the exact equilibrium equations and kinematic conditions, leading
565
to a consistent element formulation. Although it is shown in this work that the 36
566
suggested Timoshenko beam element can seamlessly model all possible scenarios,
567
the parametrized nature of the element allows also a direct transformation to an
568
Euler-Bernoulli beam element, by simple parametric changes. Additionally, the
569
suggested formulation produces constant element stiffness and hysteretic matrices
570
that do not need to be updated throughout the analysis, given that inelasticity is
571
treated through localized hysteretic evolution equations. Finally, a new compu-
572
tationally efficient solution scheme is suggested for quasi-static problems, where
573
the entire framework is presented as an ODE system and can be efficiently solved
574
without the need for any type of linearization. Several numerical examples are
575
included in this work in order to showcase the characteristics and capabilities of
576
the element, and comparisons with available experimental data are also provided.
577
Overall, the presented element possesses significant, favorable properties and can
578
be easily modeled and even incorporated in existing finite element codes. Several
579
further extensions are also available and future work will include, among others,
580
incorporation of large displacements and damage-plasticity attributes, concertedly
581
preserving its computational efficiency.
582
7. Acknowledgements
583
The authors would like to acknowledge the support of the U.S. National Science
584
Foundation, which partially supported this research under Grant No. CMMI-
585
1351591, and to thank Dr. Jeffrey W. Berman and Dr. Michel Bruneau for
586
providing the detailed experimental data.
37
587
Appendix. Stiffness and hysteretic matrices for common cases
588
CASE 1. Nonlinear Euler-Bernoulli beam element (λ = 0 and αγ = 1):
589
590
591
αu AE 0 0 − αuLAE 0 0 L 0 12αLφ3E I 6αLφ 2E I 0 − 12αLφ3E I 6αLφ 2E I 6α E I 2αφ E I 4αφ E I 6αφ E I 0 0 − φ2 2 L L L L K = α AE αu AE 0 0 0 0 − uL L 12αφ E I 6αφ E I 12αφ E I 6αφ E I 0 − L3 − L2 0 − L3 L2 6αφ E I 4αφ E I 2αφ E I 6αφ E I 0 0 − L L L2 L2
− (1−αu ) AE 0 0 − (1−αu2 ) AE 0 0 2 (1−αφ )E I (1−αφ )E I 0 0 − 0 0 L L 0 0 −(1 − αφ )E I 0 0 0 H = (1−αu ) AE (1−αu ) AE 0 0 0 0 2 2 (1−α )E I (1−α )E I φ φ 0 0 0 − 0 L L 0 0 0 0 0 (1 − αφ )E I CASE 2. Linear Timoshenko beam element (αu = αγ = αφ = 1): AE 0 L 0 12µLE3 I 0 6µ E I K = AE L 2 − L 0 0 − 12µLE3 I 0 6µ E2 I L
0
6µ E I L2 4(1+3λ)µ E I L
0 0
0
12µ E I L3 6µ E I − 2 L
−
0
AE L
0
6µ E I L2 2(1−6λ)µ E I L
0
12µ E I L3 6µ E I − 2 L
−
H = 0 ; where, µ =
592
− AE L
0
6µ E I L2 2(1−6λ)µ E I L 0 6µ E I − 2 L 4(1+3λ)µ E I L 0
1 EI ;λ = 1 + 12λ G As L 2
CASE 3. Linear Euler-Bernoulli beam element (λ = 0 and αu = αγ = αφ = 1): AE 0 0 − AE 0 0 L L I 6E I 0 − 12E3 I 6E2I 0 12E 3 2 L L L L 0 6E2I 4EL I 0 − 6E2I 2EL I L L ;H = 0 K = AE AE 0 0 0 0 − L L I 6E I 12E I 6E I 0 − 12E − 0 − L3 L2 L3 L2 0 6E2I 2EL I 0 − 6E2I 4EL I L L 38
593
References
594
[1] G. G. Deierlein, A. M. Reinhorn, M. R. Willford, Nonlinear structural analysis
595
for seismic design, NEHRP Seismic Design Technical Brief No. 4 (2010) 1–
596
36.
597
[2] D. G. Lignos, H. Krawinkler, Deterioration modeling of steel components
598
in support of collapse prediction of steel moment frames under earthquake
599
loading, Journal of Structural Engineering 137 (11) (2010) 1291–1302.
600
[3] E. Spacone, F. C. Filippou, F. F. Taucer, Fibre beam–column model for non-
601
linear analysis of R/C frames: Part I. Formulation, Earthquake Engineering
602
& Structural Dynamics 25 (7) (1996) 711–725.
603
[4] M. H. Scott, G. L. Fenves, F. McKenna, F. C. Filippou, Software patterns for
604
nonlinear beam-column models, Journal of Structural Engineering 134 (4)
605
(2008) 562–571.
606
[5] S. P. Triantafyllou, V. K. Koumousis, Small and large displacement dynamic
607
analysis of frame structures based on hysteretic beam elements, Journal of
608
Engineering Mechanics 138 (1) (2011) 36–49.
609
[6] S. P. Triantafyllou, V. K. Koumousis, An inelastic Timoshenko beam ele-
610
ment with axial–shear–flexural interaction, Computational Mechanics 48 (6)
611
(2011) 713–727.
612
613
[7] J. W. Macki, P. Nistri, P. Zecca, Mathematical models for hysteresis, SIAM Review 35 (1) (1993) 94–123.
39
614
615
616
617
618
619
[8] A. Visintin, Mathematical models of hysteresis, in: The Science of Hysteresis, vol. 1, Elsevier, 2005. [9] M. Dimian, P. Andrei, Noise-driven phenomena in hysteretic systems, vol. 218, Springer, 2014. [10] W. D. Iwan, A distributed-element model for hysteresis and its steady-state dynamic response, Journal of Applied Mechanics 33 (4) (1966) 893–900.
620
[11] L. F. Ibarra, R. A. Medina, H. Krawinkler, Hysteretic models that incorporate
621
strength and stiffness deterioration, Earthquake Engineering & Structural
622
Dynamics 34 (12) (2005) 1489–1511.
623
[12] S. Erlicher, Hysteretic degrading models for the low-cycle fatigue behaviour
624
of structural elements: Theory, numerical aspects and applications, Ph.D.
625
thesis, University of Trento, 2003.
626
627
628
629
630
631
[13] R. Bouc, Forced vibrations of mechanical systems with hysteresis, in: Proc. of the 4th Conference on Nonlinear Oscillations, Prague, 1967. [14] Y.-K. Wen, Method for random vibration of hysteretic systems, Journal of the Engineering Mechanics Division 102 (2) (1976) 249–263. [15] F. Casciati, Stochastic dynamics of hysteretic media, Structural Safety 6 (2-4) (1989) 259–269.
632
[16] M. Sivaselvan, A. Reinhorn, Nonlinear structural analysis towards collapse
633
simulation: A dynamical systems approach, Technical Report MCEER-04-
634
0005, 2004.
40
635
[17] S. P. Triantafyllou, Hysteretic finite elements and macro-elements for non-
636
linear dynamic analysis of structures, Ph.D. thesis, National Technical Uni-
637
versity of Athens, 2011.
638
[18] J. Lubliner, Plasticity theory, Dover Publications, 2008.
639
[19] T. T. Baber, Y.-K. Wen, Random vibration hysteretic, degrading systems,
640
641
642
643
644
645
646
Journal of the Engineering Mechanics Division 107 (6) (1981) 1069–1087. [20] T. T. Baber, M. N. Noori, Random vibration of degrading, pinching systems, Journal of Engineering Mechanics 111 (8) (1985) 1010–1026. [21] G. C. Foliente, Hysteresis modeling of wood joints and structural systems, Journal of Structural Engineering 121 (6) (1995) 1013–1022. [22] M. V. Sivaselvan, A. M. Reinhorn, Hysteretic models for deteriorating inelastic structures, Journal of Engineering Mechanics 126 (6) (2000) 633–640.
647
[23] J. Song, A. Der Kiureghian, Generalized Bouc–Wen model for highly asym-
648
metric hysteresis, Journal of Engineering Mechanics 132 (6) (2006) 610–618.
649
[24] K. G. Papakonstantinou, P. C. Dimizas, V. K. Koumousis, An inelastic beam
650
element with hysteretic damping, Shock and Vibration 15 (3, 4) (2008) 273–
651
290.
652
[25] P. S. Harvey Jr, H. P. Gavin, Truly isotropic biaxial hysteresis with arbitrary
653
knee sharpness, Earthquake Engineering & Structural Dynamics 43 (13)
654
(2014) 2051–2057.
655
[26] M. Ismail, F. Ikhouane, J. Rodellar, The hysteresis Bouc-Wen model, a survey,
656
Archives of Computational Methods in Engineering 16 (2) (2009) 161–188. 41
657
[27] M. Amabili, Derivation of nonlinear damping from viscoelasticity in case of
658
nonlinear vibrations, Nonlinear Dynamics (2018) doi:10.1007/s11071–018–
659
4312–0.
660
[28] M. Amabili, Nonlinear damping in nonlinear vibrations of rectangular plates:
661
Derivation from viscoelasticity and experimental validation, Journal of the
662
Mechanics and Physics of Solids 118 (2018) 275–292.
663
664
665
666
[29] M. Amabili, Nonlinear damping in large-amplitude vibrations: modelling and experiments, Nonlinear Dynamics 93 (2018) 5–18. [30] J. Rakowski, The interpretation of the shear locking in beam elements, Computers & Structures 37 (5) (1990) 769–776.
667
[31] K. J. Bathe, Finite element procedures, Prentice Hall, 1996.
668
[32] S. Erlicher, N. Point, Thermodynamic admissibility of Bouc–Wen type hys-
669
teresis models, Comptes Rendus Mecanique 332 (1) (2004) 51–57.
670
[33] J. G. Orbison, W. McGuire, J. F. Abel, Yield surface applications in non-
671
linear steel frame analysis, Computer Methods in Applied Mechanics and
672
Engineering 33 (1-3) (1982) 557–573.
673
[34] A. Charalampakis, V. Koumousis, Ultimate strength analysis of composite
674
sections under biaxial bending and axial load, Advances in Engineering
675
Software 39 (11) (2008) 923–936.
676
677
[35] Z. Friedman, J. B. Kosmatka, An improved two-node Timoshenko beam finite element, Computers & Structures 47 (3) (1993) 473–481.
42
678
[36] J. Berman, M. Bruneau, Further development of tubular eccentrically braced
679
frame links for the seismic retrofit of braced steel truss bridge piers, Technical
680
Report MCEER-06-0006, 2006.
43
Highlights:
• • • • •
A general consistent hysteretic beam finite element model is developed New displacement interpolation functions are introduced to satisfy exact equilibrium conditions Strain continuity is achieved without any shear-locking effects A new computationally efficient solution scheme is suggested Axial-moment-shear interactions and distributed plasticity are considered