Вы находитесь на странице: 1из 8

International Journal of Mechanical Sciences 133 (2017) 368–375

Contents lists available at ScienceDirect

International Journal of Mechanical Sciences


journal homepage: www.elsevier.com/locate/ijmecsci

Finite element modeling of the elastoplastic axial-torsional response of


helical constructions to traction loads
Nikolaos Karathanasopoulos a,∗, Hilal Reda b, Jean-francois Ganghoffer b
a
Computational Science and Engineering Laboratory, ETH Zürich, Clausiusstrasse 33, CH-8092 Zürich, Switzerland
b
University of Lorraine, LEMTA, Nancy

a r t i c l e i n f o a b s t r a c t

Keywords: In the current work, we present a planar elastoplastic finite element model for the simulation of the mechanical
Helix response of helical constructions to traction loads. We elaborate a numerical scheme that employs a detailed
Cable description of the helical geometry and characterizes its coupled axial and torsional response. We validate the
Plasticity
modelling results with well-established three-dimensional (3D) finite element schemes. For the validation, a
Stiffness
single layer helical strand with different construction parameters is employed. Additional analysis is carried out
Loading history
to highlight the role of hardening, providing useful insights in the effect of loading history on the structure’s
macroscopic elastoplastic response. Overall, a low computation cost elastoplastic modeling scheme, suitable for
large-scale structural analysis of helical constructions is presented.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction ticular, they inherently counterbalance the torsional moments created


by the structure’s helical constituents, leading to an equilibrated macro-
Helical constructions are extensively used structural members. They scopic mechanical response [28].
are employed in civil and mechanical engineering applications in the Analytical models have been complemented by both experimental
form of springs [1], steel cables, ropes or electricity power conductors studies [29,30] and numerical models [31–34]. The latter have been pri-
[2–5]. They demonstrate high load bearing capacities upon relatively marily based on beam or volume finite elements [35,36]. Volume mod-
small dead weights [6–8]. els have been employed to characterize the mechanical behaviour of sin-
The helix geometry plays a crucial role in the structural behaviour of gle or multilayer helical engineering strands with helical constituents of
helical assemblies [9]. Many studies have been devoted to the analytical circular, triangular or oval-shaped cross section profiles [37,38]. Lately,
characterization of the mechanical response of helical constructions to helix planar finite element models have appeared in the literature, orig-
different loading patterns, such as to axial, torsional, thermal or bending inally in the field of wave propagation [39] and later on in the lin-
loads [10–14]. Analytical formulations provide closed-form expressions ear elastic static characterization of helical constructions, providing low
for their constitutive behaviour that can be handily used for calculations, numerical-cost analysis tools [40,41].
given the construction’s geometric and material parameters [15–18]. In the last decades, applications beyond the context of engineering
Axial strains innately induce torsional loads; a mechanical behaviour strands have come to the fore. In particular, in the field of biomedical
that is reflected in most analytical approaches [19]. engineering, dedicated helically braided scaffold finite element models
The understanding and characterization of the axial-torsional load- have been presented to emulate the scaffold’s experimentally observed
ing coupling constitutes a prerequisite not only for an accurate structural global mechanical behaviour [42].
response assessment [20,21], but also for the prediction of their long Most helix closed-form expressions and numerical models consider
term mechanical performance [22–24]. In particular, the mean magni- a linear elastic helix material behaviour. However, up to now, rela-
tude of the internal loads developed throughout a helical members’ life- tively few works have accounted for a non-linear material behaviour.
time has been shown to directly affect its fatigue life expectancy [25,26]. More specifically, full three-dimensional (3D) elastoplastic finite ele-
Higher torsional loads have been related to substantially shorter fatigue ment models have been proposed [43,44], as well as reduced volume
lives [27]. Certain structural arrangements diminish the effective tor- finite element models which exploit the helical symmetry [45–47]. How-
sional loads developed in helically braided structural members. In par- ever, volume models entail relatively high computational costs, which


Corresponding author.
E-mail addresses: nkaratha@ethz.ch (N. Karathanasopoulos), hilalreda@hotmail.com (H. Reda), jean-francois.Ganghoffer@univ-lorraine.fr (J.-f. Ganghoffer).

http://dx.doi.org/10.1016/j.ijmecsci.2017.09.002
Received 14 July 2017; Received in revised form 21 August 2017; Accepted 4 September 2017
Available online 6 September 2017
0020-7403/© 2017 Elsevier Ltd. All rights reserved.
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

Fig. 1. Helical structure geometry a) Single layer helical strand, b) Normal strand section schematic representation (the helical bodies are depicted as circles, viewed in their local
Serret-Frenet nb frame), while the Cartesian system is attached to the central core.

are aggravated by the non-linear nature of the elastoplastic computa- expressions [9]:
tion. In order to provide robust computational tools, simplified phe- ⎧− cos 𝜑⎫ ⎧ 𝑏 sin 𝜑 ⎫ ⎧−𝑎 sin 𝜑⎫
nomenological uniaxial elastoplastic models have been developed [48]. ⎪ ⎪ 1⎪ ⎪ 1⎪ ⎪
𝐧 = ⎨ − sin 𝜑 ⎬, 𝐛= −𝑏 cos 𝜑⎬, 𝐭= 𝑎 cos 𝜑 ⎬
𝛾⎨ 𝛾⎨
(2)
Such models however fail to describe a distinctive attribute of heli-
⎪ 0 ⎪ ⎪ 𝑎 ⎪ ⎪ 𝑏 ⎪
cal constructions, namely the axial-torsional straining coupling [49]. ⎩ ⎭ ⎩ ⎭ ⎩ ⎭
What is more, numerical models commonly disregard the effect of hard- The angular evolution 𝜑 is related to the Curvilinear coordinate s
ening on the macroscopic strand response [50]. Notwithstanding, ex- through the parameters b and 𝛾, as follows [9]:
perimental observations have recorded considerable differences in the 𝑠 √
strain levels at which the plastic response initiates, depending on the 𝜑 = , 𝛾 = 𝛼 2 + 𝑏2 , 𝑏 = 𝛼 tan 𝜃, ℎ = 2𝜋𝑏 (3)
𝛾
loading-unloading cycle history to which the cable structure is subject
In Eq. (3), b stands for the rise along the Cartesian axis Z of the helix
to [29,51].
per unit angular evolution 𝜑, upon which the helix height h for a pe-
In the current work, we characterize the elastoplastic response of
riod evolution 2𝜋 is computed. The helix angle 𝜃 is defined as the angle
helical constructions to purely axial loads, extending a planar, two-
formed by the tangent Serret–Frenet vector t with the Cartesian plane
dimensional linear finite element modelling approach [41] to the non-
eX , eY . Fig. 1 schematically depicts the introduced parametrization:
linear material response domain (Section 2). Contrary to uniaxial phe-
Upon the helix geometric parameters a, b and 𝛾, the helix curvature
nomenological approaches [48], the elaborated approach can well sim-
𝜅 and tortuosity 𝜏 are defined as follows:
ulate the coupled axial-torsional response of tensioned cables compris-
𝑎 𝑏
ing of helical components of equal helix angle that are rotationally re- 𝜅= 𝜏= (4)
strained. In Section 3, we compute the macroscopic structural response 𝛾2 𝛾2
of single-layered metallic engineering strands and analyze the effect of For a layered helical-assembly, Eq. (1) can be used to calculate the num-
the strand’s configuration on its post-elastic mechanical behaviour. We ber of helical bodies that can be positioned within a certain layer, with-
validate our analysis by comparing the planar finite element modeling out a circumferential overlapping of the helical bodies [28]. This can
results with well-established 3D finite element models. In Section 4, we be achieved by computing the helix cross section projection on a plane
discuss the effect of hardening on the strand’s macroscopic elastoplas- that is normal to the axis Z of the helical strand. Eq. (5) describes the
tic behaviour, analyze the model’s computational merits and simulation intersection of the Cartesian plane 𝑍 = 0 with Eq. (1):
limitations and conclude. { } { } { }
𝑋 cos 𝜑∗ 𝑏 sin 𝜑∗ 𝛼
= (𝛼 − 𝑥𝑛 ) + 𝑥𝑏 , 𝜑∗ = − 𝑥𝑏 (5)
𝑌 sin 𝜑 ∗
𝛾 − cos 𝜑 ∗
𝑏𝛾
The circumferential marginal point C′ (Fig. 2b) of the helix intersec-
2. Helix geometry and finite element modeling basis tion with plane Z=0 is subsequently calculated. To that extent, the local
cross section coordinates (xn , xb ) of the point C′ are computed, using the
2.1. Helix geometry tangent point C of the circular helix cross section (Fig. 2a), as follows:
( )
𝑟
, 𝜃𝑐 = 90𝑜 − 𝜃𝑐 , 𝑥′𝑛 = 𝑟 cos 𝜃𝑐 , 𝑥′𝑏 = 𝑟 sin 𝜃𝑐
′ ′ ′
We describe the helix geometry using the Curvilinear n, b, t Serret- 𝜃𝑐 = arcsin (6)
𝛼
Frenet basis. Thereupon, the position vector X of any point along the The domain occupied by each helical body is thereafter calculated by
helical domain is defined by means of a total of three independent pa- substituting 𝑥′𝑛 and 𝑥′𝑏 in Eq. (5), upon which the Cartesian coordinates
rameters, namely: the helix cross section coordinates xn and xb , which 𝑋𝑐 ′ and 𝑌𝑐 ′ are obtained. The central angle 2 𝜓 (Fig. 2b) formed by the
follow the normal and binormal vectors n and b and the angular position margins of the helical body is then given as:
𝜑 of the helix centerline, as follows: ( )
𝑌𝑐 ′
2 𝜓 = 2 arctan (7)
𝑋𝑐 ′
⎧𝑋 ⎫ ⎧𝑎 cos 𝜑⎫ ⎧− cos 𝜑⎫ ⎧ 𝑏 sin 𝜑 ⎫
⎪ ⎪ ⎪ ⎪ ⎪ ⎪ 𝑥 ⎪ ⎪
𝐗(𝑥𝑛 , 𝑥𝑏 , 𝜑) = ⎨ 𝑌 ⎬ = ⎨ 𝑎 sin 𝜑 ⎬ + 𝑥𝑛 ⎨ − sin 𝜑 ⎬ + 𝑏 ⎨−𝑏 cos 𝜑⎬, 2.2. Helix elastoplastic finite element modeling
⎪𝑍 ⎪ ⎪ 𝑏𝜑 ⎪ ⎪ 0 ⎪ 𝛾 ⎪ ⎪
⎩ ⎭ ⎩ ⎭ ⎩ ⎭ ⎩ 𝛼 ⎭
We consider an elastoplastic material model with linear elastic mod-
−𝑟 ≤ 𝑥 𝑛 , 𝑥 𝑏 ≤ 𝑟 (1) ulus E and plastic modulus H, following the stress-strain relation of
Fig. 3. In Fig. 3 𝜎 y stands for the initial yield stress 𝜎 y,0 of the mate-
In Eq. (1), a stands for the helix centerline position and r for helix cross rial. We adopt a yield surface 𝜙 that is defined by a Von Mises yielding
section radius. For the position vector X of Eq. (1) to be defined, the criterion as follows [52]:
n and b base vectors of the Serret–Frenet basis have been employed. √
𝑝𝑙
√ 3
The local Serret–Frenet base vectors n, b, t are defined by the following 𝜙(𝝈) = 𝜎(
̄ 𝝈) − 𝜎𝑦 (𝜖̄ ), 𝜎̄ = 3𝐽2 = 𝐬∶𝐬 (8)
2
369
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

Fig. 2. Helix circumferential margins a) Cross section normal section b) Global cartesian plane Z=0 projection resulting in a non-circular shape according to Eq. (5).

control point of the descretized domain. The elastoplastic material ten-


sor is obtained using the consistency condition, requesting any stress
increment to lie on the extended yield locus [54]:
( )
𝐪𝑇 ⋅ 𝐂𝑒𝑙 ⋅ 𝐪 𝑑 𝜎̄ 𝑑 𝜎̄ 3𝐬
𝐂𝑝𝑙 = 𝐂𝑒𝑙 1 − , 𝐪= = = (12)
𝐻 + 𝐪𝑇 ⋅ 𝐂𝑒𝑙 ⋅ 𝐪 𝑑𝝈 𝑑𝒔 2𝜎̄

Performing the usual finite element discretization process, the following


discrete system of equations is obtained:

𝐊 𝐮 = 𝐟 + 𝐟0 , 𝐊= 𝐁𝑇 𝐂 𝐁 𝑑Ω, 𝐟 =− 𝐁 𝑇 𝐂𝝐 𝐟0 = 𝚽𝑇 𝝈 0
∫Ω ∫Ω ∫Ω
(13)

In Eq. (13), B stands for the linear finite element operator, 𝚽 for the
Fig. 3. Material stress-strain relation. domain shape functions and 𝝈 0 for the initial stress vector. The applied
total strain tensor 𝝐 entering Eq. (13) is analytically defined for the case
of axial strains 𝜖𝑧 = 𝛿ℎ∕ℎ by means of the helix geometric attributes,
where 𝝈 stands for the stress tensor and 𝜎̄ for the equivalent Von Mises namely: the centerline position a, pitch and tortuosity value b and 𝜏
stress, determined as a function of the second invariant J2 , thus of the respectively. The explicit expressions for the linear operator B and total
deviatoric stress tensor 𝐬 = 𝝈 − 13 𝑡𝑟𝑎𝑐𝑒(𝝈). The function 𝜙(𝝈) defined in strain tensor 𝝐 are provided in A.2.
Eq. (8) characterizes the material’s state: negative function values en- Eqs. (8)–(13) completely define the elastoplastic behaviour of the he-
tail an elastic material response, while zero function values a plastic one. lical strand. For a given strain value 𝜖 z they allow for a direct evaluation
Positive values for Eq. (8) are not defined. For the current analysis, a lin- of the finite element solution (13). The algorithm follows an incremen-
ear hardening material law that is proportional to the equivalent plastic tal implementation scheme, upon which at each loading increment the
strain 𝜖̄𝑝𝑙 (see Eqs. (10), (11)) will be employed, so that 𝜎𝑦 = 𝜎𝑦,0 + 𝐻 𝜖̄𝑝𝑙 . equivalent Von Mises stress 𝜎̄ is calculated at each integration point
The stress tensor 𝝈 is obtained upon an incremental total material strain [54]. If the equivalent Von Mises stress does not exceed the material
d𝝐 application. The latter is decomposed to an elastic 𝝐 el and a plastic yield stress limit 𝜎𝑦 (𝜖̄𝑝𝑙 ), an elastic step is carried out. Contrariwise, for
part 𝝐 pl , as follows [53]: the finite elements for which the equivalent Von Mises stress exceeds
the yield stress 𝜎 y , the elastic part m of the trial step d𝝈 tr is computed,
𝑑 𝝈 = 𝑪 ⋅ 𝑑 𝝐 𝑒𝑙 = 𝑪 (𝑑 𝝐 − 𝑑 𝝐 𝑝𝑙 ), 𝑑 𝝐 = 𝑑 𝝐 𝑒𝑙 + 𝑑 𝝐 𝑝𝑙 , 𝝈 𝑛 = 𝝈 𝑛−1 + 𝑑 𝝈 (9)
for the elastic and plastic strain contributions of the loading step to be
The constitutive Eq. (9) is defined in the general Curvilinear heli- obtained. After each plastic loading step, the equivalent Von Mises plas-
cal domain employed in Eq. (1) whose base vectors are elaborated in tic strain 𝜖̄𝑝𝑙 is calculated for the next loading increment to take place.
Appendix A.1. The relation between the deviatoric stress tensor and the Fig. 4 provides a schematic representation of the algorithmic process.
plastic strain increment d𝝐 pl can be defined using the normality rule for The effective global structural response of the helical strand is eval-
the plastic flow over the yield surface, as follows [53,54]: uated by computing the resulting force Fi and moment components Mi
components (𝑖 = 𝐠1 , 𝐠2 , 𝐠3 ) by integrating the appropriate stress compo-
𝑑𝜙 𝑑 𝜎̄ 3 2 𝜎̄
𝑑 𝝐 𝑝𝑙 = 𝑑 𝜖̄𝑝𝑙 = 𝑑 𝜖̄𝑝𝑙 = 𝑑 𝜖̄𝑝𝑙 𝐬⇒𝐬= 𝑑 𝝐 𝑝𝑙 (10) nents over the helix cross sectional area A, as follows:
𝑑𝝈 𝑑𝒔 2𝜎̄ 3 𝑑 𝜖̄𝑝𝑙
Using the principle of virtual works along with the result of Eq. (10), 𝐹𝐠1 = 𝜎11 𝑑𝐴, 𝐹𝐠2 = 𝜎22 𝑑𝐴, 𝐹𝐠3 = 𝜎33 𝑑𝐴
the relation of the equivalent plastic strain increment 𝑑 𝜖̄𝑝𝑙 to the plastic ∫𝐴 ∫𝐴 ∫𝐴
strain increment d𝝐 pl for a Von Mises yield criterion can be defined as
𝑀𝐠1 = 𝜎33 𝑥𝑏 𝑑𝐴, 𝑀𝐠2 = − 𝜎33 𝑥𝑛 𝑑𝐴,
follows: ∫𝐴 ∫𝐴
√ ( )
2 𝑝𝑙 𝑀𝐠3 = 𝜎22 𝑥𝑛 − 𝜎11 𝑥𝑏 𝑑𝐴 (14)
𝜎̄ 𝑑 𝜖̄𝑝𝑙 = 𝐬 ∶ 𝑑 𝝐 𝑝𝑙 ⇒ 𝑑 𝜖̄𝑝𝑙 = 𝑑 𝝐 ∶ 𝑑 𝝐 𝑝𝑙 (11) ∫𝐴
3
The material tensor C used for the incremental computations of Eq. (9) is The force and moment components of Eq. (14) are defined in the he-
equal either to the elastic material tensor C𝑒𝑙 (Appendix A) or to its lix local basis. For the global Cartesian components to be obtained
elastoplastic counterpart C𝑝𝑙 , depending on the material state at each Eqs. (A.4) and (A.5) need to be employed.

370
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

Table 1
Helical strand geometric and material parameters.

rc (mm) r (mm) E (GPa) H (GPa) 𝜎 y (GPa) 𝜃 (o )

1.97 1.865 188 24.6 1.56 78.2

triangular elements and 8250 nodes has been found to suffice for the ef-
fective force and moment components of Eq. (14) to differ less than 1%
with respect to considerably higher mesh discretizations. Node-to-node
contact pairs have been applied for the radial contact of the helical bod-
ies with the central core. The contact pairs have been considered to be
fully kinematically coupled. Radial friction has been shown to have neg-
ligible effects on the macroscopic axial-torsional response of the helical
strand [55]. Along the circumferential direction, no contact conditions
have been applied. Making use of Eqs. (5)–(7) we have computed for the
geometric specifications of Table 1 a central angle 2 𝜓 = 55.6𝑜 , which
corresponds to a clearance of 4.4o among the helical bodies, assuming
regular placement. We summarize the geometric and material data of
the helical considered helical strand in Table 1.
We compute the strand’s effective axial force Fz and torsional mo-
ment Mz considering force and moment equilibrium along the global
Cartesian axis Z, by summing up the contributions of the helical wires
and of the central core. For the contributions of the helical wires, the
definitions of Eq. (14) are used along with Eq. (A.6).
( )
𝐹𝑧 = 𝐹𝑐 + 6 𝐹𝑔2 cos 𝜃 + 𝐹𝑔3 sin 𝜃
Fig. 4. Elastoplastic algorithm flow diagram.
( ) ( )
𝑀𝑧 = 6 𝑀𝑔3 sin 𝜃 + 𝑀𝑔2 cos 𝜃 + 6 𝛼 𝐹𝑔3 cos 𝜃 − 𝐹𝑔2 sin 𝜃 (15)

We compare the numerically obtained results for the elastoplastic planar


FE model elaborated in Section 2.2 with a linear elastic analytical model
[11] and an elastoplastic 3D finite element model provided by Foti et al.
[49] in Fig. 6.
Fig. 6 indicates an excellent agreement between the planar and the
3D finite element model, while a plastic response is recorded at a strain
threshold of 0.8% (Fig. 6 left). We subsequently analyze the sensitivity
of the elastoplastic response to the helix angle 𝜃. In Fig. 7 (left), we note
that the axial force Fz at which plastification initiates, reduces upon a
decreasing strand helix angle 𝜃. For a helix angle of 𝜃 = 75𝑜 , the strand
yielding force is approximately 6% lower than the one corresponding
to a rather steep helix angle of 85o . However the torsional moment Mz
created considerably increases upon decreasing helix angle values 𝜃. In
the plastic response range, a slight increase in the slope of the force-
moment diagram is recorded (Fig. 7 right). For all helix angles 𝜃, the
planar finite element results are in very good agreement with the 3D
modelling results [49], with any difference to lie below 3%.
In Fig. 8 we present the stress distribution for the normal stress
𝜎 33 and for the Von Mises equivalent stress 𝜎𝑒𝑞 = 𝜎̄ for a strand load
𝐹𝑧 = 120 KN. We note a post-yield quasi constant normal stress distribu-
tion for the helical wires, with stress concentrations close to the contact
points with the central core (Fig. 8a). The core is under higher normal
stress than the helical wires, while yielding initiates at the contact re-
Fig. 5. Helical strand finite element mesh. gions and gradually extends to the entire strand cross section (Fig. 8b).

4. Discusion and conclusion


3. Results
In Section 3, a single elastoplastic loading cycle has been presented,
3.1. Elastoplastic response of a single layer metallic helical strand during which the yield stress value 𝜎 y remained constant. However,
metallic strands are commonly subject to loading-unloading cycles. The
We employ a well-documented “6+1” helical construction bench- development of plastic deformations during a loading cycle results in a
mark application for the verification of the hereby explicated elasto- modified yield surface for the subsequent loading phase, which is depen-
plastic model. In Fig. 5, we depict a finite element mesh of the helical dent on the hardening law of the material. Thus, by taking into account
strand. the structure’s loading history, a state-dependent macroscopic response
For the finite element discretization of the planar domain, prelimi- is obtained, where the yielding value is determined by the current value
nary mesh sensitivity analysis have been conducted to identify appro- of the equivalent plastic strain 𝜎𝑦 (𝜖̄𝑝𝑙 ). In Fig. 9a, we showcase the force-
priate mesh densities. A mesh density of approximately 2150 three-node strain curve for an elastoplastic loading, unloading and reloading cycle

371
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

Fig. 6. Elastoplastic force-strain (left) and force-moment response (right) for the helical strand of Table 1. Analytical elastic, planar elastoplastic and 3D elastoplastic FE modeling
comparisons [49].

Fig. 7. Sensitivity of the elastoplastic response to the strand’s helix angle 𝜃.

Fig. 8. Strand normal stress 𝜎 33 (a) and equivalent stress 𝜎 eq (b) for a strand load 𝐹𝑧 = 120 KN.

372
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

Fig. 9. a) Macroscopic force-deformation response of a 6+1 helical strand of Table 1 upon a loading, unloading and reloading cycle b) Number of nodes comparison between for the
full scale 3D model [49] and the planar model.

computation. For an equivalent plastic strain 𝜖̄𝑝𝑙 = 0.15%, the yield stress its variables (𝜕 X/𝜕 xi ), we obtain the general covariant base vector com-
increased by approximately 4% during the reloading phase. We note ponents as follows:
that the impact of hardening on the macroscopic response can be com-
parable with the effect of a structural arrangement modification, with a 𝐗,𝑥𝑛 = 𝐠1 = 𝐧
hardened response to correspond to a higher helix angle 𝜃 strand con- 𝐗,𝑥𝑏 = 𝐠2 = 𝐛
struction (Fig. 7).
𝐗,𝜑 = 𝐠3 = 𝑥𝑛 𝜏𝐛 − 𝑥𝑏 𝜏𝐧 + (1 − 𝑥𝑛 𝜅)𝐭 (A.1)
It needs to be remarked that plastification initiates at rather low
strain levels, below a total strain value of 1% for all geometric configu- [ ]
The covariant metric tensor of (A.1) is defined as 𝑔𝑖𝑗 = 𝐠𝑖 ⋅ 𝐠𝑗 given as
rations of the helical strand analysed in Section 3. The considerable dif-
follows:
ferences between the non-linear structural response and its linear-elastic
−𝜏𝑥𝑏
[ ] ⎡ ⎤
counterpart highlight the necessity for material non-linearities to be in- 1 0
tegrated in the effective global response computations. The hereby elab- 𝑔𝑖𝑗 = ⎢ 0 1 𝜏𝑥𝑛 ⎥ (A.2)
⎢ ⎥
orated approach allows not only for the simulation of helical construc- ⎣−𝜏𝑥𝑏 𝜏𝑥𝑛 𝜏 2 (𝑥2𝑛 + 𝑥2𝑏 ) + (1 − 𝜅𝑥𝑛 )2⎦
tions with circular cross section profiles, but also for non-circular ones, [ ]
The respective contravariant metric tensor, defined by 𝑔 𝑖𝑗 = 𝐠𝑖 ⋅ 𝐠𝑗 is
such as for oval-shaped that have been recently employed in construc-
given as:
tion practise [38], upon appropriate modifications of the cross-section
meshing algorithm. 𝑔 + (𝜏 𝑥𝑏 )2 −𝜏 2 𝑥𝑛 𝑥𝑏 𝜏𝑥𝑏 ⎤
[ 𝑖𝑗 ] 1 ⎡
Overall, the elaborated numerical scheme, not only accurately cap- 𝑔 = ⎢ −𝜏 2 𝑥𝑛 𝑥𝑏 𝑔 + (𝜏𝑥𝑛 )2 −𝜏𝑥𝑛 ⎥ (A.3)
tures the effective mechanical response of helical strands with a non- 𝑔⎢ ⎥
⎣ 𝜏𝑥𝑏 −𝜏𝑥𝑛 1 ⎦
linear material behaviour, but it also provides a primal robust computa-
tional framework for such computations. Its planar formulation makes where 𝑔 = (1 − 𝜅 𝑥𝑛 )2 is the determinant of the metric tensor which poses
feasible an approximate reduction of more than one order of magnitude the constrain 𝜅xn ≺ 1.
in terms of computational cost per loading increment (Fig. 9b) compared The constitutive law is provided for a system defined in a general
to full-scale 3D models [49]. However, the modeling size reduction en- Curvilinear basis in [58]. The stress tensor is written in vector notation
tails certain simulation limitations, arising from the downsizing of the as [𝜎 11 𝜎 22 𝜎 33 𝜎 23 𝜎 13 𝜎 12 ]T , while the strain tensor is accordingly writ-
3D problem to a 2D one. In particular, the proposed planar formulation ten in vector notation as [𝜖 11 𝜖 22 𝜖 33 2 𝜖 23 2 𝜖 13 2 𝜖 12 ]T [58]:
does not allow for the consideration of variations of the mechanical be- 1( ) 1( )
𝜎 𝑖𝑗 = 𝐶 𝑖𝑗𝑘𝑙 𝜖𝑘𝑙 , 𝜖𝑖𝑗 = 𝑢|
2 𝑖 𝑗
+ 𝑢𝑗 |𝑖 = 𝑢 , +𝑢 , − 𝑢𝑘 Γ𝑘𝑖𝑗 , (A.4)
haviour along the structure’s third dimension, therefore along its curvi- 2 𝑖𝑗 𝑗𝑖
linear evolution, which has been “eliminated” using helical symmetry.
In Eq. (A.1) superscripts and subscripts denote contravariant and covari-
As a result, in cases where the influence of the third dimension on the
ant base components respectively (Eqs. (A.1)–(A.3)). The elastic mate-
macroscopic structural response is expected to be significant (as for ex-
rial tensor Cel and the Christoffel symbol expressions of the second kind
ample in the case of complex contact conditions [56] or close to cable
entering Eq. (A.4) are defined as follows [59,60]:
terminations [57]), more detailed 3D modeling approaches should be
favoured. However, contrary to simplified phenomenological 1D elasto- 𝜈𝐸 𝐸 ( 𝑖𝑘 𝑗𝑙 )
𝐂𝑒𝑙 = 𝐶 𝑖𝑗𝑘𝑙 = 𝑔 𝑖𝑗 𝑔 𝑘𝑙 + 𝑔 𝑔 + 𝑔 𝑖𝑙 𝑔 𝑗𝑘 ,
plastic schemes, which fail to simulate a distinctive attribute of axially (1 + 𝜈)(1 − 2𝜈) 2 (1 + 𝜈)
loaded helical constructions, that is the creation of torsional moments Γ𝑘𝑖𝑗 = 𝐠𝑖 ,𝑗 ⋅𝐠𝑘 (A.5)
[48], the hereby elaborated planar elastoplastic numerical approach can
fully characterize the arising coupled axial-torsional behaviour. We as- where E and 𝜈 stand respectively for the material’s Young modulus and
pire that the elaborated scheme provides an effective, time-saving com- for its Poisson’s ratio value. The transformation tensor F relating the
putational framework for the elastoplastic analysis and design of heli- Curvilinear with the global Cartesian eX , eY , eZ basis is given as:
cal constructions, offering a robust modeling solution in between one-
dimensional and three-dimensional approaches. ⎡−𝛾𝐶 𝑏𝑆 𝑥𝑛 𝑆 + 𝛾𝜏𝑥𝑏 𝐶 − 𝑎𝑆 ⎤
1⎢
𝐅= −𝛾𝑆 −𝑏𝐶 −𝑥𝑛 𝐶 + 𝛾𝜏𝑥𝑏 𝑆 + 𝑎𝐶 ⎥ (A.6)
𝛾⎢ ⎥
Appendix A. Appendix ⎣ 0 𝑎 𝑏 ⎦

where C and S in (A.4) stand for the cos 𝜑 and sin 𝜑 respectively. Con-
A1. Helix constitutive relations and metric tensors sidering vector components p described in a Cartesian basis, we retrieve
contravariant (contra) vector components as follows [59]:
The constitutive material law is defined in a Curvilinear basis. By
differentiating the position vector X defined in Eq. (1) with respect to 𝐩𝑐𝑜𝑛𝑡𝑟𝑎 = 𝐅−1 𝐩 (A.7)

373
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

A2. Linear finite element operator and axial loading tensor [13] Lanteigne J. Theoretical estimation of the response of helically armored ca-
bles to tension, torsion, and bending. J Appl Mech 1985;52(June 1985):423.
doi:10.1115/1.3169064.
The Curvilinear strain tensor 𝜖 ij defined in Eq. (A.4) can be written [14] Karathanasopoulos N, Ganghoffer JF, Papailiou KO. Analytical closed-form expres-
as a sum of two distinct operators L12 and L3 , so that 𝝐 = (𝐋12 + 𝐋3 )𝒖, sions for the structural response of helical constructions to thermal loads. Int J Mech
where the second operator (L3 ) contains variations solely with respect Sci 2016;117:258–64. doi:10.1016/j.ijmecsci.2016.08.010.
[15] Sathikh S, Moorthy MBK, Krishnan M. A symmetric linear elastic model for helical
to the out of xn , xy plane coordinate (1,2), as follows: wire strands under axisymmetric loads. J Strain Anal Eng Design 1996;31(5):389–
99. doi:10.1243/03093247V315389.
⎧𝜖11 ⎫ ⎛⎡ ( ),1 0 0 ⎤ [16] Raoof M, Kraincanic I. Prediction of coupled axial/torsional stiffness
⎪ ⎪ ⎜⎢ 0 ( ),2 0 ⎥ coefficients of locked-coil ropes. Comput Struct 1998;69(1):305–19.
⎪𝜖22 ⎪ ⎜⎢ 1 ⎥ doi:10.1243/03093247V291043.
⎪𝜖33 ⎪ ⎜⎢ −Γ33 −Γ233 −Γ333 ⎥ [17] Yildirim V. Exact determination of the global tip deflection of both close-
⎨𝜖 ⎬ = ⎜⎢ −Γ1 0 1
( ),2 ⎥ coiled and open-coiled cylindrical helical compression springs having ar-
⎪ 23 ⎪ ⎜⎢ 23 1
2 ⎥ bitrary doubly-symmetric cross-sections. Int J Mech Sci 2016;115:280–98.
⎪𝜖13 ⎪ ⎜⎢ −Γ113 −Γ213 2
( ),1 −Γ313 ⎥ http://dx.doi.org/10.1016/j.ijmecsci.2016.06.022.
⎪𝜖 ⎪ ⎜⎢ 1 1 ⎥
⎩ 12 ⎭ ⎝⎣ 2 ( ),2 2
( ),1 0 ⎦ [18] Karathanasopoulos N, Kress G. Mechanical response of a helical body
to axial, torsional and radial strain. Int J Mech Sci 2015;94-95:260–5.
⎡ 0 0 0 ⎤⎞ doi:10.1016/j.ijmecsci.2015.02.022.
⎢ 0 0 0 ⎥⎟ [19] Machida S, Durelli AJ. Response of a strand to axial and torsional displacements. J
⎢ ⎥⎟⎧𝑢 ⎫ Mech Eng Sci 1973;15:241–51.
⎢ 0 0 ( ),𝑠 ⎥⎟⎪ 1 ⎪ [20] Shahsavari H, Ostoja-Starzewski M. Spectral finite element of a helix. Mech Res Com-
+⎢
0 1
( ),𝑠 0 ⎥⎟⎨𝑢2 ⎬ (A.8) mun 2005;32(2):147–52. http://dx.doi.org/10.1016/j.mechrescom.2004.05.006.
⎢1 2 ⎥⎟⎪𝑢3 ⎪ [21] Rochinha FA, Mattos HSC. Numerical modelling of the extension-
⎢ ( ),𝑠 0 0 ⎥⎟⎩ ⎭ torsion coupling in cables. Mech Res Commun 1996;23(5):511–17.
⎢ 2
⎥⎟
⎣ 0 0 0 ⎦⎠ http://dx.doi.org/10.1016/0093-6413(96)00051-1.
[22] Alani M, Raoof M. Effect of mean axial load on axial fatigue life of spiral strands.
Int J Fatigue 1997;19(1):1–11.
where Γ𝑘𝑖𝑗 stand for the Christoffel symbols defined in Eq. (A.5). The [23] Hobbs R, Raoof M. Mechanism of fretting fatigue in steel cables. Int J Fatigue
strain tensor separation of Eq. (A.8) allows for the description of the 1994;16(4):273–80. doi:10.1016/0142-1123(94)90341-7.
three-dimensional helix structural response by planar finite elements. [24] Wang D, Zhang D, Wang S, Ge S. Finite element analysis of hoisting rope and fret-
ting wear evolution and fatigue life estimation of steel wires. Eng Failure Anal
In the planar formulation, the linear operator B relates solely to the L12 2013;27:173–93. doi:10.1016/j.engfailanal.2012.08.014.
operator 𝐁 = 𝐋12 𝚽𝑇 considering invariance of the solution with respect [25] Giglio M, Manes A. Life prediction of a wire rope subjected to axial and bending
to the L3 operator, while the total strain tensor 𝝐 is given as a closed-form loads. Eng Failure Anal 2005;12(4):549–68. doi:10.1016/j.engfailanal.2004.09.002.
[26] Li H, Lan CM, Li YJDS. Experimental and numerical study of the fatigue properties
expression of the helical structure’s geometric attributes, as follows:
of corroded parallel wire cables. J Bridge Eng 2011;17(2):211–20.
[27] Chaplin CR. Interactive fatigue in wire rope applications, keynote lecture. In: Sym-
⎡ 𝜙1,1 0 0 …⎤ posium on mechanics of slender structures (MoSS 2008). Baltimore, USA; 2008.
⎢ ⎥ p. 1–12.
⎢ 0 𝜙1,2 0 …⎥ [28] Karathanasopoulos N, Angelikopoulos P. Optimal structural arrange-
⎢ 1 ⎥ ments of multilayer helical assemblies. Int J Solids Struct 2016;7879:1–8.
⎢−Γ33 𝜙1 −Γ233 𝜙1 −Γ333 𝜙1 …⎥
doi:10.1016/j.ijsolstr.2015.09.023.
𝐁 = 𝐋12 𝚽𝑇 = ⎢ 1 1 ⎥, [29] Utting WS, Jones N. The response of wire rope strands to axial tensile loads—Part i.
⎢−Γ23 𝜙1 0 𝜙
2 1,2
…⎥
Experimental results and theoretical predictions. Int J Mech Sci 1987;29(9):605–19.
⎢ 1 1 ⎥
⎢−Γ13 𝜙1 −Γ213 𝜙1 𝜙 − Γ313 𝜙1
2 1,1
…⎥ doi:10.1016/0020-7403(87)90033-6.
⎢ 1 ⎥ [30] Utting WS, Jones N. Axial torsional interactions and wire deformation
⎣ 𝜙1,2 1
𝜙 0 …⎦ in 19-wire spiral strand. J Strain Anal Eng Design 1988;23(2):79–86.
2 2 1,1 doi:10.1243/03093247V232079.
𝝐 = [0, 0, 𝑏𝜏, 𝛼𝜏, 0, 0]𝜖𝑧 (A.9) [31] Jiang WG, Yao MS, Walton JM. A concise finite element model for
simple straight wire rope strand. Int J Mech Sci 1999;41(2):143–61.
The reader is referred to [41] for the linear model derivation details. doi:10.1016/S0020-7403(98)00039-3.
[32] Karathanasopoulos N. Numerical characterization of the structural response of
helical constructions to radial and thermal loads. J Comput Methods Sci Eng
References 2016;16(4):787–800. doi:10.3233/JCM-160691.
[33] Messager T, Cartraud P. Homogenization of helical beam-like structures: ap-
[1] Kacar I, Yildirim V. Free vibration/buckling analyses of noncylindrical initially com- plication to single-walled carbon nanotubes. Comput Mech 2007;41(2):335–46.
pressed helical composite springs. Mech Based Design Struct Mach 2016;44(4):340– doi:10.1007/s00466-007-0189-3.
53. doi:10.1080/15397734.2015.1066687. [34] Fedorko G, Stanova E, Molnar V, Husakova N, Kmet S. Computer modelling and
[2] Janjic D, Pircher M, Pircher H. Optimization of cable tension- finite element analysis of spiral triangular strands. Adv Eng Softw. 2014;73:11–21.
ing in cable-stayed bridges. J Bridge Eng 2003;8(June):131–7. http://dx.doi.org/10.1016/j.advengsoft.2014.02.004.
doi:10.1061/(ASCE)1084-0702(2003)8:3(131). [35] Cartraud P, Messager T. Computational homogenization of periodic beam-like struc-
[3] Sullivan CR. Optimal choice for number of strands in a litz-wire transformer winding. tures. Int J Solids Struct 2006;43(3–4):686–96. doi:10.1016/j.ijsolstr.2005.03.063.
IEEE Trans Power Electr 1999;14(2):283–91. doi:10.1109/63.750181. [36] A. Nawrocki and M. Labrosse. A finite element model for simple straight wire rope
[4] Buln R, Hajzman M, Polach P. Nonlinear dynamics of a cablepulley system us- strands. Comput Struct 2000;77:345–59.
ing the absolute nodal coordinate formulation. Mech Res Commun 2017;82:21–8. [37] Stanova E, Fedorko G, Fabian M, Kmet S. Computer modelling of wire strands and
doi:10.1016/j.mechrescom.2017.01.001. SI: Advances/Dynamics. ropes part II: finite element-based applications. Adv Eng Softw 2011;42(6):322–31.
[5] Calim FF. Forced vibration of helical rods of arbitrary shape. Mech Res Commun doi:10.1016/j.advengsoft.2011.02.010.
2009;36(8):882–91. http://dx.doi.org/10.1016/j.mechrescom.2009.07.007. [38] Stanova E, Fedorko G, Kmet S, Molnar V, Fabian M. Finite element analysis of spiral
[6] Foti F, Martinelli L, Perotti F. A new approach to the definition of strands with different shapes subjected to axial loads. Adv Eng Softw 2015;83:45–58.
self-damping for stranded cables. Meccanica 2016;51(11):2827–45. http://dx.doi.org/10.1016/j.advengsoft.2015.01.004.
doi:10.1007/s11012-016-0444-9. [39] Treyssède F. Elastic waves in helical waveguides. Wave Motion 2008;45(4):457–70.
[7] Foti F, Martinelli L. An analytical approach to model the hysteretic bend- doi:10.1016/j.wavemoti.2007.09.004.
ing behavior of spiral strands. Appl Math Modell 2016;40(13):6451–67. [40] Frikha A, Cartraud P, Treyssède F. Mechanical modeling of helical structures ac-
http://dx.doi.org/10.1016/j.apm.2016.01.063. counting for translational invariance. Part 1: static behavior. Int J Solids Struct
[8] Foti F, Martinelli L. Finite element modeling of cable galloping vibrations. part ii: 2013;50(9):1373–82. doi:10.1016/j.ijsolstr.2013.01.010.
application to an iced cable in 1:2 multiple internal resonance. J Vibr Control 2016. [41] Karathanasopoulos N, Kress G. Two dimensional modeling of helical struc-
doi:10.1177/1077546316660017. tures, an application to simple strands. Comput Struct 2016;174:79–84.
[9] Lee WK. An insight into wire rope geometry. Int J Solids Struct 1991;28(4):471–90. doi:10.1016/j.compstruc.2015.08.016.
doi:10.1016/0020-7683(91)90060-S. [42] Laurent C, Latil P, Durville D, Rahouadj R, Geindreau C, Orgéas L, et al. Mechanical
[10] Papailiou KO. On the bending stiffness of transmission line conductors. Power De- behaviour of a fibrous scaffold for ligament tissue engineering: finite elements anal-
livery, IEEE Trans 1997;12(4):1576–88. ysis vs. X-ray tomography imaging. J Mech Behav Biomed Mater 2014;40:222–33.
[11] Costello GA. Theory of Wire Rope. New York: Springer; 1997. doi:10.1016/j.jmbbm.2014.09.003.
[12] Karathanasopoulos N. Torsional stiffness bounds of helical structures [43] Judge R, Yang Z, Jones SW, Beattie G. Full 3D finite element mod-
under the influence of kinematic constraints. Structures 2015;3:244–9. elling of spiral strand cables. Const Build Mater 2012;35:452–9.
doi:10.1016/j.istruc.2015.05.004. doi:10.1016/j.conbuildmat.2011.12.073.

374
N. Karathanasopoulos et al. International Journal of Mechanical Sciences 133 (2017) 368–375

[44] Imrak CE, Erdönmez C. On the problem of wire rope model generation with axial [54] de Borst R, Crisfield MA, Remmers JJC, Verhoosel CV. Non-Linear Finite Ele-
loading. Math Comput Appl 2010;15(2):259–68. ment Analysis of Solids and Structures. John Wiley and Sons Ltd; 2012. ISBN
[45] Jiang WG, Henshall JL, Walton JM. A concise finite element model for three-layered 9781118375938. doi:10.1002/9781118375938.
straight wire rope strand. Int J Mech Sci 2000;42:63–86. [55] Ghoreishi SR, Messager T, Cartraud P, Davies P. Validity and limitations of linear
[46] Jiang WG. A concise finite element model for pure bending anal- analytical models for steel wire strands under axial loading, using a 3D FE model.
ysis of simple wire strand. Int J Mech Sci 2012;54(1):69–73. Int J Mech Sci 2007;49(11):1251–61. doi:10.1016/j.ijmecsci.2007.03.014.
http://dx.doi.org/10.1016/j.ijmecsci.2011.09.008. [56] Jiang WG, Warby MK, Henshall JL. Statically indeterminate contacts in
[47] W.G. Jiang and J.L. Henshall. The development and applications of the helically axially loaded wire strand. Eur J Mech - A/Solids 2008;27(1):69–78.
symmetric boundary conditions in finite element analysis. Commun Numer Methods http://dx.doi.org/10.1016/j.euromechsol.2007.02.003.
Eng 1999;15:435–43. [57] Shiotani T, Oshima Y, Goto M, Momoki S. Temporal and spa-
[48] Kmet S, Mojdis M. Postelastic analysis of cable trusses. J Struct Eng 2015;141(10). tial evaluation of grout failure process with pc cable breakage by
doi:10.1061/(ASCE)ST.1943-541X.0001227. means of acoustic emission. Construct Build Mater 2013;48:1286–92.
[49] Foti F, de Luca di Roseto A. Analytical and finite element modelling of the elas- http://dx.doi.org/10.1016/j.conbuildmat.2013.04.026.
tic plastic behaviour of metallic strands under axial torsional loads. Int J Mech Sci [58] Wempner G. Mechanincs of solids with Applications to Thin Bodies. New York: Mc-
2016;115116:202–14. doi:10.1016/j.ijmecsci.2016.06.016. Graw-Hill, [1973]; 1981.
[50] Onur YA. Experimental and theoretical investigation of prestressing steel strand sub- [59] Itskov M. Tensor Algebra and Tensor Analysis for Engineers. Springer; 2007. ISBN
jected to tensile load. Int J Mech Sci 2016;118:91–100. 9783540360469.
[51] W.S. Utting and J. Norman. Tensile testing of a wire. J Strain Anal [60] Chapelle D, Bathe KJ. The finite element analysis of shells: fundamentals. Springer;
1985;20(3):151–64. 2011. ISBN 9783642164071.
[52] Wriggers P. Nonlinear Finite Element Methods. Springer Berlin Heidelberg; 2008.
ISBN 978-3-540-71000-4. doi:10.1007/978-3-540-71001-1.
[53] Simo JC, Hughes TJR. Computational Inelasticity. Springer New York; 2000. ISBN
0387975209.

375

Вам также может понравиться