Sei sulla pagina 1di 15

European Journal of Mechanics A/Solids 47 (2014) 231e245

Contents lists available at ScienceDirect

European Journal of Mechanics A/Solids


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

Analytical, numerical and experimental study of the transverse shear


behavior of a 3D reinforced sandwich structure
 a, Philippe Le Grognec a, *, Ste
phane Panier a, Christophe Binetruy a, b
Cyril Laine
a
Mines Douai, Polymers and Composites Technology & Mechanical Engineering Department, 941 rue Charles Bourseul, CS 10838,
59508 Douai Cedex, France
b
LUNAM Universit
e, Ecole Centrale de Nantes, Research Institute in Civil Engineering and Mechanics (GeM), 1 rue de la No
e, BP 92101,
44321 Nantes Cedex 3, France

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 7 November 2013
Accepted 18 April 2014
Available online 9 May 2014

Sandwich structures are known to be very sensitive to transverse shear effects when submitted to out-ofplane loads. The use of a MindlineReissner type equivalent plate model is then certainly the simplest
way to take into account these transverse shear strains that strongly inuence the global deection in
simple bending. Such a model requires the estimation of the transverse shear stiffness or of the so-called
shear correction factor. In the case of a traditional sandwich (with homogeneous foam core), this shear
correction factor is set to unity, so that the equivalent transverse shear modulus coincides with the shear
modulus of the foam core, which is fatally insubstantial. In order to improve the through-thickness
properties of sandwiches, which are governed by the core layer, use is made of thin-walled core materials or reinforcements. In these more complicated cases, the equivalent shear modulus of the core
material (in a 3D framework) highly depends on the geometry of the reinforcements and may only be
calculated numerically. Moreover, the use of this homogenized shear modulus for the heterogeneous
core layer and of a shear correction factor of unity does not generally convey to the proper value of the
transverse shear stiffness, due to the possible interactions between the reinforcements and the skins.
This paper particularly deals with sandwich structures manufactured with polymeric foam core reinforced thanks to the Napco technology (which is based on transverse needle punching) and is devoted
to obtaining their transverse shear stiffness. Bearing in mind the remarks made earlier, a one-step homogenization procedure is employed, involving simultaneously the contribution of the reinforcements to
the equivalent shear modulus of the reinforced foam core and the interactions between reinforcements
and skins. An analytical (respectively numerical) solution is derived, considering a 2D (respectively 3D)
unit cell and using the basic principle of energy equivalence. The transverse shear stiffnesses obtained by
these two simplied methods are then compared to the one obtained by a nite element numerical
computation on a whole beam-like structure for validation purposes, and nally confronted to the
experimental values resulting from 3-point bending tests performed with various volume fractions of
reinforcements.
2014 Elsevier Masson SAS. All rights reserved.

Keywords:
Transverse shear stiffness
Reinforced sandwich
Unit cell model

1. Introduction
Sandwich materials are commonly used in many applications of
aerospace, marine or transportation industries, among others, due
to the attractive combination of a lightweight and strong mechanical properties. The exural stiffness of sandwiches is indeed

* Corresponding author.
), philippe.le.grognec@
E-mail addresses: cyril.laine@mines-douai.fr (C. Laine
mines-douai.fr (P. Le Grognec), stephane.panier@mines-douai.fr (S. Panier),
christophe.binetruy@ec-nantes.fr (C. Binetruy).
http://dx.doi.org/10.1016/j.euromechsol.2014.04.006
0997-7538/ 2014 Elsevier Masson SAS. All rights reserved.

particularly signicant, thanks to the high strength of the skins and


their distance from the middle-surface of the structure. When
dealing with bending beams (respectively plates), transverse shear
effects can be neglected if the structure is almost homogeneous and
sufciently slender (respectively thin). In this case, the so-called
EulereBernoulli (respectively LoveeKirchhoff) hypotheses apply,
and only the exural stiffnesses are involved in the bending
response. Since one considers a thicker structure and/or a sandwich
composite material, these transverse shear effects can no longer be
neglected and transverse shear stiffnesses may be introduced in the
context of a Timoshenko (respectively MindlineReissner) model,
for example (Reissner, 1945; Mindlin, 1951).

232

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

The transverse shear behavior of a sandwich structure (as well


as any other out-of-plane behavior such as the through-thickness
compression) is one of the principal weaknesses of classical sandwiches, given that the corresponding stiffness is directly related to
the low mechanical properties of the soft core material. When a
sandwich structure is subjected to simple bending, the out-ofplane loads cause transverse shear strains in the core layer, that
give rise to a supplementary non-negligible deection in addition
to the classical deection associated with the exural behavior. As a
matter of fact, in the case of a homogeneous core with a very low
modulus compared to the skin one, the deection due to the
transverse shear effects may even become predominant.
In order to improve the load carrying capacity of sandwiches,
especially in the transverse shear behavior (by increasing the
equivalent transverse shear stiffnesses), without being detrimental
to lightness, the low density core layer is usually strengthened by
appropriate reinforcements. One of the simplest ways to proceed
comes down to add orthogonal reinforcements embedded in the
upper and lower skins. Among the existing methods, such as tufting,
Z-pinning and stitching (Lascoup et al., 2006), the patented Napco
technology, which is based on transverse needling, allows one to
produce tailored sandwich structures in a continuous way, while
preserving a high production efciency and a relatively low cost.
The overall purpose of this study is to analyze the mechanical
behavior of such Napco sandwich structures submitted to out-ofplane loads. The through-thickness compression has already been
 et al. (2013). Here, our
analyzed by some of the authors in Laine
attention focuses on the transverse shear behavior involved in
simple bending loading conditions. For this purpose, experimental
3-point bending tests are rst performed for various volume fractions of reinforcements (including the case of a non-reinforced
sandwich). In all cases, the signicance of the transverse shear
deection in relation to the pure exural one is emphasized, and
the increase of the equivalent transverse shear stiffness due to the
presence of reinforcements is assessed. The main objective is then
to develop analytical and numerical approaches for the determination of the effective transverse shear modulus of Napco sandwiches. Such efcient predictive tools would further be employed
to optimize the size and volume fraction of reinforcements in
relation to the transverse shear behavior.
The transverse shear behavior of beams and plates has been
studied for many years, both theoretically and experimentally. The
theoretical developments mainly concern composite laminates or
sandwiches which are most affected by transverse shear effects.
The general purpose of these works is to represent at best the
transverse shear response of such composite materials. The most
straightforward solution comes down to use a rst-order shear
deformation theory (namely a Timoshenko model for beams or a
MindlineReissner model for plates, as an example) and to derive
the corresponding equivalent transverse shear stiffness(es) of the
composite beam or plate in the context of the so-called laminate or
sandwich theory. The strong simplifying assumptions in the
transverse shear strain and stress distributions, which are supposed to be uniform in all layers throughout the thickness of the
composite structure, require the introduction of shear correction
factors that are often difcult to assess. As an alternative, higherorder shear deformation theories have gradually emerged.
Higher-order beam/plate theories are referred to as such precisely
because they involve strain and stress distributions in the section/
thickness of higher order. The objective of higher-order shear
deformation theories is to better represent the transverse shear
strain and stress elds in the beams or plates in hand in order to
better estimate the bending response of the structure without the
use of a shear correction factor. Lots of models can be found in the
literature, regarding equivalent single layer theories and rst applied

to homogeneous structures, which result most of the time from the


choice of specic kinematic hypotheses. Non-linear displacement
elds are introduced, especially in the thickness direction, using
appropriate shape functions for shear (polynomial or sinusoidal). For
instance, Barut et al. (2002) developed a higher-order plate theory
using quadratic and cubic expansions for the out-of-plane and inplane displacements, respectively. Mantari et al. (2012) recently reported many shape functions (mostly polynomial) that have been
proposed in the literature during the last century and dened a new
trigonometric one that guarantees the stress free boundary conditions on the top and bottom surfaces of the structure. As far as
laminated or sandwich structures are concerned, all these shape
functions may still be used for the overall structure in the context of
equivalent single layer theories but also in the framework of layerwise theories. In the latter case, similar shape functions are dened
for each layer of the composite material, giving rise to a global
polynomial or sinusoidal piecewise function. In both cases of rstorder and higher-order shear deformation theories, a zig-zag function can be added in order to introduce the adequate discontinuity in
the rst derivative of the displacement eld, which is called the zigzag effect. The resulting zig-zag model can also be seen as a piecewise
layer approach (see Brischetto et al. (2009) for an application of the
zig-zag theory to sandwich plates and Carrera (2003) for a general
review on the use of zig-zag functions for multi-layered plates).
Numerous other models have been developed for a more accurate determination of the transverse shear response of various
structures. Without being comprehensive, one can mention Yu
et al. (2003) who dened a 2D plate theory applicable to laminated plates based on an asymptotic analysis. Their 3D formulation
enables them to recover the 3D displacements, strains and stresses
with a very good accuracy. Nguyen et al. (2005) also proposed a 3D
approached model for thick laminates and sandwich plates. More
e and Sab (2011) presented an original Bendingrecently, Lebe
Gradient plate model which appears to be an extension of the
well-known MindlineReissner model for heterogeneous plates.
This model was successfully applied to multi-layered plates and
e and
even complex sandwich structures with cellular cores (Lebe
Sab, 2012). Moreover, such a model allows one to derive the
transverse shear stiffnesses of heterogeneous plates (including
sandwiches) using a direct homogenization procedure. Lastly,
Buannic and Cartraud (2001) made use of the asymptotic expansion
method in the context of periodic heterogeneous beams. The
consideration of higher-order terms in the developments of
displacement, strain and stress elds enabled them to improve the
classical solution corresponding to the EulereBernoulli theory
without a priori new hypotheses, unlike in standard rened beam
models.
All the higher-order beam or plate models presented above are
likely to provide the equivalent transverse shear stiffness of composite structures but also accurate information about the local
behavior at the heterogeneity scale. However, such models involve
many more degrees of freedom and give rise to increased computational costs. In this paper, the intention is not to develop or even
use a higher-order model, but rather to estimate at best the
transverse shear stiffness of a non-conventional 3D reinforced
sandwich structure, for practical use in a Timoshenko or MindlineReissner type model.
Regardless of which model is employed, lots of studies have
been carried out with the primary purpose of nding the transverse
shear stiffness of various composite structures. In the context of the
rst-order shear deformation theory, several methods have been
proposed in the literature for the determination of the shear
correction factor. These are typically homogenization or averaging
methods, based on comparisons of forces and moments between
dual problems (Altenbach, 2000) or, most often, on an energy

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

equivalence. As concerns sandwich panels, Kelsey et al. (1958)


derived rst lower and upper bounds for the transverse shear
stiffness, by applying uniform in-plane displacements or forces on
the top and bottom surfaces of the core layer, focusing on the case
of honeycomb cores. Shi and Tong (1995) also got interested in the
equivalent transverse shear stiffness of honeycomb structures using a 2D periodic unit cell. Isaksson et al. (2007) investigated the
equivalent transverse shear behavior of corrugated core structures.
Lascoup et al. (2012) derived simplied analytical expressions for
the equivalent shear moduli of stitched sandwich structures
(featuring various stitching angles), but the inuence of the foam
surrounding the stitch yarns was neglected. Also considering
structures similar to Napco sandwiches, Liu et al. (2008) analytically expressed the global stiffness tensor of pin-reinforced foam
cores (with various arrangements of pins), including notably the
transverse shear moduli. In their approach, the through-thickness
reinforcements are modeled by simply-supported beams and the
continuous core material is replaced by the superimposition of
e
horizontal and vertical elastic spring distributions. Lastly, Lebe
and Sab (2010) applied the approach from Kelsey et al. (1958) to
chevron folded cores. The same authors thoroughly discussed of
the principal questions raised by the determination of the transverse shear stiffness of sandwiches in both cases of homogeneous
e and Sab, 2012). They stated that the bounds
or cellular cores (Lebe
obtained by Kelsey et al. (1958) or any other estimations derived by
the above-mentioned approaches are not satisfactory, since the
skin effects are not (or inadequately) considered.
In the alternative of all the previous analytical solutions that only
concern relatively simple geometries and/or suffer from strong
simplifying assumptions, several semi-analytical and numerical
methods have been proposed. Xu and Qiao (2002) developed a semianalytical straightforward approach based on a multi-pass homogenization technique, using a unit cell comprising both skins and core.
With such a method, they took into account the skin-core interactions
and proved their signicance in honeycomb sandwich structures.
Hohe (2003) also suggested a direct homogenization scheme for
sandwiches instead of the classical two-step homogenization procedure where the core layer is rst homogenized separately. Finite
element computations are performed on a Representative Volume
Element of the whole sandwich and the determination of the effective stiffness matrix rests upon a strain energy equivalence principle.
Finally, Cecchi and Sab (2007) developed a MindlineReissner
equivalent plate model for orthotropic periodic plates, so as to apply
to brickwork panels. A LoveeKirchhoff model is rst identied thanks
to a numerical periodic homogenization technique, involving a 3D
unit cell. A new similar unit cell problem is then solved, whose
loading is based on the stress elds obtained in the previous calculations, in order to derive the equivalent transverse shear stiffnesses.
All the previous methods, when applied to sandwiches, properly take
into account the so-called skin effects.
This paper particularly deals with sandwich structures manufactured with polymeric foam core reinforced thanks to the

233

Napco technology and is devoted to obtaining their transverse


shear stiffness. Bearing in mind the remarks made earlier, a onestep homogenization procedure is employed, involving simultaneously the contribution of the reinforcements to the equivalent
shear modulus of the reinforced foam core and the interactions
between reinforcements and skins. First, an analytical closedform solution is sought considering a 2D unit cell model and
using the basic principle of energy equivalence. The foam core is
represented as a continuous medium whereas a beam model is
considered for the reinforcements. The skins are not directly
modeled but their presence is taken into account through
appropriate boundary conditions applied to the reinforced foam
core. Some conventions are specially suggested in order to relate
properly the 2D designed model to the real 3D conguration. A 3D
unit cell nite element model is implemented and operated for
the determination of the transverse shear stiffness based on the
same energy equivalence principle. Such a numerical model allows one to verify the relevance of some simplifying assumptions
made in the 2D analytical approach and extend the scope of the
present approach to more complicated sandwiches (with inclined
reinforcements, for instance). Finally, numerical nite element
computations are performed on complete sandwich beams in
order to validate the previous unit cell models and confront numerical reference results to the experimental values derived from
the 3-point bending tests.
2. Experimental data
2.1. Napco technology
The Napco technology is a manufacturing process of 3D
sandwich materials based on transverse needling. It consists in
strengthening the foam core of a sandwich structure by adding
orthogonal (or inclined) through-thickness reinforcements in order
to particularly enhance the out-of-plane mechanical properties. It
differs from other technologies such as stitching due to the fact that
the brous reinforcements here come from the skin material, so
that the facing fabrics (mats) and the foam core make up a
monolithic whole (see Fig. 1). In practical terms, a set of needles
regularly penetrates the sandwich structure on both sides, according to the desired pattern and density, the needles catching
and carrying yarns from the facings through the core material, as
shown in Fig. 2. Once the 3D sandwich preform is produced, it is
impregnated by a liquid resin. Among the different liquid composite molding techniques, the VARIM process (Vacuum Assisted
Resin Infusion Molding) has been retained for its efciency.
The creation of the brous reinforcements and the composite
manufacturing, associated with an experimental campaign of
measurement of geometric and material parameters, lead one to a
realistic and optimal representation of the sandwich architecture
and thus to a proper prediction of the effective mechanical properties when using appropriate analytical or numerical tools.

Fig. 1. Napco sandwiches (foam core is partly removed to show the through-thickness composite beams).

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

234

Fig. 2. Napco technology (Guilleminot et al., 2008).

Table 1
Material properties.

Young's modulus (MPa)


Poisson's ratio

transversely isotropic through-thickness reinforcing composites


(due to the unidirectional arrangement of the bers):
Polyurethane
foam

Epoxy
resin

Glass
ber

Carbon
ber

6.7
0.001

3281
0.35

72,400
0.22

290,000
0.3

2.2. Material and geometric data


The 3D sandwich specimens which will be subsequently tested
are made up of a linearly elastic isotropic closed cell polyurethane
foam (whose density is 40 kg.m3). Both facings are made of one
ply of chopped strand glass mat and one carbon [0,90] cross-ply
laminate. During the infusion process, use is made of an Epolam
5015 epoxy resin with 20% of 5015 hardener. The material data are
summarized in Table 1.
The skins are supposed to be isotropic, with equivalent Young's
modulus Es and Poisson's ratio ns. The heterogeneous throughthickness brous reinforcements are composed of aligned
isotropic bers surrounded by resin. The only case of orthogonal
reinforcements will be considered in the sequel. Therefore, such
reinforcements can be viewed as unidirectional composite columns
(UDs) and will further be represented by equivalent homogeneous
cylinders perpendicular to the skins (see Fig. 3). A preliminary
homogenization step, based on advanced mixture laws (Berthelot,
1996), is then rst performed, involving the volume fraction of
the bers within the reinforcements Vf (obtained through burn off
tests) and the material properties of both constituents (glass bers
and resin). It gives the following equivalent properties for the



EL Ef Vf Er 1  Vf


nLT nf Vf nr 1  Vf




Gf 1 Vf Gr 1  Vf



GLT Gr 
Gf 1  Vf Gr 1 Vf
0

C
C
C
C
Vf
C
C
7Gr
C

C
kr
Gr
C
3
1  Vf A

8Gr
Gf  Gr
2kr
3
Vf
KL Kr
1
1  Vf

Gf  Gr
4Gr
kf  kr
kr
3
3
2
ET
2n2
1
1

LT
2KL 2GT
EL
B
B
B
B
B
GT Gr B1
B
B
B
@

nT

ET
1
2GT

Fig. 3. Fiber reinforcements (Guilleminot et al., 2008).

(1)

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

where the quantities Ei, ni, Gi, ki and Ki represent Young's moduli,
Poisson's ratios, shear moduli, bulk moduli and transverse bulk
moduli (without longitudinal strain), respectively, and the subscripts f, r, L and T stand for the bers, the resin, the longitudinal
direction and the transverse direction. In order to simplify the
geometric representation, the cylindrical through-thickness reinforcements are supposed to have a constant circular section of
radius R.
Four different types of needle pattern have been used to create
different pile yarns densities in the nal sandwich structure. The
reference case of a sandwich material without reinforcements is
also considered for comparison purposes in further experiments
and calculations. The material and geometric parameters of the ve
panels under consideration are summarized in Tables 2 and 3,
respectively, where the subscript s stands for the skin parameters.
For information, the specic weight of each sandwich panel has
been indicated. The corresponding variation is only due to the
addition of resin during the infusion process.
2.3. 3-point bending experiments
The 3-point bending tests have been performed on a materialtesting machine (Zwick) mounted with a 10 kN-force cell,
following the NF T 54-606 standard. The samples were simply
supported near both ends and submitted to an enforced transverse
displacement at mid-span with an average speed of 1.15 mm.min1
(see Fig. 4 for the experimental set-up). For each of the ve sandwich panels, two different spans were considered, in order to
determine both the exural and transverse shear stiffnesses,
associated with the coupled pure bending and simple shear responses, respectively. For each span length, seven specimens were
tested with always the same width of 60 mm. According to the
thickness of the sandwich panels (which is approximately the same
for all densities), a total specimen length of 280 mm (respectively
500 mm) was retained for the tests with a short (respectively long)
span length of d1 230 mm (respectively d2 450 mm). All these
dimensions are sufciently large so that the specimens contain
many reinforcements, even in the case B where the volume fraction
of reinforcements is very low. Although the specimens do not
include a whole number of unit cells, which may act as representative volume elements, the volume fraction of reinforcements in
the specimens coincides thus pretty much with the theoretical
values of Table 3.
Fig. 5 plots the transverse force applied on the upper skin versus
the mid-span deection measured on the lower skin, for both short
and long specimens and for all the sandwich panels considered. In
each case, despite unavoidable imperfections, the seven curves
obtained for the seven tested specimens are very little scattered, so
much that just one curve is plotted for clarity purposes.

Panel

Density (r/dm2)

69

138

276

415

Skins
Es (MPa)

ns

11,207.5
0.364

Table 3
Geometric parameters.
Panel

Specic weight (kg/m3)


Reinforcement
radius (mm)
Reinforcement volume
fraction (%)
Foam thickness (mm)
Skin thickness (mm)

199.8
e

234.57
1.1065

277.98
1.422

350.45
1.002

428.91
1.126

11.39

16.52

20
1.138

20
1.101

2.53

20
1.132

20
1.465

9
20
1.255

The forceedisplacement curves in Fig. 5 clearly emphasize the


inuence of the volume fraction of reinforcements on the bending
behavior of the sandwich beams. Both the initial stiffness and the
failure load strongly increase with the density of brous reinforcements, whereas the failure occurs at about the same
deection, whatever the panels considered, which only depends on
the span length. The response curves of the specimens reinforced
up to 138 r/dm2 and 276 r/dm2, respectively, appear curiously very
close to each other. It is due to the presence of many duplicate
reinforcements that were observed in the panel with an expected
density of 276 r/dm2. In concrete terms, the needles arising from
both sides were not really in perfect alignment during the
manufacturing process of this particular panel, giving rise to more
numerous reinforcements but with a lower radius, whence come
the large discrepancies between the expected results and the ones
nally obtained. Lastly, the curves corresponding to the reference
sandwich panel (without reinforcements) are slightly different
from the others, as they present a plateau at the maximum load and
then a small decrease of the force before the sudden collapse. It is
due to the local core crush under the load application point and it is
all the more pronounced that the span length is important.
The exural and transverse shear stiffnesses can be derived from
the forceedisplacement curves, as explained in Dawood et al.
(2010), for instance, in the context of similar 3D glass ber reinforced polymer sandwich panels. For each panel, both tests (with
short and long specimens) are required for the determination of the
two stiffnesses. In each curve, use is made of a reference point in
the rst linear range. In all cases, the deection w at mid-span can
be viewed as the sum of two deections both depending on the
applied force F. The rst one is related to the exural behavior and
involves the corresponding exural stiffness D, and the second one
is associated to the transverse shear behavior and brings into play
the sought transverse shear stiffness S. The total deection is thus
expressed as follows:

Fd3 Fd

48D 4S

(2)

Using Equation (2) successively for both cases of a short and


long span length, with subscripts 1 and 2 respectively, one can
deduce the stiffnesses of the corresponding panel:

Table 2
Material parameters: through-thickness reinforcements and skins.

Through-thickness reinforcements
Vf (%)
e
EL (MPa)
e
nLT
e
GLT (MPa)
e
ET (MPa)
e
nT
e

235

4.01
6052.7
0.3448
1308.4
3676.9
0.4228

2.63
5098.8
0.3466
1275.5
3558.5
0.4065

3
5354.6
0.3461
1284.3
3591.3
0.4115

1.93
4615
0.3475
1259.2
3493.4
0.3957

8463.3
0.3575

8991.2
0.3616

9181.7
0.3658

8310.8
0.3615

F2 d32  d2 d21
;

48 w2  w1 FF2 dd2
1 1

F1 d1 1 
S
4 w1 

d21
d22

(3)

F d3
w2 F1 d13
2 2

The exural and transverse shear stiffnesses obtained for the


ve panels considered are listed in Table 4 and depicted in Fig. 6. In
each case, only the mean value x and the standard deviation s are
provided.
In classical sandwich structures, the exural behavior is known
to be governed by the skin behavior. It is therefore expected that

236

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

Fig. 4. 3-point bending experimental devices.

Fig. 5. Experimental forceedisplacement curves from 3-point bending tests.

the foam core, and by extension the reinforcements, has no


particular inuence on the exural stiffness D. However, experimentally speaking, this exural stiffness does not vary, as expected,
in line with the value of the skin thickness. This uncertain variation
of the exural stiffness and the associated large scattering are

probably due to the fact that, for such sandwiches, the exural part
of the deection appears negligible against the transverse shear
one, what leads to this degree of uncertainty. Conversely, the
transverse shear behavior is governed by the reinforced foam core
and thus the transverse shear stiffness should highly depend on the

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

volume fraction of reinforcements. The present results are consistent with these expectations, since the transverse shear stiffness
regularly increases with the density of the sandwich panel and,
what is more, displays a more reasonable scattering.
3. Approximate methods for the determination of the
transverse shear behavior of Napco sandwiches
3.1. Analytical resolution of the transverse shear stiffness using a 2D
unit cell
First of all, the elementary architecture of such reinforced
sandwich structures with orthogonal reinforcements allows one to
develop an analytical solution for the transverse shear stiffness.
This solution will further be compared with numerical nite
element results and confronted to the previous experimental
values for validation purposes.
3.1.1. Problem denition
In the subsequent analysis, only the transverse shear behavior is
investigated, disregarding the exural behavior also brought into
play in the simple bending response of the sandwich structure. One
thus focuses on the deformation eld in the reinforced foam core
only, considering the skins as innitely rigid. The following assumptions are then used, in order to be able to derive an explicit
expression for the sought transverse shear stiffness. A 2D repre et al. (2013), with a unit cell
sentation is retained, like in Laine
model (two half-reinforcements separated by a foam block) which
is supposed to describe the effective behavior of the global sandwich structure, once the proper periodicity conditions are prescribed (see Fig. 7). As previously mentioned, only the reinforced
foam core is rst represented, the presence of the skins being
replaced by the proper boundary conditions. Their little inuence
on the transverse shear behavior will be discussed below.
The width (2e) of the composite reinforcements is chosen in
such a way that their second moment-to-area ratio in the 2D model
is equal to the one
p of the real cylindrical reinforcements in the 3D
material e R 3=2. The same transverse shear behavior would
thus be obtained in both 2D and 3D congurations in the absence of
foam. The width (2H) of the foam block is then dened in agreement with the volume fraction of reinforcements (H e(1  Vfr)/
Vfr). This particular choice will be proven to give satisfactory results.
Finally, the global thickness (2L) is the real foam thickness
measured experimentally.
The homogeneous and isotropic foam core is considered here as
a 2D continuous solid and it is supposed to be linearly elastic (with
Young's modulus Ec and Poisson's ratio nc). The 2D model is supposed to reproduce the behavior of a panel with lateral dimensions
much larger than thickness, so that the plane strain hypothesis is
adopted. The transversely isotropic brous reinforcements (UDs)
are assumed to behave like EulereBernoulli beams, with clamped
boundary conditions, due to the entanglement of the bers into the
rigid skins. Due to these kinematic hypotheses, only the longitudinal modulus EL will be involved in the sequel among all the elastic
Table 4
Experimental values of exural and transverse shear stiffnesses.
Density

Flexural stiffness, D (N.mm2)

x
0 r/dm2
69 r/dm2
138 r/dm2
276 r/dm2
415 r/dm2

1.039
3.205
6.171
3.615
4.502







109
108
108
108
108

1.176
2.128
7.753
3.219
2.311







108
107
107
107
107

Transverse shear stiffness, S (N)


x

4682
10,060
12,025
14,028
22,184

143
141
248
458
629

237

moduli dened in Equation (1), so that the material can also be


considered as isotropic.
3.1.2. General procedure for the calculation of the transverse shear
stiffness of a sandwich structure
The calculation method is based on the classical energy equivalence principle. The unit cell is rst submitted to a prescribed
macroscopic shear strain g in such a way that only pure shear is
involved in the effective mechanical response of the sandwich.
Practically speaking, due to the extreme rigidity of the skins, this
macroscopic shear state can be achieved by enforcing two different
horizontal displacements (in the Y-direction, see Fig. 7) between the
lower and upper boundaries of the unit cell (at the interface between the foam core and the lower and upper skins, respectively).
The relative displacement between the two skins is denoted by 2d,
so as to get the relationship g d/L in small deformations (Fig. 8).
This unit cell is then successively considered as being heterogeneous (in this case, the foam block and the reinforcements are
provided with their respective moduli) and homogeneous (with an
equivalent homogenized behavior). In both cases, the total strain
energy of the unit cell is estimated, involving the mechanical parameters of the constituent materials (together with the geometric
ones) and the effective moduli, respectively. By virtue of the energy
equivalence, it is possible to express the sought effective properties,
namely the equivalent transverse shear stiffness, as a function of
the material and geometric data.
The so-called microscopic strain energy is calculated by integrating the local volume density of strain energy over the unit cell
volume U (a unit depth is retained, for simplicity purposes):

Wmicro

1
2

Z
s : 3 U

(4)

where s and 3 refer to the local stress and strain tensors,


respectively.
The energy Wmicro can be viewed as the sum of the strain energies stored in the reinforcement and the foam block. Due to the
heterogeneities, the macroscopic shear strain applied to the unit
cell may lead to any general stress/strain state at the local level, so
that all the components of tensors s and 3 must be used when
dealing with Equation (4).
Besides, the macroscopic strain energy (associated with the
equivalent material) only involves here the transverse shear
behavior, in the absence of any other macroscopic stress/strain
state. For the sake of brevity, one denes the shearing force per unit
of area Q applied to the lower and upper boundaries of the unit cell
(as represented in Fig. 8), which supposedly induces the equivalent
shear strain g. The macroscopic strain energy can then be expressed
as follows:

Wmacro

1
Q g2H 2e2L 2G* H eLg2
2

(5)

where G* stands for the effective transverse shear modulus (or


transverse shear stiffness per unit of area).
This general procedure is suitable for any sandwich conguration. It is rst applied to the case of a classical sandwich (without
reinforcements) whose results are well-known, for validation
purposes. Next, use will be made of the same method in the more
complicated case of a reinforced sandwich.
3.1.3. Case of a non-reinforced sandwich
3.1.3.1. Calculation of the effective stiffness. The main stumbling
block for the determination of the transverse shear stiffness remains the denition of a proper displacement eld within the unit

238

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

Fig. 6. Flexural and transverse shear stiffnesses for various volume fractions of reinforcements.

Fig. 7. Two-dimensional model for the analytical prediction of the transverse shear stiffness.

cell. In the case of a non-reinforced foam core (with subscript nr),


the following solution is eligible, in view of the prescribed displacements on the lower and upper boundaries:

Unr X; Y 0
Vnr X; Y gX

(6)

With such a displacement eld, the microscopic strain state is


uniform and the only non-zero strain component happens to be
3 XY g/2. Thus, the microscopic strain energy comes down to the
single transverse shear term:

Wnr

1
2

Gc g2 U 2Gc H eLg2

(7)

using the same dimensions for the unit cell as in the case of a
reinforced sandwich.
The energy equivalence principle leads to the following
expression of the effective transverse shear modulus:

G*nr Gc
Fig. 8. Description of the parameters used for the calculation of the macroscopic strain
energy.

(8)

It turns out that the effective modulus strictly coincides with the
shear modulus of the foam core. In the framework of the rst-order

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

shear deformation theory, it means that the so-called shear


correction factor is unity, which is a classical value for sandwich
structures.
3.1.3.2. Inuence of the skin thickness. The previous result (Equation (8)) shall be valid only for innitely rigid and thin faces. In the
case of Napco sandwiches, the skin modulus does not alter the
equivalent transverse shear stiffness, as it is sufciently high so that
the skins do not deform under transverse shear. Conversely, the
skin thickness t is not small enough, when compared to the core
thickness, to be neglected in the expression of the transverse shear
stiffness, whence the need for a new expression incorporating the
inuence of the skin thickness. Several corrected estimates for the
transverse shear stiffness, taking into account the thickness of the
faces, have been proposed in the literature. All of them lead to
similar numerical values. Among these, the probably most natural
solution is retained here. The basic idea is depicted in Fig. 9.
At this stage, the prescribed displacements which are responsible for the shear state in the unit cell have been enforced at the
interface between the foam core and the faces. If one considers now
the whole sandwich, including the rigid skins, the same displacements applied to the external skin boundaries may lead to the same
stress/strain distribution in the foam core and consequently to the
same microscopic strain energy. Owing to the energy equivalence
principle, the macroscopic strain energy also remains unchanged.
However, whether you include the skins in the model or not, you
may dene the macroscopic shear strain in two distinct ways, due
to the non-negligible skin thickness (Nordstrand and Carlsson,
1997). In Fig. 9, gc (previously denoted by g) corresponds to the
case where the skin effects are neglected, and gs is the new shear
strain measure dened with the skins included. Thanks to simple
geometric considerations, one obtains the following relationship:

gs

L
g
Lt c

(9)

239

3.1.4. Case of a reinforced sandwich


The same procedure is henceforth applied for the determination of the equivalent transverse shear stiffness of a reinforced
sandwich. One has to nd again an eligible displacement eld in
the reinforced foam core, which is consistent with the prescribed
boundary conditions. Owing to the high similarity between the
numerical deformed shapes of Napco sandwiches observed for
both through-thickness compression and transverse shear behaviors, the sought displacement eld here will be inspired by the
 et al. (2013) for the
buckling mode response obtained in Laine
same materials. Indeed, in both cases, the lower and upper skins
support an analogous relative horizontal displacement and the
other lateral boundaries must satisfy the same periodicity
conditions.
On one hand, along with the buckling problem, a sinusoidal
deformed shape is retained for the reinforcement. The same elds
can be used for both half-reinforcements in the unit cell, due to
the enforced periodicity conditions, so that the two half-beams
can be identied as a single entire beam. According to the
EulereBernoulli kinematics, the two following displacement elds
Ur and Vr are presupposed, standing respectively for the longitudinal and transverse displacement components on the neutral
axis:

8
< Ur X 0
(13)

: Vr X Lg sin pX
2L

The macroscopic shear strain value g appropriately appears in


the expression of the transverse displacement Vr so as to comply
with the relative horizontal displacement prescribed between the
two skins, namely 2d 2Lg. Based on this displacement eld, one
can dene the deformation eld (and therefore the stress distribution) within the reinforcement. Owing to the kinematics, the
only non-zero strain (and stress) is the axial component:

Combining the two expressions of the macroscopic strain energy (with and without the skins) leads to the following equation:

3 XX

2G* H eLg2c 2G*cor H eL tg2s

Then, the corresponding strain energy writes as follows:

(10)

since the integration volume differs between the two cases. The
new corrected expression of the equivalent transverse shear
modulus writes then:

G*cor

Lt *
G
L

(11)

In the case of a non-reinforced sandwich, one simply gets:

G*cor
nr

Lt
Gc
L

(12)

which is consistent with the suggestion from Kelsey et al. (1958).

YVr;XX

1
Wr
2

(14)

ZL Ze
EL 3 2XX YX
L e

EL p4 g2 e3
48L

(15)

still considering a unit depth.


On the other hand, solving the equilibrium equations for the
foam block with the proper boundary and continuity conditions
leads to the same displacement eld for the foam core as obtained
 et al., 2013), except for the amplitude
in the buckling analysis (Laine
which is here consistent with the displacement eld in the reinforcement:

8


Lg
pY
pY
pX
>
>
>

K
cos
K
X;
Y

sinh
Y
cosh
U
c
3
2
>
>
pH
pH
2L
2L
2L
>
>
K2 H sinh
K1 cosh
<
2L
2L


>
Lg
pY
pY
pX
>
>
>
K2 Y sinh
sin
K1 cosh
Vc X; Y
>
>
pH
pH
2L
2L
2L
>
:
K2 H sinh
K1 cosh
2L
2L

(16)

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

240

3.2. Numerical nite element resolution of the transverse shear


stiffness using a 3D unit cell

Fig. 9. Correction for the transverse shear stiffness due to the skin thickness.

with:


pH  2
pH
4L 3  4nc  eHp2 sinh
2L
2L
pH
pH
 2Lp sinh
K2 ep2 cosh
2L
2L
pH
pH
 eHp2 sinh
K3 2LpH 3e  4enc cosh
2L
2L
K1 2LpH cosh

(17)

This solution has been obtained with zero stress boundary


conditions on the lower and upper boundaries of the foam block (at
the interface with the two skins, respectively), instead of
displacement boundary conditions (namely, uniform horizontal
prescribed displacements), that would have naturally better rep et al. (2013), the present
resented the presence of the skins. In Laine
choice allowed us to derive an explicit expression for the critical
load under through-thickness compression with a good accuracy, at
least for sufciently high volume fractions Vfr  10%. Here, such a
simplifying assumption also makes it possible to obtain an
analytical solution for the transverse shear stiffness. It will also
most likely limit the validity domain to rather high volume fractions of reinforcements.
In-plane strains and stresses may be then deduced from the
previous expressions of the displacement eld, eventually resulting
in the strain energy within the foam block:

Wc

1
2

ZL ZH h

Three-dimensional numerical nite element computations


have been performed, using Abaqus software, in order to supplement the previous analytical solutions. A hexagonal arrangement of the reinforcements has been retained, according to the
experimental patterns. Thus, the overall mechanical response of
the composite reinforced foam core (and consequently of the
sandwich) is transversely isotropic. A 3D unit cell is only
considered, for efciency purposes, but here including the skins
and the full material properties of the transversely isotropic reinforcements. The geometry of the plane-parallel unit cell and
the associated nite element mesh (made up of 20-noded hexahedral elements with reduced integration) are depicted in
Fig. 10. The boundary conditions are prescribed in a similar way
than in the previous 2D analysis. Periodicity conditions are
enforced on the lateral faces of the unit cell in both directions.
Lastly, the bottom and top faces of the sandwich cell are subjected to different uniform horizontal displacements (d 6 mm
and d 6 mm, respectively), so as to produce a pure macroscopic
shear state in the unit cell (see Fig. 11 for the loading conditions
and the deformed shape of the 3D unit cell under pure transverse
shear).
With such a numerical model involving the same geometric
and material parameters together with the same boundary
conditions as in the previous section, it is possible to re-use the
same procedure for the calculation of the effective transverse
shear stiffness based on the energy equivalence principle.
Whereas the expression for the macroscopic strain energy remains almost the same (Wmacro 1/2VG*g2 where V is the unit
cell volume without considering the skins), the microscopic
strain energy is hereafter estimated thanks to a numerical
computation (by the way, the contribution of the skins in the
total strain energy is proved to be negligible). This 3D numerical
solution is then expected to be far more accurate than the 2D
analytical one. Indeed, there are not here as many simplifying
assumptions concerning the skins and the corresponding
boundary conditions, the geometry and the full material properties, and the global kinematics. Finally, the obtained effective
modulus G* is replaced by the corrected value G*cor through the
same thickness ratio as before, if necessary.



lc 2mc 3 2XX 3 2YY 2lc 3 XX 3 YY

L H

i
4mc 3 2XY Y X

(18)

 coefcients of the core material.


where lc and mc are the Lame
The total microscopic strain energy of the reinforced foam core
is obtained by simply adding the two strain energies related to the
reinforcement (Wr) and the foam block (Wc). Based on this strain
energy value, it is possible to deduce the effective transverse shear
modulus G*r by using the same denition of the macroscopic strain
energy as in the previous case of a non-reinforced foam core. This
equivalent transverse shear stiffness may only be used for particularly thin faces. Otherwise, when the skin thickness is no longer
negligible, it can be upgraded by multiplying by the thickness ratio
(L t)/L.
For all these quantities, explicit solutions have been obtained
using Maple symbolic calculation software, but they are too
cumbersome to be presented as closed-form expressions.

Fig. 10. Model for the 3D nite element computations.

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

241

Fig. 11. Transverse shear of the 3D unit cell.

4. Numerical and experimental validation


4.1. Numerical computation of the transverse shear stiffness of a
beam under simple bending
Both previous analytical and numerical approaches are based on
several simplifying assumptions and must be therefore validated
using numerical computations on complete structures. In this
section, nite element calculations are performed, involving 3D
heterogeneous beams under simple bending, for validation purposes. The post-processing of the associated numerical results
naturally leads to a new estimate for the transverse shear stiffness.

4.1.1. Basic principle


The method used here for the determination of the transverse
shear stiffness from a full 3D numerical simulation has already been
used in many studies (see, for example, Buannic et al. (2003)). Let us
consider a cantilever sandwich beam, built-in at the left-hand side
and submitted to a transverse force at the right-hand side. In the
present case, the whole beam is represented which consists of
several unit cells placed end to end. A nite element model is
implemented and allows one to determine the mean deection
along the beam. This deection can be viewed as the sum of a
deection due to the bending moment effects (only depending on
the exural stiffness) and another deection due to the transverse

Fig. 12. Three-dimensional beam model for the numerical validation.

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

242

Fig. 13. Three-dimensional sandwich beam under simple bending.

shear effects. The rst deection can be analytically expressed,


using the in-plane effective moduli of the corresponding sandwich
structure, so that the second sought deection can be obtained by
deducting the rst one from the total deection which has been
numerically evaluated. This deection is supposed to be linear with
respect to the longitudinal coordinate along the beam, in such a
way that the transverse shear stiffness can be nally derived from
the associated slope.
4.1.2. Numerical model
The complete beam consists of 10 unit cells placed side to side in
the longitudinal direction. Each unit cell is dened in the same way
as before, in terms of geometry and materials. The common surfaces between adjacent unit cells are merged in order to build a
whole structure. The global model is made up of 20-noded hexahedral elements with a coarser mesh as before, for efciency
purposes (see Fig. 12). The best compromise leads to a minimum of
34,560 elements (461,127 d.o.f.) for the most reinforced case (415 r/
dm2) and 61,440 elements (816,855 d.o.f.) in the opposite case
where the dimensions of the beam are the most important (69 r/
dm2).
The following boundary conditions are then applied. At the left
end of the beam (x 0), the whole section is xed. Conversely, at
the right end section (x l), the transverse applied force is uniformly distributed onto the skin surfaces only, with an arbitrary
amplitude. The loading and boundary conditions are depicted in
Fig. 13, together with the deformed shape of the total beam.



2
2
Erc bL3 Es b t 3 3t 2 L 3tL2
3
3

where b is the width of the beam and Erc stands for the equivalent
longitudinal modulus of the reinforced core.
In practice, the core modulus is substantially below the skin
modulus, even in the case of a reinforced foam core (see
Guilleminot et al. (2008) for more details). In addition, the core
thickness is much higher than the skin thickness, so that the general expression in Equation (21) can be simplied in the following
way:



D 2Es b t 2 L tL2

T
S

wtotal wflex wshear

and the equivalent transverse shear modulus is

The rst part is due to the bending moment and writes analytically
as follows:

wflex



T x3
x2
l
D 6
2

(20)

where T is the downward transverse force (counted as positive).


The exural stiffness D is also evaluated analytically. The most
general expression for the exural stiffness of a sandwich beam
takes the following form (Zenkert, 1992):

(22)

The values obtained when using Equation (22) are found to be in


very good agreement with numerical results deriving from a periodic homogenization nite element analysis, for all the sandwich
panels considered.
Since the exural stiffness is known, it is possible to plot the
corresponding deection wex, which is shown to be small in
comparison with the total deection. The difference between them,
namely the deection wshear due to transverse shear effects, is thus
predominant and appears to be linear, as depicted in Fig. 14. It
proves essential to determine the associated transverse shear
stiffness, using the slope of the linear deection obtained by linear
regression. This slope happens to be the macroscopic shear strain g
and is related to the transverse force T as follows:

4.1.3. Determination of the transverse shear stiffness


The estimation method of the transverse shear stiffness is
illustrated in Fig. 14 and can be summarized as follows.
The previous numerical computation allows us to plot the
average deection along the neutral axis of the beam (wtotal). This
total deection can be divided into two parts:

(19)

(21)

(23)

so that the transverse shear stiffness writes

T
g

G*

(24)

T
2gbL t

(25)

4.2. Comparison between analytical, numerical and experimental


results
Finally, the two simplied methods for the calculation of the
transverse shear stiffness, respectively analytical using a 2D unit
cell and numerical using a 3D unit cell, are validated by comparison

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

243

Fig. 14. Relative inuence of the exural and transverse shear stiffnesses on the deection of a beam under simple bending.

Fig. 15. Comparisons between transverse shear stiffnesses obtained with different analytical, numerical and experimental methods.

to both numerical results obtained with a complete beam under


simple bending and experimental 3-point bending test results. All
the transverse shear stiffnesses by unit width (expressed in
N.mm1) are plotted in Fig. 15 for the four Napco sandwiches
considered as well as the non-reinforced panel.
First, considering the non-reinforced sandwich, the four
methods give rise to very similar values. Thus, the reference
panel (without reinforcements) makes it possible to check the
consistency between the different analytical, numerical and
experimental approaches. The different hypotheses formulated in
each case are conrmed, at the very least in the absence of
reinforcements.

Regarding now the reinforced sandwiches, one can notice the


very good agreement between both numerical approaches (the
relative error has an average value of 2.4% and does not exceed 6%).
The general method based on the energy equivalence principle is
thereby validated. The analytical solution is also shown to be in
good accordance with the reference numerical results, at least for
the three higher densities of reinforcements. In these three cases,
the relative error is around 6% on average. The main difference
between the 2D analytical and 3D numerical unit cell models (other
than the dimension) lies in the consideration of the skins in the
latter. Additional numerical nite element computations have been
performed on 2D unit cells, including the skins. The results almost

244

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245

coincide with the ones obtained with 3D unit cells, what proves the
reliability of the assumptions made when switching from the 3D
conguration to the 2D one.
The inuence of the skin thickness on the transverse shear
stiffness value has already been discussed. Apart from that, the
extreme rigidity of the skins allows one to apply the same
displacement boundary conditions on the core/skins interfaces as
enforced in practice onto the external skin boundaries. Instead,
approximated boundary conditions have been retained in the
analytical approach, that do not match the real conditions in the
presence of skins, in order to simplify the kinematics used in the
analytical resolution and make possible the achievement of a
closed-form expression for the transverse shear stiffness. This
choice of boundary conditions is not so detrimental, as soon as the
volume fraction of reinforcements is about 10% or higher. On the
contrary, with the smallest density, a large discrepancy is noticed
(about 47.5% of relative error) between the analytical and numerical
predictions, what points out the limitations of the present analytical model. When brous reinforcements are much less numerous
and very distant from each other, the skin effect becomes more and
more apparent in the transverse shear deformation shape and thus
in the corresponding stiffness.
Lastly, the experimental results are confronted to the analytical
and numerical solutions. It is difcult to quantify and explain the
discrepancies observed between the experimental measurements
and the theoretical predictions, since the analytical and numerical
models are somewhat idealized. A perfect architecture is retained
in the modeling, without any imperfection, and uniform volume
fractions and mechanical properties are considered throughout the
sandwich structure. Furthermore, in the most reinforced case, the
reason why the discrepancy is so high might be the following. Due
to the numerous reinforcements, small cracks may appear in the
foam along the needle track. Then, during the infusion process,
resin may spread into the cracks and strengthen the foam core and
therefore the whole sandwich. Despite all that and independently
of the unavoidable imperfections in practice, the different approaches presented above provide good estimations for the transverse shear stiffnesses of most of the Napco sandwiches tested in
this study.
5. Conclusions
The Napco technology is a patented process that transversally
strengthens the foam core of a sandwich structure with ber
yarns taken from facings. In this study, we investigated the potential of such a reinforced sandwich in its transverse shear
behavior, which plays a signicant role in the simple bending
response of a sandwich structure. First, an analytical solution for
the transverse shear stiffness has been proposed. A 2D model was
conveniently dened in which only a unit cell of the reinforced
foam core was considered, due to the material periodicity. The
reinforcements were assumed to behave like EulereBernoulli
beams whereas the foam core was modeled as a 2D continuous
solid, without considering any simplied deformation eld. The
transverse shear stiffness was derived from the energy equivalence principle, by comparing the microscopic strain energy
induced by a macroscopic pure shear loading (using the appropriate boundary conditions) and the macroscopic strain energy of
the sought effective material. The reinforcements naturally
strengthen the foam core, especially in its transverse shear
behavior. However, the coupling effects between the reinforcements and the skins (due to the manufacturing process)
make the solution here far more complicated than the classical
one based on the equivalent shear modulus of the reinforced core
viewed as a 3D material with an innite thickness.

The same method was developed in the context of a 3D unit cell


(including the skins) using nite element calculations. This
approach, though numerical, is an efcient way to obtain the
transverse shear stiffnesses of sandwich structures, even in more
complicated cases than orthogonal reinforcements (for instance,
with other distributions and/or orientations of the throughthickness reinforcements), where analytical solutions are no more
available. Numerical computations were also performed on a
complete beam under simple bending, for validation purposes.
Experimental 3-point bending tests have been performed for
ve different sandwich panels (with various densities of reinforcements, including a non-reinforced case). Comparisons between analytical/numerical predictions and experiments were
discussed and clearly showed the accuracy of the theoretical
models. In particular, the expression obtained for the transverse
shear stiffness is proved to be suitable to properly predict the
transverse shear behavior of such sandwiches, as long as the volume fraction of reinforcements is sufciently high, say greater than
10%.
Acknowledgments
The authors are indebted to the French Ministry of Economy,
Finance and Industry (NWC-X project, Contract no. 09 2 90 6242)
for its nancial support.
References
Altenbach, H., 2000. An alternative determination of transverse shear stiffnesses for
sandwich and laminated plates. Int. J. Solids Struct. 37 (25), 3503e3520.
Barut, A., Madenci, E., Anderson, T., Tessler, A., 2002. Equivalent single-layer theory
for a complete stress eld in sandwich panels under arbitrarily distributed
loading. Compos. Struct. 58 (4), 483e495.
riaux Composites: Comportement Me
canique et Analyse
Berthelot, J.M., 1996. Mate
des Structures. Masson.
Brischetto, S., Carrera, E., Demasi, L., 2009. Improved bending analysis of sandwich
plates using a zig-zag function. Compos. Struct. 89 (3), 408e415.
Buannic, N., Cartraud, P., 2001. Higher-order effective modeling of periodic heterogeneous beams. I. Asymptotic expansion method. Int. J. Solids Struct. 38
(40e41), 7139e7161.
Buannic, N., Cartraud, P., Quesnel, T., 2003. Homogenization of corrugated core
sandwich panels. Compos. Struct. 59 (3), 299e312.
Carrera, E., 2003. Historical review of zig-zag theories for multilayered plates and
shells. Appl. Mech. Rev. 56 (3), 287e308.
Cecchi, A., Sab, K., 2007. A homogenized ReissnereMindlin model for orthotropic
periodic plates: application to brickwork panels. Int. J. Solids Struct. 44 (18e19),
6055e6079.
Dawood, M., Taylor, E., Rizkalla, S., 2010. Two-way bending behavior of 3-D GFRP
sandwich panels with through-thickness ber insertions. Compos. Struct. 92
(4), 950e963.
Guilleminot, J., Comas Cardona, S., Kondo, D., Binetruy, C., Krawczak, P., 2008.
Multiscale modelling of the composite reinforced foam core of a 3D sandwich
structure. Compos. Sci. Technol. 68 (7e8), 1777e1786.
Hohe, J., 2003. A direct homogenisation approach for determination of the stiffness
matrix for microheterogeneous plates with application to sandwich panels.
Compos. Part B: Eng. 34 (7), 615e626.
Isaksson, P., Krusper, A., Gradin, P.A., 2007. Shear correction factors for corrugated
core structures. Compos. Struct. 80 (1), 123e130.
Kelsey, S., Gellatly, R.A., Clark, B.W., 1958. The shear modulus of foil honeycomb
cores: a theoretical and experimental investigation on cores used in sandwich
construction. Aircraft Eng. Aerospace Technol. 30 (10), 294e302.
, C., Le Grognec, P., Comas Cardona, S., Binetruy, C., 2013. Analytical, numerical
Laine
and experimental study of the bifurcation and collapse behavior of a 3D reinforced sandwich structure under through-thickness compression. Int. J. Mech.
Sci. 67, 42e52.
Lascoup, B., Aboura, Z., Khellil, K., Benzeggagh, M.L., 2006. On the mechanical effect
of stitch addition in sandwich panel. Compos. Sci. Technol. 66 (10), 1385e1398.
Lascoup, B., Aboura, Z., Khellil, K., Benzeggagh, M.L., 2012. Prediction of out-of-plane
behavior of stitched sandwich structure. Compos. Part B: Eng. 43 (8),
2915e2920.
e, A., Sab, K., 2010. Transverse shear stiffness of a chevron folded core used in
Lebe
sandwich construction. Int. J. Solids Struct. 47 (18e19), 2620e2629.
e, A., Sab, K., 2011. A bending-gradient model for thick plates. Part I: Theory. Int.
Lebe
J. Solids Struct. 48 (20), 2878e2888.
e, A., Sab, K., 2012. Homogenization of cellular sandwich panels. C. R. Me
c. 340
Lebe
(4e5), 320e337.

C. Laine et al. / European Journal of Mechanics A/Solids 47 (2014) 231e245


Liu, T., Deng, Z.C., Lu, T.J., 2008. Analytical modeling and nite element simulation of
the plastic collapse of sandwich beams with pin-reinforced foam cores. Int. J.
Solids Struct. 45 (18e19), 5127e5151.
Mantari, J.L., Oktem, A.S., Soares, C.G., 2012. A new trigonometric shear deformation
theory for isotropic, laminated composite and sandwich plates. Int. J. Solids
Struct. 49 (1), 43e53.
Mindlin, R.D., 1951. Inuence of rotatory inertia and shear on exural motions of
isotropic, elastic plates. J. Appl. Mech. 18, 31e38.
Nguyen, V.T., Caron, J.F., Sab, K., 2005. A model for thick laminates and sandwich
plates. Compos. Sci. Technol. 65 (3e4), 475e489.
Nordstrand, T.M., Carlsson, L.A., 1997. Evaluation of transverse shear stiffness of
structural core sandwich plates. Compos. Struct. 37 (2), 145e153.

245

Reissner, E., 1945. The effect of transverse shear deformation on the bending of
elastic plates. J. Appl. Mech. 12, 69e77.
Shi, G., Tong, P., 1995. Equivalent transverse shear stiffness of honeycomb cores. Int.
J. Solids Struct. 32 (10), 1383e1393.
Xu, X.F., Qiao, P., 2002. Homogenized elastic properties of honeycomb sandwich
with skin effect. Int. J. Solids Struct. 39 (8), 2153e2188.
Yu, W., Hodges, D.H., Volovoi, V.V., 2003. Asymptotically accurate 3-D recovery
from Reissner-like composite plate nite elements. Comput. Struct. 81 (7),
439e454.
Zenkert, D., 1992. The Handbook of Sandwich Construction. Engineering Materials
Advisory Services.

Potrebbero piacerti anche