You are on page 1of 13

Engineering Structures 106 (2016) 209–221

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Connection systems between composite sandwich floor panels


and load-bearing walls for building rehabilitation
Mário Garrido a,⇑, João R. Correia a, Thomas Keller b, Fernando A. Branco a
a
ICIST, CEris, DECivil, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais 1, 1049-001 Lisbon, Portugal
b
Composite Construction Laboratory (CCLab), École Polytechnique Fédérale de Lausanne, BP 2225, Station 16, CH-1015 Lausanne, Switzerland

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

Article history: The use of advanced composites for building rehabilitation presents several advantages when compared
Received 6 March 2015 with traditional construction materials. When degraded building floors need to be replaced, composite
Revised 21 October 2015 sandwich panels are a potentially interesting solution, namely for buildings with load-bearing rubble
Accepted 22 October 2015
masonry walls. In this paper, connection systems between composite sandwich floors and load-bearing
Available online 11 November 2015
walls are proposed, and their behaviour under vertical loading is investigated. The systems comprise steel
angles anchored to the walls, serving as main supports of the sandwich panels, which are then adhesively
Keywords:
bonded and/or bolted to the angles. These connection systems are experimentally assessed using
Composites
Sandwich panels
sandwich panels made of glass-fibre reinforced polymer (GFRP) face sheets and cores of either polyur-
Floors ethane (PUR) foam or balsa wood, by means of flexural tests on cantilevers, which are also simulated
Building rehabilitation using non-linear finite element models. The structural response of the connection systems is determined,
Connections including the rotational stiffness conferred to the floors, the strength and the failure modes. Moment–
GFRP rotation relationships are obtained for the connection systems and sandwich panel types considered,
Polyurethane foam which provide a wide range of rotational stiffness values, from 60 to 10,856 kNm/rad per unit width
Balsa wood (m). These are then used to analytically estimate the short-term mid-span deflections of floors with
semi-rigid connections and spans ranging between 2 m and 5 m. It is shown that some of the proposed
connections allow significant floor stiffness increases compared with simply supported conditions, with
reductions in total mid-span deflection of up to 65% being achieved for a span of 4 m. The results obtained
for the proposed connections highlight (i) their potential benefits for fulfilling serviceability limit states
and (ii) the importance of considering an adequate structural model when designing sandwich floor panels.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction Lightweight systems have high potential for building rehabilita-


tion, as the additional dead load transmitted to the existing
Building rehabilitation often requires the structural strengthen- structure is limited. This is particularly relevant in the case of
ing or the replacement of structural elements. In old buildings building floors, which typically represent a very significant portion
made of masonry walls wooden floors are among the structural of the total structural mass of buildings [2,3]. The use of sandwich
elements that more frequently need to be replaced, as they often construction, characterised by two relatively thin and stiff faces
suffer from excessive deformations and/or are not able to comply and a relatively thick and lightweight core, may be an interesting
with current structural performance requirements [1]. Traditional solution for building floor rehabilitation. Currently there are
rehabilitation solutions involve the construction of new timber several types of sandwich panels commercially available for civil
floors or the adoption of different floor systems made of either engineering applications, which typically comprise steel facings
reinforced concrete (RC), steel or composite (steel–concrete or tim- and high density polymeric cores [4,5]. However, such solutions
ber–concrete) elements. However, wooden floors have limited are still considerably heavier than typical timber floors, and are
durability, whereas RC, steel and composite solutions substantially used mostly in new construction.
increase the structural mass, generally making it necessary to Fibre-reinforced polymer (FRP) composites, and particularly
strengthen the building walls, especially in seismic regions [2]. FRP sandwich panels, may be a viable alternative lightweight solu-
tion for the rehabilitation of building floors [6,7], presenting
⇑ Corresponding author. Tel.: +41 76 361 19 24. advantages over panels with steel facings due to their lightness,
E-mail address: mariogarrido@tecnico.ulisboa.pt (M. Garrido). higher durability and limited maintenance [8]. In fact, composite

http://dx.doi.org/10.1016/j.engstruct.2015.10.036
0141-0296/Ó 2015 Elsevier Ltd. All rights reserved.
210 M. Garrido et al. / Engineering Structures 106 (2016) 209–221

sandwich panels are being increasingly used in civil engineering


structural applications and have already been successfully applied
in roof structures [9,10], bridge decks [11] and building façades
[12]. Their potential use in building floors has also been suggested
by several authors [6,13,14].
In the development of a sandwich floor panel several structural
behaviour aspects must be considered, including the connections
between the floor elements and their vertical supports. In old
buildings, such supports are frequently load-bearing rubble
masonry walls. This paper addresses this subject, with the objec-
tive of: (i) proposing connection systems for composite sandwich
floor panels for use in building rehabilitation, (ii) experimentally
and numerically assessing their structural behaviour and perfor-
mance under vertical loading, and (iii) providing an analytical
method to calculate the deflections of sandwich panels supported
using the proposed connection systems. It is worth mentioning
that while the investigated connection systems are also expected
to transfer horizontal loads, their behaviour under such actions is
Fig. 1. Connection between a new timber floor and timber-framed masonry walls
beyond the scope of the current study.
using steel angles.
The different connection systems proposed in this paper com-
prise anchoring steel angles to the masonry walls, which in turn
support the sandwich floor panels via adhesively bonded, bolted, angle). Three different methods were used to join the panels and
and mixed (adhesive and bolted) connections. Sandwich panels the steel angles: (i) adhesive bonding (Fig. 2a and b), (ii) bolting
made of (i) glass-fibre reinforced polymer (GFRP) faces and (ii) (Fig. 2c and d), and (iii) a combination thereof (Fig. 2e).
cores made of polyurethane (PUR) foam and balsa wood are con-
sidered. These connections are experimentally tested under verti- 3. Experimental investigation
cal loads in a cantilever configuration, loaded at the panels’ free
edge, and connected to a closed steel frame to simulate a rigid 3.1. Test setup and materials
load-bearing wall. The tests are simulated using finite element
(FE) models, in order to understand in further depth the stress dis- A large sandwich panel of 3560 mm length  1250 mm
tributions within the connection components, and to obtain the width  134 mm thickness was manufactured by vacuum infusion.
moment–rotation (M–h) relationships that characterise each con- After curing, the sandwich panel was cut into specimens with a
nection system. Finally, an analytical method is suggested to calcu- length of 850 mm and width of 250 mm. The panels1 comprised
late the maximum deflections of sandwich panels when supported 7 mm thick GFRP face sheets (nominal dimensions) enclosing a
using the proposed connection systems. 120 mm thick core, made of either (i) rigid PUR foam, or (ii) balsa
wood. Table 1 presents a summary of the most relevant properties
of the constituent materials used in the sandwich panels (obtained
2. Description of the connection systems from material characterisation tests).
One of the extremities of the test panels was supported in a sin-
The primary function of floor-to-wall connection systems is to gle cantilever configuration and a point load was applied at the free
guarantee the transfer of vertical and horizontal (seismic and edge, as illustrated in Fig. 3. The angles that supported the sand-
wind) loads between the floors and the load-bearing walls. In addi- wich panels were connected to a closed steel frame comprised of
tion, it is also useful to guarantee some rotation restriction at these HEB 300 profiles. These had a bending stiffness equivalent to a
connections, as this reduces the floor’s flexibility, i.e. the maximum rigid rubble masonry load-bearing wall with a thickness of 1 m
deflections along the span. This can be a significant advantage and a Young’s modulus of 2 GPa (approximately 50,000 kN/m2).
given that maximum allowable deformability criteria are usually The angles consisted of L-150  12 profiles of S275 JR grade steel,
the limiting factor in the design of FRP composite sandwich floor with a leg width of 150 mm and wall thickness of 12 mm. These
panels [6]. were cut to a width of 300 mm, and bolt holes were drilled accord-
In the rehabilitation of old timber floors one of the typical con- ing to the specifications presented in Fig. 3. M10 bolts were
nection solutions comprises embedding the new joists in the load- machined and threaded to the required length from smooth S275
bearing walls and/or using steel angles to anchor them to the walls JR grade steel bars and used to connect the sandwich panels to
(Fig. 1). These steel angles provide additional support length and the steel angles. An epoxy adhesive supplied by Sika AG (Sikadur
in-plane stiffness to the floors [15,16], while also contributing to 31 EF) was used for the adhesively bonded connections, all pre-
improve the out-of-plane behaviour of the exterior masonry walls senting a thickness of 2 mm (guaranteed using appropriate
[17]. The connection systems proposed in this paper, illustrated in spacers).
Fig. 2, are based on that practice, with the steel angles acting as the Regarding the test instrumentation, vertical displacements
main supporting element. The option of embedding the panels were measured at the bottom face of the panels at the load appli-
inside the walls was discarded, due to the continuous nature of this cation point (D1) and at a cross-section distanced 135 mm
connection (as opposed to the discrete embedding of timber joists) (approximately the same as the panel thickness) from the edge
– its implementation would significantly affect the walls’ struc- of the support (D2), using TML CDP-100 displacement transducers,
tural integrity and would be very labour intensive. Therefore, to with a stroke of 100 mm and precision of 0.01 mm. Displacement
increase the rotation stiffness of the connection, a second steel
angle connected to the top face of the sandwich panels is consid- 1
For the adopted aspect ratio, the specimens behave more closely to beams (or
ered in addition to the bottom steel angle where the floor panels one-way slabs) rather than bidirectional panels. However, the current study envisages
are supported. These steel angles can be covered/embedded by the use of floor panels similar to those described in [18], and consequently the term
the footer (top angle) or the moulding/suspended ceiling (bottom ‘‘panel” is used to keep consistency with the topic.
M. Garrido et al. / Engineering Structures 106 (2016) 209–221 211

(a) (b)

A1 A2
(c) (d) (e)

B1 B2 AB2
Fig. 2. Proposed connection systems (A – adhesive, B – bolted, 1 – single bottom steel angle, 2 – top and bottom steel angles): (a) A1, (b) A2, (c) B1, (d) B2, and (e) AB2.

Table 1 0.03–0.05 kN/s for the single steel angle connections (series A1
Properties of materials used in the sandwich panels. and B1) and 0.07–0.10 kN/s for the double steel angle connections
Material Young’s modulus, E Tensile strength, rmax (series A2, B2 and AB2).
(GPa) (MPa) Table 2 details the number of specimens tested (16 in total)
GFRP according to the type of connection and core material. The two
E-glass fibres 29.4 ± 0.8 437.3 ± 28.1 steel angle configurations included the higher number of replicate
Orthophthalic polyester specimens (2 or 3), since they were expected to present the best
resin
[0/0/30/-30/90/0]S
performance and hence the highest potential for practical applica-
tions. For the remaining configurations, due to material limitations,
Shear modulus (G), Shear strength (smax), it was only possible to test either 1 or 2 specimens.
MPa MPa
PUR foam 3.2. Results and discussion
Polirígido Lda. 8.7 ± 1.0 0.32 ± 0.06
87.4 kg/m3
3.2.1. Overview
Balsa wood
Baltek SB50, 3A 48.8 ± 6.2 0.93 ± 0.19
Fig. 4 illustrates the load vs. D1 displacement curves obtained
Composites AG for all connection systems. The displacements measured at D2
101.4 kg/m3 and the rotations measured at I1 exhibited very similar qualitative
developments and were consistent with D1 displacements, hence
they are not shown. Table 3 presents a summary of the failure
at D1 was also measured using a TML DP-500E wire displacement (or ultimate) loads and initial stiffness at D1 and initial rotational
transducer (500 mm stroke, precision of 0.1 mm), in order to mea- stiffness at I1. Detailed discussion is provided in the following sec-
sure the high displacements attained in some tests. Additionally, tions for each type of connection.
the panels rotation (I1) was measured on the top face sheet at a
cross-section distanced 60 mm from the edge of the support using 3.2.2. Adhesively bonded connections (A1, A2)
a TML KB-10D inclinometer, with a range of ±10° (precision of The specimens with an adhesively bonded connection system
0.01°). The applied load was measured using a Novatech 100 kN and a single bottom angle support (A1) presented approximately
load cell (precision of 0.01 kN). linear elastic behaviour up to failure, which occurred for relatively
Load was applied with a 300 ton Enerpac hydraulic jack, low load values. The failure modes observed in this series are
reacting against the rigid steel frame. A load distribution steel shown in Fig. 5. Failure occurred in a brittle manner and for both
plate (15 mm thick, 250 mm long and 60 mm wide) and a roller types of cores it initiated at the support length, near the bottom
were positioned between the test specimen and the hydraulic jack. face sheets, either at the interface between the steel angle and
Specimens were monotonically loaded up to failure, using a the adhesive (A1-BAL-1), or at the interface between the core
manually controlled walter+bai PKNS 19D control console. The and the bottom face sheet (specimens A1-BAL-2 and A1-PUR-1).
different responses of the single and double steel angled configura- Adding a top steel angle to the adhesively bonded connection
tions, namely in terms of their stiffness values, made necessary the systems (A2) increased substantially the initial stiffness (about 9
adoption of different loading speeds for the two cases in order and 15 times for the panels with PUR foam and balsa wood cores,
to maintain consistent test durations. The experiments had respectively), as well as the ultimate load (approximately 7 and 6
durations ranging between 3 and 6 min, thus minimising effects times, respectively). The balsa wood specimens presented
stemming from the time dependency of the materials’ properties. nonlinear behaviour, characterised by a gradual loss of stiffness,
This was achieved by adopting approximate loading speeds of until the brittle failure of the foremost part of the adhesive bond
212 M. Garrido et al. / Engineering Structures 106 (2016) 209–221

Fig. 3. Experimental setup and instrumentation: (a) schematic and (b) general view.

Table 2 3.2.3. Bolted connections (B1, B2)


Specimen distribution per connection system and core material. The B1 specimens presented an approximately linear-elastic
Connection system No. of specimens response. For the panel with balsa wood core, plastic deformations
Series Description PUR core Balsa core
were observed in the steel angle (Fig. 6a) and bolts for loads above
3.5 kN. Regarding the panel with PUR foam core, the steel angle did
A1 Adhesively bonded, single steel angle 1 2
A2 Adhesively bonded, two steel angles 2 2
not present this type of deformation and the panel showed clear
B1 Bolted, single steel angle 1 1 signs of foam crushing between the bolted area and the support’s
B2 Bolted, two steel angles 3 3 edge, as illustrated in Fig. 6b. Both tests were interrupted as a
AB2 Adhesive and bolts, two steel angles 1 – safety measure in order to protect the test instrumentation;
although collapse had still not occurred, specimens exhibited very
large deformations at this stage and the bolts presented plastic
deformations at the interface zone between the steel angles and
(50–70% of the bond area) between the top face sheet and the top the bottom face sheet of the panels.
steel angle occurred, which caused a sudden load reduction. Subse- As in the adhesively bonded connection systems, adding a top
quently, the connections were still able to carry some load (yet, steel angle to the bolted connections (B2) provided significantly
lower than the failure load), while the cracking gradually opened higher stiffness (approximately 9 and 13 times higher for the
in the remaining top adhesive bond. A yielding plateau was PUR and balsa wood cored panels, respectively) and failure loads
reached when that bond was completely open, and the tests were (approximately 3 times higher for both) in comparison with the
interrupted. The PUR foam specimens also presented a non-linear single steel angle connections (B1). A gradual loss of stiffness
response characterised by a progressive and substantial loss of was observed throughout the tests, caused by the progressive dam-
stiffness (yet without load reduction), which was attributed to age accumulation within the various connection components. The
the progressive crack opening at the bond between the top face bolt holes in the GFRP faces of the sandwich panels presented bear-
sheet and the top steel angle, with no visual evidence of foam ing deformations (Fig. 6c); the bolts themselves presented very
crushing being observable throughout the test or at failure. significant plastic shear deformations (Fig. 6d), more pronounced

Fig. 4. Load vs. displacement at D1 for all tested connection systems: (a) specimens with PUR foam core and (b) specimens with balsa wood core.
M. Garrido et al. / Engineering Structures 106 (2016) 209–221 213

Table 3
Summary of experimental results.

Core material Specimen Failure load (kN) Initial stiffness, D1 (kN/mm) Initial rotational stiffness (kN/°)
Values Average ± St. Dev. Values Average ± St. Dev. Values Average ± St. Dev.
PUR foam A1-PUR-1 1.56 0.083 1.08
A2-PUR-1 9.95 11.17 ± 1.22a 0.89 0.73 ± 0.16a 11.44 8.77 ± 2.67a
A2-PUR-2 12.39 0.57 6.10
B1-PUR-1 4.81 0.051 0.61
B2-PUR-1 11.29 13.24 ± 1.74 0.42 0.45 ± 0.02 4.63 4.92 ± 0.29
B2-PUR-2 13.80 0.46 5.21
B2-PUR-3 14.63 0.46 4.93
AB2-PUR-1 18.84 0.53 6.31
Balsa wood A1-BAL-1 3.22 3.54 ± 0.32a 0.14 0.15 ± 0.01a 1.76 1.84 ± 0.08a
A1-BAL-2 3.85 0.16 1.92
A2-BAL-1 21.84 20.24 ± 1.59a 2.29 2.26 ± 0.03a 29.23 29.75 ± 0.52a
A2-BAL-2 18.65 2.23 30.27
B1-BAL-1 7.20 0.071 0.85
B2-BAL-1 16.48 18.38 ± 2.11 1.21 1.11 ± 0.10 14.52 13.00 ± 1.59
B2-BAL-2 18.01 1.00 11.35
B2-BAL-3 20.66 1.10 13.13

Italic values correspond to average and standard deviation values relative to the cells immediately to the right (grouped by categories, e.g., A1, A2, B1, B2, ...).
a
Maximum difference between individual data points and the average value.

Fig. 5. Failure modes of the A1 and A2 connection systems: (a) A1 specimen with PUR foam core, (b) A1 specimen with balsa wood core, and (c) A2 specimen with balsa wood
core.

than those observed in the bolts of the B1 specimens, indicating PUR foam core, the A1 connections were 63% stiffer than the B1
the higher capacity of the B2 connections to further explore the connections, and the A2 connections were 62% stiffer than the B2
ductility of the steel bolts. The balsa wood specimens also exhib- configuration. Among the balsa wood cored specimens, the A1
ited crack opening within the core, particularly in the direction connections had a stiffness 111% higher than the B1 configuration,
parallel to the bolts (Fig. 6e), causing the sudden load reductions whereas the A2 connections were 104% stiffer than the B2 ones.
observable in the load–displacement curves.
It is worth highlighting the ability of bolted connections to
deform plastically (either in B1 or B2 configurations) and thus 3.2.4. Adhesively bonded and bolted connection (AB2)
sustain high deformations without brittle failure. This increased In order to evaluate the potential advantages of combining the
ductility of bolted connections is a structural advantage over higher connection stiffness of adhesively bonded connections with
adhesively bonded connections, being an important feature for the plastic deformation capability and higher ductility of bolted
civil engineering design and application. However, the bolted connections, an additional test was performed in a connection
connections presented significantly lower stiffness compared to (PUR foam specimen) combining adhesive bonding and bolting
the adhesively bonded counterparts. For the specimens with a with top and bottom steel angles (AB2 configuration).
214 M. Garrido et al. / Engineering Structures 106 (2016) 209–221

Fig. 6. Failure modes for the B1 and B2 connection systems: (a) B1 specimen with balsa wood core, (b) B1 specimen with PUR foam core, (c) damage in GFRP for a B2
connection, (d) damage in the bolts for a B2 connection, and (e) B2 specimen with balsa wood core.

The load–displacement response obtained in the AB2 connec- also worth highlighting that the high stiffness and strength
tion was linear-elastic up to a load of approximately 6 kN, with together with the limited plastic deformability of the connection
the elastic stiffness being comparable to that of the A2-PUR system shifted the failure mode from the connection to the panel
connections, showing that the adhesive bonding governed the ini- (shear failure of the PUR foam core).
tial response. Above that load, cracks were initiated at the adhesive
bond between the top steel angle and the top face sheet of the 4. Numerical simulation
panel (Fig. 7a), similar to what had been observed in the A2-PUR
series. The crack opening progressed gradually (corresponding to 4.1. Objectives
a short yielding plateau in the load–deflection curve), after which
the bolts were presumably activated and conferred additional stiff- The connection systems proposed in this study were numeri-
ness to the connection. From this point onward, the load–displace- cally modelled using a nonlinear FE modelling approach in order:
ment curve presented an intermediate slope (stiffness) between (i) to simulate the behaviour observed in the experiments, with
that of the A2-PUR and B2-PUR configurations. The bolt activation emphasis on the linear-elastic response, i.e. the stiffness of the con-
seems to have limited the further propagation of the crack in the nection systems proposed; (ii) to simulate the stress distributions
adhesive, which in turn limited the shear stress within the bolts within the connection and panel components, comparing them
themselves, and consequently the extent of their plastic deforma- with the experimentally observed failure modes; (iii) to compare
tion. A maximum load of 18.8 kN was attained, when the PUR foam the deflections of the connected sandwich panels to the deflections
core failed in shear in a section located between the load applica- they would exhibit if they were to be supported in a perfect
tion point and the support (Fig. 7b). cantilever (clamped) configuration (no rotation at the support),
The combined adhesively bonded and bolted system provided so as to obtain estimates for the characteristic moment–rotation
the highest failure load among all connection configurations tested relationships of each connection system. The modelling approach
using PUR foam cored panels. Additionally, it presented high initial is presented in Section 4.2 and the results obtained are presented
stiffness (comparable with the A2 systems) and some plastic defor- in Section 4.3 (first two objectives) and Section 5 (third objective)
mation capability (albeit lower than in the B2 configuration). It is of this paper.
M. Garrido et al. / Engineering Structures 106 (2016) 209–221 215

Fig. 7. Failure in the AB2 connection system: (a) crack opening in the top adhesive layer and (b) shear failure of the PUR foam.

components of the panels and connections. Sensitivity checks were


Table 4
Orthotropic material properties of GFRP and balsa wood used for FEM. performed regarding the influence of the mesh density/refinement
on the results obtained with the FE models, leading to the selection
Material E1 (MPa) E2 (MPa) E3 (MPa) m12 G12 (MPa)
of the adopted meshes. Global element sizes (element average
GFRP 29,400 12,780 7,500 0.31 4,730 maximum edge length) of 20 mm were adopted for the faces,
Balsa wood 127 418 127 0.30 49
30 mm for the core, 10 mm for the steel angles and adhesives,
and 15 mm for the bolts.
Boundary conditions were defined by fixing all degrees of free-
4.2. Modelling approach
dom for the nodes of the steel angles at the areas where these were
bolted to the steel reaction frame used in the experiments. Load
Three-dimensional (3D) FE models were developed using the
was defined by applying a pressure over an area of the sandwich
commercial package ABAQUS/CAE (Dassault Systèmes) to simulate
panels’ top face sheet with the same dimensions and at the same
the A1, A2, B1 and B2 connection systems,2 also considering the two
location as the loading plate used in the experimental tests. Static
different core materials used in this study.
analyses were carried out using the above-mentioned FE models.
The GFRP and balsa wood parts were modelled using linear-
elastic orthotropic material properties (Table 4). The GFRP proper-
ties were obtained from Classical Laminate Theory, confirmed from 4.3. Results
results of material characterisation tests (whenever possible),
whereas the balsa wood properties were determined from testing. The load–displacement (computed at D1) curves obtained from
The PUR foam was modelled with isotropic crushable foam beha- the FE models are presented and compared with the experimental
viour, using the formulation proposed by Deshpande and Fleck results in Fig. 8. Table 5 provides a comparison between the exper-
[19], considering a shear modulus of 8.7 MPa, a Young’s modulus imental and numerical stiffness values (for the displacement at D1
of 26 MPa (Poisson’s ratio of 0.49), and an elastic–plastic constitu- and the rotation at I1). It can be seen that a reasonably good agree-
tive relation with a compressive yield stress of 0.64 MPa (obtained ment was obtained between the FE models and the experimental
from previous material characterisation tests carried out according data. In addition, to accurately predicting the initial stiffness of
to ASTM C365/C365M standard [20]). For the adhesive layer, a the systems, the models of the B2 bolted connections were able
linear-elastic isotropic material model was adopted, using the elas- to reproduce fairly closely the nonlinearity observed in the exper-
ticity modulus of 6.5 GPa indicated in the manufacturer data sheet imental load–displacement curves, corresponding to the bolts’
[21] (a Poisson’s ratio of 0.3 was assumed). The steel elements yielding and the PUR foam’s non-linear behaviour. The deviations
(bolts and angles) were modelled as linear elastic–plastic, with observed are most likely due to the occurrence of (i) slipping
an elasticity modulus of 210 GPa and yield stress of 275 MPa. between the sandwich panel and the top steel plate during the
The contacts between all adherent surfaces (face-core inter- tests (which occurs up to the point where the gaps between the
faces, adhesive bonds) were modelled using the ‘‘cohesive bolt holes and the bolts are closed), and/or (ii) bearing of the top
behaviour” option for the interaction properties in ABAQUS/CAE, GFRP face at the bolt holes, which was not taken into account in
adopting the package’s default contact enforcement method the modelling (the FE models did not incorporate any GFRP mate-
regarding traction-separation behaviour. This method does not rial damage model).
allow significant slipping or separation to occur between the The stiffness values obtained from the FE models, either regard-
adherent surfaces. The non-adherent contacts (between the bolts ing the displacement at D1 or the rotation at I1, presented an over-
and the bolt holes or between the panel’s faces and the steel angles all average relative difference of 15% to the experimental data.
in the bolted connections) were modelled using ‘‘tangential The highest relative differences in this regard were found for the
behaviour” with a friction coefficient of 0.1 and the ‘‘normal A2 connection system, with the FE models being typically 30%
behaviour” with ‘‘hard contact” for the pressure-overclosure beha- less stiff than what was experimentally observed, and also for
viour of the interaction. the B1 connection system applied to balsa wood cored panels, for
Quadratic ten-node tetrahedral solid elements (C3D10 elements which the FE model was 25–30% stiffer than in the experiments.
of the ABAQUS/CAE library) were used to model the different However, these differences are within the scatter that was
obtained in the experimental data, and also within the typical scat-
ter found for the mechanical properties of balsa wood, which has
2
It is worth mentioning that the AB2 connection system (adhesively bonded and been previously reported to be as high as 28% [22].
bolted) was not explicitly modelled, since the FE analyses were mainly focused on the
linear-elastic branch of the connections’ response. In fact, for this range of the
Fig. 9 shows the distribution of though-thickness (vertical)
response, the behaviour of the AB2 connection would not differ significantly from that stresses in the adhesive (at mid-width and mid-thickness of the
of the A2 configuration. adhesive layer) along the length of the connection for the A1 and
216 M. Garrido et al. / Engineering Structures 106 (2016) 209–221

Fig. 8. Load vs. displacement (D1) curves obtained from FEM: (a) specimens with PUR foam core and single (bottom) steel angle, (b) balsa wood core and single steel angle, (c)
PUR foam core and two (top and bottom) steel angles, and (d) balsa wood core and two steel angles.

Table 5
Comparison between FEM and experimental stiffness values at D1 and I1.

Core material Connection system Stiffness, D1 displacement (kN/mm) Rotational stiffness, I1 rotation (kN/°)
Experimental FEM Difference (%) Experimental FEM Difference (%)
PUR foam A1 0.083 0.071 14.5 1.08 0.92 14.9
A2 0.73 0.49 33.1 8.77 5.98 31.8
B1 0.051 0.046 11.1 0.61 0.56 7.0
B2 0.45 0.41 7.4 4.92 5.10 3.5
Balsa wood A1 0.15 0.16 7.4 1.84 1.98 7.8
A2 2.26 1.65 27.1 29.75 19.63 34.0
B1 0.071 0.093 30.0 0.85 1.06 24.9
B2 1.11 1.12 1.2 13.00 13.73 5.6

6 4
(a) (b)
Through-thickness

A1 connection A2 connection
Through-thickness

Load: 0.35 kN Load: 2.00 kN


3 2
stress [MPa]

stress [MPa]

0 0

-3 PUR foam core -2 PUR foam core


Balsa wood core Balsa wood core
-6 -4
0 30 60 90 120 150 0 30 60 90 120 150
Position [mm] Position [mm]

Fig. 9. Through-thickness stresses in the epoxy adhesive: (a) A1 connections and (b) top layer in the A2 connections.

A2 connections obtained from the FE models. For the A1 connec- core materials, which also agree well with the failure modes exper-
tions, Fig. 9a plots the through-thickness stresses in the adhesive imentally observed in the A1-BAL-2 specimen (at the face-core
layer bonding the bottom face sheet of the two types of sandwich interface) and in the A2-PUR-1 specimen (within the foam core).
panels (balsa wood and PUR foam cores) to the supporting steel From the FE predictions, these peeling stresses seem to be fairly
angle, for a load of 0.35 kN (for which the behaviour of both sys- independent of the type of core material, unlike the compressive
tems is well within the linear-elastic range). It can be seen that stresses at the opposite end of the adhesive layer, which are signif-
for both types of cores peeling stresses develop in the first icantly higher for the panel with the softer PUR foam core.3 This
15 mm of the connection length. This is in good agreement with
the failure initiation zone observed in the experiments for the 3
Locally, the absolute values of such stresses are inherently mesh dependent
A1-BAL-1 specimen, which failed in the adhesive bond at that same (namely the peaks values at the singularity regions) and consequently the analysis
area. These localised peeling stresses occur together with through- was focused on their qualitative development and on the relative differences
thickness tensile stress concentrations in the bottom GFRP face and encountered for the two panel types, for which similar meshes were adopted.
M. Garrido et al. / Engineering Structures 106 (2016) 209–221 217

Fig. 10. Plastic strains in the bolts: (a) B1-PUR and (b) B2-PUR.

difference most likely stems from the higher curvature of the GFRP 5. Moment–rotation relationships
face sheets associated with the overall higher flexibility of those
panels when compared to those made of balsa wood core. The 5.1. Basis of calculation
curvature of the bottom GFRP face in contact with the adhesive also
causes tensile stress peaks in the bond at a position circa 140 mm As mentioned, the proposed connection systems provide some
(Fig. 9a). This is due to the deformation compatibility between those degree of rotational restriction to the sandwich panels at the sup-
two components, since the GFRP face would exhibit a certain ports, thus contributing to reduce the overall floor flexibility. This
curvature if unrestricted, whereas the adhesive layer remains restriction may be considered equivalent to the effect of a rota-
approximately flat. tional spring, with a characteristic stiffness kh, as illustrated in
For the A2 connections, Fig. 9b shows the through-thickness Fig. 11b for the case of the experimental setup adopted in this
stresses in the adhesive (again, at mid-width and mid-thickness study (i.e., a cantilever beam configuration).
of the adhesive layer) between the top face sheet of the panels A perfectly clamped sandwich panel (Fig. 11a) would exhibit a
and the top steel angle for a load of 2.00 kN, for which the beha- deflection dc as a result of the applied point load (P). This deflection
viour of both connection systems was still linear-elastic. Once would correspond solely to the sandwich panel’s deformations due
more, the significant peeling stresses at the front of the connection to bending (dB) and shear (dS). However, as shown in Fig. 11b, due
are in good agreement with the experimentally observed failure to the existence of bending moment at the support, the connection
modes of the A2 connections, which initiated exactly at that loca- system’s flexibility causes an additional rigid body movement of
tion for all specimens (cf. Fig. 5c). In this configuration the peeling the panel (dh) characterised by the rotation h at the support. There-
stresses show a high dependence from the core material, being sig- fore, if the panel is not fully clamped, the total deflection (d) is the
nificantly higher for specimens made of PUR foam core, due to the sum of the deflection due to the rigid body rotation of the panel at
higher flexibility of these panels, as aforementioned. The compres- the support and the cantilever deflection due to the panel’s bend-
sive stress peaks in the adhesive at a position circa 140 mm are ing and shear deformations, i.e., d = dh + dc. Knowing both the total
once more due to the deformation compatibility between the top deflection of the system and the cantilever deflection, it is possible
GFRP face sheet and the adhesive layer. to determine dh and, consequently, the corresponding rotation
Fig. 10a and b plot the deformed shape and the plastic strains in angle h, given by h = arcsin(dh/L), L being the cantilever span.
the bolts used in the B1 and B2 connections used in PUR foam Finally, considering the corresponding bending moment at the
cored specimens, corresponding to the maximum loads attained support, M = P  L, a moment–rotation (M–h) relationship and elas-
with each model: (i) 4.4 kN in the A1-PUR model, and (ii) tic rotational stiffness (kh) may be obtained for the connection
14.5 kN in the A2-PUR model. It can be seen that in both cases system.
the predicted deformed shapes are in good agreement with the
experimentally observed deformations (cf. Fig. 6a, c and e), with
very significant plastic deformations in the bottom and upper parts 5.2. Cantilever models
of the bolts, respectively for the B1 and B2 connections. It can also
be seen that the magnitude of such deformations is significantly To obtain the above-mentioned relationships, the FE models
higher in the latter connections, also in accordance with the developed (and validated) in this study could be used to calculate
experiments. the rotations at the support sections and hence to estimate the dh
values. However, such estimates might be influenced by the local

Fig. 11. Deflections for a (a) cantilever beam and (b) an elastically restrained beam (with connection system).
218 M. Garrido et al. / Engineering Structures 106 (2016) 209–221

Table 6
Rotational stiffness values of the proposed connection systems.

Core Connection Rotational


material system stiffness,
kh ((kN m/rad)/m)
PUR foam A1 105
A2 8721
B1 60
B2 2447
Balsa wood A1 231
A2 10,856
B1 124
B2 3199

Fig. 12. Comparison between deformed shapes for the B2-PUR model (top) and the basically consisted of eliminating the models’ sections
corresponding PUR-Clamp cantilever model (bottom). corresponding to the connection elements (including the panel
length embedded into the connection), and fixing all degrees of
freedom of the sandwich panels’ nodes immediately after the
connection (Fig. 12).
18
5.3. Characteristic curves

12 Fig. 13 compares, as an example, the load–deflection curve for


Load [kN]

the B2-BAL connection model (balsa wood cored panel) to the


curve obtained from the corresponding cantilever model
6 (BAL-Clamp). As mentioned, the differences in the displacements
B2-BAL (FEM) calculated from the two models are due to the rotation at the sup-
Cantilever-BAL port in the former model resulting from the connection system’s
0 deformability (cf., Fig. 11b). Calculating that displacement differ-
0 15 30 45 ence (dh) and the corresponding rotation (h) as a function of the
Displacement [mm] bending moment at the support, the M–h curve may be obtained.
Fig. 14 presents these curves for each of the proposed connec-
Fig. 13. Comparison between P–d curves of the B2-BAL and BAL-Clamp models.
tion configurations and panel types. It should be noted that for
design purposes the nonlinear parts of the M–h curves should not
be considered, as progressive failure mechanisms existing within
compressibility of the sandwich panel’s core; in addition, the def- the sandwich panels and connection systems were not thoroughly
inition of a specific node to calculate the rotation might be rela- modelled in this study (e.g., crushing/bearing of the GFRP faces due
tively ambiguous. To overcome such difficulties, an alternative to the bolts, cracking of the balsa wood cores and adhesive layers).
method was adopted that consists of developing fully clamped The fact that deformations at serviceability limit states should
cantilever models for each type of sandwich panel (‘‘PUR-Clamp” remain within the linear-elastic range of the structural response
and ‘‘BAL-Clamp” models) and using those models to estimate further contributes to this point. The rotational stiffness (kh) values
the panels’ respective dc values. The numerical load–deflection (per unit width of the connection) corresponding to the linear-
curves obtained in Fig. 8 were compared with the baseline load– elastic branch of the M–h curves are presented in Table 6.
deflection curves given by the PUR-Clamp and BAL-Clamp models. The low rotational stiffness of the single steel angle connections
As discussed in Section 5.1, the difference between the two curves (A1 and B1) indicates that these are inefficient when bending
corresponds to the deflection caused by the rigid body rotation (dh) moment mobilisation at the connection is a design requirement.
resulting from the connection’s flexibility. In such cases, the addition of the top steel angle substantially
The reference cantilever models were developed by adapting improves the connection’s ability to mobilise negative bending
the previously described models of the connection systems tested, moments, thus effectively reducing the maximum deflection of
maintaining their basic features (identical meshes and elements, the sandwich panel floors (compared to a simply supported floor
material properties, loading conditions, etc.). This adaptation configuration).

12
Bend. moment [kNm/m]

Bend. moment [kNm/m]

(a) (b) 50
9 40

6 30
A1-PUR A2-PUR
20 A2-BAL
A1-BAL
3 B1-PUR B2-PUR
10
B1-BAL B2-BAL
0 0
0.0 0.1 0.2 0.3 0.00 0.02 0.04 0.06
Rotation [rad] Rotation [rad]

Fig. 14. M–h curves for connection systems (per unit width) with: (a) single steel angle (bottom only) and (b) two steel angles (top and bottom).
M. Garrido et al. / Engineering Structures 106 (2016) 209–221 219

It is also interesting to note the significant influence of the sand-


wich panel’s core material on the rotational stiffness of the connec-
tion. The stiffer core material (balsa wood) allowed for stiffer
connections in all configurations. The rotational stiffness values
of the single angle connections, A1 and B1, for balsa wood cored
panels were respectively 125% and 107% higher than those for
the PUR foam cored panels, whereas for the double steel angle con-
figurations these differences were less significant, but still impor-
tant, with the A2 and B2 configurations being respectively 24%
and 31% stiffer for the balsa wood panels than for the PUR foam
counterparts. Fig. 15. Deformed shape for the 4 m span sandwich panel, either supported using
the B2 connection (top) or simply supported (bottom).

6. Connection contribution for SLS verifications

The design of FRP sandwich panels for building floors is often


limited by serviceability limit states (SLS) requirements (maxi- Table 7
Short-term deflections from analytical and FEM models for the 4 m span sandwich
mum allowable deflection), which are defined taking into account
panel.
functionality aspects, users’ comfort and protection of non-
structural elements. The bending moment mobilisation capability Model Bending Shear Total
deflection, deflection, deflection,
offered by the connection systems proposed, particularly the A2
dB (mm) dS (mm) d (mm)
and B2 (and also the AB2) configurations, may provide a valuable
Simply supported Analytical 4.17 3.58 7.75
contribution for controlling and limiting such deflections.
FEM – – 7.64
In a preliminary design setting, these contributions may be
B2 connection Analytical 1.65 3.58 5.23
easily estimated by considering appropriate (and simple) struc-
FEM – – 4.87
tural models for the deflection calculations [23]. For instance, the
mid-span deflection (d) of a simply supported sandwich panel of
span L under a uniformly distributed load (UDL)4 p may be calcu-
lated according to Timoshenko beam theory (Eq. (1)), in which the
bending deflection (dB) is calculated using the face sheets’ Young’s results obtained for the short-term deflections are summarised in
modulus (E) and moment of inertia (I), and the shear deflection Table 7.
(dS) is estimated using the shear modulus (G) and effective shear area A very good agreement was found between the analytical
(Av) of the core: predictions and the numerical results for the short-term mid-
span displacement of the sandwich panel connected with system
5 pL4 1 pL2 B2 (relative difference of 7.5%). Additionally, the results obtained
d ¼ dB þ dS ¼ þ ð1Þ
384 EI 8 GAv also illustrate the substantial deflection reductions that may be
obtained from the sandwich floor panel-to-wall connections pro-
Considering the sandwich panel as being simply supported
posed in this paper (compared to the deflections corresponding
involves neglecting the rotational stiffness of the panel’s supports
to simply supported panels). In the present example, a deflection
(i.e., the panel-to-wall connections). To compute the panel’s mid-
reduction of approximately 35% was achieved. It is worth mention-
span deflection under a semi-rigid support situation, the following
ing that this reduction is mostly achieved by limiting the bending
closed-form equation proposed by Turvey [24] for shear deform-
flexibility of the panel, as illustrated by the equal analytical shear
able beams may be used,
deflection estimates for the simply supported and semi-rigidly
  supported configurations given in Table 7. This is in agreement
1 pL4 1 þ 48a þ 10b þ 96ab
d¼ ð2Þ with Timoshenko beam theory.
384 EI 1 þ 2b
Fig. 16 presents those reductions for spans between 2 m and
where a ¼ EI=GAv L2 is the dimensionless shear flexibility of the 5 m, estimated using Eqs. (1) and (2) for simply supported sand-
sandwich panel and b ¼ EI=kh L is the dimensionless rotational flex- wich panels and for semi-rigidly supported panels, respectively.
ibility of the connections. To allow for result comparability, the only parameter that was
In order to validate this formulation for the current application, changed was the span length, i.e., the panel cross-sectional
an FE model of a 4 m span sandwich panel was developed, based geometry was kept constant for all spans. It should be noted
on the previously presented models and considering PUR foam as that in real applications the panels’ cross-sectional dimensions
the core material and B2 floor-to-wall connection system could be adapted and optimised for each span. However, this
(Fig. 15). Symmetry simplifications were adopted in order to optimisation exercise has multiple solutions depending on
reduce the model size. Maximum (mid-span) deflections were many factors (e.g., material properties, relative cost of materials,
obtained from the numerical model and compared with the analyt- building physics requirements), being beyond the scope of this
ical predictions for: (i) a simply supported configuration under a study.
1.98 kN/m2 UDL (corresponding to a panel self-weight of For all situations presented in Fig. 16, it can be seen that the
0.38 kN/m2, other permanent loads of 1.0 kN/m2, and an estimated deflection reductions increase with the span length. This
imposed/live load of 2.0 kN/m2 with a quasi-permanent combina- is due to the increasing relative importance of bending deforma-
tion factor of 0.3, according to Eurocode 1), and (ii) considering tions to the overall panel deflection compared to shear deforma-
the B2 connection system and a sandwich panel with the same tions, which become less significant with increasing span for all
cross-sectional geometry used throughout this study. The connection configurations, as illustrated in Fig. 17. This also
explains the higher deflection reductions obtained with the balsa
4
Concentrated or point loads were not explicitly considered in the current analysis
wood cored panels when compared to those made of PUR foam
(although both panel configurations used in the experiments are able to withstand core, since balsa wood presents a substantially higher shear mod-
the design point load defined in Eurocode 1). ulus and thus lower shear flexibility (cf. Table 1).
220 M. Garrido et al. / Engineering Structures 106 (2016) 209–221

Deflection reduction [%]


50 75

Deflection reduction [%]


(a) PUR foam core (b) Balsa wood core
40 60
30 45
20 30
10 15
0 0
2 3 4 5 2 3 4 5
Span [m] Span [m]
A1 A2 B1 B2 A1 A2 B1 B2

Fig. 16. Deflection reductions for different spans using the proposed connection systems for panels with: (a) PUR foam core and (b) balsa wood core.

Fig. 17. Shear contribution to total mid-span deflection (d) for different spans using the proposed connection systems for panels with: (a) PUR foam core and (b) balsa wood
core.

(a) 16 PUR foam core (b) 12 Balsa wood core


Deflection [mm]
Deflection [mm]

12 9
8 6
4 3
0 0
2 3 4 5 2 3 4 5
Span [m] Span [m]
L/500 L/1000 Simply Sup. L/500 L/1000 Simply Sup.
A1 A2 B1 B2 A1 A2 B1 B2

Fig. 18. Deflections for different spans using the proposed connection systems for panels with: (a) PUR foam core and (b) balsa wood core.

Those deflection reductions can become very valuable for SLS most old building rehabilitation scenarios the vertical supporting
verifications, as illustrated in Fig. 18. This figure shows the total elements are load-bearing masonry walls. The following main con-
deflections obtained in each configuration and compares them clusions may be drawn from this study:
with a typical serviceability limit states deflection limit of
span/500. A curve representing half of that limit (span/1000) is also 1. The experimental and numerical investigations validated the
plotted. As may be observed in Fig. 18, the adoption of either the connection systems proposed, showing their potential for appli-
A2 or B2 connection systems allows for in service deflections cation in building rehabilitation.
that are lower than span/500 for all considered spans (in some 2. The use of a top steel angle considerably increased the rota-
cases making the difference between fulfilling or not such tional stiffness of floor panel-to-wall connections. This con-
deflection limit), and even lower than span/1000 in most cases. It structive detail may be useful when a certain clamping level
is worth mentioning the importance of guaranteeing elastic is required at the supports.
deflections at serviceability well below the SLS limit values so 3. Adhesively bonded connections typically presented lower
that when accounting for long-term deflections, namely those deformation capacity and higher stiffness than bolted connec-
due to creep, the serviceability deflection limits may still be tions, exhibiting brittle failure modes. Bolted connections pre-
fulfilled. sented plastic deformations and damage distribution through
the different connection and panel components.
7. Conclusions 4. The mechanical properties of the core material influenced the
overall connection behaviour, namely the rotational stiffness
This paper presented a study about the development and of the connections, and in some circumstances their failure
performance under vertical loading of floor-to-wall connection mode (particularly for the A2 configuration); in this regard,
systems for FRP sandwich panels to be used in building rehabilita- the best overall performance was obtained with the stiffer core
tion. The proposed connections were developed considering that in material (balsa wood).
M. Garrido et al. / Engineering Structures 106 (2016) 209–221 221

5. A method was proposed to calculate the moment–rotation References


relationship of the connection systems, which is based on
decomposing the total deflections into those corresponding [1] Costa A, Guedes JM, Varum H, editors. Structural rehabilitation of old
buildings. Berlin Heidelberg: Springer; 2013.
to a cantilever (perfectly clamped) and those due to the rigid [2] Branco M, Guerreiro LM. Seismic rehabilitation of historical masonry buildings.
body rotation about the supports; the method was duly Eng Struct 2001;33(5):1626–34.
validated by the FE models, allowing to derive characteristic [3] Correia JR, Branco FA, Ferreira J. GFRP–concrete hybrid cross-sections for floors
of buildings. Eng Struct 2009;31(6):1331–43.
M–h curves and rotational stiffness values (kh) for each [4] Harris D, Cousins T, Murray T, Sotelino E. Field investigation of a sandwich
connection system. plate system bridge deck. J Perform Constr Facil 2008;22(5):305–15.
6. An analytical method proposed by Turvey [24], based on [5] Freitas ST, Kolstein H, Bijlaard F. Sandwich system for the renovation of
orthotropic steel bridge decks. J Sandwich Struct Mater 2010;13:279–301.
Timoshenko beam theory and that takes into account the
[6] Correia JR, Garrido M, Gonilha JA, Branco FA, Reis LG. GFRP sandwich panels
rotation stiffness of the supports, was used to calculate the with PU foam and PP honeycomb cores for civil engineering structural
mid-span deflections of sandwich panels supported using applications: effects of introducing strengthening ribs. Int J Struct Integr
the connection systems developed in this study. The results 2012;3(2):127–47.
[7] Awad ZK, Aravinthan T, Zhuge Y. Investigation of the free vibration behaviour
obtained highlighted the potential influence of the support of an innovative GFRP sandwich floor panel. Constr Build Mater
conditions on the total floor deformability. 2012;37:209–19.
[8] Sousa JM, Correia JR, Cabral-Fonseca S, Diogo AC. Effects of thermal cycles on
the mechanical response of pultruded GFRP profiles used in civil engineering
The last conclusion assumes special importance given the fact applications. Compos Struct 2014;116:720–31.
that the design of sandwich panels for building floors is generally [9] Keller T, Haas C, Vallée T. Structural concept, design and experimental
governed due to deformability requirements. The correct design verification of a glass-fiber reinforced polymer sandwich roof structure. J
Compos Constr 2008;12(4):454–68.
and in-situ execution of the connections between the sandwich [10] Durand S, Lubineau G, Chapelle H. Design analysis of GRP panels for the roof of
panel floors and their supports, and the consideration of their the Haramain High Speed Railway Jeddah Station. In: Proceedings of the 10th
actual contribution to the structural behaviour can be a valuable international conference on sandwich structures (ICSS-10), Nantes, France,
August 27–29 2012.
resource for deflection control in SLS verifications. This could [11] Keller T, Rothe J, de Castro J, Osei-Antwi M. GFRP-balsa sandwich bridge deck:
potentially allow for more cost-efficient sandwich panel designs, concept, design, and experimental validation. J Compos Constr 2014;18
with thinner faces and cores, thus increasing their application (2):04013043.
[12] Sharaf T, Fam A. Analysis of large scale cladding sandwich panels composed of
potential for building rehabilitation and civil engineering projects
GFRP skins and ribs and polyurethane foam core. Thin-Wall Struct
in general. 2013;71:91–101.
It is worth mentioning that the current study did not consider [13] Mousa MA, Uddin N. Flexural behavior of full-scale composite structural
the effects of horizontal/in-plane loading of the sandwich floors. insulated floor panels. Adv Compos Mater 2011;20(6):547–67.
[14] Awad ZK, Aravinthan T, Zhuge Y, Manalo A. Geometry and restraint effects on
This aspect encompasses the study of their behaviour under hori- the bending behaviour of the glass fibre reinforced polymer sandwich slabs
zontal (seismic) loads, and the effects of the membrane stresses under point load. Mater Des 2013;45:125–34.
originating from the in-plane restriction imposed by the connec- [15] Meireles H, Bento R. Rehabilitation and strengthening of old masonry
buildings. Report DTC no. 02/2013, ICIST, IST, University of Lisbon, Lisbon;
tions along the full perimeter of the floors. Both these aspects are 2013.
potentially very relevant for the global performance of the sand- [16] Appleton J. Reabilitação de Edifícios Antigos: Patologias e Tecnologias de
wich floors and their connections, and should be addressed in Intervenção, Edições Orion, 1st ed., September 2003 [in Portuguese].
[17] Simões A, Bento R, Cattari S, Lagomarsino S. Seismic performance-based
future research. assessment of ‘‘Gaioleiro” buildings. Eng Struct 2014;80:486–500.
[18] Garrido M, Correia JR, Keller T, Branco FA. Adhesively bonded connections
between composite sandwich floor panels for building rehabilitation. Compos
Acknowledgements Struct 2015. http://dx.doi.org/10.1016/j.compstruct.2015.08.080.
[19] Deshpande VS, Fleck NA. Isotropic constitutive models for metallic foams. J
Mech Phys Solids 2000;48:1253–83.
The authors would like to acknowledge Fundação para a Ciência [20] ASTM C365/C365M. Standard Test Method for Flatwise Compressive
e a Tecnologia (FCT, research grant no. PTDC/ECM/113041/2009) Properties of Sandwich Cores. ASTM International, West Conshohocken,
and ICIST for funding the research. The first author would also like Pennsylvania, United States; 2011.
[21] SIKA. Sikadur-31 EF – Product Data Sheet, Ed. 23/01/2014; 2014.
to acknowledge FCT for the financial support through PhD scholar- [22] Osei-Antwi M, de Castro J, Vassilopoulos AP. Shear mechanical
ship no. SFRH/BD/78584/2011. The support of the company Sika characterization of balsa wood as core material of composite sandwich
Portugal SA is also gratefully acknowledged. Finally, the authors panels. Constr Build Mater 2012;41:231–8.
[23] CEN WG4. Fibre Reinforced Polymer Structures – Scientific and Technical
wish to express their gratitude to the institute PIEP – Innovation Report, submitted for approval to CEN/TC 250; 2014.
in Polymer Engineering, and particularly Luís Oliveira, for the [24] Turvey GJ. Analysis of pultruded glass reinforced plastic beams with semi-rigid
manufacturing of specimens. end connections. Compos Struct 1997;38(1–4):3–16.

You might also like