You are on page 1of 9

Construction and Building Materials 112 (2016) 84–92

Contents lists available at ScienceDirect

Construction and Building Materials


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

A new composite connector for timber-concrete composite structures


Samuel C. Auclair a,c,d, Luca Sorelli a,c,d,⇑, Alexander Salenikovich b,d
a
Department of Civil and Water Engineering, Université Laval, Canada
b
Department of Wood and Forest Sciences, Université Laval, Canada
c
Research Centre on Concrete Infrastructures – CRIB, Canada
d
NSERC Industrial Research Chair on Ecoresponsible Wood Construction – CIRCERB, Canada

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 A new elongated composite


connector has been developed for
timber concrete composite structures.
 The stiffness and strength of the
connection depend on the external
and internal diameters of the
connector, respectively.
 The non-linear Winkler beam model
well reproduces the shear connection.
 The composite connector diameters
control the deflection and the
structural ductility.

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

Article history: Timber-concrete composite (TCC) structures are emerging in several industrial applications as an efficient
Received 1 October 2015 method for optimizing the structural performance and the cost of construction. Their effectiveness
Received in revised form 4 February 2016 depends strongly on the kind of connection employed. In order to guarantee sufficient ductility to the
Accepted 9 February 2016
structure without sacrificing its stiffness and strength, the connections have to be rigid, strong and
Available online 1 March 2016
deform plastically before the brittle collapse of the timber or concrete members. This work presents a
new composite connector, which can be used to enhance the ductility of a structure without significant
Keywords:
loss of stiffness at serviceability limit states. The studied composite connector consists of a composite
Ultra-high performance fibre-reinforced
concrete
cylinder made of ultra-high performance fibre-reinforced concrete (UHPFRC) shell with a steel cylindrical
Shear connector core. The UHPFRC enhances micro-cracking resistance and energy dissipation under large deformations.
Shear connection test Performance characteristics of the connectors of various sizes have been evaluated using shear tests of
Connection strength connections. The results show that the connection stiffness is principally governed by the diameter of
Connection rigidity the concrete shell, while the connection resistance is principally governed by the diameter of the steel
Structural ductility core. A beam on a Winkler foundation model has been applied to describe the behaviour of the composite
connector in the shear tests. Finally, the composite beam theory has been applied to predict the structural
behaviour of TCC beams with different parameters of the composite connectors. The results show that the
diameters of the concrete shell and of the steel core of the connector can be conveniently varied to opti-
mize the TCC beam performance by significantly enhancing its structural ductility without significant
loss of flexural stiffness and load bearing capacity.
Ó 2016 Elsevier Ltd. All rights reserved.

⇑ Corresponding author at: Department of Civil and Water Engineering, Université Laval, Canada.

http://dx.doi.org/10.1016/j.conbuildmat.2016.02.025
0950-0618/Ó 2016 Elsevier Ltd. All rights reserved.
S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92 85

1. Introduction [18] proposed a simplified estimation of the maximum load for TCC
structures by assuming a rigid-perfectly plastic load-slip relation-
Timber-concrete composite (TCC) structures present an effi- ship (V  s) where the maximum connection shear strength is
cient method for optimizing the structural performance and the reached at the interface. In design practice, when serviceability
construction cost of buildings and bridges, as well as for retrofit limit states, such as deflection and vibration, govern the design
and strengthening of existing floors, through an intelligent use of of the floor, a simplified linear elastic calculation model, such as
the properties of both materials [1–5]. In a TCC floor, the concrete the c-method in the Annex B of Eurocode 5 [22], is suitable to
slab increases the overall stiffness and reduces the floor vibrations, design a TCC structure.
while the timber beam provides the resistance, reduces the weight By optimizing the connection it is possible to enhance ductility
and improves the environmental impact and appearance of the of a TCC structure without compromising its stiffness at the ser-
structure. viceability state. However, in the case of steel dowel connectors,
A TCC structure consists of a concrete slab supported by timber [17] found that the connection stiffness is approximately linearly
panel or beam, which may be attached by means of different types proportional to the dowel diameter, while the maximum shear
of connectors. The timber primarily resists tensile stress and the resistance of the connection is proportional to the square root of
concrete resists compressive stress generated by moments and the dowel diameter. In the context of the present discussion, it is
by the composite action. The shear connection between timber challenging to vary the diameter of a steel dowel to achieve the
and concrete generates the axial force, which greatly contributes suitable connection resistance, which allows the ductile failure of
to the total resistant moment of a TCC structure [6,7]. The beha- the structure, as it also affects the connection stiffness and, hence,
viour of the connection can be determined by means of a shear test, the flexural stiffness of the structure. The underlying idea of this
also called push-out test [8]. Fig. 1 illustrates the behaviour of dif- work is to develop a composite connector with variable properties
ferent connectors in terms of shear load vs. slip curves (V  s) from allowing optimisation of the flexural stiffness, resistance and duc-
data available in literature [9–11]. The connector law (V  s) is tility of TCC structures. The objective of the present work is two-
often highly non-linear within a slip range of 3 mm, especially in fold: (i) to develop a concept of a new composite connector and
the case of discrete connectors, like screws, studs or dowels characterize its performance by experimental shear tests; and (ii)
[12,13]. to assess the gain in the structural performance of a TCC beam
The structural behaviour of TCC floors with different types of by varying the composite connector parameters. The article is
connections has been investigated by several authors [14,10,15]. structured as follows: Section 2 introduces the connector concept
The connection parameters governing the structural behaviour of and theoretical background for analysis of the connection and of
a TCC structure are: (i) stiffness, or slip modulus, (ki ); (ii) resis- TCC beams; Section 3 presents and discusses the experimental
tance, or maximum shear resistance, (V max ); and (iii) ductility (l). tests on the proposed connections; Section 4 analyses the experi-
The connection stiffness affects the degree of composite action mental results using numerical modelling for better understanding
between the members. The resistance and ductility of the connec- the behaviour of the connection; and Section 5 predicts the struc-
tion affect the behaviour of a TCC structure only if the resistance of tural response of TCC beams with the new connectors using the
the connector is achieved before the main member collapse in numerical modelling with the emphasis on the stiffness and
bending or tension. If the connection behaves elastically when ductility of the structure.
the timber reaches the maximum tensile strength in the external
fibre, the composite structure will have a brittle failure, albeit
2. New connector concept and theoretical background
the connection load-slip relationship is ductile [11,16]. On the
other hand, when the connections yield before the collapse of the
2.1. Concept of a new composite connector
timber, the TCC structure will start deforming plastically resulting
into a non-linear, ductile behaviour [17,18].
According to the capacity design approach, in a ductile TCC
To analyse the non-linear behaviour of TCC structures in bend-
structure the connection should undergo non-linear deformation
ing, the composite beam theory has been recently extended to
before the collapse of the main member. In this study, we devel-
account for the non-linear load-slip (V  s) of connections [19–
oped a prototype of a cylindrical connector made of a concrete
21]. In the case where connections fail before the timber collapse,
shell with or without a steel core. The concrete shell diameter
governs the connection stiffness, while the steel core governs
the connection resistance. To allow large energy dissipation
120 and micro-cracking resistance special ultra-high performance
fibre-reinforced concrete (UHPFRC) is employed. Furthermore,
Long steel tube + screw + notch
the connector has an elongated shape to provide flexural
90 behaviour, which is predictable by simple models.
Shear force (V)

Steel mesh connector


2.2. Background on composite beam theory
60
Dowel with flange
For the analysis of the connections and TCC beams in this work
Toothed metal pale we employed two existing methods well established for composite
30 structures, such as follows:
Two folded steel plate

1. For the connection analysis, the Winkler model is used, which


Steel dowel connector
allows predicting the shear behaviour (V  s) of a connection
0
0 2 4 6 8 10 12 14 16 by representing the connector as a beam on an elastic founda-
tion [17,23], as schematically shown in Fig. 2, where: kc and
Slip (s) [mm]
kw are the elastic foundation with moduli of the concrete and
Fig. 1. Load vs. slip curves of various connectors (After [10]). the timber, respectively; Es Is is the flexural stiffness of the steel
86 S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92

d vn s n kc

kc E s Is E s Is
t t t t

kw kw

(a) (b) (c) (d)


Fig. 2. Timber–concrete connection with a steel dowel (after [17]) and its Winkler model in the initial state (a and b) and in the deformed state (c and d).

1500 3. Experimental tests on the new connector

F 3.1. Materials
diametre d 48 mm
f = F/tw [N/mm]

1000 The performance of the prototype connectors was evaluated


F/2 using shear tests on specimens fabricated from the following mate-
rials. The timber members were Canadian-made glulam Nordic-
d 20 tw Lam 24F-ES/NPG in air-dry condition. The mechanical properties
500
d 16 Alps red spruce of timber were taken from the manufacturer’s specification [31],
d 12 and the mean values for 10-min load duration were estimated
kw = 1300 N/mm2 using CSA O86 standard practice [32]. The concrete members were
0 fabricated from standard concrete, and its strength was measured
0 1 2 3 4 by a standard compression test [33]. The connector shells were
Wood-stud slip [mm] made of two different sizes and using two types of concrete: (i)
UHPFRC with 2% steel fibre volume, and (ii) a regular concrete mor-
Fig. 3. Experimental curves for the evaluation of wood foundation stiffness kw
([17]). tar. The properties of the UHPFRC were taken from the literature
[34,35]. The strength of the concrete mortar was measured by
standard compression tests on cylindrical samples [33]. For the
cylindrical core; t is the distance between the concrete slab and connector core two types of steel were used: (i) threaded rods
the timber beam; sn represents the slip between the two mem- (M6 or M12) and (ii) a rebar (10 M). Their mechanical properties
bers; v n is the horizontal displacement of the top end of the were determined in laboratory from standard direct tension tests
connector. The Winkler model is suitable for elongated connec- [36]. One series of tests used an UHPFRC connector without steel
tors, which mainly work in flexure, e.g., steel dowels [12,17]. As core. The average values of the mechanical properties of all mate-
experimentally determined by [17], the wood foundation mod- rials used in the tests are reported in Table 1.
ulus (kw ) is taken as non-linear relationship of the shear force
vs. slip (V  s) (see Fig. 3). The Winkler model has been recently 3.2. Fabrication of the test specimens
extended to account for the non-linear behaviour of the connec-
tor, in terms of moment and curvature (M  v) by means of a The geometry and configurations of the tested connectors are
secant stiffness approach [21,24]. The non-linearity of the shown in Fig. 4 and Table 2, respectively. All connectors were made
moment–curvature relationship, which is due to steel yielding with a square head of the same size of bc = 80 mm and lc = 40 mm
or concrete cracking, can be determined by a classical sectional using individual custom-made wood moulds with a tolerance of
analysis from the one-dimensional material law of the connec- 0.5 mm. The concrete mix was poured into the mould with the
tor materials in terms of stress and strain (r  ) [25]. Full steel core in place. The mould was removed after 24 h and the con-
details of such model, numerical implementation and validation nectors were allowed to cure for seven days at 60°C and 95% rela-
can be found in [21]. tive humidity to accelerate the hydration before installation into
2. For the structural analysis of a TCC beam, the composite beam the timber beam. To assemble the specimen the connector was
theory is used. The governing equation has been developed by inserted into the timber beam by tapping into a hole pre-drilled
[19] and validated on several applications of composite beams with the same diameter and length as the connector. Then, the
[1,26,27] under the following assumptions: (i) the timber beam concrete slab was cast in place by pouring regular concrete mix
and the concrete slab behave according to the classical Euler– into a formwork attached to the timber beam. A plastic film was
Bernoulli beam theory, which neglects the shear deformation laid under the concrete to protect the timber from wetting and
effect on the deflection; (ii) each member has the same deflec- to minimize the friction between the members during the shear
tion, rotation and curvature, i.e., there is no separation between test, as it was not taken into account in the model.
the two members; and (iii) the connection between the mem-
bers is elastic. This theory has been recently extended to 3.3. Test procedure
account for non-linear relationship of shear load vs. slip
(V  s) and the concrete damage [20,21] by means of a secant Fig. 5 shows the shear test set-up and dimensions of the speci-
stiffness approach [24,28,29]. The extended composite beam mens. Each specimen included a single connector. Two steel guards
model was implemented in a finite element method (FEM) with Teflon plates were attached to the steel frame on both sides of
and validated using experimental tests on TCC beams with a the timber beam to prevent out-of-plane rotation of the specimen
fairly satisfactory accuracy [21,30]. under load. The tests were performed according to the EN 26891
S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92 87

Table 1
Average values of mechanical properties of materials used in shear tests (MPa).

Component Product E fb ft fc fy fu

Timber beam NordicLam 24F-ES/NPG 12,400 45.9 30.8 54.8 – –


Concrete slab Standard concrete – – – 41 – –
Connector shell UHPFRC 50,000 – 14.2 150 – –
Standard mortar – – – 34 – –
Connector core Threaded rod 210,000 – – – 568 633
Rebar 207,000 – – – 434 599

E: Young’s modulus, f b : Flexural strength, f t : Tension strength, f c : Compression strength, f y : Yielding stress, f u : Ultimate strength.

Steel guard
bc dc
lc 410

200 140
lw A A
75
ds
Section A-A 100 240
Teflon plates
F
80
Fig. 4. Geometry of the composite connector.

Table 2
Configurations of the tested connectors.
250

Concrete slab
Test Concrete shell Steel core dc [mm] ds [mm] lw [mm]
#01–02 UHPFRC - 25.4 0 95 480
#03–04 UHPFRC TR⁄ 25.4 5 (M6) 95

Steel frame
#05–06 Mortar TR 25.4 5 (M6) 95 Timber
#07–08 UHPFRC TR 25.4 10.2 (M12) 95 beam
#09–10 UHPFRC RB⁄ 25.4 10 (10 M) 95
#11–12 Mortar TR 25.4 10.2 (M12) 95
#13–14 UHPFRC TR 34.9 5 (M5) 135
#15 UHPFRC TR 34.9 10.2 (M12) 135 LVDT
#16 UHPFRC RB 34.9 10 (10 M) 135

TR: Threaded Rod, RB: Reinforcement Bar.

Fig. 5. Shear test set-up.

[37] without the pre-load cycle. The average connection slip was
determined by measuring the displacements between the concrete
and timber with two LVDT’s installed on the sides of the specimen, The results show that the increase of the steel core diameter
as shown in Fig. 5. from 5 to 10 mm effectively increases the connection resistance
Different composite connectors were tested under shear force (F max ) approximately 80%. The increase of the external concrete
by varying the concrete and steel diameters as reported in Table 2. diameter from 25 to 35 mm increases the connection stiffness
The length of the connector in the wood (lw ) was about 95 and (ki ) at least 67%. The connectors with a threaded rod provided
135 mm for the connectors with a diameter of about 25 and 75% greater connection stiffness than the connectors reinforced
35 mm, respectively. with a rebar of the same diameter, possibly due to a better bond
between the concrete and the threads. Not surprisingly, the con-
3.4. Test results nectors without steel core (#01–02) showed significant dispersion
of the connection performance parameters, whereas the connec-
The connectors made of mortar were susceptible to cracking tors with steel cores showed better repeatability of results, espe-
due to concrete shrinkage and their test results are reported here cially the maximum resistance, with the average difference of
only for sake of completeness (except for the connector #06 that 4.8% between two matched specimens. However, the variation in
cracked during the installation), but disregarded in the following the connection stiffness between the matched specimens was
discussion. None of the UHPFRC connectors cracked during instal- rather significant, most likely due to the low tolerances of fabrica-
lation. Table 3 reports the test results in terms of the maximum tion of connectors, which may have lead to gaps in the assembly
load (F max ), the slip modulus (ki ), the ultimate slip (du ),and the fail- and poor initial contact with timber. To improve the fabrication
ure modes. The maximum load was measured within the slip of tolerances, steel moulds can be utilised and/or the gaps can be
3 mm. The slip modulus represents the slope of the linear portion filled with epoxy or other gap fillers.
of the load-slip curve. The ultimate slip corresponds to the signif- Table 3 indicates that the matched specimens reproduced the
icant load drop indicating the failure of the connector. The failure same failure modes. The failure mode (1), shear failure of the con-
modes were determined via examination of the connection sur- nector, was observed most frequently (10 out of 15 tests). It repre-
faces after the test, which are illustrated in Fig. 6. sents the target failure mechanism, because it depends mostly on
88 S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92

Table 3
Shear test results.

Test F max ki du Failure Test F max ki du Failure


[kN] [kN/mm] [mm] mode [kN] [kN/mm] [mm] mode
#01 8.44 22.4 3.0 (1) #09 27.85 14.4 >15 (3)
#02 12.42 13.5 5.0 (1) #10 29.00 21.6 >15 (3)
#03 16.93 28.6 7.4 (1) #11 20.38 19.9 12.9 (2)
#04 17.59 11.8 7.7 (1) #12 19.73 8.8 12.6 (2)
#05 4.15 18.1 9.7 (2) #13 28.06 48.0 1.1 (1)
#06 – – – (–) #14 27.82 36.3 1.8 (1)
#07 35.95 37.0 12.1 (1) #15 51.24 52.8 12.0 (1)
#08 31.36 26.2 11.2 (1) #16 50.11 29.9 >15 (1)

(1) Shear failure of the connector at the interface; (2) Pull-out of the steel core from the connector head; (3) Pull-out of the steel core from the connector shank.

(a) #01 (b) #03 (c) #05 (d) #07 (e) #10

(f) #11 (g) #13 (h) #15


Fig. 6. View of the connection surfaces after failure.

the properties of the prefabricated connector, which are better 4. Modelling of the shear tests
controlled and less variable than those of the connected members.
The failure mode (2), pull-out of the steel core from the connector 4.1. Model parameters
head, was observed in connectors made with regular mortar, and it
was accompanied with cracking in the concrete slab. It shows that The shear tests were analysed with the Winkler model pre-
the use of mortar for this type of connection is not desirable. The sented in Section 2.2, which is particularly suitable for a discrete
failure mode (3), pull-out of the steel core from the connector elongated connector working in flexure like a beam on an elastic
shank, was observed on connectors made with a rebar, and it foundation [17,21]. The material law of the UHPFRC in compres-
was followed up with wood crushing produced by the withdrawn sion is described by Eqs. (1)–(6) proposed by [38,39] and it is illus-
steel rod. This failure mode demonstrated the weak bond between trated in Fig. 7(a).
the rebar and the concrete shell and, along with the lower stiffness  g
of the connection discussed above, it proved to be less desirable in r ¼ f 0c ð1Þ
1;f g  1 þ  gu
this application. 1;f
S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92 89

160 16 16

140 14 14

120 12 12

Stress ( ) [MPa]
10 10
Stress ( ) [MPa]
100 w
8 0 0 8 0 14.2
80 0.000216 10.8 0.1 6.8
0.005 14.2 2.45 0
60 6 6

40 4 4

20 2 2

0 0
0 0.2 0.4 0.6 0.8 1.0 0 0.1 0.2 0.3 0.4 0.5-0 0.5 1.0 1.5 2.0 2.5

Strain ( ) [%] Strain ( ) [%] Crack opening (w) [mm]


(a) Compressive law (b) Tensile law

Fig. 7. Material laws of UHPFRC.

where, of the steel rebar and the threaded rods were experimentally deter-
" # " # mined from tensile tests according to ASTM E8 [36] and the average
02=3
k0 fc 20 curves based on three samples are shown in Fig. 9. The elastic stiff-
1 ¼ 1 þ 0:16 02
; 2 ¼ 1þ 0 1 ð2Þ ness of the Winkler foundation was determined from [17], for con-
f c þ 800 k0 fc
" # " # crete: kc ¼ 10 GPa, for wood: kw ¼ 1:3 GPa. The non-linear
ft ft
1;f ¼ 1þ4 0 1 ; 2;f ¼ 1 þ 15 0 2 ð3Þ performance parameters of the wood foundation were determined
fc fc via linear interpolation of the results shown in Fig. 3.
2;f k
X¼ ; g¼ ð4Þ
1;f k1 4.2. Analysis results
1;f Ec
k ¼ Ec 0 ; k0 ¼ 01=3
ð5Þ Fig. 10 shows the calculated load-slip curves plotted against the
fc fc corresponding experimental curves for each tested connector
(
1 if  < p (except for those made with mortar). While some sort of agree-
u¼ ln 1gþgX=0:7 ð6Þ ment between the curves can be found for the connectors without
g ln X otherwise
steel core (#01–02) and with 5 mm TR core (#03–04 and #13–14),
0 the predictions for the other connectors were rather poor, with the
The values of the strength in compression (f c ) and tension (f t ) and
initial stiffness and strength being significantly underestimated.
on the Young’s modulus (Ec ) of the UHPFRC shown in Table 1 were
Wood density of the Alpine red spruce (Picea rubens) is lower
used in the analysis. The tensile law of UHPFRC illustrated in Fig. 7
than Douglas fir (Pseudotsuga menziesii) used in the tests. Also, it
(b) was determined by inverse analysis from third-point bending
is known that the embedment stiffness and strength of wood
tests (Fig. 8), which were carried out on four UHPFRC prismatic
depend on the dowel diameter. Therefore, it is plausible that the
beams with a span of 1200 mm, height of 100 mm and width of
model parameters kw and f w assumed in Section 4.1 for the analy-
40 mm. The comparison between the experimental and calculated
sis were not representative of the tested material. To verify this
flexural behaviour presented in Fig. 8 in terms of load vs. mid-
hypothesis in the absence of the material test data, we calibrated
span deflection shows satisfactory agreement. The material laws
the wood foundation properties by inverse analysis of the model.

14
Numeric 700 700
12 Experimental
600 600
10 σ
500 500
Load (Q) [kN]

Stress (σ ) [MPa]

8
400 400
Q/2 Q/2
6 400 400 400 300 300
σ
v
4 200 200
1200
2 100 100 Threaded rod
Rebar
0 0
0 5 10 15 20 0 5 10 15-0 1 2 3 4 5 6
Deflection (v) [mm] Strain ( ) [%] (w) [mm]
Fig. 8. Bending test on UHPFRC beam. Fig. 9. Material laws of steel cores.
90 S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92

50 50
s Test
45 V 45 #16
Calculated 15
40 V Experimental 40
Test #15
35 35 #16
#7
Load (V) [kN]

Load (V) [kN]


30 #10
#8 30
#9 #13-14
25 #7-8 25
#9-10 s
V
20 20
#3-4 V
15 #4 15
#3
10 #1-2 10
#2
5 5 Calculated #13 #14
#1
Experimental
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Slip (s) [mm] Slip (s) [mm]
(a) 25 mm concrete diameter (b) 35 mm concrete diameter
Fig. 10. Calculated vs. experimental load-slip curves.

50 kw = 2500 MPa 50
#15
s fw = 1100 N/mm
45 V 45 #16
Calibrated #15
40 V Experimental 40
Test
35 #7 35
#10
Load (V) [kN]

Load (V) [kN]

30 30 #13-14
#7-8 #8
#9
25 #9-10 25
s
#3-4 V
20 20
V
15 #4 15
#3 kw = 3000 MPa
#1-2
10 10 fw = 1600 N/mm
#2
5 5 Calibrated #13 #14
#1
Experimental
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Slip (s) [mm] Slip (s) [mm]
(a) 25 mm concrete diameter (b) 35 mm concrete diameter
Fig. 11. Calibrated vs. experimental load-slip curves.

Fig. 11 shows the calibrated curves obtained using the best-fitted assembled with connectors #03, #07, #13 and #15 with their
parameters of the wood foundation, which correspond to the respective experimental (V  s) curves from Fig. 10 and uniform
experimental curves more closely. This exercise illustrated the spacing along the beam span, as shown in Table 4, selected in such
influence of the wood foundation properties on the connection a way to achieve similar initial flexural stiffness of the beams. For
behaviour. comparison, the performance of the same TCC beam assembled
with a continuous steel mesh connector with the connection law
shown in Fig. 1 was calculated.
5. Design for ductile failure of TCC beams Fig. 13 shows the calculated flexural response of the beams in
terms of load vs. mid-span deflection. The calculations predicted
This section shows the prediction of the flexural behaviour of a beam failure due the wood rupture in tension at the exterior fibre
TCC beam using the experimental connection law (V  s) and the except for the beam with connector #13 where the connectors
assessment of the possible gain in the structural ductility by would break first. The performance parameters of each analysed
employing different composite connectors. beam are presented in Table 4, where the flexural stiffness was cal-
The configuration and material properties of the analysed TCC culated with c-method [22] and the ductility ratio (l) was defined
beam are presented in Fig. 12. Timber beam was assumed to be as follows:
Nordic-Lam 24F-ES/NPG glulam with the same properties as
shown in Table 1 representing the average values adjusted to vu  ve
l¼ ; ð7Þ
10 min load duration and size effect in bending. For the concrete ve
slab, the mean values of the properties were taken in accordance
with the mechanical relationships proposed in CSA A23.3-04 [40] where v u is the deflection at failure and v e is the limit of the elastic
for regular concrete. The analyses were performed for the beams deflection.
S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92 91

500

c 100 Q/2 Q/2


Ec = 26.6 GPa 2500 mm 2500 mm 2500 mm
f = 35 MPa
w 318 fc = 3.5 MPa
t
Ew = 12.4 GPa
Fb = 45.9 MPa
Ft = 30.8 MPa 7500 mm
137

Fig. 12. Configuration and material properties of the analysed TCC beam.

Table 4
Performance parameters of TCC beams with various connectors.

Connector dc Steel Spacing ðEIÞef Q max ve vu l


[mm] core [mm] [Nmm2 1012 ] [kN] [mm] [mm] [–]

Continuous⁄ – – – 21.17 152 55.5 66.6 0.20


#03 25.4 M6 150 18.05 106 37.7 100.2 1.66
#07 25.4 M12 200 17.96 140 57.5 79.8 0.39
#13 34.9 M6 250 18.08 75 33.6 34.3 0.02
#15 34.9 M12 250 18.38 145 60.6 75.9 0.25

Continuous steel mesh connector.

composite connector (#15 with dc ¼ 35 mm, ds ¼ 10 mm) one


160 can significantly improve the structural ductility by only losing
Steel mesh connector
n o
cti

#15 15% of the flexural stiffness and the load bearing capacity of about
ea

140 #07 8% relative to the beam with continuous steel mesh connector.
t
osi
mp

120
co

#03
ect

6. Conclusion
Load (Q) [kN]

100
rf
Pe

This paper presents an original work on a new composite con-


80
nector for TCC structures, which is composed of an elongated cylin-
#13 der of UHPFRC and a steel core. Based on the presented results, the
60 tion
te ac
omposi following conclusions can be drawn:
40 No c
Q/2 Q/2
1. Experimental shear tests with the proposed composite connec-
20 tors of various configurations demonstrated that the connection
v stiffness mainly depends on the concrete shell diameter and the
0 connection resistance mainly depends on the steel core
0 20 40 60 80 100 diameter.
Deflection (v) [mm] 2. The connections with the composite connectors demonstrated
acceptable repeatability of the maximum shear resistance, but
Fig. 13. Calculated load–deflection curves of TCC beams with various connectors.
the dispersion on the connection stiffness needs to be improved
by tightening the fabrication tolerances or using gap fillers.
3. The performance of the shear connections can be effectively
analysed using the Winkler model for a beam on elastic founda-
The results confirm that by choosing the connector and the tion provided that non-linear stiffness of the wood foundation
spacing the designer can achieve desired stiffness, strength and is properly evaluated as a function of wood density and external
ductility of a TCC structure. TCC beams with similar flexural stiff- connector diameter.
ness can have remarkably different strength and ductility ratios. 4. The analysis of the flexural behaviour of a TCC beam showed
The continuous steel mesh connector provides nearly perfect com- that by choosing a proper composite connector diameter and
posite action with higher flexural stiffness and strength of the spacing, it is possible to improve the ductility of TCC beams
beam than the discrete connectors, but a very low ductility. The without significantly reducing their flexural stiffness and load-
connector #13 (dc ¼ 35 mm, ds ¼ 5 mm), which presents a rather ing bearing capacity.
brittle failure during the shear test, causes a brittle failure of the
TCC beam at a very low load. This confirms that a brittle connec-
tion would cause premature and brittle structural collapse. The
connector #03 provides the highest ductility ratio to the beam Acknowledgements
with a reduction of the maximum load of about 30% with respect
the steel mesh connector. The beams with connectors #07 We would like to acknowledge the National Sciences and Engi-
(dc ¼ 25 mm, ds ¼ 10 mm) and #15 (dc ¼ 35 mm, ds ¼ 10 mm) neering Research Council of Canada funding – Collaborative
present lower ductility ratios, because the resistance of connectors Research and Development Grants (RDCPJ 445200) – associated
was not completely achieved before the collapse of the timber. to the NSERC Industrial research chair on ecoresponsible wood
However, this example demonstrates that using an appropriate construction (CIRCERB) for supporting the present research.
92 S.C. Auclair et al. / Construction and Building Materials 112 (2016) 84–92

References [20] M.R. Salari, E. Spacone, P.B. Shing, D.M. Frangopol, Nonlinear analysis of
composite beams with deformable shear connectors, J. Struct. Eng. 124 (10)
(1998) 1148–1158.
[1] M. Van der Linden, Timber-concrete composite beams, HERON 44 (3) (1999).
[21] S. Auclair, L. Sorelli, A. Salenikovich, Simplified Non-Linear Model for Timber-
[2] A. Ceccotti, Composite concrete-timber structures, Prog. Struct. Eng. Mater. 4
Concrete Composite Beams, submitted for publication.
(3) (2002) 264–275.
[22] Eurocode 5: Design of timber structures - Part 1-1: General - Common rules
[3] N. Mascia, J. Soriano, Benefits of timber-concrete composite action in rural
and rules for buildings, Europan committee for standardization, EN 1995-1-1,
bridges, Mater. Struct. 37 (2) (2004) 122–128.
2004.
[4] A. Buchanan, B. Deam, M. Fragiacomo, S. Pampanin, A. Palermo, Multi-storey
[23] E.W. Kuenzi, Theoretical design of a nailed or bolted joint under lateral load,
prestressed timber buildings in New Zealand, Struct. Eng. Int. 18 (2) (2008)
1960.
166–173.
[24] S.W. Smith, C.A. Beattie, Secant-method adjustment for structural models,
[5] R. Le Roy, H.S. Pham, G. Foret, New wood composite bridges, Eur. J. Environ.
AIAA J. 29 (1) (1991) 119–126.
Civil Eng. 13 (9) (2009) 1125–1139.
[25] P. Paultre, Structure en béton armé, Analyse et dimensionnement, Presses
[6] E. Steinberg, R. Selle, T. Faust, Connectors for timber-lightweight concrete
internationales polytechnique, 2011.
composite structures, J. Struct. Eng. 129 (11) (2003) 1538–1545.
[26] U. Girhammar, D. Pan, Exact static analysis of partially composite beams and
[7] B. Ahmadi, M. Saka, Behavior of composite timber–concrete floors, J. Struct.
beam-columns, Int. J. Mech. Sci. 49 (2) (2007) 239–255.
Eng. 119 (11) (1993) 3111–3130.
[27] U.A. Girhammar, V.K. Gopu, Composite beam-columns with interlayer
[8] D. Yeoh, M. Fragiacomo, M. De Franceschi, K.H. Boon, State of the art on timber-
slipexact analysis, J. Struct. Eng. 119 (4) (1993) 1265–1282.
concrete composite structures: literature review, J. Struct. Eng. (2011).
[28] J.N. Reddy, An Introduction to Nonlinear Finite Element Analysis, Oxford
[9] E. Lukaszewska, H. Johnsson, M. Fragiacomo, Performance of connections for
University Press, 2004.
prefabricated timber-concrete composite floors, Mater. Struct. 41 (9) (2008)
[29] J. Mazars, A description of micro-and macroscale damage of concrete
1533–1550.
structures, Eng. Fract. Mech. 25 (5) (1986) 729–737.
[10] E. Lukaszewska, M. Fragiacomo, H. Johnsson, Laboratory tests and numerical
[30] Software DDuctileTCS, http://www.gci.ulaval.ca/DDuctileTCS, accessed: 2015-
analyses of prefabricated timber-concrete composite floors, J. Struct. Eng. 136
03-01, 2015
(1) (2009) 46–55.
[31] Technical note S01, Limite states design (CAN), Nordic, Wood Structures, 2015.
[11] L. Bathon, P. Clouston, Experimental and numerical results on semi prestressed
[32] CSA O86–14, Engineering design in wood, Canadian Standars Association,
wood-concrete composite floor systems for long span applications, in:
2014.
Proceedings of the 8th world conference on timber engineering, vol. 1, 2004,
[33] CSA A23.2-04, Methods of test and standard practices for concrete, 2004.
339–44.
[34] V.H. Perry, P.J. Seibert, The use of UHPFRC (DuctalÒ) for bridges in North
[12] A.M. Dias, L.F. Jorge, The effect of ductile connectors on the behaviour of
America: The technology, applications and challenges facing
timber-concrete composite beams, Eng. Struct. 33 (11) (2011) 3033–3042.
commercialization, in: Second International Symposium on Ultra High
[13] M.S. Mungwa, J.-F. Jullien, A. Foudjet, G. Hentges, Experimental study of a
Performance Concrete, University of Kassel, Germany, pp. 815–822, 2008.
composite wood–concrete beam with the INSA–Hilti new flexible shear
[35] G. Chanvillard, S. Rigaud, Complete characterization of tensile properties of
connector, Constr. Build. Mater. 13 (7) (1999) 371–382.
Ductal UHPFRC according to the French recommendations, in: Proceedings of
[14] M. Piazza, M. Ballerini, Experimental and numerical results on timber-concrete
the 4th International RILEM workshop High Performance Fiber Reinforced
composite floors with different connection systems, in: 6th world conference
Cementitious Composites, pp. 21–34, 2003.
on timber engineering WCTE, 2000.
[36] ASTM E8–04, Standard Test Methods for Tension Testing of Metallic Materials,
[15] A. Ceccotti, Timber-concrete composite structures, chap. E13, Timber
ASTM International, West Conshohocken, PA, 2004.
Engineering, step 2, The Netherlands, Centrum Hout, 1st ed., E13/3, 1995.
[37] EN 26891, Timber strucutre – Joints made with mechanical fasteners – General
[16] C. Zhang, P. Gauvreau, Timber-concrete composite systems with ductile
principles for the determination of strength and deformation characteristics,
connections, J. Struct. Eng. 141 (7) (2014) 1–12.
European Commitee for Standardization CEN, 1991.
[17] P. Gelfi, E. Giuriani, A. Marini, Stud shear connection design for composite
[38] S. Popovics, A numerical approach to the complete stress-strain curve of
concrete slab and wood beams, J. Struct. Eng. 128 (12) (2002) 1544–1550.
concrete, Cem. Concr. Res. 3 (5) (1973) 583–599.
[18] A. Frangi, M. Fontana, Elasto-plastic model for timber-concrete composite
[39] M.P. Collins, A. Porasz, Shear design for high-strength concrete, CEB Bulletin
beams with ductile connection, Struct. Eng. Int. 13 (1) (2003) 47–57.
d’information 193 (1989) 77–83.
[19] N.M. Newmark, C.P. Siess, I.M. Viest, Test and analysis of composite
[40] CSA A23.3-04, Concrete Design Handbook, Third edition, Canadian Standars
beams incomplete interaction, Proc. Soc. Exp. Stress Anal. 9 (1) (1951)
Association, 2008.
75–92.

You might also like