Jing Luo Et Al 2022

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Building Engineering 46 (2022) 103804

Contents lists available at ScienceDirect

Journal of Building Engineering


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

Experimental and numerical investigation into the fire


performance of glulam bolted beam-to-column connections under
coupled moment and shear force
Jing Luo a, Minjuan He a, Zheng Li a, *, Zhaozhuo Gan a, Xijun Wang a, Feng Liang a, b
a
Department of Structural Engineering, Tongji University, Shanghai, 200092, China
b
Tongji Architectural Design (Group) Co., Ltd., Shanghai, 200092, China

A R T I C L E I N F O A B S T R A C T

Keywords: The fire resistance of connection plays a crucial role in the fire performance of timber structures.
Timber connections During the past three decades, massive real-scale fire tests were performed on timber connections
Beam-to-column connection to analyse the mechanical behavior of connections under fire situation. Although different
Fire performance loading scenarios was considered in previous timber connection tests, little research focused on
Experimental study the beam-to-column connections loaded with a combination of stress caused by bending moment
Numerical model
and shear force. For post and beam structures, when structures undergo large inter-story drift, the
interforce at the connection includes not only shear force but also bending moment. In this study,
real-scale tests are carried out on bolted beam-to-column connections loaded with coupled
bending moment and shear force under normal temperature and ISO fire exposure. Three-
dimensional numerical model of the bolted beam-to-column connections are developed and
validated by experimental results. The model is capable of simulating the evolution of temper­
ature in the connections as well as their mechanical behavior. The comparison of the rotation-
time curve between experimental data and thermo-mechanical model shows comparable ten­
dency. The experiments and numerical models in this study can enrich existed experimental data
of glulam connections and can be used to calibrate the design approach for bolted beam-to-
column connections under fire exposure.

1. Introduction
Fire safety is one of the most significant issues during the design of timber structures. Connection governs the mechanical per­
formance of the structure both in normal temperature and in fire situation.
Timber connections with dowel-type fasteners are commonly used in real projects. In order to analyse the fire performance of
timber connections, many real-scale fire tests were performed on timber connections with varied connection configuration, load ratio
and fasteners type etc. Recent research was mainly focused on three-member connections such as wood-to-wood (WWW) connection,
wood-steel-wood (WSW) connection and steel-wood-steel (SWS) connection. Besides steel-to-timber connection with single slotted-in
steel plate, tests on steel-to-timber connection with multiple slotted-in steel plates loaded in tension were also performed by previous
researchers [1].
Most of dowel-type timber connections were loaded in tension parallel to grain in fire tests [2–12]. The design rules in EN 1995-1-2

* Corresponding author.
E-mail address: zhengli@tongji.edu.cn (Z. Li).

https://doi.org/10.1016/j.jobe.2021.103804
Received 21 October 2021; Received in revised form 25 November 2021; Accepted 29 November 2021
Available online 1 December 2021
2352-7102/© 2021 Elsevier Ltd. All rights reserved.
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

[13] also apply mostly to the symmetrical three-member connections with laterally loaded fasteners. In the tests by Norén [2], WWW
(wood-to-wood) nailed connection were tested and the results showed that nails yielded in both central and side members under
normal temperature. However, in fire situation, failure modes of nail connection changed due to the charring of the wood member. An
experimental program begun in France to investigate the fire performance of WWW connections and WSW connections with bolts and
dowels [3–7,11]. This experimental program intended to analyse the main influencing factors and thus to deduce empirical formulas
on fire resistance of timber connections. Fire tests for WWW connections, WSW connections, SWS connections were conducted by Lau
and Chuo [8,9]. It was noted that the temperature during the fire testing did not follow the ISO 834 curve [14]. Some conversions were
conducted to predict the fire resistance of the connections. A series of tests were carried out by Peng [10] to analyse the fire resistance
of WSW connection and SWS connection in tension loading. Hole elongation and splitting were observed for two kinds of connections.
For most of WSW connections, no bending of bolts occurred, but for SWS connections, bending of bolts were observed after the fire
tests.
Connections loaded in different directions exhibit different mechanical behavior. Therefore, besides the tension loading, other
kinds of loading type were also applied in fire tests by researchers [15–18]. In the tests by Audebert et al. [15], connections were
loaded in tension perpendicular to grain and with an angle of 45◦ under standard fire exposure. Large embedment deformation and
splitting cracks along the fastener rows were observed in these tests. Dhima et al. [16] presents tests on connections loaded in bending.
In this study, cyclic loading was also applied to some specimens before fire tests. Results showed that the post-cyclic loading which
resulting in clearance between the fasteners and the wood seemed to have little influence on the fire resistance of the connections.
Beam-to-column WSW dowelled connections loaded in the direction perpendicular to the grain were tested by Palma et al. [17,18]. A
framework of timber connection model based on the heat transfer model and Johansen-type load-carrying model was developed to
predict the resistance of the connections in normal temperature and in fire.
Real-scale fire tests on timber connections are quilt expensive and complex. Thus, numerical models have been developed by
researchers [5,11,19]. The temperature-dependent physical and mechanical properties of wood and steel are considered. Such pa­
rameters are effective because they combine the coupled mass transfer and heat. Therefore, numerical models of timber connections in
fire usually consist of heat transfer model and thermo-mechanical model. The coupling between the thermal evolution and mechanical
phenomenon is neglected [11]. The temperature field calculated from heat transfer model is then put into the thermo-mechanical
model to analyse the mechanical response of the connections in fire situation.
Although different loading directions were applied on timber connections by previous researchers, little research focused on the
beam-to-column connections loaded with a combination of stress. For post and beam structures, when the structures undergo large
inter-story drift, the interforce at the connection includes not only shear force but also bending moment [20]. Furthermore, the existed
fire tests of connections loaded with an angle to the grain were mainly conducted on dowelled connections. However, for bolted
connection, the fire resistance time will decrease compared with dowelled connection due to large exposed face of bolt head in fire
situation [21]. Thus, fire tests are needed to investigate the fire behavior of bolted connection loaded under coupled bending moment
and shear force.
Considering the architectural and strength requirements, steel-to-timber connections with slotted-in steel plate are commonly used
in glulam post-and-beam structures and thus such connections are adopted as the test specimens in this study. The steel-to-timber
connections with slotted-in steel plate have the characteristics of high strength and ductile failure mode. The slotted-in steel plate
can be protected by the side wood member in fire and thus have better fire performance especially compared with SWS (steel-wood-
steel) connections.
In this study, real-scale tests are carried out on bolted steel-to-timber beam-to-column connections loaded in coupled shear force
and bending moment under normal temperature and ISO fire exposure. Three-dimensional numerical model of the connections under
normal temperature is developed and validated by experimental data. Then, the heat transfer models are presented and validated by
the evolution of temperature from selected thermocouples in the tests. Finally, thermo-mechanical models are established based on the
mechanical model in normal temperature and temperature field in heat transfer model.

Fig. 1. Basic framework.

2
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

2. Experimental program
Glulam bolted beam-to-column bolted connections with slotted-in steel plate were designed as the tested specimens. Connections
were loaded under normal temperature and ISO fire exposure. The aim of the tests in normal temperature was to obtain the load-
carrying capacity of the connections. Then, fire tests were carried out under a certain load ratio to investigate the fire performance
of the connections. Basic framework this research program is shown in Fig. 1.

2.1. Specimen design


Configurations of the test specimens are shown in Fig. 2. Douglas fir was selected to manufacture the glulam for the tests. The
moisture content of glulam was measured as 10.46%. Material tests for the Douglas fir lumber were carried out in advance according to
ASTMD143 [22] and the results of the material tests were used as the input for the numerical models. The diameter of bolts, beam
geometrics were changed between two sets of connections. The spacing, end-distance and edge-distance of the bolts satisfy the re­
quirements of Technical code of glued laminated timber structures (GB/T50708-2012) [23]. Shown from Table 1 and Table 2, two
configurations of glulam bolted beam-to-column connections were tested in normal temperature (AC1 and AC2) and in fire condition
(ET1 and ET2). Each group had two duplicates. Therefore, in this study, 8 connection specimens were tested in total. Specimens in
group AC1 and AC2 were tested under cyclic loading in normal temperature with 2 duplicates. Specimens in group ET1 and ET2 were
tested in fire condition with 2 duplicates. Shown from Fig. 2, the geometrical dimensions for the glulam beams for connection 1
(AC1/ET1) and connection 2 (AC2/ET2) were set as 200 × 350× 1000 mm3. The geometrical configurations of glulam beams in
normal-temperature tests were the same with the glulam beams in fire tests. Besides, fasteners for connection 1 and connection 2 were
M16 and M20 high strength bolts, respectively. The grade of the bolts was 8.8s. The nominal tensile strength and the ratio of yield
strength to ultimate strength were 800 MPa and 0.9, respectively [24]. The column side had more bolts than the beam side. This was
intended to avoid the premature failure of column side, because this test was mainly to focus on the beam side to simplify the analysis
model and most connections were estimated to fail in the beam side [25–27]. The slotted-in steel plate was made by steel Q345 [28]
with thickness of 10 mm, and the yield strength of the steel plate was 345 MPa.

2.2. Tests in normal temperature


The test setup in normal temperature is shown in Fig. 3. The column was placed horizontally on the ground for its stability [27]. It is
a simplified way to neglect the influences from the column. The beam was linked to an electro-hydraulic servo actuator. The rotation of
the connection was recorded by six LVDTs (linear voltage displacement transducers). As shown in Fig. 3, LVDT1, LVDT3 and LVDT4
were selected to measure the rotation of the beam. LVDT3 and LVDT4 were symmetrically arranged on both sides of steel plate. LVDT2
and LVDT5 were selected to measure the rotation of the steel plate. LVDT6 was used to measure the displacement of the column. Cyclic
loading was applied to the tested specimens loading regime in FEMA461 [29]. Displacement-control protocol was employed for the
cyclic loading. Two cycles at each amplitude were completed. The amplitude ai+1 of the step i+1 is determined by multiply the
amplitude of the preceding step by 1.4. If the specimen has not reached the final damage at target maximum deformation amplitude,
the amplitude shall be determined by 1.3 times the target maximum deformation amplitude. 5 mm/min was selected as the rate of
displacement-control protocol employed for the cyclic loading in early circles. In addition, to save the tested time, 10 mm/min was
selected as the rate of displacement-control protocol employed for the cyclic loading in the last two steps in the loading history.
Test results and observed damage of the connections in normal temperature are shown in Fig. 4. The moment of the connection is
determined by the shear load times the distance between the load point and the beam-to-column surface. Severe embedment failure of
the wood around the bolt hole and splitting cracks along the bolt line occurred after the cyclic loading. In addition, all bolts in beam
side bended. The mean ultimate capacity of the connections for Group AC1 and AC2 were 42.68 kN · m and 37.72 kN · m, respectively.
These values were used to determine the applied load on connections in fire tests.

2.3. Tests under ISO fire exposure


The fire tests were performed on the horizontal furnace. The same configurations of connections in normal temperature were used
in fire tests. In this research, it was mainly to focus on the beam side of connections, because this part of connection would be most
exposed to fire and thus was critical for the fire resistance [25]. The connection specimen was partially protected by insulation as
shown in Fig. 5. The connection was covered by 4 layers of fireproof insulation including 1 layer of gypsum board (9.5 mm), 2 layers of

Fig. 2. Configurations of connection 1 and connection 2 as test specimens (dimensions are in millimeter).

3
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Table 1
Geometric configurations of tested specimens.

Connection type Number of Thickness of beam b Height of beam h (mm) Thickness of side wood t1 Bolt diameter d (mm)
Duplicates c (mm) (mm)

Connection 1 (AC1/ 2 200 350 95 16


ET1)
Connection 2 (AC2/ 2 200 350 95 20
ET2)
Connection type Edge distance e1 Edge distance e2 Spacing parallel to grain r Spacing parallel to grain s Thickness of steel plate t
(mm) (mm) (mm) (mm) (mm)
Connection 1 (AC1/ 120 111 85 64 10
ET1)
Connection 2 (AC2/ 140 127 100 96 10
ET2)

Table 2
Group assignment of tested specimens.

No. Connection type Test type Load ratio

AC1-1 Connection 1 Tests are under normal temperature –


AC1-2 –
AC2-1 Connection 2 –
AC2-2 –
ET1-1 Connection 1 Tests are under ISO fire exposure 15%
ET1-2 25%
ET2-1 Connection 2 15%
ET2-2 25%

Fig. 3. Test setup and instrumentation in normal temperature.

retardant cotton (20mm × 2) and 1 layer of gypsum board (9.5 mm). The whole column and half of beam were covered by fireproof
insulation. This was intended to make sure that no burning occurred in the protected area.
Test setup in fire is shown in Fig. 6. The column was inserted into the brick base for stability reason. Two steel members were placed
over the glulam column and glulam beam respectively for force transfer. Actuator and transducer were placed on the two steel
members. The steel members were protected by fire proofing coatings. Displacement transducers were used to measure the defor­
mation of the column and beam in fire tests. Displacement transducers were placed outside the furnace to avoid any negative impact of
high temperature on the equipment. Two holes had been punched in the end of steel member so that a flame-Retardant steel wire got

4
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Fig. 4. Observed damage and hysteretic curves of the test specimens.

Fig. 5. Arrangement of fireproofing area.

across the openings to connect connection specimen and displacement transducers. When the test setup was ready, the beam was
loaded to the target load level (load ratio = 15% or 25% in these tests) and the load kept constant throughout the fire test. Once the
burners started, the connection was exposed to ISO standard fire (ISO 834–1:1999 [14]). When the connection was no longer sustain
the applied load, the burners stopped and the connection was removed from the furnace and cooled with water.
The observed damage of specimens after the fire tests is shown in Fig. 7. The average value of charring rate of lateral wood members
was measured after the fire tests. The charring rates are measured as 0.72 mm/min and 0.82 mm/min for the exposed wood member in
no-bolt area and bolt area (Fig. 8), respectively. Different from the failure mode in normal temperature, severe embedment failure was
observed in wood around the bolt hole due to the charring of the wood but the bolts were not bended. Wood protected by the insulation
were not burned during the fire test, indicating that fire insulation measures were effective to avoid fire. Values of displacement
transducers from the column were quite small during the fire test, indicating that the deformation of the column was quite small and no
premature failure of the column occurred during the fire test. Results of fire resistance tests are presented in Table 3.

5
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Fig. 6. Test setup and instrumentation in fire.

Fig. 7. Observed damage of specimen after fire tests.

Fig. 8. Charring depths of the connection.

6
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Table 3
Results of fire resistance tests.

No. Thickness of side wood t1 Bolt diameter Experimental load-carrying capacity in normal Load Fire resistance tfi
(mm) d (mm) temperature ratio (min)

ET1- 95 16 42.68 15% 39.98


1
ET1- 95 16 25% 33.50
2
ET2- 95 20 37.72 15% 41.52
1
ET2- 95 20 25% 33.67
2

3. Numerical simulations
In this study, numerical models in normal temperature and in fire were performed as presented in section 3.1 and section 3.2
respectively.

3.1. Numerical models in normal temperature


A 3D finite element model of bolted steel-to-timber connection was generated using Abaqus [30]. Then, the accuracy of the model
was validated by the experimental tests. The washers were not generated between nut and wood in this model. Half of the connection
was created as the connection was symmetric. As no premature failure occurred in the column side of the connection in the experi­
mental tests and the beam side covered the ultimate capacity of the connection, the focus of this model was on the beam side of the
connection. Therefore, the column side was assumed to be a wood base with fixed end in this model but the beam side was generated in
detail. Arrangement of material properties of beam is presented in Fig. 9.
The rotational performance of bolted connections is usually strongly affected by the embedment area of the beam. Considering the
wood damage due to predrilled holes, a common approach is to reduce the material properties near the pre-drilled holes on the glulam
beam. Thus, wood foundation model was used for the wood near pre-drilled holes [31,32]. The elastic modulus and yield point of wood
foundation model were defined based on a load-embedment curve from bolt-embedment tests. As shown in Fig. 9, The radius of the
wood foundation zone was equal to 1.8 times the diameter of the bolt [33]. The behavior of wood in blue zone (Fig. 9) was assumed to
be transversely isotropic plastic with Hill yield criterion [34]. The Hill yield criterion is calculated by:
( )
f σ, σy = a1 (σ22 − σ 33 )2 + a2 (σ33 − σ11 )2
+a3 (σ11 − σ22 )2 + 2a4 τ223 + 2a5 τ231 (1)
+2a6 τ212 − σ 2y = 0

( )2 ( )2 ( )2
σy 1 σy 1 σy
a1 = − ; a2 = a3 = ;
σy22 2 σ y11 2 σy11
( )2 ( )2 (2)
1 σy 1 σy
a4 = ; a5 = a6 =
2 σ y23 2 σy12

Fig. 9. Arrangement of material properties in FEM.

7
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

y y
where σ i is the yield stress in the direction of i; σ y is equal to σ11 (yield stress parallel to the grain). The modulus of rigidity of wood Gij
was not deduced from material test but was calculated by:
√̅̅̅̅̅̅̅̅̅
Ei Ej
Gij = ( ) (3)
2 1 + υij

where Ei is the elastic modulus in the direction of i. υij is the Poisson’s ratio of wood in the direction of ij.
Material properties of wood that was far away from the pre-drilled holes was assumed to be orthotropic linear elastic [35]. As the
stress of wood in this area was quite small, such assumption could enhance the computation speed and obtain a better convergency.
Detailed parameters for steel and wood in the finite element model are shown in Table 4. Isotropic plastic material model was
applied to steel. Bilinear stress-strain relationship was defined for the constitutive behavior of bolt and steel plate. The Young’s
modulus and Poisson’s ratio of steel were set as 21000 MPa and 0.3, respectively. For the bolt with strength grade of 8.8s, the yield
strength was set as 640 MPa [24]. For the steel plate with strength grade of 345, the yield strength and ultimate strength were set as
345 MPa and 470 MPa, respectively [28].
Considering brittle failure of wood, cohesive zone model was applied to the numerical connection. Contact pairs of cohesive
surfaces were set on the predicted crack path. As splitting cracks were always observed along the bolt row, the predicted crack path was
set on the wood along the bolt row. Reference points were created to apply displacement-controlled load to the beam. Coupled
constrains were set between the reference point and a local surface of beam. Symmetry constrains were applied on middle surface of
the connection. Friction contact was defined in this numerical model. Hard contact was defined between wood and steel. The friction
coefficient between bolt and wood was set as 0.3 [36]. The surfaces of steel elements were selected as the main surfaces, while the
surfaces of wood elements were selected as the slave surfaces. Hard contact was also defined between bolt and steel plate. The friction
coefficient between bolt and steel plate was set as 0.001 [37]. Penalty contact was defined between beam surface and bottom base. The
friction coefficient for such wood-to-wood surface pairs was set as 0.5 [38]. 3D FEM meshing is presented in Fig. 10. The finite element
type was C3D8 in this model. The size of meshing was set as 10 mm. Global and local mesh constraints were adopted to control the
numerical model. The meshing was refined in the wood foundation zone and was set as concentric mesh around the bolt hole. Edges
around the bolt hole were seeded in 24 equal division.
Following these procedures, the comparison between FEM results and experimental data is shown in Fig. 11. The moment-rotation
curves of tested specimens were obtained from the skeleton curves in cyclic tests. The beginning low-stiffness section was cut in the
skeleton curve of tested specimens. Fig. 11 shows that the moment-rotation curves in numerical curves agree well with the test results.
However, the elastic stiffness obtained from numerical models is relatively larger than the experimental data. It is mainly because that
the clearance between the bolt and bolt hole were not considered in the numerical model. But the bolt clearance would decrease the
stiffness of the connection due to low contact area between bolt surface and wood surface [39,40]. The ultimate capacity of the
connection in numerical model is relatively lower than experimental results. It is mainly due to the variability in the material prop­
erties of wood in experimental tests and due to the bad convergence in large deformation of wood in numerical models. In addition, it
can be seen from Fig. 11 that the splitting failure of wood occurs in the outer bolt row for Group AC1 and AC2. The propagation of
cracks in numerical models is quite close to the failure modes in tests. The mechanical behavior and failure modes of connections in
numerical models fit well with experimental results.

3.2. Numerical models in fire


Numerical models of heat transfer and thermo-mechanical analysis were performed in this study. Thermo-physical properties of
wood and steel were considered. The thermo-related values of conductivity, specific heat and density for steel were given by EN1993-
1-2 [41]. Compared with steel, wood is a complicated material concerning coupled heat-mass transfer in fire. The moisture content,
pyrolysis gases and energy-released behavior may affect the thermo-physical properties of wood in the evolution of combustion. The
curves of thermo-physical properties of wood were proposed by different researchers [13,42–46] as shown in Fig. 12. A sensitivity
analysis was conducted to find the most appropriate thermo-physical properties of wood. The results showed that the combination of

Table 4
Detailed parameters for steel and wood in the finite element model.

Bolt Steel plate Wood Wood fooundation

Modulus of elasticity (N/mm2) E = 2.1 × 105 E = 2.1 × 105 EL = 11534 EL = 2503


ER;T = 313 ER;T = 560
Modulus of rigidity (N/mm2) / / GLT;LR = 639 GLT;LR = 432
GRT = 108 GRT = 194
Poisson ratio ν = 0.3 ν = 0.3 νLT;LR = 0.37 νLT;LR = 0.37
νRT = 0.44 νRT = 0.44
Yield stress (N/mm2) fy = 640 fy = 345 σL = 44.03 σL = 45.03
σR;T = 7.58 σR;T = 27.57
Strength (N/mm2) / fu = 470 ft,90 = 2.03 ft,90 = 2.03
fs = 2.16 fs = 2.16
Fracture energies (N/mm) [35] / / GIc = 0.24 GIc = 0.24
GIIc = 0.55 GIIc = 0.55

8
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Fig. 10. 3D FEM meshing.

thermal conductivity values by Janssens [45], density values by EN 1995-1-2 [1] and specific heat values by Fredlund and Janssens
[43,45] led to the best numerical results. In addition, the anisotropy of thermal properties for wood must be considered. The thermal
conductivity of wood in the direction parallel to grain is higher than the direction perpendicular to grain. The curves shown in Fig. 12
(a) only present the thermal-conductivity of wood in the direction perpendicular to grain. The magnifying coefficient for
thermal-conductivity of wood in the direction parallel to grain was set as 4 [47] in this study. Further work concerning this question is
still needed.
Various parts of the numerical model were generated such as the beam, bolt, and steel plate. Perfect contact between those parts
with continuous meshing was used to allow heat transfer through multiple parts. As only half of the connection was generated in this
model. symmetry plane through steel plate was adopted. The same meshing size and distribution with numerical model in normal
temperature were adopted in heat transfer model. 8-node linear heat transfer brick (DC3D8) elements were used in the generated
meshes.
In the fire test, the gas temperature followed the time-temperature curve in ISO 834–1: 1999 [14]. The net heat flux ḣnet includes the
net convective heat flux ḣnet,c and radiative heat flux ḣnet,r as shown in Eq. (4):

ḣnet = ḣnet,c + ḣnet,r (4)

The net convective heat flux ḣnet,c is calculated by:


( )
ḣnet,c = ac θg − θm (5)

where ac is the convection coefficient and is equal to 25 W · m− 2 · K− 1


[17]; θg and θm are gas temperature [◦ C] and surface temperature
[◦ C] respectively.
The radiative heat flux ḣnet,r is calculated by:
[ ]
ḣnet,r = Φ · εm · εf · σ (θr + 273)4 − (θm + 273)4 (6)

where Φ is the configuration factor of the exposed surface and is equal to 1 [17]; εm is the emissivity and is set as 0.8 [1] for wood and
0.7 for steel [41]; εf is the emissivity of the fire and is set as 1 [17]; σ is the Stephan Boltzmann constant and is set as 5.67× 10− 8 W ·
m− 2 · K− 4 [17]. θr and θm are effective radiation temperature [◦ C] and surface temperature [◦ C] respectively.
The fire-exposed surfaces are shown in Fig. 13. Surface film condition and surface radiation were directly applied to the fire-
exposed surfaces. The distribution of temperature field was calculated by heat transfer model.
Thermo-mechanical analysis was then developed using the same mesh sizes in heat transfer model. The evolution of mechanical
properties of steel and wood were dependent on temperature. The reduction factor for the mechanical properties of steel and wood
were selected according to EN 1993-1-2 [41] and EN 1995-1-2 [13] respectively. In this thermo-mechanical model, the evolution of
elastic modulus of wood in compression was used and the evolution of elastic modulus of wood was assumed to be the same in
compression and in tension. In addition, the reduction factor of elastic modulus of wood in the direction parallel to grain was used for
the reduction factor of elastic modulus of wood in the direction perpendicular to grain [11].
Contact interaction in thermo-mechanical models were the same with mechanical model of connection in normal temperature. The
constant load applied during the fire tests was also applied to the thermo-mechanical model. The same mesh size with heat transfer
model was adopted for the mesh size in thermo-mechanical model. It was to make sure that the temperature and coordinate of each
integral point were stored in heat transfer model. Then, the temperature field was inputted into the thermal-mechanical model. For
each temperature increment step, the mechanical properties of each element changed and the nonlinear mechanical calculation was
processed.
Fig. 14 presents thermocouple locations in group ET1. Fig. 15 presents the simulated and experimental temperatures for selected
thermocouple points. The results concerned Group ET1, but all other groups exhibited similar behavior. Fig. 15 (a) presents the
evolution of temperature along the surface of bolt. TC1 was located at the outer surface of bolt, while TC2 was located at 40 mm from
the surface of wood member. The temperature of TC1 is higher than TC2 as TC1 is exposed to fire. Fig. 15 (b) and 15(c) shows the
temperature versus time for selected points in wood member. Pont TC3, TC4 and TC5 was placed at 80 mm from the surface of wood

9
J. Luo et al. Journal of Building Engineering 46 (2022) 103804
Fig. 11. Comparison between FEM results and experimental data.
10
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Fig. 12. Thermo-physical properties of wood.

Fig. 13. Fire-exposed surfaces.

Fig. 14. Thermocouple locations.

11
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Fig. 15. Simulated and experimental temperatures.

member. Pont TC6, TC7 and TC8 was placed at 30 mm from the surface of wood member. TC3 and TC6 were located between two
columns of bolts. TC4 and TC7 were located near the end of steel plate. TC5 and TC8 were located far from the steel plate. The results
show that the temperature of TC3 is higher than TC4 and the temperature TC4 is higher than TC5 even though they are in the same
depth from the surface of wood member. It mainly because the temperature of TC3 is affected by the heat transfer behavior of bolt and
steel plate (TC3 was located near the steel plate and two columns of bolts). The temperature of TC4 is affected by the heat transfer
behavior of steel plate (TC4 was located near the steel plate). Similar phenomenon is also observed for point TC6, TC7 and TC8.
However, they were shallower in depth from the surface of wood member compared with TC3, TC4 and TC5. The influence of heat
transfer behavior of steel plate on the temperature of TC6, TC7 and TC8 are not significant. Thus, the evolution of temperature for TC7
and TC8 are quite similar. The temperature of TC6 is higher than TC7 and TC8. The comparison of temperature-time curves between
experimental data and numerical results shows good accuracy of heat transfer model.
After the failure of the connection, the embedment failure of wood occurs around the bolt hole while the bolts remain unbent.
Similar phenomenon was also shown in connections after fire tests (Fig. 8). Comparison between FEM results and experimental data
are shown in Fig. 16. The rotation-time curves in Fig. 16 show comparable tendency between thermal-mechanical models and
experimental data. The displacement in experimental tests reaches higher values than the displacement in thermal-mechanical model.
It is mainly because that the calculation process in thermal-mechanical model becomes non-convergent and stopped when the
connection reaches high rotation at the beginning of failure process. But it still can be used to calculate the transition between elastic
and plastic stage of the connection. The comparison of the numerical and experimental failure time in fire is presented in Table 5. It can
be observed from Table 5 that the fire resistance time in experimental tests is larger than that in thermo-mechanical model but the error
is acceptable.

4. Conclusion
The main conclusions of this research are summarized below:
1. Experimental results from the tests in normal temperature shows that the connections fail due to severe embedment deformation
and splitting cracks in the wood. The bolts are bended after cyclic loading. The numerical model of connection in normal tem­
perature is validated by experimental data. The numerical model can simulate the progressive cracks of bolted connections under

Fig. 16. Comparison between FEM results and experimental data.

12
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

Table 5
Comparison of the numerical and experimental failure time in fire.

No. Load ratio tfi (min) tfi,sim (min) Relative error

ET1-1 15% 39.98 36.64 − 8.35


ET1-2 25% 33.50 28.90 − 13.73
ET2-1 15% 41.52 39.13 − 5.76
ET2-2 25% 33.67 31.28 − 7.10

coupled shear force and bending moment. The comparison of the rotation-time curve shows that the numerical model could exhibit
nonlinear behavior of the connection. However, further improvement is needed to improve the convergence performance of the
model.
2. Experimental results from ISO-fire tests show that the bolted beam-to-column connections exhibit a ductile behavior in fire with
large embedment deformation of wood around the bolt hole. Bolts are not bended after fire tests, which is quite different with the
failure modes of connections in normal temperature. The charring rates are measured as 0.72 mm/min and 0.82 mm/min for the
exposed wood member in no-bolt area and bolt area respectively. In fire tests, no-burning area of wood were protected by one layer
of gypsum board, two layers of retardant cotton and one layer of gypsum board. Results show that the fire insulation measures were
effective to avoid fire and can be used as reference for fire tests in the future.
3. Numerical model in fire includes two stages: heat-transfer model and thermo-mechanical model. The heat-transfer model with the
selected thermal properties shows good agreement with experimental data. The same mesh size with heat transfer model was
adopted for the mesh size in thermo-mechanical model. Then, the temperature field was inputted into the thermal-mechanical
model. Similar failure modes of connections in experimental tests can be simulated by the thermo-mechanical model. The com­
parison of the rotation-time curve between experimental data and thermo-mechanical model shows comparable tendency. How­
ever, the fire resistance time in thermo-mechanical model is lower than that in experimental tests (with a maximum relative error of
− 13.73%), indicating a safe design approach towards bolted beam-to-column connections in fire. The experiments and numerical
models in this study can enrich existed experimental data of glulam connections and to calibrate the design approach for bolted
beam-to-column connection under ISO fire exposure.

Compliance with ethical standards


This article does not contain any studies with human participants or animals performed by any of the authors.

Credit authorship contribution statement


Jing Luo: Investigation, Data curation, Software, Writing - original draft, Writing - Review & Editing. Minjuan He: Supervision,
Methodology, Investigation. Zheng Li: Conceptualization, Methodology, Investigation. Zhaozhuo Gan: Investigation, Software. Xijun
Wang: Investigation, Software. Feng Liang: Validation.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgements
The authors gratefully acknowledge the National Natural Science Foundation of China (Grant No. 51878476) and National Natural
Science Foundation of China (Grant No. 52078371) for supporting this work.

References
[1] C. Erchinger, A. Frangi, M. Fontana, Fire design of steel-to-timber dowelled connections, Eng. Struct. 32 (2010) 580–589.
[2] J. Norén, Load-bearing capacity of nailed joints exposed to fire, Fire Mater. 20 (1996) 133–143.
[3] D. Dhima, Vérification expérimentale de la résistance au feu des assemblages d’éléments en bois, St. Remy Les Chevreuses : Constr. Met. (CTICM) (1999) 61.
[4] P. Racher, Résistance incendie des assemblages bois. Rapport n◦ 2002-55-1. Comportement à froid des assemblages bois-métal, Plateforme MSCG,
Polytech’Clermont-Ferrand, Décembre, 2002.
[5] P. Racher, K. Laplanche, D. Dhima, A. Bouchaïr, Thermo-mechanical analysis of the fire performance of dowelled timber connection, Eng. Struct. 32 (2010),
11481157.
[6] K. Laplanche, Comportement au feu des assemblages bois – assemblages Bois/bois et Bois/métal en double cisaillement. Rapport d’étude DSSF-ISTA/KL/MT/
ER0200511, CSTB 40 (2003).
[7] N. Ayme, D. Joyeux, Comportement au feu d’assemblages bois. Assemblages boismétal en double cisaillement, CTICM, 2003, p. 62. Rapport SRI-0.3/121-NA/PB
n◦ 030156.
[8] P.H. Lau, Fire resistance of connections in laminated vaneer lumber (LVL), University of Canterbury, 2006. Master Thesis.
[9] T.C.B. Chuo, Fire performance of connections in laminated vaneer lumber, University of Canterbury, 2007. Master Thesis.
[10] L. Peng, Performance of Heavy Timber Connections in Fire, Ottawa-Carleton Institute of Civil and Environmental Engineering, Canada, 2010. PhD Thesis.
[11] M. Audebert, D. Dhima, M. Taazount, A. Bouchaïra, Numerical investigations on the thermo-mechanical behavior of steel-to-timber connections exposed to fire,
Eng. Struct. 33 (12) (2011) 3257–3268.
[12] J. Zhang, Q. Xu, Y. Bai, J. Cai, Fire resistance of bolted steel-glulam-steel connections, J. S. China Univ. Technol. 43 (2) (2015) 58–65 (in Chinese).
[13] European Committee for Standardization, EN 1995-l-2: 2004, Eurode5—Design of Timber Structures Part1-2: General—Structures Fire Design, 2004. Brussels.

13
J. Luo et al. Journal of Building Engineering 46 (2022) 103804

[14] ISO 834-1:1999, Fire-resistance Tests – Elements of Building Construction – Part, vol. 1, General requirements, 1999.
[15] M. Audebert, D. Dhima, M. Taazount, A. Bouchaïr, Experimental and numerical analysis of timber connections in tension perpendicular to grain, Fire Saf. J. 63
(2014) 125–137.
[16] D. Dhima, M. Audebert, A. Bouchaïr, M. Taazount, Analysis of the thermomechanical behaviour of steel-to-timber connections in bending, J. Struct. Fire Engi. 5
(2) (2014) 97–111.
[17] P. Palma, A. Frangi, Modelling the fire resistance of steel-to-timber dowelled connections loaded perpendicularly to the grain, Fire Saf. J. 107 (2019) 54–74.
[18] P. Palma, A. Frangi, E. Hugi, P. Cachim, H. Cruz, Fire resistance tests on timber beam-to-column shear connections, J. Struct.l Fire Eng. 7 (1) (2016) 41–57.
[19] K. Laplanche, D. Dhima, P. Racher, Predicting the behaviour of dowelled connection in fire: fire results and heat transfer modelling, in: 8th World Conference of
Timber Engineering, Finland, 2004.
[20] Z. Shu, Z. Li, X. Yu, J. Zhang, M. He, Rotational performance of glulam dowel type bolted connections: experimental investigation and analytical approach,
Construct. Build. Mater. 213 (2019) 675–695.
[21] M. Audebert, D. Dhima, A. Bouchaïrc, Proposal for a new formula to predict the fire resistance of timber connections, Eng. Struct. 204 (2020) 110041.
[22] ASTM, Standard Test Methods for Small Clear Specimens of Timber (D143-14)[S], American Society of Mechanical Engineers, West Conshohocken, PA, 2014.
[23] Standards Press of China, Technical Code of Glued Laminated Timber Structures, GB/T50708-2012, Beijing, 2017 (in Chinese).
[24] Standards Press of China, Mechanical Properties of Fasteners Bolts, Screws and Studs, GB/T 3098.1-2010, Beijing, 2010 (in Chinese).
[25] A. Frangi, P. Palma, E. Hugi, P. Cachim, H. Cruz, Fire Resistance Tests on Beam-To-Column Shear Connections, 8th International Conference on Structures in
Fire, China, 2014.
[26] F. Lam, M. Gehloff, M. Closen, Moment-resisting bolted timber connections, Structures and Buildings 163 (2010) 167.
[27] M. Gehloff, M. Closen, F. Lam, Reduced edge distances in bolted timber moment connections with perpendicular to grain reinforcements, in: Proceedings of the
11th World Conference on Timber Engineering, 2010. Italy.
[28] Standards Press of China, Standard for Design of Steel Structures, GB50017–2017, Beijing, 2017 (in Chinese).
[29] FEMA 461-2007 Interim Testing Protocols for Determining the Seismic Performance Characteristics of Structural and Nonstructural Components..
[30] Dassault Systemes, F.E.A. Abaqus, Version 6, 13-2 [computer program], Dassault Systemes, 2013.
[31] J.P. Hong, D. Barrett, Three-dimensional finite-element modeling of nailed connections in wood, J. Struct. Eng. 136 (6) (2010) 715–722.
[32] J.P. Hong, J.D. Barrett, F. Lam, Three-dimensional finite element analysis of the Japanese traditional post-and-beam connection, J. Wood Sci. 57 (2) (2011)
119–125.
[33] J.P. Hong, Three-dimensional Nonlinear Finite Element Model for Single and Multiple Dowel-type Wood Connections, Canada: University of British Columbia,
2007.
[34] X. Sun, M. He, Z. Li, Novel engineered wood and bamboo composites for structural applications: state-of-art of manufacturing technology and mechanical
performance evaluation, Construct. Build. Mater. 249 (2020), 118751.
[35] J. Zhang, M. He, Z. Li, Numerical analysis on tensile performance of bolted glulam joints with initial local cracks, J. Wood Sci. 64 (2018) 364–376.
[36] B.H. Xu, M. Taazount, A. Bouchaïr, P. Racher, Numerical 3D finite element modelling and experimental tests for dowel-type timber joints, Construct. Build.
Mater. 23 (9) (2009) 3043–3052.
[37] J. Zhang, M. He, Z. Li, Numerical analysis on tensile performance bolted glulam joints with initial local cracks, J. Wood Sci. 64 (2008) 364–376.
[38] Y. Zhao, Research on Moment Bearing Performance and Improvement Methods of Glulam Multiple Bolted Connections in Post-and-beam Timber Construction,
Doctoral thesis, Tongji University, Shanghai, China, 2016 (in Chinese).
[39] A.J.M. Jorissen, Double Shear Timber Connections with Dowel Type Fasteners, Delft University Press, Delft, 1998.
[40] C. McCarthy, M. McCarthy, Three-dimensional finite element analysis of single-bolt, single-lap composite bolted joints: Part II––effects of bolt-hole clearance,
Compos. Struct. 71 (2) (2005) 159–175.
[41] EN 1993-1-2: 2005, Eurocode 3. Design of Steel Structures. Part 1–2: General Rules — Structural Fire Design, 2005.
[42] R. Knudson, Performance of Structural Wood Members Exposed to Fire, Doctoral thesis, University of California, Berkeley, USA, 1973.
[43] B. Fredlund, Modeling of heat and mass transfer in timber structures during fire, Fire Saf. J. 20 (1993) 39–69.
[44] A. Frangi, Brandverhalten von Holz-Beton-Verbunddecken, Doctoral thesis, Institute of Structural engineering, Zurich, Switzerland, 2001. ETH.
[45] M. Janssens, Thermo-physical Properties for Wood Pyrolysis Models. Pacific Timber Engineering Conference, Gold Coast, Australia, 1994.
[46] M. Fragiacomo, A. Menis, P. Moss, A. Buchanan, Numerical and experimental evaluation of the temperature distribution within laminated veneer lumber (LVL)
exposed to fire, Journal of Structural Fire Engineering 1 (3) (2010) 145–159.
[47] P. Cachim, J.-M. Franssen, Comparison between the charring rate model and theconductive model of Eurocode 5, Fire Mater. 33 (2009) 129–143.

14

You might also like