You are on page 1of 14

Fire Safety Journal 121 (2021) 103295

Contents lists available at ScienceDirect

Fire Safety Journal


journal homepage: http://www.elsevier.com/locate/firesaf

Thermal analysis of steel decking concrete slabs in case of fire


Fabricio Bolina a, b, Bernardo Tutikian a, João Paulo C. Rodrigues b, *
a
University of Vale do Rio dos Sinos - UNISINOS, Brazil
b
University of Coimbra, Portugal

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

Keywords: The thermal behavior of composite steel-concrete slabs in fire is more or less known, however there are still a lot
Fire of different aspects in the fire design methods that need clarification. A comparison of the temperature distri­
Slab bution in the cross-section of steel decking concrete slabs subjected to fire is made through three different
Steel decking
procedures: (a) experimental, (b) numerical and (c) analytical methods. The experimental tests corresponding to
Concrete
Rebar
eight real-scale fire tests on slabs carried out by the authors. These tests have been used to calibrate numerical
Thermal insulation models for being used with the finite element software Abaqus. The analytical methods were the ones proposed
Experimental in Annex D of EN 1994-1-2. The analytical steel decking temperatures showed convergence with the experi­
Numerical mental and numerical ones. The same was not observed for the concrete, positive and negative rebar and thermal
Analytical insulation temperatures. A new analytical approach for assessing the negative rebar temperatures and new
factors for assessing the thermal insulation performance - alternative to EN 1994-1-2 - are proposed.

1. Introduction in composite steel and concrete slabs subjected to fire [12]. Most of the
experimental and numerical research have been conducted for assessing
Fire is one of the most severe events to which a structure can be the mechanical behavior, analyzing parameters such as the concrete
subjected during its life. In composite steel and concrete structures, as strength [13], mechanical loading [14], concrete-steel decking longi­
both structural materials are responsible for the mechanical resistance, tudinal shear [15], slab-beam interaction [16], steel decking geometry
they should ensure the structural integrity of the element in case of fire. [17] and use of additional steel reinforcement [18] in case of fire. The
In steel decking concrete slabs, the profiled steel decking acts as a methods for assessing the temperature distribution in the cross-section,
continuous positive reinforcement and eliminate the need of formwork in the simplified calculation methods, lack experimental validation [19,
and steel reinforcing bars (rebars) [1,2]. When these slabs are subjected 20]. Thus, new methods are necessary for assessing the temperature
to fire on their bottom, the steel of the decking degrades leading the slab distribution in the cross-sections of steel decking concrete slabs [12].
to behave as non-reinforced. In this sense, it is necessary to experi­ The major part of the numerical researches on composite steel and
mentally study the structural response of these slabs in fire [3,4and5]. concrete elements subjected to fire has used EN1994–1.2 thermal pa­
The results of some fire tests [6] allowed the development of analytical rameters. However, some standard parameters seem to be obsolete [21]
procedures such as the ones presented at EN 1994–1.2 [3]. These because they don’t encompass all temperature dependent thermophys­
methods are behind other international standards for fire design of ical mechanisms. Some authors state that these thermal parameters are
composite structures such as the ANSI/AISC 360-05 [7], BS 5950-8 [8], unrealistic [22,23] because they were obtained in small scale tests. They
AS/NZS 2327 [9], NBR 14323 [10]. stress the need of finding new thermal parameters on full scale tests
EN 1994-1-2 has simplified analytical procedures and reference ta­ [25]. Additionally, design criteria for composite slabs subjected to fire
bles for assessing the fire resistance of composite steel concrete slabs have not been revised for decades [21] and sorely need updating.
based on a temperature distribution. They are foreseen procedures for In this sense, different authors [2,12,22] have found discrepancies
assessing temperatures on the profiled steel decking, concrete and steel between the experimental and EN 1994–1.2 material’s thermal param­
rebars. Since temperature distribution is an important factor in fire eters. These discrepancies extended down to elementary parameters
design, this must be thoroughly investigated [11]. such as the concrete thermal diffusivity and conduction. In some cases,
There are a few experimental studies on the temperature distribution the temperatures in the slab surface, not directly exposed to fire, did not

* Corresponding author.
E-mail address: jpaulocr@dec.uc.pt (J.P.C. Rodrigues).

https://doi.org/10.1016/j.firesaf.2021.103295
Received 15 July 2020; Received in revised form 2 November 2020; Accepted 25 January 2021
Available online 1 February 2021
0379-7112/© 2021 Elsevier Ltd. All rights reserved.
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

exceed 100 ◦ C. As this temperature was related to the thermal diffusivity h1 = 60 ​ mm; h2 = 59 ​ mm; c = 15 mm; Ø(+) = 6.3 or 10 mm and Ø
and temperature distribution in the cross-section, the authors concluded (-) = 6.3 or 10 mm (Fig. 1). The experimental results were used to
that the standard parameters are oversimplified and far away from the calibrate the numerical models and then to compare with the analytical
reality [26]. It is known, that standard parameters are excessively con­ results. For negative rebars and thermal insulation temperature ana­
servative, when comparing experimental with numerical and analytical lyzes, more values for h1 and c were considered beyond the ones of the
results [27]. experimental tests (Table 1).
Lamon et al. [11], in 2001, performed a numerical research on the The justification for different h1 thickness, for negative rebars and
temperature distribution in the cross-section of steel decking concrete thermal insulation, is due to the fact that they are referred to different
slabs in fire. The numerical models were calibrated with results from the things.
Cardington’s fire tests. However, these authors didn’t consider the steel For negative rebar analysis, the tested values were h1 = 40, 50, 60
decking in the simulations because they consider that it would have a and 70 mm. The lowest h1 value adopted was 40 mm, because it is the
small contribution to the mechanical strength at high temperatures. This minimum allowed at EN 1994-1.2 [3] and the maximum was 70 mm,
assumption is controversial because it is known that the steel decking because for thicknesses higher than this the negative rebars won’t be
interferes in the temperature evolution of the slab’s cross-section. Pa­ longer affected by the high temperatures, regardless the different con­
rameters such as the steel emissivity and profile’s shadow effect have crete cover thicknesses considered (c = 10, 20 or 30 mm).
influence on the heating of the slab [12]. Nonetheless, the results of this For thermal insulation analysis, the h1 values adopted were indi­
study were considered important because new thermal conductivity rectly calculated from the minimum effective slab thickness heff required
parameters were proposed [28]. Others thermal parameters have been in Table D6 of EN 1994-1.2 [3]. Taking into account the steel decking
tested so far by Kodur, in 2014, Al-Sibahy, in 2012 and Othuman, in dimensions used in the experimental tests and using what is foreseen in
2011 [25,29,30]. Table D6 for heff. The following values for h1 = 30, 50, 70, 90, 120 and
The simplified calculation methods presented in Annex D of EN 145 mm were determined. In our opinion this leads to a standard in­
1994-1-2 [3] were obtained from numerical models calibrated with congruity, because it can be used a value of h1 lower than the minimum
experimental results presented by Both, in 1998 [6]. These numerical required at EN 1994-1.2 [3]. However, we decide to use these values for
simulations used thermal parameters different from the ones proposed at h1 in order to assess the EN 1994-1.2 [3] rules.
EN 1994-1-2 for numerical modeling. This situation constitutes a stan­ The geometries of the steel decking with h2 = 59 ​ mm and h2 =
dard inconsistency of EN 1994-1-2 and needs to be uniformed and 75 ​ mm are shown in Fig. 2 (a) and (b), respectively, while Fig. 2c shows
reviewed. the size and distribution of the embossments along the web of the steel
In this paper, a thermal analysis has been performed in steel decking decking.
concrete slabs subjected to fire. This analysis has been carried with
numerical models calibrated with results of eight fire resistance tests on 2.2. EN 1994–1.2 analytical procedures
real-scale steel decking slabs, carried out by the authors of this paper.
The slabs, in the fire tests, have been tested as close as possible to a real This section describes the procedure used in the simplified calcula­
structure. The results were then compared with analytical results ob­ tion method of annex D of EN 1994–1.2 [3] for temperature calculations.
tained with the simplified calculation procedures proposed in Annex D
of EN 1994-1.2 [3]. EN 1994-1-2 simplified calculation models admit 2.2.1. Temperature in the concrete
the negative rebars fully thermally protected, which is not the case in all The temperature in the concrete is calculated according to item D.3
situations. This research studied this problem and commented its in­ of EN 1994–1.2 [3] for composite slabs. The standard procedures
fluence on the behavior of steel decking concrete slabs in case of fire. defined a critical isotherm temperature whose location was calculated
by four points with coordinates XI , XII , XIII , XIV and YI , YII , YIII , YIV ,
2. Methods for thermal analysis respectively. Portions of the slab, above the critical isotherm, were
considered not affected by high temperatures. The critical temperature
2.1. Cross-section of the steel decking concrete slabs is given by Eq. (1):

Fig. 1 shows the cross-section of the steel decking concrete slab θlim = d0 + d1 .Ns + d2 .
A 1
+ d3 .Φ + d4 . [◦ C] (1)
tested in this research. This geometry was the same in all prototypes Lr l3
used in the experimental, numerical and analytical tests. For tempera­
ture measurements, two alignments have been considered. Section A, where the factors are described in EN 1994–1.2 [3].
that is the cross-section alignment corresponding to the top flange of the
steel decking and section B corresponding to the bottom flange align­ 2.2.2. Temperature of the profiled steel decking
ment. For thermal insulation analysis of the slab, temperature measuring The analytical steel decking temperature was calculated according to
points in the slab surface were considered. item D.2 of EN 1994-1-2 [3] as shown in Eq. (2):
The cross-section dimensions in the experimental tests were: 1 A
θi = b0 + b1 . + b2 . + b3 .Ø + b4 .Ø2 [◦ C] (2)
b2 Lr

where b0 , b1 , b2 , b3 and b4 are tabulated coefficients defined for 60, 90


and 120 min of the ISO 834 [5] temperatures. The parameter Φ is a
configuration factor of the top flange of the profiled steel decking
calculated with the Hottel’s law.

Table 1
Cross-section parameters for standard and numerical temperature analysis:
negative rebars and thermal insulation of concrete slab.
Cross-section part Variables analyzed (mm)
Fig. 1. Cross-section for the standard, numerical and experimental temperature Negative rebars h1 = 40, 50, 60 ​ and ​ 70; h2 = 59 and 75; c = 10, 20 and 30
analysis: positive and negative rebars, concrete, steel decking and thermal Thermal insulation h1 = 30, 50, 70, 90, 120 ​ and ​ 145
insulation measuring points.

2
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 2. Profiled steel decking geometries for (a) h2 = 59 ​ mm, (b) h2 = 75 ​ mm and (c) embossment distribution along the steel decking web (dimensions in mm).

2.2.3. Temperature of the positive rebars The gravel used, is a dacite with plagioclase and quartz based, and had
The temperature of the positive rebars was calculated according to maximum dimensions of 9.5 mm and 19 mm. A natural quartz and an
Annex D of EN 1994-1-2 [3] as shown in Eq. (3): industrialized dacite sand were used. The concrete mix is presented in
( ) Table 3.
uf3 A 1
θs = c0 + c1 . + c2 .z + c3 . + c4 .α + c5 . [◦ C] (3) Steel rebars had a tensile strength fy = 500 MPa, used in Brazil and
hf Lr b2
classified as CA50. The rebars diameters were 6.3 mm and 10 mm. An
anti-cracking steel mesh, with cross-sectional area of 92 mm2/m, was
where z is relative to the rebar position in the cross-section, u1 , u2 and u3
used in the slabs. The tensile strength of the steel mesh was fy = 600 MPa
are the shortest geometric distances from the rebar axis to the steel
(classified as CA-60), and had 4.2 mm diameter wires welded in screens
decking and c0 , c1 , c2 , c3 , c4 and c5 are coefficients defined in Annex D of
measuring 150 mm × 150 mm. The concrete cover of the steel rein­
EN1994–1.2 [3]. The modulus of A/Lr is the ratio between the
forcement was 15 mm in all cases (positive and negative rebars and anti-
cross-sectional area of the groove of concrete around the rebar and its
cracking mesh).
surface area, whose calculation procedure is also presented in
The profiled steel decking had 0.80 mm thick, 59.0 mm profile
EN1994–1.2 [3].
height and cross-sectional area of 1137.64 mm2/m. The steel used was
an ASTM A653 with a tensile strength fy = 280 MPa. The steel decking
2.2.4. Temperature of the negative rebars
had a weight of 9.14 kg/m2 and an effective width of 915 mm.
Section D.3 of EN 1994-1-2 [3] supposes that the negative rebars can
be considered as at ambient temperature. The logic behind this is that,
2.3.2. Slab prototype characteristics
based on the critical temperature, depending of the isotherm co­
Table 4 presents the cross-sections of the steel decking concrete slabs
ordinates Xn and Yn , portions of the slab above the isotherm (which
tested. The differences between the S1–S8 slabs are the diameter of
include the negative reinforcement) could be assumed to be unaffected
positive and negative rebars. The positive rebars were along the length
by the temperature. Therefore, this study focused in determining by
of the slab while the negative rebars were over the intermediate beam
numerical and experimental procedures the temperature of the negative
only (Fig. 4b). Fig. 3 (a) shows the fabrication of the slabs. Slabs S6 up to
rebars for different positions and slab geometries (Table 1). In these
S8 were identical and used as reference tests, enabling repeatability of
procedures, the critical temperature was limited to 500 ◦ C in accordance
results.
to fib Bulletin n◦ 38 [31], temperature at which the mechanical strength
The serviceability loading on the slabs, trying to simulate a real sit­
of the steel is practically null.
uation. The value of the loading was 1.60 kN/m2 and materialized by
weight steel plates uniformly distributed over the slab (Fig. 3 b). This
2.2.5. Thermal insulation of concrete slab
loading was calculated according to the procedures of EN 1991-1.1 [24]
Thermal insulation is defined in Item D.4 of EN1994–1.2 [3] by the
and EN 1994-1.2 [3] for residential buildings. The loading had some
minimum effective slab thickness heff (Table 2), where h3 is the thickness
importance on the thermal behavior of the slab because it induced
of the mortar sheeting (if it exists). Table 2 correlates the heff values with
concrete cracking in some regions and avoid in others, thus affecting the
the fire resistance time and served as a reference in comparison to the
evolution of the temperatures in the cross-section of the slab.
values numerically obtained.
A high-stiffness auxiliary concrete frame, with an intermediate
beam, fire protected with ceramic wool, was used to create two slab
2.3. Experimental tests panels of 2300 mm each (Fig. 3 b). The intermediate beam tried to
simulate a slab with structural continuity. The hogging moments
2.3.1. Materials generated in the intermediate region of the slab changed the stress
The cement used was of high-initial resistance type that contained distribution in the concrete generating cracking and thus interfered in
fewer chemical additions. It is a Portland cement used in Brazil, classi­ the heating of the slab.
fied as CP-V ARI. No additional chemical additives were incorporated. All slabs were tested with 540 days of age. These experimental tests
were carried out at the Fire Safety Laboratory of University of Vale do
Table 2
Minimum effective slab thickness.
Table 3
Fire resistance time (min) heff (mm) Concrete mix. The average compression strength of the concrete at 28 and 540
days was 39.6 MPa and 42.4 MPa, respectively.
30 60 - h3
60 80 - h3 Material Mixing in weight (kg/m3) Mixing ratio

90 100 - h3 Cement 440 1.00


120 120 - h3 9.5 mm Gravel 400 0.91
19 mm Gravel 320 0.73
180 150 - h3
Industrialized Sand 320 0.73
240 175 - h3 Natural Sand 960 2.18
Water 200 ± 20 0.45 ± 0.05
Source: EN1994–1.2 [3].

3
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Table 4 Rio dos Sinos (UNISINOS), in Rio Grande do Sul, Brazil.


Characteristics of the steel decking slabs. Fig. 4a presents details of the test assembly while Fig. 4b shows
Specimen Positive Negative Cross-section structural drawings of prototypes S6 to S8, which are similar to all other
rebars rebars slabs (S1–S5).
diameter diameter Fig. 5 (a) shows a view of the horizontal furnace with the eight gas
(mm) (mm)
burners working. The heating curve followed the ISO 834 [5] standard
S1 No positive No negative fire curve [5]. Fig. 5b shows the specimen ready for testing.
rebars rebars Along the slab cross-section, thermocouples were installed at specific
S2 No negative
Ø6.3
heights intervals (concrete layers): 3 sensors for section A alignment
rebars
(called 2.5A, 5A and 6A, located at 25, 50 and 60 mm from the top
S3 Ø10 No negative
rebars surface, respectively) and 5 sensors for section B alignment (called 2.5B,
S4 No positive Ø6.3 5B, 6B, 8B and 11B, located at 25, 50, 60, 80 and 110 mm from the top
rebars surface, respectively) (Fig. 6a). For steel rebars, 2 sensors for each
S5 No positive Ø10.0 (positive and/or negative) were used. The rebars, steel decking and
rebars concrete layer temperatures were used to calculate the average tem­
S6 Ø6.3 Ø6.3
peratures in the slabs. In order to measure the dispersion of these
S7 Ø6.3 Ø6.3
average temperatures, a statistical analysis was performed. Thermo­
couples named “T" were used to measure the thermal insulation per­
S8 Ø6.3 Ø6.3 formance of the slab during the fire test (Fig. 6b). Thermocouples T were
installed in the slab surface not exposed to fire.
As shown in Fig. 6a, thermocouples T were arranged in Section A
thermal insulation point (i.e., the thinnest part of the slab cross-section,
Fig. 1). Fig. 6B shows that in panel 1, five thermocouples (T1-T5) were

Fig. 3. (a) Slab casting and (b) Assembly for fire resistance tests.

Fig. 4. (a) Testing assembly and (b) structural drawings of the slabs (dimensions in mm).

4
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 5. (a) Horizontal furnace and (b) Slab ready for testing.

Fig. 6. Experimental analysis: thermocouple locations (a) cross-section and (b) thermal insulation (slab surface).

uniformly distributed in Section A. In panel 2, one thermocouple T6 was


placed in the center of that area. The temperature used for the verifi­
cation of the thermal insulation was taken from thermocouples T1 to T5.
Thermocouple T6 was used only for checking porpoises.
For verification of the thermal insulation of the slabs, BS EN 13381
[32] criteria were considered, where it is established that the average
temperature cannot reach 140 ◦ C or the maximum point temperature
cannot reach 180 ◦ C, in the unexposed surface of the slab.

2.4. Numerical simulations

The numerical simulations were performed with the Abaqus software


[33]. The thermal parameters used in these analyzes, in the different
parts of the slab cross-section, are indicated in Fig. 7. On the bottom
surface of the slab, the ISO 834 [5] fire curve was considered by thermal
convection (with heat transfer coefficient α = 25 W/m2.K) and radiation
(with thermal emissivity Ɛ = 0,40, according to Ref. [12]). On the top
surface of the slab, an ambient temperature of 25 ◦ C was considered by
convection (with a heat transfer coefficient α = 9 W/m2.K).
The profiled steel decking detached from the concrete in the initial
stages of the fire exposure (Fig. 8), as observed in the experimental tests. Fig. 7. Cross-section thermal parameters.
This air layer directly affected the isotherms in the slab cross-section and
was considered in the numerical model with a thickness of 15 mm. of the tested slabs.
The steel decking was modeled with a 4-node quadrilateral shell The thermal conductivity and specific heat of the steel rebars were
(DS4), the reinforcement with a truss of 2-node link (DCC1D2) and the taken from EN 1993-1.2 [35]. Rebars and steel decking were considered
concrete with a general 3D solid 8-node linear isoparametric element with a density of 7850 kg/m3. The thermal conductivity of the steel of
(DC3D8) from the Abaqus software library. A mesh analysis was per­ the decking was taken from Craveiro et al. [1]. The specific heat of
formed and the mesh is rather refined, the size of the elements is conventional steel structures of EN 1993-1-2 [35] was used in the steel
approximately 0.5 × 0.5 × 0.5 mm. decking. Air thermal properties (i.e., specific heat, thermal conductivity
The values of the thermal conductivity and specific weight of the and specific weight) were considered according to Çengel and Ghajar
concrete with the temperature were extracted from EN 1992–1.2 [34]. [36] criterion.
The specific heat was also taken from EN 1992–1.2 for a concrete with In the numerical model, an effective heat flux (hnet ) was determined
3% humidity. The numerical models of this research used these concrete based on the convective (hnet,c ) and radiative (hnet,r ) heat fluxes. Other
parameters and showed good agreement with the experimental results

5
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 10. Numerical and experimental temperatures for section B.

Fig. 11 shows that the negative rebars were more protected from
heating than the positive ones, with a difference reaching 300 ◦ C. Figs. 9
Fig. 8. Detachment of the steel decking during test.
and 10 also show that, at the same concrete depths, the temperatures in
section A were always higher than the ones in section B, due to its
constants were defined, as the absolute zero temperature (− 273.15 ◦ C),
smaller thickness. The differences in the temperatures between the
Stefan-Boltzmann constant (σ = 5.67 × 10− 8 W/m2K4 ), gas temperature experimental and numerical values, may result from the variability of
(θg , obtained from the ISO 834), initial surface temperature (θm = 25 ◦ C) the experimental values. In the experimental tests there are phenomena
and initial radiation temperature of the surface (θr = 25 ◦ C), configu­ linked with the concrete cracking formation, mass and heat transfer and
ration factor (φ = 1.0), convection heat transfer coefficient (αc = 25 W/ chemical changes in the concrete, that are not considered in the nu­
m2K), steel decking surface emissivity (εm = 0.40) and fire emissivity merical simulations.
(εf = 1.00). Figs. 9–12 also demonstrate that there was agreement between the
experimental and numerical temperatures for thermal insulation and
3. Results and discussion thus the Eurocodes [3,34,35] thermal parameters of the materials were
adequate for numerical modeling of this type of slabs.
In this section they are first presented the calibration of the numer­ As the research that gave rise to the simplified method of Annex D of
ical models with the experimental results. Then, the experimental and EN 1994–1.2 [3] – i.e. Both [6] research – does not converge with these
numerical results were compared with those obtained by Annex D of EN material parameters, the need for correction and adjustment of EN
1994-1-2 [3] simplified calculation methods. 1994–1.2 [3] simplified calculation method is necessary. These adjust­
ments can be done, for example, in the coordinates for the definition of
the limit temperature isotherm presented in Annex D of EN 1994–1.2
3.1. Comparison between experimental and numerical temperatures [3], that is based in the temperature distribution in the cross-section of
the slab.
It can be observed in Figs. 9 and 10 that the temperatures of the steel Table 5 presents the experimental concrete layer average tempera­
decking (6A for section A and 11B for section B) were always lower than tures (AT) (as described in Fig. 6a), among all the slabs experimentally
the ISO 834 [5] fire curve temperatures. This was mainly a result of the tested (S1–S8). The standard deviation (SD) of each average tempera­
steel decking emissivity and the sink effect of the slab as a whole. ture is also presented. At the beginning of the test (0 min), the SD is of
the order of 5 ◦ C, which represents a variation around 20% of the AT. At
180 min, the SD is around 100 ◦ C, which represents a variation of the
order of 15% of the AT. Given the heterogeneity of the concrete and
randomness of the cracks, it is understood that this SD is acceptable. So,
these AT of the concrete layers were used to calibrate the numerical
models.
Table 6 shows the equivalent values of Table 5 but for the positive
and negative rebars among all tested slabs (S1–S8). As the steel has a
more uniform AT, it is noted that SD is lower throughout the fire test
time. The maximum SD was 16.6%, but generally the values were below
10%. The temperature of the rebars with a 10 mm diameter in section A
were not measured.
Table 7 shows the equivalent values but for the steel decking. The
values are only presented for the top and bottom flanges of the steel
decking because the temperatures of the web were not measured. In this
case, the SD is also low due to the homogeneity of the AT of the steel
throughout the fire test. The SD values did not exceed 10%. The SD
values demonstrate that the calibration of numerical models with AT
temperatures is so acceptable.
Fig. 9. Numerical and experimental temperatures for section A.

6
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 11. Numerical and experimental temperatures for rebars.

analytical temperature was 31.8% higher than the experimental.


Comparing the analytical and numerical temperatures, the largest
difference occurred also in the bottom flange at the same time, where
the analytical was 14.2% higher than the numerical temperatures. The
maximum difference between the experimental and numerical temper­
atures was 8%, at 60 min, while the minimum was 1.3%, at 120 min.
This agreement validated the EN 1994-1-2 [3] and Craveiro et al. [1]
thermal parameters for numerical modeling of these elements. The steel
decking temperatures practiced in EN 1994-1-2 [3] analytical method
was higher than the numerical and experimental ones, showing the
conservatism of the standard procedures. The temperatures numerically
and experimentally obtained were on average 100 ◦ C lower than the
standard ones.

3.2.2. Temperatures of the concrete


Annex D of EN 1994–1.2 [3] allows to determine the coordinates of
the temperature limit isotherm. The coordinates for the concrete were
calculated for the tested slabs (Table 9). Based on these coordinates
reference isotherms were defined in the numerical simulations for 60, 90
and 120 min. Once the numerical model was calibrated with the
experimental results, and considering the convergence between both, it
Fig. 12. Numerical and experimental temperatures for thermal insulation of
“A" section of concrete slab (see Fig. 6).
is understood that the experimental model presents the same degree of
convergence than the numerical model, showed in relation to the stan­
dard values. The temperatures calculated for these coordinates are
3.2. Comparison between the EN 1994-1-2 and the experimental and
presented in Table 10.
numerical temperatures
The temperatures in the concrete were not uniform along the cross-
section. In almost all cases of Table 10, the temperatures of the nu­
3.2.1. Temperatures of the profiled steel decking
merical model, for the same coordinates, were higher than the EN 1994-
Table 8 presents a comparison of the profiled steel decking temper­
1.2 [3] ones. However, with the time increasing, the difference
atures calculated with the EN 1994-1.2 [3] and the experimentally and
decreased to the point that the average numerical temperatures were
numerically determined ones.
lower than the analytical ones, at 120 min. In qualitative terms, the
The results of Table 8 are also shown graphically in Figs. 13–16.
numerical temperatures were 15.9% higher at 60 min, 2.6% higher at
Figs. 13–15 show comparisons of the temperatures for the steel decking
90 min, and 2.3% lower at 120 min, than the analytical ones.
bottom flange, top flange and web, respectively for 60, 90 and 120 min
of the ISO 834 [5] fire curve. Fig. 16 shows the temperatures of the web,
3.2.3. Temperatures of the positive rebars
bottom and top flanges, developed during the exposure of the slab to the
Experimental and numerical positive rebar temperatures are
ISO 834 fire curve.
compared to the analytical ones calculated with Annex D of EN1994–1.2
The results show that both the experimental and numerical tem­
[3] procedures (Table 11). The calculations were done for rebars with
peratures are lower than the ones determined with the procedures of the
6.3 mm and 10 mm diameter. The results are presented in Table 11 and
simplified calculation method of Annex D of EN 1994-1-2 [3].
Figs. 17 and 18.
Comparing the analytical with the experimental results, the largest
Figs. 17 and 18 show that analytical temperatures calculated with EN
difference occurred in the bottom flange at 60 min, time where the

7
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Table 5
Average temperature (AT) of concrete layers of slabs S1 to S8 (experimental tests) and the respective standard deviation (SD).
Concrete point ISO 834 time (min)

0 30 60 90 120 150 180

2.5 A AT (◦ C) 26.6 91.1 180.6 315.3 438.1 491.7 561.9


SD (%) 18.8 16.9 16.1 10.5 15.1 15.6 14.7
4.0 A AT (◦ C) 38.9 110.4 248.4 399.8 513.1 592.3 646.0
SD (%) 17.9 22.0 20.0 17,1 16.8 16.1 15.1
5.0 A AT (◦ C) 38.7 241.5 428.6 580.0 691.4 756.4 787.1
SD (%) 13.9 9.1 9.5 9.6 8.7 9.9 10.3
6.0 A AT (◦ C) 39.4 398.3 644.7 759.1 846.6 913.6 941.1
SD (%) 15.7 9.7 6.9 6.4 9.7 9.2 12.1
2.5 B AT (◦ C) 27.9 64.6 115.4 171.9 280.3 369.7 420.0
SD (%) 19.3 21.0 10.4 14.1 16.4 16.0 11.8
5.0 B AT (◦ C) 24.0 83.5 151.3 257.2 355.3 436.9 502.9
SD (%) 22.9 14.1 16.3 15.6 18.2 16.9 18.8
6.0 B AT (◦ C) 34.1 100.0 196.2 341.7 472.3 549.8 652.6
SD (%) 15.8 5.2 28.0 22.2 18.8 16.1 16.1
8.0 B AT (◦ C) 27.9 118.9 353.1 502.4 612.9 725.0 802.8
SD (%) 19.7 15.4 17.1 12.1 11.7 10.3 12.4
11.0 B AT (◦ C) 23.1 379.2 642.8 819.0 935.1 990.5 1023.6
SD (%) 22.1 11.0 8.1 9.7 8.9 6.3 4.7

Table 6
Average temperature (AT) of positive (+) and negative (-) rebars of slabs S1 to S8 (experimental tests) and the respective standard deviation (SD).
Steel rebar ISO 834 time (min)

0 30 60 90 120 150 180

Ø6,3 ( + ) AT (◦ C) 25.0 196.8 487.4 684.9 820.1 915.8 977.2


SD (%) 2.2 2.5 2.4 5.9 7.7 11.4 9.6
Ø10 ( + ) AT (◦ C) 27.2 109.5 350.6 553.7 634.5 787.8 848.7
SD (%) 2.0 13.0 16.6 14.3 12.2 13.1 9.8
Ø6,3 (-) Section A AT (◦ C) 25.0 72.3 157.1 296.2 418.7 480.7 523.8
SD (%) 2.7 8.2 3.0 3.8 6.6 9.8 13.2
Ø6,3 (-) Section B AT (◦ C) 24.0 57.8 96.1 174.0 256.0 350.7 433.1
SD (%) 3.2 12.9 10.7 14.2 6.3 2.3 1.2
Ø10 (-) Section B AT (◦ C) 24.8 50.9 91.8 182.4 266.0 332.9 390.1
SD (%) 3.1 4.2 7.6 6.8 5.7 9.5 8.7

Table 7
Average temperature (AT) of the top and bottom flange of the steel decking of slabs S1 to S8 (experimental tests) and the respective standard deviation (SD).
Steel deck ISO 834 time (min)

0 30 60 90 120 150 180

Top Flange AT (◦ C) 39.4 398.3 644.7 759.1 846.6 913.6 941.1


SD (%) 15.7 9.7 6.9 6.4 9.7 9.2 12.1
Bottom Flange AT (◦ C) 23.1 379.2 645.8 819.0 935.1 990.5 1023.6
SD (%) 22.1 11.0 8.1 9.7 8.9 6.3 4.7

Table 8
Steel decking temperatures: EN1994–1.2, experimental and numerical results.
Time (min) Temperatures (◦ C)

Bottom Flange Web Top Flange

EN1994-1-2 Experimental Numerical EN1994-1-2 Numerical EN1994-1-2 Experimental Numerical

60 851.42 645.80 745.33 772.38 702.15 704.77 644.70 655.70


90 952.37 819.00 854.30 899.31 812.92 843.90 759.14 757.25
120 1011.55 935.10 925.01 969.95 888.48 924.20 846.60 833.41

1994–1.2 [3] procedures were always higher than both the experimental analytical, experimental and numerical temperatures are justified by the
and numerical ones. This was similar to what occurred in the steel origin of the analytical method presented in Annex D of EN 1994–1.2
decking but different with the concrete. In the 6.3 mm rebars, analytical [3]. As already mentioned, this analytical method resulted from thermal
temperatures were more than 40% higher than the experimental ones. In results of numerical calculations carried out by Both [6]. These models
the case of 10 mm rebars, analytical temperatures were more than were established with thermal parameters for the concrete and steel
29.9% higher than experimental ones. Comparing the analytical with different from those currently indicated in the standards. This means
the numerical temperatures, the maximum difference was 26.1% higher. that slabs numerically calculated with the material thermal parameters
In the case of the positive rebars, the differences between the of EN 1992-1.2 [34] and EN 1994-1.2 [3], may present temperatures

8
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 13. Temperatures in the bottom flange of the steel decking.

Fig. 16. Numerical temperatures of the steel decking.

Table 9
Temperature limit isotherm coordinates in the concrete for the tested cross-
sections.
Time (min) Coordinates (mm)

XI YI XII YII XIII YIII XIV YIV

60 0.0 10.0 25.5 10.0 54.4 59.0 104.5 66.5


90 0.0 14.5 22.8 14.5 51.1 59.0 104.5 69.4
120 0.0 17.8 20.8 17.8 48.8 59.0 104.5 71.5

Fig. 14. Temperatures in the top flange of the steel decking. Table 10
Comparison of concrete temperatures calculated according to EN1994–1.2
analytical methods and numerical tests.
Time (min) Temperature (◦ C)

EN 1994-1-2 Numerical tests

Coordinates

XI , YI XII , YII XIII , YIII XIV , YIV Average

60 (I) 510.6 626.0 662.3 510.0 569.8 592.0


90 (II) 635.3 697.7 735.0 575.0 600.8 652.1
120 (III) 719.0 764.7 785.4 633.2 627.0 702.6

that rebars would not reach 500 ◦ C, that is the critical temperature
proposed in fib Bulletin no 38 [31].
The results for a composite slab of h1 = 40 mm and h2 = 75 mm (see
Table 1) are shown in Fig. 23. It is observed that already for h1 = 40 mm,
that is the smallest value allowed by the Eurocode, the negative rebars in
section B could be considered as thermally protected. Thus, it was
Fig. 15. Temperatures in the web of the steel decking. considered unnecessary to perform more numerical simulations for
other values of h1 and h2 .
different from those calculated with the simplified method of Annex D of Figs. 19–23 shows that the temperatures of the negative rebars are
EN 1994–1.2 [3] procedures. affected by their position in the slab cross-section. Rebars placed above
the top flange (section A) showed higher temperatures than those placed
3.2.4. Temperatures of the negative rebars above the bottom flange (section B) of the steel decking. This temper­
EN1994–1.2 [3] does not provide a methodology for assessing the ature difference between the rebars positioned in section A and B was
temperatures of the negative rebars. In Figs. 19–22 are presented the around 300 ◦ C. It should be noted that this aspect of the negative rebar
results of the numerical temperatures in function of the ISO 834 fire location within the slab is not foreseen in Annex D of EN 1994–1.2 [3]
curve exposure for composite slabs with h1 = 40, 50, 60 and 70 mm, procedures.
respectively, all with h2 = 59 mm (see Table 1). For each case, it was In order to characterize the thermal exposure of the negative rebars
used concrete rebar covers c = 10, 20 and 30 mm. No concrete slab on steel decking concrete slabs in case of fire, an equation is here pro­
thicker than h1 = 70 mm was analyzed, since previous results indicated posed (Eq. (4)). In this equation the Cf factor makes it possible to

9
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Table 11
Comparison of temperatures of positive rebars defined according to EN1994–1.2 analytical methods, experimental and numerical tests.
Time (min) Temperatures (◦ C)

Ø6.3 mm Ø10 mm

EN1994-1-2 Experimental Numerical EN1994-1-2 Experimental Numerical

60 678.19 487.40 544.70 666.78 513.20 528.82


90 847.38 657.1 679.41 836.21 689.70 665.81
120 938.25 799.60 770.42 927.45 820.10 750.30

Fig. 17. Comparison for Ø6.3 mm positive rebar temperatures.

Fig. 19. Numerical temperatures of the negative rebars for h1 = 40 mm and h2


= 59 mm.

Fig. 18. Comparison for Ø10 mm positive rebar temperatures.

understand whether there is thermal protection of the negative rebars.


The Cf coefficient is based on geometrical factors shown in Fig. 24 and
the thickness of concrete cover c for rebars - measured from upper face of
the slab.
( )
( ) η+20
kf

Cf = 0.032.(h1 − c).η.kf .
h1
≥ 1.0; forh1 ≥ 40mm (4) Fig. 20. Numerical temperatures of negative rebars for h1 = 50 mm and h2 =
10 59 mm.

where: c = 30 mm.

2.0for60min ⎧
⎨ 0.90for sectionA A coefficient Cf greater or equal to 1.0 (Cf ≥ 1.0) mean that thermal



1.6for90min
kf = Negative rebars, if: effects on negative rebars can be disregarded in fire resistance verifi­
⎩ h2 for sectionB
η =
⎪ 1.3for120min

⎩ 26.2 cations. The temperature can be taken as at ambient temperature. For
1.0for180min
values less than 1.0 (Cf < 1.0), a specific thermal analysis must be
Cf ≥ 1.0: thermally protected Cf < 1.0: thermally unprotected.
conducted to obtain a temperature distribution within the cross-section,
h1 , h2 , c in mm, with. h2 = 59 or 75mm
since the effect of high temperatures on these rebars cannot be neglec­
Eq. (4) is valid for h1 ≥ 40 ​ mm, similar to the item D.4 of
ted.
EN1994–1.2 [3]. The correlation was validated for two widely
As an illustration, Eq. (4) was applied to three practical examples,
commercially available profiled steel decking with h2 = 59 mm and h2 =
presented in Table 12. For each case, different values for h1 , c, Fire
75 mm. The maximum value for the thickness of the concrete cover was

10
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 24. Geometric parameters for the factor.Cf

Resisting Rate (FRR) required and position of the rebars in the cross-
section (section A or B) were proposed for a slab with h2 = 59 mm.
The results show that the location of the negative reinforcement for
cases 1 and 3 meets the criterion and can be considered in structural fire
design as at ambient temperature (Cf ≥ 1.0). In case 2, the negative
reinforcement was poorly placed in the cross-section (Cf < 1.0), thus its
temperature has to be considered in structural fire design.
Fig. 21. Numerical temperatures of negative rebars for h1 = 60 mm and h2 = As validation of the use of coefficient Cf , Fig. 25 shows the isotherms
59 mm. of slabs with h2 = 59 mm. Isotherms for the case in which h2 = 75 mm
are presented in Fig. 26. All isotherms refer to the time of 180 min of the
ISO 834 [5] fire curve. In all cases, it can be seen that there is a
non-uniform temperature distribution. The thinner layers of the
cross-section of the slab, section A, have higher temperatures than in
section B, that was thicker. Therefore, the criterion used in the calcu­
lation of coefficient Cf and the cross-section reference for its application
(Fig. 24) is validated.

3.2.5. Thermal insulation of concrete slab


According to Annex D of EN1994–1.2 [3], the cross-section of Fig. 1
meets the standard thermal insulation requirements (average tempera­
ture less than 140 ◦ C or maximum temperature less than 180 ◦ C in the
unexposed face of the slab) for 75 min. If equation D1 of EN 1994-1.2 [3]
is used, thermal insulation “I” should be equal to 71 min. Numerical
simulations determined that the 180 ◦ C was reached at 69.8 min in
section A and at 99.8 min in section B (see Fig. 12).
According to EN 13501-2 [37] criteria and EN1994–1.2 [3] pre­
mises, the comparison with the standard analytical procedure of
EN1994–1.2 and the numerical results, is presented in Table 13. The
effective thickness of the slab was calculated with Eq. D.15 of Annex D of
EN1994–1.2 [3]. For this study, the procedure of calculation is shown in
Fig. 22. Numerical temperatures of negative rebars for h1 = 70 mm and h2 = Eq. (5) with the effective steel decking height (heff ) equal to h2 = 59 mm
59 mm. and l1 = 126 mm, l2 = 84 mm and l3 = 84 mm related to the steel
decking geometry established according to EN 1994-1-2 [3].
( )
h2 l1 + l2
heff = h1 + . ≅ h1 + 30mm (5)
2 l1 + l3
Table 13 suggests revisions to EN1994–1.2 [3] in what concerns the
thermal insulation criterion. There is a divergence of results for

Table 12
Practical examples of coefficient Cf applications.
Case 1: Reinforcement in section A
Cross-section: h1 = 40 ​ mm; h2 = 59 ​ mm; c = 10 mm; with: FRR = 120 min; kf =
1.3; η = 0.9
Cf = 1.22
Case 2: Reinforcement in section A
Cross-section: h1 = 50 ​ mm; h2 = 59 ​ mm; c = 30 mm; with: FRR = 120 min; kf =
1.3; η = 0.9
Cf = 0.83
Case 3: Reinforcement in section B
Cross-section: h1 = 60 ​ mm; h2 = 59 ​ mm; c = 30 mm; with: FRR = 180 min; kf =
Fig. 23. Temperatures of negative rebars for h1 =40 mm and h2 = 75 mm. 1.0; η = 2.25
Cf = 2.34

11
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

Fig. 25. Isotherms for h1 = (a) 40; (b) 50; (c) 60 and (d) 70 mm (h2 = 59mm) (temperatures in ◦ C).

Fig. 26. Isotherms for h1 = 40 mm and h2 = 75 ​ mm (temperatures in ◦ C).

Table 13 Table 14
Thermal insulation: standard and numerical comparison. Proposed thermal insulation for the steel decking concrete slab.
Effective Thickness of concrete Thermal insulation time (min) FRR criteria (min) Minimum effective thickness: heff [mm]
thickness: above h1 (heff −
EN Numerical I 20 60 - h3
heff [mm] 30mm)
1994-1-
I 30 70 - h3
2
I 45 80 - h3
60 60–30 = 30 mm I 30 I 20 (max. temp. criterion
I 60 90 - h3
reached at 24.8 min)
80 80–30 = 50 mm I 60 I 45 (average temp. I 60 100 - h3
criterion reached at 50.5 I 90 110 - h3
min) I 120 120 - h3
100 100–30 = 70 mm I 90 I 60 (average temp. I 120 130 - h3
criterion reached at 83.6
I 150 140 - h3
min)
120 120–30 = 90 mm I 120 I 120 (average temp. I 180 150 - h3
criterion reached at I 180 160 - h3
121.5 min) I 240 175 - h3
150 150–30 = 120 mm I 180 I 180 (average temp.
criterion reached at
191.0 min)
175 175–30 = 145 mm I 240 I 240 (average temp. Table 14 was elaborated with a limiting average temperature of
criterion reached at 140 ◦ C or a maximum temperature of 180 ◦ C between two points (sec­
263.0 min) tion A and B thermal insulation points shown in Fig. 1). Ambient tem­
perature was taken as 25 ◦ C for all cases, because it was the average
ambient temperature on the day of the experimental tests. The results
composite slabs with effective height heff equal or less than 100 mm.
are shown in Figs. 27 and 28 for the average and maximum tempera­
Table 14 suggests a new complement for thermal insulation based on
tures, respectively.
an effective thickness of the composite slab, adjusted for a mortar
In terms of thermal insulation “I”, Figs. 27 and 28 shows the
sheeting (if it exists), given by h3 .

12
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

used in the Both [6] research, that is in the origin of the Eurocode’s
method;
• The temperatures of the negative rebars are influenced by their
location in the slab’s cross section. The concrete’s cover and rebar’s
position have influence on the rebar’s temperatures. If h1 <70 mm
the negative rebars are not properly thermal protected by the con­
crete and their temperature can easily reach 500 ◦ C, especially in
section A. This section corresponds to the smallest thickness of the
slab and obviously the temperatures will be higher than in other
sections. EN 1994-1.2 [3] does not take yet this into account in the
analytical methods;
• Therefore, a Cf coefficient has been defined. This coefficient de­
termines the ideal locations and composite slab thickness for thermal
protection of negative rebars;
• Analytical calculations of thermal insulation for steel decking slabs
with h2 = 59 mm were divergent from the experimental and nu­
merical results. This indicated the need of reviewing Table D6 of EN
1994-1.2 [3] for effective thicknesses less than 100 mm. A new table
Fig. 27. Average temperatures in the thermal insulation points in sections A is proposed, which takes into account the mentioned differences;
and B. • Amendments to the actual EN 1994-1-2 proposals should be done.
Thus, they should be considered a reduction of 150 ◦ C in the positive
rebar temperatures; maintain the actual applied concrete tempera­
tures for 90 and 120 min but increase in 80 ◦ C the temperature of the
60 min reference isotherm, use a Cf coefficient in the negative rebar
analysis and replace table D6 of EN 1994-1.2 [3] by Table 14 of the
present paper. On the other hand, the Eurocode’s temperatures for
the profiled steel decking were well matched and do not need
amendments.

Author statement

Fabrício Bolina: Conceptualization, Methodology, Validation,


Formal analysis, Investigation, Resources, Data curation, Writing –
original draft, Visualization, Bernardo Tutikian: Project administration,
Funding acquisition, Supervision, João Paulo C. Rodrigues: Conceptu­
alization, Methodology, Validation, Investigation, Writing – review &
editing, Supervision

Declaration of competing interest


Fig. 28. Maximum temperatures in the thermal insulation points in sections A
and B. 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.
influence of the location where temperatures are measured in the slab
surface. The part of the cross-section of the slab of higher thickness,
section B, has higher thermal insulation capacity when compared to Acknowledgements
section A. The thermal insulation is not uniform throughout the slab
surface as expected. For this requirement, working with an effective slab The authors would like to thank to Perfilor ArcellorMital Brazil, São
thickness for analytical calculation, as currently EN 1994-1.2 [3] admits, Paulo, for donating the profiled steel sheets for the test specimens, to
seems to be a good solution. Molder Estruturas, Garibaldi, Brazil, for their support in the construction
of the test specimens, to CKC Brazil, São Paulo, for offering consumables
to the tests and to the staff of ITT Performance Fire Safety Laboratory,
4. Conclusions
São Leopoldo, Brazil, for helping in the tests.
The following remarks and conclusions from this research may be
Appendix A. Supplementary data
drawn:

Supplementary data to this article can be found online at https://doi.


• A steel decking detachment was observed very early in the fire
org/10.1016/j.firesaf.2021.103295.
resistance tests (initial 5 min of exposure to ISO 834 [5] fire curve).
This indicates that the composite effect in the slab was not present all
over the fire test; References
• Steel decking temperatures determined analytically by EN 1994-1.2
[1] H.D. Craveiro, J.P.C. Rodrigues, A. Santiago, L. Laim, Review of the high
[3] showed convergence with the experimental and numerical ones; temperature mechanical and thermal properties of the steels used in cold formed
• Positive rebar temperatures analytically calculated with the EN steel structures - the case of the S280GD+Z steel, Thin-Walled Struct. 98 (2016)
1994-1.2 [3] procedures were higher than the experimental and 154–168, https://doi.org/10.1016/j.tws.2015.06.002.
[2] D. Pantousa, E. Mistakidis, Advanced modeling of composite slabs with thin-walled
numerical ones. This is possibly justified by the thermal parameters steel sheeting submitted to fire, Fire Technol. 49 (2) (2013) 293–327, https://doi.
proposed in the Eurocode’s method that are different from the ones org/10.1007/s10694-012-0265-x.

13
F. Bolina et al. Fire Safety Journal 121 (2021) 103295

[3] EN 1994-1-2, Design of Composite Steel and Concrete Structures. Part 1.2 - Design [20] T. Gernay, N.E. Khorasani, Recommendations for performance-based fire design of
Rules, Structural Fire Design, European Committee for Standardization, Brussels, composite steel buildings using computational analysis, J. Constr. Steel Res. 166
2011. (2020), 105906, https://doi.org/10.1016/j.jcsr.2019.105906.
[4] EN 1363-1, Fire Resistance Tests -Part 1: General Requirements, European [21] J. Jiang, A. Pintar, J.M. Weigand, J.A. Main, F. Sadek, Improved calculation
Committee for standardization, Brussels, 2015. method for insulation-based fire resistance of composite slabs, Fire Saf. J. 105
[5] ISO 834-8, Fire Resistance Tests - Elements of Building Construction, International (2019) 144–153, https://doi.org/10.1016/j.firesaf.2019.02.013.
Organization for Standardization, Geneva, 1999. [22] M. Achenbach, T. Lahmer, G. Morgenthal, Identification of the thermal properties
[6] C. Both, The Fire Resistance of Composite Steel-Concrete Slabs, Thesis, Delft of concrete for the temperature calculation of concrete slabs and columns subjected
University of Technology, 1998. to a standard fire — methodology and proposal for simplified formulations, Fire
[7] ANSI/AISC 360-05, Specification for Structural Steel Building, American Institute Saf. J. 87 (2017) 80–86, https://doi.org/10.1016/j.firesaf.2016.12.003.
of Steel Construction, Chicago, 2005. [23] Z. Xing, A.-L. Beaucour, R. Hebert, A. Noumowe, B. Ledesert, Influence of the
[8] BS 5950-8, Structural Use of Steelwork in Building, Part 8: Code of Practice for Fire nature of aggregates on the behavior of concrete subjected to elevated
Resistant Design, British Standard, London, 2003. temperature, Cement Concr. Res. 41 (4) (2011) 392–402, https://doi.org/
[9] AS/NZS 2327, Composite Steel-Concrete Construction for Buildings, Australia 10.1016/j.cemconres.2011.01.005.
Standard/New Zealand Standard, Sydney/Wellington, 2017. [24] EN 1994-1-1, Design of Composite Steel and Concrete Structures. Part 1.1 –
[10] NBR 14323, Structural Fire Design of Steel and Composite Steel and Concrete General Rules and Rules for Buildings, European Committee for Standardization,
Structures for Buildings, Brazilian Standard Association, Rio de Janeiro, 2013 [in Brussels, 2011.
portuguese]. [25] V. Kodur, in: Properties of Concrete at Elevated Temperatures, ISRN Civil
[11] S. Lamont, A.S. Usmani, D.D. Drysdale, Heat transfer analysis of the composite slab Engineering, 2014, pp. 1–15, https://doi.org/10.1155/2014/468510.
in the Cardington frame fire tests, Fire Saf. J. 36 (8) (2001) 815–839, https://doi. [26] G.-Q. Li, N. Zhang, J. Jiang, Experimental investigation on thermal and mechanical
org/10.1016/S0379-7112(01)00041-8. behavior of composite floors exposed to standard fire, Fire Saf. J. 89 (2017) 63–76,
[12] J. Jiang, J.A. Main, J.M. Weigand, F.H. Sadek, Thermal performance of composite https://doi.org/10.1016/j.firesaf.2017.02.009.
slabs with profiled steel decking exposed to fire effects, Fire Saf. J. 95 (2018) [27] L. Lin, C. Wade, Experimental Fire Tests of Two-Way Concrete Slabs. Fire
25–41, https://doi.org/10.1016/j.firesaf.2017.10.003. Engineering Research Report 02/12, University of Canterbury, School of
[13] J.J. Del Coz-Diaz, J.E. Martinez-Martinez, M. Alonso-Martinez, F.P.A. Rabanal, Engineering, Christchurch, 2002.
Comparative study of Light-Weight and Normal Concrete composite slabs behavior [28] A.S. Usmani, J.M. Rotter, S. Lamont, A.M. Sanad, M. Gillie, Fundamental principles
under fire conditions, Eng. Struct. 207 (2020), 110196, https://doi.org/10.1016/j. of structural behavior under thermal effects, Fire Saf. J. 36 (8) (2001) 721–744,
engstruct.2020.110196. https://doi.org/10.1016/S0379-7112(01)00037-6.
[14] K.M.A. Hossain, S. Attard, M.S. Anwar, Finite element modelling of profiled steel [29] A. Al-Sibahy, R. Edwards, Thermal behavior of novel lightweight concrete at
deck composite slab system with engineered cementitious composite under ambient and elevated temperatures: experimental, modelling and parametric
monotonic loading, Eng. Struct. 186 (2019) 13–25, https://doi.org/10.1016/j. studies, Construct. Build. Mater. 31 (2012) 174–187, https://doi.org/10.1016/j.
engstruct.2019.02.008. conbuildmat.2011.12.096.
[15] S. Sharma, V.T. Vaddamani, A. Agarwal, Insulation effect of the concrete slab-steel [30] M.A. Othuman, Y.C. Wang, Elevated-temperature thermal properties of lightweight
deck interface in fire conditions and its influence on the structural fire behavior of foamed concrete, Construct. Build. Mater. 25 (2) (2011) 705–716, https://doi.org/
composite floor systems, Fire Saf. J. 105 (2019) 79–91, https://doi.org/10.1016/j. 10.1016/j.conbuildmat.2010.07.016.
firesaf.2019.02.006. [31] FIB Bulletin no 38, Fire Design of Concrete Structures - Materials, Structures and
[16] J.-S. Zhu, Y.-G. Wang, J.-B. Yan, X.-Y. Guo, Shear behavior of steel-UHPC Modelling. State-Of-Art Report, Fédération Internationale du Béton (FIB),
composite beams in waffle bridge deck, Compos. Struct. 234 (2020), 111678, Lausanne, 2007.
https://doi.org/10.1016/j.compstruct.2019.111678. [32] BS EN 13381-1, Test Methods for Determining the Contribution to the Fire
[17] I. Arrayago, E. Real, E. Mirambell, F. Marimon, M. Ferrer, Experimental study on Resistance of Structural Members: Horizontal Protective Membranes. London,
ferritic stainless steel trapezoidal decks for composite slabs in construction stage, 2014.
Thin-Walled Struct. 134 (2019) 255–267, https://doi.org/10.1016/j. [33] Abaqus, Abaqus Analysis User’s Guide, Simulia, 2016.
tws.2018.10.012. [34] EN 1992-1-2, Design of Concrete Structures. Part 1.2 - Design Rules, Structural Fire
[18] L.G.F. Grossi, C.F.R. Santos, M. Malite, Longitudinal shear strength prediction for Design, European Committee for Standardization, Brussels, 2010.
steel-concrete composite slabs with additional reinforcement bars, J. Constr. Steel [35] EN 1993-1-2, Design of Steel Structures. Part 1.2 - Design Rules, Structural Fire
Res. 166 (2020), 105908, https://doi.org/10.1016/j.jcsr.2019.105908. Design, European Committee for Standardization, Brussels, 2010.
[19] H. Rezaeian, G.C. Clifton, J.B.P. Lim, Failure modes for composite steel deck [36] Y.A. Çengel, A.J. Ghajar, Heat and Mass Transfer: Fundamentals & Applications,
diaphragms subjected to in-plane shear forces – a review, Eng. Fail. Anal. 107 McGraw-Hill Education, New York, 2015.
(2020), 104199, https://doi.org/10.1016/j.engfailanal.2019.104199. [37] EN 13501-2, Fire Classification of Construction Products and Building Elements
Classification Using Data from Fire Resistance Tests, Excluding Ventilation
Services, European Committee for Standardization, Brussels, Brussels, 2016.

14

You might also like