Sie sind auf Seite 1von 16

Thin-Walled Structures 135 (2019) 521–536

Contents lists available at ScienceDirect

Thin-Walled Structures
journal homepage: www.elsevier.com/locate/tws

Full length article

Simulation of fire resistance behaviour of pultruded GFRP columns T


a b,⁎ a,⁎
Tiago Morgado , Nuno Silvestre , João R. Correia
a
CERIS, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1, 1049-001 Lisboa, Portugal
b
IDMEC, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1, 1049-001 Lisboa, Portugal

ARTICLE INFO ABSTRACT

Keywords: This paper presents a numerical investigation on the fire resistance of pultruded GFRP columns with tubular
Glass fibre reinforced polymer (GFRP) cross-section, both unprotected and protected with a passive fire protection. Three-dimensional finite element
Fire behaviour models were developed and they considered the thermo-mechanical behaviour of GFRP material (temperature-
Numerical study dependent mechanical properties) and the temperature distributions previously obtained through heat transfer
Mechanical response
analyses and fluid dynamics inside the tube cavity. The numerical results presented include the time evolution of
Stress analysis
axial and flexural deformations of the GFRP columns, as well as the stress distributions in both longitudinal and
Initial failure
Damage transversal directions of the unprotected column under one-side fire exposure and axially compressed (desig-
nated as reference column). In comparison with this reference column, the paper focuses on the evaluation of the
several effects, such as the use of fire protection system, the imposition of different fire exposure conditions and
the application of distinct load levels. The Tsai-Hill criterion is used to identify the initial failure of GFRP
columns and assess the evolution of failure index with fire exposure time, while the Hashin criterion is used to
obtain an estimate of column strength and collapse mode. It is concluded that the proposed models are able to
qualitatively capture the general trend of the experimental results, despite the quantitative differences not yet
overcome. With the consideration of creep, delamination effects and fracture, the authors are confident that
these models will soon correctly predict the complex mechanical behaviour and the fire resistance of GFRP
columns.

1. Introduction decade, further experimental and numerical studies are still needed to
better understand the fire behaviour of pultruded GFRP structures [6].
Glass fibre reinforced polymer (GFRP) composites are being in- In particular, it is known that (i) the action of compressive stresses in
creasingly used in civil engineering applications as a result of their structural members may lead to undesired nonlinearities and buckling,
particular but valuable properties, compared to traditional materials and (ii) the thermo-mechanical behaviour of GFRP in compression is
[1–3]. GFRP material presents high strength, lightness and reduced poorer than in tension. These evidences make the current experimental
maintenance costs and, therefore, constitutes a valid alternative to and numerical research on the fire resistance of pultruded GFRP
traditional materials, such as reinforced concrete, steel and timber. members under compression (columns) of utmost relevance. The few
Despite the great potential of GFRP, it presents low elasticity modulus, works reported in the literature, either experimental or numerical,
brittle failure and poor fire performance. Presently, the main drawback about the structural response of pultruded GFRP columns subjected to
of using pultruded GFRP profiles in building applications is its sensi- elevated temperatures are summarized next.
tivity to temperature (dependency of strength and stiffness), even for In case of experimental studies, Wong et al. [7] evaluated the
moderate temperatures [4]. When exposed to elevated temperatures, compressive strength of short GFRP C-section columns exposed to ele-
the mechanical properties of GFRP are significantly reduced, mainly vated temperatures (20–250 °C). Similarly, Wong and Wang [8] per-
due to the thermal degradation of the polymeric resin. Initially, for formed compressive tests of pultruded GFRP columns with C-section for
moderate temperatures (~60–140 °C), GFRP softens, creeps and distorts several temperatures (20–120 °C) and different lengths
as a result of glass transition of the organic matrix [4]. Then, for ele- (500–1500 mm). In both studies, the specimens were heated up to the
vated temperatures (~300–500 °C), GFRP releases heat and smoke due target temperature and then loaded to failure – this procedure allowed
to the polymeric matrix decomposition [5]. the evaluation of the strength reduction, failure modes and instability
Although several investigations have been reported in the last phenomena. In order to investigate the feasibility of using a water-


Corresponding author.
E-mail addresses: joao.ramoa.correia@tecnico.ulisboa.pt (N. Silvestre), joao.ramoa.correia@tecnico.ulisboa.pt (J.R. Correia).

https://doi.org/10.1016/j.tws.2018.11.022
Received 15 March 2018; Received in revised form 12 September 2018; Accepted 19 November 2018
Available online 14 December 2018
0263-8231/ © 2018 Elsevier Ltd. All rights reserved.
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

cooling system, Bai and Keller [9] subjected short pultruded GFRP (GLT(T)) moduli. Heat transfer is considered through conduction, ra-
tubes to serviceability loads (6–12% of failure load at room tempera- diation and convection of air inside the tube using fluid dynamics [20].
ture) and then exposed them to high temperatures (up to 220 °C). They The numerical results presented include: (i) the evolution of axial and
concluded that the active fire protection system allowed maintaining flexural deformations of the GFRP columns (compared to the experi-
temperatures below glass transition temperature (Tg), thus preventing mental data); (ii) the variation with time of stress distributions in both
collapse even after 2 h of fire exposure. Bai et al. [10] also reported a longitudinal and transversal directions of the unprotected column
study on the fire behaviour (full-scale) of GFRP columns protected with under one-side fire exposure (reference column); (iii) the effects of fire
flowing water. They compressed long columns with multicellular cross- protection system; and (iv) the influence of different fire exposure
section, exposed them to the ISO 834 [11] fire curve, and concluded conditions (one- vs. three-sides) and load levels. Moreover, both Tsai-
that the water-cooling system provided a very effective solution to Hill and Hashin1 criteria are used to obtain a better identification of the
improve the fire performance of GFRP columns. most damaged zones of the columns and understand their failure
Regarding the numerical studies, and besides the works presented modes. Finally, the main conclusions of this numerical study are pre-
by Gibson et al. [12] and Bausano et al. [13] at the GFRP laminate level, sented.
not much has been done in the modelling and simulation of GFRP
profiles under fire exposure. In this scope, the works reported by Wong
et al. [7] and Bai et al. [10,14] deserve to be credited. Wong et al. [7] 2. Resume of experimental study
studied the compressive behaviour of pultruded GFRP columns with C-
section for different temperatures (20–250 °C) using finite element (FE) Before presenting the numerical model, a brief resume of the ex-
models developed within the framework of ABAQUS software. Al- perimental tests on pultruded GFRP columns under fire is reported
though their models included local buckling effects and the influence of herein [18]. The columns had square tubular cross-section
geometrical imperfections, they concluded that the modelling approach (100 × 100 ×8 mm), length L = 1.5 m, and were produced by Fiber-
overestimated the experimental failure loads. Bai et al. [10] also si- line using an isophthalic polyester resin matrix reinforced with alter-
mulated the fire behaviour of GFRP multicellular columns based on nating layers of unidirectional E-glass fibre rovings (located mostly in
one-dimensional thermal analysis and using their own models [14], in the centre of the tube walls) and perimetral (inner and outer) strand
which the variation of thermo-physical properties with time and tem- mats and surface veils (69% of inorganic content in weight). Table 1
perature was considered. The evolution of lateral deformation with summarizes the mechanical (elastic and ultimate) properties (tension,
time was determined by considering the second-order deflections due to compression, flexural and shear) of the profiles used, which were de-
both thermal expansion and load eccentricity. In addition, the variation termined by means of coupon tests at room temperature (RT). Shear
of Euler buckling load and ultimate load with time was estimated using modulus (GLT) at RT was set as 3.6 GPa, based on Iosipescu tests [21].
second-order bending theory. The authors concluded that the fire re- The thermal properties of GFRP material were determined through
sistance of GFRP multicellular columns may be well predicted through dynamic mechanical analysis (DMA) and differential scanning calori-
the calculation of the Euler buckling load. metry and thermogravimetric (DSC-TGA) tests (reported in [22]). DMA
While the structural response of pultruded GFRP members exposed tests allowed the determination of the glass transition temperature (Tg
to elevated temperature has been modelled by assuming perfect elas- = 141 °C2), while DSC-TGA experiments allowed the evaluation of the
ticity, it is known that moderate to high temperatures lead to viscoe- decomposition temperature (Td = 370 °C3).
lastic effects that may be significant for moderate to long durations [6]. Additionally, calcium silicate (CS) boards with thickness of 25 mm
Despite the absence of comprehensive studies on the creep behaviour of were bonded to the GFRP columns with a thin layer of fire resistant
pultruded GFRP profiles under high temperatures, there are a few mastic to protect them against fire. The agglomerate CS boards were
works on GFRP laminates. Dutta and Hui [15] modelled the short-term produced by Promatect (type L) and presented the following thermo-
(30 min) creep response of GFRP coupons at low to moderate tem- physical properties at room temperature: dry density of ρ = 450 kg/m3,
peratures (25–80 °C) using a modified version of Findley's power law thermal conductivity of λ = 0.09 W/m °C and specific heat of Cp
calibrated through experimental values (time-dependent thermal de- = 810 J/kg °C.
formation and failure stresses). Boyd et al. [16] also performed creep The thermal degradation of the elastic properties considered in this
failure tests on one-side fire exposed GFRP laminates subjected to numerical investigation was based on experiments carried out by the
compression. In this study, an analytical model was developed to si- authors and data reported in the literature, as described next. Fig. 1
mulate the creep behaviour and compressive strength, which included depicts the variation with temperature of the longitudinal compressive
thermo-viscoelasticity and was based on Budiansky and Fleck model modulus (EL,comp(T)), in-plane shear modulus (GLT(T)), longitudinal
[17]. compressive strength (σL,comp(T)) and in-plane shear strength (τLT(T)).
In resume, the interaction between different and complex phe- The thermal degradations of GLT(T) and τLT(T) were assessed through
nomena involving mechanical and thermal anisotropy, heat transfer, experiments conducted by the authors on V-notched small-scale cou-
fluid dynamics, viscosity and delamination, still hinders the develop- pons [21]. Compressive tests [23] were also carried out on short GFRP
ment of full and rigorous simulations of GFRP structures under fire. columns to evaluate the mechanical degradation of EL,comp(T) and
These effects added to the deteriorating influence of compressive σL,comp(T) – due to difficulties in measuring deformations, the latter was
stresses, lead to an extremely complex behaviour of pultruded GFRP assumed similar to the former. The variation of σL,comp(T) was adjusted
columns that is still far from being correctly captured through numer- to take into account the influence of kinking failure mechanisms (si-
ical models. milar to the ones registered in the fire resistance tests [18]) observed by
Aiming at determining the mechanical response of pultruded GFRP Sun [24]. As it can be seen in Fig. 1, the moduli and strengths are
tubular columns exposed to elevated temperatures and simulating as
rigorously as possible fire resistance tests previously carried out and 1
Hashin damage criterion was used only in the reference column (un-
described in [18], three-dimensional FE models were developed using
protected and exposed to fire in one-side), also allowing to estimate its re-
ABAQUS Standard [19] software. In this numerical study, different
spective failure load.
complex phenomena are considered, such as thermo-mechanical ani- 2
Based on the peak value of the loss modulus curve obtained from dual-
sotropy, heat transfer and fluid dynamics, while the creep behaviour cantilever flexural setup from room temperature to 250 °C, at a heating rate of
and delamination effects are neglected. The thermo-mechanical prop- 4 °C/min.
erties are assumed temperature-dependent and different curves are 3
Based on the middle temperature of the sigmoidal mass change curve ob-
considered for the degradation of compressive (EC(T)) and shear tained from tests conducted from 30 °C to 800 °C, at a heating rate of 10 °C/min.

522
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Table 1 displacement transducers (δH1 to δH4, cf. Fig. 2). As mentioned, for
Mechanical properties of GFRP profiles at room temperature (RT). more details about the experimental procedure, the interested reader is
Mechanical property Elastic modulus [GPa] Ultimate strength [MPa] referred to [18].
From the experimental tests, it was observed that the fire protection
Flexural longitudinal EL,flex = 28.0 ± 6.1 σL,flex = 352.7 ± 76.1 system was effective in providing thermal insulation to the GFRP col-
Tensile longitudinal EL,tens = 39.9 ± 2.7 σL,tens = 336.5 ± 87.7
umns, in particular, in insulating the bottom flange; similar effective-
Compressive longitudinal EL,compa = 25.3 ± 2.5 σL,comp = 416.4 ± 22.2
Compressive transversal ET,compa = 8.3 ± 1.8 σT,comp = 93.1 ± 14.9
ness was recently reported by Zhang et al. [25,26], based on experi-
Interlaminar shear GLT = 3.6 ± 0.3 τLT = 37.0 ± 2.9 ments where different types of fire resistant panels were used to protect
modular GFRP multicellular slabs. For series S1, the use of CS protec-
a
“Apparent” modulus, as it includes the effects of local crushing at the ex- tion raised the resistance of the columns about 35 min, from 16 min
tremity sections (edges) of coupons. (unprotected) to 51 min (with CS protection). For series S2, the fire
resistance of the unprotected column was substantially reduced to
100%
EC about 6 min. The increase of load level also caused a significant de-
EL,comp (T)
G crease in fire resistance to 9 min. Regarding the failure modes, all GFRP
Normalized elastic properties [%]

GLT (T)
80% σC columns showed brittle (sudden) collapse, which was triggered by the
σL,comp (T)
τ failure of the bottom flange and/or webs (depending on the number of
τLT (T)
sides exposed to fire), as presented in [18].
60%

3. Description of numerical models


40%
Aiming at simulating the fire behaviour of pultruded GFRP columns,
a two-step procedure was adopted: the thermal response of the profiles
20%
was first obtained using numerical models within ANSYS Fluent [27]
framework and the thermal results (variation of temperature distribu-
0% tion with time) were subsequently used as input data to the FE models
0 200 400 600 800 1000 developed within ABAQUS Standard [19]6 framework to obtain the
Time [min] thermo-mechanical response of pultruded GFRP columns. In the
Fig. 1. Variation with temperature of the compressive and shear moduli/ thermal study [20], the authors developed two-dimensional models to
strengths. evaluate the evolution of temperatures in the cross-section of the un-
protected and protected GFRP profiles subjected to different fire ex-
posure (either in one- or three-sides) to the ISO 834 standard fire. The
clearly and similarly reduced with temperature, being considerably
numerical temperature curves were consistent with the experimental
affected by temperature increase even for temperatures below glass
ones, despite the several assumptions adopted and the complexity in-
transition temperature (Tg ~ 141 °C).
volved in this heat transfer problem, including conduction, radiation
This experimental campaign allowed to study the influence of (i)
and convection inside the tubular section. The interested readers should
protection system (with and without CS protection), (ii) exposure type
refer to [20] for more information regarding the thermal simulation of
(one- or three-sides), and (iii) service load levels (6.7% and 13.4% of
the GFRP profiles.
the column strength at RT4) on the fire resistance of GFRP columns. The
test programme included four fire resistance tests, divided in three
series (S1, S2 and S3): (i) S1 – one-side exposure and load of 6.7% of RT 3.1. FE model labelling and mesh
strength; (ii) S2 – three-side exposure and load of 6.7% of RT strength;
(iii) S3 – one-side exposure and load of 13.4% of RT strength. The ex- The labelling of the FE models is summarized in Table 2, including
perimental campaign is summarized in Table 2. the type of fire exposure and the applied load level considered. As
Fig. 2 illustrates the test setup used in the fire resistance tests. The shown in Table 2, columns were labelled according to the following
experiments were conducted on a vertical oven, with external dimen- nomenclature: (i) fire protection (U – unprotected; CS – CS protected);
sions of 1.35 × 1.20 × 2.10 m (long, wide and high, respectively) and (ii) fire exposure (E1 – one-side; E3 – three-side), (iii) load level (L1 –
an opening on the top, in which two different top cover sets were used: 6.7% RT axial strength; L2 – 13.4% RT axial strength). Four uniform
one allowing the fire exposure in one side (for series S1 and S3); and load and time-dependent models were considered (U-E1-L1, CS-E1-L1,
another allowing the fire exposure in three sides (for series S2). In this U-E3-L1, U-E1-L2), all with elastic material behaviour (no material
experimental study, time-temperature curve defined in ISO 834 [11] damage) and identification of initial failure through Tsai-Hill criterion.
was adopted. As illustrated in Fig. 2, the ends of the columns were not A fifth model (U-E1-LV) with the load varied from 0 to Pu (ultimate
directly exposed to heat, having remained at ambient temperature load) and progressive failure included through Hashin's damage cri-
during the entire fire exposure.5 The axial (horizontal) load was applied terion was considered. In this model, the mechanical properties were
by means of a hydraulic jack, being the compressive load measured assumed fixed for the time t = 16 min, at which the collapse of column
with a load cell, as schematized in Fig. 2. Aiming at measuring the U-E1-L1 occurred experimentally.
mechanical response, vertical displacements (flexural deformation) of In the FE model of the protected column (CS-E1-L1), the mechanical
the top flange were measured at the mid-span section using a dis- (stiffness and strength) contribution of the passive fire protection to the
placement transducer (δV), while horizontal displacements (axial de- axial and flexural deformations and stresses was neglected.7
formation) were measured at the columns’ extremities using four
6
This software did not allow the performance of two-dimensional thermal
analysis considering heat transfer though simultaneous conduction, radiation
4
These load values are 55 and 110 kN, respectively, and correspond to and convection inside the tubular section. Consequently, it was decided to
column axial shortenings of L/1500 and L/750. perform the thermal (2D) and mechanical (3D) analysis using different soft-
5
In real applications, this would correspond to situations where the supports ware, ANSYS Fluent and ABAQUS Standard, respectively.
7
are protected by an explicit fire protection system or other constructive layers Using mechanical homogenization principles, it may be concluded that the
(e.g., suspended ceilings and/or raised floors). CS protection allowed an increase of approximately 11% of the RT axial

523
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Table 2
Labelling of numerical models.
FE Models Series Protection Exposure Load level Criterion Damage

U-E1-L1 S1 Unprotected (U) 1 side (E1) P = 55 kN (L1) Tsai-Hill No


CS-E1-L1 S1 Calcium silicate (CS) 1 side (E1) P = 55 kN (L1) Tsai-Hill No
U-E3-L1 S2 Unprotected (U) 3 sides (E3) P = 55 kN (L1) Tsai-Hill No
U-E1-L2 S3 Unprotected (U) 1 side (E1) P = 110 kN (L2) Tsai-Hill No
U-E1-LV S1 Unprotected (U) 1 side (E1) 0 ≤P ≤ Pu (LV) Hashin Yes

Fig. 2. Test setup: (a) frontal view and (b) one- and (c) three-side fire exposure.

Additionally, there is also a lack of information in the literature re- thickness and height, respectively, and 750 elements along the length.
garding both GFRP-CS interface and thermo-mechanical properties of As a result, a mesh with a total of 589,568 elements and 737,940 nodes
material CS. Despite modelling the steel blocks for the reaction sup- was used.
ports, only a GFRP part was considered.
Regarding the FEs used, the time-dependent models (U-E1-L1, CS-
E1-L1, U-E3-L1, U-E1-L2) adopted solid FEs, while the load-dependent 3.2. Boundary, loading and thermal conditions
model (U-E1-LV) considered shell FEs because the solid ones do not
allow the damage progression – Hashin criterion is only available for The mechanical response of GFRP columns was modelled con-
shell FEs. For models U-E1-L1, CS-E1-L1, U-E3-L1 and U-E1-L2, the sidering symmetric conditions with respect to y-z plane (plane of
GFRP columns were discretised using eight-node solid elements with symmetry) and, therefore, displacements in x-axis were fully restrained
reduced integration (C3D8R). The cross-section of GFRP columns was (δx = 0), as shown in Fig. 3. The load (pressure force) was applied
meshed uniformly with 4 and 50 elements across the thickness and through a steel bearing plate (left side of Fig. 3) with free axial de-
along the height, respectively, and with 450 elements along the length. formation (z-axis) but restrained (null) rotations (θx = θy = θz = 0).8
To reduce the computational time of the analyses, only half of the In opposite (right) side, the steel plate was allowed to rotate freely but
profile was modelled assuming conditions of symmetry in transversal it was fixed against displacement (δz = 0), as illustrated in Fig. 3. The
direction, as schematized in Fig. 3. Then, the FE mesh had a total of interface between the GFRP profile and steel blocks was modelled using
165,232 elements and 209,250 nodes. surface-to-surface contact interaction with hard normal contact and
The model U-E1-LV was meshed using eight-node continuum shell rough tangential behaviour. Constant pressures of 5.5 MPa and
elements with reduced integration (SC8R). Continuum shell elements 11.0 MPa were applied to left bearing plate for load levels L1 and L2
were assumed as three-dimensional bodies, with 4 layers across the (see Table 2), respectively, corresponding to half of the total loads
width of each flange and 12 layers across the webs height. In each one (55 kN and 110 kN) due to symmetry simplification. For LV, compres-
of these layers, the temperature-dependent mechanical properties were sive load was incrementally increased up to Pu (ultimate load).
considered, i.e., the assign of different materials depending on the nu- The variation of nodal temperatures with fire exposure time was
merical temperatures obtained in [20] was made. The cross-section of previously obtained from the two-dimensional thermal analysis [20]
GFRP column was meshed uniformly with 4 and 50 elements across the
8
Note that during the fire resistance tests, the supports of the GFRP columns
were kept at ambient temperature; therefore, these boundary conditions should
(footnote continued) not change during the entire fire exposure. If the supports had been directly
stiffness of the unprotected columns. In fact, the axial shortening measurements exposed to fire, the heating of the GFRP material next to the supports could
registered in S1 columns after load application (before beginning fire exposure) cause changes in the rotation constrains due to material softening (in real ap-
were 0.68 mm (U-E1-L1) and 0.80 mm (CS-E1-L1). Thus, the mechanical con- plications, this can be prevented using fire protection systems, suspended
tribution of the passive fire protection system may be neglected. ceilings and/or raised floors).

524
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 3. Geometry, mesh and boundary conditions of FE models.

interaction between acting stresses and material strength components


was considered according to the following equation,
2 2 2
11 11 22 22 12
IF = + + < 1.0
S1 2 S12 S2 2 S122 (1)
where IF is Tsai-Hill failure index, 11, 22 and 12 are respectively the
longitudinal, transversal and shear stresses in a given point of the
column, and S1, S2 and S12 are respectively the longitudinal, transversal
and shear strengths of the GFRP material. When Tsai-Hill index reaches
the unit value (IF = 1.0) at a given point, it indicates that such point
achieved the onset of failure. For a given (calculated) stress state, this
criterion allows the distinction between the damaged (IF > 1.0) and un-
damaged (IF 1.0) state of the column. However, it should be empha-
sized that this failure criterion has no influence on the stiffness de-
gradation and strength of the GFRP columns: it only identifies the zones
Fig. 4. Numerical temperatures obtained from two-dimensional thermal ana-
that are more susceptible to present failure, but this “failure” does not
lysis presented in [20]. introduce any type of reduction in the mechanical properties of the
material.
In order to estimate the load carrying capacity of GFRP columns
and considered herein as input to the mechanical analyses. The evo- under fire exposure and to predict their collapse, the Hashin damage
lution of temperatures at the middle of the top flange (TF), web (W) and criterion [29] was used in the load-dependent model U-E1-LV. In op-
bottom flange (BF) for columns U-E1-L1, CS-E1-L1 and U-E3-L1 is position to the Tsai-Hill criterion, which proposes a single equation to
presented in Fig. 4. Considering the test setup used (cf. Fig. 2(a)), the predict failure initiation, Hashin criterion takes into account the ma-
temperature variation along the length of the profiles was not constant terial progressive damage by considering four failure modes (indexes):
and, therefore, longitudinal temperatures were discretized in three (i) fibre tension index (Ff,T), (ii) fibre compression index (Ff,C), (iii)
parts, and were considered uniform in each of these parts. Fig. 5 shows matrix tension index (Fm,T) and (iv) matrix compression index (Fm,C). In
the longitudinal distribution of temperatures assumed, in which three this criterion, the following equations are used,
different parts were considered: (i) 2D numerical temperatures (ex- 2 2
ˆ11 ˆ12
posed part); (ii) half of 2D numerical temperatures (unexposed part); Ff,T = + 1.0 and ˆ11 0
and (iii) initial temperature set equal to 20 °C (exterior part). Since
S T,1 S12 (2)
neither the experimental nor the numerical distribution of temperatures 2
ˆ11
along the longitudinal direction (z-axis) was known, these assumptions Ff,C = 1.0 and ˆ11 < 0
were considered in order to simulate the mechanical response of the SC,1 (3)
GFRP columns. 2 2
ˆ 22 ˆ12
Fm,T = + 1.0 and ˆ 22 0
S T,2 S12 (4)
3.3. Failure and damage criteria
2 2 2
ˆ 22 SC,2 ˆ 22 ˆ
In order to estimate the initial failure of GFRP columns under fire Fm,C = + 1 + 12 1.0 and ˆ 22 < 0
2S23 2S23 SC,2 S12 (5)
exposure, the Tsai-Hill criterion [28] was used in time-dependent
models (U-E1-L1, CS-E1-L1, U-E3-L1, U-E1-L2). In this criterion, the where ˆ11, ˆ 22 and ˆ12 are the applied effective stresses, ST,i , SC,i and

Fig. 5. Temperature distribution assumed along the length of GFRP columns.

525
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Sij are tensile, compressive and shear strengths and is a coefficient (set coupon tests in compression); shear strength τLT = 37.0 MPa (de-
by default to 0.0) that accounts for the shear stress contribution to fibre termined from coupon tests in shear). All moduli and strengths were
breakage in tension criterion. The effective stresses ˆ are computed by assumed to vary with temperature according to curves depicted in
means of, Fig. 1, while the Poisson's ratios were considered temperature-in-
dependent. The thermal expansion of the GFRP material was defined
1/(1 df ) 0 0
according to the values reported in the literature [2] and was assumed
ˆ = M , where M = 0 1/(1 dm ) 0 temperature-independent in both longitudinal (αlong = 6.7 × 10−6)
0 0 1/(1 d s) (6) and transversal (αtrans = 22.0 × 10−6) directions. These values are also
in agreement with the ones previously used in [31], where a parametric
where M is the damage operator, is applied stress and df , dm and d s
study was performed on GFRP beams. The steel material parts, corre-
are indexes that reflect the current state of fibre, matrix and shear da-
sponding to the steel blocks placed at the end sections, were defined
mage, respectively, given by,
with temperature-independent isotropic properties (Esteel = 210 GPa,
dtens
f if ˆ11 0 νsteel = 0.3).
df =
d comp if ˆ11 < 0 As mentioned, the use of Hashin damage criterion also requires the
f (7)
knowledge of GFRP fracture energies, which were defined based on
d tens some preliminary experimental tests [32]. The fracture energies as-
m if ˆ 22 0
dm = sumed for RT are: (i) FEL,comp = 45 N/mm (longitudinal compressive
d comp
m if ˆ 22 < 0 (8)
failure mode); and (ii) FET,comp = 20 N/mm (transversal compressive
comp failure mode). Since no information was found in the literature, the
(1 d tens )(1 d tens (9)
comp
dm = 1 f )(1 d f m )(1 d m )
variation with temperature of the fracture energies in compression
in which the subscripts “tens ” and “comp” correspond to tensile and (FEL,comp and FET,comp) was assumed equal to that of the longitudinal
compressive modes, respectively. The difference between stress states compressive (EL,comp(T)) modulus.
(before and after damage) in a material point should be responsible for
a sudden drop of strength of the structure to which that point belongs 3.5. Methodology and analysis
and also a decrease of its post-damage stiffness (softening stage). This
computation requires the definition of four different fracture energy In the fire resistance tests of time-dependent models (U-E1-L1, CS-
parameters, one for each failure mode. E1-L1, U-E3-L1, U-E1-L2), three steps were defined: (i) an initial step to
In Hashin damage criterion, the elasticity matrix is affected by the impose the boundary conditions and initial nodal temperatures (20 °C);
state of fibre, matrix and shear damage, which consequently reduces (ii) a mechanical load step (general, static) to apply the constant
the stresses in GFRP material; the damage evolution law (available in pressure; and (iii) a thermal load step (with a time step of 1 min) to
ABAQUS FE package) is based on the energy dissipated during the da- impose the nodal temperatures as function of the time. These are the so-
mage process and linear material softening [30]. Contrarily to Tsai-Hill called thermo-mechanical loading models. For comparison purposes for
criterion, in which the elasticity matrix is affected by temperature-de- axial and flexural deformations (Sections 4.1. and 4.2.), they were also
pendent material properties, the Hashin criterion also considers the used only for thermal loading, in which the mechanical compressive
reduction of elasticity matrix with the state of damage in GFRP mate- load was absent. In this case, the labelling excludes the capital “L” – e.g.
rial. When any of the failure indexes attains the limit value (1.0) in a U-E1-L1 is the thermo-mechanical loading model of the unprotected
given point, the stress state in that damaged point is computed by column under one-side exposure and lowest load level, while U-E1 is
means of, the same model without compressive load.
(1 df )E11 (1 df )(1 dm) 21E11 0 Transient linear analyses (time-dependent models) were carried out
considering a total time of 60 min and a time step of 1 min. It was
11 11
1
22 = (1 df )(1 dm) 12 E22 (1 df )E22 0 22
12
D 12 possible to predict the fire behaviour of the GFRP columns tested in
0 0 (1 d s)G12 D
[18], namely axial and flexural deformations. Nevertheless, the ana-
(10) lyses were not successfully completed in most of the numerical simu-
where ij are the applied stresses, ij is a strain component, E11, E22 and lations due to numerical convergence difficulties. The durations of the
G12 are longitudinal, transversal and shear moduli, 12 and 21 are numerical simulations were: U-E1-L1 (t = 39 min), CS-E1-L1
Poisson ratios, and D = 1 (1 df )(1 dm) 12 21. (t = 60 min), U-E3-L1 (t = 9 min), U-E1-L2 (t = 25 min). These “nu-
merical” durations were however higher that the corresponding ex-
3.4. Material properties perimental ones.
In the load-dependent model (U-E1-LV), a two-step procedure was
The GFRP material was modelled assuming an orthotropic beha- defined: (i) an initial step to impose the boundary conditions and to
viour using ABAQUS Standard, in which the three elastic moduli, assign the different materials (for each layer), depending on the nu-
Poisson's ratios and shear moduli corresponding to material's principal merical temperatures computed for t = 16 min (time of experimental
directions were assigned. Taking into account the mechanical proper- failure observed [18]); (ii) a mechanical load step (general, static) to
ties measured in experimental tests (Table 1), the following RT com- increase the pressure up to collapse of the column. In this model, an
pressive properties were considered (the corresponding source is in- initial imperfection10 (in both flexural and axial directions) caused by
dicated in brackets): longitudinal modulus EL = 31.0 GPa (full-section thermal expansion of GFRP column was imposed. The results obtained
elastic modulus); transverse modulus ET = 8.3 GPa (compressive from these simulations are presented and discussed in the next section.
transverse modulus); out-of-plane modulus EZ = 3.5 GPa (compressive
modulus of polymeric matrix); shear modulus GLT = 3.6 GPa (full-sec- 4. Numerical results and discussion
tion and material shear modulus); νLT = 0.27 (coupon tests)9; long-
itudinal strength σL = 416.4 MPa (determined from coupon tests in In this section, the numerical results are presented and discussed. In
compression); transversal strength σT = 93.1 MPa (determined from Sections 4.1 and 4.2, the evolution of column axial and flexural

9 10
In the other planes, this value of shear modulus was assumed, while a value The referred initial imperfection applied was determined from model U-E1
of 0.11 was adopted for the Poisson's ratio. where only thermal loading was applied (i.e., no mechanical loading).

526
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 6. Variation of axial deformation with fire exposure time for columns (a) U-E1-L1/CS-E1-L1 and (b) U-E3-L1/U-E1-L2 – mid-span section (positive-elongation;
negative-shortening).

behaviours with fire exposure time is assessed and compared with ex- following semi-empirical equation,
perimental measurements [18]. In Section 4.3, the evolution of stress (T/T0)
distributions in both longitudinal and transversal directions of the re- t
creep (t) = 0+p
ference column (U-E1-L1) is presented and the effects of fire protection, t0 (11)
fire exposure and applied load level are discussed in terms of stress
where creep is creep strain, t is time (in hours), 0 is initial strain (set as
patterns. Finally, in Sections 4.4 and 4.5, the failure and damage of
null), p is a coefficient applied to time ratio (t/t 0) and is a coefficient
GFRP columns is investigated through the use of Tsai-Hill failure cri-
applied to temperature ratio (T/T0 ). In that study, parameters p and
terion (time-dependent models) and Hashin damage criterion (load-
were set as 45 and 0.29, respectively, as a result of a curve fitting from
dependent model), respectively.
the experimental data measured in the compression creep test at 80 °C.
In order to estimate the influence of creep deformation in the pro-
4.1. Axial behaviour
tected column (CS-E1-L1, the longer test where the importance of creep
effects was expected to be higher), the instant in which the average
Fig. 6 presents the variation of axial deformation with fire exposure
temperature (Tm)11 at the mid-span section reaches 80 °C was de-
time obtained from both numerical analyses and experimental tests.
termined. Considering the evolution of numerical temperatures in
During the first 5 min, a slight axial elongation (positive) caused by the
column CS-E1-L1, it was concluded that the average temperature at the
thermal expansion of GFRP material was measured in reference column
mid-span section attains 80 °C after 23 min of fire exposure. Using Eq.
U-E1-L1 (unprotected and one-side exposed). In this period, a perfect
(11) and considering only the length directly exposed to fire, it can be
agreement was found between the results of model U-E1 (only thermal
estimated that the creep strain for Tm = 80 °C (corresponding to
loading) and experimental ones. However, the models U-E1-L1 and U-
t = 23 min) would be about 825 µm/m and the axial shortening in-
E1-L2 (both with thermal and mechanical loadings) exhibited only axial
crease would be approximately 0.8 mm. It should also be highlighted
shortening (negative), which is due to the reduction in the elasticity
that the axial shortening (obtained from model CS-E1-L1) is 0.1 mm
modulus of the material at elevated temperature. In the fire protected
after 23 min of fire exposure, and that value is 12.5% of the axial
column, the axial shortening experimentally registered was much
shortening caused by creep (determined using Eq. (11)). This prediction
higher that the numerical one. Despite the overall fair qualitative
confirms the potential behavioural influence of creep (namely, in axial
agreement between the numerical and experimental curves for the
deformations) and explains, to a certain extent, the differences found
unprotected columns, such consistency was not observed for the pro-
between some numerical and experimental results (cf. Fig. 6).
tected column: the numerical axial shortening values of CS-E1-L1 were
The unprotected column exposed to fire in three sides (U-E3-L1 and
considerably lower than those experimentally measured (CS-E1-
U-E3-L1(exp)) shortened quickly during the first 6 min, as a result of the
L1(exp)), as shown in Fig. 6(a). These differences might be due to (i) the
severe temperature increase registered – in this case, during that period,
variation of nodal temperatures with fire exposure time (the numerical
the effect of elasticity modulus reduction caused by elevated tempera-
temperatures were slightly lower than those measured, in particular at
ture was more relevant than the (competing) effect of thermal expan-
the webs); (ii) the test setup conditions (the misalignments in test setup,
sion. The overall tendency of the numerical results was consistent with
friction in supports and column geometrical imperfections might play a
the experimental data, both presenting a significant shortening during
key role); (iii) the thermal degradation of mechanical properties (e.g.
that initial period. In opposition, the increase of load level in the un-
the reduction of compressive modulus); and (iv) the creep effect. This
protected column (U-E1-L2(exp)) led to its experimental collapse after
last effect, not considered in the models, is consistent with the fact that
9 min. In this case, the numerical model U-E1 provided an initial axial
experimental deformations are generally higher than numerical ones.
elongation consistent with experimental data, while such behaviour
Aiming at quantifying the potential influence of creep effect in the
was not observed in model U-E1-L2. The collapse of columns U-E3-L1
mechanical response of the GFRP columns, a semi-empirical equation
proposed by Dutta and Hui [15] was used. As mentioned (cf. Section 1),
these authors performed an experimental investigation on the creep 11
The average temperature at mid-span section was calculated considering
rupture of GFRP composites at different temperatures (25, 50 and the average of the temperatures at the middle of bottom (P13–P14) and top
80 °C). Based on Findley's equation, Dutta and Hui [15] proposed the (P2–P3) flanges (cf. Fig. 8).

527
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 7. Variation of transversal deflection with fire exposure time for columns (a) U-E1-L1/CS-E1-L1 and (b) U-E3-L1/U-E1-L2 – mid-span section.

and U-E1-L2 occurred only 6 and 9 min after burners’ ignition, re- 4.3. Stress distribution
spectively. Therefore, few conclusions may be drawn from the com-
parison between numerical and experimental results and the differences In this section, the models were used to study the evolution of
observed may be due to the same factors referred above. stresses with fire exposure time (in practice, not possible to measure in
fire tests, especially in walls directly exposed to fire). To achieve this
goal, 15 FEs along the height of the mid-span section (see Fig. 5) of the
4.2. Flexural behaviour GFRP columns were selected and distributed as follows (Fig. 8): 4 FEs at
the centre of top flange (P1-P4), 7 FEs along the web height (P5-P11)
Fig. 7 shows the variation of transversal deflection with fire ex- and 4 FEs at the centre of bottom flange (P12-P15). In subsection 4.3.1.,
posure time obtained from the numerical analyses and measured in the the stress distribution and its time evolution are described for the re-
experimental tests. According to experimental results, the GFRP col- ference column U-E1-L1. Then, in subsection 4.3.2., the stress patterns
umns moved downwards (towards the heating source) as a result of the of the reference column U-E1-L1 are compared to those of columns CS-
temperature gradient in the cross-section's profile and the material E1-L1, U-E3-L1 and U-E1-L2, to evaluate the influence of fire protec-
thermal expansion. This initial descending movement (corresponding to tion, fire exposure and applied load level, respectively, on their beha-
the initial negative values) was less perceptible in the protected column viour.
because the thermal insulation conferred by the fire protection allowed
the reduction of the thermal gradient across the cross-section's height.
4.3.1. Unprotected column with one-side fire exposure: U-E1-L1 (reference)
That deflection downwards, associated to thermal expansion effect, was
4.3.1.1. Longitudinal stresses. Fig. 9 presents the evolution of normal
only observed only in models U-E1, CS-E1 and U-E3. Nevertheless, the
stresses in the longitudinal direction (σ11) of mid-span section of
deflections obtained from these models were significantly lower than
those measured from fire tests. With the exception of the protected
column (CS-E1-L1), the overall tendency of the remaining numerical
curves (U-E1-L1, U-E3-L1 and U-E1-L2) was different from the experi-
mental ones (cf. Fig. 7). In these models, the columns started deflecting
upwards as a consequence of the fire exposure and, subsequently, a
progressive loss of profiles’ axial stiffness occurred. The ascending
movement observed in all models is in agreement with the experimental
results [18] – this movement upwards was registered after a descending
one. Although meaningless negative vertical displacements were ob-
served in columns CS-E1-L1 and U-E3-L1 (thus, corresponding to a
descending movement), the overall trend is to revert these displace-
ments to positive ones (ascending movement) for increasing fire ex-
posure – this behaviour was qualitatively similar to that observed in the
experimental tests.
The models developed showed some inherent limitations, namely
the disregard of material progressive failure under compression, which
partly justifies the mentioned differences in terms of transversal de-
flections (particularly in the unprotected columns). Despite these ex-
pected limitations, the numerical models are able to give a reasonable
qualitative prediction of the transversal deformation/deflection of the
columns. For example, and despite the several assumptions adopted,
the curve CS-E1-L1 agreed well with that registered in the experimental
campaign, a match that may be explained by the reduced thermal Fig. 8. Finite elements (P1–P15) selected for stress distribution analysis – mid-
gradient through the profile's height. span section.

528
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 9. Evolution of longitudinal stresses (σ11) in column U-E1-L1 at (a) top flange (P1-P4), (b) web (P5-P11) and (c) bottom flange (P12-P15) – mid-span section.

column U-E1-L1 in its top flange (P1-P4, Fig. 9(a)), web (P5-P11, 4.3.1.2. Transversal stresses. Fig. 10 presents the evolution of normal
Fig. 9(b)) and bottom flange (P12-P15, Fig. 9(c)). During the first stresses in the transversal direction (σ22) of mid-span section of column
10 min, a significant reduction in the longitudinal stresses is observed U-E1-L1 in its top flange (P1-P4, Fig. 10(a)), web (P5-P11, Fig. 10(b))
in the bottom flange (P12-P15), which was caused by the rise of and bottom flange (P12-P15, Fig. 10(c)). Although being null at the
temperature. Consequently, σ11 stresses increased in the web (P6-P10), onset of fire exposure, the transversal stresses increased in the first
particularly in its lower part, due to the axial stiffness reduction of the 3 min, particularly in the bottom flange. In this flange, tensile (P12,
bottom flange – the exception is P11, whose σ11 curve is similar to those P13) and compressive (P14, P15) stresses were observed as a result of
observed in the bottom flange (P12-P15). It was also observed a slight the considerable temperature gradients imposed, which tended to zero
increase in longitudinal stresses acting on the top flange (P1-P4), after that initial period (cf. Fig. 10(c)). Additionally, the variation of the
namely the inner layer (P4), although this increase was considerably web transversal stresses is small and the minor compressive stresses
lower than that registered in the web. (lower than 1.0 MPa) tend to zero. With the exception of P1 and after an
After that initial period, the longitudinal stresses remained ap- initial insignificant variation, the σ22 stresses in the top flange tend to
proximately unaltered in the bottom flange (~1.0–9.0 MPa), while they uniform values associated to through-thickness stress gradient (tension
presented a slight decrease (between 10 and 25 min) at top flange in P1-P2 and compression in P3-P4). Taking into account these results,
(particularly in P3-P4). In parallel, the rise of temperature at the webs it can be stated that transversal stresses developed particularly in the
caused a progressive stiffness reduction and, consequently, a decrease bottom flange, but these were expectedly much lower than the
in longitudinal stresses (see Fig. 9). Then, an increase of the long- longitudinal stresses (see Fig. 9) because of the 1D nature of column
itudinal stresses in the inner layer of the bottom flange (P12) was ob- behaviour.
served, most likely due to axial stiffness reduction of the top flange. The
increase of this localized stress in the bottom flange was closely related 4.3.1.3. Shear stresses. Fig. 11 presents the time evolution of shear
to the temperature at element P12, which was significantly lower than stresses (σ12) in the web (P5-P11)13 of column U-E1-L1. Considering the
those registered at elements P13-P15.12 According to the numerical variation of σ12 stresses with fire exposure time, it can be stated that a
results, a clear stress reduction occurred in the top flange as a result of gradual increase was observed as a result of transversal deflection,
the temperature increase in the overall cross-section. Similarly, the which was caused by temperature increase and thermal degradation of
longitudinal stresses in the upper part (P6-P8) of the webs started de- the mechanical properties (particularly in the lower part of cross-
creasing due to temperature increase, thus causing an increase of σ11 section). Because the columns deflected upwards (moving away from
stresses in the web lower part (P9-P10). the oven) due to the progressive loss of axial stiffness in the bottom
Generally speaking, it can be stated that longitudinal stresses in the flange, the shear stresses expectedly increased in the web.
webs increased significantly due to the quick drop of bottom flange
axial stiffness in the first 10 min. After that, it can be seen that the
4.3.2. Influence of fire protection, fire exposure and load level
progressive heating of the cross-sections caused huge stress variations:
4.3.2.1. Effect of fire protection: CS-E1-L1 vs. U-E1-L1. Aiming at
all curves (except P5 and P11) display local minima, which denotes a
studying the effect of fire protection on the evolution of stress
certain stress transfer between the several points of the web along its
distributions, the numerical results obtained for one-side exposed
height. The material thermal expansion, considered in the numerical
columns U-E1-L1 and CS-E1-L1 (unprotected and protected with CS
models, also influences moderately the evolution of longitudinal
board, respectively) are compared. In Fig. 12(a), the evolution of
stresses, especially in the initial period (~10 min).

13
The stress evolutions in both top (P1-P4) and bottom (P12-P15) flanges are
12
For instance, after 30 min of fire exposure, the numerical temperatures at not shown herein because they are meaningless (nearly null for the entire
elements P12, P13, P14 and P15 are 411, 512, 624 and 747 °C, respectively. duration of numerical simulation).

529
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 10. Evolution of transversal stresses (σ22) in column U-E1-L1 at (a) top flange (P1-P4), (b) web (P5-P11) and (c) bottom flange (P12-P15) – mid-span section.

temperature increase at the lower part of web (P10-P11).

4.3.2.2. Effect of type of fire exposure: U-E3-L1 vs. U-E1-L1. With the
purpose of evaluating the influence of the type of fire exposure on the
GFRP columns response, Fig. 12(b) shows the evolution of longitudinal
stresses (σ11) along the profile's height (y) of columns with one-side (U-
E1-L1) and three-side (U-E3-L1) fire exposures.
In column U-E3-L1, the shape of longitudinal stress diagram change
from uniform to a highly nonlinear one after only 3 min of fire ex-
posure. In comparison with column U-E1-L1, the longitudinal stresses at
a significant part of web/bottom flange (P7-P11) decreased due to axial
stiffness reduction in this zone associated to the rapid increase of
temperatures across the tubular section. In parallel, the high stresses
migrate to the upper part of web/top flange (P5-P6), which is the less
heated zone. With the evolution of time exposure, this stress pattern
amplifies (in Fig. 12(b), see the stress distribution after 6 min). The fire
exposure also influences the neutral axis position: the web neutral axis
Fig. 11. Shear stresses (σ12) in column U-E1-L1 at web (P5-P11) – mid-span is located very close to the bottom flange in column U-E1-L1 (one-side
section.
exposure) but nearer the top flange in column U-E3-L1 (three-side ex-
posure). However, it is seen that the maximum compressive stress level
(23–26 MPa for t = 3 min, 30–34 MPa for t = 6 min) is not significantly
longitudinal stresses (σ11) along the web's height (y) of columns U-E1- influenced by the type of fire exposure.
L1 and CS-E1-L1 is presented. Initially (t = 0 min), the longitudinal
stress diagram is uniform (σ11 = 18.7 MPa) in both columns. After 4.3.2.3. Effect of load level: U-E1-L2 vs. U-E1-L1. In order to assess the
15 min of fire exposure, longitudinal stresses in column U-E1-L1 (i) effect of the applied load level on the GFRP columns response,
decreased significantly near the bottom flange (P11) due to rapid Fig. 12(c) shows the evolution of longitudinal stresses (σ11) along the
increase of temperatures, (ii) increased considerably in the central part profile's height (y) of columns under compressive loads of 6.7% (U-E1-
of web (P6-P10) as a consequence of axial stiffness loss at the bottom L1) and 13.4% (U-E1-L2) of RT axial strength.
flange, and (iii) remained approximately uniform near the top flange The shape of the longitudinal stress diagram in column U-E1-L2 (cf.
(P5). In opposition, longitudinal stress diagram in column CS-E1-L1 Fig. 12(c)) was similar to that of the reference column (U-E1-L1), while
remained approximately uniform (cf. Fig. 12(a)) for t = 15 min. the stress magnitudes were very dissimilar. As a consequence of load
Then, for t = 30 min, the longitudinal stresses at the lower part of level increase, higher initial stresses were observed (37.3 MPa, instead
U-E1-L1 web (P8-P10) continued to increase due to heating. of 18.7 MPa), as expected. The longitudinal stresses (i) increased sig-
Additionally, the longitudinal stresses increased at the upper part of the nificantly in the web (in particular, at element P10) due to the tem-
web/top flange (P5-P6, U-E1-L1), caused by the thermal degradation of perature increase in the lower part of profile, (ii) decreased near the
the mechanical properties of the top flange. On the other hand, the bottom flange (P11) for the same reason, and (iii) remained almost
longitudinal stresses in CS-E1-L1 (i) decreased near its bottom flange constant at the upper part of profile (P5). Also, it should be noted that
(P11), (ii) increased significantly in the web (in particular, at element the mentioned stress concentration at P10, where σ11 stresses increased
P10), and (iii) remained approximately constant at the top part of web from 37.3 to 64.3 MPa, might explain the premature failure (9 min)
(P5). Moreover, for t = 45 min, the longitudinal stress diagram in observed in column U-E1-L2. In resume, as expected, the applied load
column CS-E1-L1 displays a local maximum at P9 due to the level influences a lot the magnitude of longitudinal stresses, but does

530
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

y
Mid-span section

100 100 100

80 80

Web height (y) [mm]


80

Web height (y) [mm]


Web height (y) [mm]
60 60 60

40 40 40

20 20 20

0 0 0
-50 -40 -30 -20 -10 0 -50 -40 -30 -20 -10 0 -75 -65 -55 -45 -35 -25 -15 -5
Long. stress (σ11) [MPa] Long. stress (σ11) [MPa] Long. stress (σ11) [MPa]
U-E1-L1(t=0)
CS-E1-L1(t=0) U-E1-L1(t=0)
CS-E1-L1(t=0) U-E1-L1(t=0) U-E3-L1(t=0) U-E1-L1(t=0) U-E1-L2(t=0)
U-E1-L1(t=15)
CS-E1-L1(t=15) U-E1-L1(t=15)
CS-E1-L1(t=15)
U-E1-L1(t=3) U-E3-L1(t=3) U-E1-L1(t=3) U-E1-L2(t=3)
U-E1-L1(t=30)
CS-E1-L1(t=30) U-E1-L1(t=30)
CS-E1-L1(t=30)
CS-E1-L1(t=45) CS-E1-L1(t=45) U-E1-L1(t=6) U-E3-L1(t=6) U-E1-L1(t=6) U-E1-L2(t=6)

(a) (b) (c)


Fig. 12. Comparison of longitudinal (σ11) stresses evolution for columns (a) U-E1-L1/CS-E1-L1 (at 0, 15, 30 and 45 min), (b) U-E1-L1/U-E3-L1 (at 0, 3 and 6 min) and
(c) U-E1-L1/U-E1-L2 (at 0, 3 and 6 min) – mid-span section.

not affect the position of neutral axis and the shape of the stress dia- top part it did not occur during the simulation (IF ≈ 0.2 at P5, after
gram. 34 min of fire exposure). Thus, it can be stated that column's collapse
was triggered by the initial failure of bottom flange and bottom part of
4.4. Initial failure: Tsai-Hill criterion the web, mostly due to longitudinal stresses ( 11 – blue curves) and
lesser to transversal stresses ( 22 – green curves) and interaction be-
In order to investigate the initial failure of GFRP columns under fire tween them ( 11 22 – light green curves). This predicted and initial
action and to better understand some of the experimental results, the failure mode is consistent with the observed experiments.
Tsai-Hill failure criterion [28] was implemented in the time-dependent
models, as described in Section 3.3. Herein, the time evolutions of Tsai-
Hill failure index and residual section index are presented and dis- 4.4.2. Tsai-Hill index: all columns
cussed. In order to compare the numerical results obtained from the dif-
ferent time-dependent models and to predict the columns’ initial
4.4.1. Tsai-Hill index and stress ratios: reference column (U-E1-L1) failure, the variation of Tsai-Hill index (IF ) at the middle of top flange
As shown in Eq. (1), the expression of Tsai-Hill index comprises four (TF – P2-P3), web (W – P8) and bottom flange (BF – P13-P14) with fire
terms hereafter designated by quadratic stress ratios (σ/σu)2 for (i) exposure time is presented in Fig. 15.
longitudinal stresses (σ11/S1)2, (ii) transversal stresses (σ22/S2)2, (iii) The initial failure mode of column U-E1-L1 predicted by Tsai-Hill
shear stresses (σ12/S12)2, and (iv) combined longitudinal-transversal criterion (failure of bottom flange and bottom part of the web) was
stresses (σ11σ22/S1)2. Thus, the Tsai-Hill index enables the identifica- partially observed in the protected column (CS-E1-L1), in which the
tion of the stress components that contribute the most to the initial initial failure occurred in the bottom flange (IF = 1.0 at P13-P14) for
failure of the columns. Again, the analysis was started with the un- t = 27 min (well before experimental collapse), while the web initial
protected and one-side exposed column (U-E1-L1). Fig. 13 (top and failure did not occur during the simulation (IF ≈ 0.24 at P8, after
bottom flange) and Fig. 14 (web) depict the time evolution of these 60 min of fire exposure).
ratios and Tsai-Hill index (IF ) for column U-E1-L1. In opposition, different failure modes were observed in columns U-
From the observation of these figures, it may be clearly stated that E3-L1 (exposed to fire in three sides) and U-E1-L2 (subjected to an in-
the longitudinal stresses ( 11) were the most influential to failure index creased load level). The three-side fire exposure led column U-E3-L1 to
as the (blue) curves always precede the other stress curves, in terms of a premature initial failure (for t = 4 min, and before the experimental
time evolution. Regarding the evolution of Tsai-Hill index in flanges (IF collapse for t = 6 min). This failure was attained simultaneously at the
– black curves in Fig. 13), it is clear that the bottom flange initial failure bottom flange and web (at lower and central parts). The load level
(IF = 1.0 at P13-P14) occurred for t = 3 min (well before experimental increase caused also the premature initial failure of column U-E1-L2
collapse), while the top flange initial failure did not occur during the (for t = 3 min, and before the experimental collapse for t = 9 min),
simulation (IF ≈ 0.6 at P2-P3, after 34 min of fire exposure). Con- which seemed to have occurred mainly due to the bottom flange failure.
cerning the evolution of Tsai-Hill index in the web (IF – black curves in As expected, the results shown above indicate that the Tsai-Hill
Fig. 14), it is also evident that the web initial failure occurred in its criterion provided conservative estimates of the fire resistance of all
bottom part (IF = 1.0 at P11) for t = 3 min (well before experimental GFRP columns. This is explained by the fact that this criterion only
collapse), while the web initial failure occurred in its middle part considers the stress interaction and does not take into account the
(IF = 1.0 at P8) for t = 27 min (after experimental collapse) and in its material stiffness degradation when IF > 1.0.

531
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 13. Stress ratios (σ/S) and Tsai-Hill index (IF) evolution in the flanges of column U-E1-L1 at: (a) P2-P3 (TF) and (b) P13-P14 (BF) – mid-span section.

4.4.3. Residual section index the area of part i that has failed (IF > 1.0).
With the purpose of predicting the collapse of columns, the residual Fig. 16 presents the evolution of residual section index (IR) with fire
section index (IR) was also calculated. It is defined as the ratio between exposure time, contributing for a better understanding of columns’
the area of each part (top flange, web, bottom flange) of the tubular failure. According to this index and considering the experimental fire
section that remained with the Tsai-Hill index below the unit value resistances, the following conclusions can be drawn: (i) columns U-E1-
(IF < 1.0) and the corresponding total area of that part. This index was L1 and CS-E1-L1 collapsed due to bottom flange failure (IR = 0%) and
determined using the following equation, to partial web failure (IR ≈ 85%); (ii) collapse of column U-E3-L1
Ai AF , i (three-side exposed) was caused by bottom flange failure (IR = 0%), as
IR = < 1.0 well as web failure (IR = 18%); and (iii) column U-E1-L2 (subjected to
Ai (11)
the load level increase) collapsed due to bottom flange failure only (IR
where Ai is the total area of part i of the cross-section, the subscript i = 0%). It should be noted that these conclusions are in agreement with
refers to either the top flange, the web or the bottom flange, and AF , i is those previously mentioned, as well as with those observed in the

Fig. 14. Stress ratios (σ/S) and Tsai-Hill index (IF) evolution in the web of column U-E1-L1 at: (a) P5, (b) P8 and (c) P11 – mid-span section.

532
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

TF
W
BF
Mid-span section

1,0 1,0 1,0


Collapse

Collapse

Collapse

Collapse

Collapse

Collapse

Collapse

Collapse

Collapse

Collapse

Collapse

Collapse
Tsai-Hill index (IF) [-]
Tsai-Hill index (IF) [-]
Tsai-Hill index (IF) [-]

0,8 0,8 0,8

0,6 0,6 0,6

0,4 0,4 0,4

0,2 0,2 0,2

0,0 0,0 0,0


0 15 30 45 60 0 15 30 45 60 0 15 30 45 60
Time [min] Time [min] Time [min]
U-E1-L1 U-E3-L1 U-E1-L1 U-E3-L1 U-E1-L1 U-E3-L1
U-E1-L2 CS-E1-L1 U-E1-L2 CS-E1-L1 U-E1-L2 CS-E1-L1

(a) (b) (c)


Fig. 15. Comparison of Tsai-Hill index (IF) evolution at (a) top flange (P2-P3), (b) web (P8) and (c) bottom flange (P13-P14) for all columns – mid-span section.

Fig. 16. Residual section (IF < 1) evolution at top flange (TF), web (W) and bottom flange (BF) of (a) columns U-E1-L1/CS-E1-L1 and (b) columns U-E3-L1/U-E1-L2 –
mid-span section.

experimental campaign [18]. (nonlinear behaviour), (iii) initial damage occurred at the bottom
flange for the load of 76.4 kN (point II), and (iv) collapse of the column
4.5. Progressive failure: Hashin damage criterion took place for the peak load of 88.4 kN (point III).
According to the four failure modes of Hashin's criterion (see Eqs.
The results shown previously were not affected by the stress level in (2)–(5)), the collapse of column U-E1-LV occurred due to fibre com-
each point of the columns, even if these were above the GFRP material's pression. Although the failures by matrix compression and matrix ten-
strengths. Using now the fifth model (U-E1-LV), it will be possible to sion also appeared in some regions of the web, they may be neglected
perform an incremental analysis considering the mechanical properties when compared to the fibre compression failure. Therefore, the evo-
at t = 16 min, which corresponds to the experimental failure of the lution of longitudinal compressive damage is presented in Fig. 18 for
reference column, and use the Hashin damage criterion to identify the the three load levels depicted in Fig. 17 (I, II and III). In Fig. 18, the
progressive damage and collapse of that column. The results of this longitudinal compressive damage14 is shown at three different load
analysis are presented in Figs. 17–19. Fig. 17 shows the variation of levels: (I) the initial damage (P = 27.1 kN), (II) the bottom flange initial
compressive load and the axial shortening. From the observation of this damage (P = 76.4 kN) and (III) the column collapse (P = 88.4 kN).
figure, it may be concluded that (i) linear elastic behaviour is found for
compressive load up to 27.1 kN (for which initial damage occurred at
the interface between the exposed and unexposed parts of the column – 14
Red colour indicates fully damaged points (d = 1.0) while blue colour
point I), (ii) axial stiffness reduction was observed beyond this load denotes undamaged ones (d = 0.0).

533
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 17. Compressive load vs. axial shortening curve obtained from the nonlinear analysis (using Hashin damage criterion) in column U-E1-LV at failure (t = 16 min).

Fig. 18. Longitudinal compressive damaged zones in column U-E1-LV for different load levels (“I”, “II” and “III”).

Fig. 19 shows the variation of longitudinal (Fig. 19(a)) and transversal progressively emerged in the bottom flange, as shown in Fig. 19(b).
(Fig. 19(b)) stresses at bottom flange (P12-P15) with the axial short- This significant transversal stress gradient developed through the
ening of column U-E1-LV. thickness of the bottom flange would certainly cause its delamination,
For load level I, longitudinal compressive damage (region (1)) was which was observed in the experimental tests but was not considered in
observed at the interface between “exposed part” and “unexposed part” the current model (U-E1-LV). Moreover, the delaminated layers would
(previously indicated in Fig. 5). The stress concentration in this zone most likely buckle for low levels of compressive stresses due to their
was caused by the discontinuity in the material mechanical properties reduced thickness. Thus, it can be stated that the absence of delami-
between “exposed” and “unexposed” parts. Longitudinal compressive nation effects in the current model also introduces a virtual stiffening
damage was also observed at the interface between “unexposed part” effect that provides fictitious (non-realistic) strength to the column for
and “exterior part” (region (2)). Therefore, the referred damage was upper load levels. When the column load attains 76.4 kN, fully damaged
mainly caused by the longitudinal temperature distribution imposed elements (red colour, d = 1.0) were observed at the bottom flange, as
(i.e., it did not completely reproduce the experiments carried out [18]). shown in Fig. 18 (in particular, at region (3)). Fig. 18 also shows the
For the load level I, a progressive decrease in the longitudinal com- spread/increase of the damaged areas in region (4). Finally, the col-
pressive stresses at the bottom flange (elements P12-P15) was observed lapse of column U-E1-LV occurred when the compressive load reaches
(cf. Fig. 19(a)). Simultaneously, transversal tensile stresses 88.4 kN and the longitudinal compressive damage in different zones are

534
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

Fig. 19. Variation of (a) longitudinal and (b) transversal stresses with axial shortening in the bottom flange (P12-P15) of column U-E1-LV – region (3).

presented in region (5) (cf. Fig. 18). an ascending movement of the columns while the latter exhibited a
The collapse load in the experimental test was 55 kN, which is 72% descending one. Most likely, this behavioural difference was caused
of load level II (bottom flange initial damage) and 62% of load level III by the thermal expansion coefficients considered and was more
(column collapse). The 38% difference between experimental and nu- evident in the unprotected columns (higher thermal gradients).
merical collapse loads is high but considering the plethora of complex • The shape of longitudinal stress diagrams changed from uniform
phenomena involved and the uncertainty of their associated para- (room temperature) to highly nonlinear (fire exposure time). In
meters, it can be affirmed that the model U-E1-LV correctly captures the comparison with the reference column (unprotected, one-side fire
behavioural aspects of GFRP column under fire exposure. Recall that exposure, load level equal to 6.7% of axial strength), the influence of
phenomena like viscoelasticity (creep and time-dependent behaviour) fire protection, type of fire exposure and load level on the column
and delamination (through-thickness damage and physical separation) behaviour was different. The calcium silicate protection decreases
were not considered, which would certainly lower the collapse load the magnitude of longitudinal stresses to a reasonable extent, but
predicted by the numerical model. Additionally, the adopted values of does not affect too much the position of neutral axis and the shape of
fracture energy for GFRP materials are yet uncertain (unavailable in the stress diagram. The change from one- to three-side fire exposure
literature) and further studies are needed to obtain reliable data on this influenced hugely the shape of stress diagram and neutral axis po-
parameter. Nevertheless, the model agreed with the experimental re- sition, migrating this from the bottom to the top of the web.
sults reported in [18], thus confirming (i) the location of failure section However, this fire exposure change did not affect much the max-
(where longitudinal compressive damage at the bottom flange occurred imum compressive stress. The increase in applied load level influ-
at first – region (3)) and (ii) the cause of failure mode (according to this enced proportionally the magnitude of longitudinal stresses, but did
model, the collapse occurred due to bottom flange compression). not affect the position of neutral axis and the stress diagram shape.
• As expected, the Tsai-Hill criterion provided conservative estimates
5. Conclusions of the fire resistance of all GFRP columns. According to the Tsai-Hill
criterion, the initial failure of GFRP columns occurred near the mid-
This paper presents a numerical investigation on the fire resistance span section, being consistent with the experimental results. The
of pultruded GFRP columns, compared and verified against experi- longitudinal stresses were the most influential to failure index, while
mental results [18]. The paper focused on the evaluation of several transversal and shear stresses were clearly less influential. Taking
effects, such as the use of a fire protection system, the imposition of into account this failure criterion, the one-side exposed columns
different fire exposure conditions and the application of distinct load under an applied load of 6.7% of axial strength (either protected or
levels. The following conclusions are drawn: unprotected) would have collapsed when the entire bottom flange
and part of the webs (~15%) reached the limit value of failure
• GFRP columns exposed to elevated temperatures exhibited pro- criterion. The unprotected column with three-side exposure and
gressive shortening due to thermal degradation of mechanical applied load level of 6.7% of axial strength would collapse due to
properties caused by temperature increase in cross-sections. In op- bottom flange and web failure, while the unprotected column with
position to the experimental measurements, where a slight initial one-side exposure with applied load level of 13.4% of axial strength
elongation was registered in some specimens, the numerical models would collapse due to bottom flange failure only. These results were
presented only axial shortening that may have been caused by some in agreement with those observed in the experiments.
uncertainties regarding the thermal expansion coefficient adopted. • The strength and collapse mode of reference column (unprotected,
It should also be noted that the numerical models provided a lower one-side exposure, applied load level of 6.7% of axial strength) was
axial deformation rate (in particular, in the protected column) than also predicted using Hashin damage criterion. The progressive da-
the experimental tests, and this was probably caused by the non- mage analysis overestimated the load capacity of the column.
consideration of the creep phenomenon. However, it allowed the correct identification of failure location and
• Regarding the transversal deflection, a movement towards the op- column collapse mode observed in the experimental tests. Taking
posite direction of heating source was observed, which was attrib- into account the several intricate phenomena involved but not
uted to the increasing eccentricity of the applied load with respect to considered in the current models (creep and viscosity, through-
the resistant cross-section. However, the numerical results did not thickness delamination) and the uncertainty of certain parameters
follow the experimental ones in the initial stage. The former showed (e.g., fracture energies), it can be affirmed that the current models

535
T. Morgado et al. Thin-Walled Structures 135 (2019) 521–536

show a good performance and promising potential to accurately [10] Y. Bai, E. Hugi, C. Ludwig, T. Keller, Fire performance of water-cooled GFRP col-
predict the complex behaviour of GFRP column under fire exposure. umns. I: fire endurance investigation, J. Compos. Constr. 15 (3) (2011) 404–412.
[11] International Standards Organization, Fire resistance tests. Elements of building
A wealth of additional experimental data will be needed to improve construction – Part 1: General requirements, 1999.
the accuracy of such models. [12] A.G. Gibson, Y.S. Wu, J.T. Evans, A.P. Mouritz, Laminate theory analysis of com-
posites under load in fire, J. Compos. Mater. 40 (7) (2006) 639–658.
[13] J.V. Bausano, J.J. Lesko, S.W. Case, Composite life under sustained compression
Bearing in mind that these models are able to qualitatively capture and one sided simulated fire exposure: characterization and prediction, Compos.
the general trend of the experimental results, the authors also notice Part A: Appl. Sci. Manuf. 37 (7) (2006) 1092–1100.
they have current limitations (quantitative differences with respect to [14] Y. Bai, T. Keller, Y. Vallée, Modeling of stiffness of FRP composites under elevated
and high temperatures, Compos. Sci. Technol. 68 (15–16) (2008) 3099–3106.
the experimental results). However, these models show promising fea- [15] P.K. Dutta, D. Hui, Creep rupture of a GFRP composite at elevated temperatures,
tures and hopefully will be valid to accurately predict the complex Comput. Struct. 76 (1) (2000) 153–161.
mechanical behaviour and the fire resistance of GFRP columns as soon [16] S.E. Boyd, S.W. Case, J.J. Lesko, Compression creep rupture behavior of a glass/
vinyl ester composite subject to isothermal and one-sided heat flux conditions,
as the effects of creep, delamination and fracture are implemented.
Compos. Part A: Appl. Sci. Manuf. 38 (6) (2007) 1462–1472.
[17] B. Budiansky, N.A. Fleck, Compressive failure of fiber composites, J. Mech. Phys.
Acknowledgments Solids 41 (1) (1993) 183–211.
[18] T. Morgado, J.R. Correia, A. Moreira, F.A. Branco, C. Tiago, Experimental study on
the fire resistance of GFRP pultruded tubular columns, Compos. Part B: Eng. 69
The authors wish to acknowledge Fundação para a Ciência e a (2015) 201–211.
Tecnologia (FCT) (project PTDC/ECM-EST/1882/2014) and Civil [19] Dassault Systèmes Simulia Corp, Commercial software ABAQUS Standard, version
Engineering Research and Innovation for Sustainability (CERIS) for 6.13, 2013.
[20] T. Morgado, N. Silvestre, J.R. Correia, F.A. Branco, T. Keller, Numerical modelling
funding the research. The first author also wishes to thank the financial of the thermal response of pultruded GFRP tubular profiles subjected to fire,
support of FCT through scholarship SFRH/BD/94907/2013. The second Compos. Part B: Eng. 137 (2018) 202–216.
author acknowledges the support of FCT, through IDMEC, under [21] I. Rosa, T. Morgado, J.R. Correia, J. Firmo, N. Silvestre, Shear behaviour of FRP
composite materials at elevated temperature, J. Compos. Constr. 22 (3) (2018).
LAETA, project UID/EMS/50022/2013. [22] T. Morgado, J.R. Correia, N. Silvestre, F.A. Branco, Experimental study on the fire
resistance of GFRP pultruded tubular beams, Compos. Part B: Eng. 139 (2018)
References 106–116.
[23] I. Rosa, Mechanical Behaviour of FRP Composite Materials at Elevated Temperature
(in Portuguese) (MSc dissertation in Civil Engineering), IST, Technical University of
[1] J.R. Correia, GFRP Pultruded Profiles in Civil Engineering: Hybrid Solutions, Lisbon, Lisbon, 2016.
Bonded Connections and FireBehaviour (Ph.D. Thesis in Civil Engineering), IST, [24] W. Sun, Temperature Effects on Material Properties and Structural Response of
Technical University of Lisbon, Lisbon, 2008. Polymer Matrix Composites (Ph.D. Thesis in Civil Engineering), École
[2] L.C. Bank, Composites for Construction: Structural Design with FRP Materials, John Polytechnique Fédérale de Lausanne, 2015.
Wiley & Sons, 2006. [25] L. Zhang, Y. Bai, W. Chen, F.-X. Ding, H. Fang, Thermal performance of modular
[3] J.R. Correia, Pultrusion of advanced composites, in: J. Bai (Ed.), Advanced Fibre- GFRP multicellular structures assembled with fire resistant panels, Compos. Struct.
Reinforced Polymer (FRP) Composites for Structural Applications, Woodhead 172 (2017) 22–33.
Publishing Limited, 2013. [26] L. Zhang, Y. Bai, Y. Qi, H. Fang, B. Wu, Post-fire mechanical performance of
[4] J.R. Correia, M.M. Gomes, J.M. Pires, F.A. Branco, Mechanical behaviour of pul- modular GFRP multicellular slabs with prefabricated fire resistant panels, Compos.
truded glass fibre reinforced polymer composites at elevated temperature: experi- Part B: Eng. 143 (2018) 55–67.
ments and model assessment, Compos. Struct. 98 (2013) 303–313. [27] ANSYS inc, Commercial software ANSYS Fluent, version 14.5, 2012.
[5] A.P. Mouritz, A.G. Gibson, Fire Properties of Polymer Composite Materials, [28] R.M. Jones, Mechanics of Composite Materials, 2nd edition, Taylor & Francis,
Springer, Netherlands, 2006. Philadelphia, PA, 1999.
[6] J.R. Correia, Y. Bai, T. Keller, A review of the fire behaviour of pultruded GFRP [29] Z. Hashin, Failure criteria for unidirectional fiber composites, J. Appl. Mech. 47
structural profiles for civil engineering applications, Compos. Struct. 127 (2015) (1980) 329–334.
267–287. [30] ANSYS inc, Commercial software ANSYS Fluent, version 14.5, 2012.
[7] P.M.H. Wong, J.M. Davies, Y.C. Wang, An experimental and numerical study of the [31] T. Morgado, N. Silvestre, J.R. Correia, Simulation of fire resistance behaviour of
behaviour of glass fibre reinforced plastics (GRP) short columns at elevated tem- pultruded GFRP beams – Part I: models description and kinematic issues, Compos.
peratures, Compos. Struct. 63 (2004) 33–43. Struct. 187 (2018) 269–280.
[8] P.M.H. Wong, Y.C. Wang, An experimental study of pultruded glass fibre reinforced [32] L. Almeida-Fernandes, N. Silvestre, J.R. CorreiaAssessment of fracture thoughness
plastics channel columns at elevated temperatures, Compos. Struct. 81 (1) (2007) of pultruded GFRP composites, in: Proceedings of the 8th International Conference
84–95. on Thin-Walled Structures, Lisbon, 2018.
[9] Y. Bai, T. Keller, Pultruded GFRP tubes with liquid-cooling system under combined
temperature and compressive loading, Compos. Struct. 90 (2) (2009) 115–121.

536

Das könnte Ihnen auch gefallen