Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Construction and Building Materials 262 (2020) 120924

Contents lists available at ScienceDirect

Construction and Building Materials


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

Deterioration of ambient-cured and heat-cured fly ash geopolymer


concrete by high temperature exposure and prediction of its residual
compressive strength
Hongen Zhang a,b,c, Lang Li a,c, Cheng Yuan b, Qingyuan Wang a,c,d,⇑, Prabir Kumar Sarker b,⇑,
Xiaoshuang Shi a,c
a
Key Laboratory of Deep Underground Science and Engineering (Ministry of Education), School of Architecture and Environment, Sichuan University, Chengdu 610065, China
b
School of Civil and Mechanical Engineering, Curtin University, GPO Box U1987, Perth, WA 6845, Australia
c
Failure Mechanics & Engineering Disaster Prevention and Mitigation Key Laboratory of Sichuan Province, College of Architecture & Environment, Sichuan University,
Chengdu 610065, PR China
d
School of Mechanical Engineering, Chengdu University, Chengdu 610106, China

h i g h l i g h t s

 Residual properties of fly ash geopolymer after elevated heat exposure is studied.
 Higher mass loss is found for ambient curing than for initial heat curing.
 Strength loss occurred only after 600 °C for 2 h exposure.
 Residual strength decreases due to increased crushing index of coarse aggregate.
 Proposed residual strength equations correlated well with the test data.

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

Article history: This paper presents an investigation on the performance of ambient-cured and heat-cured low-calcium
Received 12 November 2019 fly ash geopolymer concrete after exposure to elevated temperatures. The concrete specimens were
Received in revised form 21 August 2020 heated to 100, 200, 400, 600, 800 and 1000 °C at a heating rate of 5 °C/min. The effect of high temperature
Accepted 12 September 2020
exposure on geopolymer concrete was studied by visual inspection, mass loss, extent of cracking, residual
strength and microstructure investigation. The total length of cross-section cracks and surface cracks
reached their peak values at 800 °C and then declined at 1000 °C. The results show that all the concrete
Keywords:
specimens could be heated at 600 °C for 2 h without strength loss. Heat-cured geopolymer concrete spec-
Ambient curing
Heat curing
imens showed higher residual compressive strengths than the ambient-cured specimens for all the expo-
Geopolymer concrete sure temperatures. A crushing index of 7.7% could be regarded as the threshold value for coarse aggregate
Crushing index to maintain the initial compressive strength of concrete up to 600 °C. Thus, the microstructural damage
Coarse aggregate as observed by SEM images, dehydration of geopolymer shown by thermogravimetric analysis and the
Elevated temperature reduced strength of coarse aggregate are considered as the contributing factors for strength losses at tem-
Predicted residual strength peratures higher than 600 °C. Finally, two prediction equations are proposed which correlated well with
the experimental results of this study and those reported in published literature.
Ó 2020 Published by Elsevier Ltd.

1. Introduction properties, low cost, good fire resistance, versatility and


convenience of maintenance over other construction materials
The increase of global infrastructures is increasing the [1,2]. The traditional binder used for producing concrete is
demand of concrete due to its advantages of desirable mechanical ordinary Portland Cement (OPC). However, manufacturing
cement contributes to a huge amount of CO2 into the atmosphere,
⇑ Corresponding authors at: Sichuan University, No. 24 South Section 1, Yihuan causing a series of environmental issues such as global warming
Road, Chengdu 610065, China, (Q. Wang); Curtin University, GPO Box U1987, Perth, [3]. Therefore, it is essential to develop alternative binders using
WA 6845, Australia, (P.K. Sarker). industrial by-products to supplement OPC in production of
E-mail addresses: wangqy@scu.edu.cn (Q. Wang), P.Sarker@curtin.edu.au concrete.
(P.K. Sarker).

https://doi.org/10.1016/j.conbuildmat.2020.120924
0950-0618/Ó 2020 Published by Elsevier Ltd.
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

Geopolymer is a newly developing environmentally friendly simulated fire exposures. The authors prepared geopolymer con-
binder which is formed by the effective chemical reaction between cretes of different strength using different heat curing regimes
an alkali and aluminosilicate sources, such as fly ash [4,5], ground for the same mixture. Based on the previous research, two mix-
granulated blast furnace slag [6] and metakaolin [7]. The com- tures were used to prepare normal strength and high strength
monly used activator solutions are Na-based [8,9] or K-based alka- geopolymer concrete with both ambient curing and heat curing
lis [10]). Some recent studies showed that geopolymer concrete methods in this study.
could provide 25–70% reduction in greenhouse gas emissions In this paper, the deterioration of geopolymer concrete by expo-
depending on the ingredients and mixture proportions [11]. sure of elevated temperatures was investigated and explained in
Apart from the environmental issues, sustainability and dura- both macro-level and micro-level by using aggregate crushing
bility of concrete is also a concern. Consideration of the perfor- index, scanning electron microscopy (SEM) images and thermo-
mance of concrete at elevated temperatures is necessary when gravimetric analysis (TGA). SEM and TGA technology have made
evaluating the sustainability and durability of concrete. The valuable contributions to the development of geopolymer compos-
mechanical properties of OPC concrete deteriorate at elevated tem- ites [29–31]. An equation to predict the residual compressive
peratures, which is attributed to the physical and chemical changes strength of concrete after exposure to elevated temperatures is
as well as the incompatibility between binder and aggregate [12]. necessary for the fire design of concrete structures [32]. However,
The research works on OPC concrete [13,14] proposed the dissoci- very limited work has been carried out to propose an equation
ation temperatures of Ca(OH)2 and CaCO3 to be 300 to 600 °C and suitable for fly ash geopolymer concrete. Therefore, this paper
700 to 1000 °C, respectively, and their melting temperature has also proposed an equation to predict the residual compressive
between 1200 and 1350 °C. Some recent research investigated strength of geopolymer concrete using the test results of this study
the performance of geopolymer paste, mortar, concrete and com- and those reported by previous studies.
posites subjected to elevated temperatures [15–18]. The behavior
of high strength and normal strength metakaolin-based geopoly-
mer concrete subjected to elevated temperatures [19] or simulated 2. Experimental work
fire [20]was reported in literature. The effect of raw materials and
alkaline activators on the thermal properties of geopolymers was 2.1. Materials
also investigated [21,22].
Kong and Sanjayan [23,24] stated that the thermal incompati- The fly ash used for making geopolymer in this investigation
bility between geopolymer matrix and aggregate was mainly was low-calcium fly ash. The chemical compositions of fly ash
responsible for the strength loss and the size of coarse aggregate was determined by X-ray fluorescence (XRF) on a dry sample basis.
affected the performance of geopolymer concrete at elevated tem- The chemical compositions and loss on ignition (LOI) of the fly ash
peratures. Crystallization and phase changes caused the strength are given in Table 1. The bulk density, granule density and water
degradation of concrete subjected to elevated temperatures content of fly ash are given in Table 2.
[6,25]. The creation and progress of micro cracks also resulted in Locally available natural siliceous sand with particle size smal-
the strength deterioration of concrete when exposed to elevated ler than 4 mm and locally crushed basalt aggregate with particle
temperatures [6,26]. The gradual change in the pore structure after size between 4.75 mm and 22 mm were used as fine aggregate
exposure to elevated temperature was also analyzed quantitatively and coarse aggregates, respectively. The main composition of sand
by CT technology [5,6]. Overall, most of the previous research is SiO2. The bulk density, granule density and water content of
focused on explaining the mechanism of strength deterioration in aggregates are presented in Table 2.
the micro-level. The alkali-activator solution was composed of commercially
Crushing index is a mechanical parameter used to measure the available liquid sodium silicate (Na2SiO3) with a modulus ratio
strength of concrete coarse aggregate [27]. As aggregates occupy (Ms) of 3.23 (where Ms = SiO2/Na2O, Na2O = 8.83%, SiO2 = 27.64%
about 75% to 80% volume of concrete, the change in crushing index by mass), and sodium hydroxide (NaOH) solution. The density of
of coarse aggregate would also affect the mechanical properties of Na2SiO3 solution at 20 °C is 1.384 g/cm3. Sodium hydroxide (NaOH)
concrete. Zhang et al. [28] investigated the effect of crushing index particles of 98% purity were dissolved in distilled water to make
on properties of recycled concrete. The results demonstrated that sodium hydroxide (NaOH) solution. The mass of NaOH particles
mechanical properties (compressive strength, flexural strength depends on the concentration of the solution expressed in terms
and modulus of elasticity) decreased with the increase of crushing of molarity. For example, NaOH solution with a concentration of
index. However, the influence of crushing index of coarse aggre- 8 M consists of 8  40 = 320 g/L of NaOH particles, where 40 is
gate on compressive strength of concrete (especially geopolymer the molecular weight of NaOH. The density of NaOH solution with
concrete) subjected to elevated temperatures has never been a concentration of 8 M and 14 M at 20 °C is 1.275 g/cm3 and
reported. Hence this investigation was conducted with one aim 1.425 g/cm3, respectively. The alkali-activator solution was pre-
to determine how the crushing index of coarse aggregate affected pared approximately 24 h prior to use.
the geopolymer concrete’s compressive strength after exposure
to elevated temperatures. The main aim of this research is to inves- 2.2. Specimen preparation and curing regime
tigate the mechanism of strength degradation at elevated temper-
atures. The crushing index of coarse aggregate was tested in all the 2.2.1. Mixture proportions
exposure temperatures except for 100 and 200 °C. Two sets of concrete mixtures used in this investigation are pre-
Sarker et al. [20] investigated the cracking, spalling and residual sented in Table 3. The mixtures are named in the form of ‘‘GPN-X”
strength of normal and high strength geopolymer concrete after or ‘‘GPH-X”, where ‘‘GP” meant geopolymer concrete, ‘‘N”

Table 1
Chemical compositions of fly ash.

Chemical composition SiO2 Al2O3 CaO Fe2O3 MgO K2O SO3 TiO2 Na2O ZnO LOIy
Fly ash (mass, %) 65.30 14.48 5.63 6.25 0.93 1.46 0.09 1.08 1.14 0.07 0.82
y
Loss on ignition.

2
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

Table 2 for about 24 h, the other batch of specimens were cured at ambient
The density and water content of fly ash and aggregates. temperature for about 72 h because of the long setting time
Materials Granule density Bulk density Water content required for these specimens [1]. After 24 h or 72 h, the specimens
(g/cm3) (g/cm3) were demoulded and cured in a controlled condition (20 °C and
Fly ash 2.6 0.6 0.5% 65% RH) until tested.
Coarse aggregate 2.6 1.5 0.3%
Fine aggregate 2.4 1.3 0.1%
2.3. Preparing samples for detecting the concrete internal structure
deterioration

represents the normal strength concrete series, ‘‘H” represents the In order to analyze deterioration of the internal structure, the
high strength concrete series and ‘‘X” is the curing condition. For concrete specimens were cut into three slices by a concrete cutting
example, ‘‘GPN-A” means ambient-cured normal strength geopoly- machine. The specimens were cut into pieces before temperature
mer concrete and ‘‘GPN-H” means heat-cured normal strength exposure in order to avoid the difficulty of cutting them after tem-
geopolymer concrete. Additionally, it is worthwhile to mention perature exposure since the specimens might suffer from explosive
that the water content shown in the table is calculated from the spalling or severe strength loss after exposure to 800 °C or 1000 °C.
water in the alkaline liquid and no extra water was added during The method consisted of two steps. Firstly, before exposure to the
the whole casting process. target temperature, the concrete specimen was cut into three
slices. Secondly, the three concrete slices were then covered by a
2.2.2. Casting and curing of test specimens layer of mortar with less than 2 mm thickness to be an artificial
The mixing of the concrete was conducted in a laboratory con- integrated concrete block (AICB). The second step could prevent
crete mixer. For the mixing procedure, all dry materials (aggregates the cross-section of concrete from being exposed to elevated tem-
and fly ash) were mixed in the pan mixer for about 3 min, then the perature directly. This method is also intuitively illustrated in
alkali-activator solution was sequentially added and the wet- Fig. 1. The mortar mixture proportions used in the second step is
mixing continued for another 4–6 min until a consistent mixture also presented in Table 3 Mixture proportions and curing condition
was obtained. The fresh concrete was poured into steel moulds of concrete and mortar. The AICB was then subjected to the target
with 100  100  100 mm to prepare cube specimens. The moulds temperature.
were filled in two layers and each layer was vibrated for about 30 s
using a vibrating table. 2.4. Heating details
Right after casting, the specimens along with the moulds were
covered with plastic film to prevent moisture loss. For each set of An electrical furnace designed for a maximum temperature of
mixture, one batch of specimens were cured in an oven at 80 °C 1600 °C was used in this study. The specimens were exposed to

Table 3
Mixture proportions and curing condition of concrete and mortar.

Mixture design of the concrete or mortar (kg/m3)


y à § –
Mix ID CA Sand Fly ash SS SH SS/SH* M Curing temperature Water content
GPN-A 1201 539 460 142.9 57.1 2.5 8 25 °C 5.54%
GPN-H 1201 539 460 142.9 57.1 2.5 8 80 °C 5.54%
GPH-A 1201 539 460 150 50 3.0 14 25 °C 5.21%
GPH-H 1201 539 460 150 50 3.0 14 80 °C 5.21%
Mortar 0 1380 460 150 50 3.0 14 40 °C –
y
coarse aggregate.
à
Na2SiO3 solution.
§
NaOH solution.
*
the mass ratio of Na2SiO3 solution to NaOH solution.

the molarity of sodium hydroxide solution.

Fig. 1. The method to prepare the artificial integrated concrete block (AICB).

3
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

elevated temperatures after the age of 90 days. The heating regime 2.5.3. Image J analysis
used in this study is presented in Fig. 2. The cube specimens were In order to carry out the analysis of crack geometry attribute
subjected to the peak temperatures of 100, 200, 400, 600, 800 and more intuitively, a free software ‘‘Image J” was utilized. The soft-
1000 °C at an incremental rate of 5 °C/min from ambient temper- ware was initially developed by the National Institute of Mental
ature (Stage I). The concrete was then held at each peak elevated Health. Fares et al. [33] investigated the porosity of the fraction
temperature for 2 h to ensure a uniform temperature within the of anhydrous elements of self-consolidating concrete using Image
specimens (Stage II). Then the specimens were left in the furnace J and the result was remarkable.
to naturally cool down to ambient temperature (Stage III).
2.5.4. Analyzing the cracks
2.5. Test methods MATLAB software was used to analyze and calculate the length
of crack on the surface and cross-section of concrete specimens
2.5.1. Mass loss test with the help of a local-binarization method provided by Li et al.
The samples exposed to each target temperature were also used [34]. At first, the cross-sections and surfaces of concrete were
for the determination of mass loss ratio of fly ash-based geopoly- transferred to grey-scale images with a local-binarization algo-
mer concrete (FAGC) using Eq. (1). The mass loss ratio value was rithm and then the total length of cracks was calculated. In this
an average of three specimens. paper, to clearly distinguish between the cracks on surface and
those in the cross-section, the cracks on the surface were defined
Mb  Ma
R¼  100% ð1Þ as EXTR-cracks (external cracks) and the cracks in the cross-
Mb
section were defined as INTR-cracks (internal cracks).
where R was the mass loss ratio and Mb referred to mass of the sam-
ple before exposure to the target temperature and Ma represents 2.5.5. Compressive strength test
mass of the sample after exposure to the temperature. The compressive strength test was conducted in accordance
with GB/T50081-2002 [35] using a YAW-2000 electro-hydraulic
2.5.2. Crushing index of coarse aggregate mechanical testing machine at a load rate of 0.5 MPa/s until the
Crushing index is a mechanical index used to measure the concrete specimens were crushed. This method was also used for
strength of concrete coarse aggregate. Kong and Sanjayan [24] testing the residual compressive strength of specimens after expo-
demonstrated the expansion behavior of aggregate under elevated sure to elevated temperatures.
temperature. Therefore, the crushing index of coarse aggregate will
also be effected after exposure to elevated temperature. Crushing 2.5.6. Scanning electron microscopic images
index of coarse aggregate was tested in accordance with GB/T Scanning electron microscopy (SEM) test was conducted using
14685-2011[27]. Three kilograms of coarse aggregate within parti- SU3500 (HITACHI, Kyoto, Japan) instrument to characterize the
cle size between 9.5 and 19 mm were used to conduct the crushing microstructures of the FAGC samples.
index test using YAW-2000 electro-hydraulic mechanical testing
machine at a load rate of 1 kN/s until the maximum load reached 2.5.7. Thermogravimetric analysis (TGA)
200kN. The maximum load was held constant for 5 s and then The aggregates were removed when preparing samples for the
unloaded to zero. Then a sieve shaker was used to obtain the resid- thermogravimetric analysis (TGA) test. Fragments of FAGC speci-
ual mass of the aggregate with particle size larger than 2.36 mm. mens unexposed to elevated temperatures were powdered to con-
The crushing index is calculated using Eq. (2). duct the TGA test through a METTLER TOLEDO TGA/DSC2/1600
G1  G2 (METTLER TOLEDO, Zurich, Switzerland) instrument in alumina
Qe ¼  100% ð2Þ crucibles in the nitrogen environment (N2 flowing at 50 mL/min),
G1
and then elevated to 1000 °C with a constant heating rate of
where Q e is the crushing index (%), G1 is the mass of coarse aggre- 10 °C/min in the same gas environment.
gate (g) and G2 is the residual mass of coarse aggregate with the
particle size lager than 2.36 mm (g).
3. Results and discussion

3.1. Residual compressive strength


1100 Stage II
Stage I 2h The compressive strengths of GPN-A, GPN-H, GPH-A and GPH-H
1000
concrete specimens before and after exposure to elevated temper-
900 atures are given in Table 4. The percentage residual strengths at
2h
800 each target temperature are also presented in Table 4. It is clear
Temperatures (°C)

from Table 4 that all the geopolymer concretes showed the highest
in

Stage III
/m

700
compressive strength value at 200 °C. At 200 °C, the compressive
°C

2h
t5

600 strength of GPN-A, GPN-H, GPH-A and GPH-H were 58.0 MPa,
ea

Nat ent tem

71.6 MPa, 69.9 MPa and 89.9 MPa, respectively. Taking GPN-H
rat

am
rua

500
ng

bi

and GPN-A geopolymer concrete for consideration, compressive


l co
ati

2h
400
He

strength of ambient-cured geopolymer concrete was lower than


olin tures

that of heat-cured specimens during the temperature range of


g to

300
per

2h 20-1000 °C. The similar trend was also shown by GPH-H and
200 GPH-A specimens.
2h At 400 °C, the percentage residual strengths of GPN-A and GPN-
100
20 H were 163.9% and 162.2%, respectively, indicating that ambient-
Time (h) cured geopolymer concrete showed marginally better strength
enhancement than oven-cured geopolymer concrete after expo-
Fig. 2. The heating regime. sure to 400 °C. The effect of curing method is more pronounced
4
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

Table 4
Compressive strength and percentage residual strength after heat exposure.

Temp. (°C) GPN-A GPN-H GPH-A GPH-H


y y y
Compr. Strength (MPa) Ratio (%) Compr. Strength (MPa) Ratio (%) Compr. Strength (MPa) Ratio (%) Compr. Strength (MPa) Ratioy (%)
20 35.2  1.2 100.0 39.40.6 100.0 43.22.4 100.0 72.82.3 100.0
100 52.0  0.7 147.7 63.5  5.6 161.2 63.9 1.43 147.9 87.6  4.2 120.3
200 58.0 2.5 164.7 71.61.0 181.7 69.92.1 161.8 89.73.3 123.2
400 54.8 2.1 163.9 63.90.8 162.2 57.96.0 134.0 80.92.4 111.1
600 46.0 2.9 130.7 50.04.3 126.9 49.23.9 113.9 72.32.2 99.3
800 17.4  0.4 49.4 22.23.4 56.3 21.7  1.9 50.2 23.12.8 31.7
1000 15.3 9.5 43.5 19.5 7.8 49.5 15.35.8 35.4 15.03.8 20.6
y
The ratio of residual compressive strength to original compressive strength.

in the percentage residual strengths of GPH-A (134.0%) and GPH-H similar to the trend of change in compressive strength. The ratio of
(111.0%). GPN-H geopolymer concrete specimens was larger than that of
It can be observed from Table 4 that the compressive strengths GPH-H specimens during all the temperature range, illustrating
of GPN-A, GPN-H and GPH-A concrete at 600 °C were still higher that normal strength geopolymer concretes experienced higher
than their original compressive strength. As for GPH-H, its residual strength growth ratio and lower strength loss ratio than high
strength at 600 °C was 72.3 MPa which was also almost equal to its
original compressive (72.8 MPa). Therefore, it could be concluded
that both ambient-cured and heat-cured geopolymer concrete

Ratio of residual compressive strength to original


could be kept at 600 °C for 2 h without loss of compressive 200
strength.
180
At 800 °C, the residual compressive strengths of the four series
of concrete were smaller than the corresponding initial compres- 160
compressive strength (%)
sive strength. The compressive strength of GPH-H even dropped
from 72.3 MPa (at 600 °C) to 23.1 MPa (at 800 °C). 140
At 1000 °C, the compressive strength of all concrete mixtures 120
was smaller than 20 MPa. The percentage residual strengths of
GPN-A, GPN-H, GPH-A and GPH-H were 43.5%, 49.5%, 31.9% and 100
20.6%, respectively.
80
Fig. 3 shows the compressive strength of different concretes
after exposure to the temperature range of 20–1000 °C. Overall, 60
the compressive strength change could be divided into two stages. GPN-A GPN-H
In the first stage, compressive strength increased with the increase 40 GPH-A GPH-H
of temperature and achieved its maximum value at 200 °C. The 20
compressive strength began to decline with different strength loss
rates between 200 and 1000 °C, which was the second stage. In the 0 100 200 300 400 500 600 700 800 900 1000 110
decline stage, the compressive strength appeared a moderate Temperatures (°C )
decrease from 200 to 600 °C and a significant decline from 600
to 800 °C, and then reduced slowly between 800 and 1000 °C. Fig. 4. The ratios of residual to original compressive strength at different elevated
temperatures.
The ratio of residual strength to original strength is also illustrated
in Fig. 4. It was observed that the trend of change in the ratio was

6.0

90 22 5.5
Crushing index of coarse aggregate (%)

5.0
80 20
4.5 Water content of GPN:5.54% 5.09±0.11 %
18
MPa

70
4.0 Water content of GPH:5.21%
Mass loss (%)

60 16
3.5
Compressive strength

14
50 3.0
10.7% 12
40 2.5
GPN-A GPN-H 10
30 2.0
GPH-A GPH-H
8 1.5
20 crushing index GPN-A GPN-H
6 1.0 GPH-A GPH-H
10
7.7%
4 0.5
0
0 100 200 300 400 500 600 700 800 900 1000 1100 0.0
0 100 200 300 400 500 600 700 800 900 1000
Temperatures (°C )
Temperature (°C)
Fig. 3. Compressive strength and crushing index of coarse aggregate at elevated
temperatures. Fig. 5. The mass loss of geopolymer concrete at elevated temperatures.

5
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

strength specimens. It was found that the percentage residual 100 °C and 200 °C, therefore, the broken line was used. It is
strength of GPH-A was higher than that of GPH-H during all the observed from Fig. 3 that the crushing index increased with differ-
temperatures, indicating that ambient-cured geopolymer concrete ent ratios during the various exposure temperatures. As discussed
performed better compressive strength enhancement and lower in section 3.1, low-calcium fly ash-based geopolymer concrete
strength deterioration than the heat-cured ones. could retain its initial compressive strength at 600 °C. Therefore,
the coarse aggregate with a crushing index of 7.7% (Fig. 3) was still
3.2. Relationship between crushing index of coarse aggregate and strong enough to work as the critical part of concrete skeleton. In
residual compressive strength other words, a crushing index of 7.7% could be regarded as the
threshold value for coarse aggregate to maintain its good strength.
Coarse aggregates play a vital role in forming the concrete With the crushing index rising to 10.7% (at 800 °C), the coarse
skeleton. Aggregate deterioration could lead to the deterioration aggregate lost most of its load capacity resulting in a significant
of concrete skeleton, which would have a detrimental impact on loss of compressive strength. After exposure to 1000 °C, the crush-
the concrete structures causing the strength loss. The results of ing index increased further to 21.1% with little influence on com-
crushing index before and after exposure to temperatures are also pressive strength change, which will be further discussed in the
illustrated in Fig. 3. The result of crushing index was not tested at following section.

Fig. 6. Surface images of geopolymer concrete at different temperatures.

Fig. 7. 8-bit grey-scale surface images of geopolymer concrete at different temperatures.

6
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

3.3. Mass loss of concrete after exposure to elevated temperatures concretes were higher than those of GPN-H specimens during the
temperature range of 100 to1000 °C, illustrating that the high
The mass loss of concrete samples resulting from the exposure strength geopolymer concretes experienced higher mass loss than
to heat was calculated through weighing the mass of specimens the normal strength concrete at elevated temperatures. At 400 °C,
before and after the target temperature exposure. The mass loss the mass loss values of the GPN-H, GPN-A, GPH-H and GPH-A sam-
values of specimens are presented in Fig. 5. It can be seen that ples were 2.32%, 2.61%, 2.82% and 3.35%, respectively. During the
ambient-cured and heat-cured geopolymer concretes presented temperature range of 600 to1000 °C, the rate of mass loss slowed
similar mass loss trend. There was an increasing trend in mass loss down in concrete samples. The mass loss values of GPN-H, GPN-
with the increasing elevated temperatures. The mass loss is gener- A, GPH-H and GPH-A at 1000 °C were 3.48%, 4.19%, 4.15% and
ally attributed to the loss of free water or bonded water according 5.09%, respectively.
to the exposure temperatures [5,6,30]. Up to 200 °C, the mass loss Overall, the high strength concrete samples suffered from
was associate with the loss of free water and partly chemically higher mass loss than the normal strength concrete samples
bonded water resulting from the dehydration of geopolymer gel. (Fig. 5). This is consistent with the results of previous study on
Beyond 200 °C, the mass loss was mainly resulted from further fly ash-based geopolymer concrete after fire exposure [20]. The
dehydration of geopolymer gel and dehydroxylation of hydroxyl maximum mass loss was smaller than the initial water content
groups [36]. Additionally, the concrete specimens cured at ambient in the concrete and the GPH-A sample suffered the highest mass
temperature showed higher mass losses than the heat-cured spec- loss value (5.09 ± 0.11%) which was very close to its initial water
imens. It was found that mass loss values of GPH-H geopolymer content (5.21%).

3.4. The change in surface of concrete specimens

The surface images of the concrete specimens before and after


exposure to different temperatures are presented in Fig. 6. The con-
crete samples showed gray or gray-white color at ambient temper-
ature. Up to 400 °C, no noticeable color difference was observed for
all the concrete samples except for GPH-A sample that presented a
light brown color at 400 °C. The color change in GPH-A at 400 °C
might be due to its highest mass loss. From the beginning of
600 °C, there was obvious color difference from the original color.
At 800 °C and 1000 °C, the concrete specimens displayed red color
which is attributed to the high iron oxide content of the fly ash
[17]. A similar change in color in high and normal strength con-
crete subjected to simulated fire was also observed by Sarker
et al. [20].
The 8-bit grey-scale images of surface obtained from the Image
J analysis software are presented in Fig. 7.The RGB color images
(Fig. 6) were converted to 8-bit grey-scale images using threshold
segmentation method. It is worthy to mention that the threshold
Fig. 8. Total length of EXTR-cracks of geopolymer concrete at elevated value of each image was not always identical, especially for the
temperatures. images at 800 and 1000 °C. It could be observed from Fig. 7 that

Fig. 9. Cross-section images of geopolymer concrete at different temperatures.

7
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

Fig. 10. 8-bit grey-scale cross-section images of geopolymer concrete at different temperatures.

there were several visual initial cracks on the surface of GPN-A 3.5. The change in the cross-section of concrete specimens
concrete specimens before exposure to elevated temperatures. This
is due to the drying shrinkage of concrete. Ridtirud et al. [37] and The original images of cross-section are presented in Fig. 9. It
Jiang et al. [38] proved that the drying shrinkage was more signif- can be seen that the color of cross-section showed negligible
icant for geopolymer materials cured at ambient temperature. No change before exposure to 400 °C. The color of coarse aggregate
obvious cracks were observed in the surfaces of GPH-A series turned from grey at 600 °C to white or brown at 800 to 1000 °C.
because the NaOH concentration in this series was up to 14 M. This is different from a previous observation by Li et al. [6] in which
Singh et al. [26] found that a higher activator concentration the color of coarse aggregate was reported to change from blue to
(14 M) released more heat than for lower activator concentration red. The difference in color change of aggregate is largely attribu-
(8 M) during geopolymerization. The released energy would pro- ted to the composition and type of aggregate [39,40]. The color
mote the geopolymerization. Therefore, 14 M-concentration NaOH change of concrete cross-section is a combined result of those of
solution could provide more appropriate temperature environ- geopolymer, sand and coarse aggregate where coarse aggregate
ment as compared to 8 M, which resulted in a more completed played the critical role because of its higher volumetric ratio [6].
geopolymerization in GPH-A series. The 8-bit grey-scale images of cross-section are illustrated in
It is observed from Fig. 7 that cracks started to become larger Fig. 10. It can be seen that the INTR-cracks were distributed in
(for GPN-A specimen) or to appear (for the other concrete speci- the cross-section of GPN-A sample at ambient temperature, indi-
mens) on the concrete surface after exposure to 200 °C. The cating the original structural defects in GPN-A samples. The
amounts of cracks and width of cracks increased with the elevated
temperature during the temperature range of 200–800 °C. For all
the four series, the amounts of cracks were most significant at 1000
800 °C. For exposure to 800–1000 °C, the amount of cracks dis-
Total length of INTR-cracks (mm)

played a decreasing trend. More impressive, seldom visual cracks


could be observed on the surface of GPH-H, while it is easy to find 800
several significant bulges on its surface at 1000 °C.
The length of EXTR-cracks at the target elevated temperatures
was calculated using MATLAB software and the results are pre-
600
sented in Fig. 8. It can be seen that the total length of EXTR- GPN-A GPN-H
cracks of GPN-A is higher than the other three samples before GPH-A GPH-H
600 °C. During the temperature range of 20–600 °C, the length of
400
EXTR-cracks showed a slowly increasing trend for the four sam-
ples. The length of EXTR-cracks of ambient-cured specimens
(GPN-A and GPH-A) increased dramatically in the range of
200
600~800 °C and reached its maximum value at 800 °C. At the same
time, the length of EXTR-cracks of heat-cured specimens (GPN-H
and GPH-H) presented a moderate growth in the length of EXTR-
0
cracks between 600 and 800 °C and reached its peak value at 0 100 200 300 400 500 600 700 800 900 1000 1100
800 °C. Beyond 800 °C, the EXTR-cracks of all the four series con-
ºC
crete declined at different ratios. At 1000 °C, the length value of
the EXTR-cracks of GPN-A was the largest, followed by GPN-H, Fig. 11. Total length of INTR-cracks of geopolymer concrete at elevated
GPH-A and GPH-H. temperatures.

8
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

amount of cracks can be seen to decrease after exposure to 200 °C The total length of INTR-cracks is also illustrated in Fig. 11. It
which is an evidence of the secondary geopolymerization of fly ash can be seen that the initial length of INTR-cracks of GPN-A is much
geopolymers after exposure to high temperature. higher than those of the specimens of other series at ambient tem-

ºC ºC

ºC ºC

ºC ºC

ºC ºC
Fig. 12. SEM micrographs of geopolymer concrete at 20 °C.

9
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

perature, which indicates that there were many defects in the con- 3.6. Microstructure investigation
crete internal macro-structure. As a result, the compressive
strength of GPN-A series was the lowest among the four concrete The SEM images of the samples all four series at 20 °C are
series. Before 600 °C, the length of INTR-cracks of GPN-A decreased shown in Fig. 12 at two different magnifications. All the images
to the bottom at 200 °C resulting from secondary geopolymeriza- display heterogeneous microstructures containing unreacted fly
tion, and then increased with the temperature. As for GPN-H, ash particles of different sizes. Micro-cracks can be seen in
GPH-A and GPH-H, the length of INTR-cracks increased rather Fig. 12 (a) and (b), which were responsible for the lower strength
slowly and the values were also much smaller than that of GPN- of GPN-A samples at ambient temperature. The micro-cracks were
A in the range of 20–600 °C. Contrary to the behavior of EXTR- believed to be caused by the drying shrinkage. It can be seen from
cracks between 600 and 800 °C, the length of INTR-cracks of Fig. 12 (a) to (d) that the microstructures of GPN-A and GPN-H
heat-cured specimens (GPN-H and GPH-H) increased dramatically samples are less cohesive than those of GPH-A and GPH-H samples
and that of ambient-cured specimens (GPN-A and GPH-A). The (Fig. 12 (e) to (h)), which explains the lower compressive strengths
same with EXTR-cracks, the peak values of INTR-cracks were still of GPN-A and GPN-H samples. Compared with the GPH-H sample,
appeared at 800 °C and the total length of the INTR-cracks of the GPH-A sample has a larger proportion of unreacted fly ash par-
GPN-A was also the largest followed by GPN-H, GPH-A and GPH- ticles, therefore, GPH-A samples showed lower compressive
H after exposure to 1000 °C. strength than the GPH-H samples.

ºC ºC

ºC ºC

ºC ºC
Fig. 13. SEM micrographs of GPH-H group geopolymer concrete at different elevated temperatures.

10
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

101 the amounts of unreacted fly ash particles after exposure to


800 °C. Additionally, after exposure to 800 °C, several micro-
100 GPN-H GPN-A
cracks are also observed in the sample. The micro-cracks (marked
99 GPH-H GPH-A as thermal cracks in Fig. 13 (c) and (d)) were believed to be induced
98 by the thermal stress. A lot of small pores and areas of high poros-
ºC ity were observed in the sample after exposure to 800 °C (Fig. 13
97 ºC (d)), indicating the low cohesion of geopolymer matrix. Further
Mass (%)

96 increasing the temperature up to 1000 °C, no unreacted fly ash par-


ticles but amounts of large pores were observed (Fig. 13 (e) and
95
(f)). The pores were possibly due to the collapse of the geopolymer
94 matrix induced by the decomposition of geopolymer [18] and
93 phase transformations [41]. Though the SEM shows a high number
of large pores after exposure to 1000 °C, the microstructure has
92 ºC fewer impurities and it reveals a smoother texture. The smooth
91 texture is believed to be due to the melting, viscous flow and sin-
tering of unreacted fly ash particles and amorphous aluminosili-
90
0 100 200 300 400 500 600 700 800 900 1000 cate gels during heating at a high temperature of 1000 °C.

ºC
3.7. Thermogravimetric analysis
Fig. 14. The TGA curves of geopolymer samples.
The TGA curves of geopolymer samples are illustrated in Fig. 14.
The derivative thermogravimetry (DTG) and differential scanning
The SEM images of GPH-H sample at 200 °C, 800 °C and 1000 °C calorimetry (DSC) curves of geopolymer concrete samples are plot-
are presented in Fig. 13. A flat and intact surface was detected at ted in Fig. 15. The mass change trend in Fig. 14 is found to be in line
200 °C. It can be seen that there was a significant reduction in with that in Fig. 5. It can be seen from Fig. 14 that geopolymer con-

Fig. 15. The DTG and DSC curves of geopolymer samples.

11
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

crete experienced four characteristic stages of mass loss. Stage I DTG curves and an endothermic peak was found in the DSC curves,
involved evaporation of the free water and part of chemically which was because of the dehydration of chemically bound water.
bound water up to 200 °C. The mass loss occurred in stage I The evaporation of free water and dehydration of the chemically
accounted for more than 50% of the total mass loss. Stage II is char- bound water was mainly responsible for the mass loss of geopoly-
acterized by the loss of chemically bound water and dehydroxyla- mer concrete up to about 200 °C. At about 950 °C, an endothermic
tion of hydroxyl groups, which occurred between 200 and 700 °C. peak was observed in the DSC curves, which was due to the melt-
The mass in stage III almost remained unchanged from 700 °C to ing of unreacted fly ash particles and geopolymer matrix. There-
950 °C. The last stage is stage IV after 950 °C where the mass loss fore, no fly ash particles were observed and large pores were
went on to increase slightly. observed instead in Fig. 13 (e) and (f).
The thermal behavior of geopolymer concrete after exposure to
elevated temperatures up to 1000 °C is an endothermic process
accompanied by exothermic process at some specific temperature, 4. Discussions
which could be obtained from the DSC curves shown in Fig. 15. It
can be seen from Fig. 15 that the peak value of mass loss rate (in Comparing Figs. 8 and 11, it was evident that the total length of
DTG curves) and the peak value of endothermic peak and exother- EXTR-cracks was much higher than the total length of INTR-cracks.
mic peak (in DSC curves) generally occurred at the same tempera- Therefore, the surface of concrete specimens suffered from much
ture. The first mass loss rate peak occurred at the temperature more deterioration than the internal structure even though the
ranging from about 43.4 °C in GPN-H sample (Fig. 15 (b)) to about concrete reached thermal equilibrium. The temperature gradient
66.8 °C in GPH-A sample (Fig. 15 (b)), which was believed to be the and gradual heat exchange that occurred during the cooling period
evaporation of free water. The different temperature values of the is considered as the main reason for this difference. It was found
first mass loss rate peak implied the different amount of free water that the length of EXTR-cracks of ambient-cured geopolymer con-
in different geopolymer concrete specimens. The higher the tem- crete was longer than that of heat-cured specimens after exposure
perature at the first peak of mass loss rate was, the larger amount to 800 °C. On the contrary, the length of INTR-cracks of heat-cured
of free water there was in geopolymer concrete. Therefore, more geopolymer concrete was longer than that of ambient-cured spec-
free water existed in the ambient-cured geopolymer concrete spec- imens after exposure to 800 °C.
imens. As a result, ambient-cured geopolymer concrete experi- Generally, the cracks (micro-cracks and macro-cracks) are
enced higher mass loss. believed to be the result of the thermal incompatibility between
At the temperature of around 75 °C (79.5 °C in Fig. 15 (c)), the the geopolymer matrix and aggregate [17,42], pore pressure and
second mass loss rate peak in the DTG curves and the exothermic temperature gradient [6]. In this paper, as concrete cured at
peak in the DSC curves were owing to the crystallization of 20 °C had a relatively less dense structure than the one cured at
geopolymer product. Some free water became the chemically 80 °C, water (free water, capillary water and chemically bonded
bound water during the crystallization, resulting in the decrease water) could move to the surface of the specimens more easily
of mass loss rate. For example, the maximum mass loss rate of causing more cracks in the concrete surface because of vapor pres-
GPN-A samples (Fig. 15 (a)) declined from 5.2  104 at 61.4 °C sure. Fig. 9 illustrated that most of the INTR-cracks occurred
to 2.8  104 at 74.8 °C. However, the exothermic peaks in GPN- around the coarse aggregates. Therefore, the evaporation (release)
H and GPH-H samples (Fig. 15 (b) and (d)) were not as obvious of water (free water, capillary water and chemically bonded water)
as those in GPN-A and GPH-A samples (Fig. 15 (a) and (c)). This is considered as the main reason to cause the cracks on the surface
is because that GPN-H and GPH-H samples were cured at 80 °C. and the thermal incompatibility between geopolymer matrix and
At about 87 °C, the third mass loss rate peak was observed in the aggregate as the main reason for the INTR-cracks generation.

Fig. 16. The schematic diagram of concrete degradation between 200 and 600 °C.

12
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

Fig. 17. Thermal macro-cracks on part of cross-section at elevated temperatures.

Fig. 18. Failure patterns at different temperatures.

13
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

It can be seen from Fig. 11 that the length of INTR-cracks was At 800 °C, the length of EXTR- and INTR-cracks reached their
nearly negligible up to 400 °C, except for GPN-A sample, which peak values, indicating the concrete structure suffered severe dam-
demonstrated that the internal macrostructure kept relatively age. Additionally, the coarse aggregate lost most of its load capac-
stable at 400 °C. The sustainable shrinkage of fly ash-based ity due to its high crushing index value at this temperature.
geopolymer matrix [23,43] and the sustainable expansion of coarse Therefore, the compressive strength dropped sharply after expo-
aggregate [23] occurred after 200 °C due to the difference in their sure to 800 °C. At 1000 °C, the length of EXTR- and INTR-cracks
thermal behavior. The theory that most of the shrinkage of slag- decreased. It appears that the concrete showed a self-healing abil-
based geopolymer is irreversible and most of expansion of aggre- ity which is related to the shrinkage, sintering and melting of
gates is reversible up to 750 °C [6] was applied for low-calcium geopolymer matrix. Previous research demonstrated that geopoly-
fly ash-based geopolymer concrete in the current study. In general, mer made from slag [44] or low-calcium fly ash [45] experienced
the expansion of coarse aggregate dominated over the shrinkage of an obvious compressive strength increase at temperatures higher
geopolymer matrix in concrete because of its larger volume than than 800 °C. Rickard et al. [45] argued that the strength increase
geopolymer. As a result, a little gap (micro-cracks or macro- at temperature higher than 800 °C was attributed to the sintering
cracks) forms between geopolymer matrix and coarse aggregate. and melting. However, the concrete strength still decreased slowly
That is how the thermal incompatibility affect the properties of because the coarse aggregate had lost most of its load carrying
concrete subjected to elevated temperatures. However, naked capacity at 1000 °C in which the crushing index of coarse aggregate
cracks were observed seldom in the cross-section (except for was 21.1%. As aggregate occupied about 73% of the total weight,
GPN-A) before 600 °C (Fig. 9), thus, the thermal incompatibility the strength of aggregate played a decisive role in the compressive
between aggregate and matrix resulted in micro-cracks. The ther- strength of concrete at this stage.
mal incompatibility would have a detrimental influence on the Figs. 17 and 18 further show that the loss of compressive
mechanical behavior of interfacial transition zone (Fig. 16) result- strength occurred at temperatures higher than 800 °C is due to
ing in the strength loss between 200 and 600 °C. However, the the strength loss of coarse aggregate. The terms (radial crack, inclu-
cracks and the crushing index value of coarse aggregate were not sion crack and tangential crack in Fig. 17) used for analyzing micro-
noticeable enough to damage the concrete structure. As a result, cracks in cement-based mortar [46] were cited to distinguish dif-
geopolymer concrete samples retained the initial strength at ferent types of cracks in geopolymer concrete. However, they are
600 °C. used here to define cracks in macro-level. The appearance of inclu-

Fig. 19. Experimental results and fitting curves of residual compressive strength versus elevated temperature.

14
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

sion cracks indicated that the coarse aggregate lost its ability to 7 MPa. Based on these results, it is expected that the concrete
stand the thermal stress at 800 and 1000 °C, which was consistent strength should be at least higher than 43.2 MPa at 800 and
with the results of crushing index (Fig. 3). 1000 °C because the behavior of coarse aggregates in Fig. 18 (c)
It can be seen from Fig. 18 (a) that the coarse aggregates were and (d) is similar to that in Fig. 18 (b). However, the tested strength
intact and well bonded with matrix (geopolymer matrix and mor- was only 23.1 MPa and 15.0 MPa at 800 and 1000 °C, respectively.
tar matrix) and the concrete strength was 43.2 MPa. Tensile frac- The results suggested that the strength of coarse aggregate
ture face and cracks appeared in some coarse aggregates as the decreased dramatically after exposure to 800 and 1000 °C, result-
concrete strength reached to 89.7 MPa (Fig. 18 (b)). Therefore, it ing in the loss of concrete strength.
could be concluded that coarse aggregates could remain intact It is observed from Figs. 17 and 18 that coarse aggregate is bet-
under the following external load (r) range: 43.2 MPa  r < 89. ter bonded with matrix at 1000 °C than that at 800 °C due to the

Table 5
Experimental and predicted residual strengths by the proposed equation.

TEM GPN-A GPN-H GPH-A GPH-H


(°C)
Exp. (Std) Pred. Exp./ Exp. (Std) Pred. Exp./ Exp. (Std) Pred. Exp./ Exp. (Std) Pred. Exp./
[MPa] (MPa) Pred. [MPa] (MPa) Pred. [MPa] (MPa) Pred. [MPa] (MPa) Pred.
20 35.2 (1.2) – 39.4 (0.6) – 43.2 (2.4) – 72.8 (2.3) –
100 52.0 (0.7) 49.9 1.04 64.1 (5.6) 55.9 1.15 63.9 (1.0) 61.3 1.04 87.6 (2.4) 84.4 1.05
200 58.0 (2.5) 53.8 1.08 71.6 (1.0) 60.2 1.19 69.9 (2.1) 66.0 1.06 89.7 (3.3) 91.2 0.98
400 57.7 (2.1) 51.1 1.13 63.9 (0.8) 57.2 1.12 57.9 (6.0) 62.8 0.92 80.9 (2.4) 86.1 0.94
600 46.0 (2.9) 37.3 1.23 50.0 (4.3) 41.8 1.20 49.2 (3.9) 45.8 1.07 72.3 (2.2) 61.4 1.18
800 17.4 (0.4) 20.9 0.83 22.2 (3.4) 23.4 0.95 21.7 (1.9) 25.7 0.84 23.1 (2.8) 33.0 0.70
1000 15.3 (9.5) 9.0 1.70 19.5 (7.8) 10.1 1.93 14.6 (5.8) 11.1 1.32 13.6 (5.2) 13.4 1.02

Fig. 20. The comparison between experimental and predicted results by the proposed equations and existing models.

15
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

melting and sintering of fly ash particles and aluminosilicate gels. 800 °C. However, the concrete strength at 1000 °C is still lower
Therefore, it is empirical to believe that the residual compressive than the strength at 800 °C (Fig. 18), which is attributed to the
strength of GPH-H sample at 1000 °C is higher than that at higher crushing index of coarse aggregate at 1000 °C (Fig. 3).

Fig. 21. Validation of the proposed equation.

16
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

Therefore, the strength loss after 600 °C is caused by the poor models (Eq. (3) and Eq. (4)) are also illustrated in Fig. 20. The mean
load capacity of coarse aggregate. The strength loss of coarse aggre- ratio of the experimental to predicted strengths and the corre-
gate is attributed to their excessive expansions. The results suggest sponding coefficient of variation (COV) are also given in Fig. 20.
that the crushing index value of coarse aggregate could be consid- It is observed that the experimental results are very close to the
ered as a contributory factor to investigate the mechanism of predicted results by the proposed model but much higher than
geopolymer concrete deterioration at elevated temperatures. the predictions by Eurocode [47] and by Xiao and Falkner [48]. Tak-
ing GPN-A for an example, the mean values of the Exp./Pred. ratio
5. Prediction of residual compressive strength after exposure to by Eurocode and Xiao and Falkner are 4.12 and 4.40, respectively,
high temperature indicating that the experimental values are more than four times
higher than the predicted values by these two equations. On the
5.1. Proposed analytical model other hand, the mean value of the Exp./Pred. ratio by Eq. (3) is only
1.16, indicating that the predicted values by Eq. (3) is close to the
With an aim to propose a prediction model, the residual com- experimental results. There is an increasing trend of the experi-
pressive strengths are plotted against the elevated temperatures mentally determined compressive strengths up to elevated tem-
for the four groups in Fig. 19. The nonlinear fitting curves for each peratures of 200 °C which was ascribed to further
group are also drawn in the figure. It can be seen that the fitting geopolymerization after being subjected temperature increase.
curves are in good agreement with the experimental data with This phenomenon is different from the trend shown by OPC con-
the R2 value ranging from 0.88 to 0.94. crete and therefore, it is not considered in the existing models.
Based on the experimental results and respective fitting curves f cT ¼ l  f c ð5Þ
(Fig. 19), Eq. (3) and Eq. (4) were recommended as the proposed
model for predicting the residual compressive strength of geopoly- where m is reduction factor recommended in Eurocode EN
mer concrete at elevated temperatures. These two equations were 1994:2005 [47].
obtained by using the Gaussian-based mathematical model and

finding the coefficients by analysing experimental results with f cT ¼ ½1:011  ðT=1900Þ  f c T 6 400 C ð6aÞ
the help of MATLAB software. Eq. (3) is used for geopolymer con-

crete with compressive strength no more than 50 MPa (GPN-H, f cT ¼ ½1:44  ðT=625Þ  f c 400 C 6 T 6 900 C ð6bÞ
GPN-A and GPH-A groups in this paper) and Eq. (4) is applicable
for geopolymer concrete with compressive strength higher than
50 MPa (GPH-H group in this paper). 5.3. Further validation of the proposed model
When fc  50 MPa
In addition to the experimental results of this study, those avail-
T262:1 2
f cT ¼ 1:547  eð 550:1Þ f
c ð3Þ able in published literature on fly ash geopolymer concrete
[20,29,31,49,50] were also employed to validate Eq. (3) and Eq.
When fc > 50 MPa (4). The details of the comparisons are also summarized in
Fig. 21. It is observed from Fig. 21(a), (b) and (g) that the mean val-
f cT ¼ 1:268  eð Þ f
T259:7 2
532:8
c ð4Þ ues of Exp./Pred. ratio are 0.67, 0.74 and 1.36, respectively. The
where, variation is attributed to the different types of fly ash, alkalis and
curing method used in the preparation of geopolymer concrete.
fcT = Residual compressive strength (MPa) The mean values of Exp./Pred. ratio are generally around 1.00 in
fc = Original compressive strength (MPa) Fig. 21(c), (d), (e), and (f), implying that the difference between
T = Elevated temperatures (°C) experiments and predicts is within the acceptable range.
e = Natural constant
6. Conclusions
5.2. Analysis of the proposed model
Ambient-cured and heat-cured low-calcium fly ash-based
The experimental results, predicted results and the value of geopolymer concrete specimens were exposed to elevated temper-
experimental to predicted residual strengths (Exp./Pred.) are given atures of 100, 200, 400, 600, 800 and 1000 °C. The mass change,
in Table 5 Experimental and predicted residual strengths by the strength change and external and internal cracks of geopolymer
proposed equation. It can be seen that the experimental compres- concrete after exposure to the target temperatures were investi-
sive strengths are generally higher than the predicted strengths gated. Further insights into the effect of high temperature were
for GPN-A, GPN-H and GPH-A. As for GPH-H, the experimental achieved by using SEM images, TGA and crushing index of coarse
strengths are larger than the predicted strengths at 100 °C, aggregates. Two equations are proposed to predict the residual
600 °C and 1000 °C. Thus, the data in Table 5 show that Eq. (3) strengths using the test results of this study and those available
and Eq. (4) is generally conservative in predicting the residual in literature. The following conclusions are drawn from the study:
compressive strengths of geopolymer concrete at elevated
temperatures. 1. The mass loss increased with the increase of exposure temper-
Eurocode EN 1994:2005 [47] provides a formula (Eq. (5)) to pre- ature. Generally, the mass loss of high strength concrete was
dict the residual compressive strength of OPC concretes grade higher than that of normal strength concrete. For the same mix-
C40/50 at various elevated temperatures. Xiao and Falkner [48] ture, ambient-cured geopolymer concretes experienced higher
recommended an empirical model (Eqs. (6a) and (6b)) to estimate mass loss than the heat-cured specimens. It was found that
residual strength of high performance OPC concrete (grade the maximum mass loss of low-calcium fly ash geopolymer
C50 ~ C100) at different temperatures. The predicted residual concrete was smaller than the initial water content in the
strength by the Eurocode [47] and by Xiao and Falkner’s model mixture.
[48] for geopolymer concretes after exposure to different elevated 2. The phenomenon that concrete surface suffered from more sev-
temperatures are presented in Fig. 20. For the ease of comparison, ere deterioration than the concrete’s internal structure is
the experimental strength and predicted results with the proposed mainly attributed to the temperature gradient and gradual heat
17
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

exchange that occurred during the cooling period. The total Appendix A. Supplementary data
length of EXTR- and INTR-cracks reached the peak values at
800 °C and then decreased after exposure to 1000 °C indicating Supplementary data to this article can be found online at
a self-healing ability which is related to the shrinkage and sin- https://doi.org/10.1016/j.conbuildmat.2020.120924.
tering of geopolymer matrix.
3. Ambient-cured geopolymer concrete generally experienced
higher strength enhancement than heat-cured concrete. Nor- References
mal strength geopolymer concrete showed higher strength
enhancement than high strength geopolymer concrete for both [1] Z. Hongen, J. Feng, W. Qingyuan, T. Ling, S. Xiaoshuang, Influence of Cement on
Properties of Fly-Ash-Based Concrete, ACI Mater. J. 114 (5) (2017).
ambient-cured and heat-cured specimens. At 400 °C, the resid-
[2] M. Behera, S.K. Bhattacharyya, A.K. Minocha, R. Deoliya, S. Maiti, Recycled
ual compressive strengths of the concretes of four series were aggregate from C&D waste & its use in concrete – a breakthrough towards
higher than the corresponding initial compressive strengths. sustainability in construction sector: a review, Constr. Build. Mater. 68 (oct.15)
(2014) 501–516.
Low-calcium fly ash-based geopolymer concrete could be
[3] A. Bagheri, A. Nazari, A. Hajimohammadi, J.G. Sanjayan, P. Rajeev, M. Nikzad, T.
exposed to 600 °C for 2 h without strength loss. Ngo, P. Mendis, Microstructural study of environmentally friendly
4. The crushing index of coarse aggregate increased with the boroaluminosilicate geopolymers, J. Cleaner Prod. 189 (2018) 805–812.
increase of exposure temperature. The coarse aggregate was [4] P.K. Sarker, R. Haque, K.V. Ramgolam, Fracture behaviour of heat cured fly ash
based geopolymer concrete, Mater. Des. 44 (2013) 580–586.
strong enough to maintain the initial compressive strength of [5] A. Kashani, T.D. Ngo, B. Walkley, P. Mendis, Thermal performance of calcium-
concrete up to 600 °C. A crushing index of 7.7% could be consid- rich alkali-activated materials: a microstructural and mechanical study,
ered as the threshold value for coarse aggregate in order to Constr. Build. Mater. 153 (2017) 225–237.
[6] Y.-L. Li, X.-L. Zhao, R.K. Singh Raman, S. Al-Saadi, Thermal and mechanical
maintain its strength. The cracks induced in coarse aggregates properties of alkali-activated slag paste, mortar and concrete utilising
by temperature gradient contributed to the strength loss of con- seawater and sea sand, Constr. Build. Mater. 159 (2018) 704–724.
crete at temperatures higher than 600 °C. [7] M. Lahoti, K.K. Wong, E.-H. Yang, K.H. Tan, Effects of Si/Al molar ratio on
strength endurance and volume stability of metakaolin geopolymers subject to
5. The SEM results showed that the geopolymer concrete has its elevated temperature, Ceram. Int. 44 (5) (2018) 5726–5734.
unique microstructural characteristic at different temperatures. [8] P. Nath, P.K. Sarker, Use of OPC to improve setting and early strength
The mass loss of geopolymer concrete during the heating pro- properties of low calcium fly ash geopolymer concrete cured at room
temperature, Cem. Concr. Compos. 55 (2015) 205–214.
cess is attributed to the evaporation of free water and chemi- [9] H. Zhang, X. Shi, Q. Wang, Effect of Curing Condition on Compressive Strength
cally bound water, and the dehydroxylation of hydroxyl of Fly Ash Geopolymer Concrete, ACI Mater. J. 115 (2) (2018).
groups, which were found in the thermogravimetric analysis. [10] H.Y. Zhang, V. Kodur, S.L. Qi, L. Cao, B.o. Wu, Development of metakaolin–fly
ash based geopolymers for fire resistance applications, Constr. Build. Mater. 55
The strength loss at temperatures higher than 600 °C is also
(2014) 38–45.
ascribed to the microstructure deterioration and the dehydra- [11] T. Stengel, D. Heinz, J. Reger, Life cycle assessment of geopolymer concrete–
tion of geopolymer gels. what is the environmental benefit, in: Proceedings of the 24th Biennial
6. Based on the experimental results, two empirical equations Conference of the Concrete Institute of Australia, 2009.
[12] G. Khoury, Compressive strength of concrete at high temperatures: a
were proposed to predict the residual compressive strength of reassessment, Mag. Concr. Res. 44 (161) (1992) 291–309.
geopolymer concrete after exposure to different temperatures. [13] G.A. Khoury, Effect of fire on concrete and concrete structures, Prog. Struct.
The proposed equations were found to correlate well with the Engng Mater. 2 (4) (2000) 429–447.
[14] I. Hager, Behaviour of cement concrete at high temperature, Bull. Polish Acad.
test data of this study and those available in the published liter- Sci.: Tech. Sci. 61 (1) (2013) 145–154.
ature on fly ash geopolymer concrete. [15] H.Y. Zhang, V. Kodur, S.L. Qi, B.o. Wu, Characterizing the bond strength of
geopolymers at ambient and elevated temperatures, Cem. Concr. Compos. 58
(2015) 40–49.
CRediT authorship contribution statement [16] H.Y. Zhang, V. Kodur, B.o. Wu, J. Yan, Z.S. Yuan, Effect of temperature on bond
characteristics of geopolymer concrete, Constr. Build. Mater. 163 (2018) 277–
Hongen Zhang: Formal analysis, Investigation, Software, Data 285.
[17] W.G. Valencia Saavedra, R. Mejía de Gutiérrez, Performance of geopolymer
curation, Writing - original draft. Lang Li: Investigation, Software, concrete composed of fly ash after exposure to elevated temperatures, Constr.
Data curation. Cheng Yuan: Methodology, Formal analysis. Qin- Build. Mater. 154 (2017) 229–235.
gyuan Wang: Conceptualization, Methodology, Resources, Super- [18] H.M. Khater, A. Diouri, N. Khachani, M. Alami Talbi, Studying the effect of
thermal and acid exposure on alkali activated slag Geopolymer, MATEC Web
vision, Funding acquisition. Prabir Kumar Sarker: Writing -
Conf. 11 (2014) 01032, https://doi.org/10.1051/matecconf/20141101032.
review & editing, Supervision. Xiaoshuang Shi: Visualization, [19] C.-S. Poon, S. Azhar, M. Anson, Y.-L. Wong, Performance of metakaolin concrete
Supervision. at elevated temperatures, Cem. Concr. Compos. 25 (1) (2003) 83–89.
[20] P.K. Sarker, S. Kelly, Z. Yao, Effect of fire exposure on cracking, spalling and
residual strength of fly ash geopolymer concrete, Mater. Des. 63 (2014) 584–
Declaration of Competing Interest 592.
[21] O. Burciaga-Díaz, J.I. Escalante-García, Comparative performance of alkali
The authors declare that they have no known competing finan- activated slag/metakaolin cement pastes exposed to high temperatures, Cem.
Concr. Compos. 84 (2017) 157–166.
cial interests or personal relationships that could have appeared [22] Y.L. Galiano, C. Leiva, C. Arenas, F. Arroyo, L. Vilches, C.F. Pereira, R. Villegas,
to influence the work reported in this paper. Behaviour of fly ash-based geopolymer panels under fire, Waste Biomass Valor
8 (7) (2017) 2485–2494.
[23] D.L.Y. Kong, J.G. Sanjayan, Effect of elevated temperatures on geopolymer
Acknowledgements paste, mortar and concrete, Cem. Concr. Res. 40 (2) (2010) 334–339.
[24] D.L.Y. Kong, J.G. Sanjayan, Damage behavior of geopolymer composites
The authors are sincerely grateful for the financial support pro- exposed to elevated temperatures, Cem. Concr. Compos. 30 (10) (2008) 986–
991.
vided by the National Natural Science Foundation of China (No. [25] S.M. Park, J.G. Jang, N.K. Lee, H.K. Lee, Physicochemical properties of binder gel
11772209 & No. 11832007 & No. 11572057), the Science & Tech- in alkali-activated fly ash/slag exposed to high temperatures, Cem. Concr. Res.
nology Support Program of Sichuan Province (No. 2019JDS0006), 89 (2016) 72–79.
[26] B. Singh, M.R. Rahman, R. Paswan, S.K. Bhattacharyya, Effect of activator
the Basic scientific research fund for Central-Government-
concentration on the strength, ITZ and drying shrinkage of fly ash/slag
supported universities (Sichuan University). The first author is sup- geopolymer concrete, Constr. Build. Mater. 118 (2016) 171–179.
ported by the China Scholarship Council (CSC), which is sincerely [27] Ministry of Housing and Urban-Rural Construction of the People’s Republic of
appreciated. Additionally, the authors also like to thank He Yi of China GBT 14685-2011. Pebble and crushed stone for construction.
[28] Z. Zhang, Y. Zhang, C. Yan, Y. Liu, Influence of crushing index on properties of
the Analytical & Testing Center (ATC) of Sichuan University for recycled aggregates pervious concrete, Constr. Build. Mater. 135 (2017)
his great help in the SEM test. 112–118.

18
H. Zhang et al. Construction and Building Materials 262 (2020) 120924

[29] H. Zhang, L. Li, T. Long, P. Sarker, X. Shi, G. Cai, Q. Wang, The effect of ordinary [40] I. Hager, Colour change in heated concrete, Fire Technol. 50 (4) (2014) 945–
Portland cement substitution on the thermal stability of geopolymer concrete, 958.
Materials 12 (16) (2019) 2501, https://doi.org/10.3390/ma12162501. [41] Y.J. Zhang, S. Li, Y.C. Wang, D.L. Xu, Microstructural and strength evolutions of
[30] T. Kovářík, D. Rieger, J. Kadlec, T. Křenek, L. Kullová, M. Pola, P. Bělský, P. geopolymer composite reinforced by resin exposed to elevated temperature, J.
Franče, J. Říha, Thermomechanical properties of particle-reinforced Non-Cryst. Solids 358 (3) (2012) 620–624.
geopolymer composite with various aggregate gradation of fine ceramic [42] F.U.A. Shaikh, V. Vimonsatit, Compressive strength of fly-ash-based
filler, Constr. Build. Mater. 143 (2017) 599–606. geopolymer concrete at elevated temperatures, Fire Mater. 39 (2) (2015)
[31] M.W. Hussin, M.A.R. Bhutta, M. Azreen, P.J. Ramadhansyah, J. Mirza, 174–188.
Performance of blended ash geopolymer concrete at elevated temperatures, [43] W.D.A. Rickard, J. Temuujin, A. van Riessen, Thermal analysis of geopolymer
Mater. Struct. 48 (3) (2015) 709–720. pastes synthesised from five fly ashes of variable composition, J. Non-Cryst.
[32] P.K. Sarker, S. Mcbeath, Fire endurance of steel reinforced fly ash geopolymer Solids 358 (15) (2012) 1830–1839.
concrete elements, Constr. Build. Mater. 90 (2015) 91–98. [44] P. Rovnaník, P. Bayer, P. Rovnaníková, Characterization of alkali activated slag
[33] H. Fares, S. Remond, A. Noumowe, A. Cousture, High temperature behaviour of paste after exposure to high temperatures, Constr. Build. Mater. 47 (2013)
self-consolidating concrete: microstructure and physicochemical properties, 1479–1487.
Cem. Concr. Res. 40 (3) (2010) 488–496. [45] W.D. Rickard, C.S. Kealley, A. Van Riessen, Thermally induced microstructural
[34] L. Li, Q. Wang, G. Zhang, L. Shi, J. Dong, P.u. Jia, A method of detecting the changes in fly ash geopolymers: experimental results and proposed model, J.
cracks of concrete undergo high-temperature, Constr. Build. Mater. 162 (2018) Am. Ceram. Soc. 98 (3) (2015) 929–939.
345–358. [46] R. Zhao, J.G. Sanjayan, Geopolymer and Portland cement concretes in
[35] B. She, Y. Ming, GT/B 50081-2002: Standard for Test Method of Mechanical simulated fire, Mag. Concr. Res. 63 (3) (2011) 163–173.
Properties in Ordinary Concrete, Standard. [47] EN 1994-1-2 (2005) (English) Eurocode 4 Design of composite steel and
[36] Y.-S. Wang, J.-G. Dai, Z. Ding, W.-T. Xu, Phosphate-based geopolymer: concrete structures – Part 1-2 General rules - Structural fire design.
formation mechanism and thermal stability, Mater. Lett. 190 (2017) 209–212. [48] J. Xiao, H. Falkner, On residual strength of high-performance concrete with and
[37] C. Ridtirud, P. Chindaprasirt, K. Pimraksa, Factors affecting the shrinkage of fly without polypropylene fibres at elevated temperatures, Fire Saf. J. 41 (2)
ash geopolymers, Int. J. Miner. Metall. Mater. 18 (1) (2011) 100–104. (2006) 115–121.
[38] C. Jiang, C. Jin, Y. Wang, S. Yan, D.a. Chen, Effect of heat curing treatment on the [49] F.U.A. Shaikh, A. Hosan, Mechanical properties of steel fibre reinforced
drying shrinkage behavior and microstructure characteristics of mortar geopolymer concretes at elevated temperatures, Constr. Build. Mater. 114
incorporating different content ground granulated blast-furnace slag, Constr. (2016) 15–28.
Build. Mater. 186 (2018) 379–387. [50] A. Nazari, A. Bagheri, J.G. Sanjayan, M. Dao, C. Mallawa, P. Zannis, S. Zumbo,
[39] E. Annerel, L. Taerwe, Revealing the temperature history in concrete after Thermal shock reactions of Ordinary Portland cement and geopolymer
fire exposure by microscopic analysis, Cem. Concr. Res. 39 (12) (2009) concrete: microstructural and mechanical investigation, Constr. Build. Mater.
1239–1249. 196 (2019) 492–498.

19

You might also like