Sie sind auf Seite 1von 12

Thermal Performance Modeling of Geopolymer Concrete

Farhad Aslani, M.ASCE 1

Abstract: Geopolymers are used for several applications due to their sustainability, low density, low cost, excellent thermal properties, and
fire resistance. Geopolymer concrete (GC) may possess superior fire resistance compared to conventional concretes with ordinary portland
cement (OPC). The proper understanding of the effects of elevated temperatures on the properties of GC is essential. In the research reported
in this paper, relationships are proven for normal and high-strength GCs at elevated temperatures to establish efficient modeling and specify
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

the fire-performance criteria for concrete structures. They are developed for unconfined GC specimens that consist of compressive strength,
modulus of elasticity, flexural strength, thermal strain, prestressed thermal strain, and the compressive stress–strain relationships at elevated
temperatures. The proposed relationships at elevated temperature are compared with experimental results and available OPC relationships.
These results are used to establish more accurate and general compressive stress–strain relationships. DOI: 10.1061/(ASCE)MT.1943-5533
.0001291. © 2015 American Society of Civil Engineers.
Author keywords: Geopolymer concrete; Fire resistance; Compressive strength; Modulus of elasticity; Flexural strength; Thermal strain;
Prestressed thermal strain; Stress-strain relationship.

Introduction Geopolymer concrete exhibits many of the characteristics of


traditional concretes, despite their vastly different chemical con-
Concrete is the single most widely used material in the world and it stituents and reactions (Aly and Sanjayan 2010; Lloyd and Rangan
has a carbon footprint to match. Manufacturing of ordinary portland 2010; Sagoe-Crentsil 2009). The mixing process, the workability
cement (OPC) involves mining, crushing, and grinding limestone and of freshly mixed GCs, and the mechanical characteristics of the
shale, burning this raw material in a rotary kiln to convert the limestone hardened material appear to be similar or better than those for
into lime via a process known as calcination, and finally grinding the traditional ordinary portland cement concretes (OPCCs). The
resulting cement clinker with gypsum. This production process is very hardening mechanism for geopolymers essentially contains the pol-
energy-intensive and involves the release of greenhouse gases such as ycondensation reaction of geopolymeric precursors, regularly alu-
CO2 and N2 O into the atmosphere (Struble and Godfrey 2004). The minosilicate oxides, with alkali polysilicates yielding a polymeric
contribution of OPC manufacture to carbon emissions is second only silicon–oxygen–aluminium framework (Davidovits 1991). Due to
to fossil fuels and is estimated to be 8 million metric tons of CO2 per their inorganic basis, geopolymers are primarily fire resistant and
year in Australia (IPCC 2007). Promoting low-emission concretes is have been revealed to have outstanding thermal strength, with very
essential in order to face the crucial challenge to reduce the environ- slight gel structural degradation detected up to 700–800°C (Duxson
mental impact of the construction sector and the concrete industry and et al. 2006). Therefore, GC might be put into a reasonable place
to limit the impact of climate change. One way of reducing these CO2 against OPCC in the construction of high fire risk infrastructures.
emissions is the use of blended cements in which a part of the portland It is a crucial to recognize the modification in compressive strength
cement clinker is replaced with supplementary cementitious materials of geopolymeric materials during thermal exposure.
(SCMs). The most common SCMs used in high-volume applications
are fly ash (FA) and ground granulated blast furnace slag (GGBFS).
Fly-ash-based and GGBFS-based blended cements are extensively Research Significance
used [AS3582.1 (AS 1998); AS3582.2 (AS 2001)], but limits are im-
posed on the OPC replacement. In most cases, blended cements still In the research reported in this paper, constitutive relationships are
contain more OPC clinker than SCM. Davidovits (1991) pioneered the proposed for normal-strength and high-strength GCs at elevated
development of a new binder termed geopolymer which can com- temperatures that are compared to available OPCC relationships
pletely replace OPC in concrete. This new binder is an inorganic and verified with previous experimental data. Regression analyses
three-dimensional (3D) polymeric material made from the reaction are conducted on present experimental data to suggest compressive
of any material rich in silica and alumina with a strong alkaline solution strength, modulus of elasticity, flexural strength, thermal strain, and
that contains sodium silicate and/or sodium hydroxide (Komnitsas prestressed thermal strain. In the research reported in this paper,
and Zaharaki 2007). Aluminosilicate materials such as FA and first, the proposed relationships for compressive strength, modulus
GGBFS have been already successfully used in the production of geo- of elasticity, flexural strength, thermal strain, and prestressed ther-
polymer concrete (GC) in Australia (Aldred and Day 2012). mal strain are corroborated with experimental data. Second, com-
pressive stress–strain relationship for GCs at elevated temperatures
1
Postdoctoral Research Fellow, Center for Infrastructure Engineering are proposed and verified with experimental data and available
and Safety, Univ. of New South Wales, Sydney, NSW 2052, Australia. OPCC relationships.
E-mail: f.aslani@unsw.edu.au
Note. This manuscript was submitted on July 27, 2014; approved on
January 28, 2015; published online on May 13, 2015. Discussion period Previous Research
open until October 13, 2015; separate discussions must be submitted for
individual papers. This paper is part of the Journal of Materials in Civil The GC and OPCC are different in the binder properties and type.
Engineering, © ASCE, ISSN 0899-1561/04015062(12)/$25.00. The OPCC binder is based on hydration reactions of calcium oxide

© ASCE 04015062-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


and silicon dioxide to form calcium silicate hydrates. Moreover, various high-temperature behaviors for GCs. To spread over the
the GC binder is manufactured by alkali activation of aluminosil- models to a specific concrete mixture precisely, it is essential to use
icate raw materials, which are transformed into reaction product only investigations that are sufficiently consistent with the applied
by polymerization in a high pH environment and hydrothermal sit- testing methodology. The GCs experimental results included in the
uations at moderately low temperatures. This chemical structure database were gathered mainly from papers presented at various
provides GC benefits over its OPCC equivalent, such as a better published articles. The database includes information regarding
strength performance when GC materials are subjected to high- the composition of the mixtures, fresh properties of GCs, testing
temperature exposure (Davidovits 1991). Studies by Davidovits methodology, and conditions. Thermal characteristics have not
(1991), Davidovits and Davidovics (1991), Davidovits et al. been investigated as much as the other characteristics of GCs.
(1994), and Barbosa and MacKenzie (2003) described very good Tables 1–4 include information about the experimental tests,
heat properties of materials prepared using sodium silicate, potas- such as geopolymer resin composition, geopolymeric binder type,
sium silicate, and metakaolin (MK), having thermal stability up to type of aggregate, composition of geopolymeric binder, mechanical
1,200–1,400°C. Krivenko and Kovalchuk (2002) investigated heat properties specimens type, mechanical properties test age, temper-
resistance of geopolymer materials manufactured using Class F FA, atures and rate of heating, and duration of heat. Table 1 shows that
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

which had good thermal resistance properties up to 800°C. general type of geopolymer resin compositions that are used in the
The previous investigations that studied the thermal properties most of the research are sodium hydroxide and sodium silicate.
of the FA geopolymers have described that these materials pos- Also, most common geopolymeric binder type is FA with different
sessed low thermal shrinkage and good strength maintenance after SiO2 ∶Al2 O3 ratio. Moreover, common fine aggregate that is used
exposure to high temperatures (Bakharev 2006; Dombrowski et al. in the database is natural river sand and type of coarse aggregate
2007; Kong et al. 2007; Kong and Sanjayan 2008). The residual is crushed basalt with different maximum aggregate size. Table 2
strength of geopolymers has been found to be influenced by raw presents composition of geopolymeric binder collected in the
material type (Kong et al. 2007), alkali cation type (Bakharev experimental database. Table 3 specifies mechanical properties
2006), different calcium-containing raw materials (Dombrowski specimen type and mechanical properties test age are used for ther-
et al. 2007), curing temperatures, and activator-to-binders ratio mal behavior analysis. Also, heating rate range is between 4 and
(Kong et al. 2008). The mechanisms affecting strength at elevated 5°C=min. The temperature that is used for different research is
temperatures have been explored by considering pore pressure varied between 20 and 1,200°C. Table 4 summarizes the curing
effects (Kong et al. 2007; Bakharev 2006) and phase transforma- regime of experimental results database.
tions (Bakharev 2006; Dombrowski et al. 2007). In most cases the
residual strength of geopolymers was observed to be higher than
initial strength (tested before the thermal exposure). However, the Compressive Strength of GC at Elevated Temperatures
introduction of aggregates decreased the strength of GC (multi- The residual compressive behavior of concrete has been under
phase) after exposure to 800°C (Kong and Sanjayan 2008). investigation since the early 1960s [see the contributions by
Pan et al. (2009) experimental results have shown that ductility
Zoldners, Dougill, Harmathy, Crook, Kasami et al., Schneider,
significantly affects the residual strength of geopolymer mortars
and Diederiches, all quoted in Schneider (1985)]. Attention has
after exposure to 800°C. Hu et al. (2009) studied the fire resistance
been focused mostly on the compressive strength (the strength at
of lightweight geopolymer mortar prepared by the alkali-activation
room temperature after a specimen has been heated to a test temper-
of MK via liquid alkaline activator consisting of Na2 SiO3 and
ature and afterward cooled) as such, on the residual strain and on
NaOH solutions and the resulted geopolymer used to bind the light-
strength recovery with time (fib Bulletin No. 46 2008). In the re-
weight fine aggregate of crushed shale haydite sand. The resulted
search reported in this paper the efforts are focused on producing
lightweight geopolymer mortars cured at 20°C for 28 days showed
compressive strength relationship for GC of different strength
a strength loss percentage up to 63% of the initial strength after
classes which integrate different geopolymer resin composition
exposed to high temperature of 950°C for 30 min. Furthermore,
and geopolymeric binder, in order to investigate their performance
investigations have been recently focused on the following: (1) ex-
after exposure at gradually increasing temperature.
perimental study on the mechanical properties of geopolymer at
In the research reported in this paper, the relationship proposed
elevated temperatures (Pan and Sanjayan 2010; Pan et al. 2012),
for the compressive strength of GC by considering SiO2 ∶Al2 O3
(2) experimentally examine on the performances of various forms
(i.e., Si:Al) ratio at the geopolymeric binder at elevated temperatures
of FA based geopolymer and geopolymer composites and their per-
is based on regression analyses on existing experimental data with
formances under elevated temperatures (Kong and Sanjayan 2010;
the results expressed as Eq. (1). Moreover, the proposed relation-
Rickard et al. 2012; Abdulkareem et al. 2014), and (3) experimen-
ship for the compressive strength of GC is compared with available
tally discover the details for the different behaviors of GC and
OPCC in fire (Zhao and Sanjayan 2011). relationships for the compressive strength of OPCC developed by
Aslani and Bastami (2011) as Eqs. (2)–(6). The key purpose of re-
gression analyses is considering the changeable experimental com-
Experimental Results Collection pressive strength of GC behaviors at different elevated temperatures
and developing the rational and simple relationships that can fit well
An experimental results database from several distributed investi- with experimental data. This proposed GC relationship is compared
gations is an actual implement for studying the applicability of the with test results and OPCC relationships, as shown in Fig. 1.

Normal-strength and high-strength GC


 
0 0
1.0 20°C
f GcT ¼ fGc ð1Þ
1.1112 − 0.0004T − 2 × 10−6 T 2 100°C ≤ T ≤ 1,000°C

© ASCE 04015062-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

Table 1. Experimental Results Database of Geopolymer Resin, Geopolymeric Binder, and Aggregate Properties
References Geopolymer resin composition Geopolymeric binder type Aggregate type

© ASCE
Cheng and Chiu (2003) Sodium silicate solution, Na2 SiO3 , weight ratio, SiO2 ∶Na2 O ¼ Granulated blast furnace slag, —
3.3% by weight; Na2 O ¼ 6–7% by weight; SiO2 ¼ 23–25% SiO2 ∶Al2 O3 ¼ 2.37; and
Potassium hydroxide, KOH, with analytical grade metakaolin, SiO2 ∶Al2 O3 ¼ 1.22
Kong and Sanjayan (2008) Sodium silicate solution, Na2 SiO3 , weight ratio, Na2 O∶SiO2 ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 1.8 Crushed basalt and slag course aggregate with a
Potassium hydroxide, KOH, flakes of 90% purity was prepared maximum aggregate size of 14 mm
to a molarity of 7.0 M/ Lyndhurst sand is the fine aggregate with fineness
modulus of 1.82
Pan et al. (2009) Sodium silicate weight ratio, SiO2 ∶Na2 O ¼ 2 Fly ash A, SiO2 ∶Al2 O3 ¼ 3.11 Fly River sand
Weight percent Na2 O ¼ 14.7 ash B, SiO2 ∶Al2 O3 ¼ 1.58
Weight percent SiO2 ¼ 29.4
Commercial grade sodium hydroxide, NaOH, pellets with 98%
purity in distilled water. Concentration of the NaOH solution
was 10 M
Dodium silicate solution to sodium hydroxide solution ratio was
fixed at 2.5
Tie-Song et al. (2009) Potassium silicate solution, 40% in mass fraction Metakaolin, SiO2 ∶Al2 O3 ¼ 5.88, —
KOH, 16% in mass fraction 1.96, and 1.20
Kong and Sanjayan (2010) Sodium silicate weight ratio, SiO2 ∶Na2 O ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 1.8 Crushed old basalt course aggregate with a
Weight percent Na2 O ¼ 14.7 maximum aggregate size of 20 mm
Weight percent SiO2 ¼ 29.4
Hydroxide solution needed for activation was prepared to a River sand is the fine aggregate
concentration of 7.0 M using potassium hydroxide, KOH, flakes
of 90% purity and distilled water
Pan and Sanjayan (2010) Sodium silicate weight ratio, SiO2 ∶Na2 O ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 1.58 —
Weight percent Na2 O ¼ 14.7

04015062-3
Weight percent SiO2 ¼ 29.4

J. Mater. Civ. Eng.


Commercial grade sodium hydroxide, NaOH, pellets with 98%
purity in distilled water. Concentration of the NaOH solution
was 10 M
Sodium silicate solution to sodium hydroxide solution ratio was
fixed at 2.5
He et al. (2010) Potassium silicate solution, prepared by dissolving amorphous Metakaolin, SiO2 ∶Al2 O3 ¼ 1.28 —
silica into a KOH solution
Zhao and Sanjayan (2011) Sodium silicate weight ratio, SiO2 ∶Na2 O ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 1.80 Basalt course aggregate with a maximum
Weight percent Na2 O ¼ 14.7 aggregate size of 14 mm
Weight percent SiO2 ¼ 29.4
Hydroxide solution was prepared to 8 and 12 M concentrations Lyndhurst sand is the fine aggregate with fineness
using a mixture of distilled water and a commercial grade of modulus of 2.19
pellets with 90% purity
Pan et al. (2012) Sodium silicate weight ratio, SiO2 ∶Na2 O ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 1.58 Crushed old basalt course aggregate with a
Weight percent Na2 O ¼ 14.7 maximum aggregate size of 14 mm
Weight percent SiO2 ¼ 29.4
Potassium hydroxide solution was prepared by dissolving the River sand is the fine aggregate
commercial grade potassium hydroxide, KOH, pellets with 98%
purity in distilled water. The percentage of KOH by mass is
28.2%. Both alkaline solutions were prepared and mixed
together 1 day prior to usage

J. Mater. Civ. Eng.


Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

Table 1. (Continued.)
References Geopolymer resin composition Geopolymeric binder type Aggregate type

© ASCE
Rickard et al. (2012) Sodium silicate weight ratio, SiO2 ∶NaOH ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 2.0, 2.5, 3.0 —
Weight percent NaOH ¼ 14.7
Weight percent SiO2 ¼ 29.4
Sodium aluminate weight ratio,
Weight percent NaOH ¼ 25.5
Weight percent Al2 O3 ¼ 19.0
Abdulkareem et al. (2014) Sodium silicate weight ratio SiO2 ∶Na2 O ¼ 3.2 Fly ash, SiO2 ∶Al2 O3 ¼ 2.85 River sand is the fine aggregate with fineness
Weight percent Na2 O ¼ 9.4 modulus of 2.83
Weight percent SiO2 ¼ 30.1
NaOH of 12 M was prepared by mixing sodium hydroxide
pellets of 97–99% purity with distilled water. Alkaline activator
prepared by mixing the Na2 SiO3 and 12 M NaOH solution at a
constant mass ratio of 1:1
Pan et al. (2014) Sodium silicate weight ratio, SiO2 ∶Na2 O ¼ 2 Fly ash, SiO2 ∶Al2 O3 ¼ 1.58 Crushed old basalt course aggregate with a
Weight percent Na2 O ¼ 14.7 maximum aggregate size of 14 mm
Weight percent SiO2 ¼ 29.4
Sodium hydroxide solution was prepared by dissolving the River sand is the fine aggregate
commercial grade sodium hydroxide, NaOH, pellets with 98%
purity in distilled water
Khater (2014) Two different alkali activators are used, i.e., (1) sodium Water cooled slag, SiO2 ∶Al2 O3 ¼ Course aggregate with a maximum aggregate size
hydroxide, and (2) sodium silicate 3.69 Air-cooled slag, of 14 mm
Sodium silicate weight ratio SiO2 ∶Na2 O ¼ 1.88 SiO2 ∶Al2 O3 ¼ 2.99
Weight percent Na2 O ¼ 17
Weight percent SiO2 ¼ 32 River sand is the fine aggregate
Analytic grade concentrated nitric and hydrochloric acid was for

04015062-4
acid resistance testing in order to prepare 2, 4, and 6 M in an

J. Mater. Civ. Eng.


equal ratio

J. Mater. Civ. Eng.


Table 2. Composition of Geopolymeric Binders
References Geopolymeric binder type SiO2 Al2 O3 Fe2 O3 CaO MgO Na2 O K2 O TiO2 SO3
Cheng and Chiu (2003) Granulated blast furnace slag 34.39 14.47 0.63 41.67 6.49 0.22 0.36 0.53 —
Metakaolin 52.26 42.83 1.01 0.02 0.09 0.02 1.56 0.13 —
Kong and Sanjayan (2008) Fly ash 48.8 27.0 10.2 6.2 1.4 0.37 0.85 1.3 0.22
Pan et al. (2009) Fly ash A 72.2 23.2 0.6 0.2 0.1 0.1 0.4 — 0.1
Fly ash B 48.4 30.6 12.1 2.7 1.3 0.2 0.3 — 0.3
Tie-Song et al. (2009) Metakaolin —
Kong and Sanjayan (2010) Fly ash 48.8 27.0 10.2 6.2 1.4 0.37 0.85 1.3 0.22
Pan and Sanjayan (2010) Fly ash 48.3 30.5 12.1 2.8 1.2 0.2 0.4 — 0.3
He et al. (2010) Metakaolin 54.64 42.53 0.97 0.12 — — 0.48 0.80 0.11
Zhao and Sanjayan (2011) Fly ash 48.8 27.0 10.2 6.2 1.4 0.2 0.9 — 0.2
Pan et al. (2012) Fly ash 48.3 30.5 12.1 2.8 1.2 0.2 0.4 — 0.3
Rickard et al. (2012) Fly ash 57.15 6.47 1.36 1.37 0.29 0.09 0.55 0.61 —
— —
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

Abdulkareem et al. (2014) Fly ash 26.4 9.25 30.13 21.6 2.58 3.07 1.3
Pan et al. (2014) Fly ash 48.3 30.5 12.1 2.8 1.2 0.2 0.4 — 0.3
Khater (2014) Water-cooled slag 36.95 10.01 1.48 33.07 6.43 1.39 0.74 0.52 3.52
Air-cooled slag 31.01 10.35 1.58 34.86 3.06 2.15 0.37 0.34 2.41

Table 3. Experimental Results Database of Mechanical Properties Specimens Type, Test Age, Temperatures and Rate of Heating, and Duration of Heat
Properties
Mechanical properties Duration
References specimens type Mechanical properties test age Temperatures and rate of heating of heat
Cheng and Chiu Cube, 50 mm Compressive strength at 1, 7, 14, Subjected to temperatures of up to 35 min
(2003) and 28 days 1,100°C
Kong and Sanjayan Cube, 25 mm Compressive strength at 3 and Subjected to temperatures of up to 1h
(2008) 7 days 800°C at a gradual incremental rate
of approximately 5°C=min
Pan et al. (2009) Cylinder, 50-mm diameter and Compressive strength at 5 days Subjected to temperatures of up to 2h
100-mm high 800°C at a gradual incremental rate
of approximately 4.4°C=min
Tie-Song et al. Prism, 4 × 3 × 36 mm Flexural strength Subjected to temperatures of 400, 2h
(2009) 600, 800, 1,000, and 1,200°C at a
gradual incremental rate of
approximately 5°C=min
Kong and Sanjayan Cylinder, 100-mm diameter and Compressive strength at 3 days Subjected to temperatures of up to 1h
(2010) 200-mm high and Cube, 25 mm 800°C at a gradual incremental rate
of approximately 4.4°C=min
Pan and Sanjayan Cylinder, 24-mm diameter and Compressive strength at 28 days Subjected to temperatures of up to 1h
(2010) 48-mm high 680°C at a gradual incremental rate
of approximately 5°C=min
He et al. (2010) Flexural strength: Prism, 4 × Flexural strength and Young’s Subjected to temperatures of 1,000, 1.5 h
3 × 36 mm and Young’s modulus at 28 days 1,100, 1,200, 1,300, and 1,400°C
modulus: Cylinder, 150-mm for 90 min in an argon atmosphere
diameter and 300-mm high
Zhao and Sanjayan Cylinder, 100-mm diameter and Compressive strength at 28 days Subjected to temperatures of up to 30 min
(2011) 200-mm high 850°C
Pan et al. (2012) Cylinder, 100-mm diameter and Compressive strength at 28 days Subjected to temperatures of up to 5h
200-mm high 800°C at a gradual incremental rate
of approximately 5°C=min
Rickard et al. (2012) Cylinder, 15-mm diameter and Compressive strength at 28 days Subjected to temperatures of up to —
30-mm high 1,000°C at a gradual incremental
rate of approximately 5°C=min
Abdulkareem et al. Cube, 100 mm Compressive strength at 27 days Subjected to temperatures of 400, 1h
(2014) 600 and 800°C at a gradual
incremental rate of approximately
5°C=min
Pan et al. (2014) Cylinder, 100-mm diameter and Compressive strength at 28 days Subjected to temperatures of up to 2h
200-mm high 550°C at a gradual incremental rate
of approximately 5°C=min
Khater (2014) Cube, 25 mm Compressive strength at 28 days Subjected to temperatures of 300 to 2h
1,000°C at a gradual incremental
rate of approximately 5°C=min

© ASCE 04015062-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


Normal-strength OPCC

0 1
1.012 − 0.0005T ≤ 1.0 20°C ≤ T ≤ 100°C
B C
B 0.985 þ 0.0002T − 2.235 × 10−6 T 2 þ 8 × 10−10 T 3 C 100°C < T ≤ 800°C
0
fcT ¼ f c0 B
B 0.44 − 0.0004T
C
C ð2Þ
@ A 900°C ≤ T ≤ 1,000°C
0 T > 1,000°C

High-strength OPCC, 55.2 MPa ≤ fc0 ≤ 80 MPa (siliceous aggregate)


0 1
1.01 − 0.00068T ≤ 1.0 20°C ≤ T ≤ 200°C
B C
B 0.935 þ 0.00026T − 2.13 × 10−6 T 2 þ 8 × 10−10 T 3 C 200°C < T ≤ 400°C
B C
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

0 B C
fcT ¼ f c0 B 0.90 þ 0.0002T − 2.13 × 10−6 T 2 þ 8 × 10−10 T 3 C 400°C < T ≤ 800°C ð3Þ
B C
B 0.44 − 0.0004T C 900°C ≤ T ≤ 1,000°C
@ A
0 T > 1,000°C

High-strength OPCC, 80 MPa < f c0 ≤ 110 MPa (siliceous aggregate)


0 1
0.8 − 0.0005T ≤ 1.0 20°C ≤ T ≤ 500°C
B C
0B
B 0.96 − 0.0008T − 5.17 × 10−7 T 2 þ 4 × 10−10 T 3 C 500°C < T ≤ 800°C
0
fcT ¼ fc B C ð4Þ
C
@ 0.44 − 0.0004T A 800°C < T ≤ 1,00°C
0 T > 1,000°C

Calcareous aggregate OPCC


0 1
1.01 − 0.0006T ≤ 1.0 20°C ≤ T ≤ 200°C
0 B C
fcT ¼ fc0 @ 1.0565 þ 0.0017T þ 5 × 10−6 T 2 − 5 × 10−9 T 3 A 200°C < T ≤ 900°C ð5Þ
0 900°C < T

Lightweight aggregate OPCC


0 1
1.01 − 0.00037T ≤ 1.0 20°C ≤ T ≤ 300°C
0 B C
f cT ¼ fc0 @ 1.0491 − 0.00036T þ 10−6 T 2 − 2 × 10−9 T 3 A 300°C < T ≤ 900°C ð6Þ
0 T ≥ 1,000°C

0
where f Gc = normal compressive strength of GC; f c0 = normal com- 37% at 600°C, and 63% at 1,000°C; and OPCC compressive
0 strength is decreased by 3.8% at 100°C, 13.5% at 300°C, 53%
pressive strength of OPCC; f GcT = compressive strength of GC at
0 at 600°C, and 96% at 1,000°C.
elevated temperatures; and f cT = compressive strength of OPCC
at elevated temperatures.
Fig. 1 makes an assessment between proposed GC and available Modulus of Elasticity of GC at Elevated Temperatures
OPCC relationships for compressive strength against available un-
stressed experimental test results (unstressed tests are as described The elastic modulus of concrete could be affected predomi-
nantly by the same factors that influence its compressive strength
next. The specimen is heated, without preload, at a constant rate to
(Malhotra 1982). A relationship is proposed to assess the elasticity
the target temperature, which is maintained until a thermal steady
modulus of GC at elevated temperatures using regression analyses
state is achieved) as per Pan et al. (2009) with Si∶Al ¼ 3.11, 1.58; conducted on experimental data and is stated as Eq. (7). Addition-
Kong and Sanjayan (2010) with Si∶Al ¼ 1.8; Pan and Sanjayan ally, the proposed relationship for the elasticity modulus of GC is
(2010) with Si∶Al ¼ 1.58; Rickard et al. (2012) with Si∶Al ¼ 2.0, compared with available relationship for the elasticity modulus
2.5, or 3.0; and Pan et al. (2014) with Si∶Al ¼ 1.58. Experimental of OPCC established by Aslani and Bastami (2011) as Eq. (8).
results show that compressive strength of GC will be raised by in- The regression analyses are considering the variable experimental
creasing the temperatures. Evaluation of proposed compressive elastic modulus of GC behaviors at different elevated temperatures
strength relationships of GC and OPCC shows that the GC com- and developing the rational and simple relationship that can fits
pressive strength is increased by 8.35% at 100°C, 14% at 300°C, well with experimental data

© ASCE 04015062-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


 
1.0 20°C
EGcT ¼ EGc −7 2
ð7Þ
0.9713 þ 0.0009T − 5 × 10 T 100°C ≤ T ≤ 1,000°C

0 1
1.0 20°C ≤ T < 100°C
B C
EcT ¼ Ec @ 1.015 − 0.00154T þ 2 × 10−7 T 2 þ 3 × 10−10 T 3 A 100°C < T ≤ 1,000°C ð8Þ
0 T > 1,000°C

where EGc = normal modulus of elasticity of GC; Ec = normal 14% at 100°C, 42% at 300°C, 77% at 600°C, and 99% at
modulus of elasticity of OPCC; EGcT = modulus of elasticity of 900°C.
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

GC at elevated temperatures; and EcT = modulus of elasticity


of OPCC at elevated temperatures. Flexural Strength of GC at Elevated Temperatures
Fig. 2 shows comparison between proposed GC and avail-
able OPCC relationships for modulus of elasticity against pub- There are few studies on flexural strength. A relationship is sug-
lished unstressed experimental test results as per He et al. (2010) gested to assess the flexural strength of GC at elevated temperatures
using regression analyses conducted on experimental data and is
with Si∶Al ¼ 1.28, and Pan et al. (2014) with Si∶Al ¼ 1.58.
expressed as Eq. (9). Likewise, the proposed relationship for the
Experimental results demonstrate that modulus of elasticity of
flexural strength of GC is compared with existing relationship
GC will be raised by increasing the temperatures. Evaluation of for the flexural strength of OPCC developed by Aslani and Samali
proposed modulus of elasticity relationships of GC and OPCC (2014b, a) as Eq. (10). The regression analyses are considering
displays that the GC modulus of elasticity is increased by 5% the changeable experimental flexural strength of GC behaviors
at 100°C, 16% at 300°C, 25% at 600°C, 27% at 1,000°C, and at different elevated temperatures and developing the coherent
20% at 1,400°C; and OPCC modulus of elasticity is reduced and simple relationships that can fits well with experimental data

 
1.0 20°C
fGcrT ¼ fGcr ð9Þ
1.0866 − 0.0018T þ 1 × 10−6 T 2 100°C ≤ T ≤ 1,400°C

 
1.0 20°C
f crT ¼ fcr ð10Þ
1.095 − 0.0012T þ 2 × 10−7 T 2 100°C ≤ T ≤ 1,100°C

where fGcr = normal flexural strength of GC; f cr = normal flexu- employing a thermal expansion coefficient, α (Li and Purkiss
ral strength of OPCC; fGcrT = flexural strength of GC at elevated 2005)
temperatures; and fcrT = flexural strength of OPCC at elevated
temperatures. εth ¼ αðT − 20°CÞ ð11Þ
Fig. 3 illustrates comparison between proposed GC and avail-
able OPCC relationships for flexural strength in contradiction For concrete with siliceous or carbonate aggregates, α can be
of unstressed experimental test results as per Tie-Song et al. taken as equal to 18×10−6 =°C or 12×10−6 =°C (Purkiss 1996).
(2009) with Si∶Al ¼ 5.88, 1.96, 1.20; and He et al. (2010) with A relationship is proposed to calculate the thermal strain of
Si∶Al ¼ 1.28. Experimental results show that flexural strength GC at elevated temperatures using regression analyses conducted
of GC will be raised by increasing the temperatures. Evaluation on experimental data and is expressed as Eq. (12). In addition, the
of proposed flexural strength relationships of GC and OPCC shows proposed relationships for the thermal strain of GC are compared
that the GC flexural strength is increased by 5% at 100°C, 30% at with available relationships for the thermal strain of OPCC devel-
300°C, 54% at 600°C, 70% at 1,000°C, and increase to 57% at oped by Aslani (2013) as Eqs. (13)–(15)
1,400°C; and OPCC flexural strength is decreased 3% at 100°C,  
25% at 300°C, 55% at 600°C, and 90% at 1,000°C. 0 20°C
εG:th ¼
0.0001 − 2 × 10−5 T − 2 × 10−9 T 2 100°C ≤ T ≤ 1,300°C
Thermal Strain of Unstressed and Prestressed GC ð12Þ
Free thermal expansion is affected predominantly by the aggre- Siliceous aggregate OPCC
gate type. It expansion is not linear with respect to temperature.
The presence of free moisture will affect the results below 150°C εth ¼ 0.00045 þ 1 × 10−6 T þ 2 × 10−8 T 2 100°C ≤ T ≤ 800°C
since the water being driven off may cause net shrinkage. Con-
ventionally, it is stated by a linear function of temperature by ð13Þ

© ASCE 04015062-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


Table 4. Curing Regime of Experimental Database
References Curing
Cheng and Chiu (2003) Specimens were sealed and set at 60°C for 3 h in a sample-drying oven. After being removed from the molds, the
samples were kept at ambient temperature for another 21 h
Kong and Sanjayan (2008) Specimens were left to cure undisturbed under ambient temperature for 24 h before being subjected to an elevated
temperature of 80°C for an additional 24 h and finally they were slowly cooled
Pan et al. (2009) Within 1 h after the specimens were prepared, they were placed in an oven preheated to the specified temperature. For
mix with Fly Ash A, curing temperature and time were between 55–80°C and 24–96 h and for mix with Fly Ash B,
curing temperature and time were between 55–60°C and 2–18 h
Tie-Song et al. (2009) Heating of the specimens to the predetermined temperature was carried out at a rate of 5°C=min in a vacuum
atmosphere then the specimens were left at that temperature for a period of 2 h and finally they were slowly cooled
Kong and Sanjayan (2010) Specimens were left to cure undisturbed under ambient temperature for 24 h before being subjected to an elevated
temperature of 80°C for an additional 24 h and finally they were slowly cooled
Pan and Sanjayan (2010) Covered specimens were cured in an oven at 60°C for 18 h. After curing, specimens were demolded and kept in the
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

laboratory until testing


He et al. (2010) —
Zhao and Sanjayan (2011) Concrete in the cylinder molds was placed in an oven at 80°C for curing. Four batches of samples were cured in the
oven for 3, 8, 48, and 96 h, respectively, to reach the different target strengths. Then the samples were placed in a
constant-temperature room 23°C, 50% relative humidity
Pan et al. (2012) After casting, the specimens were kept in the molds and covered with polyethylene sheet and placed immediately in a
preheated oven. Specimens were cured at 60°C for 24 h. After curing, the specimens were demolded
Rickard et al. (2012) Specimens were sealed and samples cured for 24 h at 70°C. Samples were demolded and stored at ambient temperature
prior to testing
Abdulkareem et al. (2014) All wrapped moulds were cured undisturbed in oven at 70°C for 24 h after casting. Molds were taken out of the oven
and left to cool at room temperature before demolding. Sealed specimens were then stored under ambient conditions
Pan et al. (2014) Specimens were kept in the moulds and covered by a polyethylene sheet and placed immediately in a preheated oven.
After curing, specimens were stored in a controlled environment kept at 23  2°C and 50  3% relative humidity
Khater (2014) All mixes were left to cure undisturbed under ambient temperature for 24 h, and were then subjected to curing
temperature of 38°C under 100% relative humidity

Carbonate aggregate OPCC where εG:th = thermal strain of GC; and εth = thermal strain
of OPCC.
Fig. 4 shows comparison between proposed GC and available
εth ¼ 0.0001 þ 5 × 10−7 T þ 2 × 10−8 T 2 100°C ≤ T ≤ 800°C
OPCC relationships for thermal strain contrary to unstressed
ð14Þ experimental test results as per Tie-Song et al. (2009) with
Si∶Al ¼ 5.88, 1.96, and 1.20; Kong and Sanjayan (2010) with
Lightweight aggregate OPCC Si∶Al ¼ 1.8; He et al. (2010) with Si∶Al ¼ 1.28; Rickard et al.
(2012) with Si∶Al ¼ 2.0, 2.5, and 3.0; and Pan et al. (2014)
with Si∶Al ¼ 1.58. Experimental results display GC expanded
εth ¼ −0.00045 þ 8 × 10−6 T 100°C ≤ T ≤ 800°C ð15Þ at temperatures below 100°C while shrank at temperatures of

4.5 2
Pan et al. (2009), Si:Al=3.11 He et al. (2010), Si:Al=1.28
Pan et al. (2009), Si:Al=1.58 Pan et al. (2014), Si:Al=1.58
Kong and Sanjayan (2010), Si:Al=1.8 1.8
4 Proposed Relationship for OPCC
Pan and Sanjayan (2010), Si:Al=1.58 Proposed Relationship for GC
Rickard et al. (2012), Si:Al=2.0 1.6
3.5 Rickard et al. (2012), Si:Al=2.5
Rickard et al. (2012), Si:Al=3.0
Pan et al. (2014), Si:Al=1.58 1.4
3 Proposed Relationship for OPCC
Proposed Relationship for GC 1.2
(EGcT / EGc )
(f'GcT / f'Gc)

2.5
1
2
0.8
1.5
0.6

1 0.4

0.5 0.2

0 0
0 200 400 600 800 1000 1200 0 200 400 600 800 1000 1200 1400
Temperature (°C) Temperature (°C)

Fig. 1. Comparison between compressive strength proposed relation- Fig. 2. Comparison between modulus of elasticity proposed relation-
ships of GC and OPCC with experimental test results ships of GC and OPCC with experimental test results

© ASCE 04015062-8 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


200–1,200°C. Comparison of the proposed thermal strain relation- of prestressed GC at elevated temperatures using regression analy-
ships of GC and OPCC shows that the GC is shrank gradually ses conducted on experimental data as per Pan and Sanjayan (2010)
between 100 and 1,200°C and OPCC is expanded between 100 with Si∶Al ¼ 1.58, and 25 and 50% prestressed) and are expressed
and 1,200°C. as Eqs. (16) and (17). Moreover, the proposed relationships for the
In the research reported in this paper, relationships are proposed prestressed thermal strain of GC are compared with available
as a function of preloading percentages of the GC compressive OPCC relationships for the prestressed thermal strain developed
0
strength at room temperature (f Gc ) to evaluate the thermal strain by Aslani (2013) and Aslani and Samali (2015) as Eqs. (18)–(21)

0
For 15–30% prestressed of the fGc
0 1
0 20°C
B 0.0027 − 0.00002T C 100°C ≤ T ≤ 300°C
B C
εG:th ¼ B C ð16Þ
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

@ −0.0512 þ 0.000165T A 300°C < T ≤ 500°C


0.31 − 0.0005T 500°C < T ≤ 800°C
0
For 30–60% prestressed of the fGc
0 1
0 20°C
B C 100°C ≤ T ≤ 300°C
B 0.0034 þ 0.0001T − 5 × 10−7 T 2 C
εG:th ¼ B C ð17Þ
@ −0.053 þ 0.00013T A 300°C < T ≤ 500°C
0.4644 − 0.0008T 500°C < T ≤ 800°C

For 10–15% prestressed of the fc0


 
0 20°C
εth ¼ ð18Þ
−0.0009 þ 1.3 × 10−5 T − 2 × 10−8 T 2 100°C ≤ T ≤ 800°C

For 15–30% prestressed of the fc0


 
0.0002 − 10−5 T 20°C ≤ T < 200°C
εth ¼ ð19Þ
−5
−0.0073 þ 5 × 10 T − 8 × 10 T −8 2 200°C ≤ T ≤ 800°C

For 30–45% prestressed of the fc0


 
0 20°C
εth ¼ ð20Þ
−0.00005 − 1 × 10−5 T − 1.5 × 10−9 T 2 100°C ≤ T ≤ 800°C

For 45–60% prestressed of the fc0


 
0 20°C
εth ¼ ð21Þ
−0.0004 − 3 × 10−6 T − 2.4 × 10−8 T 2 100°C ≤ T ≤ 800°C

Fig. 5 provides a comparison between the proposed relation- conducted recently by Aslani and Bastami (2011) and Aslani
ships for the thermal strain of prestressed GC against the experi- (2013). In the research reported in this paper, a compressive stress–
mental results of Pan and Sanjayan (2010) for 25 and 50% strain relationship for normal-strength and high-strength GC at
0
prestressed of the fGc cases, respectively. The proposed relation- elevated temperatures based on the Carreira and Chu (1985) model
ships of GC fit of the experimental results well, which indicates with several modifications is developed by using the proposed
that the GC is expanded between 20 and 200°C for both 25 and residual compression strength, modulus of elasticity, flexural
50% and shrank between 200 and 350°C for 25% and 200– strength, thermal strain, and prestressed thermal strain, which are
400°C for 50%, and expanded again between 350 and 650°C expressed as Eqs. (22)–(29)
for 25% and 400–550°C for 50%, and shrank again for both
prestressed specimens. Fig. 5 also indicates the trend of the ther- σGcT ηmT ðεεGcT
0 Þ
0 ¼ GcT
εGcT ηmT ð22Þ
mal strain–temperature curves, which is entirely a function of the f GcT ηmT − 1 þ ðε 0 Þ
GcT
preloading percentage.
ηmT ¼ ηmT;a ðfittedÞ ¼ ½1.02 − 1.17ðEp =EGcT Þ−0.74
Compressive Stress-Strain Relationship of GC at 0
Elevated Temperatures if εGcT ≤ εGcT ð23Þ

The most essential existing compressive stress-strain relationships


0
for concrete at high temperatures are summarized in research ηmT ¼ ηmT;d ðfittedÞ ¼ ηmT;a ðfittedÞ þ ðγ þ λtÞ if εGcT ≤ εGcT ð24Þ

© ASCE 04015062-9 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


2 0.04
Tie-song et al. (2009), Si:Al=60:0
1.8 Tie-song et al. (2009), Si:Al=5.88
0.02
Tie-song et al. (2009), Si:Al=1.96
1.6 Tie-song et al. (2009), Si:Al=1.20
0
He et al. (2010), Si:Al=1.28
1.4 Proposed Relationship for OPCC
-0.02

Thermal Strain
Proposed Relationship for GC
1.2
(fGcrT / fGcr)

-0.04
1

0.8 -0.06
Pan and Sanjayan (2010), Si:Al=1.58, 25%
0.6 Pan and Sanjayan (2010), Si:Al=1.58, 50%
-0.08
Proposed Relationship for OPCC, 15-30%
0.4 Proposed Relationship for OPCC, 45-60%
-0.1
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

Proposed Relationship for GC, 15-30%


0.2 Proposed Relationship for GC, 45-60%
-0.12
0 100 200 300 400 500 600 700 800 900
0
Temperature (°C)
0 200 400 600 800 1000 1200 1400
Temperature (°C)
Fig. 5. Comparison between thermal strain proposed relationships of
Fig. 3. Comparison between flexural strength proposed relationships prestressed GC and OPCC with experimental test results
of GC and OPCC with experimental test results

90

80
0.04
70
0.03
Compressive Stress (MPa)

60
0.02

0.01 50
Thermal Strain

0 40

-0.01 30

-0.02 Tie-song et al. (2009), Si:Al=60:0 20


Tie-song et al. (2009), Si:Al=5.88
Tie-song et al. (2009), Si:Al=1.96
-0.03 Tie-song et al. (2009), Si:Al=1.20 Pan and Sanjayan (2010), 23,100& 200 C
Kong and Sanjayan (2010), Si:Al=1.8
10
He et al. (2010), Si:Al=1.28 Proposed Relationship for GC
-0.04 Rickard et al. (2012), Si:Al=2.0
Rickard et al. (2012), Si:Al=2.5 0
Rickard et al. (2012), Si:Al=3.0 0 0.005 0.01 0.015 0.02
-0.05 Pan et al. (2014), Si:Al=1.58
Proposed Relationship for OPCC Strain
Proposed Relationship for GC
-0.06
0 200 400 600 800 1000 1200 1400 Fig. 6. Comparisons between proposed compressive stress-strain
Temperature (°C) relationship for GC against the experimental results from Pan and
Sanjayan (2010) at 23, 100, and 200°C
Fig. 4. Comparison between thermal strain proposed relationships of
GC and OPCC with experimental test results

Figs. 6–11 provide a comparison between the proposed high-


strength GC relationship against the experimental results of Pan
and Sanjayan (2010) at 23, 100, 200, 290, 380, 520, 575, and
γ ¼ 35 × ð12.4 − 1.66 × 10−2 f GcT
0
Þ−0.9 ð25Þ 680°C, which indicate that the proposed relationship has a good
agreement with the experimental results.
0
λ ¼ 0.75 expð−911=fGcT Þ ð26Þ
Conclusions
Ep ¼ 0
fGcT 0
=εGcT ð27Þ In this paper, constitutive relationships for the normal and high
strength GC exposed to fire are proposed, which are intended
 0
 
0 f GcT ψ to provide efficient modeling and to specific fire-performance
εGcT ¼ ð28Þ criteria of the behavior of concrete structures. The major conclu-
EGcT ψ−1
sions resulting from the research reported in this paper are as
follows:
0
ψ¼
fGcT
þ 0.8 ð29Þ • The proposed relationships for the compressive strength, elas-
17 ticity modulus, and flexural strength of unstressed GC with

© ASCE 04015062-10 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


120 140

120
100

Compressive Stress (MPa)


100
Compressive Stress (MPa)

80
80

60
60

40 40

20 Pan and Sanjayan (2010), 520 C


20
Pan and Sanjayan (2010), 290 C Proposed Relationship for GC
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

Proposed Relationship for GC 0


0 0 0.01 0.02 0.03 0.04
0 0.002 0.004 0.006 0.008 0.01
Strain
Strain
Fig. 9. Comparisons between proposed compressive stress-strain
Fig. 7. Comparisons between proposed compressive stress-strain relationship for GC against the experimental results from Pan and
relationship for GC against the experimental results from Pan and Sanjayan (2010) at 520°C
Sanjayan (2010) at 290°C
120

120
100

100
Compressive Stress (MPa) 80
Compressive Stress (MPa)

80
60

60
40

40
20
Pan and Sanjayan (2010), 575 C

20 Proposed Relationship for GC


Pan and Sanjayan (2010), 380 C 0
Proposed Relationship for GC 0 0.02 0.04 0.06 0.08 0.1
0 Strain
0 0.005 0.01 0.015 0.02
Strain Fig. 10. Comparisons between proposed compressive stress-strain
relationship for GC against the experimental results from Pan and
Fig. 8. Comparisons between proposed compressive stress-strain Sanjayan (2010) at 575°C
relationship for GC against the experimental results from Pan and
Sanjayan (2010) at 380°C
140

120
different geopolymer resin composition, geopolymeric binder
type, and aggregate type at elevated temperature are in good
Compressive Stress (MPa)

100
reasonable agreement with the experimental results;
• The free thermal strain relationships that are proposed for un- 80
stressed and prestressed GC at high temperatures are verified
well to the experimental results; 60
• The proposed compressive stress–strain relationship for GC is
made based on the well-established relationships for concrete 40
at elevated temperatures, which has a good conformity with
experimental test results of GC at different high temperatures 20 Pan and Sanjayan (2010), 680 C
(also, using these relationships in the FEM is more simple
Proposed Relationship for GC
and suitable); and 0
• The paper stressed the fact that additional tests at different 0 0.1 0.2 0.3 0.4 0.5
temperatures are needed to investigate the role of initial com- Strain
pressive and tensile stresses on the GC compressive strength,
Fig. 11. Comparisons between proposed compressive stress-strain
strain at peak stress, modulus of elasticity, free thermal strain,
relationship for GC against the experimental results from Pan and
load induced thermal strain, creep strain, transient strain, and
Sanjayan (2010) at 680°C
fire spalling.

© ASCE 04015062-11 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.


References IPCC (Intergovernmental Panel on Climate Change). (2007). “The
fourth assessment report of the Intergovernmental Panel on Climate
Abdulkareem, O. A., Al Bakri, A. M. M., Kamarudin, H., Nizar, I. K., and Change (IPCC).” Climate change 2007: The physical science basis,
Saif, A. A. (2014). “Effects of elevated temperatures on the thermal S. Solomon, et al., eds., Cambridge University Press, Cambridge,
behavior and mechanical performance of fly ash geopolymer paste, U.K., 129–234.
mortar and lightweight concrete.” Constr. Build. Mater., 50, 377–387. Khater, H. M. (2014). “Studying the effect of thermal and acid exposure on
Aldred, J., and Day, J. (2012). “Is geopolymer concrete a suitable alterna- alkali activated slag geopolymer.” Adv. Cement Res., 26(1), 1–9.
tive technology to traditional concrete?” Proc., 37th Conf. on Our Komnitsas, K., and Zaharaki, D. (2007). “Geopolymerisation: A review and
World in Concrete and Structures, CI-Premier, Singapore. prospects for the minerals industry.” Miner. Eng., 20(14), 1261–1277.
Aly, T., and Sanjayan, J. G. (2010). “Effect of pore-size distribution on Kong, D. L. Y., and Sanjayan, J. G. (2008). “Damage behavior of geopol-
shrinkage of concretes.” J. Mater. Civ. Eng., 10.1061/(ASCE)0899 ymer composites exposed to elevated temperatures.” Cement Concrete
-1561(2010)22:5(525), 525–532. Compos., 30(10), 986–991.
AS (Australian Standards). (1998). “Supplementary cementitious materials Kong, D. L. Y., and Sanjayan, J. G. (2010). “Effect of elevated temperatures
for use with portland and blended cement. Part 1: Fly ash.” AS3582.1, on geopolymer paste, mortar and concrete.” Cement Concrete Res.,
Australia. 40(2), 334–339.
AS (Australian Standards). (2001). “Supplementary cementitious materials
Downloaded from ascelibrary.org by University of Exeter on 07/22/15. Copyright ASCE. For personal use only; all rights reserved.

Kong, D. L. Y., Sanjayan, J. G., and Sagoe-Crentsil, K. (2007). “Compar-


for use with portland and blended cement. Part 2: Slag–ground granu-
ative performance of geopolymers made with metakaolin and fly ash
lated iron blast-furnace.” AS3582.2.
after exposure to elevated temperatures.” Cement Concrete Res.,
Aslani, F. (2013). “Prestressed concrete thermal behavior.” Mag. Concrete
37(12), 1583–1589.
Res., 65(3), 158–171.
Aslani, F., and Bastami, M. (2011). “Constitutive relationships for normal- Kong, D. L. Y., Sanjayan, J. G., and Sagoe-Crentsil, K. (2008). “Factors
and high-strength concrete at elevated temperature.” ACI Mater. J., affecting the performance of metakaolin geopolymers exposed to
108(4), 355–364. elevated temperatures.” J. Mater. Sci., 43(3), 824–831.
Aslani, F., and Samali, B. (2015). “Constitutive relationships for self- Krivenko, P. V., and Kovalchuk, G. Y. (2002). “Heat-resistant fly ash based
compacting concrete at elevated temperatures.” Mater. Struct., geocements.” 3rd Int. Conf. on Geopolymers, Melbourne, Australia.
48(1–2), 337–356. Li, L., and Purkiss, J. A. (2005). “Stress-strain constitutive equations
Aslani, F., and Samali, B. (2014a). “Constitutive relationships for steel fibre of concrete material at elevated temperatures.” Fire Saf. J., 40(7),
reinforced concrete at elevated temperatures.” Fire Technol., 50(5), 669–686.
1249–1268. Lloyd, N. A., and Rangan, B. V. (2010). “Geopolymer concrete with fly
Aslani, F., and Samali, B. (2014b). “High strength polypropylene fibre ash.” Proc., Second Int. Conf. on Sustainable Construction Materials
reinforcement concrete at high temperature.” Fire Technol., 50(5), and Technologies, Università Politecnica delle Marche (UNIVPM),
1229–1247. Ancona, Italy, 1493–1504.
Bakharev, T. (2006). “Thermal behaviour of geopolymers prepared using Malhotra, H. L. (1982). Design of fire-resisting structures, Surrey
class F fly ash and elevated temperature curing.” Cement Concrete Res., University Press, London.
36(6), 1134–1147. Pan, Z., Sanjayan, J., and Rangan, B. (2009). “An investigation of the
Barbosa, V. F. F., and MacKenzie, K. J. D. (2003). “Synthesis and thermal mechanisms for strength gain or loss of geopolymer mortar after expo-
behaviour of potassium sialate geopolymers.” Mater. Lett., 57(9–10), sure to elevated temperature.” J. Mater. Sci., 44(7), 1873–1880.
1477–1482. Pan, Z., and Sanjayan, J. G. (2010). “Stress-strain behaviour and abrupt loss
Carreira, D. J., and Chu, K. H. (1985). “Stress-strain relationship for plain of stiffness of geopolymer at elevated temperatures.” Cement Concrete
concrete in compression.” ACI J., 82(6), 797–804. Compos., 32(9), 657–664.
Cheng, T. W., and Chiu, J. P. (2003). “Fire-resistant geopolymer produced Pan, Z., Sanjayan, J. G., and Collins, F. (2014). “Effect of transient creep on
by granulated blast furnace slag.” Miner. Eng., 16(3), 205–210. compressive strength of geopolymer concrete for elevated temperature
Davidovits, J. (1991). “Geopolymers: Inorganic polymeric new materials.” exposure.” Cement Concrete Res., 56, 182–189.
J. Therm. Anal., 37(8), 1633–1656. Pan, Z., Sanjayan, J. G., and Kong, D. L. Y. (2012). “Effect of aggregate
Davidovits, J. (1994). “High-alkali cements for 21st century concretes.” size on spalling of geopolymer and portland cement concretes subjected
Proc., Malhotra Symp. on Concrete Technology: Past, Present and to elevated temperatures.” Constr. Build. Mater., 36, 365–372.
Future, P. Kumar Metha, ed., 383–397. Purkiss, J. A. (1996). Fire safety engineering design of structures,
Davidovits, J., and Davidovics, M. (1991). “Geopolymer. Ultra-high tem-
Butterworth Heinemann, Oxford, U.K.
perature tooling material for the manufacture of advanced composites.”
Rickard, W. D. A., Temuujin, J., and van Riessen, A. (2012). “Thermal
SAMPE, 36(2), 1939–1949.
analysis of geopolymer pastes synthesised from five fly ashes of
Dombrowski, K., Buchwald, A., and Weil, M. (2007). “The influence of
variable composition.” J. Non-Cryst. Solids, 358(15), 1830–1839.
calcium content on the structure and thermal performance of fly ash
based geopolymers.” J. Mater. Sci., 42(9), 3033–3043. Sagoe-Crentsil, K. (2009). “Role of oxide ratios on engineering perfor-
Duxson, P., Lukey, G. C., and van Deventer, J. S. J. (2006). “Thermal mance of fly-ash geopolymer binder systems.” Ceram. Eng. Sci. Proc.,
evolution of metakaolin geopolymers: Part 1—Physical evolution.” 29(10), 175–184.
J. Non-Cryst. Solids, 352(52–54), 5541–5555. Schneider, U. (1985). “Properties of materials at high temperature–
fib Bulletin No. 46. (2008). “Fire design of concrete structures—Structural Concrete.” Rep. Prepared for RILEM, Univ. of Kassel, Kassel,
behaviour and assessment.” Chapter 6, Expertise and assessment of ma- Germany.
terials and structures after fire, fib (Fédération internationale du béton/ Struble, L., and Godfrey, J. (2004). “How sustainable is concrete?” Proc.,
International Federation for Structural Concrete), Switzerland. Int. Workshop on Sustainable Development and Concrete Technology,
He, P., Jia, D., Lin, T., Wang, M., and Zhou, Y. (2010). “Effects of high- Iowa State Univ., Ames, IA, 201–211.
temperature heat treatment on the mechanical properties of unidirec- Tie-Song, L., De-Chang, J., Pei-Gang, H., and Mei-Rong, W. (2009).
tional carbon fiber reinforced geopolymer composites.” Ceram. Int., “Thermal-mechanical properties of short carbon fiber reinforced geo-
36(4), 1447–1453. polymer matrix composites subjected to thermal load.” J. Cent. South
Hu, S., Wu, J., Wen, Y., He, Y., Fa-Zhou, W., and Qing-Jun, D. (2009). Univ. Technol., 16(6), 0881–0886.
“Preparation and properties of geopolymer-lightweight aggregate re- Zhao, R., and Sanjayan, J. G. (2011). “Geopolymer and portland cement
fractory concrete.” J. Cent. South Univ. Technol., 16(6), 914–918. concretes in simulated fire.” Mag. Concrete Res., 63(3), 163–173.

© ASCE 04015062-12 J. Mater. Civ. Eng.

J. Mater. Civ. Eng.

Das könnte Ihnen auch gefallen