Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Report

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

1

1. INTRODUCTION

With the growing environmental and economic concerns associated with conventional
concrete-based building materials such as reinforced concrete structures, researchers have been
actively involved in exploring possibilities in using alternative materials to address these
concerns. For instance, alternative concrete-making materials have been trialled in reinforced
concrete structures such as recycled concrete aggregate and agriculture waste materials, in an
attempt to reduce the dependency on conventional concrete constituent materials, which are
depleting at a fast rate.

One of the primary environmental concerns from concrete-based building materials is


the high amount of carbon dioxide emission, which arises during the manufacturing of cement.
Approximately 5% of the global carbon dioxide emission is contributed by the cement industry.
In recent times, a cement-less binder for producing concrete, termed as geopolymer concrete,
is fast gaining popularity in concrete research work as this technology eliminates the need for
cement. In order to produce geopolymer, a process termed as ‘geopolymerization’ is required
which involves the reaction between aluminosilicate material and alkaline liquids. Common
aluminosilicate material used for producing geopolymer is fly ash and slag, which are both
industrial by-products and both of these materials have much lower carbon dioxide emission
factor compared to cement. It was reported that the use of geopolymer could bring down the
overall carbon dioxide emission by up to 64% in comparison with the use of cement.
Furthermore, in terms of economic consideration, due to the lower price of fly ash compared
to cement, the price of fly ash-based geopolymer concrete could be as low as 10–30% cheaper
compared to conventional cement-based concrete after taking into account the price of alkaline
liquids.

While most of the research works on geopolymer concrete focus on micro-scale


investigation, recent researchers on the use of geopolymer concrete extends to the investigation
of the structural behaviour of geopolymer concrete in load-bearing members such as reinforced
concrete beams, columns, slabs and more. The structural properties of the concrete members
is one of the most vital component in effectively introducing such concrete for actual buildings
and applications. The conformity of the performance of reinforced geopolymer concrete
2

members with existing design provisions should be ascertained in order to evaluate the
feasibility of using these design codes for geopolymer concrete members for the convenience
of structural design engineers. Due to the importance of the structural aspect of utilizing
geopolymer concrete in reinforced concrete structures, this report discusses the published
findings of research works involving geopolymer concrete structures such as beams, columns,
slabs and panels.

2. GEOPOLYMER CONCRETE

The name geopolymer was formed by a French Professor Davidovits in 1978 to represent a
broad range of materials characterized by networks of inorganic molecules. The geopolymers
depend on thermally activated natural materials like Meta kaolinite or industrial byproducts
like fly ash or slag to provide a source of silicon (Si) and aluminium (Al). These Silicon and
Aluminium is dissolved in an alkaline activating solution and subsequently polymerizes into
molecular chains and become the binder.

Professor B. Vijaya Rangan (2008), Curtin University, Australia, stated that, “the
polymerization process involves a substantially fast chemical reaction under alkaline
conditions on silicon-aluminium minerals that results in a three-dimensional polymeric chain
and ring structure”. The ultimate structure of the geopolymer depends largely on the ratio of Si
to Al (Si:Al), with the materials most often considered for use in transportation infrastructure
typically having a Si:Al between 2 and 3.5.

The reaction of Fly Ash with an aqueous solution containing Sodium Hydroxide and
Sodium Silicate in their mass ratio, results in a material with three dimensional polymeric chain
and ring structure consisting of Si-O-Al-O bonds.
The schematic formation of geopolymer material can be shown as described by
Equations (A) and (B).
3

(OH)2 O O O

Water is not involved in the chemical reaction of Geopolymer concrete and instead
water is expelled during curing and subsequent drying. This is in contrast to the hydration
reactions that occur when Portland cement is mixed with water, which produce the primary
hydration products calcium silicate hydrate and calcium hydroxide. This difference has a
significant impact on the mechanical and chemical properties of the resulting geopolymer
concrete, and also renders it more resistant to heat, water ingress, alkali–aggregate reactivity,
and other types of chemical attack.

2.1. Constituents of Geopolymer Concrete

The following are the constituents of Geopolymer concrete


 Fly Ash- rich in Silica and Aluminium
 Sodium Hydroxide or Potassium Hydroxide
 Sodium Silicate or Potassium Silicate

2.2. Properties of Geopolymer Concrete

The superior properties of Geopolymer concrete are


 sets at room temperature
 nontoxic, bleed free
 long working life before stiffening
 impermeable
 higher resistance to heat and resist all inorganic solvents
 higher compressive strength

Compressive strength of Geopolymer concrete is very high compared to the ordinary


Portland cement concrete. Geopolymer concrete also shows very high early strength. The
compressive strength of Geopolymer concrete is about 1.5 times more than that of the
compressive strength with the ordinary Portland cement concrete, for the same mix.
4

Similarly the Geopolymer Concrete shows good workability compared to the ordinary
Portland Cement Concrete.

3. REINFORCING BAR-CONCRETE BOND

The structural performance of reinforced concrete members depends on the bond


between concrete and reinforcement, in which the mechanism of bond influences the embedded
length of reinforcing bar and consequently the load-bearing capacity of structural elements,
crack opening and spacing. ACI 408R considers the bond strength as one of the structural
properties and understanding of the behaviour is critical to the eventual development of
analysis and design basis of the structural member. Because of the difference in terms of
chemical reaction and matrix formation of geopolymer concrete compared to conventional
cement concrete, the bond properties of geopolymer concrete should be clearly understood
before it is considered to be suitable to be used to replace conventional cement concrete in
reinforced concrete structures.

Due to the importance of bonding properties for structural members, researches have
been undertaken to evaluate the bond strength between reinforcement and geopolymer
concrete. Sofi et al. initiated the research on steel-geopolymer concrete bond behaviour through
beam-end testing and direct pull-out testing. It was found that on average, the bond strengths
of the fly ash-slag goepolymer concrete produced were 7.3–11.4 MPa and 10.5–14.7 MPa for
the beam-end and direct pullout testing, respectively. Based on the beam-end bond test results,
Sofi et al. concluded that the recommendations in standards such as AS 3600 , ACI 318 and
EC2 could be used to safely predict the development length of geopolymer concrete as these
codes were conservative in predicting the bond strength. Using results from lap-spliced beams,
Chang et al. also found that the code of provisions such as AS 3600 and ACI 318 gave
conservative prediction of the bond strength of the lap-spliced geopolymer concrete beams.
Kim and Park and Topark-Ngarm et al. suggested the following Eqs. (1) and (2), respectively
which would give closer match to the experimental bond strength of geopolymer concrete:

𝑐 𝜙
𝜎 = 𝑓𝑐′ (2.07 + 0.2 (𝜙) + 4.15 (𝑙 )) (1)
𝑑

σ = 2.12(ƒ'c)0.5 (2)
5

where σ is the bond strength (MPa),


𝑓𝑐′ is the cylinder compressive strength (MPa),
c is the concrete cover (mm), ϕ is the bar diameter (mm) and
Ɩd is the development length (mm).

The bonding behaviour between steel reinforcement with fly ash-based geopolymer
concrete and cement concrete in beam-end specimens was also later carried out in few
investigations. The obtained bond strengths of geopolymer concrete were between 10.6 and
19.4 MPa, the bond strength of geopolymer concrete was found to be higher than the cement
concrete. Similarly, in a separate investigation, Castel and Foster who conducted direct pull-
out test on fly ash-slag geopolymer concrete, found that the bond strength of geopolymer
concrete was on average 10% higher than cement concrete. Fernandez-Jimenez et al. observed
that the steel-geopolymer concrete bond was so strong that the failure mode observed in the
pull-out testing was through rupture of the steel bar, whereas pull-out failure was found in the
case of the steel bar bonded to normal cement-based concrete. Thus, Castel and Foster agreed
that existing bond models could be applied for the case of geopolymer concrete due to the
similarity in the bond-slip diagrams for both types of concrete.

In terms of the effect of constituent material on the bond strength of geopolymer


concrete, Sofi et al. commented that the geopolymer concrete containing coarse aggregates
gave higher bond strength than the concrete without coarse aggregate, and the former gave
similar bond strength as normal cement concrete. In addition, the types of fly ash used in the
geopolymer concrete was observed to have minor effect on the bond strength. Ganesan at el.
Investigate the effect of steel fibre addition on the bond strength of geopolymer concrete in
which he was found that depending on the cover/bar diameter ratio, the addition of steel fibres
had varying effects. For instance, for specimens with smaller bar diameter, the addition of up
to 1.0% steel fibres had positive effect in enhancing the bond strength, while in the case of
larger bar diameters, the positive effect of steel fibres diminished due to the local disturbance
caused which prevented proper compaction and thus poorer bond between the geopolymer
concrete and steel bar. Based on the experimental results, Ganesan et al. introduced a modified
bond strength equation based on Orangun et al. which takes into account the fibre reinforcing
index as follow:
6

𝜎 = 𝜎𝑡ℎ (0.009𝐼𝑓2 + 0.022𝐼𝑓 + 1.4)


where 𝐼𝑓 is the fibre reinforcing index given by 𝐼𝑓 = 𝜏𝑉𝑓 (𝑙𝑓 ⁄𝑑𝑓 ) in which 𝜏 is the matrix
interfacial bond stress (MPa), 𝑉𝑓 is the volume of fibres (%), lf is the fibre length (mm) and 𝑑𝑓
is the diameter of fibre (mm); 𝜎𝑡ℎ is the bond equation proposed by Orangun et al. given by
𝜎𝑡ℎ = 0.083[1.2 + 3(c/𝜙) + 50(𝜙/𝑙𝑑 ) + (𝐴𝑡 𝑓𝑦𝑡 /500 s𝜙)](𝑓𝑐′ )1.2 in which 𝐴𝑡 is the area of
transverse reinforcement (mm2), 𝑓𝑦𝑡 is the yield strength of the transverse reinforcement (MPa)
and s is the spacing of the transverse reinforcement (mm).

4. REINFORCED CONCRETE BEAM

Works on the structural behaviour of fly ash-based geopolymer concrete beams were
initiated by Sumajouw et al. where a total of six under-reinforced concrete beams with varying
reinforcement ratios (0.64–2.69%) were tested for flexural failure. As expected, the flexural
load-carrying capacity increased with increase in the tensile reinforcement ratio. In the
investigation by Sumajouw et al., sixteen reinforced geopolymer concrete beams (Fig. 1) with
varying tensile reinforcement ratio (0.64–2.69%) and concrete compressive strength (37–76
MPa) were tested. Sumajouw et al. reported that the general behaviour of the geopolymer
concrete beams was similar to conventional cement based concrete beams in terms of effect of
tensile reinforcement ratio on flexural capacity and ductility index.

Fig. 1. Geopolymer concrete beams by Sumajouw et al.


7

In other researches, it was also reported that under-reinforced fly ash-based


geopolymer concrete beams behaved similarly as conventional reinforced concrete beams
subjected to flexural loading. On the other hand, Yost et al. found a more brittle failure of
geopolymer concrete beams during concrete crushing compared to conventional cement-based
concrete. In contrast, Jeyasehar et al. found higher first crack load, midspan deflection and
ultimate load as well as smaller crack width for the case of reinforced geopolymer concrete
beams as compared to conventional cement-based concrete beams.

Prachasaree et al. introduced equivalent stress block parameters meant for fly ash-
based geopolymer concrete which gave good agreement with experimental findings for
geopolymer concrete beams. Prachasaree et al. reported that the proposed design parameters
could be used with the design procedure in ACI 318 and AS 3600. In the proposed method,
firstly, a simplified stress-strain model was proposed for geopolymer concrete using modified
Popovics equation below:
𝑓𝑐 ′
𝜀𝑐 𝑛𝑘
= 𝜀𝑐 𝑛 [𝜀𝑐 [𝑛 − 1 + ( ) ]]
𝑓𝑐′ 𝜀𝑐′

where 𝑓𝑐 is the compressive stress (MPa), 𝜀𝑐 is the concrete strain, 𝜀𝑐′ is the strain corresponding
to the maximum compressive stress given by 𝜀𝑐′ = 0.0051 − 4(𝑓𝑐′ )/105 based on experimental
data, n is the curve fitting factor given by n = 0.5 + (𝑓𝑐′ /14.3) − [3(𝑓𝑐′ )2 /104 ] and k is a factor
𝜀
whereby k = 1 when 𝜀𝑐′ <1 and k > 1 otherwise.
𝑐

Secondly, based on the modified stress-strain equation, Prachasaree et al. proposed the
following flexural design parameters 𝑘1 , 𝑘2 and 𝑘3 for the determination of the equivalent
stress block for the case of geopolymer concrete and hence the nominal moment capacity of
geopolymer concrete beams could be determined through standard design procedures using
these proposed parameters.
𝑓𝑐′
𝑘2 = 0.384 − ( 3 )
10
𝑓𝑐′
𝑘1 𝑘3 = 1.070 − ( ) + 9(𝑓𝑐′ )2 /105
76.3
8

where k1 and k3 are the equivalent stress block parameter and the parameter k2 defines the
centroid of the compressive forces.

Utilizing ANSYS programme to conduct numerical analysis to predict the flexural


behaviours of under-reinforced geopolymer concrete beams, Kumaravel et al. and Kumaravel
and Thirugnanasambandam found good comparison of the predicted and experimental load-
deflection relationships. Based on these researches, it was suggested that the ANSYS software
could be a useful tool in simulating the behaviour of structural members made of geopolymer
concrete, and this could benefit design engineers dealing with reinforced geopolymer concrete
members in the future.

Researchers also explored the structural behaviour of under-reinforced geopolymer


concrete beams containing different concrete materials. For instance, Andalib et al.
incorporated 30% palm oil fuel ash (POFA) into the geopolymer concrete to produce
geopolymer concrete beams and they observed similar cracking and ultimate moments as well
as crack pattern as conventional reinforced concrete beams. In another research, Srinivasan et
al. evaluated the effects of glass fibre addition on the flexural behaviour of reinforced
geopolymer concrete beams. Similarly, the flexural load capacity of the geopolymer concrete
beams was found to be increased by up to 35% in the presence of up to 0.02% volume fibres
which was attributed to the increase in tensile strain carrying capacity of the concrete. Further
addition of glass fibres to 0.04% led to reduction in the flexural load bearing capacity due to
induced voids within the concrete.

The flexural behaviour of reinforced geopolymer concrete beams subjected to


corrosion was also evaluated. Under accelerated corrosion in sodium chloride solution,
Wanchai found that the reinforced fly ash-based geopolymer concrete beams exhibited greater
degradation in the flexural capacity compared to the beam containing conventional cement-
based concrete. When immersed in sulphuric acid and combination of hydrochloric and
sulphuric acid solutions for 180 days, Kannapiran et al. observed little reduction (less than 8%)
in the flexural capacity of reinforced concrete beams and no significant changes to the load-
deflection relationship.

There are few research works carried out to evaluate the shear behaviour of reinforced
geopolymer concrete beams under flexural loading. Considering the similar crack shape and
9

failure mode, Yost et al. reported that the shear force transfer is similar in both geopolymer and
cement-based concrete beams. The shear strength of the reinforced concrete beams were also
similar for both types of concrete. On the other hand, Mourougane et al. observed higher shear
strength for the reinforced geopolymer concrete beams than the corresponding conventional
cement-based concrete beams, in the range of 5–23%.

Ng et al. found that as steel fibres were added into the geopolymer concrete, the
shear cracking of the resulting reinforced concrete beams was delayed, and more but finer
cracks were formed in the specimens. In addition, use of straight steel fibres resulted in smaller
crack width compared to the addition of hooked-end steel fibres in the geopolymer concrete
beams and this was due to the smaller diameter of the straight steel fibres.

5. REINFORCED CONCRETE COLUMN

One of the most important structural members is reinforced concrete column,


designed to carry compressive axial loading. Reinforced concrete column is generally divided
into two categories – short and slender columns. Due to architectural aesthetics and efficiency
in the use of working space, slender columns are used in building structures around the world.
In the effort to promote utilization of geopolymer concrete in reinforced concrete structural
members, several research works have been carried out in the past to evaluate the performance
of geopolymer concrete in slender reinforced concrete columns. This is considered to be vital
as the composition of geopolymer concrete is different compared to conventional cement-based
concrete and the structural performance of such concrete in compliance with codes of practices
should be determined such that the use of such concrete in actual structures could be realized.

Sujatha et al. fabricated and tested 12 slender circular reinforced concrete columns
made of geopolymer and cement-based concretes of compressive strength grades 30 and 50
MPa. The reinforced concrete columns were tested under compressive axial loading without
eccentricity. It was found that the geopolymer concrete columns had up to 34% higher load-
carrying capacity as well as having greater rigidity compared to the corresponding cement-
based concrete columns. Ganesan et al. investigated the effect of steel fibre addition on the
behaviour of slender square geopolymer concrete columns with 2.01% reinforcement ratio. In
this research, the effects of different volume of steel fibres (up to 1.0%) as well as aspect ratio
10

(l/d) of the slender columns were investigated. The slender columns were tested under
monotonic axial loading. It was found that the inclusion of steel fibres increased the load-
carrying capacity of the geopolymer concrete columns by up to 56%, and this was due to the
fibre-bridging effect which prevented early concrete cover spalling. Increase in strain at the
peak axial compressive stress and area under the stress-strain curve suggested that there was
considerable improvement in the ductility (up to 29% increase) of geopolymer concrete column
when steel fibres were added.

Research works were also carried out on slender fly ash-based geopolymer
concrete column under load eccentricity carried out in Curtin University of Technology in
Australia. The set-up of the column test are shown in Fig. 2. In these researches, the slender
columns tested had longitudinal reinforcement ratios of 1.47% and 2.95% and targeted concrete
strength grades of 40 and 60 MPa; the column specimens were tested at specified varying load
eccentricities from 15 to 50 mm (Table 1).

Fig. 2. Slender geopolymer concrete column test


11

Table 1. Details of slender geopolymer column test under load eccentricities


Column Concrete Load Longitudinal reinforcement
compressive eccentricity Bars Ratio (%)
strength (MPa) (mm)
GCI-1 42 15 4Y12 1.47
GCI-2 42 35 4Y12 1.47
GCI-3 42 50 4Y12 1.47
GCI-4 43 15 8Y12 2.95
GCI-5 43 35 8Y12 2.95
GCI-6 43 50 8Y12 2.95
GCII-1 66 15 4Y12 1.47
GCII-2 66 35 4Y12 1.47
GCII-3 66 50 4Y12 1.47
GCII-4 59 15 8Y12 2.95
GCII-5 59 35 8Y12 2.95
GCII-6 59 50 8Y12 2.95

The columns had similar failure mode characterized by crushing of concrete


in the compressed face near the mid-height of the columns (Fig. 3). Brittle failure mode was
observed in columns with smaller load eccentricity, higher concrete strength and higher
reinforcement ratio. Similarly, the load-carrying capacity of the columns were increased with
decrease in load eccentricity, increase in concrete strength and longitudinal reinforcement ratio.
On the other hand, the mid-height deflection of the tested geopolymer columns increased with
the increase in load eccentricity, decrease in concrete strength and reinforcement ratio.
In the experimental works carried out by Sumajouw et al., the load-carrying
capacity of the slender geopolymer concrete column was compared with the design provisions
from AS 3600 and ACI 318 which demonstrated that the design provisions can be used in the
case of the slender geopolymer concrete column.
12

Fig. 3. Failure mode of slender geopolymer concrete columns

6. CONCRETE FILLED STEEL TUBULAR (CFT) COLUMN

Concrete filled steel tubular (CFT) columns have been increasingly used in structures
such as bridges, high-rise buildings, transmission towers and warehouses etc. This is due to the
excellent structural behaviour of CFT columns such as high strength, high ductility, high
stiffness and full usage of construction materials. Shi et al. utilized geopolymer recycled
concrete as concrete fill in steel tubular columns and tested the structural behaviour of the CFT
columns. It was reported that in the geopolymer CFT columns, the load capacity was reduced
in the increased ratio of recycled concrete as coarse aggregates, and this reduction was more
significant compared to in cement-based CFT columns. On the other hand, the effect of
recycled concrete in geopolymer CFT columns was found to be more sensitive in enhancing
the ductility of the column compared to the corresponding cement-based CFT columns.
13

Espinos et al. demonstrated that when geopolymer concrete was used as concrete infill in
conventional CFT columns, there was no particular effect; however, when used as outer core
concrete infill in an innovative double tube CFT (Fig. 4), the fire resistance time was
significantly delayed in comparison with using conventional concrete. The improvement in the
fire resistance was caused by the delay in the temperature rise in the inner core conventional
concrete as the outer core geopolymer concrete had lower thermal conductivity. On the other
hand, poorer performance was observed when geopolymer concrete was used as inner core and
conventional concrete in the outer core of the double tube CFT due to the quicker rise of
temperature in the outer tube.

Fig. 4. Double tube CFT by Espinos et al.

7. REINFORCED CONCRETE PANEL

Reinforced concrete wall panels are considered as important structural member as


beams, slabs and columns. However, reinforced concrete wall has slenderness effect, leading
to stability problems. Therefore, Ganesan et al. investigated the strength performance of
reinforced concrete wall panel made using geopolymer concrete. It was observed that the main
difference between the reinforced geopolymer concrete and cement-based concrete wall panels
when subjected to axial loading was that the geopolymer concrete showed greater deflection
14

even though the load-carrying capacity was similar. It was noted that ACI 318 gave
conservative prediction of the ultimate strength of the geopolymer concrete wall panels and
hence Ganesan et al. proposed the following equation to predict the ultimate strength of the
geopolymer concrete wall panels.
ℎ ℎ 2 ℎ
𝑃𝑢 = 0.585[𝑓𝑐′ 𝐿𝑡 + (𝑓𝑦 − 𝑓𝑐′ )𝐴𝑠𝑐 ] [1 + (40𝑡) − (30𝑡) ] [1 − (18𝑡)]

where 𝑃𝑢 is the ultimate load (kN), L is the length of panel (mm), t is the thickness of panel
(mm), 𝑓𝑦 is the strength of steel reinforcement (MPa), 𝐴𝑠𝑐 is the area of steel reinforcement
(mm2), h is the height of panel (mm).

Sarker and Macbeath evaluated the fire endurance performance of reinforced fly
ash-based geopolymer concrete panels and found that the heat transfer rate of the geopolymer
concrete panel was greater than the corresponding cement-based concrete panel when exposed
to high temperature of up to 1000 °C. However, the damage to the specimen after high
temperature exposure was less severe in the case of the geopolymer concrete panel. Because
of this, the residual-to-original strength ratios of the geopolymer concrete panels were higher
at 0.61–0.71 compared to those of the cement-based concrete panels which were about 0.50–
0.53. The results suggested superior fire endurance of reinforced geopolymer concrete panels
compared to the cement-based concrete specimens. The conformity of structural geopolymer
concrete panels towards fire resistance requirements in AS 1530 was also reported by Aldred
and Day.

8. OTHER REINFORCED CONCRETE STRUCTURES

8.1 Reinforced Ferrocement Slab

Mohana and Nagan carried out flexural tests on geopolymer reinforced ferrocement
slabs and compared the performance with conventional ferrocement slabs. It was found that
the cracking load, yielding load and ultimate load of the geopolymer ferrocement slabs were
all higher compared to the corresponding conventional ferrocement slabs. Also, the
geopolymer ferrocement slabs could sustain larger deflection at yield and failure. In terms of
cracking behaviour, the number of cracks was observed to be more in the case of the
geopolymer ferrocement slab while the average crack width and spacing were higher in the
15

case of conventional ferrocement slab. A series of 24 geopolymer ferrocement slabs with


varying volume fractions of reinforcement and type of reinforcement were tested under drop
weight impact loading in an experimental investigation carried out by Nagan and Mohana. It
was found that there were similar failure patterns of the slabs made of conventional ferrocement
and geopolymer ferrocement. It was noted that the best performance was observed in the
geopolymer ferrocement slab reinforced with a combination of 4 layers of chicken mesh and 1
layer of weld mesh.

8.2 Reinforced Concrete Pipes

When reinforced concrete pipes were tested under three-edge bearing test,
Shrestha found that the pipe made of geopolymer concrete exhibited more uniformly
distributed crack lines with smaller crack widths, as compared to the wider crack line of the

Fig. 5. Failure mode of (a) reinforced geopolymer concrete pipe and (b) conventional
reinforced concrete pipe
16

that sewer pipes could be fabricated using reinforced geopolymer concrete and the pipes passed
the Australian Standard for load-carrying strength, with the ability to withstand considerable
internal hydrostatic pressures using geopolymer concrete with 7-day compressive strength of
40–60 MPa.

8.3 Prestressed Concrete Railway Sleepers

Gourley and Johnson highlighted that high load bearing precast railway sleepers
made of geopolymer concrete (compressive strength of 60–80 MPa) passed the requirements
of Australian Standard for static and cyclic loading with ease. An advantage observed in the
use of geopolymer concrete railway sleepers was that the steel-concrete bond was so great that
there was no steel slippage at the ultimate load whereas in conventional design, steel wire failed
in tension before slippage could occur. However, the geopolymer prestressed sleeper had
comparatively poorer demoulding performance compared to that for prestressed sleeper from
conventional cement-based concrete, thus requiring automatic demoulding machine with
additional hammering vibration.

8.4 Median Barrier

Dhakal et al. fabricated a New Jersey type median barrier and tested the barrier to
failure. It was found that the performance of geopolymer median barrier was in compliance
with AASHTO and therefore could be used in most roads. In addition, the investigation showed
that the use of structural analysis and design approach for conventional cement-based concrete
structures can be adopted for geopolymer concrete structures.

9. FIELD APPLICATION

Reinforced geopolymer concrete application had been utilized for actual field
application, by using both precast geopolymer concrete elements and casting in-situ of the
geopolymer concrete. Examples of these include precast retaining wall, precast bridge decks
and boat ramp (Fig. 6).
17

Fig.6 (a) Precast retaining wall

Fig.6 (b) Precast bridge deck

Fig.6 (c) Precast boat ramp


18

In 2013, being the first application of geopolymer concrete in multi-storey building, precast
geopolymer concrete floor beams were utilized as structural floor elements in the construction
of The University of Queensland’s Global Change Institute, Australia. (Fig. 7)

Fig.7 (a) Precast geopolymer concrete beam

Fig.7 (b) Global Change Institute


19

10. CONCLUSION

It is concluded that the geopolymer concrete members such as beams and columns
could be designed using design codes for conventional reinforced concrete members as most
of the codes gave conservative estimation of the ultimate load-capacity of the geopolymer
concrete members.

In addition, due to the similarity in the general structural behaviours such as load-
deflection, cracking characteristics and failure mode of the geopolymer concrete members with
conventional concrete members, it can be concluded that geopolymer concrete members could
be designed in the same way as conventional concrete members.

Although there are a number of design equations meant for geopolymer concrete
structures have been proposed, but still these are fairly limited and therefore there are still
opportunities in further researching the structural behaviour of geopolymer concrete in order
to develop a standard design method for reinforced geopolymer concrete member which are
more economic, effective and realistic. This is essential in order to fully introduce geopolymer
concrete for large-scale structural applications in the future, which would ultimately result in a
more environmental-friendly and sustainable construction industry.
20

REFERENCES

1. Australian Standard AS 3600, Concrete Structures, Standards Australia, Sydney, 2004.


2. ACI Committee 318, Building Code Requirements for Structural Concrete (ACI 318-
02), American Concrete Institute, Farmington Hills, Mich., 2002.
3. M. Sofi, J.S.J. van Deventer, P.A. Mendis, G.C. Lukey, Bond performance of
reinforcing bars in inorganic polymer concrete (IPC), J. Mater. Sci. 42 (2007) 3107–
3116.
4. E.H. Chang, P. Sarker, N. Lloyd, B.V. Rangan, Bond behaviour of reinforced fly ash-
based geopolymer concrete beams, in: R.I. Gilbert (Ed.), Concrete Solutions 09, The
24th Biennial Conference of the Concrete Institute of Australia, Luna Park, Sydney,
Concrete Institute of Australia, 2009.
5. J.S. Kim, J.H. Park, An experimental evaluation of development length of
reinforcements embedded in geopolymer concrete, Appl. Mech. Mater. 578-579 (2014)
441–444.
6. P. Topark-Ngarm, P. Chindaprasirt, V. Sata, Setting time, strength and bond of high-
calcium fly ash geopolymer concrete, J. Mater. Civ. Eng. (2014)
7. A. Castel, S.J. Foster, Bond strength between blended slag and class F fly ash
geopolymer concrete with steel reinforcement, Cem. Concr. Res. 72 (2015) 48–53.
8. A.M. Fernandez-Jimenez, A. Palomo, Lopez-Hombrados, Engineering properties of
alkali-activated fly ash concrete, ACI Mater. J. 103 (2) (2006) 106–112.
9. N. Ganesan, A. Santhakumar, P.V. Indira, Bond behaviour of reinforcing bars
embedded in steel fibre reinforced geopolymer concrete, Mag. Concr. Res. 67 (1)
(2015) 9–16.
10. C.O. Orangun, J.O. Jirsa, J.E. Breen, A reevaluation of test data on development length
and splices, ACI J. Proc. 74 (3) (1977) 114–122.
11. D.M.J. Sumajouw, D. Hardjito, S.E. Wallah, B.V. Rangan, Behaviour and strength of
reinforced fly ash-based geopolymer concrete beams, in: Proceedings of K.H. Mo et al.
/ Construction and Building Materials 120 (2016) 251–264 263 Australian Structural
Engineering Conference, 11–14 September, Newcastle, Australia.
12. J.R. Yost, A. Radlinska, S. Ernst, M. Salera, N.J. Martignetti, Structural behaviour of
alkali activated fly ash concrete. Part 2: structural testing and experimental findings,
Mater. Struct. 46 (2013) 449–462.
21

13. C.A. Jeyasehar, G. Saravanan, M. Salahuddin, S. Thirugnanasambandam,


Development of fly ash based geopolymer precast concrete elements, Asian J. Civ. Eng.
14 (4) (2013) 605–615.
14. W. Prachasaree, S. Limkatanyu, A. Hawa, A. Samakrattakit, Development of
equivalent stress block parameters for fly-ash-based geopolymer concrete, Arab. J. Sci.
Eng. 39 (12) (2014) 8549–8558.
15. S. Kumaravel, S. Thirugnanasambandam, C.A. Jeyasehar, Flexural behaviour of
geopolymer concrete beams with GGBS, IUP J. Struct. Eng. 7 (1) (2014) 45–54.
16. R. Andalib, M.W. Hussin, M.Z.A. Majid, M. Azrin, H.H. Ismail, Structural
performance of sustainable waste palm oil fuel ash-fly ash geo-polymer concrete
beams, Environ. Treat. Technol. 2 (3) (2014) 115–119.
17. S. Srinivasan, A. Karthik, S. Nagan, An investigation on flexural behaviour of glass
fibre reinforced geopolymer concrete beams, Int. J. Eng. Sci. Res. Technol. 3 (4) (2014)
1963–1968.
18. T. Sujatha, K. Kannapiran, S. Nagan, Strength assessment of heat cured geopolymer
concrete slender column, Asian J. Civ. Eng. 13 (5) (2012) 635–646.
19. N. Ganesan, S.D. Raj, R. Abraham, K. Namitha, Effect of fibres on the strength and
behaviour of GPC columns, Mag. Concr. Res. (2015), http://dx.doi.org/
10.1680/macr.15.00049.
20. X. Shi, Q. Wang, X. Zhao, F.G. Collins, Structural behaviour of geopolymeric recycled
concrete filled steel tubular columns under axial loading, Constr. Build. Mater. 81
(2015) 187–197.
21. N. Ganesan, P.V. Indira, A. Santhakumar, Prediction of ultimate strength of reinforced
geopolymer concrete wall panels in one-way action, Constr. Build. Mater. 48 (2013)
91–97.
22. P.K. Sarker, S. Mcbeath, Fire endurance of steel reinforced fly ash geopolymer concrete
elements, Constr. Build. Mater. 90 (2015) 91–98.
23. R. Mohana, S. Nagan, An experimental investigation on the flexural behaviour of
geopolymer ferrocement slabs, J. Eng. Technol. 3 (2) (2013) 97–103.
24. P. Shrestha, Development of geopolymer concrete for precast structures (M.Sc. Eng.
thesis), The University of Texas at Arlington, 2013.
25. M. Dhakal, M. Al-Masud, S. Alam, C. Montes, K. Kupwade-Patil, E. Allouche, A.
Saber, Design, fabrication and testing of a full-scale geopolymer concrete median
barrier Retrieved from <http://www.flyash.info/2013/076-Dhakal- 2013.pdf>2013.
22

You might also like