Sie sind auf Seite 1von 11

Materials and Structures (2015) 48:721–731

DOI 10.1617/s11527-014-0268-9

ORIGINAL ARTICLE

Long term durability properties of class F fly ash


geopolymer concrete
David W. Law • Andi Arham Adam • Thomas K. Molyneaux •

Indubhushan Patnaikuni • Arie Wardhono

Received: 9 August 2013 / Accepted: 7 February 2014 / Published online: 19 February 2014
Ó RILEM 2014

Abstract The environmental impact from the pro- scanning electron microscopy and energy dispersive
duction of cement has prompted research into the X-ray spectroscopy. The results showed that both the
development of concretes using 100 % replacement geopolymer concretes with activator modulus 1.00
materials activated by alkali solutions. This paper and 1.25 gave durability parameters comparable to
reports the assessment of a number of key durability Ordinary Portland and blended cement concretes of
parameters for geopolymer concrete made from fly ash similar strength, while the geopolymer concrete with
activated with sodium silicate and sodium hydroxide. an activator modulus of 0.75 displayed lower durabil-
Properties investigated have included workability, ity performance. However, there is a concern over the
compressive strength, water sorptivity, carbonation, long term performance of the geopolymer concretes
chloride diffusion and rapid chloride permeability. with activator modulus of 1.00 and 1.25 when
Microstructure studies have been conducted using considering chloride induced corrosion of reinforcing
steel due to the initial pH and long term chloride
diffusion coefficient.
D. W. Law (&)  T. K. Molyneaux  I. Patnaikuni 
A. Wardhono
Keywords Geopolymers  Durability 
School of Civil, Environmental and Chemical
Engineering, RMIT University, Melbourne, VIC 3000, Microstructure  Permeability  Carbonation 
Australia Chloride
e-mail: david.law@rmit.edu.au
T. K. Molyneaux
e-mail: tom.molyneaux@rmit.edu.au
I. Patnaikuni 1 Introduction
e-mail: indu.patnaikuni@rmit.edu.au
A. Wardhono It is widely known that the production of Portland
e-mail: arie.wardhono@rmit.edu.au cement consumes considerable energy and at the same
time contributes a large volume of CO2 to the
A. A. Adam
Department of Civil Engineering, Tadulako University, atmosphere. The calcination of CaCO3 to produce
Palu, Indonesia 1 ton of ordinary Portland Cement (PC) releases
e-mail: adam.arham@gmail.com 0.53 tons of CO2 into the atmosphere, and if the
energy source used in the production of PC is carbon
A. Wardhono
Department of Civil Engineering, State University of fuel then another 0.45 tons of CO2 are produced [33].
Surabaya, Surabaya, Indonesia Therefore the production of 1 ton of PC produces
722 Materials and Structures (2015) 48:721–731

approximately 1 ton of CO2 in the atmosphere. Table 1 Chemical composition of class F fly ash
Increased environmental concerns coupled with Component Percentage
increased development and use of concrete in con-
struction has prompted a search for more environ- SiO2 49.45
mentally friendly materials. Al2O3 29.61
A possible alternative is the use of alkali-activated Fe2O3 10.72
binder using industrial by-products containing silicate CaO 3.47
materials [25, 40]. One of the most common industrial MgO 1.3
by-products used as binder materials is fly ash (FA). K2O 0.54
FA has been widely used as a pozzolanic material to Na2O 0.31
enhance the physical, chemical and mechanical prop- TiO2 1.76
erties of cements and concretes. It has been estimated P2O5 0.53
that the energy requirement of geopolymers is 60 % Mn2O3 0.17
less than that of PC [34]. SO3 0.27
Previous research has shown that it is possible to LOI 1.45
use 100 % FA as the binder by activating it with an
alkali component, such as; caustic alkalis, silicate
salts, and non silicate salts of weak acids [9, 48]. experiments analysing the strength gain and durability
Activation of FA involves using a highly alkaline properties of geopolymer concrete.
solution. This reaction forms an inorganic binder
through a polymerization process [14, 19]. The term
‘‘Geopolymeric’’ is used to characterise this type of 2 Materials and methodology
reaction. The geopolymeric reaction differentiates
geopolymers from other types of alkali activated A range of geopolymer concrete and mortar specimens
materials (such as; alkali activated slag). were cast varying the mix composition and activator
Research has shown that geopolymer concrete can modulus. Testing included compressive strength,
achieve comparable strengths to Ordinary Portland sorptivity, chloride diffusion, rapid chloride perme-
(OP) and blended cement concretes [10, 26, 27, 30, 39, ability and carbonation. In addition, scanning electron
41, 49]. Durability studies have investigated acid microscopy (SEM) and energy dispersive X-ray
attack [11, 46], sulphate attack [12] and fire resistance analysis were also undertaken.
[32], while Miranda has investigated the chloride The FA used to manufacture the geopolymer
resistance of geopolymer mortar specimens [35]. The concrete was a low calcium FA (class F fly ash)
research has shown that geopolymer concrete has the conforming to AS 3582.1-1998 [4], Table 1. The
potential to be a suitable alternative material to OP and fineness of the FA had 86.82 % passing a 45 lm sieve.
blended cement concretes. However, the long term The SO3 is less than 1 % which should ensure high
durability of the material has yet to be established, volume stability which is desirable for good durability.
particularly with reference to the protection of rein- The alkaline activator used in this study was a sodium
forcing steel. silicate based solution containing sodium silicate
In a normal exposure environment with proper solution (wt. ratio: Na2O/SiO2 = 2) and 10 M sodium
design and production, concrete made with PC can be hydroxide. The chemical composition of the sodium
a durable material. However, it has been long recog- silicate solution by mass was Na2O = 14.7 %,
nized that traditional concrete can suffer from deteri- SiO2 = 29.4 % and water = 55.9 %.
oration due to the attack from aggressive agents such The blended sodium silicate and sodium hydroxide
as chloride and acidic gases such as CO2 that initiate solutions are characterized by the activator modulus
corrosion in the reinforcing steel. In order to be a (Ms), which is defined as the mass ratio of SiO2 to
successful alternative to PC and blended cement Na2O; and the dosage, corresponding to the percentage
concretes, geopolymer concrete must show similar Na2O. The adopted nomenclature for the specimens
durability properties and be able to resist attack from was GX-Y, where X is dosage and Y is Ms, i.e. G7.5-
these aggressive agents. This paper reports a set of 1.0 for a 7.5 % dosage and a 1.0 activator modulus.
Materials and Structures (2015) 48:721–731 723

Table 2 Mix proportion of fly ash-based geopolymer concrete


Mix Fly ash (kg) Aggregate (kg) Activator (kg) Added water (kg) w/b
Sand (7 mm) (10 mm) Na2SiO3 (liquid) NaOH (10 M)

G7.5-0.75 1,050 1,728 763 1,528 198 209 88 0.34


G7.5-1.00 1,030 1,728 763 1,528 262 165 84 0.32
G7.5-1.25 1,016 1,728 763 1,528 324 123 79 0.32
Quantities for kg/m3

Different dosage and activator modulus are 90 °C. This temperature can be reduced by using high
achieved by varying the ratio of sodium silicate and calcium (class C) FA [27] or by adding calcium
sodium hydroxide. The sodium hydroxide solution compound [49] as the higher calcium content can
was prepared in a fume cupboard by dissolving produce C–S–H which can be cured at room temper-
sodium hydroxide pellets in deionised water at least ature. As these experiments used a 100 % class F fly
1 day prior to mixing. ash, a curing temperature of 90 °C was adopted [3].
Both coarse and fine aggregate were prepared in The slump test was undertaken in accordance with
accordance with AS 1141.5-2000 [5]. The moisture Australian Standard AS 1012.1-1993 [6]. Compres-
condition of the aggregate was in a saturated surface sive strength tests were performed with a loading rate
dry condition. The fine aggregate was river sand in of 20 MPa/min according to AS 1012.9-1999. Sorp-
uncrushed form. The coarse aggregate was crushed tivity tests were undertaken with 100 mm diameter
basalt aggregate with a specific gravity of 2.99. and 50 mm height specimens in accordance with
The mix design was based on previous tests on ASTM C1585-04 [7]. The sides of the specimens were
mortar specimens [3]. The total aggregate in the coated with epoxy to allow free water movement only
concrete was kept to 64 % of the entire mixture by through the bottom face (unidirectional flow). The
volume for all mixes. The ratio of ingredients results were plotted against the square root of the time
(cementitious materials, chemical activator, aggre- to obtain a slope of the best fit straight line. According
gate, and water) was calculated based on the absolute to Hall [28], the penetration of water under capillary
volume method [37], as a result, the total weight of action (sorptivity, S) can be modeled by:
binder and water was varied to keep the volume of I ¼ A þ St1=2 ð1Þ
material and water/binder or water/solid ratio con-
stant. Table 2 summarizes the mix details for the where I is the cumulative absorbed volume after time
geopolymer concrete. t per unit area of inflow surface, I = Dw/ar, Dw being
The concrete mix was designed to provide a 28 day the increase in weight, a the cross-sectional area and
compressive strength of 40 MPa. This target strength r the density of water.
was chosen to replicate the strength for standard site Chloride diffusion tests were based on AASHTO
concrete as specified in AS 3600. To achieve this T259 [1] and rapid chloride permeability testing
target strength a 7.5 % Na2O dosage was selected. The (RCPT) was performed according to ASTM C1202-
water in the mix was taken as the sum of water 07 [8] and AASHTO T277 [2]. Accelerated carbon-
contained in the sodium silicate, sodium hydroxide ation tests were undertaken in a purpose built carbon-
and added water. The solid is taken as the sum of FA, ation chamber, Fig. 1 at a CO2 concentration of 5 %, a
the solid in the sodium silicate solution and the sodium temperature of 20 ± 1 °C and a relative humidity of
oxide pellets. Liquid sodium silicate and sodium 70 ± 1 % for 28 days. The CO2 concentration was
hydroxide were blended in different proportions, correlated relative to the O2 concentration in the
providing an activator modulus (Ms) ranging from chamber. Testing for carbonation depth was under-
0.75 to 1.25. In general class F fly ash based taken using Phenolphthalein indicator. Pore water
geopolymer concretes exhibit longer setting times from mortar samples was obtained using a purpose
and slower strength development at room temperature built pore press. The pH was measured electronically
[50], with curing generally being between 60 and using a pH electrode.
724 Materials and Structures (2015) 48:721–731

Fig. 1 Schematic of
accelerated carbonation
chamber

The microstructure was observed using SEM imag- The strengths of the FA-based geopolymer concrete
ing employing both secondary and backscatter electron are shown in Table 3. There was a significant increase
detectors. To prepare the samples for SEM analysis the in strength between MS = 0.75 and 1.0. However,
specimens were cut using a diamond saw to a size of there was only a small further increase to the
2–4 mm in height and 5–10 mm in diameter. The MS = 1.25 concrete. This is attributed to an increase
samples were subsequently gold coated for imaging. in the alkali modulus leading to an increase in soluble
Samples were mounted on the SEM sample stage with silicates and consequently an increase in the reaction
conductive, double-sided carbon tape. A total of 3 rate. The proposed reaction mechanism for geopoly-
samples were investigated for each mix. merisation is a multistage process [23] consisting of
five stages: (1) dissolution, (2) speciation equilibrium,
(3) gelation, (4) reorganization, and (5) polymeriza-
3 Results and discussion tion and hardening. The dissolution process starts with
an attack on the FA particles by alkaline solution [24].
3.1 Material properties As a result the reaction product is generated both
inside and outside the shell of the FA particle sphere
The FA-based geopolymer displayed a very high until the ash particle is completely or almost com-
slump with all mixes giving a slump in excess of pletely consumed. At the same time, precipitations of
200 mm. This is attributed to the spherical shape of FA reaction products occur as the alkaline solution
particles combined with the lubricating effect of penetrates the larger sphere and fills up the interior
sodium silicate solution increasing the flowability. space with reaction product, forming a dense matrix.
The high slump across the mix types was not crucial Due to the precipitation of reaction products, some
as the basic requirements were that the fresh concrete portions of smaller particles are covered with the
did not segregate when vibrated and little bleeding products providing crust which prevents contact with
occurred, both of which requirements were fulfilled by the alkaline solution resulting in unreacted FA parti-
the geopolymer concrete. Following heat curing the cles. The increase in strength observed between
geopolymer mix produced a concrete of comparable MS = 0.75 and MS = 1.0 is attributed to an increase
appearance to a standard OP concrete. in the dissolution process, thus resulting in a higher
Materials and Structures (2015) 48:721–731 725

Table 3 Compressive strength of geopolymer concrete 2.5


G7.5-0.75 (56 days)
Mix Compressive strength (MPa) G7.5-1 (56 days)
2.0
G7.5-1.25 (56 days)
7 days 28 days 90 days G7.5-0.75 (90 days)

i (mm3/mm2)
1.5 G7.5-1 (90 days)
G7.5-0.75 39.1 ± 3.5 44.4 ± 3.4 46.1 ± 2.1 G7.5-1.25 (90 days)
G7.5-1.00 51.3 ± 5.2 53.3 ± 2.6 53.6 ± 5.5
1.0
G7.5-1.25 52.5 ± 4.6 56.9 ± 3.3 57.3 ± 2.0

0.5

reaction rate and in fewer unreacted FA particles. The


relatively small variation between MS = 1.0 and 1.25, 0.0
0 5 10 15 20
suggests that little further dissolution of the FA arises 1/2 1/2
t (min )
from the increase in activator modulus. This would
indicate that either all the FA has been dissolved once Fig. 2 Absorption (i) versus square root of time for geopolymer
a MS = 1.0 has been reached, or that the increase does concrete at 56 and 90 days
not result in any further dissolution of the protective
crust on the FA particles. Thus the formation of the Table 4 Sorptivity correlation parameters
[Mz(AlO2)x(SiO2)ynMOHmH2O] gel, the dominant
Mix Sorptivity parameters
step in formation of an amorphous structure of
geopolymers, which relies on the extent of dissolution 56 days 90 days
of alumino-silicate materials, reaches a limiting value Si (mm/min1/2) R Si (mm/min1/2) R
as the activator modulus increases above 1.0.
G7.5-0.75 0.101 0.996 0.101 0.997
As would be expected with heat curing there was only
G7.5-1 0.078 0.996 0.075 0.998
a small increase in strength from 28 to 90 days for all the
specimens, irrespective of activator modulus. Both the G7.5-1.25 0.071 0.997 0.066 0.998
MS = 1.0 and 1.25 concretes achieved a strength in
excess of 50 MPa at 28 days. The compressive The increase in activator modulus from 1.0 to 1.25
strengths observed are in agreement with those observed showed little change in the sorptivity values. These
by other authors using class F fly ash [10, 26, 41, 43]. changes are consistent with those for the strength data,
which also showed an improvement in performance
3.2 Sorptivity from MS 0.75 to 1.0, but little further improvement
from MS 1.0 to 1.25. This improvement can be
The sorptivity curve of the geopolymer concretes at 56 accounted for by an increase in the SiO2 content and
and 90 days are given in Fig. 2. The results show a increased dissolution of the FA, which results in a
non-linearity in the initial stages, which is in agreement higher rate of reaction, leading to a denser structure
with previous research [22], which found that the with a reduction in the porosity of the geopolymer
sorptivity curve of geopolymer concrete was less linear concrete. The correlation coefficients, R, of all the
compared to that of OP and blended cement concretes. sorptivity data exceed 0.98. Overall the sorptivity
During the setting period the geopolymer concrete parameters of the geopolymer concretes are compara-
exhibited an increased degree of bleeding compared to ble to those of OP and blended cement concretes of
the OP concretes. This would be expected to result in similar strengths [22, 38].
high quantities of cement paste at the surface. This high
concentration of cement paste is hypothesised as 3.3 Chloride diffusion
leading to a rapid saturation of the outer paste layer.
Once this layer is saturated the area of absorption is The results are presented in Table 5. The apparent
reduced due to the presence of aggregates. The level of diffusion coefficient (Da) and surface concentration
bleeding was higher in the specimens with an MS of (Cs) were calculated by plotting the chloride profiles
0.75, which corresponds to the highest value of and determining the best fitted curve using Fick’s 2nd
sorptivity, Table 4, and the least linear plot. Law of Diffusion [18].
726 Materials and Structures (2015) 48:721–731

Table 5 Apparent chloride diffusion coefficient and surface Table 6 Rapid chloride permeability test data
chloride value
Mix Solution temperature °C Duration (Min)
Mix Cs (%) Da 9 10-11 m2/s
Initial Final
G7.5-0.75 0.16 3.1
G7.5-0.75 26.1 60 60
G7.5-1 0.17 3.1
G7.5-1 25.1 60 200
G7.5-1.25 0.14 3.7
G7.5-1.25 22.4 60 270

The geopolymer specimens display a similar diffu- to evaluate the performance of geopolymer concrete.
sion coefficient for all activator moduli. The diffusion The RCPT is often used as a standard test to assess the
coefficients of the FA geopolymer concretes are chloride resistance of concrete in severe exposure
comparable with those for OP and blended cement conditions. The heating observed in these tests indi-
concretes, indeed being in the lower range of values cates that the RCPT should not be applied as a method
reported in the literature [13, 17]. This data would for assessing the chloride resistance of geopolymer
indicate a high level of resistance to chloride ingress concretes and an alternative method should be used to
for geopolymer concretes. However, one factor that assess their performance such as the Nord tests NT
should be taken into account when considering long BUILD 433.
term performance is that OP and to a greater extent
blended cement concretes show a reduction in the 3.4 Carbonation
chloride diffusion coefficient with time. This is
represented as the maturity factor, m [13]. This The geopolymer concrete displayed no clear border
improved performance is attributed to on going between the coloured and colourless area when
hydration of the concrete with time. For geopolymer sprayed with phenolphthalein indicator. There was a
concrete, which is produced by heat curing, little if any graduation in colour as the outer part was lighter in
further reaction will take place, illustrated by a colour compared to the inner part. In this case the
minimal increase in compressive strength with time carbonation ‘front’ was not clear, as such it is not
when compared to OP and blended cement concretes possible to measure accurately the penetration depth
cured at ambient temperatures. Hence, it may be using a phenolphthalein indicator. Even after full
expected that little improvement in the diffusion carbonation, some light pink colour was found in the
coefficient will occur over time when compared with outer part of the specimens.
OP and blended cement concretes. This would mean When compared to carbonated OP and blended
that the chloride diffusion coefficient after 20 years concrete the overall colour of non-carbonated geo-
would be similar to that observed in these tests. Thus at polymer concrete was lighter than that observed in
20 years the diffusion coefficients for geopolymer uncontaminated OP or blended cement concretes. This
concretes may be higher than those for OP and blended suggests that the pH of the pore solution of the
cement concretes. carbonated geopolymer concrete was lower and the
RCPT were undertaken on the geopolymer speci- pH was not significantly affected by the CO2. The
mens. However, rapid heating of the specimens was colouring observed is in agreement with Davidovits
observed, Table 6, with all specimens reaching 60 °C who stated that the pH of geopolymer concrete is in the
before the conclusion of the test. The flow of electric region 11.5–12.5 and that the carbonation products of
current through a conductor generates heat which will geopolymer concrete have a minimum pH of 10–10.5
in turn increase the mobility of the ions that carry the which is much higher than the pH from the calcium
current, which will itself raise the total current flow carbonate which can have a pH lower than 9 [20].
producing more heat in a cyclic process. Given the In order to asses the pH of the geopolymer material
rapid rises in temperature observed it can no longer be mortar specimens using a dosage and Ms the same as
assumed that Ohm’s law (V = IR) applies, which is the concrete specimens, together with those with a
the basic principle of the RCPT method. Hence the dosage of 15 % used in previous mortar tests were
RCPT is assessed as not giving data which can be used investigated [3], Table 7. The pore water was obtained
Materials and Structures (2015) 48:721–731 727

Table 7 Mix proportion of fly ash-based geopolymer mortar produced. The OP cements produce Ca(OH)2 and
Mix Fly Fine Activator (kg) Added
C–S–H gel, while the geopolymers produced
ash sand water [Mz(AlO2)x(SiO2)ynMOHmH2O] gel. The pH of
(kg) (kg) Na2SiO3 NaOH (kg) fresh, OP and blended concrete, is above 13 and that
(liquid) (15 M)
of carbonated concrete below 9 [29], due to the
G7.5-0.75 0.523 1.440 0.128 0.101 0.082 formation of calcium carbonate.
G7.5-1.0 0.522 1.431 0.124 0.133 0.064 In the fresh OP cement concrete the Ca(OH)2 and
G7.5-1.25 0.521 1.438 0.167 0.046 0.108 C–S–H gel provide buffering to maintain the pH above
G15-1 0.505 1.388 0.193 0.148 0.046 13. In the geopolymer concrete the [Mz(AlO2)x
G15-1.25 0.500 1.376 0.255 0.117 0.033 (SiO2)ynMOHmH2O] will not provide a similar
G15-1.5 0.496 1.364 0.316 0.087 0.020 buffering. The pH from the mortars indicate that
3 following the geopolymeric reaction a pH in the pore
Quantities for kg/m
solution of around 12 for all geopolymer mortars is
achieved despite the variation in hydroxyl ion
Table 8 Carbonation data for geopolymer mortar specimens, concentrations in the initial mix. This would
pH
indicate that there is no buffering by the [Mz(AlO2)x
Mix pH (SiO2)ynMOHmH2O] comparable to that provided
0 days 3 days 7 days 28 days by the C–S–H gel. Rather the pH of the geopolymers
produced is controlled by the sodium hydroxide from
G7.5-0.75 11.86 11.88 11.01 10.88 the activator that is contained in the pore solution. A
G7.5-1.0 11.94 11.91 11.35 10.46 number of factors are hypothesised as controlling this
G7.5-1.25 11.73 11.71 11.39 10.73 pH, including the pH of the activator, the degree of
G15-1 11.96 11.97 11.50 11.05 reaction that occurs and hence the residual activator
G15-1.25 11.99 11.88 11.50 11.00 remaining and the chemical composition of the binder
G15-1.5 11.97 11.98 11.77 11.23 material, which will determine the composition of the
geopolymeric material produced.
In the geopolymer concrete the carbonation is
from the mortar specimens using a pore press and the hypothesised as the reaction of the sodium hydroxide
pH was tested at different time periods until full with CO2 forming sodium carbonate hydrates. The
carbonation of the specimens had been achieved after result of this is only a minimal reduction of pH to a value
28 days exposure. Spraying with phenolphthalein of approximately pH 11. The geopolymer mortars with
gave similar colours to that obtained for the concrete the lower dosage (7.5 %) having a slightly lower final
specimens. The results are given in Table 8. pH than those with the higher dosage (15 %).
The mortar specimens all had an initial pH of These results are comparable with those reported
between 11.75 and 12, with little variation between the for alkali activated slag concretes which have not
G7.5 or G15 mixes. However, higher initial pH values shown any detrimental effects due to carbonation [21,
have also been reported depending on the activator pH 44] and that final a pH of between 10 and 12, could be
and the raw materials used [36, 42, 45]. The pH expected [15].
reduced slowly with time until full carbonation was Given the results from the concrete and mortar
achieved after 28 days. At 3 days no discernable specimens, coupled with other reported results sug-
difference in pH was detectable, by 7 days a reduction gests that a final pH in the region of pH 11 can be
in pH was clearly measured and by 28 days the G7.5 expected for geoploymer concrete. This is a value that
specimens had a pH in the range 10.4–10.9 and the could provide protection to reinforcing steel following
G15 in the range 11–11.25. These results show a carbonation.
similar initial pH to that reported by Davidovits,
though the final pH is higher than that reported [20]. 3.5 Microstructure
The variation in the initial and final pH recorded for
the mortars compared with that of OP and blended The microstructure of all three geopolymer concrete
cements is attributed to the difference in the materials mixes were similar, Fig. 3. All the geopolymer
728 Materials and Structures (2015) 48:721–731

Fig. 3 Microstructure of the three geopolymer concrete mixes

specimens contained both unreacted FA and silica.


Some particles of FA were also found to have been
partially dissolved by alkali Fig. 4. The most distinc-
tive difference in the microstructure was a denser pore
structure as the modulus increased. Analysis of the
number of unreacted grains in each sample noted a
slight decrease in the number of unreacted FA parti-
cles, with an increase activator modulus, while the
number of partially reacted grains observed remained
similar as the activator modulus increased. This would
indicate that there is an additional dissolution of the FA
due to the increase in activator modulus but that the
increase does not result in any further dissolution of the
protective crust on the FA particles. The decrease in the
number of unreacted FA particles was more pro-
nounced between the G7.5-0.75 to G7.5-1 mixes, than
from the G7.5-1 to the G7.5-1.25 mixes.
Fig. 4 Partially dissolved fly ash grain
According to Steveson and Sagoe-Crentsil [47],
unreacted components in FA-based geopolymer
binder make up a significant proportion of the total 3.6 Long term durability
volume of the binder. These components are compos-
ites, hence the strength of the unreacted particles, the The compressive strength of the geopolymer concrete
interface between them and geopolymer matrix is is in excess of the specified minimum strength in AS
expected to have a significant bearing on the overall 3600 for the design of concrete structures in exposure
strength of the material. The reduction in the number categories A1 and A2 and B1 and B2. This indicates
of unreacted particles of FA as the activator modulus that from a structural perspective the use of geopoly-
increases would explain the higher strength and lower mer concrete is a feasible alternative to OP and
sorptivity observed. blended concretes used at present.
Materials and Structures (2015) 48:721–731 729

However, the long term durability of reinforced similar strength OP and blended cement concretes.
concrete is dependent upon the protection of reinforc- The MS = 0.75 gave the highest sorptivity value while
ing steel in concrete which is provided by the passive the MS = 1.0 and 1.25 gave similar values.
layer formed on the steel surface due to the high pH in The apparent chloride diffusion values were similar
OP and blended cement concretes [16]. It is the for all the geopolymer concretes. The values were
breaking down of this layer by carbonation or favorably comparable to similar OP and blended
chlorides that results in the corrosion of the reinforce- cement concretes, though due to the heat curing there
ment. Thus the initial pH would be expected to have a may be no long term reduction in the diffusion
significant factor in the chloride induced corrosion of coefficient for geopolymer concretes, as compared to
reinforcing steel. The Cl/OH ratio is regarded as one of OP and blended cement concretes. The RCPT results
the controlling factors in the initiation of corrosion in rapid heating in geopolymer concrete.
[31]. The pH produced following carbonation will also Carbonation testing indicated that the initial pH of
be a significant factor in the long term durability of geopolymer concrete is less than that of OP and
geopolymer concrete. blended cement concretes but is higher after carbon-
The data obtained suggests that carbonation of ation. The final pH after carbonation may be sufficient
geopolymers may not be as potentially deleterious as to provide protection for reinforcing steel in carbon-
carbonation of OP and blended cement concretes as ated geopolymer concrete.
the pH remains at a level that can provide protection to SEM analysis showed a denser pore structure with
the reinforcing steel. However, for chloride induced increasing activator modulus, consistent with the
attack the long term protection provided by geopoly- durability test results.
mer concrete may be lower than for OP and blended The durability test program indicates that both the
cement concretes. The lower initial pH may lead to a MS = 1.0 and 1.25 geopolymer concretes gave dura-
lower concentration of chloride ions being required at bility parameters comparable to OP and blended
the rebar to initiate corrosion and, as discussed, the cement concretes of similar strength. However, con-
long term chloride diffusion coefficient may be higher cern exists over the long term performance when
than for OP and blended cement concretes. To considering chloride induced corrosion of reinforcing
determine this effect more clearly longer term testing steel due to the lower initial pH and potentially
of geopolymer samples is required to determine what constant chloride diffusion coefficient.
the maturity factor is for geopolymer concrete. Due to the different composition of the geopolymer
concrete compared to OP and blended concretes the
direct application of the current standards and dura-
4 Conclusions bility test methods is not considered appropriate and
before the use of geopolymer concrete is adopted the
There is a significant increase in strength from the development of a specific standard accounting for the
MS = 0.75 to the MS = 1.0 and 1.25 concrete which mix design and testing of the long term performance is
is attributed to an increase in the dissolution of the FA recommended.
grains and a resultant increase in the reaction rate.
Minimal variation in strength is observed between the Acknowledgments The authors wish to express their thanks
to PQ Australia for the supply of materials for the research
MS = 1.0 and 1.25 concretes and little increase in
project, and Cement Australia Ltd for providing the FA. The
strength is observed for any of the geopolymer authors also wish to acknowledge the facilities, and the scientific
concretes with time, which is attributed to the heat and technical assistance, of the Australian Microscopy &
curing. Microanalysis Research Facility at the RMIT Microscopy &
Microanalysis Facility, at RMIT University.
The geopolymer concretes display a non-linearity
in the sorptivity data in the initial stages compared to
OP and blended cement concretes. This non linearity References
is hypothesized as being due to an increased bleeding
1. AASHTO (2005a) Standard method of test for resistance of
in geopolymer concretes giving a cement rich surface concrete to chloride ion penetration’’, T259. American
layer allowing higher initial absorption. The sorptivity Association of State Highway Transport Officials, Wash-
values of the geopolymer concretes are comparable to ington DC
730 Materials and Structures (2015) 48:721–731

2. AASHTO (2005b) Standard method of test for rapid and simple model for the promotion and understanding of
determination of the chloride permeability of concrete’’ green-chemistry. In: Davidovits J (ed) Proceedings of geo-
T277. American Association of State Highway Transport polymer, green chemistry and sustainable development
Officials, Washington DC solutions. Institut Géopolymère, Saint-Quentin, pp 9–15
3. Adam AA, Patnaikuni I, Law DW, Molyneaux TCK (2007) 21. Deja J (2002) Carbonation aspects of alkali activated slag
Strength of mortar containing activated slag and fly ash. In: mortars and concretes. Silic Ind 67(1):37–42
Proceedings of the 23rd biennial conference of the Concrete 22. DeSouza SJ (1996) Test methods for the evaluation of the
Institute of Australia. Adelaide, Australia, pp 353–358 durability of covercrete. M.A.Sc. Thesis, University of
4. AS (1998) Supplementary cementitious materials for use Toronto, Toronto, Canada
with portland and blended cement—fly ash. AS 3582.1- 23. Duxson P, Lukey G, van Deventer J (2007) Physical evo-
1998, Standards Australia lution of Na-geopolymer derived from metakaolin upto
5. AS (2000) Methods for sampling and testing aggregates— 1,000 °C. J Mater Sci 42(9):3044–3054
particle density and water absorption of fine aggregate. AS 24. Fernandez-Jimenez A, Palomo A, Criado M (2005) Micro-
1141.5-2000, Standards Australia structure development of alkali-activated fly ash cement: a
6. AS (1993) Methods of testing concrete—Sampling of fresh descriptive model. Cem Concr Res 35(6):1204–1209
concrete. AS 1012.1-1993, Standards Australia 25. Gjorv OE (1989) Alkali activation of a Norwegian granu-
7. ASTM (2004) Standard test method for measurement of rate lated blast furnace slag. In: Proc. third international con-
of absorption of water by hydraulic-cement concretes. ference on FA, silica fume, slag, and natural Pozzolans in
ASTM C1585-04. ASTM, West Conshohocken concrete, Trondheim, Norway, pp 1501–1518
8. ASTM (2004) Standard test method for electrical indication 26. Gruskovnjak A, Lothenbach B, Holzer L, Figi R, Winnefeld
of concrete’s ability to resist chloride ion penetration. F (2006) Hydration of alkali-activated slag: comparison
C1202-10, ASTM, West Conshohocken with ordinary Portland cement. Adv Cem Res 18(3):
9. Bakharev T, Sanjayan JG, Cheng Y-B (1999) Alkali acti- 119–128
vation of Australian slag cements. Cem Concr Res 29(1): 27. Guo X, Shi H, Dick WA (2010) Compressive strength and
113–120 microstructural characteristics of class C FA geopolymer.
10. Bakharev T (2005) Geopolymeric materials prepared using Cem Concr Comp 32(2):142–147
class F fly ash and elevated temperature curing. Cem Concr 28. Hall C (1989) Water sorptivity of mortars and concretes: a
Res 35(6):1224–1232 review. Mag Conc Res 41(147):51–61
11. Bakharev T (2005) Resistance of geopolymer materials to 29. Ho DWS, Lewis RK (1987) Carbonation of concrete and its
acid attack. Cem Concr Res 32(3):658–670 prediction. Cem Concr Res 17(3):489–504
12. Bakharev T (2005) Durability of geopolymer materials in 30. van Jaarsveld JGS, van Deventer JSJ (1999) Effect of alkali
sodium and magnesium sulphate solutions. Cem Concr Res metal activator on the properties of Fly-Ash based geo-
35(6):1233–1246 polymer. Ind Eng Chem Res 38:3932–3941
13. Bamforth P, Pocock D (2000) Design for durability of 31. Kayyali OA, Haque MN (1995) The Cl-/OH- ratio in
reinforced concrete exposed to chlorides. Workshop on chloride contaminated concrete—a most important crite-
structures with a service life of 100 years or more, 6th rion. Mag Concr Res 47(172):235–242
international conference on deterioration and repair of 32. Kong DLY, Sanjayan JG (2010) Effect of elevated tem-
reinforced concrete in the Arabian Gulf, Bahrain peratures on geopolymer paste, mortar and concrete. Cem
14. Barbosa VFF, MacKenzie KJD, Thaumaturgo C (2000) Concr Res 40(2):334–339
Synthesis and characterisation of materials based on inor- 33. Lawrence CD (2003) The production of low-energy
ganic polymers of alumina and silica: sodium polysialate cements. Lea’s chemistry of cement and concrete, 4th edn.
polymers. Int J Inorg Mater 2(4):309–317 Butterworth-Heinemann, Oxford, pp 421–470
15. Bernal SA, Provis JL, Brice DG, Kilcullen A, Duxon D, van 34. Li Z, Ding Z, Zhang Y (2004) Development of sustainable
Deventer JSJ (2012) Accelerated carbonation testing of cementitious materials. In: Proceedings of international
alkali-activated binders significantly underestimates service workshop on sustainable development and concrete tech-
life: the role of pore solution chemistry’. Cem Concr Res nology, Beijing, China, pp 55–76
42(10):1317–1326 35. Miranda JM, Fernandez-Jimenez A, Gonxales JA, Palomo
16. Broomfield JP (2007) Corrosion of steel in concrete, 2nd A (2005) Corrosion resistance in activated FA mortars. Cem
edn. Taylor and Francis, London and New York Concr Res 35(6):1210–1217
17. Cairns JJ, Law D W (2003) Prediction of the ultimate state 36. Lloyd RR, Provis JL, van Deventer JSJ (2010) Pore solution
of degradation of reinforced concrete structures. In: Pro- composition and alkali diffusion in inorganic polymer
ceedings of the integrated lifetime engineering of buildings cement. Cem Concr Res 40(9):1386–1392
and civil infrastructures, Kuopio, Finland 37. Neville AM (1996) Properties of concrete, 4th edn. Wiley,
18. Crank J (1975) The mathematics of diffusion, 2nd edn. New York
Clarendon Press, Oxford 38. Olivia M, Nikraz H, Sarker P (2008) Improvements in the
19. Davidovits J (1994) Properties of geopolymer cements. In: strength and water penetrability of low calcium FA based
Proceedings of the first international conference on alkaline geopolymer concrete. In: Proceedings of the third ACF
cements and concretes. Kiev State Technical University, international conference, Vietnam, 2008
Kiev, Ukraine, pp 131–149 39. Palomo A, Grutzeck MW, Blanco MT (1999) Alkali-acti-
20. Davidovits J (2005) Geopolymer chemistry and sustainable vated fly ashes: a cement for the future. Cem Concr Res
development. The poly(sialate) trminology: a very useful 29(8):1323–1329
Materials and Structures (2015) 48:721–731 731

40. Philleo RE (1989) Slag or other supplementary materials? sulphuric acid attack. In: Proceedings of the international
In: Proceedings of the third international conference on FA, conference on durability of building materials and compo-
silica fume, slag, and natural Pozzolans in concrete, nents, Lyon, France, pp 369–375
Trondheim, Norway, pp 1197–1208 47. Steveson M, Sagoe-Crentsil K (2005) Relationships
41. Provis JL, Yong CY, Duxson P, van Deventer JSJ (2009) between composition, structure and strength of inorganic
Correlating mechanical and thermal properties of sodium polymers. J Mater Sci 40(16):4247–4259
silicate-fly ash geopolymers. Colloid Surf A 336(1-3):57–63 48. Talling B, Brandstetr J (1989) Present state and future of
42. Puertas F, Fernández-Jiménez A, Blanco-Varela MT (2004) alkali-activated slag concretes. In: Proceedings of the third
Pore solution in alkali-activated slag cement pastes. Rela- international conference on fly ash, silica fume, slag, and
tion to the composition and structure of calcium silicate natural Pozzolans in concrete, Trondheim, Norway,
hydrate. Cem Concr Res 34:139–148 pp 1519–1545
43. Rangan VB (2008) Studies on fly ash A-based geopolymer 49. Temuujin J, van Riessen A, Williams R (2009) Influence of
concrete. Malays Constr Res J 3(2):1–20 calcium compounds on the mechanical properties of FA
44. Shi C, Krivenko PV, Roy DM (2006) Alkali activated geopolymer pastes. J Hazard Mater 167(1–3):82–88
cement concretes. Taylor and Francis, Abingdon 50. Winnefeld F, Leemann A, Lucuk M, Svoboda P, Neuroth M
45. Song S, Jennings HM (1999) Pore solution chemistry of (2010) Assessment of phase formation in alkali activated
alkali-activated ground granulated blast-furnace slag’. Cem low and high calcium fly ashes in building materials. Constr
Concr Res 29:159–170 Build Mater 24(6):1086–1093
46. Song X, Munn R, Marosszeky M, Brungs M (2005) Dura-
bility of fly ash based geopolymer concrete against

Das könnte Ihnen auch gefallen