Sei sulla pagina 1di 7

Setting Time, Strength, and Bond of High-Calcium

Fly Ash Geopolymer Concrete


Pattanapong Topark-Ngarm 1; Prinya Chindaprasirt 2; and Vanchai Sata 3

Abstract: In this paper, setting time, strength and bond of high-calcium fly ash geopolymer concrete were investigated. The high-calcium fly
Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

ash was from Mae Moh power plant in northern Thailand. Both sodium silicate solution and sodium hydroxide solution were used as alkali
activators in every mix. Sodium hydroxide solution with 10 M, 15 M, and 20 M concentrations, sodium silicate to sodium hydroxide ratios
of 1.0 and 2.0, alkaline liquid to fly ash ratio of 0.5 and two curing regimes viz., heat curing at 60  2°C for 24 h and room temperature curing
at 23  2°C were used. The results indicated that fresh geopolymer concrete had short setting time of 28–58 min due to the presence of
high calcium content of fly ash. In general, strengths and modulus of elasticity increased with the increase in NaOH concentration.
For compressive strength, the optimum Na2 O content was around 12% of fly ash. The high-strength geopolymer concrete with 28-day
compressive strength of 54.4 MPa was obtained for mix with 15M NaOH. The modulii of elasticity of geopolymer concrete were related
to the compressive strengths and comparable to those of portland cement concrete. The tensile splitting strength and bond strength were also
related to the compressive strength and the values were higher than those of portland cement concrete. In particular, the bond strengths were
significantly higher than those given by the current design code. DOI: 10.1061/(ASCE)MT.1943-5533.0001157. © 2014 American Society
of Civil Engineers.
Author keywords: High-calcium fly ash; Geopolymer concrete; Mechanical properties; Bond strength.

Introduction material rich in SiO2 and Al2 O3 , activated by alkaline solution such
as sodium hydroxide (NaOH) or potassium hydroxide (KOH) and
The global production of portland cement (PC) increases every year sodium silicate (Na2 SiO3 ) solution with the help of elevated tem-
and has shown no sign of slowing down (USGS 2013). Large perature (Davidovits 2008). Geopolymers, in general, exhibit good
amount of natural resources such as limestone, fossil fuels, electric- mechanical properties, high durability, and strength maintenance in
ity, and natural gas are required in PC production. High tempera- extreme conditions such as elevated temperatures, high corrosion,
tures are required in the production of PC, and this has resulted in a and high radiation (Binici 2013; Kong and Sanjayan 2008; Sata
larger amount of carbon dioxide (CO2 ) emission into the atmos- et al. 2012; Sufian Badar et al. 2014).
phere. In Thailand, approximately 33 million tons of portland Fly ash is commonly used to produce geopolymers due to its
cement were produced in 2012 (USGS 2013). availability around the world. Fly ash is classified as either class C
In an effort to reduce the growth of PC consumption, alternative or class F according to its chemical composition. It has been shown
pozzolanic materials such as fly ash, slag, palm oil fuel ash, bagasse that low-calcium fly ash-based geopolymer concretes (LCGC) have
ash, and rice husk ash are used as alternative cementitious materials similar mechanical properties to concretes produced with PC
to replace part of portland cement (Chindaprasirt et al. 2007c; (Hardjito and Rangan 2005). In Thailand, approximately three
Sata et al. 2007; Tangchirapat et al. 2009). These materials are by- million tons of fly ash is produced from lignite coal-fired Mae
products from industrial processes and thus require less energy to Moh power station in Lampang province. This fly ash contains
produce compared to PC production. These pozzolanic materials relatively high calcium oxide content, typically around 12–25% by
contain rich silica (SiO2 ) and alumina (Al2 O3 ) and can also be used weight. According to previous researches, geopolymer paste and
to produce geo-binder when mixed with alkaline solutions. Geopol- mortar produced from high-calcium fly ash exhibited good strength
ymers or “chains or networks of mineral molecules linked with co- and durability (Chindaprasirt et al. 2010; Chindaprasirt et al. 2007c;
valent bonds” can be produced from aluminosilicate, an inorganic Wongpa et al. 2010). Rattanasak and Chindaprasirt (2009) reported
that leaching of silica and alumina was enhanced at high concen-
1
Ph.D. Candidate, Dept. of Civil Engineering, Faculty of Engineering, tration of NaOH and thus a high strength geopolymer was normally
Khon Kaen Univ., Khon Kaen 40002, Thailand. E-mail: p.toparkngarm@ obtained with a high concentration NaOH solution.
gmail.com One of the problems with high-calcium fly ash is the fast setting
2
Professor, Sustainable Infrastructure Research and Development due to the presence of calcium and the formation of calcium silicate
Center, Dept. of Civil Engineering, Faculty of Engineering, Khon Kaen hydrate and/or calcium aluminate silicate hydrate (Chindaprasirt
Univ., Khon Kaen 40002, Thailand (corresponding author). E-mail: et al. 2012). The use of high NaOH concentration in the high-
prinya@kku.ac.th calcium fly ash geopolymer can prolong the setting by limiting the
3
Assisstant Professor, Sustainable Infrastructure Research and Develop- leaching of calcium and allows normal geopolymerization process
ment Center, Dept. of Civil Engineering, Faculty of Engineering, Khon
to control the setting of paste (Hanjitsuwan et al. 2014).
Kaen Univ., Khon Kaen 40002, Thailand. E-mail: vancsa@kku.ac.th
Note. This manuscript was submitted on March 6, 2014; approved on In the utilization of geopolymer concrete, the reinforcement is
July 1, 2014; published online on August 19, 2014. Discussion period open an important aspect (Esfahani and Rangan 1998). Since concrete is
until January 19, 2015; separate discussions must be submitted for indivi- weak in tension, bond strength between rebar and concrete is thus a
dual papers. This paper is part of the Journal of Materials in Civil Engi- critical strength to transfer externally applied load from concrete to
neering, © ASCE, ISSN 0899-1561/04014198(7)/$25.00. rebar. Sarker (2011) and Sofi et al. (2007a) indicated that bond

© ASCE 04014198-1 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198


strengths between low-calcium fly ash geopolymer concrete and Table 1. Mixing Proportion of Geopolymer Concrete by Weight (kg=m3 )
reinforcement were higher than those with the PC system. This Aggregate
is supported by the work of Songpiriyakij et al. (2011), which Fly NaOH Sodium Na2 O
Mixes Coarse Fine ash (Molar) silicate (moles=m3 )
reported that high-calcium geopolymer paste had higher bond
strength at 3 days than that of traditional epoxy agent. S=H 1-10M 1091 588 414 104 (10M) 104 682.9
Fresh and hardened properties of geopolymer concrete are af- S=H 1-15M 1091 588 414 104 (15M) 104 831.7
fected by several factors such as alkaline liquid to fly ash ratio, S=H 1-20M 1091 588 414 104 (20M) 104 965.5
NaOH concentration, ratio of alkaline solutions, curing conditions, S=H 2-10M 1091 588 414 69 (10M) 138 637.7
S=H 2-15M 1091 588 414 69 (15M) 138 736.4
and admixtures (Sathonsaowaphak et al. 2009). A number of re-
S=H 2-20M 1091 588 414 69 (20M) 138 825.2
ports (Hardjito and Rangan 2005; Rangan et al. 2006b) on the prop-
erties of geopolymer concrete and the effects the factors previously
mentioned were based on low-calcium fly ash. In this research,
fresh properties, and strengths including bond strength of high- and initial molar ratios of starting materials. Total moles of Na2 O
Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

calcium fly ash geopolymer concrete (HCGC) were investigated. were calculated from Na2 O in fly ash, sodium hydroxide, and
sodium silicate solutions.

Experimental Details Mixing Procedure


The mixing procedure was adopted from previous researches
Materials on high-calcium geopolymer mortar (Chindaprasirt et al. 2010;
Rattanasak and Chindaprasirt 2009). The geopolymer concretes
were mixed in a 23  2° C controlled room. The procedure started
Fly Ash
by mixing fly ash with NaOH solution for 5 min. Coarse and fine
High-calcium fly ash (FA) from Mae Moh power station in
aggregates were then added to the mixer and mixed for another
Lampang province, Thailand was used in this study. The fly ash
5 min. Sodium silicate solution was then added to the mixture
was greyish-brown color and consisted of 45.23% SiO2 , 19.95%
and mixing was continued for additional 5 min.
Al2 O3 , 13.15% Fe2 O3 , and 15.51% CaO, which was analyzed us-
ing X-ray fluorescence (XRF). The particles were mainly spherical Casting of Specimens
with median particle size of 57 microns as determined by Particle For compressive strength and splitting tensile strength tests, geo-
Size Analyzer Laser model Mastersizer S (Malvern Instruments polymer concrete was cast into 100 × 200 mm cylindrical steel
and 38% was retained on 45 μm sieve. The fly ash was a high- molds in accordance with ASTM C192 (ASTM 2007). For the
calcium fly ash according to the chemical composition requirement modulus of elasticity test, geopolymer concrete was cast into 150 ×
per ASTM C618 (ASTM 2008a). 300 mm cylindrical steel molds. For the pullout bond test, geopol-
ymer concrete was cast into 100 × 150 mm cylindrical steel molds
Aggregate
with rebar placed vertically at the center. A 10-mm-thick acrylic
Commercially available limestone with maximum size of 20 mm
circular plate with a hole cut at the center was placed inside
and specific gravity of 2.65 in saturated surface dry (SSD) condi-
the steel mold to help align the rebar. The specimen was cast
tions was used as coarse aggregate. For fine aggregate, river sand
upside-down to ensure smooth contact surface between concrete
with specific gravity of 2.58 and fineness modulus of 2.9 in SSD
and bearing plate.
conditions was used.

Alkaline Solutions Curing of Specimens


For alkali activated solutions, sodium hydroxide (NaOH) and Two types of curing conditions were used in this study. The first
sodium silicate (Na2 SiO3 ) with 15.32% Na2 O, 32.87% SiO2 , was oven curing at 60  2°C for 24 h after a 2 h delay time. The
and 51.81% H2 O were used. For the NaOH solution, NaOH pellets delay time was the time after the casting of concrete. The delay time
were mixed with distilled water and stirred until all the pellets allows concrete to be finished and reach initial setting before trans-
were completely dissolved. The solution was then left for 24 h portation to the curing oven or curing rack. The second curing con-
before use. dition was in a 23  2°C controlled room. All concrete cylinders
were removed from the molds after 24 h curing and wrapped with
Rebar a vinyl sheet and cured in a 23  2°C controlled room until testing.
The steel bar used in this experiment was DB16 SD 40 with cross-
sectional area of 201 mm2 and perimeter of 50.3 mm. The rebar
conformed to ASTM A615 grade 60 [A615/A615M-09b (ASTM Testing of Specimens
2009a)]. The average yield strength and ultimate strength were
520 MPa and 650 MPa, respectively. Slump Flow and Setting Time
A slump flow test was performed on a plastic sheet placed on flat
Mix Proportions, Mixing Procedure, Casting and wooden board according to ASTM C1611 (ASTM 2009b) within
Curing Details 5 min after mixing. The reported slump flow values were the aver-
age of four diagonal measurements. The setting time of geopolymer
Mix Design concrete was measured according to ASTM C807 (ASTM 2008b)
The combined aggregate was 73% of total weight of the mix. The using mortar collected from fresh concrete by passing fresh con-
ratio of coarse aggregate to fine aggregate was 65∶35 and the alkali crete through a #4 sieve to remove coarse aggregate. The setting
liquid to fly ash (L∶A) ratio by weight was fixed at 0.5 for all mixes. time was measured from the time the fly ash made a contact with
The concentrations of NaOH were 10 M, 15 M, and 20 M and two NaOH solution until the penetration of a 2 mm diameter needle was
sodium silicate to NaOH ratios (S∶H) by weight of 1.0 and 2.0 were less than 10 mm. The reported setting time results were the average
used. Table 1 shows the mix proportions of geopolymer concrete of two specimens.

© ASCE 04014198-2 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198


Compressive Strength and Splitting Tensile Strength from fly ash increased (Rattanasak and Chindaprasirt 2009), which
improved the geopolymerization and thus shortened the setting
The concrete cylinders for compressive strength test were capped at
time. For the mixes with low NaOH (S∶H ¼ 2), the setting times
both ends to ensure smooth surfaces. Concrete cylinders were
were slightly increased with the increase in NaOH concentrations.
tested for compressive strength according to ASTM C39 (ASTM
At low NaOH content, setting was controlled by the presence of
2001) at the ages of 7 and 28 days. The reported results were the
calcium, and the high NaOH concentrations hindered the dissolu-
average of three concrete samples. The splitting tensile strengths of
tion of calcium to the system and thus setting was slightly delayed
concrete at 7 and 28 days were tested according to ASTM C496
(Chindaprasirt et al. 2012).
(ASTM 2010).

Modulus of Elasticity Compressive Strength

The modulus of elasticity was determined at the age of 28 days The results of compressive strength of HCGC are shown in
Table 3. The compressive strengths of HCGC at 7 and 28 days were
Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

using 150 × 300 mm concrete cylinders in accordance with ASTM


C469 (ASTM 2002). The reported results were the average of three 19.9–42.9 MPa and 32.6–54.4 MPa, respectively. It was clear that
concrete samples. high curing temperature influenced the early strength of HCGCs.
For HCGCs cured at high temperature, the compressive strength
increased rapidly until 7 days and continued at slower rate to
Bond Strength 28 days. This characteristic of HCGC is similar to the strength gain
The bond strength between geopolymer concrete and reinforcement of high strength concrete reported by the American Concrete Insti-
was tested at the age of 7 days in accordance with ASTM C234 tute (ACI) in ACI 363R-92 (ACI 1997).
(ASTM 2000). The reported results were the average of three con- The compressive strengths increased with the increase in NaOH
crete samples. concentration. At high NaOH concentrations, the leaching of
alumina and silica was enhanced and this resulted in increased
geopolymerization, and thus the strength increased (Rattanasak
Results and Discussions and Chindaprasirt 2009). For the mix series with S∶H ratio of 1.0,
the high NaOH concentration of 20 M resulted in a reduction of
strength, which were in line with other researches (Chindaprasirt
Slump Flow and Setting Time et al. 2007b; Hardjito and Rangan 2005; Temuujin et al. 2009;
Fresh characteristics, slump flow, and setting time of concrete Wongpa et al. 2010). The reduction of strength at high NaOH con-
with different mix designs are shown in Table 2. Flows of fresh con- centration was due to the high concentration of hydroxide ions
crete were 470–620 mm. For the mixes with high NaOH content (OH─), which caused aluminosilicate gel precipitation at the early
(S∶H ¼ 1), the flow values decreased with the increase in NaOH stage of development (Somna et al. 2011).
concentrations. The flow values of mixes with 10, 15, and 20 M HCGC continued to gain strength after temperature curing. On
NaOH were 580, 560, and 470 mm, respectively. The increase in the other hand, for low-calcium fly ash cured at high temperature,
NaOH concentration increased the viscosity and leaching of silica the strength gain after curing is not large (Rangan et al. 2006a). The
and alumina from fly ash particles and thus reduced the workability presence of calcium also contributed to the strength gain after some
(Rattanasak et al. 2011). For the mixes with low NaOH content temperature curing (Chindaprasirt et al. 2012; Dombrowski et al.
(S∶H ¼ 2), the flow values (500–530 mm) remained relatively un- 2007; García-Lodeiro et al. 2010; Guo et al. 2010; Temuujin et al.
affected with the change in NaOH concentration. In this system, the 2009). For high temperature curing at 60°C, the hydration was ac-
amount of NaOH was low and thus the effect of NaOH on flow was celerated and the moisture was consumed. The strength gain at later
less than that of the former system. age was therefore not large. For curing at 25°C, HCGC continued to
The setting times of concrete were 28–68 min and were also gain strength from 7 to 28 days due to continued hydration and the
influenced by S∶H ratio. The setting times of geopolymer mortar availability of H2 O.
in this research were relatively short compared to those of geopol- Fig. 1 shows the plot of 7 and 28 days compressive strengths
ymer concrete produced from low-calcium fly ash (Hardjito and against mass ratio of Na2 O to fly ash in the matrix. The optimum
Rangan 2005). The short setting time of HCGC was associated with compressive strengths were obtained with total mass proportion of
the high calcium content of the system (Rattanasak et al. 2011). For Na2 O to fly ash of approximately 12% for both curing regimes. The
mixes with high NaOH content (S∶H ¼ 1), the setting times de- result in this study was in line with that reported by Guo et al.
creased with the increase in the NaOH concentrations. The setting (2010) on geopolymer produced with class C fly ash. The 28 days
times of mixes with 10, 15, and 20 M NaOH were 50, 36, and compressive strength of HCGC could be expressed as function of
28 min, respectively. At high NaOH concentrations, setting was total moles of Na2 O as shown in Eqs. (1) and (2) with R2 equal to
controlled by normal geopolymerization (Chindaprasirt et al. 2012). 0.82 and 0.84.
At high NaOH concentrations, the dissolution of Si4þ and Al3þ ions For 25°C curing

0
fc28 ¼ −1.13ðMNa2 O Þ2 þ 28.1ðMNa2 O Þ − 130 ð1Þ
Table 2. Flow and Setting Time of Geopolymer Concrete
Mixes Flow (mm) Setting time (min) For 60°C for 24 h curing
S∶H 1-10M 580 50
S∶H 1-15M 560 36 0
S∶H 1-20M 470 28
fc28 ¼ −1.27ðMNa2 O Þ2 þ 31.2ðMNa2 O Þ − 140 ð2Þ
S∶H 2-10M 500 52
S∶H 2-15M 550 68 0
where f c28 = the 28-day compressive strength; and M Na2 O = mass
S∶H 2-20M 530 58
proportion of Na2 O to fly ash (%) of starting material.

© ASCE 04014198-3 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198


Table 3. Properties of Geopolymer Concrete
Compressive strength Tensile strength
γ 7d 28d 7d 28d Ec uc
3
Mixes Curing (kg=m ) MPa (Std.) MPa (Std.) MPa (Std.) MPa (Std.) MPa (Std.) MPa (Std.)
S∶H 1-10M Room 2,409 19.89 (0.81) 39.67 (1.12) 2.11 (0.23) 3.87 (0.29) 30,400 (1200) 9.04 (0.31)
Oven 2,432 40.73 (3.01) 46.67 (5.56) 2.80 (0.09) 4.15 (0.05) 31,000 (2000) 13.41 (0.66)
S∶H 1-15M Room 2,417 22.30 (1.74) 45.34 (2.13) 2.77 (0.18) 4.36 (0.24) 34,800 (600) 8.73 (0.36)
Oven 2,447 42.91 (4.94) 54.40 (2.79) 3.59 (0.25) 4.91 (0.48) 37,800 (1800) 13.57 (0.53)
S∶H 1-20M Room 2,439 20.50 (0.89) 37.64 (2.57) 2.48 (0.22) 3.96 (0.46) 38,400 (2800) 10.05 (0.59)
Oven 2,450 41.20 (10.15) 43.42 (3.77) 3.73 (0.39) 4.43 (0.46) 38,000 (4200) 14.59 (0.86)
S∶H 2-10M Room 2,384 21.27 (2.29) 33.80 (2.10) 2.55 (0.19) 3.07 (0.43) 23,400 (2600) 7.85 (0.59)
Oven 2,393 37.69 (3.70) 40.09 (3.15) 3.52 (0.42) 3.62 (0.18) 24,200 (1200) 13.83 (0.20)
Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

S∶H 2-15M Room 2,403 22.99 (4.24) 39.02 (0.29) 3.02 (0.19) 3.47 (0.12) 26,800 (800) 10.39 (0.06)
Oven 2,412 38.85 (4.78) 48.18 (3.58) 3.59 (0.50) 4.03 (0.73) 31,000 (2200) 14.47 (0.44)
S∶H 2-20M Room 2,421 23.06 (0.63) 46.69 (1.17) 3.15 (0.04) 3.84 (0.24) 35,400 (2600) 8.80 (0.58)
Oven 2,409 39.27 (4.74) 49.50 (1.43) 3.92 (0.08) 3.99 (0.27) 31,800 (1200) 14.13 (1.16)
Note: Std. = standard deviation.

Fig. 1. Compressive strength of geopolymer concrete versus total Fig. 2. Splitting tensile strength of geopolymer concrete versus com-
content of Na2 O (%) pressive strength

Splitting Tensile Strength 363R-92 (ACI 1997) suggest Eqs. (5) and (6) for the calculation
of splitting tensile strengths of normal and high-strength concretes,
The results of splitting tensile strength are shown in Table 3. The
respectively
splitting tensile strength of HCGC at 7 and 28 days were 2.11–3.92
and 3.07–4.91 MPa. In general, the average splitting tensile fct ¼ 0.17ðf c0 Þ0.75 ð4Þ
strengths of concrete were affected by the concrete’s mix propor-
tions, a finding similar to that for compressive strength. It appears
that the splitting tensile strength at 7 days for the series with S∶H fct ¼ 0.56ðf c0 Þ0.5 ð5Þ
ratio of 1 was generally lower than that of the series with S∶H ratio
of 2. The strength continued to gain at the higher rate up to 28 days. fct ¼ 0.59ðf c0 Þ0.5 ð6Þ
This was true for both curing regimes.
Fig. 2 shows the plot of splitting tensile strength against the where f ct = splitting tensile strength (MPa); and f c0 = compressive
square root of compressive strength. The splitting tensile strength strength (MPa).
of concrete increased with increasing compressive strength. The These equations are plotted along with experimental data as
mean splitting tensile strengths in this experiment were between shown in Fig. 2. It can be seen that splitting tensile strengths of
6.9 and 13.7% of its mean compressive strength, which was similar HCGC in this study were generally higher than those reported by
to those of PC of 8–14% (Sagoe-Crenstil et al. 2001). A number of Sofi et al. (2007b) and those obtained from Eq. (5) for low-calcium
researchers described the relationship between the splitting tensile fly ash geopolymer concrete. This was due to strong interfacial
strength and compressive strength of concrete using Eq. (3) transition zone (ITZ) between geopolymer paste and aggregate re-
ported by several authors (Chen et al. 2013; Lee and van Deventer
fct ¼ kðfc0 Þn ð3Þ 2004; Ryu et al. 2013; Temuujin et al. 2010). Lee and van Deventer
(2004) reported that geopolymer produced with high hydroxide
where k and n are constants obtained from a regression analysis content and higher silicate solution tended to have higher ITZ
of experimental data. Ryu et al. (2013) suggested Eq. (4) for the strength due to lower porosity in geopolymer paste. In the geopol-
calculation of splitting tensile strength of low-calcium fly ash- ymer system with the presence of high calcium compound, it was
based geopolymer concrete. ACI 318-08 (ACI 2008) and ACI suggested that voids in aluminosilicate matrix were filled with

© ASCE 04014198-4 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198


calcium silicate hydrate (CSH) products, resulting in a low porosity
geopolymer paste with increased strength (Hanjitsuwan et al.
2014). The splitting tensile strength results in this study were gen-
erally higher than the strengths calculated using standard design
codes. However, it was reported that the equation given in ACI
318-08 (ACI 2008) generally overestimated the splitting tensile
strength of high compressive strength concrete (Arioglu et al.
2006). The splitting tensile strengths of HCGC in this study were
related to the compressive strengths as shown in Eq. (7) with R2
equal to 0.77

f ct ¼ 0.45ðfc0 Þ0.57 ð7Þ


Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

Modulus of Elasticity Fig. 4. Modulus of elasticity of geopolymer concrete versus square


The results for the modulus of elasticity are shown in Table 3. The root of compressive strength
values were 23,400–41,000 MPa. The concrete with S∶H ratio of 1
showed higher modulus of elasticity values than those with S∶H
ratio of 2. Fig. 3 shows the plot of modulus of elasticity against follows and in line with other reports on low-calcium geopolymer
Na2 O content. It can be concluded that modulus of elasticity in- concretes
creased with increasing amounts of Na2 O. Similar to compressive
Ec ¼ 0.043ðwc Þ1.5 ðfc0 Þ0.5 ð9Þ
strength development, with S∶H ratio of 1, high content of the so-
dium hydroxide solution enhanced the leaching of silica and alu-
mina ions, hence a higher degree of geopolymerization was Ec ¼ 3,320ðf c0 Þ0.5 þ 6,900ðwc =2,320Þ1.5 ð10Þ
obtained. The relationship between modulus of elasticity and
Na2 O is shown in Eq. (8) with R2 ¼ 0.79 Bond Strength
Ec ¼ 2,700ðMNa2 O Þ þ 400 ð8Þ The results for bond strength between HCGC and rebar are shown
in Table 3. The bond strengths were 8.73–14.59 MPa. All speci-
The results of modulus of elasticity were plotted against the mens in this experiment failed by concrete rupture and thus the
square root of compressive strength as shown in Fig. 4. It can ultimate bond strength was reached. In general, the bond strength
be seen that the modulus of elasticity of concrete generally in- of the HCGC was related to the compressive strength of the con-
creased as compressive strength increased, which was similar to crete. Many equations have been proposed to estimate bond
the current design code, ACI 318-08 (ACI 2008) and alternative strength between deformed bar and concrete as shown below.
equation for high strength concrete ACI 363R-92 (ACI 1997) as Orangun et al. (1977) proposed Eq. (11) for normal concrete
shown in Eqs. (9) and (10), respectively. In fact, the results of this without transverse reinforcement, which is the basis for current
investigation were between those given by the two codes. It is development length calculation in ACI 318 [ACI 408R-03
widely known that the modulus of elasticity of concrete depends (ACI 2003)] as follows:
on mix proportion, properties of the aggregates, and the curing con-
ditions of the concrete. Several researchers (Hardjito and Rangan uc ¼ ð0.101 þ 0.268cmin =db þ 4.4db =ld Þðf c0 Þ0.5 ð11Þ
2005; Sofi et al. 2007b; Yost et al. 2013) reported on geopolymer
concrete with a lower modulus of elasticity when compared to those where uc = ultimate bond strength (MPa); cmin = minimum con-
calculations using a normal concrete equation, with the exception crete cover (mm); db = nominal bar diameter (mm); ld = develop-
of Joseph and Mathew (2012) who reported a higher modulus of ment length (mm); and f c0 = compressive strength (MPa). Esfahani
elasticity. It can be seen that the results in this study were close to and Rangan (1998) suggested that bond strength between reinforc-
the values calculated with the ACI 363R-92 (ACI 1997) equation as ing steel and concrete was related to splitting tensile strength of
concrete and proposed Eq. (12) as follows:
uc ¼ 4.9½ðcmin =db þ 0.5Þ=ðcmin =db þ 3.6Þðf ct Þ ð12Þ

where f ct = tensile strength of concrete taken as 0.55ðfc0 Þ0.5 (MPa);


and fc0 = compressive strength (MPa) less than 50 MPa. Hadi
(2008) proposed Eq. (13) for bond strength between high strength
steel and high strength concrete based on the pullout test as follows:
uc ¼ 0.083045ð22.8 þ 0.208cmin =db − 38.212db =ld Þðfc0 Þ0.5
ð13Þ

The bond strengths in this study can be written as shown in


Eq. (14) with R2 ¼ 0.85 to obtain
uc ¼ 2.12ðfc0 Þ0.5 ð14Þ

Fig. 5 shows a plot of bond strength of geopolymer concrete


Fig. 3. Modulus of elasticity of geopolymer concrete total content of
against the square root of compressive strength along with the re-
Na2 O (%)
sults of LCGC (Chang et al. 2009; Sofi et al. 2007a). The bond

© ASCE 04014198-5 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198


approximately 12% by weight of fly ash. Optimum compressive
strength of 54.4 MPa was obtained for the mix with 15 M NaOH.
The splitting tensile strength and modulus of elasticity of HCGC
were related to the compressive strength. The splitting tensile
strengths of HCGC were relatively high in terms of percentage
of compressive strength, which was probably due to the strong in-
terfacial zone of the aggregate and the high calcium geopolymer
matrix. The bond strengths between HCGC and rebar were signifi-
cantly higher than those of normal portland cement concrete given
by the current ACI 318 design code. There was also an indication
that the bond strength of HCGC was slightly higher than that of the
low-calcium fly ash geopolymer concrete.
Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

Acknowledgments
Fig. 5. Ultimate bond strength between rebar and geopolymer concrete
versus square root of compressive strength This work was financially supported by the Thailand Research
Fund (TRF) under the TRF Senior Research Scholar, Grant No.
RTA5480004. It was also supported by the Higher Education
Research Promotion and National Research University Project
strengths of the HCGC and reinforcing bar were significantly of Thailand, Office of the Higher Education Commission, through
higher than those found in Orangun et al. (1977), Esfahani and the Advanced Functional Materials Cluster of Khon Kaen Univer-
Rangan (1998), and Hadi (2008). In fact, Eq. (14) from this inves- sity AFM-2554-Ph.D. 54402.
tigation gave the bond strengths nearly double the values of those
obtained using Eq. (11) from Orangun et al. (1977). The high bond
strength was due to the relatively high splitting tensile strength as
reported by several authors on bond strength between deformed bar References
and LCGC (Sarker 2011; Sofi et al. 2007a). The high splitting ten- American Concrete Institute (ACI). (1997). “State-of-the-art report on high-
sile strength was related to ITZ strength and was also influenced by strength concrete.” 363R-92, Farmington Hills, MI.
the amount Na2 O and Na2 SiO3 . The NaOH concentration played American Concrete Institute (ACI). (2003). “Bond and development of
an important role in bond strength. Songpiriyakij et al. (2011) re- straight reinforcing bars in tension.” 408R-03, Farmington Hills, MI, 56.
ported the bond strength of 4.41 and 6.38 MPa of fly ash geopol- American Concrete Institute (ACI). (2008). “Building code requirements
ymer paste activated with NaOH concentrations of 10 and 18 molar. for structural concrete (ACI 318-08) and commentary.” 318-08,
This was due to the leaching ability of silica and alumina in high Farmington Hills, MI.
alkaline solution, as stated in previous sections. Arioglu, N., Girgin, Z. C., and Arioglu, E. (2006). “Evaluation of ratio
For the bond strength of low-calcium fly ash concrete, the re- between splitting tensile strength and compressive strength for
concretes up to 120 MPa and its application in strength criteria.”
sults of Sofi et al. (2007a) were from pullout test and results of
ACI Mater. J., 103(1), 18–24.
Chang et al. (2009) were from a splice-beam test. In general, ASTM. (2000). “Standard specification for comparing concrete on
splice-beam tests give more realistic bond behavior between rebar basis of the bond developed with reinforcing steel.” C234, West
and concrete while the pullout test is simple to set up and can be Conshohocken, PA.
used for comparison purposes [ACI 408R-03 (ACI 2003)]. The ASTM. (2001). “Standard test method for compressive strength of
bond strengths from the pullout test reported by Sofi et al. cylindrical concrete specimens.” C39, West Conshohocken, PA.
(2007a) were also high but slightly lower than the bond strengths ASTM. (2002). “Standard test method for static modulus of elasticity
obtained in this study. The superior bond strength characteristics of and poisson’s ratio of concrete in compression.” C469, West
the HCGC was due to the good ITZ characteristics of geopolymer Conshohocken, PA.
matrix and may be due in part to the presence of calcium ASTM. (2007). “Standard test method for making and curing concrete test
specimens in the laboratory.” C192, West Conshohocken, PA.
and additional calcium silicate hydrate. This aspect needs further
ASTM. (2008a). “Standard specification for coal fly ash and raw or
investigation. calcined natural pozzolan for use in cement.” C618, West
Conshohocken, PA.
ASTM. (2008b). “Standard test methods for time of setting of hydraulic
Conclusion cement mortar by modified vicat needle.” C807, West
Conshohocken, PA.
Based on the obtained data in this experiment, the following ASTM. (2009a). “Standard specification for deformed and plain
conclusions can be made. The results in this study indicated that carbon-steel bars for concrete reinforcement.” A615/A615M-09b, West
high-calcium fly ash was suitable to use in producing high strength Conshohocken, PA.
geopolymer concrete with high bond strength between concrete and ASTM. (2009b). “Standard test methods for slump flow of self-
rebar. The fresh concrete had short setting time of 28–58 min due to consolidating concrete.” C1611, West Conshohocken, PA.
the presence of high calcium content in the system. For compres- ASTM. (2010). “Standard test method for splitting tensile strength of
sive strength, curing at 60°C for 24 h accelerated the strength de- cylindrical concrete specimens.” C496, West Conshohocken, PA.
Binici, H. (2013). “Engineering properties of geopolymer incorporating
velopment as indicated by the high 7-day compressive strengths.
slag, fly ash, silica sand and pumice.” Adv. Civ. Environ. Eng.,
The strength of HCGC continued to develop with time similar 1(3), 108–123.
to portland cement concrete, which was due to the presence of Chang, E. H., Sarker, P., Lloyd, N., and Rangan, B. V. (2009). “Bond
calcium and the formation of additional calcium silicate hydrate. behaviour of reinforced fly ash-based geopolymer concrete beams.”
The compressive strength of HCGC was related to the NaOH con- 24th Biennial Conf. of the Concrete Institute of Australia, R. I. Gilbert,
centration and the Na2 O content. The optimum Na2 O content was ed., Concrete Institute of Australia, Luna Park, Sydney, 10.

© ASCE 04014198-6 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198


Chen, X., Wu, S., and Zhou, J. (2013). “Influence of porosity on compres- Rattanasak, U., and Chindaprasirt, P. (2009). “Influence of NaOH solution
sive and tensile strength of cement mortar.” Constr. Build. Mater., 40, on the synthesis of fly ash geopolymer.” Miner. Eng., 22(12),
869–874. 1073–1078.
Chindaprasirt, P., Chareerat, T., Hatanaka, S., and Cao, T. (2010). “High- Rattanasak, U., Pankhet, K., and Chindaprasirt, P. (2011). “Effect of chemi-
strength geopolymer using fine high-calcium fly ash.” J. Mater. Civ. cal admixtures on properties of high-calcium fly ash geopolymer.” Int.
Eng., 10.1061/(ASCE)MT.1943-5533.0000161, 264–270. J. Miner. Metall. Mater., 18(3), 364–369.
Chindaprasirt, P., Chareerat, T., and Sirivivatnanon, V. (2007a). “Workabil- Ryu, G. S., Lee, Y. B., Koh, K. T., and Chung, Y. S. (2013). “The mechani-
ity and strength of coarse high calcium fly ash geopolymer.” Cem. cal properties of fly ash-based geopolymer concrete with alkaline
Concr. Compos., 29(3), 224–229. activators.” Constr. Build. Mater., 47, 409–418.
Chindaprasirt, P., Chotithanorm, C., Cao, H. T., and Sirivivatnanon, V. Sagoe-Crenstil, K. K., Brown, T., and Taylor, A. H. (2001). “Performance
(2007b). “Influence of fly ash fineness on the chloride penetration of of concrete made with commercially produced coarse recycled concrete
concrete.” Constr. Build. Mater., 21(2), 356–361. aggregate.” Cem. Concr. Res., 31(5), 707–712.
Chindaprasirt, P., Homwuttiwong, S., and Jaturapitakkul, C. (2007c). Sarker, P. K. (2011). “Bond strength of reinforcing steel embedded in fly
“Strength and water permeability of concrete containing palm oil fuel ash-based geopolymer concrete.” Mater. Struct./Mater. Constr., 44(5),
Downloaded from ascelibrary.org by Technische Universiteit Delft on 11/22/19. Copyright ASCE. For personal use only; all rights reserved.

ash and rice husk-bark ash.” Constr. Build. Mater., 21(7), 1492–1499. 1021–1030.
Chindaprasirt, P., De Silva, P., Sagoe-Crentsil, K., and Hanjitsuwan, S. Sata, V., Jaturapitakkul, C., and Kiattikomol, K. (2007). “Influence of poz-
(2012). “Effect of SiO2 and Al2 O3 on the setting and hardening of high zolan from various by-product materials on mechanical properties of
calcium fly ash-based geopolymer systems.” J. Mater. Sci., 47(12), high-strength concrete.” Constr. Build. Mater., 21(7), 1589–1598.
4876–4883. Sata, V., Sathonsaowaphak, A., and Chindaprasirt, P. (2012). “Resistance
Davidovits, J. (2008). Geopolymer chemistry and application, 2nd Ed., of lignite bottom ash geopolymer mortar to sulfate and sulfuric acid
Institute Geopolymer, Saint-Quentin, France. attack.” Cem. Concr. Compos., 34(5), 700–708.
Dombrowski, K., Buchwald, A., and Weil, M. (2007). “The influence of Sathonsaowaphak, A., Chindaprasirt, P., and Pimraksa, K. (2009).
calcium content on the structure and thermal performance of fly ash “Workability and strength of lignite bottom ash geopolymer mortar.”
based geopolymers.” J. Mater. Sci., 42(9), 3033–3043. J. Hazard. Mater., 168(1), 44–50.
Esfahani, M. R., and Rangan, B. V. (1998). “Bond between normal strength Sofi, M., Deventer, J. S. J., Mendis, P. A., and Lukey, G. C. (2007a). “Bond
and high-strength concrete (HSC) and reinforcing bars in splices in performance of reinforcing bars in inorganic polymer concrete (IPC).”
beams.” ACI Struct. J., 9. J. Mater. Sci., 42(9), 3107–3116.
García-Lodeiro, I., Fernández-Jiménez, A., Palomo, A., and MacPhee, Sofi, M., van Deventer, J. S. J., Mendis, P. A., and Lukey, G. C. (2007b).
D. E. (2010). “Effect of calcium additions on N-A-S-H cementitious “Engineering properties of inorganic polymer concretes (IPCs).” Cem.
gels.” J. Am. Ceram. Soc., 93(7), 1934–1940. Concr. Res., 37(2), 251–257.
Guo, X., Shi, H., and Dick, W. A. (2010). “Compressive strength and Somna, K., Jaturapitakkul, C., Kajitvichyanukul, P., and Chindaprasirt, P.
microstructural characteristics of class C fly ash geopolymer.” Cem. (2011). “NaOH-activated ground fly ash geopolymer cured at ambient
Concr. Compos., 32(2), 142–147. temperature.” Fuel, 90(6), 2118–2124.
Hadi, M. N. S. (2008). “Bond of high strength concrete with high strength Songpiriyakij, S., Pulngern, T., Pungpremtrakul, P., and Jaturapitakkul, C.
reinforcing steel.” Open Civ. Eng. J., 2, 5. (2011). “Anchorage of steel bars in concrete by geopolymer paste.”
Hanjitsuwan, S., Hunpratub, S., Thongbai, P., Maensiri, S., Sata, V., and Mater. Des., 32(5), 3021–3028.
Chindaprasirt, P. (2014). “Effects of NaOH concentrations on physical Sufian Badar, M., Kupwade-Patil, K., Bernal, S. A., Provis, J. L., and
and electrical properties of high calcium fly ash geopolymer paste.” Allouche, E. N. (2014). “Corrosion of steel bars induced by accelerated
Cem. Concr. Compos., 45, 9–14. carbonation in low and high calcium fly ash geopolymer concretes.”
Hardjito, D., and Rangan, B. V. (2005). Development and properties Constr. Build. Mater., 61, 79–89.
of low-calcium fly ash-based geopolymer concrete, Curtin Univ. of Tangchirapat, W., Jaturapitakkul, C., and Chindaprasirt, P. (2009). “Use of
Technology, Perth, Australia. palm oil fuel ash as a supplementary cementitious material for produc-
Joseph, B., and Mathew, G. (2012). “Influence of aggregate content on ing high-strength concrete.” Constr. Build. Mater., 23(7), 2641–2646.
the behavior of fly ash based geopolymer concrete.” Sci. Iran., 19(5), Temuujin, J., van Riessen, A., and MacKenzie, K. J. D. (2010). “Prepara-
1188–1194. tion and characterisation of fly ash based geopolymer mortars.” Constr.
Kong, D. L. Y., and Sanjayan, J. G. (2008). “Damage behavior of geopol- Build. Mater., 24(10), 1906–1910.
ymer composites exposed to elevated temperatures.” Cem. Concr. Temuujin, J., van Riessen, A., and Williams, R. (2009). “Influence of cal-
Compos., 30(10), 986–991. cium compounds on the mechanical properties of fly ash geopolymer
Lee, W. K. W., and van Deventer, J. S. J. (2004). “The interface between pastes.” J. Hazard. Mater., 167(1–3), 82–88.
natural siliceous aggregates and geopolymers.” Cem. Concr. Res., U.S. Geological Survey (USGS). (2013). “U.S. geological survey, mineral
34(2), 195–206. commodity summaries.” Reston, VA, 38.
Orangun, C. O., Jirsa, J. O., and Breen, J. E. (1977). “A reevaluation of test Wongpa, J., Kiattikomol, K., Jaturapitakkul, C., and Chindaprasirt, P.
data on development length and splices.” ACI J., 74(3), 114–122. (2010). “Compressive strength, modulus of elasticity, and water per-
Rangan, B. V., Hardjito, D., Wallah, S. E., and Sumajouw, D. M. J. (2006a). meability of inorganic polymer concrete.” Mater. Des., 31(10),
“Properties and applications of fly ash-based concrete.” Mater. Forum, 4748–4754.
30, 170–175. Yost, J. R., RadliÑska, A., Ernst, S., and Salera, M. (2013). “Structural
Rangan, B. V., Wallah, S., Sumajouw, D., and Hardjito, D. (2006b). “Heat- behavior of alkali activated fly ash concrete. Part 1: Mixture design,
cured, low-calcium, fly ash-based geopolymer concrete.” Indian Concr. material properties and sample fabrication.” Mater. Struct., 46(3),
J., 80(6), 47–52. 435–447.

© ASCE 04014198-7 J. Mater. Civ. Eng.

J. Mater. Civ. Eng., 2015, 27(7): 04014198

Potrebbero piacerti anche