8
Indian Journal of Engineering & Materials Sciences Vol. 24, December 2017, pp. 499-506 Effect of water-to-powder ratios on the compressive strength and microstructure of metakaolin based geopolymers Abideng Hawa a *, Woraphot Prachasaree b & Danupon Tonnayopas c a Department of Civil Engineering, Faculty of Engineering, Princess of Naradhiwas University, AmphurMuang, Narathiwat 96000, Thailand b Department of Civil Engineering, Faculty of Engineering, Prince of Songkla University, Hat Yai, Songkhla 90112, Thailand c Department of Mining and Materials Engineering, Faculty of Engineering, Prince of Songkla University, Hat Yai, Songkhla 90112, Thailand Received 28 July 2016; accepted 22 February 2017 This study investigated the properties of geopolymer mortars prepared from metakaolin (MK) partially replaced with oil palm ash (OPA). The geopolymer blend had an alkaline activator (sodium silicate, sodium hydroxide, and water). The main parameters studied were compressive strength, microstructure, and sulfuric acid resistance (losses of strength and mass). Geopolymers were made varying the heat curing and the water to powder ratio (w/p set at 0.45, 0.55, and 0.65). The geopolymer samples were cured at 80°C for 0.5, 1, 2, or 4 h and kept at ambient temperature until testing. The partial substitution of MK by OPA was set at 5, 10, and 15 wt% in the mixtures. The compressive strength was measured after 2, 6, and 24 h and 7 and 28 days. The matrix structure effects were examined using scanning electron microscopy (SEM), X-ray diffraction (XRD), and Fourier transform infrared spectroscopy (FT-IR). The results revealed that the geopolymer mortar with 5% OPA and water to powder ratio 0.45 was superior to the other samples. SEM imaging indicated that it had a dense-compact matrix structure, which contributed to the highest compressive strength. Key words: Geopolymer, Compressive strength, Metakaolin, Oil palm ash Geopolymers or alkali activated materials have the advantage that waste materials can be used in them, either as the main component or as a partial replacement. These products have high early strength with good mechanical properties and durability 1,2 . Initially the raw materials for geopolymer synthesis were fly ash (FA), metakaolin (MK), and volcanic ash (VA) 3 , with later use of waste pozzolans such as oil palm ash (OPA) 4 , rice husk ash (RHA) 5,6 and bagasse ash (BA) 7 by partial replacement. When the main raw materials have high calcium content, the resulting material is usually not considered a geopolymer. In this current study, MK with very low calcium content was used. Geopolymers are resistant to acid solutions and have less drying shrinkage than OPC materials. Under high temperature curing of geopolymer, the binder paste contributes to the low drying shrinkage. Some researches 8,9 indicate that fly ash based geopolymer mortars with alkali activator undergo very low drying shrinkage, compared with OPC concrete. Slaty et al. 10 showed that kaolinite based geopolymers with sand filler present excellent stability with very low shrinkage, and no additional shrinkage is observed up to 180 days. The main product of geopolymerization reactions is alkali aluminosilicate, which contributes to the good mechanical properties and resistance to chemical attacks in geopolymers. However, the physical and mechanical properties of component phases do not sufficiently explain the behavior of geopolymers, instead microstructural analysis of geopolymers is necessary. The information about the structure of geopolymers obtainable by X-ray diffraction (XRD) is rather limited. Consequently, scanning electron microscopy (SEM) and Fourier transform infrared (FT-IR) were considered for additional information in this study. Some studies 11-13 have investigated the geopolymer structure by SEM and FT-IR. Generally, excessive water content in cement mortar or concrete blend reduces its final strength. However, in the case of geopolymers the effects of water-cement ratio (or water-powder ratio in the current study) are poorly known. Both the mechanical properties and the structure of MK based geopolymers may be affected by the w/p ratio. Currently, OPA is a waste by-product of the palm oil industry, remaining a problem for the agro industries in Thailand and other South East Asia. It has been estimated that the annual industrial solid OPA ————— *Corresponding author (E-mail: [email protected])

Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

Embed Size (px)

Citation preview

Page 1: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

Indian Journal of Engineering & Materials Sciences

Vol. 24, December 2017, pp. 499-506

Effect of water-to-powder ratios on the compressive strength and microstructure

of metakaolin based geopolymers

Abideng Hawaa*, Woraphot Prachasareeb & Danupon Tonnayopasc

aDepartment of Civil Engineering, Faculty of Engineering, Princess of Naradhiwas University, AmphurMuang, Narathiwat 96000, Thailand

bDepartment of Civil Engineering, Faculty of Engineering, Prince of Songkla University, Hat Yai, Songkhla 90112, Thailand

cDepartment of Mining and Materials Engineering, Faculty of Engineering, Prince of Songkla University, Hat Yai, Songkhla 90112, Thailand

Received 28 July 2016; accepted 22 February 2017

This study investigated the properties of geopolymer mortars prepared from metakaolin (MK) partially replaced with oil

palm ash (OPA). The geopolymer blend had an alkaline activator (sodium silicate, sodium hydroxide, and water). The main

parameters studied were compressive strength, microstructure, and sulfuric acid resistance (losses of strength and mass).

Geopolymers were made varying the heat curing and the water to powder ratio (w/p set at 0.45, 0.55, and 0.65).

The geopolymer samples were cured at 80°C for 0.5, 1, 2, or 4 h and kept at ambient temperature until testing. The partial

substitution of MK by OPA was set at 5, 10, and 15 wt% in the mixtures. The compressive strength was measured after 2, 6,

and 24 h and 7 and 28 days. The matrix structure effects were examined using scanning electron microscopy (SEM), X-ray

diffraction (XRD), and Fourier transform infrared spectroscopy (FT-IR). The results revealed that the geopolymer mortar with

5% OPA and water to powder ratio 0.45 was superior to the other samples. SEM imaging indicated that it had a dense-compact

matrix structure, which contributed to the highest compressive strength.

Key words: Geopolymer, Compressive strength, Metakaolin, Oil palm ash

Geopolymers or alkali activated materials have the

advantage that waste materials can be used in them,

either as the main component or as a partial replacement.

These products have high early strength with good

mechanical properties and durability1,2. Initially the raw

materials for geopolymer synthesis were fly ash (FA),

metakaolin (MK), and volcanic ash (VA)3, with later use

of waste pozzolans such as oil palm ash (OPA)4, rice

husk ash (RHA)5,6 and bagasse ash (BA)7 by partial

replacement. When the main raw materials have high

calcium content, the resulting material is usually not

considered a geopolymer. In this current study, MK with

very low calcium content was used.

Geopolymers are resistant to acid solutions and have

less drying shrinkage than OPC materials. Under high

temperature curing of geopolymer, the binder paste

contributes to the low drying shrinkage. Some

researches8,9 indicate that fly ash based geopolymer

mortars with alkali activator undergo very low drying

shrinkage, compared with OPC concrete. Slaty et al.10

showed that kaolinite based geopolymers with sand

filler present excellent stability with very low shrinkage,

and no additional shrinkage is observed up to 180 days.

The main product of geopolymerization reactions is

alkali aluminosilicate, which contributes to the good

mechanical properties and resistance to chemical

attacks in geopolymers. However, the physical and

mechanical properties of component phases do not

sufficiently explain the behavior of geopolymers,

instead microstructural analysis of geopolymers is

necessary. The information about the structure of

geopolymers obtainable by X-ray diffraction (XRD) is

rather limited. Consequently, scanning electron

microscopy (SEM) and Fourier transform infrared

(FT-IR) were considered for additional information in

this study. Some studies11-13 have investigated the

geopolymer structure by SEM and FT-IR. Generally,

excessive water content in cement mortar or concrete

blend reduces its final strength. However, in the case

of geopolymers the effects of water-cement ratio

(or water-powder ratio in the current study) are poorly

known. Both the mechanical properties and the

structure of MK based geopolymers may be affected by

the w/p ratio.

Currently, OPA is a waste by-product of the palm

oil industry, remaining a problem for the agro

industries in Thailand and other South East Asia. It has

been estimated that the annual industrial solid OPA —————

*Corresponding author (E-mail: [email protected])

Page 2: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

INDIAN J. ENG. MATER. SCI., DECEMBER 2017

500

waste in Thailand amounted to about 280,000 tons in

201114. Moreover, this production rate is likely to

increase, as the Thai government policy is to increase

the nationwide arable area. However, SiO2 is the main

chemical component in OPA, making it a potentially

valuable cementitious material.

In this study, OPA was used to replace MK in

geopolymer mortars at four levels, namely the OPA

content was varied at 0, 5, 10 and 15%. The mortars

were subjected to four types of heated curing, lasting

0.5, 1, 2 and 4 h, and had three levels of water to

powder ratios at 0.45, 0.55 and 0.65. The main

parameters studied were the compressive strength,

microstructure, and sulfuric acid resistance (losses of

strength and mass). Analytical methods used to

investigate the microstructure of the geopolymers

included scanning electron microscopy (SEM), X-ray

diffraction (XRD) and Fourier transform infrared

spectroscopy (FT-IR).

Experimental Program

Materials

The metakaolin (MK) used was collected from

Narathiwat province in southern Thailand. It was

calcined in an electric furnace at 750ºC for 2 h. The

X-ray diffractogram (XRD) of MK is shown in Fig. 1.

The crystalline phases of MK consisted of quartz, illite,

and microcline. Figure 2 shows the XRD of oil palm

ash (OPA), in which the crystalline phases detected are

quartz, calcite, and sylvite. The OPA was collected

from a palm oil mill in the Krabi province of southern

Thailand.

Particle size distributions of the ground MK and

OPA were measured from wet dispersions using the

Malvern Hydro 2000 MU volume sample dispersion

unit, available for the Mastersizer 2000 granulometer.

The particle size distribution parameters d10, d50 and d90

of the MK, as determined by laser diffraction, were

1.35, 6.31 and 88.81 µm, while those for the OPA were

4.32, 19.30 and 100.11 µm. The measurements gave

the average particle sizes (d50) 6.31 µm for MK and

19.31 µm for OPA. The cumulative size distribution

curves by volume are shown in Fig. 3. The chemical

compositions analyzed by X-ray fluorescence (XRF)

of the cementitious materials are given in Table 1. The

BET specific surface area (SSA), determined by nitrogen

adsorption using the Quantachrome Autosorb-1-MP,

was 13.61 m2/g for MK and 13.06 m2/g for OPA.

The alkaline solution used was a mixture of sodium

hydroxide (NaOH) in flakes of 98% purity, sodium

silicate (Na2SiO3), and water. The sodium silicate

solution had a composition by weight of 14.14% Na2O,

Fig. 1 — XRD pattern of MK

Fig. 2 – XRD pattern of OPA

Fig. 3 — Particle size of MK and OPA

Table 1 — Chemical compositions (wt%) of MK and OPA

Oxide SiO2 Al2O3 Fe2O3 CaO TiO2 MgO K2O other LOI

MK 50.30 41.02 1.05 0.33 1.50 - 4.08 - 1.72

OPA 38.37 1.48 3.01 13.84 0.21 3.00 14.09 5.57 20.43

Page 3: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

HAWA et al.: METAKAOLIN BASED GEOPOLYMERS

501

27.67% SiO2, and 56.28% H2O. The composition was

analyzed using inductively coupled plasma-optical

emission spectrometer (ICP-OES) for Na2O and SiO2,

and by drying at 103-105ºC for H2O. River sand was

used as a fine aggregate in the manufacture of

geopolymer mortars. The river sand had specific

gravity of 2.51 and fineness modulus of 3.12.

Synthesis of geopolymer mortars

Geopolymer mortar blends were mixed from the

raw materials (MK and OPA), river sand, and

alkaline activators (Na2SiO3, NaOH, and H2O). First,

sodium hydroxide was mixed into sodium silicate

and added to water for 5 min, and the average

mixture temperature became 74±2°C from reaction

heat. Based on our experiments, the mixtures with

different water to powder ratios show differences in

temperatures for a narrow range of 2°C. When the

raw materials and river sand were added for final

mixing of 5 min, the mixture temperature became

48±2°C. The texture of the samples was quite

sticky and fast setting, and to cast them in an acrylic

mold required effort. All of the samples were heated

in an electric oven at 80°C after casting. After

curing for 0.5, 1, 2, or 4 h, they \were removed

from the oven. After unwrapping and demolding,

the samples were stored at ambient temperature

(air curing) until they were tested. The

temperature-time profile of geopolymer process is

shown in Fig. 4. The details of mixture proportions,

modulus of activating solution and total oxides

molar ratios in the mixture composition are shown

in Table 2. The samples were mixed manually, not

with a motorized mixer. The geopolymer slurries

were poured into acrylic 50×50×50 mm molds to set,

and to form samples for compressive strength

and sulfuric acid resistance testing. The samples

were compacted as described in ASTM 109/C109M15.

Test and analysis methods

Compressive strength

The compressive strength tests were conducted

according to ASTM C109/C109M15. For each mixture

the prepared of specimens were 50×50×50 mm cubes

that were tested at 2 h, 6 h, and 24 h, and at 7 days and

28 days.

Sulfuric acid resistances

The geopolymer mortar samples for sulfuric acid

resistance were handled in accordance with ASTM

C109/C109M15. After heat curing in an oven for 2 h at

80°C, the 50×50×50 mm cubic specimens were

submerged in 5% sulfuric acid solution. The mass loss

and compressive strength were tested after immersion

in sulfuric acid for 28, 56, and 91 days.

Scanning electron microscopy (SEM)

SEM analyses were performed using a JMS-5800

LV model scanning electron microscope (JEOL,

Japan) to assess the microstructures of geopolymer

mortars. Small scraps of the samples were tested using

scanning electron microscopy.

X-ray diffraction (XRD) analysis

X-ray diffraction (XRD) tests were performed to

investigate the phase composition and the crystalline

content in unexposed MK, OPA and in geopolymer

Table 2 — Mix proportions of geopolymer mortars (1,000 g)

Sample name SiO2/Al2O3

(molar ratio)

Na2SiO3

(g)

NaOH

(g)

MK

(g)

OPA

(g)

Sand

(g)

Water

(g)

Heat time

(h)

G-0.5-0.45 2.77 112 44.8 189.4 - 568.3 85.3 0.5

G-1-0.45 2.77 112 44.8 189.4 - 568.3 85.3 1

G-2-0.45 2.77 112 44.8 189.4 - 568.3 85.3 2

G-4-0.45 2.77 112 44.8 189.4 - 568.3 85.3 4

G-2-0.55 2.77 112 44.8 189.4 - 568.3 85.3 2

G-2-0.65 2.77 112 44.8 189.4 - 568.3 85.3 2

G5-2-0.45 2.88 112 44.8 179.9 9.5 568.3 85.3 2

G10-2-0.45 3.01 112 44.8 170.5 18.9 568.3 85.3 2

G15-2-0.45 3.15 112 44.8 161.0 28.4 568.3 85.3 2

Fig. 4 —The temperature-time profile of geopolymer process

Page 4: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

INDIAN J. ENG. MATER. SCI., DECEMBER 2017

502

paste. Powder XRD was conducted using an X’Pert

MPD X-ray diffractometer (PHILIPS, Netherlands) at

angles from 5° to 40° (2𝜃) using the clay and rock

0.4 program.

Fourier transform infrared spectrophotometer (FT-IR) analysis

FT-IR analysis was performed using the KBr pellet

method on EQUINOX 55 spectrometer (Bruker,

Germany), across the wavenumber range 4000-400 cm-1.

Results and Discussion

Compressive strength

Almost all tested specimens set and formed hard

samples almost instantly when mixed at ambient

temperature. The compressive strength testing was

focused on the influences of curing time and different

water-powder ratios. Figure 5 presents the effects of

curing time on the compressive strength of geopolymer

mortars at 2, 6, and 24 h, and at 7 and 28 days after heat

curing. It should be noted that the specimens were under

heat curing during 0.5 to 4 h. The specimens were not

cured at 80°C up to 28 days. After heating curing

(0.5, 1, 2, or 4 h for different mixtures), the specimens

were left at ambient temperature until strength testing.

The early 2 h compressive strength at ambient

temperature is high for all mixtures, partly because the

samples were prepared from hot mixtures before curing

in an oven. However, the sample cured for 0.5 h (G-0.5-

0.45) had the lowest compressive strength, and it is very

different from those with longer curing at 80°C. The

curing may increase the degree of geopolymerization

because of increased reaction extent or different

reactions that form further (longer chained) products.

On the other hand, at longer ages, the geopolymerization

slightly developed further strength; especially those

specimens cured for 2 and 4 h.

The effects of water-powder ratio on compressive

strength of the MK-based geopolymer mortars cured at

80°C for 2 h are illustrated in Fig. 6. For all mixtures,

a low water-powder ratio with curing at an elevated

temperature was found to improve the compressive

strength. However, the G-2-0.45 and G-2-0.55 samples

with slightly different values were cured at ambient

temperature. The compressive strength at all ages of

these samples significantly improved to the range from

62.71 to 67.51 MPa. However, the decreased

compressive strength with the highest water-powder

ratio was clearly due to low extent of reactions. Similar

to OPC mortar or concrete, the G-2-0.65 sample had

low concentration of alkaline activator and high

amount of free water. It was observed that the 0.45 and

0.55 ratios showed slightly different values. This may

be attributed to the fact that the MK had comparatively

high specific surface area (13.61 m2/g).

The compressive strength results of the mixes with

varied OPA and w/p are presented in Fig. 7. It can be

observed that the compressive strengths of the

specimens without OPA (G-2) were slightly low,

especially for the 0.45 and 0.55 w/p samples.

Nevertheless, the compressive strength of samples

having partial replacement of MK with OPA clearly

Fig. 5 — Compressive strengths of geopolymer mortars across

various heat curing times

Fig. 6 — Compressive strengths of geopolymer mortars heat cured

at 80°C for 2 h

Fig. 7 — Compressive strengths of geopolymer mortars containing

OPA, heat-cured at 80°C for 2 h

Page 5: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

HAWA et al.: METAKAOLIN BASED GEOPOLYMERS

503

decreased with OPA content. On comparing mortar

samples with 5% OPA binder but varying w/p ratio, a

higher w/p ratio gave lower strength. This is because

high w/p ratio increases the free water, while

the average particle sizes (6.31𝜇m for MK against

19.31 𝜇m for OPA) correlate to the specific surface

area (SSA). Moreover, a higher w/p ratio increased the

porosity and the flakes in the binder matrix (see Fig. 8).

Chen et al.16 determined the effects of different

water-binder ratios on geopolymers with reservoir

sludge as a partial replacement of metakaolin, and high

water-binder ratios reduced the strength.

Scanning electron microscopy (SEM)

Microstructures of the MK-based geopolymers,

containing 5% OPA with diverse water to powder

ratios, were imaged by SEM with results shown in

Fig. 8. Some crude materials, which had not reacted,

were partly covered with flakes on the specimen

surfaces. It was evident that the G5-2-0.55 and G5-2-

0.65 matrices were not homogeneous and contained

little pores. The G5-2-0.65 specimen had numerous

flakes and the largest cracks (see Fig. 8c), while in the

G5-2-0.45 sample a lower content of unreacted crude

materials was apparent in the samples (see Fig. 8a).

It was likewise found that the G5-2-0.45 sample had

good homogeneity and the least number of pores across

these cases, and it had the least unreacted MK and

OPA. This case likewise delivered the highest 68 MPa

compressive strength, whereas the G5-2-0.55 and G5-

2-0.65 samples had in comparison greatly reduced

strength. This suggests that the dispersion of the

aluminosilicate materials from geopolymerization

gave the best compressive strength to the G5-2-0.45

case. The pores, flakes, and cracks in the geopolymer

reduce its compressive strength, and these are apparent

in Figs 8(b) and 8(c).

X-ray diffraction (XRD)

The X-ray diffractograms of the geopolymer

samples containing 5% OPA with various w/p ratios

are shown in Fig. 9. The crystalline phases are similar

Fig. 8 — SEM micrographs of geopolymer mortar containing 5% OPA: (a) G5-2-0.45, (b) G5-2-0.55 and (c) G5-2-0.65

Page 6: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

INDIAN J. ENG. MATER. SCI., DECEMBER 2017

504

in all cases, consisting of quartz, illite and microcline.

These cases are clearly distinct from the native MK.

However, it was observed that the geopolymer samples

had lost calcite and sylvite from the primary crystalline

phases of OPA. This is because of the small amount of

OPA used, and the OPA had only little calcite and

sylvite. It is indicative that the reactions between

powder and alkaline activator produce quartz, illite and

microcline minerals, which contributed to the

hardening of the geopolymer. At the high 0.55 and 0.65

w/p, the mineral phases in the sample paste remained

relatively unchanged. However, a significant increase

in the peak intensity of microcline was noted around

2θ= 27°, from the remaining microcline in MK.

Fourier transform infrared spectroscopy (FT-IR)

Figures 10 and 11 show the FT-IR spectra of the

MK, OPA, and the samples G5-2-0.45, G5-2-0.55, and

G5-2-0.65 for purposes of comparison. The major band

of MK at approximately 1,083 cm-1 is associated with

the asymmetric stretching vibrations of Si-O-(Al or Si).

Duxson et al.17 show that this band is typical for

amorphous aluminosilicate. The peak band is observed

at frequencies near those reported in literature18. The

broad adsorption bands of MK around 810 and 464 cm-

1 were assigned to Al-O vibrations and Si-O stretching

vibrations, respectively. In the OPA spectra, the -OH

groups were also detected at 1422 cm-1, and the band at

approximately 789 cm-1 indicates Al-O or Si-O-Al

groups. The main peak of OPA at around 1030 cm-1 is

attributed to the stretching of Si-O and Al-O.

Figure 11, the FT-IR spectra of the geopolymer

samples had major bands at around 3453, 1659, 1000,

723, 589, and 446 cm-1. The interesting findings from

geopolymerization products include the bands located

at approximately 3453 and 1659 cm-1 for the O-H

stretching vibrations and the H-O-H bending

vibrations, respectively. These bands indicate hydroxyl

groups18,19. The principal band associated with the Si-

O-(Al or Si) stretching vibrations at 1000 cm-1 had a

spectral peak for the geopolymer materials. The Si-O-

(Al or Si) bond in the reaction products has also been

found in earlier research20. Previous studies with fly

ash21,22 and metakaolin based geopolymers23-25

reported that the geopolymer materials has an obvious

peak at approximately 1000 cm-1. The principle groups

in the geopolymers were related to the base materials

in the MK, across the 400 and 2000 cm−1 range in the

FT-IR spectra. It has been formerly accounted that

some amount of unreacted crude materials will remain

in geopolymerized items26. The binder aluminosilicate

matrix was formed by geopolymerization to an extent

dependent on the w/p ratio, for these MK and OPA

based materials, and the alkaline activator affects the

FT-IR patterns. The impact of the w/p ratio on the

nanostructure may be rather small, based on the FT-IR

spectra in Fig. 11. The patterns in the FT-IR spectra

were slightly different in a previous study27 which

reported that fly ash based geopolymers with different

Fig. 9 — XRD patterns of geopolymers containing 5% OPA, heat

cured for 2 h

Fig. 10 — FT-IR spectra of the raw materials

Page 7: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

HAWA et al.: METAKAOLIN BASED GEOPOLYMERS

505

w/p ratios have very similar FT-IR spectra, suggesting

similar nanostructures.

Three samples showed common behavior in the

FT-IR spectra. To investigate this further, changes in

the chemical bonds from the FT-IR analysis are shown

in Fig. 12. It is observed that the binder system gives a

strong peak at approximately 1000 cm-1 which can be

attributed to the asymmetric stretching mode of Si-O-

(Al or Si) bond, as indicated previously in Fig. 12. The

compressive strength of the geopolymer mortars is

associated with high absorption in this band, indicating

that the quality of geopolymer can to some extent be

assessed from the FT-IR results. In addition, the area

under FT-IR curves for the 0.45, 0.55, and 0.65 w/p

samples, with 5% OPA and heat cured for 2 h, is shown

in Fig. 12. The areas under the curve were 156,054.5,

160,678.1, and 161,305.6 for 0.45, 0.55, and 0.65 w/p,

in this order. It is observed that the area under FT-IR

curve significantly correlates with the compressive

strength (see Fig. 7) of the geopolymer.

Sulfuric acid resistance

Compressive strength

Figure 13 shows the effects on compressive strength

from exposure to sulfuric acid. The G-2-0.45 sample

had the highest compressive strength before treatment

with sulfuric acid. However, after 28 days of

immersion the compressive strength of G-2-0.45 had

lost 62.8%. In contrast, the G15-2-0.45 sample had

only 40.7% strength loss. It may be that OPA limits the

aluminosilicate bonding in geopolymer mortars,

and these bonds are vulnerable to sulfuric acid.

After 28 days the compressive strengths of the

geopolymers were steady with no more changes.

Mass loss

The sulfuric acid resistances of geopolymer mortars

were tested after heat curing in the oven at 80°C for

2 h. Figure 14 presents the mass losses of geopolymer

Fig. 11 — FT-IR spectra of geopolymer samples containing 5%

OPA, heat-cured at 80°C for 2 h

Fig. 12 — FT-IR spectra of geopolymer samples

Fig. 13 — Compressive strengths of geopolymer mortars after

exposure to 5% H2SO4 solution

Page 8: Effect of water-to-powder ratios on the compressive strength …nopr.niscair.res.in/bitstream/123456789/43704/2/IJEMS 24... · 2018-03-09 · This study investigated the properties

INDIAN J. ENG. MATER. SCI., DECEMBER 2017

506

mortars exposed to sulfuric acid for up to 91days. The

results show that the control samples had the highest

mass loss of 4% at 28 days of exposure, while the

samples with OPA lost only about 2%. After 28 days

of exposure, the remaining mass losses were very

small. This can be attributed primarily to the

geopolymerization in geopolymer mortars. The

geopolymer prepared using sodium hydroxide and

sodium silicate had the best performance, which is

attributed to its stable cross-linked aluminosilicate

polymer structure25. In the present study, the

compressive strengths of samples containing OPA had

significant losses across all exposure times.

Conclusions

The effects of OPA content in MK-based

geopolymers with various w/p ratios were

determined on compressive strength, durability, and

microstructure, with the following conclusions:

(i) The geopolymerization can be accelerated with

long heat curing, even though compressive

strength develops rapidly early on.

(ii) The compressive strength of geopolymers is

significantly affected by the w/p ratio. It decreases

with the w/p ratio, showing similar pattern as

earlier found for OPC mortar and concrete.

(iii) The compressive strength of geopolymer increases

with relative density. The effects of w/p ratio are

attributed to the pore structure and the flakes in the

material.

(iv) The OPA can decrease the mass loss of MK based

geopolymers under exposure to sulfuric acid. OPA

also protected against strength loss at long

exposure times.

Acknowledgments

The authors would like to thank anonymous

reviewers for their valuable comments. Special thanks

go to Dr. Seppo Karrila, Research and Development

Office (RDO), Prince of Songkla University for

reviewing and correcting the presentation.

References 1 Temuujin J, Minjigramaa A, Lee M, Chen-Tan N & van

Riessen A, Cem Concr Compos, 33 (10) (2011) 1086-1091.

2 Sata V, Sathonsawaphak A & Chindaprasirt P, Cem Concr Compos, 34 (5) (2012) 700-708.

3 Tchakoute H K, Elimbi A, Yanne E & Djangang C N, Cem Concr Compos, 38 (2013) 75-81.

4 Ranjbar N, Mehrali M, Alengaram U J, Metselaar H S C & Jumaat M Z, Constr Build Mater, 65 (2014) 114-121.

5 Detphan S & Chindaprasirt P, Int J Min Metall Mater, 16 (6) (2009) 720-726.

6 Hwang C L & Huynh T P, Constr Build Mater, 101 (1) (2015) 1-9.

7 Amin N, Faisal M, Muhammad K & Amin W, R S C Adv, 83 (2015) 67814-67819.

8 Al-Zahrani M M, Maslehuddin M, Al-Dulaijan S U & Ibrahim M, Cem Concr Compos, 25 (4-5) (2003) 527-537.

9 Hardjito D & Rangan B V, Technical Report, Curtin University of Technology, Perth, 2005.

10 Slaty F, Khoury H, Rahier H & Wastiels J, Appl Clay Sci, 104 (2015) 229-237.

11 Puertas F, Palacios M & Vazquez T, J Mater Sci, 41 (10) (2006) 3071-3082.

12 Kani E N & Allahverdi A, J Mater Sci, 44 (12) (2009) 3088-3097.

13 Kriskova L, Pontikes Y, Zhang F, Cizer Ö, Jones P T, van Balen K & Blanpain B, Cem Concr Res, 55 (2014) 59-68.

14 Hawa A, Tonnayopas D, Prachasaree W & Taneerananon P, Ceram Silikáty, 57 (4) (2013) 319-327.

15 ASTM Standard Test Method for Compressive Strength of

Hydraulic Cement Mortars (Using 2-in. or [50-mm] Cube Specimens).C109/C109M-07, Philadelphia, 2007.

16 Chen J H, Huang J S & Chang, Y W, Cem Concr Compos, 33 (5) (2011) 602-610.

17 Duxson P, Provis J L, Lukey G C & van Deventer J S J, in

Proc of the World Congress Geopolymer, J Davidovits (Ed.),

2005, 21-25.

18 van Jaarsveld J G S, van Deventer J S J & Lukey G C, Mater

Lett, 57 (7) (2003) 1272-1780.

19 Frost R L & Vassallo A M, Clays Clay Miner, 44 (5) (1996)

635-651.

20 Granizo M L, Blanco-Varela M T & Mart´ınez-Ram´ırez S,

J Mater Sci, 42 ( 9) (2007) 2934-2943.

21 Bakharev T, Cem Concr Res, 35 (6) (2005) 1233-1246.

22 Valcke S L A, Pipilikaki P, Fischer H R, Verkuijlen M H W & van Eck E R H, Mater Struc, 48 (3) (2015) 557-569.

23 Bernal S A, Provis J L, Rose V & Mejía de Gutierrez R, Cem Concr Compos, 33 (1) (2011) 46-54.

24 Kouamo H T, Elimbi A, Mbey J A, Sabouang C J N & Njopwouo D, Constr Build Mater, 35 (2012) 960-969.

25 Aredesa F G M, Camposa T M B, Machadoa J P B, Sakaneb K K, Thima G P & Brunellia D D, Ceram Int, 41 (6) 7302-7311.

26 Wang H, Li H & Yan F, Colloids Surf A: Physicochem Eng Aspects, 268 (1-3) 1-6 (2005).

27 Ismail I, Bernal S A, Provis J L, Hamdan S & van Deventer J

S J, Mater Struc, 46 (3) (2013) 361-373.

Fig. 14 — Mass loss of geopolymer mortars after exposure to 5%

H2SO4 solution