You are on page 1of 9

This article was downloaded by: [American Public University System]

On: 26 February 2014, At: 03:09


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Sustainable Cement-Based


Materials
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tscm20

Drying shrinkage and creep


performance of geopolymer concrete
a b a a
Kwesi Sagoe-Crentsil , Trevor Brown & Alan Taylor
a
CSIRO Materials Science and Engineering , Highett , Australia
b
Faculty of Engineering & Industrial Sciences , Swinburne
University of Technology , Hawthorn , Australia
Published online: 20 Feb 2013.

To cite this article: Kwesi Sagoe-Crentsil , Trevor Brown & Alan Taylor (2013) Drying shrinkage and
creep performance of geopolymer concrete, Journal of Sustainable Cement-Based Materials, 2:1,
35-42, DOI: 10.1080/21650373.2013.764963

To link to this article: http://dx.doi.org/10.1080/21650373.2013.764963

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Journal of Sustainable Cement-Based Materials, 2013
Vol. 2, No. 1, 35–42, http://dx.doi.org/10.1080/21650373.2013.764963

Drying shrinkage and creep performance of geopolymer concrete


Kwesi Sagoe-Crentsila,b*, Trevor Browna and Alan Taylora
a
CSIRO Materials Science and Engineering, Highett, Australia; bFaculty of Engineering &
Industrial Sciences, Swinburne University of Technology, Hawthorn, Australia
Downloaded by [American Public University System] at 03:09 26 February 2014

The creep behavior and drying shrinkage performance of fly ash geopolymer (GP)
concrete mixtures have been investigated using equivalent grade 40-MPa ordinary
Portland cement (OPC) concrete as the reference system. Drying shrinkage values,
measured in accordance with AS 1012.13-1992, up to one year fell well below the
nominal 700 μstrain limit, with GP concrete values typically less than 400 μstrain at
1 year. Variations in basic mean creep coefficient of concrete (Фcc.b), measured as
the ratio of the creep strain to elastic strain for a specimen loaded at 28 days under
constant stress of 0.4fc (fc – 28 day compressive strength) was monitored for up to
52 weeks. Values of Фcc.b obtained for steam-cured GP concrete was found to be of
the order of 45% lower than corresponding OPC concrete. Based on these results,
the effects of GP concrete mixtures on load-dependent (creep) deformations appear
to be negligible. The paper further discusses factors that contribute to observed GP
concrete creep deformation and drying shrinkage performance.
Keywords: geopolymer; fly ash; drying shrinkage; creep; compressive strength

1. Introduction are available concerning the effects of


The prediction of concrete creep and dry- each of these variables on creep and
ing shrinkage remains a critical parameter shrinkage behavior of concrete made with
in concrete specification and is of crucial OPC and blended cements.[3–4] Drying
importance for durability and long-time shrinkage, generally leads to cracking
serviceability of concrete structures.[1] and, even though it may not affect the
While creep of concrete is associated structural integrity, durability problems
with deformation of hardened concrete are generally increased and is perhaps
due to applied constant load, drying one of the most deleterious properties of
shrinkage refers to internal moisture loss Portland cement concrete.[5] Correspond-
of the hardened concrete. In ordinary ing studies investigating various aspects
Portland cement (OPC) systems, both of mix composition and engineering per-
creep and drying shrinkage are known to formance of geopolymer (GP) systems
be affected by variable factors including including shrinkage and creep are now
cement type, w/b ratio, aggregate type only beginning to emerge [5–9].
and content, age, temperature relative Hardjito et al. [10], in their study of
humidity of the environment, and age creep of medium strength GP concrete,
and size of member.[2] In the literature, a observed lower creep coefficient values
considerable amount of experimental data compared to equivalent grade OPC

*Corresponding author. Email: Kwesi.Sagoe-Crentsil@csiro.au

Ó 2013 Taylor & Francis


36 K. Sagoe-Crentsil et al.

concretes. These researchers furthermore 2. Experimental


showed that the specific creep of GP con- 2.1. Raw materials
crete decreased as the compressive
Table 1 shows chemical composition of
strength increased. This observation cor-
ASTM Type I OPC used as reference for
relates well with conventional OPC con-
concrete mixtures and Class F fly ash
crete.[11–14] Typical literature values of
used for the GP mixtures. Commercial
specific creep for conventional 50–60-
“N” grade sodium silicate solution with
MPa Portland cement concrete after one
SiO2/Na2O molar ratio of 3.22 and
year are of the order of 50–60 micro-
28.7% SiO2 content with total solids con-
strain,[12] with this value decreasing for
tent of a 38% supplied by PQ Australia
higher strength concretes. In the case of
was used in mixtures.
Downloaded by [American Public University System] at 03:09 26 February 2014

high-performance high-volume fly ash


River sand and local 9-mm Hornfels
concrete, for instance, Malhotra and
coarse aggregate was used. The coarse
Mehta [15] reported specific creep was in
aggregate fraction was received in a satu-
the region of 30 microstrain/MPa after
rated surface-dry condition.
one year.
While the roles of curing on
mechanical properties of GP concretes
2.2. Specimen preparation and curing
are yet to be fully established, it has
been observed that the drying shrinkage Table 2 provides concrete mix propor-
strains of heat-cured GP concrete are tions of the concrete formulations. Both
generally lower compared to equivalent the reference OPC and GP concrete
values observed for concretes cured mixes were designed to achieve target
under ambient conditions,[16] This 28-day strength of 40 MPa.
effect was attributed to water which is The reference OPC concrete samples
released during the chemical reaction were prepared in a similar fashion as GP
process of ambient-cured GPs which concrete and molded according to stan-
subsequently evaporate over a period of dard requirements for compressive
time causing significantly large drying strength measurements. After casting, all
shrinkage strains particularly within the samples were placed in the steam curing
initial two weeks. chamber at 65 °C for 6hrs. Thereafter, all
Other material-related parameters that samples, were stored at 23 °C and 100%
can influence engineering performance of RH until required for testing. Standard
GP binders, such as the pore network concrete cylinders of 100  200 mm were
distribution, have also been investigated.
[17,18] Kriven et al. [17] examined the
pore network distribution of GP binders Table 1. Chemical analysis of OPC and fly
and observed that there exist several ash (mass %).
clusters of pore sizes and found them
not too dissimilar to those characterizing Oxide OPC Fly ash
OPC systems. In this paper, the early- SiO2 20.2 47.19
age basic creep behavior of GP concrete Al2O3 4.16 29.79
is investigated using equivalent grade Fe2O3 5.30 13.93
Portland cement concrete as the refer- CaO 64.8 3.29
MgO 1.29 1.38
ence system. The experiments assess the Na2O 0.22 0.24
influence of age and loading on the K2O 0.42 0.49
actual creep in early age, while measur- TiO2 – 1.77
ing drying shrinkage response of con- SO3 2.67 0.13
crete specimens. LOI 1.34 1.3
Journal of Sustainable Cement-Based Materials 37

Table 2. Concrete mix proportions. accordance with a modified procedure of


AS 1012.1 -1993. The cast specimens
OPC were stored in the fog room (set at 23 °C
concrete (kg/ GP concrete
Material m3 ) (kg/m3) and 100% RH) for 6 days and then stored
in a standard temperature and humidity
Fly ash – 269 controlled room (at 23 °C and 50% RH)
OPC 346 –
Coarse 1112 1163 and measurement taken at regular inter-
aggregate vals.
(9 mm)
Silicate – 148
solution/ 2.5. Creep
NaOH
Downloaded by [American Public University System] at 03:09 26 February 2014

Creep measurements were carried out on


Dry sand 753 784
Water 186 27
the prisms cut from panels after 28 days
of casting, measuring 150  150 
300 mm, in accordance with Australian
Standard as 1012.16-1996 (methods of
used for compressive strength measure-
testing concrete – determination of creep
ments. Table 3 provides data on mea-
of concrete cylinders in compression).
sured fresh concrete properties.
The load for determining creep was
applied after 2 days and readjusted after 7
2.3. Compressive strength and 28 days. The loads in all creep tests
The compressive strength was measured corresponded to 40% of the compressive
in accordance with AS 1012.9-1999 [19] strength of the specific concrete mixtures
at a loading rate of 20 MPa/min – Stan- at the time of loading. Total strain was
dard Test Method for Compressive measured on specimens in the creep test
Strength of Cylindrical Concrete Speci- rig, while drying shrinkage strain was
men. Specimens were capped using con- obtained from parallel unloaded speci-
ventional sulfur capping method in mens in a separate complementary creep
accordance with the Australian Standard test rig under same conditions as for the
AS 1012.9 (1999) for concrete compres- creep measurements. The calculated load-
sive strength tests at 1, 3, 7, and 28 days. induced creep strain at any time is the
Tests were performed in a 2500 kN ELE difference between the average strain val-
Universal testing machine. ues of the loaded and control specimens.
The creep loads used in the current test
were calculated based on 40% of 28-day
2.4. Drying shrinkage strength.
The standard drying shrinkage measure-
ment of the aerated GP products was per-
formed on 75  75  300-mm prisms in 3. Results and discussion
3.1. Compressive strength
The mean compressive strength develop-
Table 3. Properties of fresh concrete. ment of steam-cured OPC and GP con-
cretes is shown in Table 4. The results in
Property OPC concrete GP concrete Table 4 show that GP and OPC samples
Mix temp. (°C) 15.7 25.5 varied slightly from the target strengths
Flow (mm) – 260 of 40 MPa at 28 days. It is also observa-
Slump (mm) 120 160 ble that different strength development
Wet density 2380 2430
(kg/m3)
trends occur for the two heat-cured con-
crete types up to 28 days.
38 K. Sagoe-Crentsil et al.

Table 4. Compressive strength development strength increase occurs between 7 days


of ambient cured OPC and GP concretes. and 28 days for GP concretes, which is
typical of this class of materials, suggest-
Compressive strength (MPa)
ing that significant polymerization reac-
Age (days) OPC steam-cure GP steam-cure tion occurs during the initial heat curing
1 22.5 42.3 regime compared to equivalent hydration
7 28.0 42.3 reactions characterizing OPC systems
28 38.1 44.0 (see Figure 1).

The heat-cured OPC specimens typi-


cally show greater strength development, 3.2. Drying shrinkage
Downloaded by [American Public University System] at 03:09 26 February 2014

achieving nearly 40% increase between The drying shrinkages of the mixes are
1 day and 28 days during fog-room stor- shown in Figure 2, where it can be seen
age. However, only a corresponding mar- that the OPC mixtures showed excellent
ginal increase in compressive strength of drying shrinkage properties both at early
less than 5% is observed between 1 day ages and at ages up to one year. It is evi-
and 28 days for the heat-cured GP sam- dent from the trends shown in the plots
ples. This strength increase is a result of that the drying shrinkage steam-cured
residual polymerization reactions that samples of both GP and OPC concrete
occur beyond the initial heat exposure. fell well below the 700 μstrain limit
Furthermore, only a relative modest specified in AS 3600-2001.[20]

Figure 1. (a) Uniaxially loaded specimens under constant temperature and humidity conditions.
(b) Close-up of loaded creep specimens.
Journal of Sustainable Cement-Based Materials 39

0
0 50 100 150 200 250 300 350 400

-100
GP
OPC
-200
MICROSTRAIN

-300

-400

-500
Downloaded by [American Public University System] at 03:09 26 February 2014

-600
TIME (DAYS)

Figure 2. Drying shrinkage for 40-MPa GP and OPC concrete.

The drying shrinkage trends of the are accelerated by heat curing, the associ-
OPC and GP concrete at early ages (i.e. ated initial drying shrinkage was signifi-
<7 days) were very similar. At this early cantly reduced, signifying the completion
age, as expected, both binders register of initial condensation reactions and
rapid increase in drying shrinkage. Given moisture loss from capillaries, as also
the 7-day moist curing of the specimens confirmed by the associated rapid
prior to shrinkage measurements, the initial strength development.
high shrinkage of the GP samples could The lower drying shrinkage observed
reflect the time taken for them to equili- for GP concretes may also be partly due
brate with the surrounding environment. to a finer disconnected capillary network
The average change in shrinkage values structure. Correspondingly, contributions
between 7 and 56 days of the GP samples of chemical or autogenous shrinkage fac-
was of the order of 250 μstrain. This value tors of GP concrete shrinkage also remain
is much lower compared to the value of unknown. Chemical shrinkage has a pro-
400–450 μstrain experienced by Portland found effect on the macroscopic deforma-
cement concrete in agreement with obser- tion of the concrete volume. For instance,
vations of Hardjito et al. [21] for heat- in OPC concrete, the dimensional
cured fly ash-based GP concrete after one changes caused by chemical shrinkage
year. are a result of the hydration reactions that
Similar to OPC concrete, it may be occur when the anhydrous cement phases
expected that the main mechanism for react with available water. Thus, the vol-
drying shrinkage in GP concrete is the ume change caused by chemical shrink-
build-up of negative pressure within the age occurs when the concrete is in a
capillary network of the bulk cement plastic state. Overall, early age concrete
paste as meniscii form. The resulting shrinkage is of increasing concern, as it
stresses lead to contraction of the con- can be responsible for cracking when the
crete unless restrained.[22] The observed concrete has not gained significant
results suggest that increased condensa- strength to withstand internal stresses.
tion processes occurring in steam-cured
GP concretes may have implications on 3.3. Creep
the shrinkage process. Thus, when the Figure 3 shows variation of the basic
condensation reaction of GP concretes creep coefficient with time up to
40 K. Sagoe-Crentsil et al.
Downloaded by [American Public University System] at 03:09 26 February 2014

Figure 3. Average basic creep coefficient for steam cured 40-MPa grade OPC and GP concrete.

52 weeks. As shown in the figure, basic basic creep coefficient, determined from a
creep coefficient for 40-MPa grade GP standard test, can where necessary, be
concrete is 40–60% lower than the corre- extrapolated to a final test value by any
sponding OPC concrete. Creep is the acceptable mathematical model for creep
time dependant increase in strain under behavior, calibrated such that Фcc.b is also
sustained stress for a given material. At predicted by the chosen model.
the application of a load, the instanta- From the instantaneous creep, the
neous strain includes elastic strain and basic creep factor has been calculated for
some creep strain. test specimens, as illustrated in Figure 3.
Basic creep is taken as the increase in After adding trend line of best fit, the
strain ½ðtðtÞ above the initial elastic resulting mathematical formula or model,
strain under sustained loading, minus for Basic Creep coefficient is Фcc.b =
shrinkage strains over the same time 0.2687ln(x)  0.168. The creep coeffi-
period. Therefore, cient, defined as the ratio of creep
strain-to-elastic strain after one year of
cbðtÞ ¼ TðtÞ  EðtÞ  sðtÞ ð1Þ loading for the GP concrete with 40 MPa
compressive strength is of the order of
where at any time t after loading: 0.10. The values are well within the rec-
cbðtÞ = basic creep strain ommended range for conventional con-
E = elastic strain upon loading crete as specified in the Australian
sðtÞ = shrinkage Standard AS3600 for Portland cement
Australian Standard AS3600-2009 concrete.[20]
(Concrete structures) Section 3.1.8.2 [16] Wallah [7] obtained creep coefficient
gives the basic creep coefficient of con- values between 0.6 and 0.7 for heat-
crete (Фcc.b), as the ratio of the creep strain cured GP concrete after one year for
to elastic strain for a specimen loaded concrete with compressive strength of
at 28 days under constant stress of 0.4fc 40, 47, and 57 MPa and 0.4 for GP
(fc – 28 day compressive strength). Thus, concrete with compressive strength of
basic creep coefficient is an indicator of 67 MPa.
concrete deformation due to loading; thus, While the dependence of creep strain
lower creep coefficient represents a higher on factors such as binder loading, aggre-
resistance to deformation under load. The gate type and ratio, air entrainment as
Journal of Sustainable Cement-Based Materials 41

well as the age of the concrete at the time 4. Conclusions


of loading, are well established; the rela- The following observations have been
tive effects of these factors on GP con- made based on laboratory drying shrink-
crete remain uncertain. Hence, in age and creep tests for nominal 40-MPa
comparing current results with published GP and OPC concretes samples up to
work, such as that of Hardjito et al. [21], 52 weeks:
it is vital to recognize contributions aris-
ing from differences in GP mix composi- (1) Test results obtained from this
tions and cure conditions. However, study disclose that primary struc-
notwithstanding results obtained from the tural properties of GP concretes,
current study are in good agreement with namely, compressive strength, dry-
Downloaded by [American Public University System] at 03:09 26 February 2014

the limited published data for GP con- ing shrinkage, and creep behavior
cretes which fall well within specification broadly conforms to existing stan-
range of OPC concretes as stipulated, for dards and guidelines for OPC sys-
example, in the Australian standard AS tems.
3600.[20] (2) All mixes displayed excellent dry-
The scope of the current investiga- ing shrinkage properties, as mea-
tion did not allow for more comprehen- sured in accordance with AS
sive testing to determine the influence 1012.13-1992. Shrinkage values
of stress levels or other materials and obtained at both early and latter
process parameters. Hence, further test- ages up to one year fall well
ing need to be undertaken to develop a below the nominal 700 μstrain
more comprehensive data-set for creep limit (Australian Standard for
behavior and also to benchmark against Concrete Structures AS3600 -
the performance of conventional con- 2005), with GP concrete values
cretes. typically less than 400 μstrain at
It is further worth noting that similar one year.
to the current results, the measured creep (3) Similar to OPC concrete, the main
strains of fly ash-based GP concrete mechanism for drying shrinkage
obtained by Wallah [7] were significantly in GP concrete appears similar to
smaller than predicted equivalent grade the build-up of negative pressure
Portland cement concrete. While the within the capillary network of the
exact cause for this trend remains bulk cement paste as meniscii
unclear, the observed smaller creep form which generate stresses
strains of fly ash-based GP concrete may resulting in contraction of OPC
be also due in-part to unreacted or concrete. However, acceleration of
partially reacted fly-ash particle residues condensation reactions of heat-
acting as ‘micro-aggregates’ in the sys- cured GP concretes suggest
tem, as suggested by Davidovits.[9] significant mitigation of early age
Hence, the additional restraining action drying shrinkage arising from the
due to micro-aggregates could potentially combined effects of nearly
contribute to smaller creep compared to completed condensation reactions
traditional OPC concrete for which as confirmed by associated rapid
cement hydration reaction is relatively initial strength development.
more complete. Particularly, given that (4) Variations in basic mean creep
the aggregate fraction remains primarily coefficient of concrete (Фcc.b)
responsible for counteracting creep and obtained for steam-cured GP con-
drying shrinkage deformations of the bin- crete was found to be of the order
der phases.
42 K. Sagoe-Crentsil et al.

of 45% lower than corresponding [10] Hardjito D, Wallah SE, Sumajouw DMJ,
OPC concrete. The combined Rangan BV. On the development of fly
ash-based geopolymer concrete. ACI
effects of greater shrinkage occur-
Mater. J. 2004;101:467–472.
ring within OPC material, on the [11] Neville AM. Properties of concrete
one hand, and possible restraining (Fourth and Final ed.). Essex: Pearson
effects due to unreacted fly ash Education, Longman Group; 2000.
residue particles acting as micro- [12] Warner RF, Rangan BV, Hall AS, Faul-
kes KA. Concrete structures. Melbourne:
aggregates in GP concretes appear
Longman; 1998.
to be major contributors to [13] Neville AM, Dilger WH, Brooks JJ.
observed differences in creep per- Creep of plain and structural concrete.
formance of the two concrete London: Construction Press, Longman
Downloaded by [American Public University System] at 03:09 26 February 2014

types. Group; 1983.


[14] Gilbert RI. Time effects in concrete
structures. Amsterdam: Elsevier; 1988.
[15] Malhotra VM, Mehta PK. High-perfor-
References mance, high-volume fly ash concrete: mate-
[1] Gilbert RI. Creep and shrinkage models rials, mixture proportioning, properties,
for high strength concrete - proposal for construction practice, and case histories.
inclusion in AS3600. Austr. J. Struct. Ottawa: Supplementary Cementing Materi-
Eng. 2002;4:95–106. als for Sustainable Development; 2002.
[2] Hanson W. Early age creep and stress [16] Rangan VB. Developments in Porous, bio-
relaxation tests. In: Bentur A, Kovler K, logical and geopolymer ceramics: ceramic
editors. Early age cracking in cementi- engineering and science. In: Manuel Brito,
tious systems. Cachan Cedex: RILEM Eldon Case and Waltraud M. Kriven edi-
Publications S.A.R.L; 2002. 257–65. tors. Proceedings, vol. 28, issue 9, Devel-
[3] Standards-Australia, methods of testing opments in porous, biological and
concrete - Determination of creep of con- geopolymer ceramics. The American Cera-
crete cylinders in compression, AS mic Society; Daytona Beach, Fl 2008.
1012.16-1996. [17] Kriven WM, Bell JL, Mallicoat SW,
[4] Nawy Edward G, editor. Chapter 26 in Gordon M. Intrinsic microstructure and
concrete construction engineering hand- properties of metakaolin-based geopoly-
book. 2nd ed. New York (NY): CRC mers’. In: International workshop on
Press; 2007. geopolymer binders, interdependence of
[5] Collins F, Sanjayan JG. Cracking composition, structure and properties;
Tendency of Alkali-Activated Slag Con- Weimar, Germany. Edited by Bauhaus-
crete Subjected to Restrained Shrinkage, Universitat; 2006, p. 71–86.
Cement and Concrete Research, Vol. 30, [18] Steveson M, Sagoe-Crentsil K. Relation-
No. 5, 2000, pp. 791–798. ships between composition, structure and
[6] Krivenko PV. Alkaline cements: termi- strength of inorganic polymers. J. Mater.
nology, classification aspects of durabil- Sci. 2005;40:2023–2036.
ity. In: Proceedings of 10th International [19] Australian Standard AS 1012.9-1999
Congress on the Chemistry of Cement; Methods of testing concrete - determina-
1997; Goteborg (Sweden). p. 4iv046– tion of the compressive strength of con-
4iv050. crete specimens. Standards Australia.
[7] Wallah SE. Creep behaviour of fly ash- 19999.
based geopolymer concrete. Civil Eng. [20] Standards-Australia, concrete structures,
Dimen. 2010;12:73–78. Draft for Public Comment Standards Aus-
[8] Davidovits J. Properties of geopolymer tralia (Revision of AS 3600–2001); 2005.
cements. In: Krivenko PV, editor. Alkaline [21] Hardjito D, Wallah SE, Sumajouw DMJ,
cements and concretes, Vol. 1. Vipol Rangan BV. Fly ash-based geopolymer
Stock. Ukraine: Kiev; 1994. 131–49. concrete. Aust. J. Struct. Eng.
[9] Davidovits J. Soft mineralurgy and geo- 2005;6:77–86.
polymers. In: Proceedings of Geopoly- [22] Bissonnette B, Pierre P, Pigeon M. Influ-
mer ‘88, First European Conference on ence of key parameters on drying shrink-
Soft Mineralurgy; Compiegne, France: age of cementitious materials. Cement
The Geopolymer Institute; 1988. Concr. Res. 1999;29:1655–1662.

You might also like