You are on page 1of 46

Test methods for sulfate resistance

of concrete and mechanism


of sulfate attack

State-of-the-art Review
Aimin Xu, Ahmad Shayan and Pud Baburamani

producti on pf this report


rted by the National
is up
spo Service program
Interest
NIS

Review Report 5

arOb
Transport
Research
Test methods for sulfate resistance
of concrete and mechanism
of sulfate attack

A state-of-the-art review

Aimin Xu
Ahmad Shayan
Pud Baburamani

Production of this report is supported by the


National Interest Service program
NIS

ARRB Transport Research Ltd


Review Report 5
September 1998
Information Retrieval

XU. A, SHAYAN, A and BABURAMANI, P. (1998): TEST METHODS FOR SULFATE RESISTANCE
OF CONCRETE AND MECHANISM OF SULFATE ATTACK: a state-of-the-art review. ARRB
Transport Research Ltd. Review Report 5. A review of worldwide practice, current literature and
research. XX pages including figures and tables.

ABSTRACT: Sulfate attack on concrete has been studied worldwide for more than 60 years.
However, the mechanisms of attack are still not entirely understood, and deterioration of
concrete from sulfates still occurs.

The source of the sulfates may be either 'external' or 'internal'. External sources are the naturally
occurring sulfates in the environment or those sulfates that are the product of industrial
processes or various human activities (e.g. fertilisers often release sulfates into the soil and
groundwater). Internal sources of sulfate may include the sulfates introduced in the cements
from which concrete is made.

Standard tests have been developed to evaluate the resistance of concretes to sulfate attack.
Some, but not all of these tests, take into account the mechanisms of sulfate attack so far
discovered in various research work.
One of the first attempts to increase the resistance of concrete to sulfate attack was based on
the finding that calcium aluminates in cement react with sulfates to cause expansion of
concrete. Some countries, as a result, started to regulate the composition of cements used to
make concrete.
The recent incorporation of pozzolanic materials in concrete, though having some beneficial
effects, complicates the study of the mechanisms of sulfate attack in concrete.
This report comprises a review of the standard tests developed to date for sulfate attack, and
how the test methods have evolved. It also reviews laboratory research that has been done on
the mechanisms of sulfate attack.

About the authors Aggregates and Rocks for


Aimin Xu Engineering purposes. He has
Research Engineer 18 years of experience in
Review Report 5 research and consulting on the
Aimin's career started in 1982 as
September 1998 various aspects of AAR, and
an engineer at the Electric Power
developed the accelerated 21-
Construction Research Institute
ISBN 0 86910 772 0 day AAR test for Australian
after graduating from Nanjing
ISSN 1441-3973 aggregates. Dr Shayan has
Institute of Technology, China, in
successfully led in excess of 90
1982 with a Bachelor (Hons)
consultancy projects in the area
Degree in civil engineering. He
of concrete durability and
Any material may be reproduced earned his Ph.D in concrete science
published more than 80 papers
without permission provided the at Chalmers University of
in the international scientific
source is acknowledged. Technology, Sweden in 1992, and a
media.
Licentiate of Engineering Degree in
Although the Report is believed to the same university in 1990. During Pud Baburamani
be correct at the time of the period of his study in Sweden, Senior Research Scientist
publication, ARRB Transport Aimin spent six months at the
Pud obtained his Ph.D. degree
Research Limited, to the extent University de Sherebrook, Canada,
from the Department of
lawful, excludes all liability for loss as a Visiting Scientist researching
Materials Engineering, Monash
(whether arising under contract, concrete microstructure. He joined
University in 1985. As a
tort, statute or otherwise) arising ARRB Transport Research in 1997
materials engineer/scientist,
from the contents of the Report or as a Research Engineer in the
Pud has over 15 years of
from its use. Where such liability Concrete Area. Prior to starting his
career in Australia, Aimin worked as experience in the
cannot be excluded, it is reduced to characterisation of
the full extent lawful. Without an inspection engineer with STK
microstructure, mechanical
limiting the foregoing, people Inter Test AB, Sweden.
properties and deformation
should apply their own skill and behaviour of engineering
Ahmad Shayan
judgement when using the materials. He joined ARRB
Chief Scientist
information contained Transport Research in 1989,
in the Report. Dr Ahmad Shayan, a well-known
and his current position is
international expert and Australia's
Senior Research Scientist with
Wholly prepared and printed by leading investigator in the field of the Materials Technology
ARRB Transport Research Ltd alkali-aggregate reaction, heads up
Group.
500 Burwood Highway our AAR team. He is currently the
Vermont South VIC 3133 Chairman of the Standards,
AUSTRALIA Australia Committee CE/12, on
Review Report 5 arOb
Test methods for sulfate resistance of concrete and mechanism of
Transport
Research
sulfate attack

Contents
EXECUTIVE SUMMARY vi

1. INTRODUCTION 1

2. MECHANISMS OF SULFATE ATTACK 1


2.1. Source of Sulfate 2
2.2. Action of Sulfates 3
2.2.1. Sodium sulfate 3
2.2.2. Calcium sulfate 4
2.2.3. Magnesium sulfate 4
2.2.4. Ammonium sulfate 5
2.2.5. Effect of carbonic acid and chloride 5
2.3. Deterioration mechanisms 7
2.3.1. Volume Change 7
2.3.2. Decomposition of paste compounds 8
2.3.3. Salt exfoliation 8
2.4. Internal Sulfate Attack 9

3. TEST METHODS 11
3.1. Aggressive Sulfate Content in Water and Soil 11
3.2. 2. Standards and Test Methods 12
3.3. Research on Test Methods 17
3.3.1. Cement Composition Requirements 17
3.3.2. Tests for Internal Sulfate Attack 19
3.3.3. Performance-orientated Tests 21 ARRB Transport
Research Ltd
3.3.3.1. The pH of the sulfate solution 21
ACN 004 620 651
3.3.3.2. Specimens 24
HEAD OFFICE:
3.3.3.3. Pre-test curing 24 500 Burwood Highway
3.3.3.4. Duration of sulfate exposure and assessment of failure Vermont South
VIC 3133
25
AUSTRALIA
3.3.3.5. Different types of sulfates and concentrations 26 Tel: (03) 9881 1555
3.3.3.6. Influence of temperature 27 Fax: (03) 9887 8104
3.3.3.7. Influence of loading 27 Email: info@arrb.org.au
Internet: www.arrb.org.au
3.3.3.8. Influence of other reactive agents 28
PERTH OFFICE:
3.3.3.9. Precision of various tests 29 Street address:
Unit 5, 4 Brodie Hall Drive
4. CONCLUSIONS 30 Technology Park
Bentley, WA 6102
AUSTRALIA
5. REFERENCES 31 Postal address:
PO Box 1068
Bentley, WA 6982
AUSTRALIA
Tel: (08) 9472 5544
Fax: (08) 9472 5533
Email: maryl@arrb.org.au
Review Report 5 arGlb
Transport Researct
Test methods for sulfate resistance of concrete and mechanism of
sulfate attack

Executive Summary

Sulfate attack on concrete has been studied worldwide for more than 60 years.
However, the mechanisms of attack are still not entirely understood, and deterioration
of concrete from sulfates still occurs.

The source of the sulfates may be either 'external' or 'internal'. External sources are
the naturally occurring sulfates in the environment or those sulfates that are the product
of industrial processes or various human activities (e.g. fertilisers often release sulfates
into the soil and groundwater). Internal sources of sulfate may include the sulfates
introduced in the cements from which concrete is made.

Standard tests have been developed to evaluate the resistance of concretes to sulfate
attack. Some, but not all of these tests, take into account the mechanisms of sulfate
attack so far discovered in various research work. The tests range from those that
monitor changes in the strength of concrete specimens after set periods of immersion
in known composition sulfate solutions, to those that use X-ray diffraction to examine
concrete specimens for expansive products (e.g. ettringite and thaumasite) that have
resulted from sulfate attack.

One of the first attempts to increase the resistance of concrete to sulfate attack was
based on the finding that calcium aluminates in cement react with sulfates to cause
expansion of concrete. Some countries, as a result, started to regulate the composition
of cements used to make concrete.

The recent incorporation of pozzolanic materials in concrete, though having some


beneficial effects, complicates the study of the mechanisms of sulfate attack in concrete.

The results of 'accelerated' tests on concrete specimens (the type of test usually done
in the laboratory) are of limited use in interpreting the mechanisms of sulfate attack in
the field. Also, concrete specimens prepared and studied in the laboratory are not
necessarily of the same composition as concretes that have suffered sulfate attack in
the field.

This report comprises a review of the standard tests developed to date for sulfate
attack, and how the test methods have evolved. It also reviews laboratory research
that has been done on the mechanisms of sulfate attack.
1
Review Report 5

1. Introduction
Sulfates occur naturally and are used in industry. They can enter concrete in solution form
from the external environment, or they may be mixed into concrete (e.g. in cement). They
react with cement compounds to form expansive products. Large amounts of sulfate may
have deleterious effects on concrete. Concrete resistance to sulfate attack has been studied
worldwide in order to explain the mechanisms of sulfate attack and to evaluate the durability
of concrete in the sulfate-rich environments.
An early attempt to increase concrete resistance to sulfate attack was based on the
finding that calcium aluminates in cement were the major constituents reacting with sulfate to
cause excessive expansion. In many countries national standards were established to regulate
the composition of cement for concrete used in aggressive environments. Laboratory tests
were established to estimate directly the sulfate resistance. In most of these texts concrete
specimens were stored in sulfate-containing solutions and were monitored for changes in
their properties over time. Some of these methods have been adopted as standard test
methods by standards organisations or they have become national codes.
Although sulfate attack on concrete has been studied for more than 60 years, some
details of the mechanisms of attack are still not known, and concrete failure still occurs.
Furthermore, the use of new and supplementary cementing materials in recent decades (e.g.
pozzolans), while enhancing some aspects of concrete durability, makes the picture of sulfate
resistance more complicated because of the varied effects ranging from beneficial to
deleterious.
`Accelerated' (the usual type of laboratory tests) on the sulfate resistance performed
under various conditions give complex and sometimes confusing results. While there is
insufficient understanding of the mechanisms of attack, it is difficult to devise suitable test
procedures and to interpret results satisfactorily. Efforts have been made, sometimes at
national level, to modify or establish standard methods that may more reliably evaluate
concrete resistance to sulfate attack.
This report reviews some of the mechanisms that are believed to account for the
deterioration of concrete in the presence of various sulfates. The test methods established as
ASTM and AS standards etc. are described (including their evolution), and those tests used in
research are also discussed.

2. Mechanisms of Sulfate Attack


`Sulfate attack' can be defined as 'the deterioration of concrete as a result of physical—
chemical interactions between the minerals in hydrated portland cement paste and sulfate
from the environment.' (Mehta et al, 1992) The interactions cause expansion, cracking,
spalling or even disintegration of concrete. It may be noted that attack of the same nature can
also occur as a result of internal reactions; e.g. the formation of the same types of compounds
can occur where there is reaction between cement/concrete constituents and sulfates. The
common feature is the appearance of excess sulfate-bearing phases representable by
ettringite*in the deteriorated concrete. Recent reviews of the actions of sulfate and their
damaging effects on concretes are provided by Lea (1998) and St John et al. (1998).

* Ettringite (3CaO•Al2O3.3CaSO4.32 H2O) is a type of AFt (calcium alumino-ferrite trisulfate)


compound with a typical needle or long rod shape/crystals. AFt has the general formula [Ca3 (Al,
Fe)(OH)6] X3 xH2O where X represents a formula unit of doubly charged anion e.g. SO4 2-. Amorphous
ettringite is often found in concrete that has been attacked by sulfates.

ARRB Transport Research Ltd


2
Review Report 5

`External' sulfate attack is often associated with other deterioration mechanisms such as
the leaching of cement hydration products (which causes softening of concrete), salt
exfoliation (which is more common for the sulfate salts), or organic acids attack, etc. These
concurrent degradations make the cause-and-result relationship more complicated, and the
mechanisms of sulfate deterioration less easy to study. The category of sulfate attack is thus
broadened, and therefore many of the degradation types have been investigated concurrently
with the sulfate-related deterioration.
`Internal' sulfate attack is mainly due to the formation and growth of ettringite in the
hardened cement paste. It is caused by excessive sulfate in the cementing material, and also
by early age encounter with temperatures higher than 60°C (e.g. when concrete is steam
cured). The latter is known as 'delayed ettringite formation'. Use of sulfate or sulfide-
contaminated aggregate may also lead to internal sulfate attack.

2.1. Source of Sulfate


Seawater and ground water are the main sources of external sulfates. Sewage water contains
sulfates and also affects concrete structures. According to a review by Mehta, (1993)
seawater contains an average of 2700 mg/L SO4 (0.028 M), but the sulfate content in ground
water varies according to the cation type in the soil, i.e. the sulfate content ranges from high
for alkali-sulfates to low for alkaline earth metal sulfates.
The average amounts of the major constituents for a selection of seawaters are given in
Table 1. These seawaters vary in content of the different salts, depending on where the
samples were taken; e.g. the Gulf Sea, which receives relatively small inflows from rivers and
is in a hot arid climate, has a high content of salts. Interestingly, the ratio of sulfate to
chloride is almost identical, i.e. 0.14 (mass / mass) for the waters shown in Table 1.
Sulfate in ground water mainly comes from the natural minerals in soils, which may be
enriched due to evaporation, particularly in hot climates (Al-Amoudi 1995). Acid sulfate
soils are produced when ground is excavated and there is oxidation of iron sulfides that are
present in soil. In many cases human activities are the source of sulfates, e.g. when
(NH4)2SO4-containing agricultural fertiliser is placed on soil, when high-sulfur fuels are burnt
by industry and cause acid rain; or when some chemical industries release sulfate salts into
the surrounding areas. In outfall sewers, sulfate attack on concrete pipes is caused by
hydrogen sulfide in the sewage gas (especially when temperature is higher than 23°C and the
ventilation is poor). In chimneys and railway tunnels there is also sulfate attack (Taylor
1977). In addition, concrete structures in industrial areas and on busy highways are exposed
to and absorb high amounts of sulfur on their surfaces, which results in the formation of
gypsum and reduction in pH of pore solutions (Belazs et al. 1997). Waste disposal areas may
contain strong sulfate brine (Bonen 1997).
Some types of hydrogen-sulfide producing bacteria and sulfur-oxidising bacteria which
exist in soil or in water drainage systems are the cause of some reported concrete
deterioration (Tazawa et a/. 1994; Sand et a/. 1994); in these cases sulfate attack is occuring
under acidic conditions.

ARRB Transport Research Ltd


3
Review Report 5

Table 2. Main composition of some seawaters (grams per litre)


Sea Na K Mg Ca CI SO4 Reference
Mediterranean 11.56 0.42 1.78 0.47 21.38 3.06 Lea 1998
Atlantic* 9.95 0.33 1.50 0.41 17.83 2.54
10.57 0.08 1.26 0.57 19.10 2.60 Regourd
1981
Gulf Sea 20.70 0.73 2.30 0.76 36.90 5.12 Al-Amoudi
1995
12.60 0.47 1.60 0.48 23.40 3.30 Damghani
and Vakili
Hormuz Strait 2.19 0.07 0.26 0.05 3.96 0.58 1995
Mean seawater 11.00 0.40 1.33 0.43 19.80 2.76 Lea 1998
* NaC1 26.9 g/L, MgC12 3.2 g/L MgSO4 2.2 g/L, CaSO4 1.2 g/L, CaC12 0.6 g/L, KHCO3 0.2 g/L.

The source of 'internal' sulfate is usually the gypsum (CaSO4.2f 2O) added into cement
and/or the sulfates brought in with aggregates and some mineral admixtures. Crammond
(1984) studied sulfate-containing aggregates; he noted that there was petrographical evidence
that coarsely crystalline gypsum present as aggregate reacts with cement compounds to form
plates of portlandite. Shayan (1988), has shown that sulfides present in some aggregate may
later oxidise to sulfuric acid which attacks Ca(OH)2 as well as CSH and causes concrete
deterioration in the process through the formation of hydrated sulfate phases. In principle, the
source of sulfate can be controlled by careful selection of an appropriate cement, aggregate,
and water. Unless the amount of sulfate is too high, e.g. > 4% as SO3 by mass of cement,
significant expansion of the concrete does not occur (Crammond 1984). It has been shown,
especially in the past 10 years or so, that the harmful effect of internal sulfate is potentially
greater if concrete is cured at high temperature in its early age.

2.2. Action of Sulfates


The chemical reactions between sulfates and cement paste constituents have been recognised
as the major cause of the deterioration of concrete. In principle, the sulfate salt ionises in
water to attack the cement hydrates, typically Ca(OH)2, calcium silicate hydrate (CSH) and
calcium aluminate hydrates. The actual reaction process depends on the type of sulfate salt
and the reaction products, and especially on the solubility of each constituent - which, in
turn, is a function of the pH value of the solution.

2.2.1. Sodium sulfate


The most vulnerable phase of the cement hydration products is portlandite (Ca(OH)7),which
reacts with sodium sulfate to form gypsum (CaSO4.2H20):

Ca(OH)2 +2Na+ + +10H2 0 = CaSO, -2H2 0 +2Na+ +20H- +8H20


(1)
(portlandite) (gypsum)

Gypsum has a lower solubility (0.01 mole/L at 25°C) than Ca(OH), (0.02 mole/L at 25°C)
and precipitates out, while Ca(OH)2 dissolves to replenish Ca2+ . This reaction can proceed to
completion if the sodium hydroxide produced can be removed, e.g. by flowing water.
Otherwise, the accumulation of alkali in the pore solution will suppress the solubility of
Ca(OH)2 due to the common ion effect, and the above reaction will not continue. Reaction (1)
is unlikely to take place between sulfate and CSH (calcium silicate hydrate, the other main
hydration product of cement), because the solubility of CSH is lower than that of

ARRB Transport Research Ltd


4
Review Report 5

gypsum,(Lea 1998) though some researchers believe that the reaction may occur to some
extent (Lawrence 1990).
If the concentration is high enough, Na2SO4 can react with the aluminium-bearing phases
to form ettringite (3Ca0•A1203•CaSO4.32H2O), which is stable in a certain range of the salt
concentration: i.e. at a concentration of Na+ 0.25 mole/L, ettringite is stable when the
concentration of S042+ is 0.00067 to 0.13 mole/L (Damidot et al. 1992) In this case, the
presence of Ca(OH)2 in the system is essential for supplying the Ca2+ ions. It has been
reported that carbonated concrete might be more resistant to sodium sulfate attack (Sersale
1997).
The mechanisms, which can be expressed as the reaction between environmental sulfates
and calcium aluminate hydrate (Equation 2) or calcium monosulphoaluminate hydrate in the
hydrated cement (Equation 3), both result in the formation of ettringite (Mehta et al. 1992;
Lawrence 1990).

4CaO • A12 03 • 19H2 0 + 3S042- + 2Ca(OH) 2 + 1 1 H, 0 =3Ca0 • A12 03 • 3CaSO4 • 32H2


(calcium aluminate hydrate) (ettringite)
(calcium trisulfoaluminate)
(2)

3Ca0. A1,03 • CaSO4 • 12H2 0 + 2SO4'2- +2Ca(OH)2 +18H20


(calcium monosulphoaluminate)
(3)
= 3CaO • A1203 .3CaSO4 .32H2O
(ettringite)

2.2.2. Calcium sulfate


Calcium sulfate present in soil or resulting from the reaction between Ca(OH)2 and sodium
sulfate (secondary gypsum), can react with the calcium aluminate hydrate, the same as that in
fresh cement paste or concrete, to produce ettringite (Lea 1998):
3(CaSO4 .2H20)+4Ca0 • A1203 •19H20 + 17H20
(secondary gypsum) (calcium aluminate hydrate)
(4)
= 3CaO A1203 -3CaSO4 •32H20+ Ca(OH) 2
(ettringite)

Ettringite is a type of AFt (aluminate ferrite tri-sulfate) compound with a typical needle or
long rod shape. Amorphous ettringite is also very often seen in concrete attacked by sulfate.
The hypotheses proposed to explain the concrete deterioration are based on the growth of
ettringite, causing inner stress in the surrounding paste, and that amorphous ettringite absorbs
a large number of water molecules which leads to further volume expansion. (Mehta et al.
1992; Mehta 1973).

2.2.3. Magnesium sulfate


Magnesium sulfate reacts with Ca(OH)2 and CSH to produce secondary gypsum
(CaSO4.2H20) and silica gel (Si02):

ARRB Transport Research Ltd


5
Review Report 5

Ca(OH) 2 + Mg 2+ +SO4- +2H2 0 =CaSO4 .2H2 0 + Mg(OH) 2


(5)
(gypsum) (brucite)

3Ca0 .2Si02 +3Mg 2+ ± 3S042- 9H2 0 = 3( CaSO4 2H2 0)+3Mg(OH) 2 2Si02

(gypsum) (brucite) (silica gel)


(6)

The latter reaction, according to Lea, (1998) results in a low pH value with the saturation of
Mg(OH)2 (brucite, 0.0002 mole/L, equilibrium pH = 10.25). At this pH, the CSH phases are
not stable and Ca(OH)2 will be released into solution to establish the equilibrium pH. In the
presence of MgSO4, the above reactions proceed until gypsum can precipitate out. Regourd
(1981) showed the gypsum can be in a flat prism shape growing adjacent to Ca(OH)2. At the
same time, Mg can substitute Ca in the CSH to form CMSH or MSH, a soft white material
(Regourd 1981; and 1988; Crammond & Haliwell 1995; Cole 1953; al-Amoudi 1994; Bonen
1994). Some studies have established that the white material may have a composition of
4Mg0-Si02-8.5H20 (Lea 1998). Because of its destructive effect on CSH phases, magnesium
sulfate has caused greater expansion of portland cement concrete than has sodium sulfate
(Cohen & Bentur 1988).
The action of MgSO4 on the aluminates is similar to that of other sulfates. Regourd
(1981) showed that nearly amorphous ettringite or a thick fan-shaped mass resulted from the
MgSO4 attack on concrete. However, the low pH associated with MgSO4 attack leads to a
gradual dissociation of ettringite (unstable at pH < 10.6 — see Lea 1998) into gypsum and
gibbsite (Al(OH)3). The dissociation of ettringite, on the other hand, may in part ease the
destructive expansion, which may explain the observation that magnesium sulfate is less
aggressive to calcium aluminate cements than sodium sulfate (Lea 1998).

2.2.4. Ammonium sulfate


Ammonium sulfate in solution reacts with Ca(OH)2 in concrete to produce gypsum
(CaSO4.2H20) and release ammonium into solution:
(NH4 ) 2 SO4 ± Ca(OH) 2 +2H20 CaSO4 .2H,0+ 2NH:-
(7)
(gypsum)

The resulting NH4+ will further interact with OW in solution to form NH3 and water:

2NH: + 20H - 2NH3 T +2H2 0 (8)


This process neutralises concrete pore solution, which will in turn cause the cement hydration
products to become unstable and eventually decompose. This is especially hazardous to the
concrete that contains pozzolanic materials as a partial replacement for cement, (there is a
very low Ca(OH)2 content to buffer the acidic type of attack.

2.2.5. Effect of carbonic acid and chloride


Carbonic acid ions in water percolating into concrete or formed through the dissolution of
atmospheric CO2 in concrete pore water may react with calcium, silica and sulfate to form
thaumasite, CaSiO3•CaCO3.CaSO4-15H20, a fibrous crystalline phase.
In cement paste, ettringite is transformed into thaumasite in the presence of carbon
dioxide and silica liberated by the carbonation as stated by Lukas (1975) and Regourd (1981),
or else the thaumasite is crystallised over the ettringite as speculated by Taylor (1977).
Thaumasite often coexists with ettringite. (Cf. sulfate attack on aggregate, where thaumasite

ARRB Transport Research Ltd


6
Review Report 5

is the dominant product formed - see Lea 1998). An exception reported by Baronio and Berra
(1986) was where a sample of deteriorated concrete in contact with aggressive water of pH
below 7 had developed a large quantity of thaumasite in the absence of ettringite. This could
be attributed to the fact that ettringite is not stable at such a low pH. The reaction can be
expressed as:

3CaO A1203 • 3CaSO4 • 32H2 0 +3CO2 +2S/O2 + 2Ca 2+


(ettringite) (source of carbonic ions)
(9)
=2CaSiO3 • CaCO3 • CaSO4 • 15H20 + A12 03 + CaCO3 + CaSO4 • 2 H2 0
(thaumasite)

The silica for forming thaumasite may come from deteriorated aggregates, e.g. mica and
feldspar, and from CSH (Baronio and Berra 1986). Through field investigations and
laboratory work, Crammond and Halliwell (1995) found that concrete containing finely
powdered calcium or magnesium carbonate was potentially more vulnerable to thaumasite-
caused sulfate attack. Apparently the carbonate dust helps supply carbonic ions for the
reaction.
Thaumasite is more stable than ettringite, and its formation is favoured at low
temperatures (0-5°C) and high relative humidity, e.g. 90% RH (Crammond 1985; Gaze 1997).
St John et al. (1998) warned: in tunnels and other underground works where high humidity
and low temperature prevail the possible presence of thaumasite should always be
considered.
The occurrence of thaumasite in concrete that had undergone sulfate attack was reported
by Erlin and Stark. Regourd (1981) showed thick bands of thaumasite in concrete samples
degraded by seawater, Baronio and Berr (1986; 1987) identified thaumasite in badly
deteriorated tunnel concrete, some of which was reduced to a whitish soft mass. Regourd
(1981) also reported the presence of ettringite and thaumasite in the cracks of certain
feldspathic aggregates containing micaceous layers.
Concrete with intensive formation of thaumasite, like concrete with ettringite-induced
expansion, may become softened and completely disintegrate. Whereas ettringite results from
the reaction between aluminates and the sulfatious, thaumasite can be due to the reaction
between silica-bearing phases and sulfates; in the latter case the sulfate-resisting cement is
not effective in preventing sulfate attack.
Another major type of ion that often coexists with sulfate is chloride, such as occurs in
seawater (sodium chloride or brine). Experiments have demonstrated that high concentrations
of chloride in sulfate solutions significantly mitigate the deterioration of concrete (Al-
Amoudi 1995). In a study of a CaO—A1203—CaSO4—CaC12—H20 system, Damidot and Glasser
(1997) calculated the concentrations of Calf, AO+, S042-, CF and OH in equilibrium with
various possible precipitates. They demonstrated that Ca(OH)2 is not stable at high C1-
concentrations and is converted to 3CaO•CaC12.15H20, which dissolves to increase the
concentration of Ca2+ in solution so as to maintain electro-neutrality. However, the high
aqueous a- concentrations stabilise ettringite so that it is stable at lower pH (pH=9.5) in the
presence of CF as compared to the minimum pH 10.5 in the system without chloride.
Surveys of reinforced concrete structures in Arabian Gulf by Matta (1993) revealed the
sulfate content in the concrete was generally higher than 4% SO3 by mass of cement, the
highest value being 27%. In spite of such high sulfate being present in the concrete, very few
instances of deteriorations were found to be related to sulfate attack. Laboratory experiments
of Al-Amoudi (1995) have confirmed that in the presence of chloride, concrete did not
exhibit either physical or chemical deterioration. The higher solubility of ettringite in the
seawater (pH 8) and the high ambient temperature in the area were considered to be the main
reasons. The high concentration of a- ions increased the solubility of Ca(OH)2 and the higher

ARRB Transport Research Ltd


7
Review Report 5

temperature (concrete under summer sun may have a temperature of about 70°C) reduced the
expansion force of the ettringite. This also seems to be related to the concentration of sodium
in seawater, because certain amounts of sulfate and chloride in the water are necessary to
maintain the equilibrium with sodium; i.e. a higher sulfate concentration is required to
stabilise ettringite when the concentration of sodium is higher (Damidot & Glasser 1997).
However, these observations by no means prove that marine concrete structures generally are
exempt from sulfate attack.

2.3. Deterioration mechanisms


2.3.1. Volume change
Observations of sulfate attack have proved that deteriorated concrete has undergone
significant expansion, which is in accordance with the reaction mechanisms mentioned
above. The reactions form a significantly larger volume of solid phases which cannot be
accommodated by the capacity of the pore system. Internal stresses increase and the concrete
swells and then cracks.
Although, as pointed out by many researchers (Mehta et al. 1992; Lea 1998) the
expansion has rarely shown a quantitative relationship with the expansive phases, e.g. with
the amount of ettringite formed, it is still pertinent to note the theoretical volume changes that
may occur under optimum conditions. Table 2 lists molecular volumes of some of the main
constituents in the reaction mechanisms. It should be mentioned that the expansion is
possible only if the sulfates are presented in the solution phase. In other words, if the sulfates
were in the solid phase as reactants, then some of the reactions would lead to shrinkage, as
mentioned by Mehta (1973). From the data in Table 2, it can be estimated that transformation
of portlandite to gypsum by reaction with Na2SO4 will double the solid volume, whereas it
will be tripled if MgSO4 is the reactant. The expansive compounds listed in this table,
however, have several varieties depending on the number of water molecules in their
formula; e.g. ettringite may be dried to a water content of 8I- 2O without a great change in
crystalline structure (Taylor 1990) or it may be in some poor crystalline form. Further
absorption of water by such types of ettringite will lead to additional stresses in the concrete.
Ettringite growth is not rapid because the supply of sulfate is a diffusion-controlled
process, and the crystals can find their way along the capillaries or grow in air voids. A case
similar to this, as has been demonstrated by Scrivener and Lewis (1997), who studied the
heat-cured mortars and delayed ettringite formation, is where ettringite can crystallise in the
space made available by the expansion of the cement paste (i.e. the ettringite does not cause
expansion, for a certain period of time). With the filling of air voids, further expansion of
concrete may be expected to be more rapid, and thus it will be more quantitatively related to
the amount of ettringite formed.

Table 2 Molecular volumes of some compounds that may be present in concrete A

Compound Volume (cm'/mole)


Ca(OH)2, portlandite 33.1
Mg(OH)2, brucite 24.6
CaSO4.2H20, gypsum 74.2 _
Na2SO4.10H20, mirabilite 219.8
MgSO4.7H20, epsomite 146.8
4Ca0-A1203.19H20, tetra-calcium aluminate 19-hydrate 369.2
3CaO•A1203•CaSO4.12H20, mono-sulfate 312.7
3CaO•A1203.3CaSO4.32H2O, ettringite 714.9 —
Al(OH)3, gibbsite 32.0
K2Ca(SO4)2.6(OH)2, syngenite 127.8
data from CRC Handbook of Chemistry and Physics (1996) and Lea's Chemistry of Cement and
Concrete. Source: (Lea 1998).

ARRB Transport Research Ltd


8
Review Report 5

2.3.2. Decomposition of paste compounds


As indicated in the reaction equations, sulfate attack causes the cement hydration products to
decompose, and reduces concrete strength. The decreased pH of the pore solution, resulting
for example from the attack by magnesium sulfate and ammonium sulfate, is primarily
responsible for these effects. The effects of sulfate attack will be amplified by other factors
such as soft water or acidic water percolation, which cause more leaching.
Mehta et al (1992) have stressed that in most cases the loss of adhesion and strength,
rather than expansion and cracking, were the primary manifestations of sulfate attack. These
were evidenced by mineralogical studies, which revealed the decomposition of cement
hydration products such as CSH and Ca(OH)2. This explains why the expansion hypothesis
does not always provide a satisfactory explanation of the concrete deterioration. Low water
permeability and the ionic diffusivity of concrete, for example, are also vitally important for
concrete to resist sulfate attack.

2.3.3. Salt exfoliation


The solubility of sulfate salts such as Na2SO4 and MgSO4 increases markedly with
temperature, as shown in Figure 1. Note that the hydration rates (number of the combined
water molecules) of these salts are also very sensitive to temperature. Concrete pores filled
with salt solution may become supersaturated when the temperature goes down, e.g. at night,
and salt crystals may grow rapidly and exert stresses on the concrete inner structure.
Supersaturated salt solutions can suddenly be crystallised due to vibration, e.g. by a passing
vehicle; this might explain the `pop-out' of a concrete highway surface after the temperature
dropped overnight, as noted by Sayward (1984).
Haynes et al. (1996) reported several cases of damage to concrete in contact with ground
water; severe surface scaling occurred at the evaporation surface. These authors pointed out
that field concrete damaged by sulfate attack commonly shows a loss of adhesion and
strength, even though the amount of ettringite is not significantly high. They speculated that
the crystallisation of gypsum and sodium sulfate hydrate might be the main cause; they
proposed that the salts might become dehydrated with the evaporation and re-hydrated when
water is again available. The crystallisation of the hydrated form of the salts causes
expansion.
Similarly, St John (1982) showed that severe salt exfoliation had occurred in tunnels
where ground water containing sulfate salts was percolating through concrete, or was drawn
up by capillary forces. The evaporation of moisture enriches the salt content in the pore
solution, which in combination with solubility changes with temperature decreases leads to
the formation of salt crystals. However, the most destructive effect is due to the sudden
crystallisation of hydrated salts, e.g. Na2SO4-10H20 and MgSO4.7H20, associated with
temperature fluctuation, whereas the gradual formation of salt crystals by evaporation alone
provides more seed crystals, thus limiting the supersaturation degree and the size of crystals
(Taylor 1996).

ARRB Transport Research Ltd


9
Review Report 5

35

30

25 -

Solubility (wt %)
20 - MgSO4•7H20

15 -

10 -

CaSO4 .2H20

ox
0 5 10 15 20 25 30 35
Temperature (°C)

Figure 1. Solubility of some salts in water. (Source: St John 1982)

2.4. Internal Sulfate Attack


Internal sulfate attack occurs when deterioration of concrete is caused by sulfate that has been
introduced with cementitious materials and/or with aggregates. Delayed ettringite formation
occurs when the primary ettringite, that forms very early in the hydration of portland cement,
is decomposed due to heat curing at early age, and precipitates much later after the concrete
has hardened; 'secondary ettringite formation' refers to the formation of ettringites from
previous forms of ettringite (eg. primary ettringite) through dissolution, redistribution and
precipitation in other locations (eg. pores and cracks).
Delayed ettringite formation, i.e. formation and growth of ettringite in concrete long
after hardening as a result of later reactions between concrete constituents, has largely been
observed for hydrothermally treated concrete (Taylor 1996; Scrivener 1996; Klemm & Miller
1997) and for concrete rich in cement with low w/c* which underwent high temperature
curing (minimum 60°C) and drying—wetting (20°C) cycles during service (Start et al. 1992).
Babuskin et al (1972) calculated the Gibbs function (AG) of formation of AFt and AFmT
through the reaction of CA* and gypsum, and showed that at above 60°C AFm had a more
negative AG than AFt (Figure 2). Ettringite is thus not thermodynamically stable at high
temperatures and may be converted to calcium monosulphoaluminate hydrate (AFm)
(3CaO.A1203.CaSO4-12H,0) and gypsum. The dilemma of this is that the reverse reaction
occurs at lower temperatures, and will occur slowly after the hardening of the concrete — to
form ettringite, This will cause expansion. Odler (1997) categorised this type of ettringite as
`delayed ettringite' and 'secondary ettringite', and pointed out that the latter mainly
crystallises in cracks where water is more available. Somewhat higher critical temperatures
have been suggested by various authors based on test results, e.g. 70°C by Heinz et al.
(1989).

* `w/c' is short for 'water to cement' ratio.


Afm is a group name for calcium-aluminoferrites with the general formula [Ca4A1,Fe)(OH)61X.xH,0
where x represents one formula unit of a singly charged anion, or half a formula unit of a doubly
charged anion. Calcium monsulfoaluminate (3CaO.A1203•CaSO4 .12H2O) belongs to this group.
C.A = 3CaO.A1,03

ARRB Transport Research Ltd


10
Review Report 5

-20

-25 —

-30 —

AFm + gypsum
59.6°C

-45 —

-50 —
AFt

-55
200 250 300 350 400
Temperature (K)

Figure 2. The dependence of AG of phases on temperature for C3A + 3CaSO4 + H2O


system. (Source: After Babuskin et al 1972)

In recent years, researchers have hypothesised that during cement hydration, sulfate ions
are absorbed by CSH especially at high temperatures; these sulfate ions will be released
slowly into the pore solution after the concrete has cooled down (Sayward 1984; Fu &
Beardoin 1996). It was therefore speculated that SO3 in clinkers contributes to delayed
ettringite formation (Hime et al. 1996). Scrivener (1996) demonstrated that expansion of
concrete cured at normal temperatures, in relation to internal ettringite formation, depended
upon the sulfur/alumina ratio of the CSH; i.e. no expansion was observed when the mole ratio
is near unity, whereas if the ratio is close to three, expansion occurred. Hime et al. (1996)
stated that while clinker sulfate levels of 1.5% (as SO3) might cause delayed ettringite
formation, the lower SO3 limit for high-alkali cements might be 2% or more. Note that this is
lower than the commonly accepted sulfate level in cement, e.g. maximum SO3 content =
2.5% for the sulfate-resisting portland cement set by BS4027.
However, such assumptions were not always confirmed by experimental results. Klemm
and Miller (1997) tested 33 clinkers with SO3 contents ranging from 0.03% to 3.0% with SO3
to alkali ratios of 0.05 to 2.7 and sulfur to alumina ratios of 0.01 to 1.2; they showed that
most of the sulfate in clinkers had gone into solution to form ettringite when the paste was
still plastic.
Michaud et al. (1997) showed that ettringite or more general calcium sulfoaluminates
are not stable in contact with alkali-silica solutions when the concentration of silica is more
than a few millimoles per litre. Sulfate ions are not incorporated in the solid phases but
remain in the pore solution, migrating and forming new ettringite where the concentration of
silica is low enough.
In many cases of concrete deterioration, delayed ettringite formation (DEF) has been
associated with alkali-silica reaction, and researchers have argued (Shayan & Ivanusec, 1996)
as to which one is the dominating mechanism causing the deterioration. In fact, AAR is a
faster process than DEF, and occurs much earlier. Shayan and Quick (1991/92) proved that
secondary ettringite, which develops only after a long period of moist curing, mainly forms in
the cracks created by the AAR, and does not contribute to the deterioration observed after
short-term exposure (62 days). The same authors also showed that the formation of ettringite
may have caused OH" ions to be released from the Ca(OH)2, which in turn enhances the AAR
(Shayan & Quick 1992).

ARRB Transport Research Ltd


11
Review Report 5

Johansen et al. (1994) have reported that cracks had formed in some concrete railway
sleepers. Although the concrete therein had some AAR-reactive aggregates, the quantities
might not have been sufficient to cause the total amount of expansion that was observed.
They hypothesised that the cracks, especially those around the aggregates (`ring cracks')
were created by overall expansion in the paste due to delayed ettringite formation. They
stated that the width of the fissure should be proportional to the aggregate size. More detailed
microstructural examinations would be necessary to prove this hypothesis.

3. Test Methods
Some national codes and international standards organisations have established specifications
for cement/concrete that is in service in sulfate-aggressive environments. Such specifications
may include criteria for different severities of sulfate attack, as well as standard test methods
to evaluate the performance of concrete. It should be noted that standard test methods vary
from country to country in terms of the test environments and assessment of various types of
concrete failure; many test methods are still being modified.

3.1. Aggressive Sulfate Content in Water and Soil


National codes of different countries (or organisations) give different limits for the aggressive
sulfate content for the safe service of concrete structures. The international organisation
CEMBUREAU set 600 mg 50427L (600 ppm) in water and 6000 mg S042-/kg (6000 ppm) in
soil as the boundary between moderate and severely aggressive sulfate media where sulfate-
resisting cement should be used. Based on these limits of sulfate content, the European
Prestandard ENV 206 states 500 ppm in water and 3000 ppm in soil as its criteria (Lawrence
1990). However, ACI recommended 1500 ppm S042- in water and 2000 ppm 5042- in soil
respectively (Fagerlund 1987). Usually more detailed recommendations are made, e.g. such
as that summarised by BRE for concrete in service in sulfate-containing soil and ground
water (see Tables 3, 4, 5 and 6).

Table 3 Sulfate aggressiveness to concrete, CEMBUREAU


Aggressiveness
Class 1 Class 2 Class 3 Class 4 Class 5
Water SO4 (ppm) < 200 200 - 600 600 - 3000 3000 - 6000 > 6000
Soil SO4 (%) < 0.2 0.2 - 0.6 0.6 - 1.2 > 1.2
w/c, max. 0.55 0.55 I 0.50 0.50 0.45 0.45 + protection
Cement type — — Sulfate - resistin• cement

Table 4 Recommendations for sulfate resistance, ACI 201.2R-77


A ressiveness
Mild Moderate Severe Very severe
Water SO4 (ppm_) 0 - 150 150 - 1500 1500 - 10 000 > 10 000
Soil SO4 (%) 0.00 - 0.10 0.10 - 0.20 0.20 - 2.00 >2.00
w/c, max. 0.50 0.45 0.45
Cement type II, IP(MS), IS(MS) V V + Pozzuolana or Slag

ARRB Transport Research Ltd


12
Review Report 5

Table 5 Recommendations for concrete exposed to sulfate attack


Sulfate content (as SO3) Minimum Cement Content
Class Soil Ground Cement (kg/m3) w/c
Total 2:1 water/soil water Type Aggregate size, max (mm)
(%) extract (g/1) (ppm) 40 20 10
-
1 < 0.2 - < 300 OPC, P-BLF 240 280 330 0.55
2 0.2 - 0.5 - 300 - 1200 OPC 290 330 380 0.50
P-BLF 240 280 330 0.55
SR or supersulfated 270 310 360 0.50
3 0.5 - 1.0 1.9 - 3.1 1200 - 2500 SR or supersulfated 290 330 380 0.50
4 1.0 - 2.0 3.1 - 5.6 2500 - 5000 SR or supersulfated
5 > 2.0 > 5.6 > 5000 330 370 420 0.45
+ protective coatings

(Source: Taylor 1977).

Table 6 Requirements for well compacted cast-in-situ concrete 140 mm to 450 mm in


thickness exposed on all vertical faces to a permeable sulfate soil/fill, BRE Digest 363
Concentration of sulfate and magnesium Minimum
Class Soil or fill Ground water Cement type cement A content Maximum
By acid By 2:1 water / (see Table 4a) (max. aggregate water/cement
extraction (%) soil extract (g/1) (g/1) size 20 mm) ratio
3
SO4 SO4 Mg SO4 Mg (kg/m )
1 <0.24 <1.2 <0.4 A-L Note 6 0.65
2 > 0.24 classify 1.2-2.3 0.4-1.4 A-G 330 0.50
on basis of H 280 0.55
2:1 extract I-L 300 0.55
3 2.3-3.7 1.4-3.0 H 320 0.50
I-L 340 0.50
4 3.7-6.7 < 1.2 3.0-6.0 < 1.0 H 360 0.45
I-L 380 0.45
3.7-6.7 > 1.2 3.0-6.0 > 1.0 H 360 0.45
5 > 6.7 < 1.2 > 6.0 < 1.0 As for Class 4 plus surface protection
> 6.7 > 1.2 > 6.0 > 1.0
A Cement content includes supplementary cementing materials, such as fly ash and slag. To maintain the
mortar fraction, the cement contents given may be increased by 40 kg/m3 for 10 mm maximum nominal
size aggregate and may be decreased by 30 kg/m3 for 40 mm maximum nominal size aggregate (Table 8
of BS 5328: Part 1).
B Minimum value 275 kg/m3 for unreinforced concrete (BS 8110:1985 and BS 5328:Part 1:1990); 300
kg/m3 (BS 8110) and maximum w/c 0.60 for reinforced concrete.

Table 6a Type of cement stated in Table 6

Code Type or combination Code Type or combination


A Portland cement (BS 12) H Sulfate-resisting portland cements (BS 4027)
B Portland blast furnace cements (BS I High- slag blast furnace cement (BS 4246)
146) containing not less than 74% slag by mass of
C High slag blast furnace cement (BS nucleus
4246) J Combinations of portland cements and blast furnace
D Combinations of portland cements (BS slag containing not less than 70% slag and not more
12) and blast furnace slag (BS 6699) than 90% slag by mass of slag plus cement
E Portland PFA cements (BS 6588) K Portland PFA cements containing not less than 26%
F Combinations of portland cements (BS PFA by mass of nucleus
12) and PFA (BS 3892:Part 1) L Combination of portland cements and PFA
G Pozzolanic PFA-cements (BS 6610: containing not less than 25% PFA and not more than
1991) __ 40% PFA by mass of PFA plus cement

3.2.2. Standards and Test Methods


Because one of the main mechanisms of sulfate attack involves the formation of expansive
compounds from the reaction between sulfate ions and calcium aluminate hydrates, standards

ARRB Transport Research Ltd


13
Review Report 5

have been established that specify the content of C3A in cement used in sulfate-containing
environments must be low. European Standard ENV 206 specifies the need to use a sulfate-
resisting cement in moderate-to-severe sulfate environments, but it leaves the choice of actual
cement type to individual countries. In the UK a sulfate-resisting cement, according to
BS4027, should not have a C3A content greater than 3.5% by mass. In the USA, ASTM C150
specifies that cements used for moderate and high sulfate resistance should not contain more
than 8.0% and 5.0% C3A respectively. Australian Standard AS 1315 specifies a maximum
5.0% C3A for sulfate resisting cement.
A lower level of C3A may be achieved by using more iron rich raw materials so that the
alumina is consumed to form C4AF.* However, this still does not guarantee the behaviour of
the concrete; cements low in C3A but high in C4AF have been known to develop excessive
expansion when the specimens were exposed to sulfate solution (see Gonzales & Irassar
1997).
Some specifications not only restrict C3A to a maximum of 5%, but they also restrict
combinations of cement composition, e.g. C3A + C4AF or C3St + C3A or C2F + C4AF, (see
Jackson and Lawton 1992).
Although limiting cement compositions is sometimes adequate to prevent severe sulfate-
induced deterioration, as addressed in the BS4027 standard, it may not necessarily lead to
satisfactory performance in terms of a concrete structure's durability. Like other
environmentally caused deteriorations, sulfate attack of concrete is first of all a diffusion or
solution permeation process, as pointed out by Mehta et al. (1992): they state that in this
respect 'the coefficient of permeability is more important than modifications in the chemistry
of portland cement.'
Furthermore, a controversy has been noted: although the presence of C3A promotes
sulfate deterioration, it is also beneficial in delaying chloride penetration into concrete, and
the latter is important in protecting the reinforcing steel from corrosion (Gerwick 1990). The
use of pozzolanic blends in cement/concrete has made such composition restrictions less
relevant (their properties are strongly dependent on the quantity and quality of the blending
materials), and as noted in AS 3972-1997, the chemical limitation of C3A < 5% 'penalises
innovative solutions to the problem of sulfate attack.' Clearly, performance-orientated
standard testing methods are needed to evaluate the applicability of cements or cement blends
in sulfate-containing environments.

ASTM C452-95, originally published in 1960, sets up a standard method for testing the
potential soundness of portland cements to sulfate attack by determination of the expansion
of mortar bars (25 x 25 x 285 or 160 mm) made from a mixture of cement and additional
gypsum. The total amount of sulfur trioxide content in the mixture is 7.0% by mass, and the
mortar bars are stored in water at 23°C. ASTM C150 notes that, for sulfate resistance, the
cement (Type V) mortar so tested should not have an expansion of more than 0.040% at 14
days. This method is unsuitable for testing slag blended cement (AS2350.14).

ASTM C265-91, originally published in 1951, aims at testing the water-extractable SO3 in
hardened portland cement mortar to represent the unreacted CaSO4 remaining in the mortar.
Mortar cubes of 24 hours age are crushed to pass through a No.8 sieve (2.36 mm), then
extracted in water for 2 minutes, followed by filtering in less than 2 minutes. The filtrate is
tested for SO3 by the precipitation method (i.e. using BaC12). This sulfate may have
significance in relation to the internal sulfate attack.

* C4AF = 4CaO•Al2O3•Fe2O3
f C3S = 3CaO•SiO2

ARRB Transport Research Ltd


14
Review Report 5

ASTM C1038-95, originally published in 1985, is similar to ASTM C452 excepting that no
additional gypsum is added to the cement. The test applies only to portland cements. It notes
that an expansion limit of 0.020% in 14 days of water immersion is the criterion in Canadian
Standards document CAN 3-A5-M83.

ASTM C1012-95, originally published in 1984, specifies testing the length change of
specimens stored in a sulfate solution (5.0% by Na2SO4). This test is for assessment of sulfate
resistance of concrete and mortars made using portland cement, blends of portland cement
with pozzolans or slags, and blended portland cements. The current version of this standard,
ASTM C1012-95, is based on research carried out since the 1940s (Rosner et al. 1982;
Mather 1957, 1978; Biczok 1967; Miller & Snyder 1945; Wolochow 1952a and b; Mehta
1975; Polivka & Brown 1958; Mehta & Polivka 1975; Regourd 1975; Mehta & Haynes 1975;
Campus 1963; Bakker 1981; Miller & Manson 1940; Lea 1970; Brown 1981; Patzias 1987,
1991). The test is performed on specimens with a moderate compressive strength (20 MPa at
7 days) stored in a very aggressive sulfate solution kept at pH value of 6-8. The details of
concrete behaviour in controlled test conditions are discussed in the next section.
There are no acceptance criteria specified in the 1995 version, though in the earlier
version of this standard an expansion of 0.10% at 6 months indicated failure in the test. In
ASTM C 1157M-97 the performance specification for blended cements, the requirement for
moderate sulfate resistance (MS) is that the maximum expansion at 6 months is 0.10%; and
for high sulfate resistance (HS) expansion at 6 months and 1 year should not exceed 0.05%
and 0.10% respectively.

Australian Standard AS 2350.14-1996, was developed to evaluate the potential use of


cements in a sulfate-contaminated environment. It modified the ASTM C452 and C1012
methods by employing ISO/CEM prisms (40 x 40 x 160 mm) to perform the test, and it was
also used develop a more rapid testing procedure.
Test results demonstrated that, compared to the ASTM prisms (25 x 25 x 160 mm) or
the German prisms (10 x 40 x 160 mm), the ISO/CEM prisms showed very small expansion.
This is more likely to result in large errors and poor reproducibility of the test. In evaluating
the effect of pre-curing on experimental results, tests were made on the German prisms (using
five different cements) by employing three water curing schemes: (a) cured to compressive
strength 20 MPa, (b) 3-day curing, and (c) 7-day curing. The three curing regimes did not
give different orders of expansion. The test results also showed that after 16 weeks of
immersion in a 5% Na2SO4 solution it was possible to distinguish between the different
cement types.
Based on the above conclusions, AS 2350.14-1996 specifies using a modified ISO/CEM
prism (15 x 40 x 160 mm) for the test. A 7-day water curing regime is recommended as being
the most convenient and practical. Further work was carried out by a number of different
laboratories, using samples already made with different types of cement (including some
cements of internationally known performance) and distributed to the participating
laboratories to evaluate the new procedure. The method was found to correctly characterise
the resistance of various cements to sulfate attack.
The criterion for sulfate-resistant cement, Type SR, according to AS 3972-1997 is that
maximum expansion at 16 weeks should not exceed 900 prn/m (0.09%). The reproducibility
of this test was found to be up to 300 micro-strain for specimens after 16 weeks exposure.
Thus, it is noted that the target expansion should be set to 600 micro-strain, so that no single
result of the test should exceed the maximum limit set in the standard.

ARRB Transport Research Ltd


15
Review Report 5

Other Standard Codes are very similar in their testing procedures and conditions, however
diverse they may be in their assessments and criteria for failure. For example, the Chinese
standards GB 2420-81 and GB 749-65 evaluate sulfate resistance from the relative flexural
strength of mortar prisms (10 mm square section) immersed in a sulfate solution (3% Na2SO4
or else the same as a real-life environment) and identical mortar prisms stored in plain water.
The pH value of the sulfate solution is maintained at 7.0 by titrating with 1 N H2SO4 for GB
2420. Exposure periods of 28 days and up to 6 months are specified by the respective
Standards. It should be noted that because GB 2420 was designed as an accelerated test to
evaluate different cements, it was recommended that this test should be run in parallel to GB
749 test, to give more reliable results. GB749 states that after exposure to sulfate solution,
mortar bars with a relative flexural strength of 0.80 or less may be expected to have poor
sulfate resistance.
In research, various methods have been used by laboratories; examples are the lean
mortar bar test by Wolochow (1952b), and the test methods summarised by Mehta (1993) in a
review and by Frearson and Higgins (1995). A brief summary of test methods is shown in
Table 7. It appears that a main difference in the various codes is the specimen cross-section,
which varies from 25 x 25 mm in the ASTM methods to 10 x 10 mm in the GB methods.
Although most standards use length change of the mortar prism to assess the influence of
sulfate on cement, the Appendix of ASTM C 1012 notes that other methods used or proposed
include changes in: (a) pulse velocity (ASTM C 597); (b) resonant frequency (ASTM C 215);
(c) compressive strength (ASTM C 109 and C 349); (d) flexural strength (ASTM C 348); (e)
mass; and (f) hardness. Kalousek et al. (1972) mentioned a criterion for concrete failure in
sulfate solution set by the Bureau of Reclamation (USA) in the 1970s was 0.5% expansion
(or 40% loss in elastic modulus (3 x 3 x 16.5 inch prisms)).
When strength change is used to assess the sulfate resistance of concrete large amounts
of specimens are necessary, especially for prolonged tests (see Frearson and Higgins 1995).
Length change measurement is the easiest non-destructive test to perform, and has been
adopted in most of the standard methods.
Because specimen sizes for strength tests using the ASTM standard methods are
sometimes too large to undergo a rapid deterioration in the case of moderate-strength
concrete, these methods have been modified in some laboratories. Depending on the need to
accelerate a test or to simulate specific requirements (as noted in ASTM C1012), the test can
be run with modifications to the test conditions. In research, binary or ternary sulfate
solutions with very high concentrations have often been adopted, to enhance the sulfate
attack as well as to simulate real-life conditions.
In recent years, more attention has been paid to delayed ettringite formation resulting
from curing at elevated temperatures. Tests performed involve steam curing of specimens and
monitoring length change. Usually microstructural and mineralogical examinations are used
to provide evidence for ettringite or gypsum formation. Although there is no standard test
method specified for testing delayed ettringite formation, procedures similar to the testing of
alkali-aggregate reaction (AAR) have been used.
Klemm and Miller(1997) summarised some of the test methods used by researchers,
including the procedure of the 'Duggan test'. In this test, concrete cores of 22 mm diameter
and 50 mm length, are water cured for three days, then they are subjected to three cycles of
heating and cooling over a period of seven days. The cores are then placed in plain water and
their lengths measured. Expansion of 0.05% or more after 30 days might indicate potential
for deleterious expansion.
The testing of the 'slowly soluble sulfate' in cement clinker was discussed by Klemm
and Miller (1997). They reasoned that a dissolution procedure involving large volumes of
water, such as the ASTM C114 procedure for water soluble alkalis, would render the highly

ARRB Transport Research Ltd


16
Review Report 5

soluble sulfates as slowly soluble sulfates; whereas the quick extraction procedure of ASTM
C265, using low-volumes of water, would not result in the dissolution of any slowly soluble
sulfate. They used selective dissolution techniques to separate cement clinker, and then
determined the sulfate-bearing phases by X-ray diffraction (XRD), i.e. salicylic acid—
methanol solution treatment to dissolve calcium silicates, or KOH—sucrose water solution
treatment to dissolve calcium aluminates, so that the sulfates presenting in the various
clinkers could be separated.

Table 7 Summary of Test Methods


Standard Specimens Initial Curing Subsequent Property
or reference Composition Size (mm) Storage condition determined
ASTM C265 cement:sand 50 mm cube sealed in mould 400g crushed mortar SO3 in the filtrate
(c:s)= 1:2.75 and placed in water mixed with 100g as un-reacted
w/c = 0.50, bath for 24h at water and digested CaSO4 in the
using 23±0.15° C, for 2 min, then filter mortar
distilled temperature of
water fresh mix: 22.4-
23.6°C
ASTM C452 cement A ' U : 25 x 25 x 285 moist closet for 22 in fresh water at Length change
sand = (nominal gage - 23 h; followed by 23°C ± 1.7°C; compared to
1:2.75 length 250) in water for 30 replenish the water nominal gage
w/c = 0.485 or for routine min, at 23°C ± every 7 days length. Measure
or 0.460 for test: 1.7°C after 14 days
air entraining 25 x 25 x 160
cements
ASTM ibid without ibid ibid in lime saturated ibid.
C1038 adding water, 23°C±1.7°C
gypsum for 14 days
ASTM ibid u ibid u seal in mould which in 0.352 M (5.0%)_ ibid. Measuring
C1012 is immersed in Na2SO4 solution r at intervals varying
water at 35°C ± 23°C ± 1.7°C, pH = according to
3°C, 231/2h ± 30 6.0 - 8.0. Volume of length change
min; followed with solution to mortar = rate. Review
in lime water at 4:1 data at 15 week
23°C ±1.7°CE
AS w/c = 0.500 15 x 40 x 160 2-day in sealed ibid ibid. Up to 16-
2350.14 (nominal gage mould and 5-day in week continual
length 130) lime water, monitoring.
23°C±2°C
Koch (1960) c:s = 1:3 10 x 10 x 80 1-day in mould, in 0.704 M (10%) flexural strength
w/c = 0.60 then in water at Na2 SO4 solution at
20°C till 21-day 20°C
Locher ibid ibid ibid in 0.310 M (4.4%) ibid
(1956) Na2SO4 solution at
20°C
Smolczyk ibid 10 x 40 x 160 1-day in mould, ibid expansion
(1972) then in water at
20°C till 14-day
German c:s = 1:3 ibid ibid in 0.1666 M (2.366 ibid
(Frearson & w/c = 0.50 %) Na2SO4 solution
Higgins at 20°C
1995)
Frearson & c:s w/c ibid ibid in 0.310 M (4.4%) ibid
Higgins 1:3 0.60 Na2SO4 solution at
(1995) 1:2.3 0.50 20°C
1:3 0.45
Wolochow c:s=1:4 25 x 25 x 285 2-day in mould at in 0.352 M (5.0%) ibid
(1952a) standard 23°C, then in water Na2SO4 solution at
consistency at 23°C till 7-day 23°C
Kollek & c:s = 1:4 ibid 1-day in mould, various: 0.042, ibid
Lumley w/c = 0.60 then in water at 0.084, 0.169, 0.296
(1990) 20°C till 7-day (0.6, 1.2, 2.4, 4.2%),
at 20°C
French c:s = 1:3 20 x 20 x 160 1-day in mould, in 0.1666 M (2.366 ibid
(Frearson & w/c = 0.50 then in water at `)/0) Na2SO4 solution
Higgins 20°C till 28-day at 20°C
1995)

ARRB Transport Research Ltd


17
Review Report 5

Standard Specimens Initial Curing Subsequent Property


or reference Composition Size (mm) Storage condition determined
GB749 C: S = 1: 10 x 10 x 30 1 day in covered sulfate solution Relative flexural
G
3.50; mould at 20°C, (concentration not strength after 6
standard then 14 days water specified) 20 ± 5°C months
consistency _ curing
GB2420 C:S=1: 10 x 10 x 60 24h ± 2h in mould in 3.0% Na2SO4 Relative flexural
G
2.50 at 20 ± 3°C, then in solution, 20°C. strength after
w/c = 0.50 water at 50°C for 7 Volume of solution 28 days
d to mortar > 30 : 1
A Add gypsum so that the total sulfur triox'de (SO3) in the cementing material is 7.0% by mass.
B Including admixtures, such as gypsum.

c For blends of portland cement with a pozzolan or slag, w/c shall develop a flow within ±5 of that of
portland cement at w/c of 0.485.
D Test specimen for concrete: 100-mm square cross section if the maximum aggregate size = 50.0 mm,
or 75-mm square cross section if the maximum aggregate size = 25.0 mm.
E Until the compressive strength of the cube samples (made of the same mortar or concrete, stored at
same conditions) reaches 20 MPa.
F Other sulfate solutions can also be used. It is noted that a solution containing 0.176 M Na2SO4 and
0.176 M MgSO4 is shown to adversely affect the blended cement containing slag, due to magnesium
ion attack other than sulfate attack (see Biczok 1967; Miller & Snyder 1945).
Flexural strength of the mortar prisms exposed to sulfate solution compared to that store in water.

3.3. Research on Test Methods


Many research projects have been undertaken in the past to clarify the mechanisms of sulfate
attcck, and to develop reliable test methods for this purpose. The standard methods
mentioned above have been developed on the basis of extensive studies of factors that
influence the test results, or more specifically the rate of deterioration of concrete under
controlled test conditions. Test methods were reviewed by Locher (1956), Wittekindt (1960),
Calleja (1980), Mehta (1980 & 1993), Osborne (1989), Lawrence and Brown (1981).
Frearson and Higgins (1995) commented on the relative significance of various factors in
testing procedures.

3.3.1. Cement composition requirements


Cement composition is easily determined, by either chemical analysis (ASTM C114;
BS4550), or X-ray diffraction (XRD). Therefore, limitations on cement composition have
been used as the initial criteria for the safe service of concrete structures in sulfate
environments.
Gollop and Taylor (1994) made a microanalytical study of a sulfate-resisting portland
cement (SRPC) with ferrite as the only aluminium-baring phase. Its approximate composition
was Ca2A10.8Fea8Mg0, 2Sio.205, and the A1203 content of the clinker was 3.8%. The main
hydration products were CSH and Ca(OH)2. Although the cement contained no C3A and the
iron-rich ferrite was of low reactivity, ettringite was formed initially and persisted for at least
a year. The latter was attributed to the relatively high S03/A1,03 ratio in the cement. It was
speculated that persistent ettringite in the system could contribute to the resistance to sulfate
attack, as the A1203 in the ettringite could not react with sulfate from external sources to form
more ettringite. Although there may be alumina present in the CSH and other phases which
are vulnerable to sulfate attack, the low total A1203 content of the SRPC greatly reduces the
amount available for the reactions.
However, questions about the actual behaviour of concrete made with cement meeting
the requirement still remain. Even the sulfate-resisting cement is not resistant to all types of
sulfate attack, and further tests are necessary. Lawrence (1990) discussed various aspects of

ARRB Transport Research Ltd


18
Review Report 5

sulfate attack as well as test methods and codes. The results of his tests on mortar prisms
exposed to a 5% Na2SO4 solution demonstrate a close relationship between mortar bar
expansion and the C3A content (Figure 3); these data have been used for ASTM limits for
portland cements.
While this relationship is largely true for Portland cements (not entirely, as that portion
of the C3A forming primary ettringite at early hydration would not respond to sulfate attack,
thus the critical C3A content is also related to SO3 content in the cement — see Jackson &
Lawton 1992), limiting C3A content is less reliable for other cements such as slag cements
and fly ash cements, which have a poorly defined relationship between expansion and the
C3A content. Nevertheless, the content of C3A (< 5%) or total amount of calcium aluminates
(about 20%) in cement is a primary guideline for the performance of concrete in a sulfate
environment.

• Portland cement
2000
Best fit line, r =0.970 + Slag mixes
/(portland cements only)
1000
800
600 Proposed limit for
•+
400 type V equivalence

- Proposed limit for


200 type II equivalence

100 • •
80 •
60
40

20 AST M type ll ASTM type I

10
0 2 4 6 8 10 12 14
Bogue C3A, 0/0
Figure 3. Time for ASTM mortar prisms to expand 0.1% in a 5% Na2SO4 solution.
(Source: Lawrence 1990)
(ASTM Type I = portland cement for general use; Type II = moderate sulfate-resistant portland cement;
Type V = high sulfate-resistant portland cement).

Kalousek et al. (1972) summarised the work done at the Bureau of Reclamation so as to
predict concrete service life in sulfate-containing environments. Prisms of 3 x 3 x 161/4 inches
were stored in a 2.1% Na2SO4 solution or water. A wet—dry process was used: 16 hours in
solution (or water) followed by 8 hours in 50% RH at 23°C. The cements used include Type
V cements, and blended cements containing pozzolans such as fly ash and pumice, etc. By
plotting the test results against exposure time, and extrapolating to the criteria set by the
Bureau (0.5% expansion, or 40% loss in elastic modulus), they postulated that the Type V
cements tested would have a service life of less than 50 years, whereas the effects of
pozzolanic materials in cement ranged from deleterious to very beneficial. They pointed out
that limitings on C3A and C4AF content did not necessarily prevent or reduce sulfate attack.
Other research on low C3A cements indicated that C3S content also affects cement's
resistance to sulfate. Domic and Droljc (1986) thus showed that reaction between Ca(OH)2
and sulfate ions leads to gypsum-induced expansion. This fact has been taken into account by
some Standard specifications to limit the content of C3S, the major source of Ca(OH)2 (e.g.

ARRB Transport Research Ltd


19
Review Report 5

C3S + C3A 5_ 58% for moderately sulfate-resisting portland cement (see Jackson & Lawton
1992).

3.3.2. Tests for internal sulfate attack

One type of test represented by the ASTM C452 and C1038 Standards involves the addition
of extra gypsum to the specimens and observation of their length change after a certain
duration. The specimens are cured at normal temperature. The other type of internal sulfate
attack is delayed ettringite formation, which is essentially an 'early material history'-induced
problem. Testing for delayed ettringite formation usually involves steam curing the
specimens at elevated temperatures (Klemm &Miller 1997), though there is not yet any
standard test method specified. Mielenz et al. (1995) reported the presence of ettringite in
distressed railway ties made of concrete. The ettringite occurred irrespective of the curing
temperature. These authors suggested that the clinker contained a large amount of low
solubility sulfate (possibly in the form of anhydrite), which gradually went into pore solution
to produce ettringite. The last case seems to be difficult to test by either of the ASTM C452
or C1038 test methods.

Mehta et al. (1979) studied the long-term exposure of concrete (34 MPa 28-day strength)
specimens to a 5% Na2SO4 + 5% MgSO4 solution. Although the cement contained no C3A,
and its expansion by ASTM C452 test was very low even after 75 days, the specimens
developed severe surface spalling after 6 years of exposure. XRD analysis showed huge
amounts of gypsum in the deteriorated zone. Compared with the reference specimens stored
in water, the sulfate-attacked specimens had a 50% strength reduction. It was concluded that
the ASTM C452 method (water immersion of gypsum-enriched mortar bars) was not
appropriate for predicting sulfate attack involving surface spalling and loss of cohesion
resulting from gypsum formation after long exposure to sulfate solutions.

Using the ASTM C1038 method, Fu et al. (1997) reported that specimens which
satisfied the criterion of less than 0.02% expansion after 14 days, had shown unacceptable
expansion after 5 years. Moreover, larger specimens (3 x 3 x 10 inches) had a higher
expansion than smaller specimens (1 x 1 x 6 inches) after 80 days exposure. This might be in
part attributable to the 'delayed ettringite formation', which occurs at later ages.

In spite of such findings, ASTM C452 provides a rapid means of eliminating unsuitable
cementing materials in terms sulfate resistance. The similar ASTM C1038 method screens
only extremely unsuitable materials; but the short exposure of the test is not representative of
the extensive exposure to sulfates that is encountered in practice.

Efforts have been made to clarify the role of delayed ettringite formation in concrete
deterioration. The sources of sulphate involved in this formation have been stated to be:

(a) Sulphate incorporated into the lattice of calcium silicates; mainly C2S which may
contain 2.5 - 3.4% SO3 (Taylor 1996). This sulfate will go into solution only with the
hydration of the clinkers.

(b) Unhydrous CaSO4 which has a dense structure and very low solubility. This will
dissolve very slowly to provide sulfate ions in the pore solution of concrete.

(c) Sulfate arising from high temperature curing of concrete at early ages. Primary
ettringite, which forms at very early stages of hydration, decomposes at high
temperature to form Afm, leaving a considerable amount of sulfate ions in the pore
solution of concrete.

These forms of sulfate will become available over extended periods of time to react with the
aluminate phases in concrete to cause expansion and cracking.

ARRB Transport Research Ltd


20
Review Report 5

Joins (1996) used dilute sodium carbonate solution to treat cement clinker, and measured
the 'insoluble' sulfate content. Klemm and Miller (1997) showed that the sulfur-bearing
phases in different clinker minerals can be determined by XRD in combination with
treatment of the cement by salicylic acid—methanol or potassium hydroxide—sucrose water
solution. The sulfur-bearing phases in the KOH-sucrose water solution treatment residue
were expected to represent the slowly soluble sulfate in calcium silicates. However, the same
authors showed that the 'slowly soluble sulfate' seemed to form ettringite before the concrete
hardened; thus their contribution to the late ettringite was doubtful.
High temperature curing at the early age of concrete, on the other hand, is recognised as
the main factor for delayed ettringite formation that causes expansion and even cracking of
concrete (Diamond 1996; Glasser 1996; Stark & Bollmann 1997). Several studies of delayed
ettringite formation or alkali-aggregate reaction (AAR) involving elevated temperature
treatment of concrete and subsequent exposure to wet or humid conditions, have
demonstrated the presence of large quantities of ettringite in deteriorated concrete (Scrivener
& Lewis 1997; Stark et al. 1992; Michaud et al. 1997; Kelham; Camarini; Cincotto; Meland
et al.: Bournazel et al. 1996). These prove the effect of high temperature treatment on the
stability of ettringite, and on the formation of ettringite long after concrete hardening.
However, high temperatures may not always generate deleteriorous effect, which appears to
be true for the concrete containing AAR active aggregate.
Shayan and Quick (1992) discussed internal microcracking of precast concrete railway
sleepers reported in several countries. They believe that the conclusions which ascribe
cracking to delayed ettringite formation and the like were not convincing. They investigated
the concrete sleepers in question, and concluded that the internal cracking was due to AAR,
as well as resulting from the differential expansion of the interior and exterior of the sleepers,
whereas ettringite only precipitated in already existing cracks.
However, for concrete made with sound aggregates, delayed ettringite formation may
still cause cracking, and tests have been conducted for its investigation. A procedure used by
Scrivener and Lewis (1997) is representative of this type of test. The cement mortar, made
with w/c* = 0.50 and sand/cement = 3, was cast into 25 mm cubes and 16 x 16 x 160 mm
prisms. The mortars were pre-cured for 4 hours, then steam cured at 20°, 80° and 90°C for 12
hours (They were heated to the required temperature over a period of 1 hour). The specimens
were cooled to room temperature, demoulded and placed in water at room temperature. The
expansion of the prisms was measured weekly in the first month and monthly thereafter. The
mortar samples were then examined by XRD, backscattered electron (BSE) imaging and
energy dispersive X-ray analysis (EDX). They found in the heat cured pastes there was
virtually no evidence of ettringite after heat treatment. Over the subsequent periods, the
formation of ettringite could be clearly observed. They also suggested the possibility that
ettringite formed within CSH gel and caused expansion.
Stark and Bullmann (1997) considered that both the drying—wetting and the high
temperature contribute to the damage induced by delayed ettringite formation. They carried
out tests to simulate climatic conditions experienced by concrete pavements, e.g. sunshine,
day—night temperature variations, rain and frost. Three processes involving drying—wetting
cycles were used:
(1) 2 weeks drying at 20°C, 65% RH followed by 8 weeks water immersion at 20°C;
(2) 2 weeks: 24 hours drying at 60°C, 0% RH followed by 24 hours at 20°C, <40% RH;
then 8 weeks storing in water at 20°C; 24 hours 20°C to —20°C (in NaC1 solution)
freeze—thaw cycle;

`w/c' is short for 'water to cement ratio'.

ARRB Transport Research Ltd


21
Review Report 5

(3) 2 weeks drying at 60°C, 0% RH followed with 8 weeks water immersion at 20°C, then
the 24 hours freeze-thaw cycle.
The samples were 100 x 100 x 400 mm prisms taken from a concrete pavement within 2
weeks of placing. The physical properties tested included mass, water absorption, ultrasonic
pulse velocity and modulus of elasticity. Microstructural examinations were also made. It was
reported that at least three cycles of treatment were necessary to cause obvious damage, and
that the third process above produced the largest expansion. These authors showed that the
primary ettringite in concrete may dissolve in the pore solution and recrystallise in the large
available spaces of the pores and cracks, which can be enhanced by the heating up to 60°C. In
addition, the high temperature drying created microcracks. It seems that the freeze-thaw
cycle was the major factor in the destruction of the concrete tested by these processes, while
the ettringite contributed to filling air voids, thus reducing the concretes' resistance to the
frost damage. Similar procedures without freeze-thaw used by Soroka and Carmel (1987)
showed spalling of the mortars after some one and a half years of testing; the spalling was
said to result from salt crystallisation rather than chemical attack. In these cases delayed
ettringite formation had a secondary role in the concrete damage. This observation agrees
with the conclusions of Shayan and Quick (1991/92) and Shayan (1995).

3.3.3. Performance-orientated tests


ASTM C1012 (a sulfate solution immersion test) and similar tests have been established so
that conditions are more similar to those experienced by concrete in service, i.e. in real life.
Although it is difficult to compare methods because the mechanisms involved may be
different, it appears that this particular method has been increasingly used in research. The
details of tests used in research laboratories vary with respect to sulfate solution (e.g. pH,
sulfate concentration and type), exposure duration, temperature, specimen size, deterioration
assessment, etc.

3.3.3.1. The pH of the sulfate solution


Brown (1981) reported a comprehensive study of the testing of sulfate resistance under
various controlled conditions, as well as a review of previous work. The earliest test as such
was the Merriman Slab Test (1993). It used paste specimens of size 2 x 4 x 0.25 inches,
stored in a 10% Na2SO4 solution maintained at neutral pH by daily titration with sulfuric acid
(H2SO4); visual estimation was made at 28 days. The same test procedure was adopted later
by Mehta and Gjory (1993); they used automatic pH control devices to measure the pH and
they introduced 1 N sulfuric acid to maintain the solution at near neutral pH.
Brown experimented on mortar prisms made with Type I cements, stored in Na2SO4
solutions maintained at various pH conditions, i.e. pH = 11.5, 10.0, 8.9 and 6.0, which were
compared with their counterparts in solutions of uncontrolled pH and those in plain water.
Consumption of sulfate ions, expansion, and compressive strength were all monitored.
Figures 4-8 show the main results of this work. It was concluded that control of pH by the
addition of a 1 N H2SO4 :
(1) ensured a constant concentration of sulfate ions, which would otherwise vary,
depending on the cement type and thus make any comparison difficult (as well as
delaying the deterioration process);
(2) maintained a constant pH condition which is more likely to occur in real environment.
It was observed that the sulfate ions consumption rate was twice as high as the liberation of
hydroxyl ions, i.e.:

[SO4 it = [ SO4 1 - -11[0H -11

ARRB Transport Research Ltd


22
Review Report 5

where [S042-], and [OH-], are the respective concentrations of S042- and 0I-F at time t. This
was consistent with the stoichiometry of the formation of gypsum or sulphoaluminate. The
consumption of sulfate ions increased with decreasing pH, indicating that there was leaching
of the hydration products under these conditions. However, after the initiation of expansion,
it appeared to be a linear function of sulfate ion consumption (Figure 7) indicating that more
of the reaction products were forming with increasing expansion. Interestingly, the expansion
per sulfate consumption is highest for the solution with pH = 11.5, the condition under which
ettringite is more stable, than under lower pH conditions. Apparently this proves that the
expansion is mainly due to the formation of ettringite.
Ion Concentration in Solution, M x 100

pH=12.7

10% C3A cement


Hydroxyl

S:C:W=2.75:1:0.5
Volume of bars to volume of 0.2 M
Na2SO4 = 3 to 1
pH=6.12
0 7 14 21 28
Time, Days
Figure 4. The rate of hydroxyl ion liberation. (Source: Brown 1981)

0.201
Sulfate Ion Concentration in Solution, M

0.197

0.193

0.189

0.185 -

0.181 -

0.177 -

0.173
0 7 14 21 28
Time, Days
Figure 5. Rate of sulfate ion consumption, 10% C3A cement mortar bar in 0.2 M
Na2SO4 . (Source: Brown 1981)

ARRB Transport Research Ltd


23
Review Report 5

0.7
Figure 6.
Rates of mortar bar pH=10.0
0.6 pH=6.0
expansion under 0.35 M Na2SO4
different experimental
0 0.5
pH conditions.
pH=11.5
(Source: Brown a 0.4 pH not controlled
1981). x
w
a) 0.3

m
a_ 0.2

0.1
Distilled water
0
0 2 6 8 10 12 14
Time, Weeks

Figure 7. 0.7
Relation between mortar
bar expansion and the 0.6
sulfate ion consumption.
(Source: Brown 1981) 0.5
Percent Expansion

0.4

0.3

0.2

0.1

0
0 10 20 30
rvillimoles of Sulfate Consumed / dm2 of Sample Surface

140
Figure 8. Distilled water
Variation in cube strength
with time in different 120
experimental pH conditions
Percent ofInitial Strength

in 0.35 M Na2SO4.
pH not controlled
compared with the distilled
100
water immersion. (Source:
Brown 1981)
pH=11.5

80 pH=10.0

pH=6.0

60

I I I I [
2 4 6 8 10
Time, Weeks

ARRB Transport Research Ltd


24
Review Report 5

3.3.3.2. Specimens

Specimen composition, shape and size have varied according to the purpose of the
assessment of mortar and the acceleration needed for the test. The early research of Locher
(1956) and Koch and Steinegger (1960) used small prisms of square section (10 x 10 x 80
mm), for which flexural strength was measured. Later, length change was used as the main
assessment for the test, and the specimen prisms were longer, e.g. 10 x 40 x 160 mm and 25
x 25 x 285 mm in Europe and the USA respectively.
The composition of the mortar has been similar for other types of standard test; e.g.
ASTM C1202 specified a w/c of 0.485, and sand to cement ratio of 2.75:1, which is the same
as that specified in the ASTM C109 test for strength. Higher water - to-cement ratios (w/c =
0.60, 0.50) and lower cement content (sand/cement = 3 or 4) were used in other laboratories.
To compare the effects of various factors on the test results, Frearson and Higgins
(1995) carried out a series of tests. They varied the aggregate to cement ratio, specimen
shape, initial curing period, specimen compaction, initial curing deficiencies, early
carbonation, concentration of sulfate solution, and type of sulfate solution to see how these
parameters affected expansion. They observed the following:
Aggregate to cement ratio: when the w/c ratio was the same, specimens richer in cement
had a greater expansion at one year. After three years, this trend persisted for only some of
the mortars tested. It appears that these results are subject to two opposing effects of cement
content: (a) the richer the cement mortars the greater the potential for expansion as well as
the greater number of capillary pores per unit volume of the mortar, and (b) the richer the
mortar the better the workability and thus increased compaction.
Specimen shape: Comparison of prisms of section size 10 x 40 mm with 20 x 20 mm
showed that the flat-sectioned mortars (the former prisms) produced an earlier and greater
expansion than their square-sectioned counterparts (the latter prisms).
Laboratory test specimens were made using standard sand of specified grading, e.g. as
specified by DIN 1164 and ASTM C778. In some cases, the sulfate resistance test was used
to test mortar that had been separated from concrete by sieving (Frearson & Higgins 1995;
Fidjestol & Frearson 1994).

3.3.3.3. Pre-test curing


The ASTM C1012 standard specifies a curing scheme: the test starts when the water-cured
specimens reach 20 MPa in compressive strength. As an accelerated test to compare different
cements, this method is designed so that the deterioration of the mortars occurs after a
comparatively short period of exposure to sulfate solution. To shorten the pre-curing time, the
mortar prisms are cured in warm water at 35°C for the first day, which, according to Mehta
(1980), is equivalent in effect to 7 days curing at 23°C in water.
Many researchers prefer the specimens be cured in water at room temperature; e.g.
German test method described by Smolczyk and Blunk (1970) cures the mortar prisms (sand
/cement = 3, w/c = 0.60) of 10 x 40 x 160 mm in water (20°C) for 14 days. Frearson and
Higgins (1995) modified this method by using a lower w/c so that the test samples would be
more representative of the field concrete. The other conditions are similar to the ASTM
C1012 method, except that the sulfate solution temperature is 20°C.
Frearson and Higgins (1995) showed that varying the initial water curing period from 1
to 4 weeks had greater influence on the early expansion of the mortars tested, and that this
effect was much less significant for specimens exposed to sulfate solution for 3 years. On the
other hand, they demonstrated that specimens compacted by vibration had reduced entrapped
air contents by 1%, 1.5% and 3.5% for mortars with sand to cement ratios 3, 4, and 4.7

ARRB Transport Research Ltd


25
Review Report 5

respectively. This resulted in significant differences in expansion under sulfate attack. In


testing the sensitivity of the test procedure to accidental drying, they observed that early age
drying (accidental exposure to air for 3 and 17 hours during initial curing period) led to lower
expansion of the mortars (made with SRPC) after six months and one year of sulfate attack.
Another significant factor was drying with carbonation, which seemed to protect the mortar
from the attack for the period (one year). The same result was also reported by Osborne
(1989).
In establishing the AS 2350.14 standard method, tests carried out by cement industry
laboratories in Australia indicated that 7-day water curing at room temperature might also
result in a suitable mortar maturity for the test, and it was easier to perform, compared to that
set by the ASTM C1012. Bucea et al (1997) carried out tests to compare the two methods and
found that the porosity of the mortars cured by the ASTM scheme was 18.04% with a
standard deviation of 0.34%, whereas the AS curing resulted in porosity of 19.48% with a
standard deviation of 1.62%. A narrower deviation in mortar porosity was demonstrated to be
beneficial in distinguishing the sulfate resistance of various binders, and in this respect the
ASTM procedure was preferred.
On the other hand, equal maturity as defined by strength, is more applicable to blended
cements, since concrete/mortar made with such cements has variable maturity development
depending on the curing temperature and time, compared to that made with portland cements.

3.3.3.4. Duration of sulfate exposure and assessment of failure


It is not possible to set the acceptance criterion without specifying the test duration. Earlier
research used a relatively short test periods, e.g. Koch and Steineger monitored the flexural
strength of mortar bars for up to 77 days immersion in a 0.2 M Na2SO4 solution; Markestad
observed the compressive strength of concrete immersed for up to 70 days in a 0.6 M MgSO4
solution; and Merriman tested paste immersed in a pH regulated 0.7 M Na2SO4 solution and
made visual estimations at 28 days (see Mehta 1993). More recent test results, e.g. Hogan and
Meusel (1983) and Bucea et al (1997), demonstrated that expansion at 10 weeks was still at
an stage, and was not sufficient to classify the cements under test. Gonzalez and Irassar
(1997) showed that mortars made with low C3A (< 1%) but high C4AF (15%) cements
developed expansion after 180 to 720 days when exposed to 5% Na2SO4 solution.
As shown in Figure 3, different cements have different expansion rates under sulfate
attack; e.g. in 5% Na2SO4 solution, the times for 0.10% expansion (previous ASTM C1012
criterion) are about 100, 200 and 360 days for Types I, H and V cements respectively. This
type of trend is complicated by the inclusion of pozzolanic materials such as fly ash, blast
furnace slag or silica fume in the cementing material. In such cases, expansion occurs much
later (Frearson & Higgins 1995; Douglas et al 1991; Dicic & Miletic 1986; Mather). Clearly,
it may be more realistic to vary the acceptance levels for different cementing systems.
Australian standard AS 2350.14 specifies a maximum expansion of 900 µm/m at 16
weeks for sulfate-resisting cement and the equivalent. This has been criticised as
inappropriate (Bucea et al 1997), based on the fact that the C3A content in cements varies,
and that pronounced expansion (tested using either ASTM C1012 or AS 2350.14 methods)
occurs only after 20 - 30 weeks of exposure for most cements with moderate C3A content (see
Figure 9). A long exposure period (6 months or more) was recommended for distinguishing
the behaviours of different cementing materials (Cao 1997; Butler).
Assessments other than the amount of expansion or change in strength can often provide
useful information or help set more reliable criteria for concrete failure. Although
quantitative X-ray analysis of mortar samples gives more direct evidence of deterioration

ARRB Transport Research Ltd


26
Review Report 5

mechanisms, assessment of mass change, or visual observation of scaling and microcracking,


easily monitor the development of deterioration (Hogan & Meusal 1981).

8000

7000 - Cement C3S C3A C4AF


1 51 6.6 11
6000 - 2 57 4.3 14
3 63 5.3 13
Expa nsion (p m/m:

4 34 4.6 15
5000 -
Cement 1
-- - Cement 2
4000 -
- - -Cement 3
- -9 - - Cement 4
3000 -

2000 -

1000 -

0 10 20 30 40 50
Weeks

Figure 9. Expansion of portland cement mortars in normal 5% Na2 SO4 solution.


(Source: Cao et al. 1997 )

Although expansion is an easily measured parameter, in some cases it does not reliably
reflect the extent of deterioration. Cao et al (1997) showed that high expansion might not
necessarily indicate specimen failure, whereas specimens made of some binders having least
expansion in sulfate solution exhibited early failure by crumbling.
As strength reduction is also the ultimate result of sulfate damage, it has been studied to
provide supplementary data or used as an index for the failure. Sulfate attack in most real
environments is associated with the decomposition of cement hydration products, particularly
in neutral or lower pH conditions. The strength reduction results, in part, from this
decomposition.
Brown (1983) reported that over time concrete strength increased to a maximum then
steadily decreased. The early strength increase could be attributed to continuous hydration
and filling up of pores by compounds formed as a result of the reactions between sulfate and
cement hydration products. Similar observations were reported by other workers (Lawrence
1990; Bucea et al. 1997; Douglas et al. 1991; Samanta & Chattergee 1982). Brown noted that
strength decline occurred earlier than expansion (cf Figure 8), attributable to the leaching and
conversion of cement hydrates such as Ca(OH)2 into gypsum, whereas the expansive reaction
products were merely filling up the pores at the early stages of sulphate attack (Brown 1981).

3.3.3.5. Different types of sulfates and concentrations


Various sulfate solutions have been used - the most common being sodium sulfate and
magnesium sulfate. Because of their high solubility these salts can provide a high S042-
environment. In some tests, the solution contained both sulfates in equimolar quantities
(Mehta 1993; Mehta et al. 1979), or in proportions that simulated the real environment
(Bonen 1997). Ammonium sulfate has occasionally been used in some research.
A solution concentration of 5% (0.35 mole/L for Na2SO4, 0.41 mole/L for MgSO4) has
been most commonly used. In some research work a very high concentration sulfate solution,

ARRB Transport Research Ltd


27
Review Report 5

e.g. 10% Na2SO4 , has been used; this may be close to or exceed the solubility of some of the
sulfate salts, and the deterioration mechanism may be quite different from the chemical attack
which is supposed to simulate. In comparison, 10% MgSO4 solutions can be used without
inducing salt crystallisation effects, and have been used for accelerated tests (Wafa 1994).
Because of the reaction between Mg2+ ions and cement hydration products, it has been
reported that MgSO4 results in the highest expansion of portland cement mortars (Sersale et
a/. 1997; Al-Amoundi et a/. 1994; Ben-Yair 1974; Al-Amoundi et al. 1995). Although
concretes containing pozzolanic materials, such as silica fume, have beneficial effects on the
concretes' resistance to sodium sulfate attack, they are specially vulnerable to MgSO4 attack
(Bonen 1993; Locher 1966). Giergiczny (1997) reported that mortars with 10% silica fume
had a significant strength reduction after one year of storage in a 0.123 mole/L MgSO4
solution, and a substantial amount of gypsum was found in the specimens. The reactions are
represented by Equation (5).
Significantly, the sulfate-resisting portland cement (SRPC) had little effect in resisting
the magnesium sulfate attack, and the beneficial effect of replacing 70% of the cement with
blast-furnace slag did not last long (Frearson & Higgins 1995). Similarly, using silica fume to
partially replace cement has been shown to be detrimental for mortars under magnesium
sulfate attack (Bonen & Cohen a & b; Yeginobali & Dilek 1995). In contrast, concrete made
using calcium aluminate cement is more durable in MgSO4 than in Na2SO4 (Lea 1998). This
partly proves that the decalcification of CSH phases and formation of brucite (Mg(OH)2) are
important degradation processes.
Although there are only limited data available on ammonium sulfate attack, it has been
shown that this salt is more deleterious to concrete than is sodium sulfate (Lea 1998). In
ammonium sulfate solutions, cement containing silica fume performed worse than low-C3A
portland cement (Bate 1984).

3.3.3.6. Influence of temperature


Usually in chemical reactions, increasing the temperature helps to overcome the activation
energy and also increases the rate of reaction. In most cases, dissolution is an endothermic
process (i.e. absorbs heat) and the solubility of salts increases with temperature (exceptions
include Ca(OH)2 and CaSO4.1/21120). The reverse is also true: that a solution may become
oversaturated with decreasing temperature.
Of particular interest in the case of sulfate attack is to the thermal instability of ettringite
at temperatures higher than 60°C (see Figure 2), and the thaumasite increased stability of
temperatures in the range of 0-5°C.
The solubilities of some sulfate salts became substantially higher at higher temperatures,
e.g. Na2SO4.10H20 and MgSO4.7H20, (see Figure 1). It was observed that in testing concrete
immersed in moderately high concentration of the sulfate solution, a sudden crystallisation of
salt occurred when the temperature was lowered (e.g. in 30% Na2SO4, cooling from 30°C to
5°C at a rate of 5°C per hour - see Follard & Sandberg 1994, causing physical damage to the
specimens (St John 1982; Folliard & Sandberg 1994; Novak & Colville 1989).
Thus in laboratory tests it is important to control the temperature and the solution
composition/concentration, so as to simulate the real environment. This is especially
important when high concentration sulfate solutions are used.

3.3.3.7. Influence of loading


In practice, concrete as a structural material is subject to stresses, e.g. tension, torsion, shear,
and compression. Stress affects the microstructure of concrete in terms of cracking and creep,

ARRB Transport Research Ltd


28
Review Report 5

depending on the level of the stress/strength ratio; stress also affects the apparent behaviour
of concrete, e.g. its expansion under sulfate attack.
Schneider and Piasta (1991) reported the results of sulfate attack on concrete under
uniaxially sustained loads for more than 3 years. The time-dependent deformation of concrete
under sulfate attack did not appear to follow the theory of concrete creep at the stress/
strength level of 0.65; the deformation under this loading was accelerated in a 5% Na2SO4
solution, and was associated with reductions in residual strength and development of
microcracking in the specimens tested. At stress/strength ratios of 0.35 and 0.50 (below the
linear creep limit) the length change stopped after the initial deformation (about 2 months)
and the concrete behaved according to the creep theory; for example, when the stress/strength
ratio was lower than 0.20, only swelling was observed.
Care should be taken in the analysis of the results obtained under uniaxial compressive
load because the specimen under compression will simultaneously be subjected to tension on
the side surface near the ends, due to the Poisson Effect. This may partly contribute to the
rapid deterioration of the specimens under a compression load in laboratory. Nevertheless,
tests of the behaviour of the concrete under simultaneous load and sulfate attack are
important in determining both material durability and structure mechanics.

cp/R
3000 —
0.0
2500 —

2000 —

1500 —
0
X 1000 — 0.20

500 -r
- imp;
gove
l
0III a I 0.35
1" = 400 600 800 1000 1200
-500 °- 0.50
ihr 41h

-1000

-1500
0.65
-2000 Time, Days

Figure 10. Length change of mortar bars under compression, stored in a 5%


Na2SO4 solution. (Source: Schneider and Piasta 1991)
(G = applied stress; R = concrete strength).

3.3.3.8. Influence of other reactive agents


Cloride ions: CF ions inhibit the formation of gypsum and portlandite due to the
increased solubility of these compounds. In general, this may have a beneficial effect on
concrete made from portland cement, whereas this benefit is only marginal for the blended
cement concretes (Al-Amoudi 1 994).113 On the other hand, the higher solubility of Ca(OH)2
will cause more lime to leach, which may result in softening of the concrete in the long run,
especially for blended cement concretes.
Al-Amoudi et al. (1994) tested strength (using 25 mm cubes) and made a mineralogical—
microscopical study of cement paste specimens (12.5 mm cubes) stored in solutions with a
high content of CF ions and various sulfate contents (0%, 0.55%, and 2.1% of Na2SO4 and
MgSO4 solutions), at 25°C. The results showed that gypsum and portlandite were not present

ARRB Transport Research Ltd


29
Review Report 5

in mortars in contact with high concentration Cr solutions, whereas the formation of


Friedel's salt (3Ca0.A1203.CaC12.12H20) was not significant. The absence of these phases
may have been related to the experimental conditions that they employed (high temperature
and salt concentration). They observed that blended cements (with silica fume or blast
furnace slag) showed more magnesium sulfate attack when compared to plain cement, which
had gypsum- and ettringite-induced attack.
The opposite conclusion was reported by Ben-Yair (1974), who used 0.4% sulfate
solutions (Na2SO4, K2SO4, MgSO4) containing up to 3.5% NaCl equivalent (NaC1, CaC12 and
MgC12). He monitored the expansion of normal portland cement concrete for long periods
(1.5 and 8 years) and found that. in most cases the expansion in chloride-containing solutions
increased due to the formation of Friedel's compound. Ben-Yair used lower concentration
sulfate solutions than Al-Amoudi; maybe at higher sulfate concentrations, Friedel's salt
formation is inhibited.

CO2 and C032-: Sawicz and Heng (1996) tested concrete containing limestone
powder in a 5% Na2SO4 solution. The cement contained 12.1% C3 4, and expansion
developed fairly quickly; i.e. expansion at 150 days was about 630 pm/m (max. 1105 vim/m).
They found that the presence of limestone powder increased the specimens' resistance to
sulfate. e.g. binder containing 13% limestone with a water to binder ration of w/b=0.55 gave
the lowest expansion. Based on XRD results showing that various carbonate containing
compounds, such as 3Ca0.A1203.CaCO3.11H20 and 3CaO.A1203.1/2CaCO3.1/2Ca(OH)2.12H20
were formed, these authors suggested that S042- ions in the inter-layer of ettringite crystals
were replaced by CO3-2 ions, and the monosulfate phase was not stable in the presence of
CO2 or CO3-2. The mono-carbonate and hemi-carbonate were more resistant to sulfate attack;
consequently, the specimens showed lower expansion. These results were confirmed by
further tests (See Piasta et al.).
On the other hand, CO-, in solution may cause the formation of thaumasite, which
happens in marine concrete structures in contact with seawater. Sarkar and Malhotra (1995)
studied concrete test panels stored in seawater at a seaside. They reported that the concrete
was carbonated due to the CO2 dissolved in the water, and that thaumasite was observed.
However, laboratory experimental procedures for testing thaumasite attack have not been
established.

3.3.3.9. Precision of various tests


The precision of different test procedures varies according to the number of parameters
involved in sample preparation, curing and testing conditions. In ASTM C1012, it is stated
that the precision of the ASTM procedure evaluated by cooperative testing has been found to
vary with the type of cement. The details are summarised in Table 8. For low expansions
tested according to ASTM C1038, the deviation is proportionally smaller, as shown in Table
9.

Table 8 Precision of ASTM C1012 method for expansion 0.04% to 0.07%*


Single-operator Multi-laboratory
Cement type Standard Max. difference Standard Max. difference
deviation of two tests deviation of two tests _
Blended. Type IP or IS 0.010 % 0.028 % 0.020 % 0.056 %
Type II cements 0.005 % 0.014 % 0.020 % 0.056 %
Type V cement 0.003 % 0.009 % 0.010 % 0.028 %
* At least three mortar bars.

ARRB Transport Research Ltd


30
Review Report 5

Table 9 Precision of ASTM C1038 method for expansion 0.0075% to 0.0115%*


Single-operator Multi-laboratory
Standard deviation Max. difference of two tests Standard deviation Max. difference of two tests
0.00165 % 0.005 % 0.00287 % 0.008 %
* Difference in length change at 24 h and at 14 days.

There are few data about other test procedures; however, individual reports often show
that the expansion of mortar bars tested in duplicated tests is very similar, as discussed by
Frearson and Higgins (1995).

4. Conclusions
Performance-orientated tests are most commonly used for evaluating the resistance of
concrete to sulfate attack. With the development of test methods and extensive research many
deterioration mechanisms have been proposed, which in turn have been incorporated into the
test methods.
Some of the important factors influencing the rate of concrete deterioration, e.g. the pH
of the sulfate solution, have now been incorporated into the standard test method (the effect
of pH was noticed three decades ago!). However, some known phenomena are yet to be taken
into account in research studies, (e.g. the dramatic change in solubility of sodium sulfate with
temperature, or the stability change of ettringite with temperature). As a result misleading or
erroneous conclusions may be made for some tests. It should also be noted that standard
length-change monitoring cannot satisfactorily assess concrete deterioration attributed to low-
pH caused damages. Other means of assessment, such as elastic modulus and strength
measurements (as specified for freeze—thaw resistance) should be formally adopted for
sulfate resistance tests.
Despite decades of research there is insufficient understanding of the mechanisms of
sulfate attack in real conditions and there is also confusion about many observed phenomena.
One confusing example is delayed ettringite formation; ettringite is found in cracks but the
cracks themselves may not have been initiated by the ettringite. Unfortunately, the test
methods so far reported for this are in favour of producing AAR and thermally induced
cracks, both are more rapid in creating cracks than the delayed formation of ettringite
crystals.
Another important aspect is thaumasite-induced deterioration; although occasionally
reported to occur in field, this type of deterioration has not been successfully reproduced in
the laboratory. Consequently, no standard test exists for this. With the ever-increasing use of
pozzolans as supplementary cementing materials, which are theoretically more vulnerable to
the thaumasite-induced deterioration, a test method for this is necessary.
Finally, the standard test methods discussed in this review are mainly laboratory
accelerated tests performed on cement mortar samples. Attention has in the past been mainly
paid to obtaining comparable results for different cement products prepared under standard
conditions. These methods, however, cannot be directly adopted for testing concrete
specimens from real structures without known specifications, or for predicting the service life
of concrete. For the latter, more investigations are required.

ARRB Transport Research Ltd


31
Review Report 5

5. References

AL-AMOUDI, O.S.B. (1995). "Durability of reinforced concrete in aggressive Sabkha


environments," ACI Materials Journal, May-June, p. 236-245.
AL-AMOUDI, O.S.B., (1994). "A discussion of the paper 'A microstructural study of the
effect produced by magnesium sulfate on plain and silica fume-bearing portland cement
mortars' by David Bonen," Cement and Concrete Research, Vol.24(2), p.371-372.
AL-AMOUDI, O.S.B., MASLEHUDDIN, M., and SAADI, M.M. (1995). "Effect of
magnesium sulfate and sodium sulfate on the durability performance of plain and blended
cements," ACI Materials Journal, p.15-24.
AL-AMOUDI, RASHEEDUZZAFAR, MASLEHUDDIN, M., and ABDULJAUWARD, S.N.
(1994). "Influence of chloride ions on sulfate deterioration in plain and blended cements,"
Magazine of Concrete Research, Vol.46(167), p.113-123.
BABUSKIN, V.J., MATVEEV, G.M., and MCHEDLOV-PETROSYAN, O.P. (1972).
Thermodynamics of Silicate, 3rd edition. Moscow.
BAKKER, R. (1981). "On the cause of increased resistance of concrete made from blast
furnace cement to the alkali silica reaction and to sulfate corrosion," (English Translation of
doctoral thesis at RWTH 1980), Maastricht, 144 pages.
BALAZS, G., CSANYI, E., and TAMAS, F. (1997). "Effect of air pollution on portland
cement concrete — Case studies and model experiments," Proceedings, 10th International
Congress on the Chemistry of Cement, Gothenburg, Sweden, Vol.4, 4iv002.
BARONIO, G., and BERRA, M. (1986). "Concrete deterioration with the formation of
thaumasite - analysis of the causes," it cemento, Vol.3, p169-184.
BATE, S.C.C. (1984). "High alumina cement concrete in existing building superstructures,"
Building Research Establishment Report 5040, London, HMSO.
BEN-YAIR, M. (1974). "The effect of chloride on concrete in hot and arid regions," Cement
and Concrete Research, Vol 4, p.405-416.
BERRA, M.,and BARONIO, G. (1987). "Thaumasite in deteriorated concretes in the
presence of sulfates," Concrete Durability: Katharine and Bryant Mather international
conference, Detroit, USA, ACI SP-100, edited by J.M. Scanlon. p.2073-2089.
BICZOK, I. (1967). Concrete Corrosion, Concrete Protection, Chemical Publishing
Company, New York, p.178.
BONEN, D. (1993). "A microstructural study of the effect produced by magnesium sulfate on
plain and silica fume bearing portland cement mortars," Cement and Concrete Research,
Vol.23(3), p.541-553.
BONEN, D. (1994). "Reply to A discussion of the paper 'A microstructural study of the
effect produced by magnesium sulfate on plain and silica fume-bearing portland cement
mortars' by David Bonen," Cement and Concrete Research, Vol.24(2), p. 373-374.
BONEN, D. (1997). "The microstructure of concrete subjected to high-magnesium and high-
magnesium sulfate brine attack," Proceedings, 10th International Congress on the Chemistry
of Cement, Gothenburg, Sweden, Vol.4, 4iv022.
BONEN, D., and COHEN, M.D. (1992a). "Magnesium sulfate attack on portland cement
paste — I. Microstructural analysis," Cement and Concrete Research, Vol.22, p.169-180.

ARRB Transport Research Ltd


32
Review Report 5

BONEN, D., and COHEN, M.D. (1992b). "Magnesium sulfate attack on portland cement
paste — H. Chemical and mineralogical analyses," Cement and Concrete Research, Vol.22,
p.707-718.
BOURNAZEL, J.P., MORANVILLE-REGOURD, M., and SELLIER, A. (1996).
"Microstructure of steam cured concretes deteriorated by alkali-silica reaction," Alkali-
aggregate reaction in concrete, edited by A. Shayan, Proceedings of the 10th International
Conference, Melbourne, Australia. 18-23, p.949-956.
BROWN, P.W. (1981). "An evaluation of the sulfate resistance of cements in a controlled
environment," Cement and Concrete Research, V ol.11(5/6), p.719-727.
BUCEA, L., CAO, H.T., FERGUSON, 0., and MATEO, J. "Sulfate resistance of cement
and concrete - Materials selection," Proceedings of the Austroads 1997 Bridge Conference,
Sydney, Australia, edited by G.J. Chirgwin. p.131-153.
BUTLER, W.B. (1997). "The use of fly ash blended cement for sulfate resistance," ditto.
p.141-144.
CALLEJA, J. (1980). "Durability," Proceedings, 7th International Congress on Chemistry of
Cement, Paris, Vol.1, p.VII-2/1-2/48.
CAMARINI G. and CINCOTTO, M.A. "Delayed ettringite formation in steam cured OPC
and slag cement pastes and its effects on mortar compressive strength," ditto. Vol.4, 4iv063.
CAMPUS, F. (1963). "Essais de resistance des Mortiers et Betons a la Mer (1934-1964),"
Silicates Industriel, Vol 28, p.79-88.
CAO, H.T., BUCEA, L, and FERGUSON, 0. (1997). "Sulfate resistance of cementitious
materials — Mechanisms, deterioration processes, testing and influence of binder," Concrete
'97, Proceedings of 18th Biennial Conference, Adelaide, p.263-268.
CAO, H.T., BUCEA, L., RAY, A., and YOZGHATLIAN, S. (1997). "The effect of cement
composition and pH of environment on sulfate resistance of portland cements and blended
cements," Cement & Concrete Composites Vol .19, p.161-171.
COHEN, M.D., and BENTUR, A. (1988). "Durability of portland cement-silica fume pastes
in magnesium sulfate and sodium sulfate solutions," ACI Materials Journal Vol.85(3), p.148-
157.
COLE, W.F., (1953). "A crystalline hydrated Mg silicate formed in the breakdown of a
concrete sea wall," Nature, Vol.171, p.353-355.
CRAMMOND, N.J. (1984). "Examination of mortars containing varying percentages of
coarsely crystalline gypsum as aggregate," Cement and Concrete Research, Vol.14. p.225-
230.
CRAMMOND, N.J. (1985). "Thaumasite in failed cement mortars and renders from exposed
brickwork," Cement and Concrete Research, Vol.15(6), p.1039-1050.
CRAMMOND, N.J., and HALLIWELL, M.A. (1995). "The thaumasite form of sulfate attack
in concretes containing a source of carbonate ions — A microstructural overview," Advances
in Concrete Technology, edited by V.M. Malhotra; Proceedings, Second CANMET/ACI
International Symposium, Las Vegas, Nevada, USA, p.357-380.
CRC Handbook of Chemistry and Physics, 76th edition. D.R. Lide, editor-in-chief, CRC
Press, 1996.
DAMGHANI, M., and VAKILI, M.B. (1995). "Deterioration of concrete hydraulic structures
in the Persian Gulf region," Concrete Under Severe Conditions: Environment and loading,
Edited by K. Sakai, N. Banthia and O.E. Gjorv. E & FN Spon, Vol.1. p.839-849.

ARRB Transport Research Ltd


33
Review Report 5

DAMIDOT, D., and GLASSER, F.P. (1997). "Thermodynamic investigation of the Ca0-
A1203-CaSO4-CaC12-H20 system at 25°C and the influence of Na2O," Proceedings, 10th
International Congress on the Chemistry of Cement, Gothenburg, Sweden, Vol.4, 4iv066.
DAMIDOT, D., ATKINS, M., KINDNESS, A., and GLASSER, F.P. (1992). "Sulphate attack
on concrete: limits of the AFt stability domain," Cement and Concrete Research, Vol.22,
p.229-234.
DIAMOND, S. (1996). "Delayed ettringite formation — Processes and problems," Cement
and Concrete Composites Vol.18, p.205-215.
DOMIC, D., and DROLJC, S. (1986). "The influence of alite content on the sulfate resistance
of portland cement," Proceedings, 8th International Congress on the Chemistry of Cement,
Rio de Jenero, Vol.V, p.195-199.
DOUGLAS, E., VAN HUYSSTEEN, E., and MALHOTRA, V.M. (1991). "Sulfate resistance
of mortars made with high volumes of Class F fly ash or granulated blast-furnace slag —
Progree report," Supplementary Papers, .2nd CANMET/ACI International Conference on
Durability of Concrete, Montreal, p.749-796.
DUCIC, V., and MILETIC, S. (1986). "Sulfate corrosion resistance of blended cement
mortars," Supplementary Papers, 2nd ACl/CANMET International Conference on Fly Ash,
Silica Fume, Slag and Natural Pozzolans in Concrete, Madrid, p.12-26.
ERLIN, B., and STARK, D.C. (1965). "Identification and occurrence of thaumasite in
concrete," Highway Research Record No.113, p.108-113.
FAGERLUND, G. (1987). Betongkonstruktioners Bestandighet — En oversikt. Uppsala,
Sweden.
FIDJESTOL, P., and FREARSON, J.P.H. (1994). "High performance concrete using blended
and triple blend binders," Proceedings, ACI Conference on High Performance Concretes,
Singapore.
FOLLIARD, K.J., and SANDBERG, P. (1994). "Mechanisms of Concrete Dterioration by
Sodium Sulfate Crystallization", Proceedings, 3rd CANMET/ACI International Conference
on Durability of Concrete, Nice, France, ACI SP-145, p.933-945.
FREARSON, J.P.H., AND HIGGINS, D.D. (1995). "Effect of test procedures on the
assessment of the sulfate resistance of slag cements," Proceedings of Fifth CANMET/ACI
International Conference on Fly Ash, Silica Fume, Slag, and Natural Pozzolans in Concrete,
Milwaukee, Wisconsin, USA, ACI SP-153. Vol.2. p.975-993.
FU, Y., and BEAUDOIN, J.J. (1996). "On the distinction between delayed and secondary
ettringite formation in concrete," Cement and Concrete Research, Vol.26(6), p.979-980.
FU, Y., DING, J., and BEAUDOIN, J.J. (1997). "Expansion of portland cement mortar due to
internal sulfate attack," Cement and Concrete Research, Vol.27(9), p.1299-1306.
GAZE, M.E. (1997). "The effects of varying gypsum content on thaumasite formation in a
cement," Cement & Concrete Research, Vol.27(2), p.259-265.
GERWICK, JR., B.C. (1990). "International Experience in the performance of marine
concrete," Concrete International, p.47-53.
GIERGICZNY, Z. (1997). "Sulfate resistance of cements with mineral admixtures,"
Proceedings, 10th International Congress on the Chemistry of Cement, Gothenburg, Sweden,
Vol.4, 4iv019.
GLASSER, F.P. (1996). "The role of sulfate mineralogy and cure temperature in delayed
ettringite formation," Cement & Concrete Composites Vol.18, p.187-193.

ARRB Transport Research Ltd


34
Review Report 5

GOLLOP, R.S., and TAYLOR, H.F.W. (1994). "Microstructural and microanalytical studies
of sulfate attack. II. Sulfate-resisting portland cement: Ferrite composition and hydration
chemistry," Cement and Concrete Research, Vol.24(7), p.1347-1358.
GONZALEZ, M.A., and IRASSAR, E.F. (1997). "Ettringite formation in low C3A portland
cement exposed to sodium sulfate solution," Cement and Concrete Research, Vol.27(7),
p.1061-1306.
HAYNES, H., O'NEILL, R., and MEHTA, P.K. (1996). "Concrete deterioration from
physical attack by salts," Concrete International, p.63-68.
HEINZ, D., LUDWIG, U., and RUDIGER, I. (1989). "Delayed ettringite formation in heat
treated mortars and concretes," Concrete Precasting Plant and Technology, Vol.11, p.56-61.
HIME, W.G., CONNOLLY, J., and MARUSIN, S. (1996). "Delayed ettringite formation,"
Proceedings, 18th International Conference on Cement Microscopy, Houston, TX, p.378.
HOGAN F.J. and MEUSEL, J.W. (1981). "Evaluation for durability and strength
development of ground granulated blast furnace slag," Cement, Concrete and Aggregates,
3(1), p.40-52.
JACKSON, P.J., and LAWTON, J.M. (1992) "The development of standards for cements,"
Proceedings, 9th International Congress on the Chemistry of Cement, New Delhi, Vol.V, p.
157-164.
JOHANSEN, V., THAULOW, N., and IDORN, G.M. (1994). "Expansion reactions in mortar
and concrete," Zement-Kalk-Gips, Vol.47(5), E150-155.
JONS, E.S. (1996). "Measuring the solubility of sulfate in cement," World Cement Research
and Development Section, p.65-68.
KALOUSEK, G.L., POR1ER, L.C., and BENTON, E.J. (1972). "Concrete for long-time
service in sulfate environment," Cement and Concrete Research, Vol.2(1), p.79-89.
KELHAM, S. "Effects of cement composition and hydration temperature on volume stability
of mortar," ditto. Vol.4, 4iv060.
KLEMM, W.A., and MILLER, F.M. (1997). "Plausibility of delayed ettringite formation as a
distress mechanism — Considerations at ambient and elevated temperatures," Proceedings,
10th International Congress on the Chemistry of Cement, Gothenburg, Sweden, Vol.4,
4iv059.
KOCH, A., and STEINEGGER, H. (1960). "A rapid method for testing the resistance of
cements to sulfate attack," "Sulfate-resistant cements and their testing," Zement-Kalk-Gips,
Vol.13(7), p.317-324.
KOLLEK, J.J., and LUMLEY, J.S. (1990). "Comparative sulfate resistance of SRPC and
porland slag cements ," Blue Circle Technical Services Division Report.
LAWRENCE, C.D. (1990). "Sulphate attack on concrete," Magazine of Concrete research,
Dec., Vol.42(153). p.249-264.
LEA, F.M. (1970). The Chemistry of cement and Concrete, 3rd Edition, Chemical Publishing
Co., Inc., New York, NY, p.727.
Lea's Chemistry of Cement and Concrete, 4th Edition. Edited by P.C. Hewlett. (1998).
Arnold.
LOCHER, F.W. (1956). "Testing the sulfate resistance of cements," Zement-Kalk-Gips,
Vol.9, p.204-210.
LOCHER, F.W. (1966). "The problem of the sulfate resistance of slag cements," Zement-
Kalk-Gips, Vol.19, p.395-401.

ARRB Transport Research Ltd


35
Review Report 5

LUKAS, W. (1975). "Betongzerstorung durch SO3 - Angriff under bildung von thaumasit and
woodfordit," Cement and Concrete Research, Vol.5(5), p.503-518.
MANGAT, P.S., and EL-KHATIB, J.M. (1992). "Influence of initial curing on sulfate
resistance of blended cement concrete," Cement and Concrete Research, Vol.22(6),. p.1089-
1100.
MATHER, B. (1957). "Laboratory tests of portland blast-furnace slag cements," Journal of
the American Concrete Institute, Proceedings, Vol 54, p.205-232.
MATHER, K. "Current research in sulfate resistance at the waterways experiment station,"
Proceedings of George Verbeck Symposium on Sulfate Resistance of Concrete, ACI SP-77.
p.63-74.
MATHER, K. (1978). "Tests and evaluation of portland and blended cements for resistance
to sulfate attack," ASTM STP 663, p.74-86.
MATTA, Z.G. (1993). "Deterioration of concrete structures in the Arabian Gulf," Concrete
International, July, p.33-36.
MEHTA, P.K. (1973). "Mechanism of Expansion Associated with Ettringite Formation,"
Cement and Concrete Research, Vol.3(1), p.1-6.
MEHTA, P.K. (1975). "Evaluation of sulfate - resisting cements by a new test method,"
Proceedings of the American Concrete Institute, Vol 72, p.573-575.
MEHTA, P.K. (1980). "Performance tests for sulfate resistance and alkali-silica reactivity of
hydraulic cements," Durability of Building Materials and Components, ASTM STP 691, P.J.
Sereda and G.G. Litvan, Eds., ASTM STP 691, p.336-345.
MEHTA, P.K. (1993). "Sulfate Attack on Concrete: A Critical Review," Durabilidad del
Concreto (Concrete Durability), edited by R.R. Villarreal, University of Autonoma de Nuevo
Leon, p.107-132.
MEHTA, P.K., and HAYNES, H.H. (1975). "Durability of concrete in sea water,"
Proceedings of the American Society for Civil Engineers, Vol 101, No. ST 8, p.1679-1686.
MEHTA, P.K., and POLIVKA, M. (1975). "Sulfate resistance of expansive cement
concretes," AC/ SP-47, p.367-379.
MEHTA, P.K., PIRTZ, D., and POLIVKA, M. (1979). "Properties of Alite Cement," Cement
and Concrete Research, Vol.9(4), p.439-450.
MEHTA, P.K., SCHIESSL, P., AND RAUPACH, M. (1992) "Performance and durability of
concrete systems," Proceedings, 9Th International Congress on the Chemistry of Cement, New
Delhi, India, Vol.I. p. 571-659.
MELAND, I., JUSTNES, H., and LINDGARD, J. "Durability problems related to delayed
ettringite formation and / or alkali aggregate reactions," ditto. Vol.4, 4iv064.
MICHAUD, V., NONAT, A., and SORRENTINO, D. (1997). "Experimental simulation of
ettringite formation in alkali silica solutions, produced by alkali-silica reaction, in concrete,"
Proceedings, 10th International Congress on the Chemistry of Cement, Gothenburg, Sweden,
Vol.4, 4iv065.
MIELENZ, R.C., MARUSIN, S.L., HIME, W.G., and JUGOVIC, Z.T. (1995). "Investigation
of prestressed concrete railway tie distress," Concrete International, p.62-68.
MILLER, D.G., and MANSON, P.W. (1940). "Tests of 106 commercial cements for sulfate
resistance," Proceedings, ASTM, Vol 40, p.988-1001.

ARRB Transport Research Ltd


36
Review Report 5

MILLER, D.G., and SNYDER, C.G. (1945). "Report on comparative short -time tests for
sulfate resistance of 121 commercial cements," Report of Committee C-1 on Cement,
Appendix III, Proceedings, ASTM Vol 45,. p.165-194.

NOVAK, G.A., and COLVILLE, A.A. (1989). "Efflorescent mineral assemblages associated
with cracked and degraded residential concrete foundations in southern California", Cement
and Concrete Research, Vol.19, p.1-6.
ODLER, I. (1997). "Letter to the editor: Ettringite nomenclature," Cement and Concrete
Research, Vol.27(3), p.473-474.
OSBORNE, G.J. (1989). "Determination of the sulfate resistance of blast-furnace slag
cements using small-scale accelerated methods of test," Advances in Cement Research, No.5,
p.22-27.
PATZIAS, T. (1987). "Evaluation of sulfate resistance of hydraulic- cement mortars by the
ASTM C1012 test method." Concrete Durability, Katharine and Bryant Mather International
Conference, A CI SP-100, Vol.2, p.2103-2120.
PATZIAS, T. (1991). "The development of ASTM C1012 with recommended acceptance
limits for sulfate resistance of hydraulic cement," Cement, Concrete, and Aggregates, CCA
GDP, Vol 13(1), Summer. p.50-57.
PIASTA, W.G., SAWICZ, Z., KOPROWSKI G., and OWSIAK, Z. (1997). "Influence of
limestone powder filler on microstructure and mechanical properties of concrete under
sulfate attack," Proceedings, 10th International Congress on the Chemistry of Cement,
Gothenburg, Sweden, Vol.4, 4iv018.
POLIVKA, M., and BROWN, E.H. (1958). "Influence of various factors on sulfate resistance
of concrete containing pozzolan," Proceedings, ASTM, Vol 58, p.1077-1100.
REGOURD, M. (1975). "The action of sea water on cements," Annales de L'Institut
Technique du Batiment et des Travaux Publics, Vol 329, p.86-102.
REGOURD, M. (1981). "Chemical durability of concrete," Proceedings of Contemporary
European Concrete Research, Stockholm, June 9-11,. p.121-142.
REGOURD, M. (1988). "Physiochemical studies of cement pastes, mortars and concretes
exposed to sea water," Performance of Concrete in Marine Environment, edited by V.M.
Malhotra, p.63-82.
ROSNERL, J.C., CHEHOVITS, J.G., and WHARBURTON, R.G. (1982). "Sulfate resistance
of mortars using fly ash as a partial replacement for portland cement," Proceedings, 6th
International Conference on Utilization of Fly Ash, Reno.
SAMANTA, C., and CHATTERJEE, M.K. (1982). "Sulfate resistance of portland-pozzolanic
cements in relation to strength," Cement and Concrete Research, Vol.12(5/6),. p.726-734.
SAND, W., DUMAS, T., and MARCDARGENT, S. (1994). "Accelerated Biogenic Sulfuric
Acid Corrosion Test for Evaluating the Performance of Calcium - Aluminate Based Concrete
in Sewage Applications," Micro - biologically Influenced Corrosion Testing, edited by J.R.
Kearns and B.J. Little, ASTM STP-1232, p.234-249.
SARKAR, S.L., and MALHOTRA, V.M. (1995). "Microstructure durability of concrete
exposed to Artic conditions," Concrete Under Sever Conditions-Environment and Loading,
Edited by K. Sakai, N. Banthia and O.E. Gjorv. E&FN SPON, p.828-838.
SAWICZ, Z., and HENG, S.S. (1996). "Durability of concrete with addition of limestone
powder," Magazine of Concrete Research, Vol.48(175), p.131-137.
SAYWARD, J.M. (1984). "Salt action on concrete," Special Report 84-25. US Army Corps
of Engineers, Cold Regions Research & Engineering Laboratory.

ARRB Transport Research Ltd


37
Review Report 5

SCHNEIDER, U., and PIASTA, W.G. (1991). "The behaviour of concrete under Na2SO4
solution attack and sustained compression or bending," Magazine of Concrete Research,
Vol.43(157), p.281-289.
SCRIVENER, K., and LEWIS, M. (1997). "A microstructural and microanalytical study of
heat cured mortars and delayed ettringite formation," Proceedings, 10th International
Congress on the Chemistry of Cement, Gothenburg, Sweden, Vol.4, 4iv061.
SCRIVENER, K.L. (1996). "Delayed ettringite formation and concrete railroad ties"
Proceedings, 18th International Conference on Cement Microscopy, Houston, TX, p.375-377.
SERSALE, R., CIOFFI, R., DE VITO, B., FRIGIONE, G., and ZENONE, F. (1997).
"Sulphate attack of carbonated and uncarbonated portland and blended cement mortars,"
Proceedings, 10th International Congress on the Chemistry of Cement, Gothenburg, Sweden,
Vol.4, 4iv017.
SHAYAN, A. (1988) Deterioration of a concrete surface due to the oxidation of pyrite
contained in pyritic aggregates. Cement & Concrete Research, Vol.18(5), p. 723 -730.
SHAYAN, A. (1995). "Behaviour of precast prestressed concrete railway sleepers affected by
AAR," Real World Concrete, Proceedings of R.N. Swamy Symposium, Milwaukee, USA,
p.35-56.
SHAYAN, A., AND IVANUSEC, I. (1996). "An experimental clarification of the association
of delayed ettringite formation with alkali-aggregate reaction," Cement and Concrete
Composites, Vol.18, p.161-170.
SHAYAN, A., and QUICK, G.W. (1991/92). "Relative importance of deleterious reactions in
concrete: formation of AAR products and secondary ettringite," Advance in Cement
Research, Vol.4(16). p.149-157.
SHAYAN, A., and QUICK, G.W. (1992). "Microscopic features of cracked and uncracked
concrete railway sleepers," ACI Materials Journal, Vol.89(4), p.348-361.
SMOLCZYK, H., and BLUNK, G. (1972). "The behaviour of very young concrete in sulfate
solution," Beton Information, Vol. 1, p.2-9.
SOROKA, I., and CARMEL, D. (1987). "Durability of external renderings in a marine
environment," Durability of Building Materials, Vol.5,. p.61-72.
ST JOHN, D.A. (1982). "An unusual case of ground water sulphate attack on concrete,"
Cement and Concrete Research, Vol.12, p.633-639.
ST JOHN, D.A., POOLE, A.W., and SIMS I. (1998) Concrete Petrography: A handbook of
investigative techniques, Arnold.
STARK, J., and BOLLMANN, K. (1997). "Ettringite formation — A durability problem of
concrete pavements," Proceedings, 10th International Congress on the Chemistry of Cement,
Gothenburg, Sweden, Vol.4, 4iv062.
STARK, J., BOLLMANN, K., and SEYFARTH, K..(1992). "Investigation into delayed
ettringite formation in concrete," Proceedings, 9th International Congress on the Chemistry
of Cement, New Delhi, India, V ol.V , p. 348-354.
TAYLOR, H.F.W. (1977) Concrete Technology and Practice, 4th edition, McGraw-Hill Book
Company, Sydney.
TAYLOR, H.F.W. (1990). Cement Chemistry, Academic Press, London.
TAYLOR, H.F.W. (1996). "Ettringite in cement paste and concrete," Beton du Materiaux
la Structure, Conference in honour of M. Moranville-Regourd, Arles, France.

ARRB Transport Research Ltd


38
Review Report 5

TAZAWA, E.I., MORINAGA, T., and KAWAI, K. (1994). "Deterioration of concrete


derived from metabolites of micro organisms," Durability of Concrete, edited by V.M.
Malhotra, ACI SP-145, p.1087-1097.
WAFA, F.F. (1994). "Accelerated sulfate attack on concrete in hot climate," Cement,
Concrete, and Aggregate, Vol.16 (1), p.31-35.
WITTEKINDT, W. (1960). "Sulfate-resistant cements and their testing," Zement-Kalk-Gips,
Vol.13(12), p.565-572.
WOLOCHOW, D. (1952b). "A lean mortar bar expansion test for sulfate resistance of
portland cement," "Determination of the sulfate resistance of portland cement," ASTM
Proceedings, Vol.52, p.250-266.
WOLOCHOW, D. (1952b). "Determination of the sulfate resistance of portland cement,"
Report of Committee C-1 on Cement, Appendix, Proceedings, ASTM, Vol 52, p.250-363.
YEGINOBALI, A., and DILEK, F.T. (1995). "Sulfate resistance of mortars containing silica
fumes as evaluated by different methods," Proceedings of Fifth CANMET/ACI International
Conference on Fly Ash, Silica Fume, Slag, and Natural Pozzolans in Concrete, Milwaukee,
Wisconsin, USA, ACI SP-153. Vol.2, p.795-813.

Standards

ACI 201.2R-77: "Guide to Durable Concrete," American Concrete Institute, 1977.


AS 1315-1982: "Portland Cement," Standards Australia.
AS 2350.14-1996 "Length change of portland and blended cement mortars exposed to a
sulfate solution," Standards Australia.
AS 3972-1997 "Portland and Blended Cements," Standards Australia.
ASTM C150-97: "Standard Specification for Portland Cement," Annual Book of ASTM
Standards, Vol 04.01.
ASTM C265-91: "Standard Test Method for Calcium Sulfate in Hydrated Portland Cement
Mortar," Annual Book of ASTM Standards, Vol 04.01.
ASTM C452-95: "Standard Test Method for Potential Expansion of Portland - Cement
Mortars Exposed to Sulfate," Annual Book of ASTM Standards, Vol 04.01.
ASTM C1012-95a: "Standard Test Method for Length Change of Hydraulic-Cement Mortars
Exposed to a Sulfate Solution," Annual Book of ASTM Standards, Vol 04.01.
ASTM C1038 - 95: "Standard Test Method for Expansion of Portland - Cement Mortar Bars
Stored in Water," Annual Book of ASTM Standards, Vol 04.01.
ASTM C1157M-97: "Standard Performance Specification for Blended Hydraulic Cement,"
Annual Book of ASTM Standards, Vol 04.01.
BS 4027:1991 "Specification for Sulfate-resisting Portland Cement." British Standard
Institution, 1991.
BRE Digest 363 "Sulphate and acid resistance of concrete in the ground," BRE Digest
Concise reviews of building technology, Series 802 Digest 363, July 1991. 8 pages.
CEMBUREAU "Use of concrete in aggressive environments," Cembureau Recommendation
1 st Edition, Paris 1978.

ARRB Transport Research Ltd


Recent publications I arP13
Research Report No. 315 Transport
Research
Practical relationships for the assessment of road feature treatments
J. McLean

Research Report No. 316


Trial of stored value cards for parking operations on the Gold Coast.
G. Giummarra

Research Report No. 317


The National Protocol System in 1997: a cooperative management system to maintain
biodiversity
Q. Farmer-Bowers

Research Report No. 318


Toward a methodology for comparative resource consumption:
modal implications for the freight task
N. Houghton, J. McRobert

Research Report No. 320


Relationships between accidents and access conditions
R. Brindle

Research Report No. 321


Roundabouts: Capacity and performance analysis
R. Akcelik

Research Report No. 323


Night-time noise levels: a state-of-the-art review
H. Lansdell, C. Cameron

Review Report 1
The use of accident costs for countermeasure evaluation in Australia
N. Mabbott, D. Swadling

Review Report 2
Examples of recent developments in specialised transport technologies
F Clerk Edited by R. Brindle

Review Report 3
A review of transport resources for people with disabilities
J. Evans, M. White

Maintenance Specifications
Supplement to the Unsealed Roads Manual

Special Report No. 56 ARRB Transport Research Ltd


publishes a large number of
State-of-the-art pavement performance modelling at a network and project level
technical reports and
T Martin manuals.

SIDRA 5 User Guide To order these or other ARRB


R. Akcelik, M. Besley Transport Research
Publications, or a free
catalogue contact:
AUSTROADS/ARRB Transport Research Reports ARRB Transport
Research Ltd
APRG Report No. 18 ACN 004 620 651
Selection and Design of Asphalt Mixes: Australian Provisional Guide
500 Burwood Highway
Vermont South
APRG Report No. 19 VIC 3133
Austroads specification framework for polymer modified binders AUSTRALIA

APRG Report No. 21 Tel: (03) 9881 1555


A guide to the design of new pavements for light traffic Fax: (03) 9887 8104

International
APRG Report No. 22/ARR 322 Tel: +61 3 9881 1555
The performance of insitu stabilised marginal sandstone pavements Fax: +61 3 9887 8104

Email: info@arrb.org.au
Internet: www.arrb.org.au
AUSTROADS publications - a selection
AP-11 Traffic Engineering Practice

AP-17 Guide to the Structural Design of Road Pavements


AUSTROADS
AP-34 Design Vehicles and Turning Path Templates

AP-41/96 Bitumen Sealing Safety Guide

AP-42 Benefit Cost Analysis Manual

AP-43/98 The Australian and New Zealand Road System and


Road Authorities National Performance Indicators 1997

AP-44/97 Asphalt Recycling Guide

AP-48/97 Australia at the Crossroads: Roads in the Community - a Summary

AP-49/97 Roads in the Community: Are they doing their job? (in press)

AP-50/97 Roads Serving the Community: Towards Better Practice (in press)

AP-51/98 Electronic Toll Collection Standards Study: a report by ITS Australia for
AUSTROADS

AP-55/98 Principles for strategic planning

AP-117 Travel Demand Management

AP-118 Urban Speed Management in Australia

AP-119/97 Value of travel time savings

AP-120 National Guidelines for Alcohol Interlock Programs

AP-121 Novice Car Driver Competency Specification

AP-123/97 Economic Effects of Investment in Road Infrastructure

AP-124/97 Benchmarking Framework

AP-125/97 Austroads/Transit New Zealand partnership plan background paper


AUSTROADS publishes a
AP-126/97 A Minimum Common Dataset for the Reporting of Crashes large number of guides anc
reports.
on Australian Roads
For a full list with prices, or
Concrete Structures Durability, Inspection and Maintenance Procedures to place orders, please
AP-127/97
contact:

AP-128 Alcohol impaired pedestrians ARRB Transport


Research Ltd
ACN 004 620 651
AP-129 Responsibilities for local roads
500 Burwood Highway
Vermont South
AP-130 Competition in the supply of roadworks to government
VIC 3133
AUSTRALIA
AP-131 Private sector financing in roads
Tel: (03) 9881 1555
Fax: (03) 9887 8104
AP-201/93 Environmental Impact Assessment of Major Roads in Australia
International
Tel: +61 3 9881 1555
Fax: +61 3 9887 8104

Email: info@arrb.org.au
Internet: www.arrb.org.au
Test methods for sulfate resistance of concrete and mechanism of sulfate attack
0 LA
0.
a)
cc
a)
a)

ISBN 0 869 10 7720

You might also like