You are on page 1of 31

4.

3 Sulphate attack

J. Trägårdh
Cement and Concrete Research Institute (CBI), Drottning Kristinas väg 26, Stockholm,
Sweden.

F. Bellmann
F.A. Finger-Institute for Building Material Science, Bauhaus University Weimar,
Coudraystrasse 11, 99423 Weimar, Germany.

4.3.1. Introduction

Portland cement concrete can be attacked by solutions containing sulphate ions, such
as some natural or polluted ground waters. Soils containing sodium, potassium, magnesium,
and calcium sulphate are the main sources of sulphate ions in groundwater [1, 2]. Sulphate
attack on concrete has been reported to be the cause of severe damage to concrete for over a
century. Attack can lead to strength loss, expansion, spalling of surface layers and, ultimately,
disintegration [3, 4].
Ettringite (C3A.3CaSO4.32H2O), thaumasite (CaCO3.CaSO4.CaSiO3.15H2O), and
gypsum (CaSO4.2H2O) are three minerals, which are found among the deterioration products
of cementitious materials exposed to sulphate attack [3-8]. Ettringite formation is considered
to be the cause of most of the expansion and disruption of concrete structures affected by
sulphate attack [3]. However, not every attack necessarily results in the formation of ettringite
[9]. The following forms of sulphate attack have been recognized in the literature [10]:
• The classic form of sulphate attack associated with ettringite and/or gypsum formation,
• Thaumasite form of sulphate attack (TSA),
• Physical sulphate attack associated with the crystallization of sulphate- containing salts
at or near an evaporative surface,
• Delayed ettringite formation (DEF),
• Sulphate attack associated with the formation of AFm phases.
This chapter will review the classic form of sulphate attack, the thaumasite form of
sulphate attack (TSA), and delayed ettringite formation (DEF). The focus is on self-
compacting concretes (SCC) and how this type of concretes may differ from traditional
concretes (TC) with respect to the ingress of sulphate-rich solutions. Physical resistance to

89
sulphate attack is especially important, as it may be regarded as one of the main factors which
differentiate SCC from TC. Physical resistance and durability properties of SCC are
influenced by the large amounts of added filler materials, which in turn have an influence on
the microstructure [11, 12] and the connectivity/fineness of the capillary pore system. Other
factors which are treated and which influence the nature of sulphate attack are chemical
resistance (type of cement), composition of the sulphate solution and addition of pozzolan
materials.
Sulphate attack is treated from the perspective of concrete exposure to near-neutral
ground waters in soil and bedrock and marine waters. Acidic attack containing sulphates is
treated in chapter 4.7. In sections 4.3.3.1 and 4.3.4 previously unpublished results are
presented by the first author.
In recent years SCC has been more and more widely used in many countries around
the world. SCC is used for house building and infrastructures such as bridges, tunnel works
etc. One of the most commonly used filler materials is limestone filler and the amounts can
sometimes reach the same levels as the cement content. This makes SCC potentially
vulnerable to TSA if the amount of sulphates in the environment exceeds certain threshold
values. In order to evaluate the risk of TSA in SCC, as well as TC, several laboratory
investigations have been carried out in different countries. Some examples of investigations
are [13- 25].
According to German specifications (DIN 1164) Portland cements with an Al2O3
content of maximum 5 wt %, or a maximum C3A content of 3 % (SRPC, CEM I) or blast
furnace slag cements with a minimum slag content of 66 wt % (CEM III/B) are sulphate
resistant cements by definition. Guidelines for the use of concrete in aggressive ground are
given in DIN 1045 which stipulates the use of sulphate resistant cement and dense concrete.
In Germany, the ground is considered aggressive with respect to sulphate attack when ground
waters have sulphate concentrations exceeding 600 mg/l, or when the sulphate content of the
soil exceeds 3000 mg/kg. This corresponds to class XA2 (moderately aggressive) in EN 206-
1.
During the 1990-ties, a number of cases of TSA have occurred in the U.K. which
resulted in concrete deterioration of buried members of highway bridge structures containing
limestone aggregates. The concrete members were embedded in pyritic clays containing
ground waters with high sulphate concentrations. This caused some alarm and the Building
Research Establishment (BRE) published guidelines for the use of concrete in aggressive
ground in 2001 [26]. According to the BRE guidelines a risk for TSA or classic sulphate
attack exists when the sulphate concentration of ground waters exceeds 400 mg/l (sulphate
class DC-2).
Some limitations must be put on the use of Portland limestone cements and
limestone fillers in SCC regarding specific environments and new or complementary EN-
guidelines are needed in the future. In the UK, the experience of TSA has lead to a
complementary British standard [27] to EN 206-1, as a consequence of the CEM II-class
cements allowing up to 15% and 30% limestone filler.

4.3.1 Classic form of sulphate attack

The classic form of sulphate attack is determined by the chemical interaction of


sulphate-rich soil or water with calcium aluminate hydration products (C-A-H) in the cement

90
paste formed by the hydration of Portland cement (4CaO.Al2O3.13H2O, 3CaO.Al2O3.6H2O,
C3A.CaSO4.12-18H2O, etc.) [5].
The sulphate attack that leads to expansive ettringite formation is best described by the
following ionic equation [3]:
2+ 2- - -
6Ca + 3SO4 + 2Al(OH)2 + 4OH + 26H2O 3CaO.Al2O3.3CaSO4.32H2O (4.3.1)

The formation of ettringite can be accompanied by the precipitation of gypsum according to


the equation:
2+ 2-
Ca + SO4 + 2H2O CaSO4.2H2O (4.3.2)

Ettringite is not stable at low pH levels (pH 11.5-12) where it could decompose to form
gypsum [28].
The reasons for the destruction of the material and possible measures to prevent
damages have been investigated for more than 100 years. Most of the knowledge was
obtained in studies on conventionally vibrated concrete (TC), whereas the resistance of SCC
has only been investigated in few studies [11, 13- 16, 29].
Important differences between TC and SCC with respect to sulphate attack are
addition of a high amount of mineral admixtures (e.g. limestone, fly ash) and increased
impermeability of the concrete. In the following section the sulphate attack on hardened
concrete is reviewed, with a special focus on SCC where data is yet available.

4.3.2.1 Formation of Ettringite (C3A·3CaSO4·32H2O)


Ettringite is a hardly soluble salt that is composed of columns in which three calcium
ions alternate with one aluminium ion. These positively charged ions are connected within the
columns by hydroxide ions. The columns themselves are linked by sulphate ions and water
forming a trigonal structure (ICSD 16045). Ettringite crystals usually have a long hexagonal
prismatic shape when observed by Scanning Electron Microscopy (SEM). Substitutions of
foreign ions for aluminium and sulphate in the crystal structure were reported and it was
suggested to use the term AFt for ettringite and solid solutions of this mineral [30]. Relevant
details of the structure were reviewed by [31]. Most of the water is only weakly bound
causing the loss of structural water at relatively low temperatures and low partial pressures for
water. Thus, ettringite can shrink and swell by the loss and gain of crystal water [32, 33] but it
is doubtful if this behaviour is responsible for the damage of the hardened cement paste
during sulphate attack [34].
The formation of ettringite takes place during sulphate attack on hardened concrete. It
is also formed during cement hydration by the reaction of the aluminate (C3A) and the ferrite
phase (C4AF) with the set regulator. Therefore, the mere identification of ettringite is not a
proof that the tested mortar or concrete was damaged by sulphate attack. Nevertheless, the
formation of ettringite by sulphate attack on concrete can lead to severe damage. This
problem was recognized as early as 1890 by [35].
The ability of hydrated clinker phases to withstand sulphate attack was first studied by
[36]. Mortar prepared from C3S and C2S showed a very good resistance, whereas samples
containing C3A were badly damaged within a very short time after exposure. The expansion

91
recorded for specimens containing C4AF was clearly lower than for mortars containing C3A.
These results were confirmed by extensive field and laboratory investigations [37- 49]. It was
found that the resistance of concrete to sulphate attack can be increased by lowering the
permeability (low water cement-ratio, high cement content, good grading curve of the
aggregate) and by the use of sulphate resisting cements. Cements with a high chemical
resistance to sulphate attack are Portland cements with a low C3A content and binders
containing pozzolanic and latently hydraulic admixtures. The studies mentioned above
indicated fairly similar C3A contents of about 5% that can be present in sulphate resisting
Portland cements (SRPC).
By limiting the C3A content in SRPC, the amount of ettringite that can be formed is
lowered. Most of the C3A reacts already during the early hardening stage with sulphate from
the set regulator and from the alkali sulphates to ettringite. Ettringite itself is hardly
susceptible to sulphate attack and the formation of ettringite in the early hardening stage has
no detrimental effect. After the consumption of the sulphate by the formation of ettringite,
some of the C3A still remains. A reaction of this remaining C3A gives AFm phases, which is a
series of calcium aluminate hydrates with different anions in the structure. Monosulphate
(C3A·CaSO4·12H2O), monocarbonate (C3A·CaCO3·11H2O) and calcium aluminate hydrate
(C3A·Ca(OH)2·12H2O) are the most relevant of these AFm phases. Structural details were
reviewed by [31] and the formation of solid solutions between these minerals was discussed
by [50]. AFm phases are susceptible to sulphate attack since they can react with sulphate ions
from the aggressive solution to ettringite. The basic idea in the development of SRPC with a
low content of C3A was to lower the amount of AFm phases after hydration because these are
vulnerable to sulphate attack. But it needs to be considered that there are also other sources
for aluminium besides C3A. Especially relevant is ferrite (C4AF). This clinker phase is a solid
solution between C2F and C2A in which the Al-rich end member does not exist. The amount
of C3A in a clinker and the Al/Fe ratio in the ferrite phase both depend on the alumina ratio in
the raw meal. The lower the Al/Fe ratio in ferrite, the less reactive is this phase [51, 52].
Therefore, SRPC is produced with a low alumina ratio in the raw meal ensuring that the
clinker has low content of C3A and a hardly reacting C4AF. Ferrite has a tendency to be zoned
during cooling with Fe-rich cores and Al-rich outer regions [31]. It is evident that ferrite also
contributes to the formation of ettringite but much slower and only to a minor extent due to its
low rate of reaction.
The aforementioned considerations were taken into account during the discussion of
standards for the production of SRPC. ASTM C150 has a limit for C3A in type V cements
( 5%) and also for the amount of C4AF (2·C3A+C4AF 25%). The German standard DIN
1164:2000-11 allows 3% C3A and 5% Al2O3 in SRPC. Other nations have broadly similar
requirements for SRPC.
The specific reaction by which ettringite is formed leading to expansion and micro
cracking has been discussed for a long time. In the following section some relevant ideas
appearing in the literature are reviewed.
• Ettringite possesses a higher volume than the solid phases that are consumed during
the formation of this mineral. Thus, there is an increase in solid volume during a
reaction of monosulphate, calcium aluminate hydrate or other AFm phases to
ettringite. This increase in solid volume can lead to a crystallization pressure if the
mineral precursor of ettringite is tightly bound within the microstructure and water and

92
sulphate ions are transported via gel pores [53]. The crystallization pressure exerted to
the pore walls during the growth of ettringite depends on the supersaturation with
respect to this mineral. It can be higher than the tensile strength of hardened cement
paste [54]. The reaction of sulphate prone phases in a dense microstructure to ettringite
by a dissolution-precipitation mechanism can lead to expansion and the formation of
micro cracks, preferentially around aggregates [34, 55, 56]. It was discussed that the
higher solubility of aluminium in the absence of portlandite [57] after a pozzolanic
reaction can result in a transport of aluminium over longer distances and therefore
ettringite can preferentially be formed in large empty pores [34]. This would help to
explain why pozzolanic materials can increase the resistance to sulphate attack.
• It was supposed that ettringite crystals that are formed at high pH-values have only
very small sizes. These small crystals could eventually be able to attract water
generating an expansion of the surrounding paste [58, 59]. However, this idea neglects
that ettringite has a crystal structure and therefore a continuous uptake of water by this
mineral is not possible. Also an adsorption of water on the surface of ettringite is not
likely because of the (compared to C-S-H phases) small specific surface which is
relevant for adsorption [34].
• A topochemical formation of ettringite, first discussed by [60], is not likely because
the crystal structure of ettringite is different from the one of its precursor (AFm) and
the reaction proceeds via a dissolution-precipitation-mechanism [34, 54, 61].

4.3.2.2 Formation of Gypsum (CaSO4·2H2O)


Gypsum is the mineral form of calcium sulphate that contains the highest amount of
crystal water. The mineral can often be found in nature in saline sediments together with
anhydrite (CaSO4), clay minerals, carbonates, halite and other minerals. Gypsum is composed
of Ca-O-S layers that are connected by water molecules. Due to its structure, gypsum exhibits
a good cleavage. The unit cell is monoclinic (ICSD 27221). At temperatures between 100 and
140°C gypsum decomposes to bassanite (CaSO4·0.5H2O) that reacts at higher temperatures to
anhydrite. The solubility of gypsum depends on the temperature. About 2 g/l can be dissolved
at 20°C.
A first systematic study of the deterioration of mortar samples made from alite and
belite by sulphate attack was made by [36]. It was shown that a formation of gypsum can be
observed after the action of sodium sulphate solutions. The expansion recorded during the
formation of gypsum was very limited in that study when compared to the one encountered
during the formation of ettringite. In contradiction to this, the formation of gypsum was
accompanied by a strong expansion when the samples were exposed to magnesium sulphate
solution. This can be attributed to the formation of gypsum from portlandite and the co-
precipitation of brucite (Mg(OH)2). [62] showed that the formation of gypsum in sodium
sulphate solution can lead to expansion and also to micro cracking. It was further
demonstrated that the expansion can be delayed and is observed in the late stage of the attack.
These results were confirmed by other studies [63- 66]. The formation of gypsum can be
avoided by the transformation of portlandite into C-S-H phases by a pozzolanic reaction prior
to sulphate attack [62]. The reaction of portlandite to gypsum does not only lead to expansion,
but can also result in softening of the surface of the exposed samples [28]. Biczok [67] cites

93
experimental studies which suggest that the formation of gypsum requires relatively high
sulphate concentrations when compared to the formation of ettringite. It is also noted that the
formation of gypsum is more destructive than the formation of ettringite at very high sulphate
concentrations. Such very high sulphate concentrations are rarely found under field condition.
But according to [68], soils with such high sulphate concentrations exist in some regions of
North America. In these alkaline soils, magnesium sulphate and alkalis sulphates are enriched
due to the transport of these salts from the ground water to the surface by evaporation. Very
high sulphate concentrations can also be due to oxidation of pyrite (FeS2). Nevertheless, such
extremely high sulphate concentrations are rare under field conditions. Despite of this, they
are used to accelerate the destructive process in laboratory test procedures [69-71], and
ASTM C 1012. It was shown that the formation of gypsum is favoured by the conditions in
the laboratory test procedures whereas under most field condition the formation of gypsum
can not be expected [72].
During a reaction of portlandite with concentrated alkali sulphate solution to gypsum,
the formation of easily soluble alkali hydroxide is observed as a by-product:

Ca (OH)2 + 2 Na+ + SO42- + 2 H2O CaSO4·2H2O + 2 Na+ + 2OH- (4.3.3)

The solubility of gypsum increases with the concentration of alkali hydroxide in the
solution [73]. This can explain the observation of [74] that the addition of alkali hydroxide to
sodium sulphate solution lowered the expansion of test samples when compared to plain
sodium sulphate solution.
Since the formation of gypsum requires the presence of portlandite in the hardened
cement paste, cements with high alite contents are more vulnerable to a destructive formation
of gypsum [65] because portlandite is a by-product of the reaction of alite with water to C-S-
H phases. In contradiction to this detrimental effect of a high alite content of the cement it has
to be considered that a high content of this clinker mineral results in high strength and low
permeability of the hardened cement paste which leads to an improved physical resistance
against sulphate attack.
The damage of hardened cement paste exposed to sulphate attack by the
transformation of portlandite into gypsum has led to the suggestion to restrict the alite content
in SRPC [38, 75-79]. Such limitations apply for example in Russia and Argentina.
The formation of ettringite and gypsum at high sulphate concentrations can be avoided
by the application of sulphate resisting Portland cement in combination with a mineral
admixture having pozzolanic properties (e.g. fly ash or blast furnace slag). The low content of
C3A in the cement limits the amount of ettringite that can be formed during sulphate attack.
The pozzolanic reaction of the mineral admixture lowers the amount of portlandite in the
microstructure and makes the hardened cement paste less vulnerable to the formation of
gypsum. In agreement with this idea, mixtures containing sulphate resisting Portland cement
and pozzolanic material showed the best performance under field conditions [40, 41, 45, 80].
In addition to the good chemical resistance there is also a good physical resistance against
sulphate attack due to the late reaction of the pozzolanic or latently hydraulic admixtures.

94
4.3.2.3 Mineral Admixtures
Experimental investigations have shown that the addition of mineral admixtures can
increase the sulphate resistance of concrete. Ground granulated blast furnace slag, coal fly ash
and silica fume are the most relevant among these materials. Others, such as metakaolin or
natural pozzolanas have only limited importance. Mineral admixtures can be added to the
clinker during the production of blended cements or during the production of concrete in the
ready-mix plant. A general overview of their properties and impact on the microstructure is
given for example by [31, 81, 82]. The impact of limestone on the formation of thaumasite is
discussed in section 4.3.3.
Ground granulated blast furnace slag can be used for the production of sulphate
resisting cements. It is recommended that the cement produced from this material and
Portland cement clinker contains more than 65% slag (CEM III/B). The resistance of blast
furnace slag cement can be increased by using higher replacement levels of slag and clinker
with a low content of C3A [83].
The general ability of coal fly ash to increase the sulphate resistance of concrete has
been demonstrated in many studies [38, 40, 44, 84, 85]. However, there are differences in fly
ash quality [39]. There is at present no agreement on a test procedure to predict the ability of a
given fly ash to increase the sulphate resistance of concrete. Several suggestions for testing
procedure exist. They deal with the chemical composition of the fly ash [86] or with the
mineralogical composition of the hardened cement paste before the exposition to sulphate
attack [87]. In most studies and field applications, coal fly ash with a low CaO-content is
used, whereas lignite fly ash with a high CaO-content from is rarely applied.
The addition of pozzolanic and latently hydraulic admixtures can increase the physical and the
chemical aspect of the sulphate resistance:
• During the pozzolanic reaction, portlandite is converted to additional C-S-H phases.
Since portlandite is important for the formation of gypsum, the amount of gypsum that
can be formed is decreased by the addition of pozzolanic materials [62, 84]. It has also
been discussed if the presence of portlandite has an influence on the formation of
ettringite [34, 54]. The aforementioned statements refer to the exposition to sodium
sulphate. If the concrete is exposed to magnesium sulphate, the addition of pozzolanic
material can have a detrimental effect on sulphate resistance. Magnesium sulphate
reacts with portlandite to gypsum and brucite (magnesium hydroxide). In the absence
of portlandite, the C-S-H phases are attacked. This can lead to a disintegration of the
hardened cement paste [28]. The higher reactivity, SiO2-content, and added amount of
the pozzolanic material and the lower the alite content in the cement, the more
portlandite can be converted into C-S-H phases. Latently hydraulic materials with a
high CaO content such as blast furnace slag have a lower ability to reduce the
portlandite content than pozzolanic materials with low CaO content such as coal fly
ash and silica fume. A pre-hydration period of at least 4-6 weeks is recommended
before exposition to sulphate attack if fly ash is used. During this period, a substantial
degree of hydration of the fly ash can be gained [88]. High temperatures during pre-
hydration favour a pozzolanic reaction and increase the sulphate resistance.
• The content of clinker minerals whose reaction products are vulnerable to sulphate
attack (e. g. aluminate and ferrite) is lowered by the addition of mineral admixtures
due to dilution. Some very fine mineral admixtures serve as a surface on which

95
additional C-S-H phases can grow. This leads to a higher relative degree of hydration
of the clinker phases and therefore the impermeability of the hardened cement paste
can be increased [89-91].
• The hydration of pozzolanic and latently hydraulic materials proceeds slower than the
hydration of Portland cement. Due to the low rate of reaction, micro-cracks and large
capillary pores can be sealed by the reaction product formed in the late stage. This
leads to higher diffusion coefficients of hardened cement paste containing pozzolanic
or latently hydraulic materials [84, 92, 93].
Beside the mentioned general effects, there are other details relevant to the resistance
to sulphate attack. Addition of silica fume without proper mixing with the cement can lead to
large inclusions of SiO2 [94] that can promote the formation of thaumasite. Furthermore, the
water demand of cement can increase by the addition of silica fume. Reactive aluminium
contained in fly ash can, under certain circumstances, participate in the formation of ettringite.
The use of fly ash reduces bleeding.

4.3.2.4 Microstructure
Changes in the microstructure and the phase composition during sulphate attack on
hardened cement paste, mortar and concrete have been studied extensively. In most of the
investigations, polished sections of cement paste and mortar were examined by SEM using
backscattered electron imaging and X-ray microanalysis after exposure to sulphate attack at
room temperature [95-102]. These studies demonstrate that ettringite, gypsum and brucite are
formed during sulphate attack. The chemical reactions are accompanied by the formation of
micro-cracks and leaching of calcium.
After exposure to highly concentrated test solutions in the laboratory, gypsum can be
found in the microstructure in form of small inclusions and also in form of veins running
parallel to the surface of the sample. These single- or multi-layers of gypsum have a thickness
between 10 and 30 m. Their distance to each other and to the surface varies between 100 and
200 m. More gypsum veins are found in magnesium sulphate solution than in sodium
sulphate solution. Micro cracks are observed to run parallel and perpendicular to the gypsum
bands. Parallel cracks are often observed at the edge between gypsum and the surrounding
matrix. Perpendicular cracks often run from the gypsum veins into the core of the test sample.
It is not possible to distinguish if the cracks are due to ettringite formation or to gypsum
formation because both processes occur almost simultaneously. Since the formation of
ettringite requires a lower sulphate concentration than the formation of gypsum, the second
mineral is formed only in a minor depth of the sample when compared to ettringite. In the
outer part of the specimens, where gypsum is formed, the Ca/Si-ratio of the C-S-H phases is
lowered to about 1.1, whereas in greater depths it retains its original value of about 1.7.
Portlandite is in most cases depleted in the area close to the surface being converted to
gypsum or leached from the microstructure. Thus, there is an increase of porosity in the
concerned region. The leaching of calcium is accompanied by the formation of a skin of
calcium carbonate on the surface of the sample. There are indications that the amount of
ettringite is reduced in the attacked region. This can be due to carbonation or to a reaction of
ettringite with sulphate ions to gypsum and aluminium hydroxide.
Below the region that is damaged by leaching and the formation of gypsum, C-S-H
phases, portlandite and ettringite are commonly identified. With respect to the undamaged

96
core of the sample, only AFm has reacted with sulphate ions to ettringite. No other transition
is detected, the Ca/Si-ratio of the C-S-H phases remaining at its original level. The depth of
the sample in which AFm has reacted to ettringite depends on permeability, concentration of
the aggressive solution, duration of exposition, degree of pre-hydration and also on
composition and fineness of the cement.
The formation of ettringite and gypsum is characterized by the fact that the volume of
the products is higher than the volume of the solid reactants. This increase in volume of solid
phase leads to a densification of the microstructure and a temporal increase of strength of the
sample [103]. However, after the formation of micro- and macro-cracks due to sulphate
attack, the strength is reduced below the strength of water stored samples [48, 104]. The
expansion of test samples is accelerated by the appearance of the first cracks [105, 106]. This
is especially relevant for sodium sulphate solution [28]. The expansion of the outer region by
the formation of ettringite and gypsum in the test samples results in the generation of a tensile
stress in the core of the samples [107]. The expansion is earlier detected in thinner samples
because the inner tensile core resisting the expansion is smaller [108].
There occurs also formation of a layer of brucite and gypsum on the surface of the
samples during storage in magnesium sulphate solution increasing the impermeability of the
specimens. The expansion of the samples leads to the formation of micro-cracks within the
paste and also within the protecting layer of brucite and gypsum.
The above described microstructural changes during sulphate attack were derived after
the examination of hardened cement pastes, mortars and concretes damaged by sulphate
attack under laboratory conditions. During the real exposition of concrete under field
conditions it has to be considered that cycles of wetting and drying can accelerate the
transport of sulphate ions. Furthermore, there can be additional transport due to evaporation of
water at the top of porous concrete elements. In such extreme cases, efflorescence of easily
soluble salts such as thenardite (Na2SO4) is reported [109, 110]. Another difference between
hardened cement paste and concrete is the presence of an interfacial transition zone (ITZ)
around the aggregates that possesses a higher porosity. Subsequently, an increased diffusion
of sulphate ions in the ITZ was reported [111] and sulphate phases can be preferentially
formed around aggregates. The ITZ is mechanically weaker than the surrounding matrix.
Cracks are readily formed during expansion in such weak parts of the microstructure [55].

4.3.2 Thaumasite form of sulphate attack (TSA)

Thaumasite may be formed through the reaction of water with calcium carbonate, any
sulphate salt, and hydrated calcium silicates, which are present in concrete or mortar mixtures
containing Portland cement. Bensted [112-114] proposed two alternative or parallel reaction
routes for the formation of thaumasite, the direct C-S-H route and the woodfordite route via
the formation of intermediate solid solution compounds, where the end members are ettringite
and thaumasite. The intermediate compounds form a discontinuous series and Barnett et al.
[115] identified a solid solution gap. According to [115] this indicates that there could be a
range of unstable compositions in the solid solution series. A low temperature of about 5°C
favours thaumasite formation, however, cases of its formation at temperatures at 20°C or
higher have been reported [5, 23, 116]. Investigations [17, 23, 117,] found evidence for a
more rapid deterioration at temperatures below 15 °C.

97
According to [112-114] sulphate attack, leading to thaumasite formation at
temperatures below 15°C, can be exemplified by the following equations:
Direct C-S-H-route: Ca3Si2O7⋅3H2O (C-S-H) + 2CaSO4⋅2H2O (gypsum) + 2CaCO3 (calcite) +
24H2O Ca6 [Si (OH)6]2(CO3)2(SO4)2⋅24H2O (thaumasite) + Ca(OH)2 (4.3.4)

Woodfordite route: Ca6[AlxFe(1-x) (OH)6]2(SO4)3⋅26 H2O (ettringite) + Ca3Si2O7⋅3H2O


(C-S-H) + 2CaCO3 (calcite) + 4H2O Ca6[Si(OH)6]2(CO3)2(SO4)2⋅24H2O (thaumasite) +
CaSO4⋅2H2O (gypsum) + 2⋅Al(OH)3 + 2(1-x)Fe(OH)3⋅4Ca(OH)2 (4.3.5)

The calcium hydroxide formed could eventually carbonate to calcite from interaction
with carbon dioxide. A comprehensive study by Köhler et al. [117] could not verify any
transformation of ettringite into woodfordite or vice versa. Woodfordite was only identified in
two cases out of 120. They suggested that this is a clear indication that thaumasite is not
formed via the woodfordite route. The authors proposed that thaumasite formation occurs
through nucleation on the surfaces of ettringite crystals. The direct formation of thaumasite
without ettringite seems to be a rare case as it requires low temperatures and highly
supersaturated solutions with high ionic concentrations. Investigations reported by [117] and
[21] could not find any thaumasite in samples where aluminium was absent. The sulphate
concentration in [117] was as high as 3000 mg/l SO42- and the temperature 5 °C. It has been
discussed that the Woodfordite route (considering a reaction of C-S-H phases with ettringite
and calcite) implies the formation of thaumasite without sulphate attack from the
environment. This is in contradiction with field experience [118]. Also, thermodynamic
calculations indicate that a formation of thaumasite via the Woodfordite route is not likely at
8°C [119].
According to [6], calcium hydroxide is a reactant rather than a reaction product. The
source of sulphate and carbonate necessary for the reaction can alternatively come from an
aggressive solution containing dissolved sulphate and carbonate ions (or from atmospheric
carbon dioxide). [6] proposed the following reaction in the presence of magnesium ions:

C3S2H3 + 3Ca(OH)2 + 2CaCO3 + 4MgSO4 + 32H2O = 2C3SCSH15 + 2CaSO4 x 2H2O


(gypsum) + 4Mg(OH)2 (4.3.6)

Thaumasite formation is very detrimental because it leads to the breakdown of C-S-H


and CH as well as the hydrated aluminate phases, C-A-H and ettringite, of concrete and a
significant strength loss of hardened concrete can then be noticed. The characteristics of its
formation as deterioration product in buried concrete, is to transform the surface of the
concrete into a white pulpy mass.

4.3.2.1 Physical resistance and capillary porosity


It is known that the rate of deterioration and the durability is directly related to the
permeability of concrete which depends on the w/c-ratio. This has been reported by [120]
among others and in the UK guidelines for concrete in aggressive ground [26] the w/c ratio is
specified for certain sulphate classes. Increasing the cement content could also have the effect

98
of reducing the degree of deterioration but not eliminating it [121]. The investigation reported
in [122] showed the effect of lowering the w/c-ratio on the strength of mortar specimens
exposed to a sodium sulphate solution. The specimens with decreased permeability (w/c ratio
= 0.35) showed significantly improved performance. According to [31, 94] this could be
explained by the reduction of pore size and the decreasing diffusion of sulphate ions into
concrete.
Similarly, the incorporation of large amounts of inert filler material, for example
limestone powder, into SCC may influence the microstructure [11, 12], the capillary pore size
distribution and the permeability. The effect may be negible, positive or negative depending
on the coupling between w/c-ratio, cement powder ratio (c/p), filler cement ratio (f/c) and
water powder ratio (w/p) [123, 124].
The influence which the incorporation of limestone filler in SCC may have on the
capillary pore size distribution is exemplified by figure 4.3.1 and table 4.3.1 below. Compared
with a traditional concrete specimen (TC), the SCC specimen containing 200 kg/m3 limestone
filler showed a reduction of the pore size. The samples had the same w/c ratio (0.45) and total
porosity measured by mercury intrusion porosimetry (MIP).

Table 4.3.1 Mix proportions of SCC and TC specimens in figure 4.3.1

Type w/p- w/c- f/c- Lime filler Cement Water Porosity


ratio ratio ratio (kg/m3) (kg/m3) (l/m3) (MIP, %)
CEM I
SCC 0.30 0.45 0.50 200 401 181 15.70
TC - 0.45 - - 430 194 15.40

Pore size distribution

0,07
TC
0,06
SCC
Pore volume [ml/g]

0,05

0,04

0,03

0,02

0,01

0,00
0,001 0,01 0,1 1 10 100 1000

Pore radius [ m]

Figure 4.3.1 MIP-measurements showing the capillary pore size distribution in SCC
compared with TC at w/c-ratio 0.45.

99
The interdependence between the w/c-ratio, w/p-ratio and f/c-ratio and the influence it
may have on the capillary pore size distribution and the permeability, is exemplified by water
absorption curves (figures 4.3.3 and 4.3.5) and chloride diffusion coefficients (table 4.3.2) of
six SCC mix proportions. The pore size distributions are given in figures 4.3.2 and 4.3.4.

Table 4.3.2 Mix proportions, porosity and chloride diffusion of SCC samples shown in
figures 4.3.2 and 4.3.5

SCC w/p- w/c- f/c- Water Cement Lime MIP- D, chloride


mix ratio ratio ratio (l/m3) (kg/m3) filler Porosity (x 10-12
CEM I (kg/m3) (%) m2/s)
1 0.24 0.40 0.70 150 375 262 11.81 3.08
2 0.21 0.40 0.89 140 350 312 10.72 3.04
3 0.19 0.40 1.07 132 330 352 12.06 2.64
4 0.28 0.44 0.58 165 375 217 14.97 3.48
5 0.28 0.47 0.69 165 350 240 15.27 4.33
6 0.28 0.50 0.78 165 330 258 15.13 4.37

SCC mixes 1-3 had a constant w/c ratio of 0.40 but gradually decreasing w/p ratios
and increasing f/c ratios. As seen in figure 4.3.2 it resulted in a reduction of the pore sizes.
Although SCC mix 3 had the largest total porosity, the increased amount of limestone filler
created a finer capillary pore system. As a result, the chloride diffusion coefficient (table
4.3.2) was decreased suggesting a lower diffusion rate of ions into the concrete. The
difference between SCC mix 1 and 2 is marginal, indicating that a threshold value of the f/c-
ratio must be reached in order to achieve a significant reduction of the pore size.

Pore size distribution

0,07 3) f/c-ratio = 1.07


0,06
Pore volume [ml/g]

0,05
2) 0.89
0,04

0,03
1) 0.70
0,02

0,01

0,00
0,001 0,01 0,1 1 10 100 1000
Pore radius [ m]

Figure 4.3.2 Pore size distribution in SCC mixes 1-3 determined by MIP. Constant water-
cement ratio (0.40). For details see table 4.3.2.

100
capillary suction

1,8
water uptake (kg/m2) 1,6 1
1,4 2
1,2 3
1
0,8
0,6
0,4
0,2
0
0 5 10 15 20 25 30
square root of time (hours)

Figure 4.3.3: Water absorption curves for SCC mixes 1-3.

The finer pore size distribution with increasing f/c-ratio at constant w/c-ratio and
decreasing w/p ratio in SCC mixes 1-3 and the following increasing resistance to water
penetration, finds support by the water absorption curves shown in figure 4.3.3. Again, the
significant difference is between SCC mix 3 and the other two mixes.
In SCC mixes 4-6 (table 4.3.2) the w/p ratio was kept constant while the w/c- and f/c-
ratios gradually increased. In this case, a coarser pore size distribution is noticed from SCC
mix 4 to 5 due to the increased w/c ratio.

P ore size distribution

0,14 6) f/c-ratio = 0.78


0,12
Pore volume [ml/g]

0,10
5) 0.69
0,08 4) 0.58
0,06

0,04

0,02

0,00
0,001 0,01 0,1 1 10 100 1000
Pore radius [ m ]

Figure 4.3.4 Pore size distribution in SCC mixes 4-6 determined by MIP. Constant w/p-ratio
and increasing w/c- and f/c-ratios.

101
This is however not the case changing from mix 5 to 6, where an increased pore
volume is noticed instead, whiles the pore size distribution is about the same. The effect could
be attributed to the increased amount of limestone filler in mix 6 and therefore a higher f/c-
ratio. Due to the increased pore volume in mix 6, a higher diffusion rate was measured (table
4.3.2) and a higher water absorption rate (figure 4.3.5).

capillary suction

3
6
2.5 5
water uptake (kg/m2)

4
2

1.5

0.5

0
0 5 10 15 20 25 30
square root of time (hours)

Figure 4.3.5 Water absorption curves for SCC mixes 4-6.

The influence of w/p-ratio and f/c-ratio on sulphate attack and thaumasite formation
with magnesium- or mixed magnesium-sodium sulphate solutions has been treated by some
authors. In the case of attack on concretes and mortars by magnesium sulphate, or even mixed
magnesium and sodium sulphate solutions, some investigations report a more detrimental
effect when w/c-ratio or w/b-ratio is decreased [40, 122, 125]. This means, somewhat
unexpected, that a finer pore size distribution and a denser microstructure could promote
degradation when exposed to magnesium-rich sulphate solutions but not, according to [122],
when exposed to sodium-rich sulphate solutions. [122] attributed this effect to the limited
pore space for the expansive reaction products to occupy, which in this case was gypsum, M-
S-H and ettringite.
Evidence on the contrary has been reported by [14] who investigated SCC specimens
containing different amounts of limestone filler exposed to a weak-moderately strong
magnesium sulphate solution (1400 mg/l SO42-). They observed less deterioration correlating
with denser microstructure and finer capillary pore size. The reaction products formed were
primarily thaumasite and gypsum. The two SCC mixes in the study had the same w/c ratio
(0.40), but one of the SCC mixes (8 in table 4.3.3) contained a larger amount of limestone
filler and consequently had a lower w/p ratio and higher f/c ratio (see table 4.3.3). From water
absorption curves the resistance to capillary transport was calculated and, as shown in table
4.3.3, mix 8 had the higher resistance and finer capillary pore sizes.

102
Table 4.3.3 Mix proportions and resistance to capillary transport for SCC mixes 7 and 8.

SCC w/p- w/c- f/c- Lime filler Cement Water Resistance to


mix ratio ratio ratio (kg/m3) (kg/m3) (l/m3) capillary transport,
CEM I (x 106 s/m2)
7 0.33 0.40 0.22 90 415 166 73,8
8 0.28 0.40 0.43 180 415 166 124,1

The diagram in figure 4.3.6 shows degradation versus time for the two SCC mixes.
SCC mix 8 showed no signs of degradation while SCC mix 7 (higher w/p ratio) degrades due
to thaumasite and gypsum formation.

Degradation vs. time

6 7) w/p = 0. 33
f/c = 0. 22
5
Wear rating (mm)

3
8) w/p = 0. 28
2 f/c = 0. 43

0
0 500 1000 1500 2000
Exposure time (days)

Figure 4.3.6 Degradation vs. time for the two SCC mixes in table 4.3.3. Degradation was
measured as mm loss from corners. Chemical composition of the cement is given in table
4.3.5

4.3.2.2 Chemical resistance - influence of cement type


According to the reactions put forward by [6,112] and others, the thaumasite form of
sulphate attack is able to consume not only the hydrating aluminate phases (C3A) and ferrite
phases (C4AF) but also the hydrated C-S-H phases. This is different from the classical
sulphate attack, which primarily attacks the hydrating C3A (C-A-H).
Investigations carried out by [19, 21, 117, 121, 126] indicate a strong correlation
between the C3A and aluminium content of the cement and the amount and rate by which
thaumasite is formed. [21] demonstrated by laboratory experiments that the amount of
thaumasite is directly proportional to the total amount of Al2O3. [117] prepared synthetic
suspensions based on alite clinker with gypsum, calcite and various amounts of C3A. After

103
one year storage in water, samples without aluminate showed total absence of thaumasite as
well as ettringite, while samples containing small amounts of aluminate only showed traces of
thaumasite and samples which had the highest aluminate content (3.7 %) showed thaumasite
formation accompanied by ettringite already after 2 weeks. The authors concluded that the
formation of thaumasite without ettringite is extremely slow or unlikely. They found that
ettringite controls the rate of thaumasite formation.
The studies reported in [19, 121] concluded that the faster rates of deterioration due to
thaumasite formation coincided with increased C3A contents. The studies also showed that
SRPC can be susceptible to the thaumasite form of sulphate attack.
The above is illustrated by results reported in [13] where SCC containing two
different Portland limestone cements with different C3A contents where investigated. Figures
4.3.7 and 4.3.8 show that the faster rate of deterioration due to TSA occurred in samples
which contained the cement with the highest C3A content. Mix proportions and composition
of the cements are given in table 4.3.4 and 4.3.5. The rates of deterioration were measured as
the sum of mm loss from corners during 1500 days of exposure to a 1400 mg/l SO42-
magnesium sulphate solution. Two series of concrete cube samples were tested, one without
extra limestone filler and one with 50 kg/m3 limestone filler. In each series, the w/c-ratio,
w/p-ratio, filler/cement ratio and cement content were kept constant. The limestone filler was
added at the expense of the fine aggregate content.
In the series without extra limestone filler (only CEM II A/L), a viscosity agent was
used in order to obtain a proper SCC. The results indicated that the degradation was worse in
the case were limestone filler was added but it should be noted that severe TSA was also
observed on the samples only containing Portland limestone cements with 10-12 % calcite.

Degradation vs. time


Type 1
30

25
Wear rating (mm)

Type 2
20

15

10

0
0 500 1000 1500 2000
Exposure time (days)

Figure 4.3.7 Degradation vs. time for SCC mixes 1 and 2 in table 4.3.4. Portland limestone
cements containing different amounts of C3A. From [13].

104
Degradation vs. time

30
25
Type 1
Wear rating (mm)

20

15 Type 2
10
5
0
0 500 1000 1500 2000
Exposure time (days)

Figure 4.3.8 Degradation vs. time for SCC mixes 3 and 4 in table 4.3.4. Portland limestone
cements containing different amounts of C3A. From [13].

Table 4.3.4: Mix proportions of SCC specimens in figure 4.3.7 and figure 4.3.8.

SCC w/p- w/c- f/c- Lime Blended Water


mix ratio ratio ratio filler limestone cement (l/m3)
(kg/m3) CEM II/A-L
type 1 and 2
(kg/m3)
1 0.56 0.65 0.17 50 300 (type 1) 195
2 0.56 0.65 0.17 50 300 (type 2) 195
3 - 0.65 - 0 300 (type1) 195
4 - 0.65 - 0 300 (type 2) 195

105
Table 4.3.5: Composition of Portland limestone cements in figure 4.3.7 and 4.3.8 (type 1 and
2). CEM I is the composition of CEM I in table 4.3.3. From [13, 14].

Analysis,wt % CEM I SR/LA CEM II/A-L CEM II/A-L


oxide (type 1) (type 2)
CaO 64.2 62.0 63.7
SiO2 22.1 18.5 19.6
Al2O3 3.5 5.0 3.6
Fe2O3 4.4 2.9 2.8
MgO 0.85 1.2 2.7
Na2O 0.06 0.1 0.23
K 2O 0.57 1.2 0.92
SO3 2.3 3.6 3.5
Component, wt %
Calcite - 10.9 10.3
(interground)
C3A 1.8 8.0 4.6
C4AF 13.2 8.7 9.0
C3S 55.6 63.3 59.9
C2S 21.4 5.7 13.6
Specific area 316 469 474
(Blaine), m2/kg

4.3.2.3 Influence of pozzolan materials


Incorporation of pozzolan materials into concrete such as silica fume, fly ash and blast
furnace slag is an effective way to densify the microstructure and reducing the capillary pore
size. In most cases this physical effect is slowing down the rate of deterioration significantly.
However these materials contain reactive components such as silica and aluminium which
could provide a source for reactants in thaumasite formation, if the physical effect is not
considered. A somewhat contradictory result has been reported in the literature.
In studies reported by [21, 117] on cement pastes mixed with fly ash, gypsum and
calcite placed in cold water found non or only minor detrimental effect of fly ash on
thaumasite formation. [21] reported that when micro silica or slag was mixed into the pastes
instead of fly ash they found that these admixtures accelerated thaumasite formation and the
pozzolanic reaction was retarded. However, when nanosilica was incorporated instead of
microsilica, a complete pozzolanic reaction took place and as a result no thaumasite formation
occurred. They attributed this to the fact that the total amount of calcium hydroxide was
bound due to the pozzolanic reaction.
A study reported by [20] on concretes and mortars stored in cold sulphate solutions
concluded that OPC and Portland limestone cements blended with fly ash did not escape the
thaumasite form of sulphate attack. [21] concluded that a longer pre-storage time, which is
allowing more time for the pozzolanic reaction, hinders the attack but after a while the
specimens deteriorated at a fast rate.

106
The investigation reported in [126] studied the influence of fly ash, silica fume and
slag on the sulphate resistance of limestone cement pastes immersed in cold magnesium
sulphate solution. It was found that addition of silica fume to limestone cement significantly
improved its sulphate resistance. Fly ash and slag did not significantly change the sulphate
resistance of the limestone cement pastes.
From a survey on different concretes containing limestone aggregates [19, 121] it was
concluded that blast furnace slag cements with 70 and 90 % replacement of Portland cement
did not suffer thaumasite attack after three years of exposure to cold magnesium and sodium
sulphate solutions, while concretes with SRPC and Portland cements blended with fly ash (25
% and 40 % replacement levels) degraded due to thaumasite formation.
From the chemical point of view, the silicon-rich C-S-H phases resulting from the
reaction of pozzolanic or latently hydraulic materials are more resistant than calcium-rich C-
S-H phases present in hydrated Portland cement. It was derived by means of thermodynamic
calculation, that concrete can be protected at low sulphate concentrations against the
formation of thaumasite by the addition of a sufficient amount of pozzolanic or latently-
hydraulic material [127].

4.3.2.4 Influence of the composition of the sulphate-rich solution


Sulphate attack on concrete is primarily attributed to sodium-, potassium-,
magnesium- and calcium sulphates. Due to the limited solubility of calcium sulphate in water
at normal temperatures, sulphate attack is normally caused by increased concentrations of
magnesium- and/or sodium/potassium sulphate [122]. The solubility of calcium sulphate at
normal temperatures is approximately 1400 mg/l SO42-.
The solubility of calcium compounds (eg. calcium carbonates) is increasing with
decreasing temperatures which could be one of the factors controlling the effectiveness of
thaumasite formation at low temperatures.
Magnesium sulphate salt, MgSO4, appears to be more detrimental than sodium
sulphate salt, Na2SO4, due to the conversion of portlandite, Ca(OH)2, to gypsum and brucite,
Mg(OH)2, which leads to breakdown of the C-S-H gel [6, 116]. When magnesium-rich
sulphate solutions attack sulphate resistant Portland cements (SRPC), with low C3A and alkali
contents, where pH in the pore solutions is dominated by Ca(OH)2 at about 12.5, the pH will
drop to approximately 10-11 near the surface and magnesium will precipitate as brucite at the
expense of calcium hydroxide according to the reaction:

Mg2+ + SO42- + Ca(OH)2 + 2H2O = CaSO4 2H2O + Mg(OH)2 (4.3.7)

If instead SRPC is exposed to sodium-rich sulphate solutions, the pH will increase (>12.5 -13)
and the following reaction will take place described by [35]:

2Na+ + SO42- + Ca(OH)2 + 2H2O = CaSO4 2H2O + 2 Na+ + 2 OH- (4.3.8)

Gypsum will react and convert to ettringite due to the high pH-levels. According to
[128] the stability of thaumasite is affected by the pH of the solution. At high pH levels (
12.5-13.0) thaumasite is stable and at low pH levels, it reacts with the ions present in the
solution [128]. In the case where low alkali SRPC containing limestone filler is exposed to a
magnesium-rich sulphate solution, gypsum rather than thaumasite would be the predominant

107
reaction product when pH eventually drops to about 10 near the surface. At early stages of the
attack thaumasite may be formed but would destabilise as the attack continues. Evidence
which supports this reaction has been reported by [14].
However, if cements high in alkalis and C3A, for example some limestone Portland
cements, are exposed to a magnesium-rich sulphate solution, the risk of TSA is increased due
to the longer time span required before pH drops below the stability of thaumasite. Severe
TSA on limestone Portland cements by magnesium sulphate solutions has been reported by
[6, 13, 23] and others. In this case, magnesium-rich sulphate solutions seem to be more
detrimental than sodium-rich solutions [6].
In Norway sprayed concrete linings in tunnels and mortar prisms stored in tunnels
containing limestone filler has been reported to deteriorate due to TSA from sulphate-bearing
waters from alum shale [22]. The water had a sulphate concentration of about 200-300 mg/l
SO3 (pH 3-4). In Germany one of the destructive mechanisms to 30 year old shotcrete linings
and in-situ cast concrete members in a Bavarian tunnel was found to be thaumasite formation
[129]. Thaumasite was formed in spite of low sulphate concentrations in the water which was
in contact with the concrete, about 20 mg/l SO42- (pH 7.8). The concretes were permeable and
carbonated and of a rather low quality considering the environment. Oxidation of pyrite in
carbonated shotcrete tunnel linings could cause similar degradation reported by [136]. In a
survey reported in [17, 121] it was found that the degree of TSA in limestone filled mortars
did not depend on the concentrations of sulphate ions in the magnesium sulphate solutions if
the solutions were replenished simulating mobile water conditions. On the contrary, when
stagnant water conditions were simulated (solutions not replenished) a correlation between
sulphate concentration and the degree of attack were found. The above suggests that if the
right conditions are met (e.g. percolation and constant supply of sulphate ions), TSA can
occur in concrete and mortars in contact with waters having a low sulphate concentration and
the physical resistance against ingress of sulphate ions is important.
Seawater attack on bedding mortar and concrete in harbour wall steps, Wales, has
been reported to produce early stages of TSA, showing that seawater is capable of inducing
TSA in short period of time, about 3 years [130]. The sulphate concentration of the seawater
was about 4000 mg/l SO42- (0.4 %). The mortar had a cement/lime ratio of about 1:1.

4.3.4 Potential for delayed ettringite formation in scc compared with tc


considering hydration heat production

The incorporation of large amounts of limestone filler in SCC could lead to a


somewhat higher heat of hydration and a higher cumulated heat production in SCC than in TC
[12, 131]. As a consequence of this, temperatures higher than 65 ºC could be obtained within
the hardening concrete, which could result in the formation of delayed ettringite. Results from
[12, 131] indicate that limestone filler could contribute with additional reaction heat and that
the maximum heat production rate is dependent on the cement/powder ratio. The fineness and
the grain-size distribution of the used limestone filler probably affect the heat of hydration as
well. However, [131] showed that the incorporation of quartz filler instead of limestone filler
in SCC did not generate additional hydration heat. Another factor which also could influence
the heat of hydration to a greater extent in SCC than in TC, is the type and amount of
chemical admixtures.

108
A data program for calculating temperature development in concrete members has
been developed by JEJMS Concrete AB and Luleå Technical University in co-operation with
the construction industry in Sweden [132, 133]. The data program is based on adiabatic and
semi-adiabatic hydration tests on commonly used SCC and TC for civil engineering, obtained
by the Luleå Technical University. Apart from the calorimetric tests, results from
compressive-, pull-, creep- and shrinkage tests were obtained.
In the following section an example of the heat development in a SCC and a
comparable TC is given from two field concrete mix designs for civil engineering purposes.
The objective was to evaluate if SCC containing limestone filler has a larger potential for
developing delayed ettringite formation (temperatures > 65 ºC) than a comparable TC,
considering the maximum heat production. The heat production is exemplified by a wall
construction with five different thicknesses. The mix designs of the SCC and TC are given in
table 4.3.6 below. The materials used in the two concretes were the same. The comparison is
performance–based which means that the cement content and w/c-ratio were allowed to
deviate between the concretes in order to achieve approximately the same 28 day compressive
strength. In spite of lower cement content and higher w/c-ratio, the SCC reached a higher 28
days compressive strength (see table 4.3.6). The development of the SCC mix design (w/c-
ratio = 0.47) as a replacement for TC (w/c-ratio = 0.44) has been further described in [134,
135].

Table 4.3.6 Mix design proportions of the SCC and TC in figure 4.3.9.

Components TC SCC
Aggregates [kg/m3] 1765.9 1555.0
CEM I [kg/m3] 395 364
Limestone Filler [kg/m3] - 211
Superplasticizer [kg/m3]
- Sikament BV40 0.79
- Glenium 51 1.11
Air Entraining Agent [kg/m3] 0.198 0.042
Water [kg/m3] 172 171
Air-void Content [%] 4.0 4.8
Slump Flow [mm] - 740
w/c 0.44 0.47
w/p 0.44 0.30
c/p 1.0 0.63
Compressive Strength (28 d) 57 68
[Mpa]

The maximum hydration heat production and heat production rates are shown in
figure 4.3.9 below. From figure 4.3.9, it is concluded that the SCC has a faster hydration heat
development which is in accordance with results obtained by [131]. In thinner wall
geometries (0.2-0.8 m) a slightly higher maximum heat production is calculated in SCC, but

109
when the concrete mass is further enlarged due to an increased wall thickness of 3.2 m the TC
shows a higher maximum heat production than the SCC. This is explained by the higher
cement content in the TC. The example shows that the difference in maximum heat
production between a SCC with limestone filler and a performance-based equivalent TC is
marginal in the region of 65 ºC (max. 66 ºC for TC and max. 64.5 ºC for SCC). In fact, a
further reduction of the cement content in the SCC could be allowed considering the higher 28
day compressive strength.

70

60

50
d = 3.2 m
Temperature [°C]

40
TC
Serie1
SCC
Serie6
30
d = 0.4 m
d = 0.8 m d = 1.6 m
d = 0.2 m
20

10

0
0 100 200 300 400 500 600
Time [h]

Figure 4.3.9 Heat production rate in two comparable SCC and TC mix proportions,
exemplified by five different thicknesses of a wall geometry (0.2 – 3.2 m). The mix
proportions of the two concretes are given in table 4.3.6.

From this it is concluded that an increased potential for delayed ettringite formation in
SCC with limestone filler compared with TC is small, because any extra reaction heat from
limestone filler or superplastisizers in SCC is overshadowed by dominant factors such as
cement content, type of cement and volume of concrete. However, if the cement content is the
same in SCC and TC and the SCC contains high amounts of limestone filler, an increased risk
for DEF in SCC is possible considering that the maximum heat production is increasing with
decreasing c/p-ratio, which has been shown by [131].

110
4.3.5 Conclusions

The following conclusions can be made:

• The initiation time for sulphate attack is prolonged in SCC compared with an
equivalent TC due to a denser microstructure in SCC. A finer capillary pore system is
obtained in SCC due to incorporation of filler materials which results in a better ability
to withstand sulphate attack. The capillary pore size distribution, connectivity and total
capillary porosity in SCC is dependant on factors such as w/p-, c/p-, f/c-ratios.
Interdependence between these factors exists regarding the nature of the capillary pore
size distribution.
• The deterioration rate and grade during sulphate attack are influenced by the C3A
content of the cement. The incorporation of high amounts of limestone filler in SCC
makes it more vulnerable to thaumasite form of sulphate attack (TSA). In the case of
TSA, ettringite is commonly formed as a precursor mineral and nucleus for subsequent
thaumasite formation. Direct formation of thaumasite without ettringite seems to be a
rare case requiring supersaturated solutions with high ionic concentrations and low
temperatures. SRPC does not make SCC immune against TSA but the nature of attack
is different, the rate is much slower and the degree of attack is smaller.
• Incorporation of pozzolan materials into concrete such as silica fume, fly ash and blast
furnace slag is an effective way to densify the microstructure and reducing the
capillary pore sizes. In most cases this physical effect is slowing down the rate of
deterioration significantly. Furthermore, the calcium hydroxide is consumed during
the reaction of these materials and the calcium-silicon ratio of the C-S-H phases is
decreased making these phases more resistant against the formation of thaumasite. On
the other hand, pozzolanic and latently-hydraulic materials contain reactive
components such as silica and aluminium which could provide a source for reactants
in ettringite and thaumasite formation, Contradictory results have been reported in the
literature regarding TSA and pozzolan materials. Further research is needed in this
area.
• In the case of TSA, a correlation exists between sulphate concentration and the degree
of TSA in concretes with limestone filler, if stagnant water conditions are prevailing.
In this case sulphate concentrations somewhere between 700 and1400 mg/l MgSO4 are
able to promote TSA in approximately five years time. On the other hand, if mobile
water conditions are prevailing, field evidence reported in the literature indicates that
no such correlation exists. The evidence suggests that if the right conditions are met
(e.g. percolation and constant supply of sulphate ions), TSA can occur in concrete and
mortars in contact with waters having a low sulphate concentration and that the
physical resistance against ingress of sulphate ions is important. Sulphate solutions
which are dominated by magnesium ions seem to be more detrimental than sodium or
calcium dominated solutions.

111
• From a performance-based comparison between SCC and TC, it is concluded that an
increased risk concerning delayed ettringite formation in SCC with limestone filler is
small. This comes from the fact that any extra heat production from limestone filler or
superplastisizers in SCC is overshadowed by dominant factors such as cement content,
type of cement and volume of concrete. However, in the case where the cement
content is equal in SCC and TC and the SCC contains high amounts of limestone
filler, an increased risk for DEF in SCC is possible considering that the maximum heat
production is increasing with decreasing c/p-ratio.

References

1. Building Research Establishment, Sulphate and acid resistance of concrete in the ground BRE
Digest 363, 1996.
2. Department of Environment, Transport and the Regions: The thaumasite form of sulphate attack:
risks, diagnosis, remedial works and guidance on new construction, Report of the thaumasite
expert group. DETR, January 1999.
3. ACI Committee 201: Durable Concrete, ACI Manual of Concrete Practice, Part 1, Detroit,
Michigan, 1992, 39 p.
4. Collepardi, M.: Ettringite formation and sulfate attack on concrete. In: Malhotra, V. M. (editor):
Proceedings of the Fifth CANMET/ACI International Conference on Recent Advances in
Concrete Technology (ACI SP-200), 2001, p. 21-37.
5. Collepardi, M.: Thaumasite formation and deterioration in historic buildings. In: Cement and
Concrete Composites, Volume 21 (1999), p. 147-154.
6. Hartshorn, S. A.; Sharp, J. H.; Swamy, R. N.: Thaumsite formation in Portland-limestone cement
pastes. In: Cement and Concrete Research 29 (1999), p. 1331-1340.
7. Hime, W. G.; Mather, B.: Sulfate attack, or is it? In: Cement and Concrete Research, Volume 29
(1999), p. 789-791.
8. Barnett, S. J.; Halliwell, M. A.; Crammond, N. J.; Adam, C. D.; Jackson, A.R.W., Study of
thaumasite and ettringite phases formed in sulfate/blast furnace slag slurries using XRD full
pattern fitting. In: Cement and Concrete Composites, Volume 24 (2002), p. 339-346.
9. Mehta, P. K.: Sulfate attack on concrete – A critical review. In: Materials Science of Concrete III.
The American Ceramic Society, 1993, p. 105-130.
10. Brown, P. W.: Thaumasite formation and other forms of sulfate attack. In: Cement and Concrete
Composites, Volume 24 (2002), p. 301-303.
11. Trägårdh, J.: Microstructural Features and Related Properites of Self- Compacting Concrete. In:
Proceedings of the 1st International Conference on Self-Compacting Concrete, Stockholm, 1999,
p. 175-186.
12. Ye, G.; Poppe, A. M.; De Schutter, G.; van Breugel, K.: Numerical simulation of the hydration
process and the development of microstructure of self-compacting cement paste containing
limestone as filler, In: Material and Structures, in press.
13. Trägårdh, J.; Kalinowski, M.: Investigation of the Conditions for a Thaumasite Form of Sulfate
Attack in SCC with Limestone Filler. In: Proceedings of the 3rd International Conference on Self-
Compacting Concrete, Reykjavik 2003, p. ,.
14. Kalinowski, M.; Trägårdh, J.: Thaumasite and gypsum formation in SCC with sulfate resistant
cement exposed to a moderate sulfate concentration. In: Proceedings of the 2nd North American
Conference on the Design and Use of Self-Consolidating Concrete and the 4th International
RILEM Symposium on Self-Compacting Concrete, Chicago 2005, p. 319-327.

112
15. Nehdi, M.; Pardhan, M.; Koshowski, S.: Durability of self-consolidating concrete incorporating
highvolume replacement composite cements. In: Cement and Concrete Research, Volume 34
(2004), p. 2103-2114.
16. Persson, B.: Sulphate resistance of self-compacting concrete. In: Cement and Concrete Research,
Volume 33 (2003), p. 1933-1938.
17. Halliwell, M.; Crammond, N.; Barker, A.: The thaumasite form of sulfate attack in limestone-
filled cement mortars. Building Research Establishment Report BR 307, 1996.
18. Halliwell, M; Crammond, N.: Avoiding the thaumasite form of sulfate attack: two year report.
Building Research Establishment Report BR 385, 2000.
19. Crammond, N; Halliwell, M.: The thaumasite form of sulfate attack in laboratory prepared
concretes. Building Research Establishment Report BR 306, 1996.
20. Lipus, K.; Sylla H.-M.: Investigations in Germany of the thaumasite form of sulphate attack. In:
Proceedings of the First International Conference on Thaumasite in Cementitious Materials.
Building Research Establishment, 2002.
21. Nobst, P.; Stark, J.: Investigations on the influence of cement type on the thaumasite formation.
In: Proceedings of the First International Conference on Thaumasite in Cementitious Materials.
Building Research Establishment, 2002.
22. Justnes, H.: Thaumasite formed by sulfate attack on mortar with limestone filler. In: Proceedings
of the First International Conference on Thaumasite in Cementitious Materials. Building
Research Establishment, 2002.
23. Hartshorn, S. A.; Sharp, J. H.; Swamy, R. N.: The thaumasite form of sulfate attack in Portland-
limestone cement mortars stored in magnesium sulfate solution. In: Cement and Concrete
Composites, Volume 24 (2002), p. 351-359.
24. Tammimi, A.; Sonebi, M.; Tagnit-Hamou, A.; Saric-Coric, M.: Durability investigation of self-
compacting concrete using scanning electron microscope. In: Proceedings of the Second IMS
Conference on Applications of traditional and High Perfomance materials in harsh environment,
Jan. 17-18, 2006, Dubai, UAE, 24 p.
25. Bellmann, F., The formation of the mineral thaumasite resulting from sulphate attack on concrete,
PhD thesis 2006, Bauhaus University Weimar, Germany.
26. BRE Special Digest 1: Concrete in aggressive ground, Part 1-4. Part 1: Assessing the aggressive
chemical environment. Building Research Establishment, 2001.
27. British Standard Institution, BS 8500:2002, Concrete – Complementary British standard to BS
EN 206-1. Part 1: method of specification and guidance for specifiers, Part 2: Specification for
constituent materials and concrete.
28. Santhanam, M.; Cohen, M. D.; Olek, J.: Sulfate attack research- whither now. In: Cement and
Concrete Research, Volume 31 (2001), p. 845-851.
29. Friebert, M.: Der Einfluss von Betonzusatzstoffen auf die Hydratation und Dauerhaftigkeit
selbstverdichtender Betone. PhD thesis, Bauhaus-University Weimar, 2005.
30. Smolczyk, H.-G.: Die Ettringit-Phasen im Hochofenzement. In: Zement-Kalk-Gips, Volume 14
(1961), p. 277-284.
31. Taylor, H.F.W.: Cement chemistry. – 2nd edition. - London: Thomas-Telford, 1997.
32. Mehta, P. K.; Hu, F.: Further evidence for expansion of ettringite by water absorption. In: Journal
Of the American Ceramic Society, Volume 61 (1978), p. 179-181.
33. Mehta, P. K.; Wang, S.: Expansion of ettringite by water absorption. In: Cement and Concrete
Research, Volume 12 (1982), p. 121-122.
34. Brown, P. W.; Taylor, H. F. W.: The role of ettringite in external sulfate attack. In: Marchand, J.;
Skalny, J. (ed.): Materials science of concrete – Special Volume: Sulfate attack mechanisms.
Westerville (Ohio): American Ceramic Society, 1999, p. 73-97.
35. Candlot, E.: Sur les propriétés des produits hydraulique. In: Bulletin de la Societé d’
Encouragement pour l’ Industrie Nationale, Volume 89 (1890), p. 685-716.

113
36. Thorvaldson, T.: Chemical aspects of the durability of cement products. In: Third International
Symposium on the Chemistry of Cement, 15.9.-20.9.1952, London, p. 436-466.
37. Miller, D. G.; Manson, P. W.: Long-time tests of concretes and mortars exposed to sulfate waters.
Technical Bulletin 194. Minnesota: University of Minnesota- Agriculture Experiment Station,
1951.
38. Monteiro, P. J. M.; Kurtis, K. E.: Time to failure for concrete exposed to severe sulfate attack. In:
Cement and Concrete Research, Volume 33 (2003), p. 987-993.
39. Kurtis, K. E.; Monteiro, P. J. M.; Madanat, S. M.: Empirical models to predict concrete expansion
caused by sulfate attack. In: ACI Materials Journal, Volume 97 (2000), p. 156-161.
40. Kalousek, G. L.; Porter, L. C.; Benton, E. J.: Concrete for long-time service in sulphate
environment. In: Cement and Concrete Research, Volume 2 (1972), p. 79-89.
41. Kalousek, G. L.; Porter, L. C.; Harboe, E. M.: Past, present, and potential developments of
sulphate resisting concretes. In: Journal of Testing and Evaluation, Volume 4 (1976), p. 347-354.
42. Hill, E. D.: A note on the history of Type V cement development. In: Marchand, J.; Skalny, J.
(ed.): Materials science of concrete – Special Volume: Sulfate attack mechanisms. Westerville
(Ohio), American Ceramic Society, 1999, p. 207-210.
43. Hughes, C. A.: The durability of cement mortars- the cement and method of testing major
variables. In: Annual meeting of the American Society for Testing and Materials, Band 33 (1933),
p. 511-537.
44. Stark, D.: Longtime study of concrete durability in sulfate soils. In: George Verbeck Symposium
on sulfate attack on concrete (ACI Special publication, 77). Detroit: American Concrete Institute,
1982, p. 21-40.
45. Harboe, E. M: Longtime studies and field experiences with sulfate attack. In: George Verbeck
Symposium on sulfate attack on concrete. (ACI Special publication, 77). Detroit: American
Concrete Institute, 1982, p. 1-20.
46. Merriman, T.: Durability of Portland cement. In: Engineering News Record, Volume 104 (1930),
p. 62-64.
47. Mather, B.: Field and laboratory studies of the sulphate resistance of concrete. In: Swenson, E. G.
(ed.): Performance of concrete- A Symposium in honour of Thorbergur Thorvaldson. Toronto:
University of Toronto Press, 1968, p. 68-76.
48. Dahl, L. A.: Cement performance in concrete exposed to sulfate soils. In: Journal of the American
Concrete Institute, Volume 46 (1950), p. 257-272.
49. Schröder, H. T.; Hallauer, O.; Scholz, W.: Beständigkeit verschiedener Betonarten im
Meerwasser und in sulfathaltigem Wasser. (Deutscher Ausschuß für Stahlbeton, Heft 252).
Berlin: Verlag Wilhelm Ernst & Sohn, 1975.
50. Glasser, F. P.; Kindness, A.; Stronach, S. A.: Stability and solubility relationships in AFm Phases.
Part I. Chloride, sulfate and hydroxide. In: Cement and Concrete Research, Volume 29 (1999), p.
861-866.
51. Neubauer, J.; Götz-Neunhoeffer, F.; Lindner, I.: Untersuchung des Hydratationsverhaltens
synthetischer Kalziumaluminatferrate bei Anwesenheit von C3A, CA und Kalziumsulfat. In:
Proceedings of the 4. Tagung Bauchemie, 30.9.-1.10.2002, Weimar, p. 13-19.
52. Chen, Y.; Shi, L.: Rate of ettringite formation from calcium aluminoferrite hydration. In: 10th
International Congress on the chemistry of cement and concrete, 2.-6.6.1997, Göteborg, 2ii031
53. Glasser, F. P.: Reactions between cement paste components and sulfate ions. In: Marchand, J.;
Skalny, J. (ed.): Materials science of concrete – Special Volume: Sulfate attack mechanisms.
Westerville (Ohio): American Ceramic Society, 1999, p. 99-122.
54. Xie-Ping; Beaudoin, J. J.: Mechanism of sulfate expansion I. Thermodynamic principle of
crystallization pressure. In: Cement and Concrete Research, Volume 22 (1992), p. 631-640.
55. Johansen, V.; Thaulow, N.; Skalny, J.: Simultaneous presence of alkali-silica gel and ettringite in
concrete. In: Advances in Cement Research, Volume 5 (1993), p. 23-29.

114
56. Erlin, B.: Ettringite- whatever you may think it is. In: 18. International Conference on Cement
Microscopy. Duncanville (Texas): International Cement Microscopy Association, 1996, p. 380-
381.
57. Damidot, D.; Glasser, F. P.: Thermodynamic investigation of the CaO-Al2O3-CaSO4-H2O
systemat 25°C and the influence of Na2O. In: Cement and Concrete Research, Volume 23 (1993),
p. 221-238.
58. Mehta, P. K.: Mechanism of expansion associated with ettringite formation. In: Cement and
Concrete Research, Volume 3 (1973), p. 1-6.
59. Mehta, P. K.: Mechanism of sulfate attack on Portland cement concrete- another look. In: Cement
and Concrete Research, Volume 13 (1983), p. 401-406.
60. Lafuma, H.: Théorie de l’expansion des liants hydraulique. In: Revue des matériaux de
construction, Volume 243 (1929), p. 441-444.
61. Glasser, F. P.: The pore fluid in Portland cement: its role and composition. In: 11th International
Congress on the Chemistry of Cement, 11.-16.5.2003, Durban, p. 19-30.
62. Tian, B.; Cohen, M. D.: Does gypsum formation during sulfate attack on concrete lead to
expansion? In: Cement and Concrete Research, Volume 30 (2000), p. 117-123.
63. Mehta, P. K.; Pirtz, D.; Polivka, M.: Properties of alite cements. In: Cement and Concrete
Research, Volume 9 (1979), p. 439-450.
64. Lea, F. M.: The mechanism of sulphate attack on Portland cement. In: Canadian Journal of
Research, Volume 27 (1949), p. 297-302.
65. Gonzáles, M. A.; Irassar, E. F.: Ettringite formation in low C3A Portland cement exposed to
sodium sulfate solution. In: Cement and Concrete Research, Volume 27 (1997), p. 1061-1072.
66. Santhanam, M.; Cohen, M. D.; Olek, J.: Effects of gypsum formation on the performance of
cement mortars during external sulfate attack. In: Cement and Concrete Research, Volume
33(2003), p. 325-332.
67. Biczok, J.: Betonkorrosion – Betonschutz. Berlin: Verlag für Bauwesen, 1968.
68. Lea, F. M.: The chemistry of cement and concrete. London: Arnold, 1956.
69. Wittekindt, W.: Sulfatbeständige Zemente und ihre Prüfung. In: Zement-Kalk-Gips, Volume
13(1960), p. 565-572.
70. Koch, A.; Steinegger, H.: Ein Schnellprüfverfahren für Zemente auf ihr Verhalten bei
Sulfatangriff. In: Zement-Kalk-Gips, Volume 13 (1960), p. 317-324.
71. Mulenga, D. M.; Nobst, P.; Stark, J.: Praxisnahes Prüfverfahren zum Sulfatwiderstand von Beton
und Mörtel mit und ohne Flugasche. In: Beiträge zum 37. Forschungskolloquium des Deutschen
Ausschuss für Stahlbeton, 7.-8.10.1999, Weimar, p. 197-213.
72. Bellmann, F.; Möser, B.; Stark, J.: Influence of sulfate solution concentration on the formation of
gypsum in sulfate resistance test specimen. In: Cement and Concrete Research, Volume 36
(2006), p. 358-363.
73. Hansen, W. C.; Pressler, E. E.: Solubility of Ca(OH)2 and CaSO4 2H2O in dilute alkaline
solutions. In: Industrial and Engineering Chemistry, Volume 39 (1947), p. 1280-1282.
74. Heller, L.; Ben-Yair, M.: Effect of sulphate solutions on normal and sulphate-resisting Portland
cement. In: Journal of Applied Chemistry, Volume 14 (1964), p. 20-30.
75. Dimic, D.; Droljc, S.: The influence of alite content on the sulfate resistance of Portland cement.
In: 8th International Congress on the Chemistry of Cement, 22.-27.9.1986, Rio de Janeiro, p. 195-
199.
76. Thorvaldson, T.; Wolochow, D.; Vigfusson, V. A.: Studies on the action of sulphates on Portland
cement. IV- The action of sulphate solutions on mortars prepared from binary and ternary
compounds of lime, silica, alumina and iron. In: Canadian Journal of Research, Volume 6 (1932),
p. 485-517.

115
77. Davis, R. E.; Hanna, W. C.; Brown, E. H.: Cement investigations for Boulder dam- results of test
on mortars up to age of 10 years. In: Journal of the American Concrete Institute, Volume 18
(1946), p. 21-47.
78. Irassar, E. F.; Gonzáles, M.; Rahhal, V.: Sulphate resistance of type V cements with limestone
filler and natural pozzolana. In: Cement and Concrete Composites, Volume 22 (2000), p. 361-
368.
79. Cao, H. T.; Bucea, L.; Ray, A.; Yozghatlian, S.: The effect of cement composition and pH of
environment on sulfate resistance of Portland cements and blended cements. In: Cement and
Concrete Composites, Volume 19 (1997), p. 161-171.
80. Mather, K.: Research in sulfate resistance at the Waterways Experiment Station. In: George
Verbeck Symposium on sulfate attack on concrete. (ACI Special publication, 77). Detroit:
American Concrete Institute, 1982, p. 63-74.
81. Locher, F. W.: Zement: Grundlagen der Herstellung und Verwendung. Düsseldorf: Verlag Bau
und Technik, 2000.
82. Massazza, F.: Pozzolana and pozzolanic cements. In: Hewlett, P. C. (ed.): Lea’s chemistry of
cement and concrete. – 4th edition. - London: Arnold, 1998, p. 471- 631.
83. Locher, F. W.: Zur Frage des Sulfatwiderstandes von Hüttenzementen. In: Zement-Kalk-Gips,
Volume 19 (1966), p. 395-401.
84. Torii, K.; Taniguchi, K.; Kawamura, M.: Sulfate resistance of high fly ash content concrete. In:
Cement and Concrete Research, Volume 25 (1995), p. 759-768.
85. Thomas, M. D. A.; Bleszynski, R. F.; Scott, C. E.: Sulfate attack in a marine environment. In:
Marchand, J.; Skalny, J. (ed.): Materials science of concrete – Special Volume: Sulfate attack
mechanisms. Westerville (Ohio): American Ceramic Society, 1999, p. 301-313.
86. Dunstan, E. R. Jr.: A possible method for identifying fly ashes that will improve the sulphate
resistance of concrete. In: Cement, Concrete and Aggregates, Volume 2 (1980), p. 20-30.
87. Mehta, P. K.: Effect of fly ash composition on sulfate resistance of cement. In: ACI Materials
Journal, Volume 83 (1986), p. 994-1000.
88. Mehta, P. K.: Sulfate attack on concrete- A critical review. In: Skalny, J. (ed.): Materials science
of concrete III. Westerville (Ohio): American Ceramic Society, 1992, p. 105-130.
89. Dhir, R. K. ; Hubbard, F. H. ; Munday, J. G. L.; Jones, M. R.; Duerden, S. L.: Contribution of pfa
to concrete workability and strength development. In: Cement and Concrete Research, Volume 18
(1978), p. 277-289.
90. Péra, J.; Husson, S.; Guilhot, B.: Influence of finely ground limestone on cement hydration. In:
Cement and Concrete Composites, Volume 21 (1999), p. 99-105.
91. Hornung, D.; Gathemann, B.; Stark, J.: Leistungsoptimierte integrierte Bindemittel für leicht- und
selbstverdichtende Betone mit hoher Dauerhaftigkeit. In: Proceedings of the 15. Internationale
Baustofftagung- ibausil. Weimar, 24.-27.9.2003, p. 2/195-2/218.
92. Feldman, R. F.: Significance of porosity measurements on blended cement performance. In:
Malhotra, V. M. (ed.): First International Conference on the use of fly ash, silica fume, slag and
other mineral byproducts in concrete. (ACI Special publication, 79). Detroit: American Concrete
Institute, 1983, p. 415-433.
93. Manmohan, D.; Mehta, P. K.: Influence of puzzolanic, slag, and chemical admixtures on pore size
distribution and permeability of hardened cement pastes. In: Cement, Concrete, and Aggregates,
Volume 3 (1981), p. 63-67.
94. Cohen, M. D.; Bentur, A.: Durability of Portland cement-silica fume pastes in magnesium
sulphate and sodium sulfate solutions. In: ACI Materials Journal, Volume 85 (1988), p. 148-157.
95. Gallop, R. S.; Taylor, H. F. W.: Microstructural and microanalytical studies of sulfate attack. I.
ordinary Portland cement paste. In: Cement and Concrete Research, Volume 22 (1992), p. 1027-
1038.

116
96. Gallop, R. S.; Taylor, H. F. W.: Microstructural and microanalytical studies of sulfate attack.
II.Sulfate resisting Portland cement: ferrite composition and hydration chemistry. In: Cement and
Concrete Research, Volume 24 (1994), p. 1347-1358.
97. Gallop, R. S.; Taylor, H. F. W.: Microstructural and microanalytical studies of sulfate attack.
III.Sulfate-resisting Portland cement: reactions with sodium and magnesium sulfate solutions. In:
Cement and Concrete Research, Volume 25 (1995), p. 1581-1590.
98. Gallop, R. S.; Taylor, H. F. W.: Microstructural and microanalytical studies of sulfate attack. IV.
Reactions of a slag cement paste with sodium and magnesium sulfate solutions. In: Cement and
Concrete Research, Volume 26 (1996), p. 1013-1028.
99. Gallop, R. S.; Taylor, H. F. W.: Microstructural and microanalytical studies of sulfate attack.
V.Comparison of different slag blends. In: Cement and Concrete Research, Volume 26 (1996), p.
1029, 1044.
100. Bonen, D.; Cohen, M. D.: Magnesium sulfate attack on Portland cement paste. I. Micro-structural
analysis. In: Cement and Concrete Research, Volume 22 (1992), p. 169-180.
101. Bonen, D.; Cohen, M. D.: Magnesium sulfate attack on Portland cement paste. II. Chemical and
mineralogical analysis. In: Cement and Concrete Research, Volume 22 (1992), p. 707-718.
102. Bonen, D.: A microstructural study of the effect produced by magnesium sulfate on plain and
silica fume-bearing Portland cement mortars. In: Cement and Concrete Research, Volume 23
(1993), p. 541-553.
103. Brown, P. W.: An evaluation of the sulfate resistance of cements in a controlled environment.
In:George Verbeck Symposium on sulfate attack on concrete. (ACI Special publication, 77).
Detroit: American Concrete Institute, 1982, p. 83-91.
104. Irassar, E. F.: Sulfate resistance of blended cements: Prediction and relation with flexural
strength. In: Cement and Concrete Research, Volume 20 (1990), p. 209-218.
105. Lagerblad, B.: Long-term test of concrete resistance against sulfate attack. In: Marchand, J.;
Skalny, J. (ed.): Materials science of concrete - Special Volume: Sulfate attack mechanisms.
Westerville (Ohio): American Ceramic Society, 1999, p. 325-336.
106. Clifton, J. R.; Frohnsdorff, G.; Ferraris, C.: Standards for evaluating the susceptibility of cement-
based materials to external sulfate attack. In: Marchand, J.; Skalny, J. (ed.): Materials science of
concrete -Special Volume: Sulfate attack mechanisms. Westerville (Ohio): American Ceramic
Society, 1999, p. 337-355.
107. Santhanam, M.; Cohen, M. D.; Olek, J.: Mechanism of sulfate attack: A fresh look. Part 2:
Proposed mechanisms. In: Cement and Concrete Research, Volume 33 (2003), p. 341-346 (b).
108. Ferraris, C. F.; Clifton, J. R.; Stutzman, P. E.; Garboczi, E. J.: Mechanisms of degradation of
cement based systems by sulfate attack. In: Scrivener, K. L.; Young, J. F.: Mechanisms of
chemical degradation of cement-based systems. London: E&F Spon, 1997, p. 185-192.
109. Brown, P. W.; Badger, S.: The distributions of bound sulfates and chlorides in concrete subjected
to mixed NaCl, MgSO4, Na2SO4 attack. In: Cement and Concrete Research, Volume 30 (2000),
p. 1535-1542.
110. Diamond, S.; Lee, R. J.: Microstructural alterations associated with sulfate attack in permeable
concretes. In: Marchand, J.; Skalny, J. (ed.): Materials science of concrete – Special Volume:
Sulfate attack mechanisms. Westerville (Ohio): American Ceramic Society, 1999, p. 123-173.
111. Yang, S.; Zhongzi, X.; Mingshu, T.: The process of sulfate attack on cement mortars. In:
Advanced Cement based Materials, Volume 4 (1996), p. 1-5.
112. Bensted, J.: Thaumasite – direct, woodfordite and other possible formation routes. In:
Proceedings of the First International Conference on Thaumasite in Cementitious Materials.
Building Research Establishment, 2002.
113. Bensted, J.: Mechanism of thaumasite sulphate attack in cements, mortars and concretes. In: ZKG
International, Volume 53 (2000), p. 704-709.

117
114. Bensted, J.: Thaumasite – background and nature in deterioration of cements, mortars and
concretes. In: Cement and Concrete Composites, Volume 21 (1999), 117-121.
115. Barnett, S. J; Adam, C. D.; Jackson, A.R.W.: Solid solutions between ettringite and thaumasite.
In: Journal of Materials Science, Volume 35 (2000), p. 4109-4114.
116. Irassar, E. F.; Bonavetti, V. L.; González, M.: Microstructural study of sulfate attack on ordinary
and limestone Portland cements at ambient temperature. In: Cement and Concrete Research,
Volume 33 (2003), p. 31-41.
117. Köhler, S.; Heinz, D.; Urbonas, L.: Effect of ettringite on thaumasite formation. In: Cement and
Concrete Research, Volume 36 (2006), p. 697-706.
118. Bensted, J.: A discussion of the paper “On the formation of thaumasite
CaSiO3⋅CaSO4⋅CaCO3⋅15H2O: Part II.” and reply by F. Bellmann. In: Advances in Cement
Research, Volume 18 (2006), p. 129-134.
119. Bellmann, F.: On the formation of thaumasite CaSiO3 CaSO4 CaCO3 15H2O (Part II). Advances
in Cement Research, Volume 16 (2004), p. 89-94 and Volume 17 (2005), p. 46.
120. Al-Almoudi, O.S.B.; Maslehuddin, M.; Rasheeduzzafar: Permeability of concrete – influential
factors. In: Proceedings of the 4th International Conference on deterioration and repair of
reinforced concrete in the Arabian Gulf, Bahrain, Volume II, 1993, p. 717-33.
121. Crammond, N. J.; Halliwell, M. A.: Assessment of the conditions required for the thaumasite
form of sulphate attack. In: Scrivener, K. L.; Young, J. F. (editors): Mechanisms of chemical
degradation of cement-based systems. 1997, p. 193-200.
122. Al-Amoudi, O.S.B.: Attack on plain and blended cements exposed to aggressive sulfate
environments. In: Cement and Concrete Composites, Volume 24 (2002), p. 305-316.
123. Bessa, A.; Bigas, J. P.; Gallias, J. L.: Influence of cement type and mineral admixtures on
transport properties and chemical resistance of mortars. In: Oh, B. H. et al. (editors): Concrete
under severe conditions: Environment and Loading CONSEC ’04, Seoul 2004, p.
124. El-Alfi, E. A.; Radwan, A. M.; Abed El- Aleem, S.: Effect of limestone fillers and silica fume
pozzolana on the characteristics of sulphate resistant cement pastes, In: Ceramics- Silikàty,
Volume 48 (2004), p. 29-33.
125. Hughes, D. C.: Sulfate resistance of OPC, OPC/fly ash and SRPC pastes: pore structure and
permeability. In: Cement and Concrete Research, Volume 15 (1985), p. 1003-12.
126. Vuk, T.; Gabrovsek, R.; Kaucic, V.: Influence of mineral admixtures on sulfate resistance of
limestone cement pastes aged in cold MgSO4 solution. In: Cement and Concrete Research,
Volume 32 (2002), p. 943-948.
127. Bellmann, F.: Stark, J.: Formation of thaumasite at low sulfate concentrations. In: Proceedings of
the 1st International Workshop on the Service Life Design for Underground Structures, 19.-
20.10.2006, Shanghai, p.117-127.
128. Jallad, K. N.; Santhanam, M.; Cohen, M. D.: Stability and reactivity of thaumasite at different pH
levels. In: Cement and Concrete Research, Volume 33 (2003), p. 433-437.
129. Heinz, D.; Urbonas, L.: Concrete-damaging thaumasite formation? – Examination of a case of
damage. In: Cement International, Volume 4 (2006), p..
130. Sibbick, T.; Fenn, D.; Crammond, N.: The occurrence of thaumasite as a product of seawater
attack. In: Proceedings of the First International Conference on Thaumasite in Cementitious
Materials. Building Research Establishment, 2002.
131. Poppe, A-M.; De Schutter, G.: Analytical hydration model for filler rich self-compacting
concrete. In: Journal of Advanced Concrete Technology, Volume 4 (2006), p. 259-266.
132. Jonasson, J-E.: Modelling of temperature, moisture and stresses in young concrete. Luleå
University of Technology, Division of Structural Engineering. Doctoral Thesis 1994:153 D.
ISSN: 0348-8373. 1994.
133. ConTeSt Pro – Users manual: A program for temperature and stress calculations in concrete. A
publication from Cementa AB, Stockholm, 2003

118
134. Utsi, S.; Jonasson, J-E.; Wallin, K.: Development of SCC for civil engineering purposes. In:
Proceedings of the RILEM 3rd International Symposium on Self-Compacting concrete, Reykjavik,
2003, p. 9.
135. Utsi, S.; Jonasson, J-E.; Wallin, K.: Use of SCC in a tunnel lining for a railway tunnel in Sweden.
In: Proceedings of the RILEM 3rd International Symposium on Self-Compacting concrete,
Reykjavik, 2003, p. 10.
136. Bellmann, F.; Röck, R.; Stark, J.: Thaumasite damage in a shotcrete tunnel lining. In: Cement
International, Volume 3 (2005), p. 102-109.

119

You might also like