11.05.2017 Views

CCA Technical Support CD_2005

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Course for Cement Applications<br />

<strong>Technical</strong> Documentation<br />

Version <strong>2005</strong>


Guidelines<br />

This <strong>CD</strong> rom contains the current version of the<br />

<strong>Technical</strong> Documentation developed by Holcim<br />

Group <strong>Support</strong> Ltd Divisions on which the Course<br />

for Cement Applications is based.<br />

It is divided into 2 sections:<br />

• Cement Section<br />

Clicking on the section will link the user to<br />

the contents page of that section. Chapters<br />

within each section can be accessed by<br />

clicking on the chapter title.<br />

Other sections and chapters can be<br />

accessed directly by clicking on the<br />

navigation bar in the header of each page.<br />

• Concrete Section


Cement<br />

Concrete<br />

Cement - Contents<br />

Chapter No.<br />

1. CEMENT HYDRATION .................................................................................................... 2<br />

2. CEMENT PROPERTIES ................................................................................................ 30<br />

3. CLINKER OPTIMIZATION ............................................................................................. 48<br />

4. MINERAL COMPONENTS............................................................................................. 67<br />

5. CEMENT ADMIXTURES.............................................................................................. 121<br />

6. CEMENT GRINDING.................................................................................................... 125<br />

7. PROPERTIES AND APPLICATION OF COMPOSITE CEMENTS ............................. 147<br />

8. SPECIAL CEMENTS.................................................................................................... 177<br />

9. CEMENT STANDARDS ............................................................................................... 194<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 1 of 220


Cement<br />

Concrete<br />

Cement Hydration<br />

Author: S. Montani<br />

1. INTRODUCTION ................................................................................................................... 3<br />

2. MAIN FEATURES OF CEMENT HYDRATION .................................................................... 3<br />

3. HYDRATION REACTIONS ................................................................................................... 4<br />

3.1 ...Hydration of the individual clinker components.......................................................... 4<br />

3.2 ...Mineral additives ........................................................................................................ 6<br />

4. MECHANISM AND KINETICS.............................................................................................. 6<br />

4.1 ...Basic theories............................................................................................................. 6<br />

4.2 ...Reaction sequence during cement hydration............................................................. 7<br />

4.3 ...Acceleration or retardation of cement hydration ........................................................ 9<br />

5. HEAT OF HYDRATION ...................................................................................................... 11<br />

6. ROLE OF GYPSUM IN CEMENT ....................................................................................... 14<br />

6.1 ...General aspects ....................................................................................................... 14<br />

6.2 ...Calcium sulphate modifications................................................................................ 14<br />

6.3 ...Gypsum and setting of cement ................................................................................ 15<br />

6.4 ...Effect of gypsum on strength and volume stability................................................... 18<br />

6.5 ...Gypsum substitutes.................................................................................................. 20<br />

7. MICROSTRUCTURE AND PROPERTIES OF HARDENED CEMENT<br />

PASTE .............................................................................................................................. 21<br />

7.1 ...Basic physical features ............................................................................................ 21<br />

7.2 ...Morphological features............................................................................................. 22<br />

7.3 ...Evolution of pastes during hydration ........................................................................ 25<br />

7.4 ...Influence of the aggregates on the structure............................................................ 28<br />

8. LITERATURE...................................................................................................................... 29<br />

8.1 ...General overview on cement hydration.................................................................... 29<br />

8.2 ...Gypsum and setting of cement ................................................................................ 29<br />

8.3 ...Microstructure and properties of hardened cement paste........................................ 29<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 2 of 220


Cement<br />

Concrete<br />

1. Introduction<br />

The term cement hydration applies to all the reactions of cement with water. These reactions<br />

determine to a great extent the concrete properties, as it is finally the mixture of cement with<br />

water which is the binding agent in concrete. With regard to the cement application, it is thus<br />

essential to have a basic understanding of the processes occurring during cement hydration.<br />

To a minor extent, hydration reactions can also take place before the cement is applied in<br />

concrete during the storage of clinker and the grinding and storage of cement. Even this<br />

minor surface hydration may cause serious changes of the physical properties of cement.<br />

Table 1 gives an idea of the possible hydration of cement from clinker storage up to cement<br />

application.<br />

Table 1:<br />

Occurrence of cement hydration<br />

Storage of clinker 0 to 10%<br />

Grinding of cement 0 to 1%<br />

Storage of cement 0 to 4%<br />

Concrete mix 0 to 100%<br />

Percentage hydration<br />

The study of the mechanisms and phenomena of cement hydration has a long tradition in<br />

cement research. First basic theories to explain setting and hardening of cement have been<br />

established by Le Châtelier and Michaelis at the end of the last century. The actual more<br />

refined theories still base on the work of these two researchers. The open questions<br />

remaining are principally related to the reaction of C 3 S and C 3 A at early stages.<br />

2. Main Features of Cement Hydration<br />

As for any chemical reaction, main features of interest with regard to the cement hydration<br />

are the hydration reactions, the mechanisms and kinetics and the heat of hydration. In the<br />

following, a brief overview on these three aspects shall be given:<br />

♦ Hydration<br />

reactions<br />

The nature of the hydration products is decisive for the mechanical properties of the<br />

hardened concrete. In Portland cement, the hydration is predominantly a reaction of the<br />

calcium silicates with water, producing a gel-like calcium silicate hydrate and calcium<br />

hydroxide:<br />

Calcium silicates + H 2 O → CSH gel + Ca(OH) 2<br />

In blended cements, the reaction schemes get more complex. Nevertheless, the<br />

hydration products are in general quite similar.<br />

♦ Mechanisms and kinetics<br />

The understanding of the mechanisms of cement hydration gives valuable indications on<br />

the reaction behaviour to be expected, including the kinetics. The knowledge of the<br />

kinetics of cement hydration and the influencing factors is very important for the concrete<br />

practice. The reaction speed of the cement hydration must be slow enough to allow the<br />

placement of concrete; after that, a rapid reaction is desired.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 3 of 220


Cement<br />

Concrete<br />

♦ Heat of hydration<br />

The heat liberated during cement hydration may improve or impair the quality of the<br />

concrete. For ordinary Portland cement, the heat of hydration is typically 380 J/g at 28<br />

days. This value is lower for the other special Portland cements with the exception of<br />

rapid hardening cement. In case of the blended cements, generally less heat is<br />

developed.<br />

Besides the above mentioned chemical aspects, considerable changes in the physical<br />

properties of the cement are associated with the cement hydration. The principal changes<br />

are:<br />

♦ Development of strength<br />

From the practical point of view of course the most important change. As long as the<br />

hydration proceeds, the cement is continuously gaining strength.<br />

♦ Increase in specific surface<br />

The transformation of the low surface cement to a gel-like hydrated product of extremely<br />

high surface is one of the most striking changes of cement hydration. After complete<br />

hydration, typically a 1000 fold increase in specific surface is obtained.<br />

♦ Increase in solid volume<br />

The total volume occupied by the hydration products is roughly twice the volume<br />

occupied originally by the unhydrated cement. That means that 1 cm 3 of cement will give<br />

more or less 2 cm 3 of hydrated cement.<br />

More detailed and specific indications on the different aspects of cement hydration will be<br />

given in the following chapters.<br />

3. Hydration Reactions<br />

3.1 Hydration of the individual clinker components<br />

3.1.1 Calcium silicates<br />

The hydration reactions of the two calcium silicates in clinker - C 3 S and C 2 S - can be<br />

represented by the following chemical equations:<br />

2 (3CaO • SiO 2 ) + 6H 2 O 3CaO • 2SiO 2 • 3H 2 O + 3Ca(OH) 2<br />

Weight fraction 100 + 24 75 + 49<br />

(in symbols of cement chemistry: 2C 3 S + 6H C 3 S 2 H 3 + 3CH)<br />

2 (2CaO • SiO 2 ) + 4H 2 O 3CaO • 2SiO 2 • 3H 2 O + Ca(OH) 2<br />

Weight fraction 100 + 21 99 + 22<br />

(in symbols of cement chemistry: 2 C 2 S + 4H C 3 S 2 H 3 + CH)<br />

The reactions of both silicates are thus stochiometrically very similar and require more or<br />

less the same amount of water. The main difference between the two reactions is that C 3 S<br />

produces more than twice as much calcium hydroxide as C 2 S.<br />

The principal hydration product of the calcium silicates is a gel-like or microcrystalline<br />

calcium silicate hydrate. The formula C 3 S 2 H 3 is only approximate, because the composition<br />

of this hydrate is actually variable over quite a wide range. In contrast, the calcium hydroxide<br />

formed is a crystalline material with a fixed composition.<br />

It is believed that calcium hydroxide is not significantly contributing to strength and that the<br />

calcium silicate hydrates are in the first place responsible for the strength development. The<br />

importance of the calcium hydroxide being a strong base is mainly related to its effects on<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 4 of 220


Cement<br />

Concrete<br />

the passivation of the steel reinforcement, the leaching of concrete and the reaction with<br />

reactive silica and alumina from pozzolanic additives.<br />

3.1.2 Tricalciumaluminate<br />

The hydration of aluminates is heavily influenced by the presence of gypsum. In the absence<br />

of gypsum, the reaction of C 3 A with water is very violent and leads to immediate stiffening of<br />

the paste (known as flash set) due to the rapid formation of hexagonal calcium aluminate<br />

hydrates (C 2 AH 8 and C 4 AH 13 ). These hydrates are not stable and later convert into the cubic<br />

form C 3 AH 6 , so that the final form of reaction can be written as follows:<br />

3CaO • Al 2 O 3 + 6H 2 O 3CaO • Al 2 O 3 • 6H 2 O<br />

Weight fraction 100 + 40 40<br />

(in symbols of cement chemistry: C 3 A + 6H C 3 AH 6 )<br />

The theoretical capacity of C 3 A to combine with water is such that 100 parts by weight of C 3 A<br />

combine with 40 parts by weight of H 2 O; it is thus nearly double that of the silicates.<br />

When gypsum is present in the cement, flash set can be avoided and the C 3 A reacts first with<br />

the gypsum to form calcium trisulfate-aluminate hydrate (ettringite):<br />

3 CaO • Al 2 O 3 + 3CaSO 4 + 32H 2 O 3 CaO • Al 2 O 3 • 3CaSO 4 • 32H 2 O<br />

Weight fraction 66 + 100 + 137 203<br />

(in symbols of cement chemistry: C 3 A + 3CS + 32H C 3 A • 3 CS • H 32 )<br />

This reaction continues until all the available gypsum is used up. After the depletion of<br />

gypsum, ettringite is converted into the monosulfate:<br />

3CaO • Al 2 O 3 • 3CaSO 4 • 32H 2 O + 2(3CaO • Al 2 O 3 ) + 4H 2 O<br />

(3CaO • Al 2 O 3 • CaSO 4 • 12H 2 O)<br />

(C 3 A • 3CS • H 32 + 2C 3 A + 4H 3C 3 A • CS • H 12 )<br />

Calcium aluminate hydrate continues to be formed from any remaining unhydrated C 3 A and<br />

the final product of C 3 A hydration is generally a monosulphate-calciumaluminate hydrate<br />

solid solution.<br />

3.1.3 Ferrite phase<br />

The C 4 AF forms the same sequence of hydration products as does C 3 A, with or without<br />

gypsum. The reactions are, however, slower; C 4 AF never hydrates rapidly enough to cause<br />

flash set and gypsum retards C 4 AF hydration more drastically than it does C 3 A.<br />

3.1.4 Calcium oxide and magnesium oxide<br />

The uncombined lime (calcium oxide) and periclase (magnesium oxide), if present in large<br />

quantities in cement, may cause expansion due to a slow hydration reaction after the setting:<br />

CaO + H 2 O Ca(OH) 2 ; MgO + H 2 O Mg(OH) 2<br />

The actual degree of expansion depends on the state and distribution of these oxides in<br />

cement, in particular the large crystals of periclase or hard burnt free lime produce<br />

unsoundness.<br />

3.1.5 Alkalis<br />

Alkalis are present as alkalis sulphates or are incorporated in the main clinker phases. The<br />

first type of alkalis go readily into solution upon mixing with water and accelerate the early<br />

hydration reactions. The effect of alkalis contained in the clinker phases is not well known;<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 5 of 220


Cement<br />

Concrete<br />

there is at least some dissolution with time which has some effect on later hydration<br />

reactions.<br />

The presence of alkalis in the pore solution generally does not give any special problems<br />

except when certain aggregates are used that can participate in the alkali-aggregate reaction<br />

leading to expansion and disruption of the concrete. The aggregates giving raise to such a<br />

reaction contain either reactive silica or dolomite minerals.<br />

3.2 Mineral additives<br />

3.2.1 Blast furnace slag<br />

Granulated blast furnace slag does not hydrate per se in water, but in the presence of<br />

activators like lime, alkalis, sulphate and Portland cement, it shows hydraulic properties. The<br />

activators appear to function by the removal of passive surface films that act as a barrier to<br />

significant hydration. The glass in the granulated slag is the hydraulic component in the<br />

presence of Portland cement.<br />

The hydration reactions of blast furnace slag are not easy to follow and it is thus difficult to<br />

establish clear reaction equations. Studies on slag cements have revealed that the hydration<br />

products of the slag hydration are similar to those of the calcium silicate reaction. It is<br />

believed that the main difference to Portland cement hydration is the lower C/S ratio of the<br />

calcium silicate hydrates and the reduced amount of calcium hydroxide present.<br />

3.2.2 Pozzolanic additives<br />

The natural and artificial pozzolans contain reactive siliceous and aluminous materials which<br />

on their own do not possess cementitious properties. They react, however, in the presence of<br />

water with dissolved calcium hydroxide released from the hydration of the Portland cement<br />

clinker. The reaction products are calcium silicate and calcium aluminate hydrates similar to<br />

those found in the hydration of Portland cement clinker. Favourable for this pozzolanic<br />

reaction seems to be the presence of alkalis.<br />

3.2.3 Limestone filler<br />

Finely ground limestone can react during hydration with the C 3 A from the clinker to form<br />

monocarboaluminate (C 3 A • CaCO 3 • 11H 2 O). The presence of limestone in the cement can<br />

thus to a certain extent retard the hydration of C 3 A, but it is by far less effective than gypsum<br />

and can not avoid flash set. Moreover, fine limestone filler can have an indirect influence on<br />

the hydration reaction of C 3 S.<br />

4. MECHANISM AND KINETICS<br />

4.1 Basic theories<br />

As mentioned in the introduction, the first theories to explain setting and hardening of cement<br />

were advanced by Le Châtelier and Michaelis. Le Châtelier attributed the development of<br />

cementing action to the passage of the anhydrous cement compounds into solution and the<br />

precipitation of the hydration products as interlocking crystals. In the theory put forward by<br />

Michaelis, cohesion is considered to be the result of the formation of a colloidal gelatinous<br />

mass. The formation of the gel may take place without the cement compounds going into<br />

solution by a direct „topochemical“ or „solid state“ reaction.<br />

Both hydration mechanisms take place within in the cement paste. Cementitious reaction<br />

initially takes place by some dissolution and precipitation, whilst the hydrated material formed<br />

has only a small degree of crystalline order being gel-like and of colloidal dimensions. The<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 6 of 220


Cement<br />

Concrete<br />

colloidal theory can also explain the phenomena of swelling, shrinkage, creep and<br />

selfhealing of cracks.<br />

4.2 Reaction sequence during cement hydration<br />

Hydration of cement is a sequence of overlapping chemical reactions between clinker<br />

components, mineral additives, calcium sulphate and water, leading to continuous cement<br />

paste stiffening and hardening. The hydration reactions described before proceed<br />

simultaneously at differing rates and influence each other.<br />

A simplified schematic presentation of the hydration process of Portland cement is given in<br />

Figure 1. The three main stages which can be distinguished in this presentation are:<br />

First stage<br />

In the first stage, ettringite and calcium hydroxide are formed. The ettringite covers the<br />

aluminate particles and avoids flash set of the cement paste. After these first reactions, the<br />

so-called „dormant“ period of typically four to six hours starts where no considerable further<br />

hydration takes place (certain increase in Ca 2+ concentration).<br />

The dormant period is usually explained by the behaviour of the C 3 S, the main constituent of<br />

Portland cement. The two principal theories are based on the protective layer and the<br />

delayed nucleation concept respectively. The first ascribes the „dormant“ period to the<br />

formation of a protective layer on the C 3 S particles which is destroyed with time and the latter<br />

regards the retardation in C 3 S reaction as the cause of a delayed nucleation of the<br />

corresponding hydration phases.<br />

Still in the first stage, the cement paste is setting. This stiffening is attributed by some<br />

researchers to the recrystallisation of the initially finely divided ettringite. Others believe that<br />

the setting is caused by the loss in plasticity due to some first formation of calcium silicate<br />

hydrate from C 3 S.<br />

Figure 1:<br />

Schematic representation of the formation of hydrate phases in Portland<br />

cement paste<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 7 of 220


Cement<br />

Concrete<br />

Second stage<br />

The second stage after the dormant period is characterised by the restarting of the hydration.<br />

Precipitation of CSH in form of long fibres and calcium hydroxide takes place on the surface<br />

of the silicate phases and the reaction proceeds rapidly for all clinker phases. This stage<br />

ends with the termination of the ettringite formation and the development of a basic matrix<br />

after about 24 hours.<br />

Third stage<br />

In the third stage, the spaces between the solid particles are overbridged and filled with short<br />

fibres of the CSH phase. The initial matrix, formed by first hydrates of the aluminate, ferrite<br />

and silicates, is thus densified and the strength of the cement paste increases. The calcium<br />

hydroxide formed in large crystals is built into the CSH gel matrix. At this stage, ettringite is<br />

converted into monosulfate and corresponding hydration products are formed from the still<br />

not hydrated aluminate and ferrite phases.<br />

As the hydration proceeds, the reactions get more and more diffusion controlled and the<br />

overall rate of reaction is decreasing. Hydration reactions, primarily those of C 2 S, will<br />

continue as long as the reactants and space permit it; this can go on for years. In practice,<br />

there is usually never a complete hydration of the cement and unhydrated cement is nearly<br />

always remaining in the hardened cement paste.<br />

A good overview on the hydration reactions in Portland cements can also be obtained by<br />

taking a look at the heat evolution curves. Figure 2 shows such a curve with the indications<br />

of the main reactions going on in the process of cement hydration.<br />

Figure 2:<br />

Heat evolution curve of Portland cement<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 8 of 220


Cement<br />

Concrete<br />

For blended cements, the reactions sequences get more complex. In case of the active<br />

mineral additives, the main differences to Portland cement hydration can be seen in more<br />

formation of CSH phase at later stage and a decrease in the amount of calcium hydroxide<br />

and the pore space. For the inert mineral additives, the effect will primarily be a „dilution“ of<br />

the Portland cement matrix.<br />

A general effect observed in blended cement is a certain delay of setting and hardening at<br />

early stages. Normally, the active mineral additives start to contribute to cement hydration<br />

only after more than 3 days, when suitable activation has taken place (blast furnace slag) or<br />

sufficient calcium hydroxide has been formed (pozzolans) for their reaction.<br />

4.3 Acceleration or retardation of cement hydration<br />

The process of hydration can be accelerated or retarded by different factors:<br />

♦ Use of accelerators and retarders<br />

Until now, the most efficient accelerator is calcium chloride (CaCl 2 ), but the problem with<br />

this product is the corrosion of the steel reinforcement. The exact mechanism of<br />

acceleration by CaCl 2 is not known; it is believed that CaCl 2 is acting like a catalysator.<br />

Retarders for cement hydration are for instance phosphates, zincates and carbohydrates<br />

(including sugar). The retarding action is simply explained by the formation of a<br />

monomolecular protective film around the cement particles which slows down their<br />

reaction with water.<br />

♦ Hydration temperature<br />

The rate of hydration is strongly influenced by the temperature. Rate of hydration at early<br />

age roughly doubles when the temperature increases by 10°C. The effect of temperature<br />

on cement hydration is of particular importance in extreme climates and for steam curing.<br />

♦ Cement fineness<br />

The higher the fineness of the cement, the more extended is the zone of reaction. This<br />

leads of course to a higher rate of hydration.<br />

♦ Cement composition<br />

The activity of the clinker used in the cement affects considerably the rate of hydration. A<br />

high rate of hydration can be expected of clinkers rich in C 3 A and C 3 S (see Figure 3) and<br />

with a high alkali content.<br />

Important for the rate of hydration is also the content of mineral additives in the cement.<br />

The rate of hydration generally is the lower, the higher the dosage of mineral additives in<br />

the cement.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 9 of 220


Cement<br />

Concrete<br />

Figure 3:<br />

Rates of hydration of compounds in Portland cement<br />

Increasing the rate of hydration, and hence the early strength development, is always done<br />

at the expense of the ultimate strength. An explanation is that the microstructure formed<br />

during the accelerated hydration is usually more coarse and less favourable for strength<br />

development. Contrary to this, retarded, slow hydration leads to the formation of a more<br />

refined microstructure which results in better final strength.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 10 of 220


Cement<br />

Concrete<br />

5. Heat of Hydration<br />

The hydration reaction is an exothermic process, that means during the reaction of cement<br />

with water heat is liberated. The quantity of liberated heat is quite appreciable (typically 380<br />

J/g for OPC at 28 days) and can lead to a significant temperature rise in mass concrete<br />

construction, where the heat is not allowed to escape (see Figure 4). During cooling of the<br />

hardened concrete, thermal gradients may develop and give problems with crack formation.<br />

That is the reason why special low heat cements have to be used for mass concrete<br />

applications.<br />

Figure 4:<br />

Temperature rise in 1:9 (weight) concrete under adiabatic conditions for<br />

different Portland cement types<br />

The main measure to control the heat of hydration of Portland cement is the adjustment of<br />

clinker composition. The clinker minerals contributing most to the heat of hydration are C 3 S<br />

and C 3 A (see Table 2), so that it is necessary to limit these compounds to reduce heat<br />

development during hydration. On the other hand, it is also possible to keep the heat<br />

liberation low by not grinding the cement too fine.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 11 of 220


Cement<br />

Concrete<br />

Table 2:<br />

Heat of hydration of the clinker components at 21°C (J/g)<br />

Clinker component 3 days 28 days 6 ½ years<br />

C 3 S 245 380 490<br />

C 2 S 50 105 225<br />

C 3 A 890 1380 1380<br />

C 4 AF 290 495 495<br />

The effect of clinker composition and cement fineness on the heat of hydration can be well<br />

seen in Figure 5 which shows the development of heat of hydration for the different Portland<br />

cement types. The low heat cements used for mass concrete - type II and IV ASTM - with<br />

low heat of hydration of about 250 to 300 J/g after 28 days contain less C 3 S and C 3 A and are<br />

fairly coarse. The rapid hardening cement - type III ASTM - with increased heat of hydration<br />

of about 420 J/g after 28 days is usually richer in C 3 S and C 3 A and of higher fineness,<br />

compared with the ordinary Portland cement - type I ASTM.<br />

Figure 5:<br />

Development of heat of hydration for the different Portland cement types<br />

cured at 21°C (w/c-ratio of 0.40)<br />

For practical purposes, it is not only the total heat of hydration that matters but also the rate<br />

of heat evolution. The same total heat produced over a long period of time can be dissipated<br />

to a greater degree, consequently producing a smaller rise in temperature in the concrete. In<br />

ordinary Portland cement about one half of the total heat is liberated between 1 and 3 days<br />

and about three quarters in 7 days. It is common to specify the heat of hydration after 7 and<br />

28 days, giving a reasonable indication of both total heat of hydration and its rate of<br />

liberation.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 12 of 220


Cement<br />

Concrete<br />

In addition to the composition and fineness of cement, the rate of heat evolution is also<br />

greatly influenced also by the temperature of hydration (see Table 3). The ambient conditions<br />

thus play also a decisive role with regard to the effects of heat of hydration on the concrete<br />

properties.<br />

Table 3:<br />

Heat of hydration (J/g) developed after 3 days at different temperatures<br />

Cement type<br />

Temperature of cement hydration<br />

4°C 24°C 32°C 41°C<br />

I 154 285 309 335<br />

III 221 348 357 390<br />

IV 108 195 192 214<br />

A very effective means to reduce the amount and rate of heat of hydration of Portland<br />

cement is the use of mineral additives in the cement (blended cements). By the replacement<br />

of the clinker by the less reactive or inert mineral additives, the heat development of the<br />

cement can be easily controlled. As an example, the heat development curves of slag blends<br />

with different Portland cement contents are presented in Figure 6. The advantage of the<br />

blended cement over the low heat Portland cements is that the same clinker as for the<br />

ordinary Portland cement can be used.<br />

Figure 6:<br />

Heat of hydration of mixtures of OPC (P.C.) and ground blast furnace<br />

slag (S); isothermal method.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 13 of 220


Cement<br />

Concrete<br />

6. Role of Gypsum in Cement<br />

6.1 General aspects<br />

Ever since cement has been produced in greater quantities, gypsum (CaSO 4 •2H 2 O) has<br />

been ground into the cement. Gypsum is added to the cement mainly for the purpose of<br />

regulating its setting time. It prevents flash setting and makes the concrete workable for<br />

hours.<br />

The gypsum influences not only the setting but also other cement properties such as<br />

grindability, sensitivity to storage, volume stability and strength. Gypsum is an extremely<br />

important part of cement. It is, however, often neglected to pay proper attention to it in the<br />

production of cement.<br />

Here, only the effect of gypsum on setting, volume stability and strength shall be elucidated<br />

more in detail. For the better understanding of the influence of gypsum on cement, it is first<br />

necessary to take a look at the different modifications in which calcium sulphate can be<br />

present. In a final part of this chapter, the possible gypsum substitutes shall be discussed.<br />

6.2 Calcium sulphate modifications<br />

In cement, it is possible to find at least five basic modifications of calcium sulphate (see<br />

Table 4). The stable modifications, dihydrate (gypsum) and anhydrite II (natural anhydrite),<br />

can be found in nature and both are used as additives to clinker in the grinding process. The<br />

metastable modifications, hemihydrate and anhydrite III, are not available in nature, but can<br />

easily form during grinding and storage of cement at elevated temperatures by dehydration<br />

of gypsum. The high temperature modification anhydrite I can form during burning of the<br />

clinker in the cement kiln.<br />

Table 4:<br />

Modifications of calcium sulphate in cement<br />

Designation Formula Crystal<br />

water<br />

Density<br />

Range of<br />

stability<br />

Solubility at<br />

20°C<br />

Occurrence<br />

% g/cm 3 °C % nature cement<br />

Dihydrate CaSO 4 •2H 2 O 20.92 2.32 < 40 0.20 yes yes<br />

Hemihydrate CaSO 4 •1/2H 2 O 6.21 2.70 metast. 0.95 no yes<br />

(α,β)<br />

Anhydrite III CaSO 4 0 2.50 metast. 0.95 no yes<br />

(α,β)<br />

Anhydrite II CaSO 4 0 2.98 40-1180 0.20 yes yes<br />

Anhydrite I CaSO 4 0 -- > 1180 -- no (~clinker)<br />

The different mode of action of the individual modifications on the properties of cement can<br />

be mainly attributed to their different solubilities in water. In Figure 7, it can be seen that the<br />

solubility of the hemihydrate and anhydrite III is appreciably higher than that of dihydrate or<br />

anhydrite II. The high solubility of hemihydrate and anhydrite III can lead to anomalous<br />

setting, as discussed later. The solubility of dihydrate and anhydrite II in the ambient<br />

temperature range is approximately the same, but their effect on properties of cement can be<br />

very different. The reason for this is the different rate of solubility which is greater for the<br />

dihydrate than for anhydrite II.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 14 of 220


Cement<br />

Concrete<br />

Figure 7:<br />

The Solubility of gypsum, hemihydrate and anhydrite<br />

6.3 Gypsum and setting of cement<br />

The reactions which take place immediately after the addition of water to cement are of<br />

decisive importance for the setting of cement. Directly after the mixing of cement with water,<br />

sulphate dissolves and reacts with aluminate to form ettringite. This vigorous initial reaction<br />

ceases after a few minutes and then the so-called dormant period starts.<br />

According to the classical theory, the set retardation is due to the fact that the ettringite forms<br />

a cohesive cover around the aluminate particles and in this way inhibits their further reaction<br />

(see Figure 8). The reaction inhibiting effect of the ettringite cover is ended only when this<br />

cover is broken open by the pressure of crystallisation and not sufficient sulphate is left to<br />

close the burst section. Thereafter, the reaction of C 3 A can continue unhindered until<br />

complete hydration.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 15 of 220


Cement<br />

Concrete<br />

Figure 8:<br />

Schematic description of set retardation due to C 3 A sulfate interaction<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 16 of 220


Cement<br />

Concrete<br />

The reasons for the setting of cement are still not exactly known. Some researchers explain<br />

the setting of the cement by the recrystallisation of the initially finely divided ettringite into<br />

bigger crystals which are building bridges between the cement particles. Other researchers<br />

attribute the initial set of cement as corresponding to that stage in the hydration sequence<br />

when a sufficient amount of tricalcium silicate has been converted to calcium silicate hydrate.<br />

This initial removement of water causes a partial loss of the plasticity of the cement paste.<br />

The setting of cement is mainly influenced by the reactivity of the clinker (i.e. C 3 A and alkali<br />

content) and by the type and amount of the calcium sulphate added. The cement fineness,<br />

the dosage of mineral additives, use of chemical admixutres and the hydration temperature<br />

play also an important role. The effect of C 3 A content, gypsum dosage and hydration<br />

temperature on the initial setting of Portland cement is illustrated in Figure 9.<br />

Figure. 9:<br />

Influence of gypsum dosage, C 3 A content and hydration temperature on<br />

initial setting of Portland cement<br />

In order to obtain normal setting, it is essential to adjust the supply of soluble sulphate to the<br />

reactivity of the clinker during the first minutes and hours of hydration (see Figure 10). If<br />

there is no proper balance between sulphates and clinker reactivity, abnormal setting can<br />

occur:<br />

Flash set is caused by the formation of aluminate hydrate, due to a high content of reactive<br />

C 3 A and/or a too small amount of easily soluble calcium sulphate.<br />

False set can be traced back to high temperature in the mill, causing a dehydration of the<br />

gypsum to an easily soluble hemihydrate and anhydrite III. Hemihydrate and anhydrite III<br />

recrystallize to gypsum whose crystals grow into each other and form a solid framework<br />

which affects the stiffening of the cement paste. This framework of gypsum crystals is broken<br />

down again through reaction with C 3 A. For this reason the cement paste can regain its<br />

previous plasticity by remixing.<br />

Quick set: The high content of easily soluble sulphate and reactive clinker minerals can also<br />

have an accelerating effect on the formation of sulfo-aluminate hydrate and calcium silicate<br />

hydrate which can likewise lead to early stiffening of the cement (quick set).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 17 of 220


Cement<br />

Concrete<br />

Figure 10:<br />

Formation of rigid structure during setting of Portland cement<br />

Effect of gypsum on strength and volume stability<br />

The addition of gypsum - in addition to its principle purpose of regulating setting - has a<br />

significant effect on the strengths and volume stability. Up to a certain limit, which depends<br />

on the clinker composition, the gypsum addition increases the strengths and prevents<br />

shrinkage. If the gypsum addition surpasses this limit, it causes considerable swelling which<br />

can lead to expansion. This is why an upper limit for the SO 3 content is given in the cement<br />

standards (see Fig. 11).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 18 of 220


Cement<br />

Concrete<br />

Figure 11: Shrinkage and compressive strength of cements as a function of the SO 3<br />

Content<br />

The gypsum required to obtain optimum strength and volume stability grows with the C 3 A<br />

content of the clinker, cement fineness, alkali content and application temperature. Other<br />

variables influencing the optimum SO 3 content are the form (and solubility) of the calcium<br />

sulphate in the cement, the nature and amount of mineral additives and also the presence of<br />

chemical admixtures. Often the optimum SO 3 content would be even above the maximum<br />

value given by the standards.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 19 of 220


Cement<br />

Concrete<br />

6.4 Gypsum substitutes<br />

There has never been lack of effort to substitute gypsum by other set regulators. In particular<br />

in countries with few natural gypsum deposits, the cement industry is searching intensively<br />

for alternative materials. The main alternatives up to date are the use of natural anhydrite<br />

and limestone and the application of by-product gypsum coming form other industries.<br />

The use of natural anhydrite as partial substitute for gypsum is already well established in<br />

practice. In general, it appears that blends of 40 to 60 percent gypsum and 60 to 40 percent<br />

natural anhydrite can be safely used with all clinkers. The advantage of such blends over<br />

pure gypsum is the reduction of the risk of false setting and the improved storage stability<br />

and flowability of the cement.<br />

In some places, limestone is used as replacement for gypsum. For very reactive clinkers, the<br />

replacement is typically limited to 25%. In cements with less reactive clinkers, replacement<br />

levels of up 50% may be possible without negative effects on the setting of cement.<br />

With respect to by-product gypsum, there are a number of commercial processes which<br />

produce such materials. An overview on the main sources of by-product gypsum is given in<br />

Table 5; the most important sources regarding quantities are phosphogypsum and<br />

desulphogypsum. Another type of by-product not mentioned in the table are used plaster<br />

moulds, which are for instance used in casting certain clay products.<br />

Table 5:<br />

Main sources of by-product gypsum<br />

Product/process By-product % sulphate* % impurities<br />

orthophosphoric acid phosphogypsum 95-98 (G) 0.2-1.5 P 2 O 5 , up<br />

to 1.5 fluoride<br />

hydrofluoric acid fluorogypsum 80-95 (HH or AH) 2-3 CaF 2 ,<br />

unreacted<br />

silicofluorides<br />

citric, formic and organogypsum --- ---<br />

tartaric acid<br />

boric acid borogypsum 40-50 (G), 20-30 (AN) 7-10 H 3 BO 3<br />

flue gas SO 2 removal desulfogypsum 90-95 (G), sulfite for mod. ---<br />

process<br />

titanium white titanogypsum 90-95 (G) 0.4 TiO 2<br />

* G = CaSO 4 • 2H 2 0, HH = CaSO 4 • ½H 2 O, AN = CaSO 4<br />

The main problem for the use of such by-products as subsitute for natural gypsum is the<br />

presence of impurities, which may have some harmful effect on setting, strength develpment<br />

or some other property of the cements. Moreover, the amount of impurities can fluctuate<br />

quite a lot so that the variability in cement properties may also be higher.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 20 of 220


Cement<br />

Concrete<br />

7. Microstructure and Properties of Hardened Cement Paste<br />

7.1 Basic physical features<br />

In this section we deal with the physical structure of hydrated cement paste (HCP), as well as<br />

the effect of its morphology on the important properties of hardened concrete.<br />

We will be concerned with the main chemical reaction taking place during hydration:<br />

Calcium Silicates + H 2 O ---> C-S-H + Ca(OH) 2<br />

where C-S-H are the hydrates of the Calcium Silicates (C 3 S and C 2 S), also called "C-S-H<br />

gel".<br />

The most important physical aspect of that reaction is the fact that the total volume occupied<br />

by the hydration products is, roughly, twice the volume occupied originally by the unhydrated<br />

cement (see Figure 12):<br />

Figure 12:<br />

Basic physical feature of cement hydration<br />

It must be mentioned that, out of that total volume, 28 % are very small pores (gel pores) and<br />

the rest (72 %) is solid phase:<br />

Gel pores = 0.28 Vhp<br />

Figure 13 shows schematically the growth of crystals during hydration of two pastes, one of<br />

low and the other of high water/cement ratio (w/c). The w/c ratio is a measure of the degree<br />

of dispersion of the cement grains in water.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 21 of 220


Cement<br />

Concrete<br />

Figure 13:<br />

Structure formation in cement paste during hydration<br />

High w/c<br />

Low w/c<br />

0 h<br />

No setting<br />

Setting<br />

2 - 4 h<br />

High age<br />

Cement hydration can be seen as a continuous process by which the capillary pores (space<br />

originally occupied by water) are being gradually filled with hydration products.<br />

7.2 Morphological features<br />

Figure 14 presents a model (Feldman and Sereda) of the general structure of HCP, showing<br />

its main features.<br />

In the hydrated cement paste, there are two distinctive phases:<br />

♦ Solid Phases<br />

♦ Pores (at least partially filled with water)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 22 of 220


Cement<br />

Concrete<br />

Figure 14:<br />

Model of Feldmann and Sereda for the general structure of hydrated<br />

cement paste<br />

Legend<br />

x Water in interlayer regions<br />

o Water absorbed on surfaces<br />

C Capillary<br />

pore<br />

— C-S-H sheets<br />

7.2.1 Solid phases<br />

7.2.1.1 CALCIUM SILICATE HYDRATES-(C-S-H)<br />

The C-S-H particles occupy 50 - 60 % of the total solid volume in hydrated paste; it is also<br />

called C-S-H gel. The particles are very small (colloidal dimensions) and sheet-like shaped<br />

(see Figure 15)<br />

Figure 15:<br />

Basic feature of C-S-H particle<br />

The C-S-H particles are arranged as a highly disordered layered structure (see Figure 14)<br />

with very high specific surface ≈ 3.000.000 cm 2 /g.<br />

The strength of the HCP is attributed to the attraction forces between the C-S-H crystals over<br />

their enormous surface.<br />

The space left between the particles is called gel pores; these pores are extremely small<br />

(about 1.5 nm), of the order or magnitude of a water molecule (0.25 nm).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 23 of 220


Cement<br />

Concrete<br />

7.2.1.2 CALCIUM HYDROXIDE (CH)<br />

Compound with a definite stoichiometry: Ca(OH) 2 , forming large crystals (0.01 - 0.1 mm) of<br />

hexagonal shape (Portlandite).<br />

These crystals constitute 20 - 25 % of the solid volume in the hydrated paste.<br />

Their contribution to strength is low (low surface area); however, their presence is very<br />

important to passivate the embedded steel rebars.<br />

7.2.1.3 CALCIUM SULFOALUMINATES<br />

Occupy 15 - 20 % of solid volume in hydrated paste; they only play a minor role in strength.<br />

There exist two forms:<br />

Ettringite: Elongated needle shaped crystals<br />

Monosulfate: Hexagonal plates or "rosettes"<br />

Iron oxide can replace aluminium oxide in the Crystal structure.<br />

7.2.1.4 UNHYDRATED CLINKER GRAINS<br />

Smaller grains react faster and coarser grains (> 30 µm) tend to remain partially unhydrated.<br />

7.2.2 Pores<br />

7.2.2.1 LNTERLAYER SPACE IN C-S-H (GEL PORES)<br />

Occupy 28 % of bulk volume of C-S-H gel<br />

Size: 0.5 nm - 2.5 nm (H 2 O molecule = 0.25 nm)<br />

They are always present in HCP but, due to their small size do not adversely affect strength<br />

and permeability. However, the movement of water - that is firmly held within the gel pores -<br />

is the main reason for drying shrinkage and creep.<br />

7.2.2.2 CAPILLARY PORES<br />

Represent the space originally occupied by water, not filled with hydration products. Their<br />

size and volume depend on the w/c ratio and the degree of hydration:<br />

♦ well hydrated paste, low w/c: 10 - 50 nm<br />

♦ at early ages, high w/c: up to 10 µm<br />

7.2.2.3 MACROPORES<br />

Air bubbles naturally entrapped or intentionally entrained during mixing (10 - 100 µm).<br />

"Compaction" voids: mm or cm<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 24 of 220


Cement<br />

Concrete<br />

7.3 Evolution of pastes during hydration<br />

Figures 16, 17 and 18 show the evolution of the structure of cement pastes, made with the<br />

same amount of cement (1 kg) but with different w/c ratios, during hydration (under water).<br />

Figure 16:<br />

Development of Hydration of Cement Paste made with 1 kg of cement;<br />

w/c = 0.70<br />

Figure 17:<br />

Development of Hydration of Cement Paste made with 1 kg of cement;<br />

w/c = 0.50<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 25 of 220


Cement<br />

Concrete<br />

Figure 18:<br />

Development of Hydration of Cement Paste made with 1 kg of cement;<br />

w/c = 0.30<br />

lnitially, all pastes contain about 320 cm 3 of unhydrated cement (1 kg), but different volumes<br />

of water, 700, 500 and 300 cm 3 for w/c ratio 0.7, 0.5 and 0.3, respectively. Consequently,<br />

pastes with higher w/c ratio will have larger total volumes. The volume of water corresponds<br />

to the initial volume of capillary pores.<br />

As hydration proceeds, the total volume remains basically constant, but the amount of<br />

unhydrated cement is gradually reduced. At degree of hydration 0.5, half of the original<br />

cement has been consumed and replaced by hydration products. The latter occupy a volume<br />

about double that of the original cement that hydrated, thus reducing the volume of capillary<br />

pores. From the total volume of hydrates, 72% are solid phase and 28% are gel pores.<br />

If we compare the pastes with w/c 0.50 and 0.70 we see that, as they contained initially the<br />

same volume of cement, at any stage both pastes contain the same amount of hydration<br />

products. However, the paste with higher w/c ratio will contain more pores because the<br />

cement grains were more dispersed.<br />

The paste with w/c ratio 0.30 can never reach full hydration because there is no room left to<br />

accomodate new hydration products. High-strength concretes, with very low w/c ratios<br />

always contain unhydrated cement for that reason.<br />

Clearly, pastes and concretes with lower w/c ratios will have less and smaller capillary pores,<br />

thus explaining why they are stronger and less permeable (more durable), see Figures 19<br />

and 20.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 26 of 220


Cement<br />

Concrete<br />

Figure 19: Strength vs. w/c-ratio<br />

Figure 20:<br />

Permeability vs. w/c-ratio<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 27 of 220


Cement<br />

Concrete<br />

7.4 Influence of the aggregates on the structure<br />

Figure 21 shows that the presence of the aggregates distorts the structure of the HCP in their<br />

vicinity. In this contact or transition zone (about 20-25 µm thick), there exists a higher<br />

capillary porosity and a higher proportion of Ca(OH) 2 than normal, making it weaker than the<br />

rest of the paste. This explains why, in the case of rounded and smooth aggregates the<br />

failure of concrete tends to happen at the interface between the coarse aggregate and the<br />

paste.<br />

Figure 21:<br />

Model of the contact zone between cement stone and aggregate<br />

To achieve high-strength concretes it is necessary to overcome this weakness of the contact<br />

zone. The increased porosity has to be compensated by using high dosages of a<br />

superplastizer and the preferential alignment and presence of Ca(OH) 2 crystals can be<br />

solved by adding microsilica or fly-ash.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 28 of 220


Cement<br />

Concrete<br />

8. Literature<br />

8.1 General overview on cement hydration<br />

Author Title, Publisher pp<br />

Taylor, H.F.W. Cement chemistry, Academic Press, London, 1990<br />

Jawed, I., Skalny,<br />

J. and Young<br />

Mindess, S. and<br />

Young J.F.<br />

Lea, F.M.<br />

J.F, Hydration of Portland cement, in: Structure and<br />

Performance of Cements, Ed. P. Barnes, Applied Science<br />

Publishers, Londond and New York, 1983<br />

Concrete, Prentice Hall, Inc., Englewood Cliffs<br />

The chemistry of cement and concrete, Edward Arnold Ltd.,<br />

Glasgow, 1970<br />

237 - 316<br />

8.2 Gypsum and setting of cement<br />

Author Title, Publisher pp<br />

Frigione G. Gypsum in Cement, in: Advances in Cement Technology,<br />

Ed. S.N. Ghosh, Pergamon Press, Oxford, 1983<br />

307 - 347<br />

Roy, D.M. and<br />

Grutzeck, M.W.<br />

Locher, F.W.,<br />

Richartz, W.,<br />

Sprung, S. et al.<br />

Gypsum & anhydrite in Portland cement, 3rd Edition, United<br />

States Gypsum Company<br />

Setting of cement<br />

Part I Reaction and development of structure, Zement-Kalk-Gips,<br />

1976, 10<br />

Part II Effect of adding calcium sulphate, Zement-Kalk-Gips, 1980,<br />

6<br />

Part III Influence of clinker manufacture, Zement-Kalk-Gips, 1982,<br />

12<br />

Part IV Influence of composition of solution, Zement-Kalk-Gips,<br />

1983, 4<br />

8.3 Microstructure and properties of hardened cement paste<br />

435 - 442<br />

271 - 277<br />

669 - 676<br />

224 - 231<br />

Author Title, Publisher pp<br />

Diamond, S. The microstructures of cement paste in concrete, 8th 122 - 147<br />

International Congress on the Chemistry of Cement, Rio de<br />

Janeiro, 1986, Vol. I<br />

Mehta, P.K. Hardened cement paste - microstructure and its relationship 113 - 121<br />

to properties, dto., Vol.I<br />

Taylor, H.F.W. Structure and composition of hydrates, 7th International pp 2 - 13<br />

Congress on the Chemistry of Cement, Paris, 1980, Vol. I,<br />

Subtheme II-2<br />

Diamond, S. Cement paste microstructure - an overview at several pp.2 - 30<br />

levels, in: Hydraulic Cement Pastes: Their Structure and<br />

Properties, Cement and Concrete Association, Wexham<br />

Springs, Slough, UK, 1974<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 29 of 220


Cement<br />

Concrete<br />

Cement Properties<br />

1. INFLUENCE OF CEMENT COMPONENTS ON CEMENT PROPERTIES ........................ 31<br />

2. LITERATURE......................................................................................................................47<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 30 of 220


Cement<br />

Concrete<br />

1. Influence of Cement Components on Cement Properties<br />

1.1 General<br />

Due to the great variety of factors involved, it is difficult to describe precisely the relationship<br />

between the cement components and the cement properties. The available models for the<br />

prediction of cement performance usually only reflect the general trends.<br />

The effects on the cement properties are best understood for the clinker and gypsum. Least<br />

knowledge is available in case of the mineral components and the chemical admixtures, so<br />

that virtually the only way to assess their influence is to carry out performance tests.<br />

1.2 Clinker<br />

The composition of clinker gives some indications on the properties of cement to be<br />

expected, as it influences the rate of hydration reaction and thus the setting and hardening<br />

rate of cement The composition of clinkers control the quantity and rate of heat evolved<br />

during hydration and the resistance of cement to sulphate attack; therefore, limiting values<br />

are specified.<br />

In this section, the influence of composition of clinker on the following properties of cement<br />

shall be discussed:<br />

♦ water requirement of standard paste and consistency of concrete<br />

♦ stiffening rate and setting time of standard paste and slump loss of concrete<br />

♦ heat of hydration<br />

♦ strength of mortar and concrete<br />

♦ sulphate resistance<br />

♦ other properties of concrete<br />

A summary on the relationship between clinker composition and the principal cement<br />

properties workability (water demand, setting) and strength is given in Table 2.<br />

Table 2: Effect of clinker composition on water requirement and setting time of<br />

standard paste and compressive strength of ISO mortar (general trends)<br />

Clinker Water<br />

requirement<br />

Setting time Strength<br />

early final<br />

C 3 S -- -- <br />

C 2 S -- -- <br />

C 3 A <br />

C 4 AF -- -- <br />

K 2 O -- <br />

Na 2 O -- <br />

SO 3 -- <br />

P 2 O 5 -- --<br />

increasing decreasing -- no effect<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 31 of 220


Cement<br />

Concrete<br />

1.2.1 Water requirement of standard paste and consistency of concrete<br />

The water requirement of the standard paste of normal consistency depends primarily on the<br />

aluminate and alkali content of clinker and on the fineness of cement.<br />

The relation between the water requirement of standard paste and the composition of<br />

cement cannot be applied to concrete, as there is a rather weak relationship between the<br />

water requirement of paste and water/cement ratio of concrete (Figure 1)<br />

Po - Water reducing admixture (Pozzolith)<br />

Me - Superplasticizer (Melment)<br />

Lu 2 ... Ho 2 - various OPC<br />

Gm, Du, etc. - various Group plants<br />

Figure 1: Water requirement of cement and w/c-ratio of concrete<br />

The effect of cement on the consistency or water requirement of concrete is rather small<br />

compared to other factors, such as sand, admixtures and temperature. An exception is<br />

concrete with a very short mixing time, where cement with false set may seriously impair the<br />

consistency of concrete.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 32 of 220


Cement<br />

Concrete<br />

1.2.2 Rate of stiffening and setting time of standard paste and slump loss of<br />

concrete<br />

The stiffening rate or the „Vicat“ setting time of the standard paste is significantly influenced<br />

by the composition of clinker. The sulphates and phosphates of clinker usually delay,<br />

whereas aluminate shortens the setting time of cement.<br />

The relation between the stiffening rate or setting time of standard paste and the stiffening<br />

rate - expressed as slump loss - of concrete is, just as for the water requirement, rather poor.<br />

Therefore, it is difficult to estimate the stiffening rate of concrete on the basis of composition<br />

or fineness of cement.<br />

1.2.3 Heat of hydration<br />

The effect of the clinker composition on heat of hydration has already been discussed in<br />

detail in the paper on cement hydration. The principal way to control the heat evolution of the<br />

clinker is the adjustment of the C 3 S and C 3 A content.<br />

1.2.4 Strength of mortar and concrete<br />

The rate of strength development of mortar or concrete depends on the type (or composition)<br />

of cement. The general tendency of cements with a slow rate of hardening is to have a<br />

slightly higher ultimate strength.<br />

The ASTM type IV cement, with low content of C 3 S, has the lowest early strength, but<br />

develops the highest ultimate strength (Figure 2). This agrees with the influence of individual<br />

clinker components on the rate of strength development measured on pure clinker minerals<br />

(see Figure 3).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 33 of 220


Cement<br />

Concrete<br />

Figure 2:<br />

Strength development of concrete made with different cement types<br />

Figure 3:<br />

Compressive strength of cement compounds<br />

The two calcium silicates develop the highest strength, but at different rates. The aluminate<br />

develops little strength, despite a high rate of hydration.<br />

The rate of strength development of mortar or concrete depends on the clinker composition<br />

as follows:<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 34 of 220


Cement<br />

Concrete<br />

a) Calcium silicates. The different rates of hydration of C 3 S and C 2 S affect the rate of<br />

hardening in a significant manner: A convenient rough rule assumes that C 3 S contributes<br />

the most to the strength development during the first four weeks and C 2 S afterwards. In<br />

general, somewhat higher ultimate strengths are reached by cements with lower calcium<br />

content, i.e. rich in C 2 S. This observation corresponds with the assumption that the<br />

strength of cement depends on the specific surface of its hydration products. C 2 S<br />

produces more colloidal CSH gel and less of the crystalline Ca(OH) 2 than the C 3 S.<br />

b) Aluminates and ferrites. The influence of the other two major components on the<br />

strength development is still controversial. Presumably, the C 3 A contributes to the<br />

strength of the cement paste during a period of one to three days. In general, both<br />

aluminate and ferrite contribute to the strength of cement to a minor extent, but<br />

significantly influence the hydration process of the silicates and thus have an indirect<br />

effect on the rate of hardening.<br />

c) Of the minor components, the alkali sulphates exert the greatest influence on the rate of<br />

hardening. The alkali sulphate - mostly present as easily soluble potassium sulphate or<br />

calcium-potassium sulphate with a molar ratio of 2:1 to 1:2 - accelerates the rate of<br />

hardening, improving the early strength and decreasing the 28 day and ultimate strength<br />

(see Figure 4). Of the other minor components, fluorine accelerates, whereas the<br />

phosphorous compound delays the rate of hardening.<br />

d) Clinker characteristics other than chemical composition. Particularly the burning and<br />

cooling conditions influence the rate of hardening of a particular clinker composition.<br />

Frequently, clinkers of the same chemical composition have different strengths and<br />

clinkers of different chemical composition have the same strength. A simple experiment<br />

proves that the very same clinker composition may have rates of hardening which vary<br />

considerably. Reburning of a clinker in a laboratory furnace changes the rate of<br />

hardening, but does not affect the chemical composition of clinker (see Figure 5).<br />

Figure 4: Effect of soluble K 2 O on the compressive strength of ISO mortar<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 35 of 220


Cement<br />

Concrete<br />

A = Clinker of high activity<br />

B = Clinker of low activity<br />

Maturity ≅ Degree of hydration<br />

Figure 5: Model of strength development of mortar and concrete made with two<br />

clinkers of same chemical composition and different activity<br />

A general guide on the necessary amount of the main clinker phases to achieve optimum<br />

strength development is given in Table 3. The most essential point is to have a high C 3 S<br />

content (in the order of 60%) and to adjust the C 3 A content.<br />

Table 3: "Ideal" composition of the clinker for optimum strength development<br />

Clinker phase "Ideal" content (%)<br />

C 3 S 55 - 65<br />

C 2 S 15 - 25<br />

C 3 A 7 - 10<br />

C 4 AF 7 - 10<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 36 of 220


Cement<br />

Concrete<br />

The influence of the clinker composition on the standard mortar strength is noticeably<br />

reduced in concrete. Depending on the quality of cement, sand and aggregate, the<br />

proportioning of concrete or mortar components, curing temperature, specimen dimension,<br />

the rate of hardening in mortar and in various concrete compositions is quite different.<br />

Moreover, the use of admixtures in concrete - which is common practice today - makes the<br />

relation even more complicated. Due to different hardening rates of mortar and concrete, the<br />

relation between concrete and mortar strength at various ages varies and depends on the<br />

above mentioned factors. Concluding, the cement properties, as demonstrated through<br />

standardised testing methods, do not show their effect in the same way in concrete.<br />

1.2.5 Sulphate resistance<br />

The sulphate resistance of concrete depends primarily on the C 3 A content of clinker. The<br />

ferrite phases (C 4 AF) affect the sulphate resistance to a much lesser degree. The higher the<br />

C 3 A content of clinker, the more susceptible the concrete is to sulphate corrosion (Figure 6).<br />

E 28 after 28 days of regular curing = 100%<br />

E 180 after 28 days of regular curing and 180 days of<br />

exposure to 10% sodium sulphate solution<br />

Figure 6: Sulphate resistance of cement measured on ISO mortar specimen (55 OPC)<br />

Influence of C 3 A content on the loss of Young’s modulus of elasticity E (determined<br />

from ultrasonic pulse velocity measurements)<br />

The rate of sulphate corrosion depends - apart from the C 3 A content of clinker - on factors<br />

other than cement:<br />

♦ composition of concrete, particularly the water/cement ratio<br />

♦ age of concrete at the time of the first exposure to sulphates<br />

♦ type and concentration of sulphate solution<br />

♦ duration and mode of sulphate exposure<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 37 of 220


Cement<br />

Concrete<br />

1.2.6 Other properties<br />

The other properties of concrete, such as<br />

♦ freeze - thaw - resistance<br />

♦ permeability<br />

♦ cracking<br />

♦ shrinkage and creep<br />

are only slightly influenced by the composition of clinker and quality of cement. Other<br />

influencing factors, such as air content, w/c-ratio, curing conditions, are decisive.<br />

The cement exerts only an indirect influence on these properties by its effect on the water<br />

requirement and rate of hardening.<br />

1.3 Mineral components<br />

The effect of the mineral components on cement performance can be related mainly to their<br />

activity. The three main classes of materials in this respect are latent hydraulic (e.g. blast<br />

furnace slag), pozzolanic (e.g. fly ash and natural pozzolans) and inert (e.g. limestone).<br />

In case of the active mineral components (latent hydraulic and pozzolanic), the general<br />

effects with respect to cement properties are as follows:<br />

♦ lower water requirement (except for natural pozzolans)<br />

♦ delay in setting times<br />

♦ lower heat of hydration<br />

♦ lower early strength<br />

♦ higher long term strength<br />

♦ lower permeability<br />

♦ improved resistance to sulphate and other chemical attacks<br />

♦ lower sensitivity for alkali-aggregate reaction<br />

The actual influence on the cement properties will of course still depend on the individual<br />

nature of each material. A more detailed comparison on the effects of the main active mineral<br />

components blast furnace slag, fly ash and natural pozzolan (at same dosage) is made in<br />

Table 4.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 38 of 220


Cement<br />

Concrete<br />

Table 4: Effect of main active mineral components on cement properties (general<br />

trends)<br />

Blast furnace slag Fly ash (class F) Natural pozzolan<br />

Water requirement 0 <br />

Setting time <br />

Heat of hydration <br />

Early strength <br />

Final strength <br />

Sulphate resistance <br />

Permeability (chloride) <br />

Alkali-aggregate<br />

<br />

reactivity<br />

Shrinkage 0 0 <br />

<br />

<br />

O neutral effect<br />

increase<br />

decrease<br />

The inert mineral components like limestone do exert similar influences as the active<br />

materials in terms of water requirement, setting and heat of hydration, but they will not<br />

improve the final strength and durability characteristics of the cement.<br />

Other cement properties than the above mentioned are generally not affected to a great<br />

extent by the addition of mineral components.<br />

1.4 Gypsum<br />

The main function of the gypsum in cement is to regulate the cement setting, but the gypsum<br />

also influences other cement properties such as grindability, flowability and storage stability,<br />

volume stability and strength.<br />

The use of anhydrite instead of gypsum helps to reduce the risk of false setting and to<br />

improve the storage stability and flowability of the cement at high grinding temperature (see<br />

also chapter 5.6). In case of highly reactive clinkers, proper set retardation may, however, be<br />

a problem and blends with gypsum have to be used.<br />

The substitution of natural gypsum by by-product gypsum may sometimes cause problems<br />

with setting and strength development due to potential presence of impurities in such type of<br />

materials.<br />

1.4.1 Specific surface area<br />

The specific surface area of cement is usually determined by the Blaine method. The Blaine<br />

value is calculated from the air permeability of a cement sample compacted under defined<br />

conditions. The resistance to air flow of a bed of compacted cement depends on its specific<br />

surface. The Blaine specific surface is not identical with the true specific surface of the<br />

cement, but it gives a relative value, which suffices for practical purposes.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 39 of 220


Cement<br />

Concrete<br />

An absolute measurement of the specific surface can be obtained by the nitrogen (or water<br />

vapour) absorption method - BET. In this method, the "internal“ area is also accessible to the<br />

nitrogen molecules and the measured value of the specific surface is therefore considerably<br />

higher than that determined by the air permeability method:<br />

Method Blaine Nitrogen Absorption (BET)<br />

Cement A 2’600 cm 2 /g 7’900 cm 2 /g<br />

Cement B 4’150 cm 2 /g 10’000 cm 2 /g<br />

The Blaine value can sometimes be misleading, especially in the case of outdoor stored<br />

clinker, composite cements - consisting of a more easily grindable component - and clinkers<br />

containing underburnt material which is easier to grind. The properties of such cements can<br />

often be poorer compared to other ground to the same specific surface.<br />

1.4.2 Particle size distribution<br />

Cements of the same specific surface may have different PSD and different properties. Thus,<br />

the specific surface is not the only fineness criterion determining the properties of a particular<br />

cement composition.<br />

The determination of the PSD can be carried out by the following methods:<br />

♦ mechanical sieving (residues on sieves of a definite size (e.g. 32, 45 and 60 µm))<br />

♦ laser and sedigraph (residues over the whole range of particles sizes)<br />

The mechanical sieving is usually applied in the cement plants. Due to the limitations in sieve<br />

sizes, this method does not allow to measure the whole range of particle sizes.<br />

The overall particle size distribution of cement is commonly analysed by means of the<br />

theoretical distribution according to Rosin-Rammler-Sperling (RRS), which is described be<br />

the following formula:<br />

ln [ln (100/R d )] = n [ln (d) - ln (d')]<br />

being:<br />

R d = % of particles with diameter greater than d (residue)<br />

d = particle size in µm<br />

d' = characteristic diameter in µm (36.8% of the particles greater than d')<br />

n = slope of RRS straight line<br />

The data obtained in the particle size analysis is accordingly plotted in a so-called RRSdiagram<br />

(see Figure 7), having a double logarithmic ordinate (y-axis) and a logarithmic<br />

abscissa (x-axis). After linear regression of the particle size distribution, the slope n of the<br />

straight line and the characteristic diameter d' (at 36.8% residue) can be calculated.<br />

The slope n and diameter d' are the significant values for the particle size distribution. The<br />

first characterises the degree of distribution (wide-narrow), whereas the second one states its<br />

location and is an indicator for the overall fineness. High n values results from a narrow PSD<br />

and low d' values from a high overall fineness.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 40 of 220


Cement<br />

Concrete<br />

Differences in the PSD of cement can also be seen in the relation between the traditionally<br />

measured Blaine values and sieve residues. At same sieve residue, the Blaine tends to be<br />

lower with higher n values (see also Figure 8).<br />

Figure 7: Particle size distribution of cement in RRS-diagram<br />

Figure 8: Correlation between Blaine and residue 32 µm for different n values (data<br />

FLS)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 41 of 220


Cement<br />

Concrete<br />

1.5 Grinding of Portland cements<br />

1.5.1 Influence of Blaine fineness<br />

Since the hydration starts on the surface of the cement particles, it is the specific surface<br />

area of Portland cement that largely determines the rate of hydration and thus the setting and<br />

hardening rate. To achieve a faster hydration and strength development, rapid-hardening<br />

cements are ground finer than ordinary Portland cement. It is common practice to produce<br />

cement of various strength classes from one clinker by altering the fineness to which it is<br />

ground. The Blaine value of cement varies between 2’500 cm 2 /g for ordinary Portland cement<br />

(type I, ASTM) and 5’000 cm 2 /g for high early strength cement (type III, ASTM).<br />

The rate of hydration is slowed down by the presence of cement gel and if a large quantity of<br />

gel is formed rapidly, because of a large cement surface, the inhibiting action of the gel soon<br />

takes place. For this reason, extra fine grinding is efficient only for the early strength up to 7<br />

days. Moreover, the rate at which the strength of concrete increases is substantially lower<br />

than that of mortar (see Figure 9).<br />

Considering the energy consumption for grinding, the fine grinding is often not economically<br />

feasible. In those cases, where high early strength is not required, fine grinding is of little<br />

value (see Figure 10). A large number of concrete applications are unable to exploit the<br />

effects of fine grinding.<br />

The relations between the Blaine fineness of cement and concrete properties can be<br />

summarised as follows:<br />

1) Increasing the fineness of cement, reduces the amount of bleeding in concrete (see<br />

Figure 11).<br />

2) Increasing the fineness of cement above 3000, increases somewhat the water<br />

requirement of concrete. Compared to the influences other than cement on the water<br />

requirement of concrete, the influence of cement fineness is considerably smaller.<br />

3) The strength of concrete is influenced by the fineness of cement. The early compressive<br />

strength increases with an increase in cement fineness. The difference in compressive<br />

strength due to the difference in fineness of cement, is considerably smaller at 28 days<br />

and at later age (see Figure 9).<br />

4) The fineness of cement influences the drying shrinkage of concrete. When the water<br />

content is increased because of fineness, the drying shrinkage is increased.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 42 of 220


Cement<br />

Concrete<br />

Figure 9: Effect of cement fineness on strength of mortar and concrete<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 43 of 220


Cement<br />

Concrete<br />

Figure 10: Relative specific energy consumption and compressive strength<br />

development<br />

Figure 11: Effect of cement fineness on bleeding of concrete (non air-entrained concrete,<br />

w/c-ratio = 0.57)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 44 of 220


Cement<br />

Concrete<br />

1.5.2 Influence of particle size distribution<br />

The influence of the fineness on the cement properties can be described more precisely,<br />

when taking into account the PSD of the Portland cement. The PSD is of particular<br />

importance with respect to workability and strength development.<br />

The workability of Portland cement and concrete may impair when the PSD becomes<br />

narrower (at constant specific surface). On one hand, the water requirement for a certain<br />

consistency tends to increase and, on the other hand, the faster conversion of aluminate at<br />

narrow PSD may lead to early stiffening problems.<br />

The mentioned stiffening problems may especially occur if clinkers of high reactivity (high<br />

C 3 A and alkali content) are ground together with the gypsum at low temperatures (little<br />

formation of easily soluble sulphates), as it is the case in the modern grinding systems. With<br />

such clinkers, the proper adjustment of the wideness of the PSD and/or of the calcium<br />

sulphate carrier is therefore important.<br />

The effect of the PSD of Portland cement on strength development is not always clear. The<br />

general trends can be summarised in the following way:<br />

♦ The most valuable particles for early strength are the ones between 0 - 8 µm. The Blaine<br />

value is thus a good indicator in this respect, as it is proportional to the portion in this<br />

fraction.<br />

♦ The 28 day strength is mainly controlled by the amount of particles in the range between<br />

2 - 24 µm, which is proportional to steepness n of the PSD.<br />

The increase in the steepness n at a given Blaine is accordingly an effective means in<br />

improving the strength potential at 28 days as illustrated in Figure 12. The positive effects of<br />

higher n values are, however, less pronounced on concrete.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 45 of 220


Cement<br />

Concrete<br />

Figure 12: 2 day and 28 day compressive strength of Portland cement, as a function of<br />

the specific surface area and the slope of the RRS-distribution of the cement<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 46 of 220


Cement<br />

Concrete<br />

2. Literature<br />

Cement components and cement properties<br />

Gebauer, J., Kristmann, M., The influence of the composition of industrial clinker on cement<br />

and concrete properties, World Cement Technology, March 1979, pp. 46 - 51<br />

Wolter, H., Production, properties and applications of Portland slag cements and blast<br />

furnace cements, Concrete Workshop, Queensland Cement Ltd., 1994<br />

Malhotra, V.M., Ramezanianpour, A.A., Fly ash in concrete, Second edition,CANMET,<br />

Ontario, 1994, 307 pp.<br />

Massazza, F., Pozzolana and pozzolanic cements, in: Lea's Chemistry of Cement and<br />

Concrete, Fourth Edition, Arnold, 1998, pp. 471 - 631<br />

Cochet, G, Sorrentino, F., Limestone filled cements: properties and uses, in: Progress in<br />

Cement and Concrete, Volume 4, Mineral admixtures in cement and concrete, ABI, New<br />

Delhi, 1993, pp. 266 – 295<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 47 of 220


Cement<br />

Concrete<br />

Clinker Optimization<br />

Dr. J.A. Imlach<br />

1. INTRODUCTION............................................................................................................... 49<br />

2. CLINKER COMPOSITION ASPECTS.............................................................................. 50<br />

2.1 Main Elements (SiO2, Al2O3, Fe2O3 and CaO)...................................................... 50<br />

2.2 Principal Accompanying Elements (MgO, Mn 2 O 3 , SrO, and TiO 2 ,) .......................... 51<br />

2.3 Alkalis, Suphur and Chloride.................................................................................... 52<br />

2.4 Disturbing Minor Species P 2 O 5 B2O3, and (F)......................................................... 53<br />

2.5 Trace Heavy Toxic Elements. .................................................................................. 55<br />

3. CLINKER PRODUCTION ASPECTS ............................................................................... 56<br />

3.1 Kiln Type .................................................................................................................. 56<br />

3.2 Raw Meal Preparation.............................................................................................. 56<br />

3.3 Alite Size .................................................................................................................. 56<br />

3.4 Alite Polymorphism .................................................................................................. 57<br />

3.5 C 3 A Polymorphic Modification.................................................................................. 57<br />

3.6 Clinker Microstructure .............................................................................................. 58<br />

3.7 Rate of Clinker Cooling ............................................................................................ 59<br />

3.8 Reducing Conditions ................................................................................................ 60<br />

4. IMPLEMENTATION OF FINDINGS.................................................................................. 60<br />

4.1 Introduction .............................................................................................................. 60<br />

4.2 Barriers to Implementation ....................................................................................... 60<br />

4.3 Fine Tuning of Material Aspects............................................................................... 61<br />

4.4 Fine Tuning of Process Aspects .............................................................................. 62<br />

5. LITERATURE CITED........................................................................................................ 64<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 48 of 220


Cement<br />

Concrete<br />

1. Introduction<br />

As a result of ever increasing environmental and economical pressures, cement companies<br />

are now producing composite cements, with ever-increasing degrees of clinker substitution.<br />

Whilst this has many benefits, there are also some drawbacks, one of these being a poorer<br />

strength development at early ages, e.g. 1 and 3 days.<br />

As a means of compensating for this reduced early strength performance, one possibility is<br />

to increase the early strength performance of the clinker being used. The present report is an<br />

attempt to review, based on a limited number of the many thousands of reports available, the<br />

influence of those factors which influence strength development.<br />

The present survey has had to be of necessity limited in scope and is restricted to<br />

determining the influence of clinker alone on strength development. Other equally important<br />

and highly relevant aspects such as fineness of grinding, grain size distribution, optimisation<br />

of set retarder, use of chemical admixtures etc. lie outside the terms of reference.<br />

It has furthermore been attempted to give priority to that literature concerning clinkers<br />

produced at least from plant raw meals and preferably in industrial kilns. This has not always<br />

been possible. Articles referring to clinker or individual clinker phases produced from pure<br />

oxides in laboratory furnaces have been on the whole not used.<br />

To bring order into this report it has been necessary to break up the literature review into<br />

those individual aspects having an influence on early strength development. Thus the review<br />

of any one paper may lead to entries under several individual topics.<br />

In the present review the following have been chosen as topics / sub-topics:<br />

Chemical Composition Characteristics<br />

♦ main elements (SiO2, Al2O3, Fe2O3 and CaO), with those characteristics derived from<br />

them such as LSF, SR and AR and the Bogue-calculated C3S, C2S, C3A and C4AF<br />

♦<br />

♦<br />

♦<br />

♦<br />

principal accompanying elements MgO, TiO2, Mn2O3<br />

soluble, volatile elements SO3, K2O, Na2O and Cl,<br />

disturbing minor elements B, P2O5 and (F)<br />

trace heavy elements such as Cr, Ni, Pb, Zn, etc.<br />

Clinker Production Characteristics:<br />

♦ kiln type<br />

♦ raw meal preparation<br />

♦ clinker microstructure (alite size, crystallisation of liquid phase, etc.)<br />

♦<br />

♦<br />

♦<br />

polymorphic modifications<br />

rate of clinker cooling<br />

kiln atmosphere<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 49 of 220


Cement<br />

Concrete<br />

2. Clinker composition aspects<br />

2.1 Main Elements (SiO2, Al2O3, Fe2O3 and CaO)<br />

Many articles exist on the influence of the main oxides on strength development properties<br />

and so those now quoted are only a selection of those available.<br />

One of the earliest and still most comprehensive statistical evaluations was performed by<br />

Blaine et al. (1965, 1968) at the US National Bureau of Standards on 199 commercially<br />

available Portland cements (Types 1 to 5) in the 1960’s. Here not only the main elements<br />

were taken into consideration but almost all of the other elements dealt with in Chap. 2.<br />

From Blaine’s work the most relevant in the present case is the compressive strength<br />

development at 1 and 3 days according to ASTM specifications. This was based on 2-inch<br />

mortar cubes that had been stored in water after an initial 24 hrs in their moulds in a moist<br />

cabinet. For strength development at 1 day Blaine (1968) reported 8 different equations<br />

which had been drawn up. The most dominant factors which appeared in these equations<br />

were C3S, SO3, insoluble residue, and K 2 O. For example in his equation 1, the positive<br />

contributing factors were as follows:<br />

+ 18.29xC 3 S +225.8xSO 3 + 376.1xK 2 O<br />

Also included were a negative constant and negative factors for L.o.I., insoluble residues.<br />

The work of Blaine is so comprehensive that for the full benefit it must be read in its entirety.<br />

Alexander et al. (1968) performed a statistical investigation into the strength development of<br />

commercially available Portland cements in Australia. The latter had a wide range of<br />

chemical composition, surface areas and a wide range of strength development<br />

characteristics. These authors considered that the main differences in strengths, when tested<br />

under the following conditions:<br />

♦ curing time between 1 day and 6 months<br />

♦ W/C ratios from 0.35 to 0.80<br />

♦ curing temperatures from 5 to 75 °C<br />

were most sensitive to the C 3 A content. Other factors such as particle size had a smaller but<br />

still significant effect, whilst differences in C 3 S were considered to be of minor importance.<br />

With specific regard to curing times of 7 days and less, Alexander et al. did consider that the<br />

C 3 S content was of importance.<br />

Gebauer et al. (1980), on the basis of the “Comparative Study 1976/77” evaluation of 55<br />

“Holderbank” Group plant clinkers, found that the chemical composition of clinker is the<br />

decisive factor influencing the properties of cement. The compressive strength development<br />

of laboratory cements prepared from them is correlated with the quantity of the principal<br />

clinker phases, and the alkalis and SO 3 as follows:<br />

CS 2-day (ISO) = -2.3 + 0.22 C 3 S + 0.40 C 3 A + 3.8 KNS (r = 0.839)<br />

CS 2-day (ISO) = -1.42 + 0.19 C 3 S + 0.35 C 3 A + 8.5 K sol (r = 0.876)<br />

(KNS represents K 2 O + Na 2 O + SO 3 and K sol soluble K 2 O)<br />

Stürmer et al. (1990) reported that the C 3 A content of clinker in the range 0 to 13 % had no<br />

influence on compressive strength development at 1 and 3 days. Only at 28 days was a<br />

definite strength increase noted up to 10 % C 3 A, thereafter a drop occurred. (Here It must be<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 50 of 220


Cement<br />

Concrete<br />

stated that modification of the original industrial raw mix reduced the original Bogue<br />

calculated C 3 S from 75 % down to 57 %).<br />

Munoz et al. (1995), when considering hourly average clinkers from a single plant, found that<br />

the C 3 A content had a negative influence on strength development and claim that this might<br />

be due to the hydration products being poorly crystallised.<br />

Ghosh et al. (1998) investigated the factors influencing variations in day to day strength<br />

development for two cement plants in India, using quality data gathered over a period of 3 to<br />

4 months. Here the naturally occurring variations in chemical composition, fineness of<br />

grinding etc. were taken into consideration. They found that the principal factors influencing<br />

strength development within each plant were CaO content, fineness of grinding and clinker<br />

microstructure. For these two plants, the 3-day strength was increased by 15 to 30 % when<br />

the clinker CaO was increased by 2 %, e.g. from 63.5 % to 65.5 % with the C 3 S increasing by<br />

15 %-age points. It was reported that plant B exhibited a higher strength not only because of<br />

a slightly higher C 3 S content, but also due to its microstructure.<br />

2.2 Principal Accompanying Elements (MgO, Mn 2 O 3 , SrO, and TiO 2 ,)<br />

2.2.1 MgO<br />

It is well known that clinker contains varying amounts of MgO, mainly due to the presence of<br />

dolomite in the raw materials. Up to 2.0 % MgO can be incorporated into the clinker minerals,<br />

mainly in the alite, before MgO occurs as an undesirable phase, responsible for expansion in<br />

mature concrete. The presence of up to 2 % MgO in clinker, as a substitute for CaO, can<br />

thus increase early strength development in that its presence increases the effective LSF<br />

and hence the quantity of C 3 S actually present. In clinkers with very high LSF values in the<br />

order of 100, the presence of MgO can lead to over-saturation and the presence of free CaO.<br />

Blaine (1968) did not consider MgO to influence strength development at early ages.<br />

2.2.2 Mn 2 O 3<br />

On average Mn 2 O 3 is present in clinker at about the 0.1 % level. Unlike the other transition<br />

elements V, Cr, Co, Ni and Cu, Mn along with Fe and Ti is one of the transition elements not<br />

associated with toxic properties. From an environmental viewpoint no reason is present to<br />

limit the quantity of Mn 2 O 3 and in previous years one Group plant (Westport) produced a<br />

clinker with up to 1.2 % Mn2O3. This was due to the high natural level of this element in the<br />

quarry materials.<br />

As presented by Moir (1992), Mn 2 O 3 is mainly located in the ferrite phase and so its presence<br />

in large quantities would increase the quantity of ferrite and decrease that of aluminate. It<br />

thus in general reduces early strength development as reported by Knöfel et al. (1983)<br />

although these authors did indicate that a strength increase was obtained for 0.5 % MnO 2 at<br />

up to 28 days. Although Knöfel’s work was based on cements produced from pure oxides,<br />

Alexander et al. (1968) found, on the basis of a statistical evaluation of industrial clinkers,<br />

that a significant correlation existed between compressive strengths at less than 7 days and<br />

the level of Mn 2 O 3 up 0.28 %. The statistical work of Blaine (1968) indicated that Mn 2 O 3<br />

negatively influenced strength development at 3 days but not at 1 day.<br />

2.2.3 SrO<br />

According to Group plant data SrO is on average present in clinker in concentrations of 0.11<br />

%, but can be as high as 0.51 %. Evidence is available (Bürki 1993) that SrO at higher levels<br />

can reduce early strength development potential of Portland cement clinker. At Italcementi’s<br />

Vibo Valenta plant, the limestone contains SrSO 4 , which is limited so as to result in a<br />

maximum of 1.5 % SO 3 in clinker. With increasing quantities of Sr the alite becomes unstable<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 51 of 220


Cement<br />

Concrete<br />

and results in CaOf that cannot be reduced by harder burning. By the addition of fluorspar to<br />

the raw mix it is possible to compensate for the presence of SrO and obtain a normal free<br />

CaO level. According to Blaine (1968), SrO is associated with lower strengths at 1 and 3<br />

days when present in the range up to 0.4 %.<br />

2.2.4 TiO 2<br />

The average level of TiO 2 in clinker is about 0.28 %, but values of up to 0.75 % have been<br />

found at the Joliette plant. TiO 2 is preferentially located in the ferrite phase (up to 2. 0 %, but<br />

can be found at lower concentrations in belite (0.75 %) and at the 0.4 % level in alite and<br />

aluminate.<br />

With regard to its influence on phase development and compressive strength development<br />

Knöfel (1977) reported that for laboratory produced clinkers increasing the TiO 2 content up to<br />

1 % increased compressive strengths did cause an increase in strength in spite of a<br />

decrease of C 3 S, but only at 7 days and more. At 2 days no increase was observed.<br />

Blaine (1968) does not report an influence of TiO 2 on strength development at 1- and 3-days.<br />

2.3 Alkalis, Suphur and Chloride<br />

According to the statistical investigation of Blaine (1968), K 2 O and SO 3 are principal factors<br />

contributing to early strength development at 1- and 3-days. Na 2 O was not considered to<br />

have an influence. Osbaek (1979) determined the strength development characteristics of<br />

laboratory cements having the same fineness and gypsum content, produced from industrial<br />

clinkers from 31 plants. He found a better correlation between soluble alkalis than the total<br />

alkali content. Whereas alkalis have a positive influence on 1- and 3-day strengths, they<br />

have a negative influence at 28 days where a 1 % soluble K 2 O eqvt decreases compressive<br />

strengths by 10 MPa. When 2 % K 2 SO 4 was added to a chosen cement the strength<br />

developments changed as follows:<br />

♦<br />

♦<br />

♦<br />

♦<br />

1-day: increase from 10 to 16 MPa<br />

3-day: increase from 22 to 30 MPa<br />

7-day: no change<br />

28-day: reduction from 50 down to 40 MPa.<br />

Strunge et al. (1985) investigated the influence of higher levels of SO 3 in clinker on clinker<br />

mineralogy and strength development. Both laboratory and industrial clinkers were<br />

examined, with SO 3 contents ranging around 3 % by weight. On the basis of their<br />

investigations they found that:<br />

♦<br />

♦<br />

♦<br />

increasing SO 3 levels in clinker results in lower alite and higher belite contents (at same<br />

LSF)<br />

higher SO 3 levels reduce compressive strengths, but less than expected from the<br />

reduction of the alite content<br />

at more than 2 % SO 3 in clinker and alkalis > 1 %, a normal setting behaviour can be<br />

obtained without the addition of a retarder.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 52 of 220


Cement<br />

Concrete<br />

2.4 Disturbing Minor Species P 2 O 5 B2O3, and (F)<br />

2.4.1 P2O5<br />

P 2 O 5 can occur in some areas (e.g. Uganda) in limestone and other natural raw mix<br />

components as well as in industrial by-products. When present in clinker in sufficient<br />

quantities P 2 O 5 has a deleterious effect on early strength development.<br />

The explanation, as reported by Gutt (1968) and Moir (1992), is that P 2 O 5 preferrentially<br />

enters the belite lattice and forms a complete solid solution series with C3P (3CaO.P 2 O 5 ), but<br />

is only incorporated to a minor extent (1.1 wt. %) into the alite lattice. With increasing<br />

quantities of P 2 O 5 in a raw mix, the amount of alite is reduced by approximately 10 % C 3 S for<br />

every 1 % P 2 O 5 in the clinker. This effect can be partially compensated by increasing LSF.<br />

Moir (1992) quoting unpublished Blue Circle data, shows the influence of P 2 O 5 content on 3-<br />

day compressive strengths. Basically in a cement with a Na 2 O eqvt of 0.8 % compressive<br />

strengths of approx. 20 N/mm 2 were observed up to 1 % P 2 O 5. At 2.0 % P 2 O 5 the 3-day<br />

strength had dropped to only 10 N/mm 2 . Increasing the Na 2 O eqvt resulted to 1.5 % results in<br />

the 3-day strengths increasing to about 15 N/mm 2 .<br />

According to Blaine (1968) small quantities (< 0.5 % P) can have a negative influence on<br />

early strength development.<br />

2.4.2 B 2 O 3<br />

No information is currently available at present on the levels of B 2 O 3 naturally present in<br />

clinker, but it is assumed that these are quite low That B 2 O 3 cannot be determined by XRF is<br />

perhaps one reason why this should be the case.<br />

From phase diagram studies it is known that B 2 O 3 is similar to P 2 O 5 in that with increasing<br />

quantities it influences phase equilibrium and reduces the quantity of C 3 S (Fletcher and<br />

Glasser 1993). It hence reduces early strength development. Again B 2 O 3 is appreciably<br />

soluble in C 2 S (17 mole %), but only poorly soluble in C 3 S. The presence of B 2 O 3 is therefore<br />

not compatible with the production of high alite, high early strength cements.<br />

According to Moir (1992) the presence of 2.8 % B 2 O 3 results in alite becoming so unstable at<br />

increasing temperatures that at 1700 °C it has no primary phase field in the phase diagram.<br />

On the basis of it being a well known flux in the ceramic industry it was hoped that it must<br />

also have the same effect on clinker burning. Tests at CTS/MT laboratories (Imlach 1994)<br />

showed that whilst the use of colemanite (2CaO.3B 2 O 3 .5H 2 O) did improve burnability, as<br />

determined by the size of the alite crystals, it did result in increasing quantities of free CaO<br />

and a corresponding reduction of C 3 S.<br />

The use of borate glasses as “alkali correctives” is therefore not to be undertaken without<br />

previous evaluation of the expected reduction of C 3 S.<br />

2.4.3 Fluoride<br />

The influence of fluoride is much more complex than that of P 2 O 5 and B 2 O 3 and depending<br />

on the amount present - and on burning conditions - can have the following effects on early<br />

strength development:<br />

♦<br />

Increase strength by acting as a flux and hence enabling manufacture of high alite<br />

clinkers<br />

♦ Increase early strength development by the production of clinker containing C 11 A 7 CaF 2<br />

as aluminate phase (Jetset cement)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 53 of 220


Cement<br />

Concrete<br />

♦ Offset strength reductions due to B 2 O 3 and P 2 O 5<br />

♦ Reduce early strength development by acting as a powerful set retarder<br />

2.4.4 Use as Burning Flux<br />

This is the effect that is presently being used by a few plants of the “Holderbank Group”,<br />

(Apasco Group and Cerro Blanco). Here the use of fluorspar, which results in clinker levels of<br />

0.22 to 0.28 % F, is being used to manufacture products with LSF’s in the range 101 to 103,<br />

which contain about 75 % alite and no belite as calcium silicates. Burning and cooling<br />

regimes are normal.<br />

2.4.4.1 PRODUCTION OF JETSET CEMENT<br />

The addition of fluorspar to raw meals rich in Al 2 O 3 can, under some circumstances, result in<br />

a clinker containing C 11 A 7 CaF 2 , and not the normal C 3 A, as aluminate phase. This change<br />

was made use of in a 1969 PCA patent application for the production of Jetset cement, which<br />

developed high very early strengths by containing some 0.5 % F and some 15 % C 11 A 7 CaF 2 .<br />

In laboratory trials at Holcim Group <strong>Support</strong>, Imlach (1974) showed that a lower burning<br />

temperature of 1350 °C was possible, but that a slow cooling of the liquid phase was<br />

necessary if the required C 11 A 7 CaF 2 was to be obtained instead of the normal C 3 A.<br />

Due to its lower hydraulic activity than C 3 A, it was considered that the setting and early<br />

strength development of Jetset cement would be controllable. This was found not to be the<br />

case and was very temperature sensitive. Set controllers containing K 2 SO 4 and citric acid<br />

were used at Holnam’s former Okay plant. Because of such problems the production of<br />

Jetset cement never became established.<br />

2.4.4.2 COMPENSATION FOR B 2 O 3 , P 2 O 5 AND SRO<br />

As previously reported both B 2 O 3 and P 2 O 5 can result in reduced early strength development.<br />

Both Moir (1992) and Gutt (1986) report on the use of F to compensate for the loss of<br />

strength incurred, this being achieved by destabilisation of the belite solid solution. However,<br />

this is a method that is not normally practised and will not be discussed further. The only<br />

practical application of fluoride to stabilise alite thus appears to be at Italcementi’s Vibo<br />

Valente plant to offset the effects of SrO as already indicated.<br />

2.4.4.3 RETARDING ACTION<br />

One of the certain drawbacks of producing clinker with fluoride is the increase in setting time<br />

that it causes. This is said to be due (Moir 1992) to the formation of an insoluble barrier layer<br />

of CaF2 on the surface of the hydrating clinker minerals. Whilst the alite structure can contain<br />

up to at least 2 % F (Paul and Glasser 1999), adding such levels to clinker would be both<br />

uneconomic and would result in an excessive setting time. According to Moir (1992)<br />

increasing the F content to 1.75 % results in setting times of around 10 hrs. During<br />

production trials at Torredonjimeno, Bachmann (1996) found that at 0.5 % % F in clinker,<br />

setting times of > 8 hrs were determined under laboratory test temperature of 20 °C.<br />

At lower temperatures the influence of F becomes even more troublesome. According to Moir<br />

(1992) the setting time at 5°C can be given by the equation:<br />

Setting time (min) = 360 + 500x F – 30xCaOf – 0.3 surface area.<br />

It is known that at one Blue Circle plant (Hope), which has F naturally present in the raw<br />

materials, the F content of the raw mix must be changed according to the time of year.<br />

According to Heinemeyer the presence of only 0.17 % F in clinker was sufficient to produce<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 54 of 220


Cement<br />

Concrete<br />

concrete which had excessive de-moulding times cements when ambient air temperatures<br />

dropped to the region of °C.<br />

2.5 Trace Heavy Toxic Elements.<br />

Even when produced from only natural raw components, Portland cement clinkers always<br />

contain trace amounts of heavy toxic elements in quantities roughly proportional to their<br />

concentration in the Earth’s crust. Several of the toxic elements are of importance in that<br />

their emission from the exhaust gas stack is limited by strict regulations. These elements are<br />

split into classes and their emission levels are for example limited by the German TA Luft<br />

regulations as follows:<br />

♦ Class 1 (Cd, Hg, Tl) 0.2 mg/Nm 3<br />

♦ Class 2 ( As, Co, Ni, Se, Te) 1.0 mg/N/m 3<br />

♦ Class 3 (Cr, Cu, Pb, Pd, Pt, Rh, Sb, Sn, V) 5.0 mg/Nm 3<br />

♦ Unclassified (Zn) no limit<br />

The levels of these elements in clinker normally range from < 0.01 ppm to round about 100<br />

ppm (Imlach 1997). On the basis of cements investigated in the mid 1960’s, Blaine (1968)<br />

concluded that the heavy toxic elements in the concentration levels naturally present had no<br />

assured influence on strength development, especially at early ages.<br />

With the use of magnesia chrome bricks, petcoke, tyres etc, the level of Cr can be as high as<br />

320 ppm, the level of Ni 400 ppm and the level of V 135 ppm and the level of Zn 530 ppm.<br />

Heavy toxic elements at these concentrations are not considered to influence strength<br />

development potential (Stephan 1999), either positively or negatively.<br />

In the early 1970’s Onoda did patent a “doped-alite” cement which was reported to have a<br />

superior early strength development. The basis of this was the addition of chromite ore which<br />

resulted in the clinker having a Cr 2 O 3 content of > 1.0 %. This clinker was never produced on<br />

a routine basis due to the increased costs by using chromite ore and due to the higher<br />

content of soluble Cr6 + . The latter is associated with a mason’s skin disease (chrome<br />

eczema) and in Scandinavian countries the quantity of soluble Cr6 + in clinker is limited to < 2<br />

mg/ l, i.e. < 2 ppm.<br />

In laboratory based trials various authors from academic circles have investigated the<br />

influence of heavy metals on setting time and strength development. Here levels of up to 2.5<br />

% and above have been considered. Stephan et al. (1999) found that that Cr, Ni and Zn only<br />

had an influence on clinker properties when present in quantities > 5000 ppm, and therefore<br />

much higher than is normally the case. Schmidt (1980) showed that ZnO and PbO had no<br />

influence on strength development properties up to 7 days at a dosage rate of 1.0 % when<br />

incorporated into the lattice of the main clinker minerals. This is in contrast to the presence of<br />

soluble Pb and Zn salts in the gauging water.<br />

Today, even if a heavy toxic element were shown to vastly improve strength development<br />

characteristics, its routine implementation would definitely be opposed by environmental<br />

Groups. This would be on the grounds of increased stack emissions and on the leaching of<br />

such metals as Al, Cr etc. from concrete drinking water pipes.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 55 of 220


Cement<br />

Concrete<br />

3. Clinker Production Aspects<br />

3.1 Kiln Type<br />

According to Gebauer et al. (1980) the type of the kiln does not have a direct influence on<br />

cement properties<br />

3.2 Raw Meal Preparation<br />

Moir (1997) reports that in Blue Circle Industries models have been developed which<br />

satisfactorily predict the day to day variations in compressive strength development,<br />

especially at 28 days. However, significant differences exist between the predicted strength<br />

levels at different cement plants if a common expression is used, this being attributed to<br />

differences in microstructure.<br />

Moir’s study, performed on materials from 15 cement plants with 6 different kiln types,<br />

attempted to find out the influence of kiln feed heterogeneity and estimated burning<br />

temperature on clinker strength potential. No clear relationship was found between<br />

combination temperature and the main chemical parameters LSF and SR. A better<br />

correlation (r = 0.923) was found for the following expression:<br />

Temp (°C) for 1 % free CaO =<br />

414 + 21x acid soluble 90 µm residue+10 x LSF +3 x 150 µm residue + 32 x AR.<br />

Expressions were also derived for the strength of Type 1 (pure Portland) cements from the<br />

same 15 plants. Here the 2 day strengths according to EN 196-1 ranged from 24.8 to 31.9<br />

N/mm 2 and the 28 day values from 56.3 to 66.3 N/mm 2 . Some cements were “strength over<br />

achievers” i.e. had higher values than predicted, whilst others were “strength under<br />

achievers”.<br />

The “over achievers” were always those cements in which the 90 µm-residue of the raw mix<br />

had a higher LSF value than that of the bulk raw mix itself. The lower strength potential of<br />

clinkers with a lower LSF residue on 90 µm may be associated with the presence of belite<br />

clusters. No significant correlation was found between strength difference and estimated<br />

burning temperature.<br />

3.3 Alite Size<br />

Gebauer et al. (1980) consider that the clinker microstructure is of secondary importance<br />

regarding compressive strength development. Increasing the size of the alite usually leads to<br />

slower stiffening and setting times. For one extremely alite rich clinker (Höver), increasing the<br />

natural alite size by tempering the clinker in the laboratory had, with the same chemical<br />

composition, the following results:<br />

Original clinker<br />

Tempered clinker<br />

% alite < 20µm 80.9 63.6<br />

Penetration begin (hr) 1.37 1.88<br />

Setting begin (hr) 2.17 2.50<br />

2-day CS ISO (N/mm 2 ) 23.7 22.5<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 56 of 220


Cement<br />

Concrete<br />

28-day CS ISO (N/mm 2 ) 48.2 51.3<br />

Munoz et al. (1995), on the basis of a statistical investigation of 29 hourly average clinker<br />

samples from a single plant, considered that the crystallite size and strains of alite, as<br />

determined by x-ray diffraction by applying the variance procedure on the (021) peak,<br />

influenced compressive strength development. They reported that the crystallite size of has a<br />

positive influence on strength development that substantially decreases with age of testing.<br />

They also claim that strains in the alite crystals have an adverse effect on compressive<br />

strength development at all ages, even when increasing lattice defects and surface fineness<br />

result in faster rates of hydration. The latter was explained as being due to a “memory effect”<br />

of the defect content of the original alite.<br />

In his cement industry based PhD thesis, Platzer (1998) reports on factors influencing clinker<br />

reactivity. Starting point of his study was the undesirable situation that the introduction of<br />

tyres as secondary fuel (20 % of heat) had definitely resulted in a deterioration of cement<br />

properties, especially in the colder part of the year. In the short term the symptoms of the<br />

problem were relieved with a changed clinker composition and finer cement. This, however,<br />

brought higher costs and incurred logistic (storage) problems.<br />

On the basis of the ensuing study Platzer showed the influence of the following factors, at the<br />

same basic clinker chemistry, on compressive strength development:<br />

♦<br />

the Zn introduced by the tyres was not the reason, as the quantity was less than the<br />

demonstrated 0.5 % influence threshold<br />

♦ the mean alite size was increased from 31.5 µm to 39.7 µm by the use of tyres and<br />

showed a different size distribution<br />

♦ the increase in size was due to the much longer (too long) sinter zone<br />

♦<br />

the increased quantities of MgO, TiO 2 and Zn assist by acting as mineralisers.<br />

The temperature profile of the kiln, especially with the use of secondary fuel, therefore has to<br />

be matched with the alite size and the clinker quality desired. Here the use of microscopy is<br />

unavoidable.<br />

3.4 Alite Polymorphism<br />

Ono et al. (1968) considered that the high strength development associated with slow cooling<br />

down to 1200 °C could be explained as follows. “The crystalinity of the R-form of alite rising<br />

at the lower part of its temperature stability range, and the inversion of modifications may be<br />

disturbed, and alite is in an active state”.<br />

On the basis of laboratory produced alite and alite + C 3 A mixtures, Aldous (1983) found<br />

evidence to indicate that pure, synthetic alite doped with MgO, Al 2 O 3 , Fe 2 O 3 , TiO 2 and F -<br />

exhibited differing compressive strengths according to their polymorphic modification. Higher<br />

strengths were observed at 2 days and longer for the rhombohedral modification over the<br />

triclinic form, the latter displaying strengths similar to a commercial OPC. With slightly lower<br />

levels of dopants, triclinic modifications were formed with lower reactivities as determined by<br />

compressive strength, heat development and by X-ray diffraction determination of the degree<br />

of hydration. On comparing strength development after the same degree of hydration, Aldous<br />

found that rhombohedral alite still exhibited a higher value.<br />

3.5 C 3 A Polymorphic Modification<br />

The aluminate (C 3 A) phase can exist in either a cubic or an orthorhombic modification,<br />

depending on its content of the alkalis Na 2 O and K 2 O. In the earlier literature the alkali<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 57 of 220


Cement<br />

Concrete<br />

containing aluminate was given the formula NC 8 A 3 and KC 8 A 3 . According to Spierings<br />

(1976), the inclusion of Na 2 O into the C 3 A lattice results in a reduction of hydraulic activity as<br />

determined by heat evolution experiments.<br />

3.6 Clinker Microstructure<br />

Svinning et al. (1996, 1998) statistically investigated the influence of material and process<br />

parameters on the microstructure, and later compressive strength of clinkers from two<br />

Norwegian plants. Here the microstructure was partially investigated indirectly by X-ray<br />

diffraction using the peaks in the following 2 theta regions:<br />

♦<br />

♦<br />

32.4 °– 32.8 ° (overlap of alite and belite peaks)<br />

32.9 ° - 34.1 ° (principal aluminate and ferrite peaks)<br />

In addition the distribution of the size of the alite crystals (> 60 µm, 30 – 60 µm, 10 – 30 µm<br />

and < 10 µm) was determined by Scanning Electron Microscopy (SEM) .<br />

The materials and process parameters included in the evaluation were as follows:<br />

Materials Parameters:<br />

♦ raw meal fineness<br />

♦<br />

♦<br />

♦<br />

♦<br />

degree of raw meal pre-calcination<br />

coal parameters(x4)<br />

organic liquid waste burnt<br />

chemical composition parameters (x4)<br />

Process Parameters:<br />

♦<br />

♦<br />

♦<br />

♦<br />

♦<br />

temperatures (x6)<br />

pressures (x8)<br />

environmental parameters (x6)<br />

feed of hot meal, liquid waste, oil, tyres, etc (x7)<br />

“others” (x6)<br />

To complete the evaluation Svinning (1996, 1998) determined both setting time and<br />

compressive strength development from 1 up to 28 days.<br />

Svinning’s studies were therefore very comprehensive in nature and a model was drawn up<br />

to predict cement properties at the Norwegian plants involved. The model involves X-ray<br />

diffraction profile at 32.9 and 34.1 ° 2-theta, mineralogy / microstructure of clinker and<br />

gypsum. A sufficiently good prediction of the clinker mineralogy / microstructure was possible<br />

by including only free CaO and XRD the profile. The explained variance of compressive<br />

strength at 1-day was 86 %. To obtain the full benefit of Svinning’s work, reference must be<br />

made to the original papers.<br />

Ghosh et al. (1998), studying two plant clinkers A and B, report higher early strengths at 3<br />

and 7 days for plant B material which is characterised by:<br />

♦<br />

♦<br />

♦<br />

better developed alite and belite crystals<br />

no twinning of belite<br />

no decomposition rims around alite<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 58 of 220


Cement<br />

Concrete<br />

♦<br />

better micro-homogeneity,<br />

when compared to clinker from plant A.<br />

3.7 Rate of Clinker Cooling<br />

Ono et al. (1968) investigated the influence of 4 different cooling regimes on various clinker<br />

types corresponding to the ASTM Types 1, 2 and 3. These were burnt at 1450 °C (Type 3 at<br />

1500 °C) for 30 min. and cooled at rates varying between fast cooling from 1450 °C to very<br />

slow cooling down to 800 °C. It was found that the best strength development for the Type 3<br />

clinker was obtained by cooling from the burning temperature down to 1200 ° in 15 min.<br />

Sylla (1975), aware of conflicting findings of previous authors, reported on trials in a cyclone<br />

and in a grate pre-heater kiln in which the effect of changing the clinker cooling rate, both<br />

within the kiln and after discharge from the kiln, was studied. The former was varied by<br />

changing the shape of the flame and position of the burner and the clinker temperature at the<br />

kiln outlet used as a measure of the rate of kiln-internal cooling. Clinker outlet temperatures<br />

of between 1410 and 1450 °C (alkali rich clinker in grate pre-heater kiln) and between 1320<br />

and 1450 °C (low alkali clinker in cyclone pre-heater kiln) were obtained. From this work<br />

Sylla found that:<br />

♦<br />

♦<br />

♦<br />

♦<br />

♦<br />

the rate of cooling at high temperatures in the kiln (> 1310 °C) had much more<br />

influence than the rate of external cooling<br />

for both clinkers (and kilns) the higher the clinker temperature at the kiln outlet the<br />

lower the strength development at 1 day<br />

this loss in strength with temperature was more pronounced for the alkali-rich clinker<br />

for the low alkali clinker a 6 N/mm 2 increase in 1-day strength was observed when the<br />

clinker outlet temperature was reduced from 1440°C to 1320°C<br />

the slower the cooling after the kiln also resulted in higher 1 day strengths for the alkalirich<br />

clinker but not for the low alkali clinker: Differences in the order of 6 N/mm 2 were also<br />

observed<br />

♦ the 28 day strength development increased for both clinkers with lowering kiln outlet<br />

clinker temperature. Again the alkali rich clinker showed the greatest change with<br />

temperature<br />

♦ the 28-day strength did not correlate with the rate of cooling outside of the kiln.<br />

Jepson (1976) on the basis of an evaluation of 14 industrial kilns fitted with grate and<br />

planetary coolers, showed that the type of cooler and hence rate of the cooling after the<br />

clinker leaves the kiln has no real influence on strength development. He also was of the<br />

opinion that the cooling rate within the kiln was the more critical factor.<br />

Chatterjee et al. (1980) reviewed the literature on the influence of cooling rate on cement<br />

properties. With respect to strength development he reported that in general rapid cooling<br />

produces lower earlier strengths and higher 28-day strengths.<br />

Ichikawa et al. (1995) considered that the cooling rate from 1300 °C can be estimated by x-<br />

ray diffraction. This is deduced from the angle (2-theta) and the width at half peak height of<br />

the ferrite 020 peak at around 12 °. The slower the rate of cooling the higher the 2 theta<br />

angle and the narrower the peak width at half height. For commercial clinkers the cooling<br />

rate, as is to be expected depends on the clinker nodule size.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 59 of 220


Cement<br />

Concrete<br />

Misteli (1999) considers that the degree of crystallisation of the liquid phase, and thus the<br />

temperature at which the clinker leaves the kiln is a critical factor. In clinker quenched from<br />

the maximum burning temperature, the aluminate and ferrite are finely dispersed and so the<br />

aluminate grains are “protected” by ferrite to the largest possible extent against hydration.<br />

For slowly cooled clinker, removed from the kiln after the crystallisation of the liquid phase,<br />

and so achieving complete separation of the aluminate and ferrite, this shielding of the<br />

aluminate in reduced to a minimum.<br />

3.8 Reducing Conditions<br />

According to Sylla (1981) on the basis of laboratory produced clinkers, reducing conditions -<br />

whether local or caused by the kiln atmosphere - changes the composition of the ferrite, and<br />

increases the content of C 3 A and its reactivity. They also decrease the stability of C 3 S and<br />

cause a lighter coloured clinker to be formed. To achieve the best cement quality the clinker<br />

should leave the kiln at a temperature above 1250 °C.<br />

Long (1985) reports that reducing conditions promote the replacement of Ca 2+ by Fe 2+ and<br />

results in decreased hydraulic activity.<br />

4. Implementation of findings<br />

4.1 Introduction<br />

The preparation and communication of the present literature review is not to be considered<br />

as an end in itself. Here the final aim is the successful industrial implementation of the<br />

present findings to produce clinkers with the highest possible early strength development.<br />

As was indicated in Chap. 1, the factors influencing strength development can be classified<br />

as being “material” or “process” in nature. In general the literature cited up to about 1975 /<br />

1980 was more concerned with the influence of “material” factors, the work performed since<br />

then being focussed more on “process” aspects. This is most probably due to the fact that<br />

easy to use laboratory instrumentation was first widely available for chemical analysis (e.g.<br />

XRD), whereas equipment for X-ray diffraction, reflected light microscopy, scanning electron<br />

microscopy is less widely available and requires specialists to interpret the findings. Whereas<br />

all cement plants (old or modern) possess an XRF, only a few posses the other instruments<br />

mentioned, which are mostly to be found at research institutes. The fine tuning with respect<br />

to maximum strength development of the chemical composition will therefore be easier to<br />

accomplish than fine tuning of the process.<br />

4.2 Barriers to Implementation<br />

Most of the findings presently reported have been known for many decades and the question<br />

thus arises as to why they have not been generally implemented. The reason for this is that<br />

various factors have been (and still are) operative which resist the formation of any change<br />

whatsoever. These can be classified as being influenced by the following aspects.<br />

♦<br />

increased clinker production costs<br />

• use of external materials (e.g. fluorspar, high grade limestone, high alkali<br />

clay, etc.)<br />

• selective use of quarry (e.g. high grade limestone)<br />

• higher burning temperatures and higher fuel and refractory costs<br />

• higher quality control costs, etc.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 60 of 220


Cement<br />

Concrete<br />

♦<br />

♦<br />

♦<br />

Changes in Cement Properties<br />

• longer setting times<br />

• higher alkali content<br />

• higher C 3 A content and hence lower sulphate resistance<br />

• different colour<br />

• changed workability, etc.<br />

Clinker production aspects<br />

• reduction of kiln output<br />

• more difficult to burn mix<br />

• increased pre-heater blockages<br />

• insufficient raw meal homogeneity, etc<br />

Personnel Aspects<br />

• change to established routine<br />

• increased demand on production personnel , etc.<br />

4.3 Fine Tuning of Material Aspects<br />

On the basis of the literature reviewed in Chap. 2.1 and 2.3, it can be seen that many authors<br />

have reported on the influence of chemical composition, especially the main elements (SiO 2 ,<br />

Al 2 O 3 , Fe 2 O 3 and CaO) and the soluble volatile elements (K 2 O, Na 2 O, SO3 and Cl). Authors<br />

cited include Blaine, Alexander, Gebauer, Stürmer, Munoz and Ghosh.<br />

Basically these authors findings can be summarised by the following type of equation for<br />

which that of Gebauer will be chosen as representative for the group:<br />

CS 2-day (ISO) = -2.3 + 0.22 C 3 S + 0.40 C 3 A + 3.8 KNS (r = 0.839)<br />

CS 2-day (ISO) = -1.42 + 0.19 C 3 S 0.35 C 3 A + 8.5 K sol (r = 0.876)<br />

(KNS represents K 2 O + Na 2 O + SO 3 and K sol soluble K 2 O)<br />

As presented it is considered that the chemical composition is the major factor influencing<br />

compressive strength development.<br />

4.3.1 Principal Clinker Oxides<br />

This is essentially the easiest, most effective and proven way of improving clinker strength<br />

development characteristics and is possible using easy to control changes in chemical<br />

composition.<br />

On the basis of the just given model, reaching the highest early strength possible should be<br />

attempted by maximising the LSF and hence C 3 S content. (From recent experience this is<br />

possible up to a maximum of about 103). decreases the belite down to zero in the final<br />

instance. At present such extreme LSF clinkers are produced using fluorspar only at plants of<br />

the Apasco Group (since 1995) and at Cerro Blanco (since 1998).<br />

Various comments have been made on the influence of C 3 A, ranging from it being the most<br />

important factor (Alexander 1968) to it being of no influence on early strengths (Stürmer<br />

1990) to it having a negative influence (Munoz 1995). The influence of C 3 A will therefore<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 61 of 220


Cement<br />

Concrete<br />

have to be decided on a plant to plant basis, and is probably influenced by the rate of clinker<br />

cooling (Chap 3.7).<br />

If it is required to increase strengths by increasing C 3 A the addition of bauxite to the raw mix<br />

is an already established method, although the increase will necessitate a better control of<br />

setting times.<br />

4.3.2 Soluble, Volatile Elements<br />

Referring to the results of Gebauer (1980) and Osbaek (1979) as basis, early strengths could<br />

be improved by increasing the content of the alkalis and SO 3 in clinker, with the soluble<br />

alkalis present as alkali sulphate being more important than alkalis and SO 3 incorporated into<br />

the clinker minerals’ lattices.<br />

Increasing the SO 3 content of the clinker is quite easy in that lower cost, high-S fuel can be<br />

used, e.g. petcoke.) Increasing the quantity of the alkalis is very difficult and is mostly<br />

impossible, as no cheap, easily available “alkali corrective” has been shown to exist. Suitable<br />

materials such as K 2 SO 4 are both too expensive and normally unavailable.<br />

4.4 Fine Tuning of Process Aspects<br />

If the findings of Gebauer (1980) are correct, only modest, secondary increases in early<br />

strength are to be expected by optimising process parameters. These will also require a<br />

higher, more sophisticated input in quality control in comparison to the easy measurement of<br />

chemical composition. Microscopy and x-ray diffraction become almost essential.<br />

Fortunately (Gebauer 1980), the type of kiln does not exert a direct influence on strength<br />

development characteristics.<br />

4.4.1 Raw Meal Preparation<br />

Moir (1997) considered that plants with a raw mix having a higher 90 µm residue LSF than in<br />

the bulk mix, resulted in cements with very high strength levels, so called “over achievers”.<br />

This was considered to be due to raw mixes having a lower LSF in the 90 µm residue being<br />

higher in SiO 2 and forming belite nests and less than expected alite. This raw mix property<br />

might be able by avoiding the use of quartz sand as a corrective and could be easily checked<br />

by sieving followed by chemical analysis.<br />

4.4.2 Control of Alite Size<br />

Controlling the size of the alite crystals almost certainly requires the use of clinker<br />

microscopy, or more complicated still x-ray diffraction.<br />

Gebauer (1980), however, was able to show that by tempering samples of an industrially<br />

produced clinker in a laboratory kiln to increase the size of the alite, the 2-day strengths<br />

dropped by 1.2 N/mm 2 whilst the 28-day strength increased by 3.1 N/mm 2 . Platzer (1997)<br />

demonstrated on an industrial scale that the alite size has to be matched with the quality of<br />

clinker desired.<br />

On the basis of the above it would definitely seem to be possible to increase compressive<br />

strength development by ensuring that the average size of the alite crystals are as small as<br />

possible. Achieving this would necessitate optimising the temperature profile of the kiln to<br />

achieve a shorter effective burning zone.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 62 of 220


Cement<br />

Concrete<br />

4.4.3 Control of Polymorphic Form of Alite<br />

According to Ono (1968) and Aldous (1983) higher strength development is achieved when<br />

alite is present in the rhombohedral form. That this form is present would require the use of<br />

x-ray diffraction as a quality control tool.<br />

It is known the that addition of small quantities of F can lead to the stabilisation of the<br />

rhombohedral modification, although this alone did not lead to improved strength<br />

development at the Cerro Blanco plant in Ecuador.<br />

4.4.4 Clinker Cooling<br />

On the basis of the work of Ono (1968), Sylla, and Jepson (1976), it would appear that<br />

clinker cooling at temperatures down to 1200 °C within the kiln, are much more important<br />

than cooling outside the kiln. An optimum cooling may increase compressive strengths by up<br />

to 6 N/mm 2 .<br />

The optimum cooling regime, which appears to involve a slow cooling within the kiln, can be<br />

obtained by inserting the burner pipe as far into the kiln as possible so as to obtain<br />

separation of the ferrite and aluminate phases. Methods for determining the rate of cooling<br />

would necessitate clinker microscopy, or alternatively x-ray diffraction to determine the width<br />

of the ferrite peak at around 12 ° 2 theta.<br />

4.4.5 Reducing Conditions<br />

Prevention of reducing conditions will ensure that no lowering in compressive strengths takes<br />

place due to this factor. Inspection of the colour of the inner core of large clinker grains or the<br />

use of the Magottaux test will demonstrate freedom from reducing conditions.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 63 of 220


Cement<br />

Concrete<br />

5. Literature Cited<br />

Aldous. R.T.H., “ The Hydraulic Behaviour of Rhombohedral Alite”, Cement and Concrete<br />

Research, Vol. 13, pp 89 – 96, 1983.<br />

Alexander, K.M., Taplin, J.H, and Wardlaw, J., “ Correlation of Strength and Hydration with<br />

Composition of Cement”, Proceedings of the 5 th International Congress on the Chemistry of<br />

Cement, Tokyo, 1968, Supplementary Paper III-72, Volume 3, pp 152 – 166.<br />

Bhatty, J.I., “Role of minor elements in cement manufacture and use”, PCA Research and<br />

Development Bulletin RD 109T, 1995.<br />

Blaine, R.L., Arni, H.T., Foster, B.E., et al., “Interrelations between cement and concrete<br />

properties, Part 1 – “Materials, techniques, water, requirements and trace elements”, United<br />

States Department of Commerce, National Bureau of Standards, Building Science Series 2,<br />

August 20, 1965, pp 1 – 35.<br />

Blaine, R.L., Arni, H.T., DeFore, M.R., “Interrelations between cement and concrete<br />

properties, Part 3 -”Compressive strengths of Portland cement test mortars and steam-cured<br />

mortars”, United States Department of Commerce, National Bureau of Standards, Building<br />

Science Series 8, April, 1968, pp 1 – 98.<br />

Bürki, P., Private communication, 1993.<br />

Chatterjee, T.K. and Ghosh, S. N., “ The effect of cooling rate on cement properties- a<br />

review”, World Cement Technology, June 1980, pp 252 – 257.<br />

Fletcher, J.G. and Glasser, F.P., “Phase Relations in the System CaO - B 2 O 3 - SiO 2 ”, J.<br />

Materials Science,, vol. 28, 1993, pp 2677 – 2686.<br />

Gebauer, J. et al., “Influence of Clinker Quality upon the Properties of Cement and Concrete,<br />

HMC Report MA 80/2712/E, 1980.<br />

Ghosh, S.P. and Mohan, K., “Interrelationship among Lime Content of Clinker, its<br />

Microstructure, Fineness of OPC Grinding and Strength Development of Hydrated Cement at<br />

Different Ages”, Proceedings of the 10 th International Congress on the Chemistry of Cement,<br />

Gothenburg, 1998, Volume 2, ii018, 4 pp.<br />

Gutt, W., “Manufacture of Portland cement from phosphatic raw materials”, Proceedings of<br />

the 5 th International Congress on the Chemistry of Cement, Tokyo, 1968, Volume 1, pp 93 -<br />

105.<br />

Heinemeyer, H. (private communication referring to Hardegsen plant)<br />

Ichikawa, M., and Komukai, Y., “ Estimation of clinker cooling rate by XRD pattern<br />

ddecomposition of ferrite phase and its correlation with strength development”, JCA<br />

Proceedings of Cement and Concrete, No 49, 1995, pp 8 – 13.<br />

Imlach, J.A., “The influence of heating conditions on the production of Fluuorine-containing<br />

Portland cement clinker”, Cement Technology, Vol. 5, No. 4,July/August 1974, pp 403 – 406.<br />

Imlach, J.A., “ Evaluation of the suitability of boron as a mineraliser in the production of<br />

Portland cement clinker”, HMC Report MA 94/3320/E.<br />

Imlach, J. A. “Sources and reductions of heavy metal emissions”, HMC Report MA<br />

97/13205/E, 1997.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 64 of 220


Cement<br />

Concrete<br />

Jepsen, O.L., “Zementfestigkeit und ihre Bezeihung zur Kühlgeschwindigkeit und Kühlertyp”,<br />

Zement –Kalk-Gips, Nr. 2, 1976, pp 62 - 64<br />

Knöfel, D., “Beeinflussung einiger Eigenschaften des Portlandzementklinkers und des<br />

Portlandzementes durch TiO 2 ”, Zement-kalk-Gips, vol. 4, 1977, pp 191 – 196.<br />

Knöfel, D., “Effect of manganese on the properties of Portland cement clinker and Portland<br />

cement”, Zement-kalk-Gips, vol. 7, 1983, pp 402 – 408.<br />

Long, G. R., Phil. Transactions. Royal Society London, 1985, A310, p 43.<br />

Misteli, B., CTS/MT, private communication.<br />

Moir, G. and Glasser, F.P, “Mineralisers, Modifiers and Activators in the Clinkering Processl”,<br />

Proceedings of the 9 th International Congress on the Chemistry of Cement, New Dehli, 1992,<br />

Volume 1, pp 125 - 153.<br />

Moir, G., “Influence of Raw Mix Heterogeneity on Ease of Combination and Clinker Strength<br />

Potential”, Proceedings of the 10 th International Congress on the Chemistry of Cement,<br />

Gothenburg, 1997, Volume 1, 1i041, 8 pp.<br />

Munoz,M., et al., “ Influence of the mineralogical composition, specific surface area and<br />

strains crystallite size of alite on the compressive mechanical strength of portland mortars. II.<br />

Clinkers of high tricalcium aluminate content”, Cement and Concrete Research, 1995, Vol.<br />

25,No 5, pp 1103 – 1110.<br />

Ono, Y., Kawamura, S., and Soda, Y., “Microscopic observations of alite and belite and the<br />

hydraulic strength of cement”, Proceedings of the 5 th International Congress on the<br />

Chemistry of Cement, Tokyo, 1997, Vol. 1, supplementary paper 1-79.<br />

Osbaeck, B., “Der Einfluss von Alkalien auf die Festigkeitseigenschsften von<br />

Portlandzement”, Zement-Kalk-Gips, Nr. 2, 1979, pp 72 – 77.<br />

Paul, M. and Glasser, F.P., “Mineralised Clinkers”, Unpublished HMC study.<br />

Platzer, E., “Untersuchungen zur Klinkerreaktivität im Werk Retznei der Lafarge Perlmooser<br />

AG, Gesteinshüttenkolloquium 1998, 23 Oktober 1998, Institut für Gesteinshüttenkunde,<br />

Montanuniversität, Leoben.<br />

Schmidt, O., “Einfluss von ZnO und PbO auf die Eigenschaften von Portlandklinker und<br />

Portlandzement”, Dissertation, Fakultät für Bergbau, Technische Universität Clausthal,<br />

Januar 1980.<br />

Spierings, G.A.C.M. and Stein, H.N., “The influence of Na 2 O on the hydration of C 3 A”,<br />

Cement and Concrete Research, Vol. 6, 1976, pp 265 – 272.<br />

Stephan,D., Maleki, H., Knöfek, D., Eber, B. and Härdtl, R., “ Influence of Cr, Ni and Zn on<br />

the properties of pure clinker phases, Part 1 – C3S, Cement and Concrete Research, vol. 29,<br />

1999, pp 545 – 552.<br />

Strung, J., Knöfel, D. and Dreizler, I., “Einfluss der Alkalien und des Schwefels aud die<br />

Zementeigenschaften”, Zement-Kalk-Gips, Nr. 3, 1985, pp 150 – 158.<br />

Stürmer S.et al., “ Einfluss des Trikalziumaluminat (C 3 A-) – Gehaltes im Portlandzement auf<br />

Brennverhalten und Phasenbestandder Klinker sowie Festigkeitsentwicklung der Zemente<br />

Svinning, K.and Bremseth, S.K., “ The Influence of Microstructure in Clinker and Cement on<br />

Setting Time and Strength Development until 28 Days”, Proceedings of the 18 th International<br />

Conference on Cement Microscopy, Houston, 1996, pp 514 – 533.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 65 of 220


Cement<br />

Concrete<br />

Svinning, K.and Bremseth, S.K., “X-ray Diffraction Studies on variations in Microstructure in<br />

Portland Cement ClinkerCorrelated to Variations in Production Conditions in the Kiln”,<br />

Proceedings of the 18 th International Conference on Cement Microscopy, Houston, 1996, pp<br />

382 – 403.<br />

Svinning, K.and Justnes, S., “ Application of Partial Squares Regression Analysis in<br />

Examination of Correlations between Production Conditions, Microstructures in Clinker and<br />

Cement and Cement Properties”, Proceedings of the 10 th International Congress on the<br />

Chemistry of Cement, Gothenburg, 1998, Volume 1, 1i038, 8 pp.<br />

Sylla, H.M., “Einfluss der Klinkerkühlung auf Erstarren und Festigkeit von Zement”, Zement-<br />

Kalk-Gips, H 9, Sept. 1975, pp 357 – 362.<br />

Sylla, H.M., “Einfluss reduzierenden Brennens auf die Eigenschaften des Zementklinkers,<br />

Zement-Kalk-Gips, H 12, December 1981, pp 618 – 630.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 66 of 220


Cement<br />

Concrete<br />

Mineral Components<br />

1. INTRODUCTION TO MINERAL COMPONENTS............................................................... 68<br />

1.1 Definition of Mineral Components ............................................................................ 68<br />

1.2 Types of Mineral Components ................................................................................. 68<br />

1.3 Industrial MIC - Production Worldwide.................................................................... 69<br />

1.4 Opportunities of MIC ................................................................................................ 72<br />

1.5 Threats of MIC ......................................................................................................... 72<br />

1.6 Mineral Components - A Core Business of Holcim ................................................. 73<br />

2. MINERAL COMPONENTS STRATEGY........................................................................... 74<br />

2.1 Introduction .............................................................................................................. 74<br />

2.2 Concept of a MIC Strategy....................................................................................... 74<br />

3. MINERAL COMPONENTS - OVERVIEW......................................................................... 76<br />

3.1 Mineral Components in Cementitious Applications.................................................. 76<br />

3.2 Chemical Composition of Mineral Components ....................................................... 77<br />

3.3 Performance of Mineral Components ...................................................................... 78<br />

3.4 Standards for Mineral Components ......................................................................... 82<br />

3.5 Testing Methods....................................................................................................... 83<br />

4. MAIN MINERAL COMPONENTS FOR PRODUCTION OF BLENDED<br />

CEMENTS.........................................................................................................................83<br />

4.1 Blast-Furnace Slag and other Slag Types ............................................................... 83<br />

4.2 Fly Ash ..................................................................................................................... 99<br />

4.3 Pozzolans............................................................................................................... 110<br />

5. REFERENCES.................................................................................................................. 120<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 67 of 220


Cement<br />

Concrete<br />

1. Introduction to Mineral Components<br />

1.1 Definition of Mineral Components<br />

Mineral Components are materials with binding properties, which are not derived from the<br />

clinker manufacturing process.<br />

Mineral Components can be used<br />

- as alternative material for clinker (clinker substitute) or<br />

- for the manufacturing of concrete (cement substitute).<br />

Other terms used for Mineral Components are mineral additions, mineral additives, mineral<br />

admixtures, cement extenders, or for chemically inert material, fillers.<br />

1.2 Types of Mineral Components<br />

According to their reactivity, Mineral Components are subdivided in three groups:<br />

♦ latent hydraulic material (e.g. granulated blast-furnace slag)<br />

♦ pozzolanic component (e.g. natural pozzolana, fly ash, silica fume, burnt clay)<br />

♦ filler (usually limestone).<br />

Latent hydraulic materials have a natural hydraulic potential (i.e. they develop silicate<br />

hydrates by hydration reactions). To accelerate the reaction with water they require an<br />

activator such as:<br />

♦ Lime<br />

♦ Portland cement clinker<br />

♦ Gypsum<br />

♦ Chemical activator<br />

Blast-furnace slag is the most typical latent hydraulic material.<br />

Pozzolanic Materials can give rise to the formation of silicate hydrates only in presence of<br />

lime (calcium hydroxide, Ca(OH)2), as a consequence of its reaction with the amorphous<br />

(glassy) silica contained in the pozzolana. The hydration reaction of Portland cement clinker<br />

with water will provide the necessary lime to make the pozzolanic reaction happen.<br />

Natural pozzolanas, generally of volcanic origin, and low-calcium fly ashes are the most<br />

commonly used materials belonging to this group.<br />

Inert additions (fillers) possess no hydraulic or pozzolanic activity, but contribute to cement<br />

properties other than strength development.<br />

Limestone filler is the most commonly used, particularly in the production of fillerized<br />

cements, with up to 35% clinker replacement, and in the manufacture of masonry cements<br />

(up to 75% limestone).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 68 of 220


Cement<br />

Concrete<br />

Mineral Components can also be grouped according to their origin into natural and artificial<br />

materials. Artificial Mineral Components are either derived from industrial production as byproducts,<br />

or can be obtained by thermal activation of natural materials or by-products.<br />

Table 1 shows an overview of the classification of Mineral Components.<br />

Table 1 Classification of Mineral Components according to reaction mechanism<br />

Example: SLAGS<br />

Example: NATURAL POZZOLANS, FLY ASH<br />

artificial natural artificial<br />

* granulated blast-furnace<br />

slag (gbfs)<br />

* volcanic ashes * activated natural materials<br />

(burnt clays, shales, and volcanic<br />

rocks, metakaolin, Kalsin, etc.)<br />

* all other rapidly cooled slags * tuffs * industrial by-products (fly ash,<br />

silica fume, rice husk ash, etc.)<br />

* opaline cherts and<br />

shales<br />

* diatomaceous earth<br />

* activated industrial by-products<br />

(calcined red mud, waste, pulverized<br />

clay bricks or tiles, etc.)<br />

* rhyolites, etc.<br />

The properties of the two different groups of Mineral Components are described in detail in<br />

chapter 4. Owing to their wide application and strategic importance fly ashes are discussed<br />

separately from pozzolans in section 4.2.<br />

1.3 Industrial MIC - Production Worldwide<br />

The global volume of industrial MIC produced each year amounts to approximately 830<br />

million metric tons, compared to a total cement output of 1’500 million metric tons.<br />

Figure 1 shows the annually produced volumes of the main industrial MIC. Fly ash makes up<br />

for more than half of the total quantity, followed by blast-furnace slag. Almost two thirds of<br />

this slag is granulated and is hence suitable for cement addition. A yearly volume of 6 million<br />

metric tons of granulated blast-furnace slag is traded by ship and can potentially travel to any<br />

destination port around the globe.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 69 of 220


Cement<br />

Concrete<br />

Granulated<br />

Blast Furnace Slag<br />

Gbfs<br />

124<br />

Seaborne Slag Trading (6)<br />

60<br />

Steel Slag<br />

65<br />

Fly Ash<br />

490<br />

120<br />

Bottom Ash<br />

Global volumes: MIC cement = 614 mio t/y<br />

MIC aggregates = 245 mio t/y<br />

Cement production = 1.500 mio t/y<br />

Holcim = 142 mio t/y<br />

Figure 1 Global production of slag and fly ash (2002)<br />

Figures 2 and 3 give an overview on the geographic distribution of the produced blastfurnace<br />

slag and fly ash volumes.<br />

Rest of Eastern<br />

Europe / CIS<br />

Bfs<br />

volume<br />

Gbfs<br />

(mio t/y)<br />

9 China<br />

17<br />

Western Europe<br />

USA/Canada<br />

Ukraine<br />

16<br />

5<br />

Japan<br />

5<br />

6 4 6 7<br />

36<br />

19<br />

6<br />

5<br />

Rest of Latin 0.3<br />

7<br />

17<br />

America<br />

2<br />

Turkey, Middle East<br />

0.5<br />

Africa<br />

and Indian<br />

5<br />

Subcontinent<br />

8<br />

1.5<br />

0.7<br />

6<br />

Rest of Asia,<br />

Brazil<br />

Australia<br />

Air cooled bfs<br />

(mio t/y)<br />

Total bfs: 184<br />

Total gbfs: 124<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 70 of 220


Cement<br />

Concrete<br />

Figure 2 Global production of gbfs and ungranulated bfs (2002, mio t/y)<br />

Fly ash<br />

production<br />

Used<br />

Directly usable<br />

Beneficiable<br />

Non-usable<br />

Global prod. : 490<br />

Utilization : 167<br />

Directly usable : 173<br />

Beneficiable : 90<br />

Non usable : 60<br />

USA/Canada<br />

73 / 33%<br />

10 3225<br />

6<br />

CIS<br />

Eastern Europe/Tk<br />

56 / 37%<br />

62 / 16%<br />

8<br />

12<br />

21 China<br />

Western Europe 14 10<br />

15<br />

100 / 60%<br />

39 / 33%<br />

16<br />

NE-Asia<br />

22<br />

6 4<br />

23 / 67%<br />

17<br />

18<br />

11<br />

30 60<br />

13<br />

5 16<br />

India<br />

80 / 10%<br />

Africa<br />

10<br />

8<br />

1 / 10%<br />

SE Asia<br />

20<br />

42<br />

7 / 47%<br />

Latin America<br />

6 / 16%<br />

3 3<br />

2<br />

1<br />

South Africa<br />

Australia<br />

3<br />

30 / 6%<br />

6<br />

12 / 16%<br />

2<br />

2<br />

6<br />

3<br />

8<br />

14<br />

1<br />

Total fly ash<br />

Current utilization<br />

production<br />

rate %<br />

Figure 3 Global production of fly ash (2000, mio t/y)<br />

Today, industrial MIC have become valued commodity products. The following trends<br />

contribute to this development:<br />

• Changes in environmental legislation / constraints<br />

Dumping of industrial by-products is being restricted or not allowed any more. At the<br />

same time, disposal costs increase.<br />

• Reduced margins on core products<br />

Increased competition in the steel sector and deregulation of power industry put the<br />

margins under pressure. Steelworks and power plants find new value potential in the<br />

commercialisation of their by-products.<br />

• High technical competence, sophisticated processes<br />

The improvements in their core processes also result in a more consistent quality of the<br />

by-products. Consequently, larger volumes of good quality slags and fly ashes can be<br />

expected.<br />

• Lower trading barriers<br />

• Lower transportation costs<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 71 of 220


Cement<br />

Concrete<br />

1.4 Opportunities of MIC<br />

The integration of MIC offers considerable opportunities, ranging from environmental benefits<br />

to product performance, marketing aspects and financial profits.<br />

The main opportunities with MIC are:<br />

• Reduction of CO 2 : As a clinker substitute, MIC is the most important means to<br />

considerably reduce the output of CO 2<br />

• Reduction of operational costs<br />

• Reduction of net operating assets (NOA)<br />

• Extension of lifetimes (e.g. quarries, machines) and/or delay in investments<br />

• Differentiation of products<br />

• opportunity to produce tailored products<br />

• improvement of cement properties for selected applications<br />

• Entrance into new geographic markets<br />

Although the opportunities mentioned above rarely appear simultaneously, in most<br />

companies the significance of even one single opportunity justifies the integration of MIC.<br />

With CO 2 taxes soon to be imposed in many countries, the use of MIC as most effective<br />

means to reduce CO 2 emissions will gain further importance.<br />

1.5 Threats of MIC<br />

Figure 4 visualizes the integration of MIC in the cement process. The MIC deriving from<br />

pozzolan quarries, coal power stations or integrated steel mills can be introduced at different<br />

stages of the value chain:<br />

• As raw meal component for clinker production<br />

• As cement constituent in the cement mill (‘clinker substitute’)<br />

• Pure as concrete addition (‘cement substitute’)<br />

From a cement manufacturer’s point of view the MIC flows should be restricted to the first<br />

two channels, shown with green arrows in figure 4. The red arrows indicate the direct supply<br />

of MIC to the concrete producers. In this case the MIC bypass the cement process and<br />

partially substitute cement in the concrete mix.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 72 of 220


Cement<br />

Concrete<br />

Figure 4: Integrated Cement and MIC Business Structure<br />

This can have the following impacts:<br />

• Partial replacement of cement in concrete lowers our cement sales<br />

• MIC as traded commodities lower the entrance barriers for new competitors in the<br />

cement markets<br />

• New non-cement competitors try to control the MIC sources and the direct MIC<br />

channels to concrete producers and<br />

• will substantially reduce the cement price level by introducing low priced MIC in order to<br />

gain important market shares<br />

• Lower priced MIC blended cements compete against OPC (in spite of the same or better<br />

performance)<br />

• MIC enables competitors to react much faster in the markets<br />

1.6 Mineral Components - A Core Business of Holcim<br />

In 1997, MIC was declared part of Holcim’s core business. Thus, it must be integrated in the<br />

business plan of a Group company.<br />

Since then, a number of projects related to MIC strategies have been realized by the Group<br />

companies, ranging from sourcing strategies (e.g. Holdercim Brasil) to blast-furnace slag<br />

granulators (e.g. Alsen and Holnam) and grinding stations dedicated to composite cements<br />

(e.g. Juan Minetti).<br />

The HMC-Corporate Product Development division assists Group companies in developing<br />

their strategies, co-ordinates and acts as information turntable.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 73 of 220


Cement<br />

Concrete<br />

2. Mineral Components Strategy<br />

2.1 Introduction<br />

The development of an appropriate MIC strategy is an important part of any cement business<br />

plan. It should aim at revealing and preventing potential threats at an early stage and making<br />

best use of the opportunities offered by MIC.<br />

2.2 Concept of a MIC Strategy<br />

A comprehensive MIC strategy covers aspects along the entire value chain of cement<br />

business and comprises all MIC materials (blast-furnace slag, fly ashes, and other<br />

pozzolans).<br />

Following the value chain, the focus lies on five major topics:<br />

1. Sourcing<br />

The MIC sources are usually not owned by the cement industry, but are inter-linked with<br />

other production processes. Analysing and understanding the MIC supply market,<br />

keeping track of the developments of upstream industries and knowing the relevant<br />

sources are crucial factors in order to evaluate the accessible MIC volumes in the market.<br />

Furthermore, the quality of MIC material and its consistency over time need to be<br />

analysed carefully to decide on its suitability for cement production.<br />

2. Logistics<br />

An extensive use of MIC is linked with an increase in logistic flows. The optimization of<br />

these flows by using synergies becomes most important in order to reduce expenses.<br />

The MIC strategy process should comprise inbound, interplant and outbound logistics.<br />

3. Manufacturing<br />

The impact of MIC integration on cement production can achieve various dimensions:<br />

extension of quarry lifetime, modification of clinker properties, increase of cement<br />

capacity, elevation of production complexity. Therefore the MIC strategy process should<br />

take into account all activities from raw material exploitation to clinker and cement<br />

production and estimate the required investments. Environmental questions, like the CO 2<br />

issue, should also be addressed.<br />

4. Product<br />

Mineral Components are an adequate tool for product differentiation. Understanding the<br />

application segments and the resulting concrete property requirements within the market<br />

is essential for the development of customer-focussed products and services.<br />

5. Market<br />

Analysing and understanding the composite cement market, scrutinising the product mix<br />

and identifying competitors and their strategies are prerequisites for the assessment of<br />

the MIC potential in a market and the basis for the development of an accurate product<br />

mix and a suitable marketing concept.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 74 of 220


Cement<br />

Concrete<br />

HMC-CPD has developed a tool that visualizes the process of strategy shaping: the MIC<br />

Pyramid.<br />

Figure 5: The MIC Pyramid<br />

The functional structure of the MIC Pyramid follows the value chain from sourcing to<br />

marketing. The base of the pyramid is composed of 24 boxes containing the key topics for a<br />

thorough analysis of the MIC situation in a company.<br />

In the vertical dimension the pyramid is organized according to the concept of MIC strategy<br />

development. The analysis is followed by a synthesis, where opportunities, threats and key<br />

issues of the functional sections are recognized. In a next step MIC options are developed,<br />

which are interrelated to sourcing and marketing scenarios. The evaluation of the options<br />

regarding impacts and risks results in the definition of the most appropriate MIC strategy for<br />

the Group company.<br />

Going from top down, the MIC Pyramid visualizes the implementation process of the MIC<br />

strategy from the approved strategy to the individual action plan at operational level.<br />

MIC strategies can have different objectives, ranging from reduction of clinker factor to<br />

expansion of production capacities, product differentiation or prevention of market entrance<br />

by a competitor. In any case, the MIC strategy is part of the cement business plan and as<br />

such impacts on all issues covered by the latter, such as procurement, logistical set-up,<br />

investments, product-mix, financial projections etc.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 75 of 220


Cement<br />

Concrete<br />

3. Mineral Components - Overview<br />

3.1 Mineral Components in Cementitious Applications<br />

Slags have been known since ancient civilizations. It has been illustrated that ancient<br />

Egyptians produced glass slags, rich in vesicles (air bubbles) and containing crystals of the<br />

minerals melilite and tridimite, by burning wheat straw. This contains opal, an amorphous<br />

form of silica, that combines during burning with alkali to form glassy slag.<br />

More than 2000 years ago, the Romans discovered the activity of the volcanic ash from<br />

Pozzuoli when it was mixed with lime. This was used in many structures that lasted through<br />

the centuries and are still existing today, such as aquaducts, bridges, historical buildings like<br />

the Pantheon Figure 6.<br />

Vitruve, architect and engineer in the 1st century AC was the first who tried to explain the<br />

reactions that transform some rocks into cementitious materials.<br />

Figure 6<br />

Pantheon, Rome (2 nd century).<br />

More recently, the production of blended cements has been practiced for many decades in<br />

many countries. The latent hydraulic reactivity of vitreous blast-furnace slag was discovered<br />

by Emil Langen in 1862. The first use of slag in cement dates back to 1865, when in<br />

Germany a slag-limecement was commercially produced. In 1883 slag was used as raw<br />

material for Portland cement manufacture. The first production of a Portland blast-furnace<br />

type of cement by grinding together Portland cement and granulated slag occurred in<br />

Germany in 1892.<br />

Fly ash cements has been used since the 1930ies in cementitious applications. Cements<br />

containing natural pozzolanas have been firmly installed in the Italian and Greek markets for<br />

decades, and are used in practically all applications involving concrete construction.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 76 of 220


Cement<br />

Concrete<br />

Table 2 gives a rough overview on the present use of composite cements worldwide:<br />

Table 2<br />

Production of Composite Cements worldwide.<br />

Addition<br />

Situation<br />

<br />

Slag Ash Pozzolan Limestone<br />

established EC countries Europe Italy Europe<br />

Latin America Oceania Turkey Morocco<br />

East. Europe South Africa Latin America<br />

Japan<br />

Greece<br />

South Africa<br />

increasing Australia Canada Middle East East. Europe<br />

Brazil Far East Far East New Zealand<br />

Canada<br />

India<br />

3.2 Chemical Composition of Mineral Components<br />

Mineral Components exhibit in many respects very different properties. However, they have<br />

one thing in common: They consist essentially of the same major chemical elements as PC<br />

clinker: SiO2 - Al2O3 - CaO.<br />

The respective proportions regarding the CaO - SiO2 - Al2O3 diagram are shown in Table 3<br />

and represented in Figure 7.<br />

Table 3 Ranges of the chemical composition of Mineral<br />

Components (on the basis of dried and calcined materials)<br />

Elements<br />

Blast-furnace<br />

Slag<br />

Natural and/or<br />

Artificial Pozzolan<br />

Class F Fly Ash<br />

SiO 2 30 - 40 50 - 75 40 – 65<br />

Al 2 O 3 (TiO 2 ) 8 - 25 15 - 25 15 - 40<br />

Fe 2 O 3 0.5 - 1.5 3 - 10 3 - 17<br />

CaO 35 - 45 1 - 15 1 - 10<br />

MgO 1 - 18 < 5 - 3<br />

SO 3 1.5 - 6 0.1 - 1.5 0.3 - 3<br />

K 2 O 0 - 0.5 0.5 - 4 2 - 3<br />

Na 2 O 0 - 0.5 0.5 - 4 0.4 - 2<br />

MnO 0.2 - 2 - < 0.3<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 77 of 220


Cement<br />

Concrete<br />

Figure 7<br />

CaO-SiO-AlO-System<br />

3.3 Performance of Mineral Components<br />

Mineral Components have to be activated for adequate hardening.<br />

Slag has a high silica and calcium oxide content. If activated by OH - or SO 4 -- ions, it displays<br />

hydraulic activity (cementitious properties). In practice, the activation is realized by addition<br />

of lime, Portland cement clinker (OH - ) or gypsum (SO 4 -- ).<br />

Silica, and alumina at a lower degree, are the main components of pozzolan, whereas the<br />

calcium oxide is low. Pozzolan supplies the silica (or alumina) that reacts with added lime or<br />

with the lime (Portlandite) produced by hydration of Portland cement clinker.<br />

When analyzing the calcium hydroxide content in mortar as a function of time, it can be<br />

observed that the quantity of CH decreases significantly for a a blended cement compared<br />

to pure Portland cement Figure 8.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 78 of 220


Cement<br />

Concrete<br />

Figure 8<br />

Calcium Hydroxide Content of Pozzolanic Cement and Portland Cement<br />

Mortar<br />

The main hydration products of Mineral Components when mixed with OPC and water are<br />

essentially the same as those formed during the hydration of Portland cement, namely nearly<br />

amorphous calcium-silicate hydrates (CSH) of different stoichiometric composition, the socalled<br />

AFt and AFm-phases (Aluminate-Ferrite-tri-sulphate, hydroxide etc.; C 3 (A,F) . 3(CaSO 4 ,<br />

Ca(OH) 2 ) . aq and Aluminate-Ferrite-mono-sulphate, hydroxide, chloride, carbonate, etc.;<br />

C 3 (A,F) . (CaSO 4 , Ca(OH) 2 , CaCO 3 , CaCl 2 ) . aq, and Ca(OH) 2 .<br />

The amount of CSH gel of cements containing slag, fly ash or natural pozzolans is<br />

substantially higher than that of Portland cement, because less calcium hydroxide is<br />

generated due to the lower free lime content of blended cements compared with OPC.<br />

That leads to<br />

- a shift in pore size to smaller pores Figure 9 and 10 and<br />

- the blocking of corresponding pores ("pore blocking effect", Figure 10)<br />

In concrete the porous interfacial zone around the aggregates is considerably smaller when<br />

Mineral Components are used.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 79 of 220


Cement<br />

Concrete<br />

Figure 9<br />

Pore size distribution in hydrated binder vs. content of ground<br />

granulated blast-furnace slag.<br />

Figure 10<br />

Pore size reduction and pore blocking.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 80 of 220


Cement<br />

Concrete<br />

The denser microstructure of concrete containing Mineral Components significantly<br />

- decreases the permeability of the concrete and thus<br />

- increases the resistance to chemically aggressive environments<br />

- increases the strength at long-term (because of the lower hydration rate of Mineral<br />

Components)<br />

The CSH-phases of blended cements are able to incorporate alkalis. The "trapped" alkalis<br />

are no more available to form alkali silicates, which owing to volume expansion can damage<br />

the concrete. Concretes containing Mineral Components are less sensitive to alkali silicate<br />

reaction.<br />

Mineral Components also improve the properties of fresh concrete. They<br />

- improve workability<br />

- decrease the water demand.<br />

The rate of hydration of Mineral Components is lower compared with Portland cement.<br />

Therefore MIC additions result in lower early strength. When the concrete contains MIC<br />

proper curing is very important.<br />

The lower rate of hydration of the MIC is coupled with a lower heat development, which<br />

makes Mineral Components additions beneficial in mass concrete construction.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 81 of 220


Cement<br />

Concrete<br />

3.4 Standards for Mineral Components<br />

The increased use of Mineral Components in cement and concrete brought to the<br />

development of specifications describing the characteristics of the different materials and the<br />

testing methods applied to evaluate suitability of use in cementitious blends.<br />

These features are usually enclosed in the main cement standards (e.g. EN 197, ASTM C-<br />

595 and 1157). When the mineral component can be directly used in the mix design of<br />

concrete, specific standards are usually drafted.<br />

A comprehensive description of national and international specifications for blended cements<br />

and their mineral constituents is given in the chapter on cement standards.<br />

As an example, some of them are listed below.<br />

♦ ASTM C 618 “Coal Fly Ash and Raw or Calcined Natural Pozzolan for Use as a Mineral<br />

Admixture in Portland-Cement Concrete”<br />

♦ EN 450 “Fly Ash for Concrete”<br />

♦ ASTM C 989 “Specification for Ground Granulated Blast-furnace Slag for Use in<br />

Concrete and Mortar”<br />

♦ ASTM C 1240 “Specification for Silica Fume for Use in Hydraulic Cement Concrete and<br />

Mortar”<br />

♦ EN 13263 "Silica Fume for Concrete".<br />

Often the question is raised whether it is more advisable to use Mineral Components in the<br />

manufacture of blended cements rather than in the preparation of concrete mixes. Basically,<br />

a concrete prepared with a blended cement and another prepared by addition of the mineral<br />

component directly into the concrete mixer, prove to be equivalent in performance. However,<br />

the use of blended cements offer the folowing advantages: a well established quality control<br />

on materials, the availability of large storage facilities that smoothen quality fluctuations, and<br />

the possibility to adjust the cement composition and production parameters to the desired<br />

final binder performance.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 82 of 220


Cement<br />

Concrete<br />

3.5 Testing Methods<br />

The most currently used testing methods for the determination of the quality of an addition<br />

are:<br />

♦ Chemical analysis including heavy metals and insoluble residue<br />

♦ Mineralogical analysis by means of XRD and microscopy<br />

♦ Calorimetric methods such as DTA, DSC or TGA<br />

♦ Particle shape (microscopy)<br />

♦ Density<br />

♦ Fineness and particle size distribution (Blaine, sieve analysis, laser granulometry)<br />

♦ Grindability (Holcim grindability test)<br />

♦ Activity index and pozzolanic activity determination:<br />

• Keil Index (hydraulic)<br />

• Strength Activity Index (SAI) according to ASTM C 311<br />

• CEN pozzolanic activity test<br />

♦ Test on standard laboratory cement<br />

♦ Specific accelerated mortar test, etc.<br />

A combination of 2 or 3 of the above methods is generally necessary for the current quality<br />

control of acceptance of an addition.<br />

4. Main Mineral Components for Production of Blended<br />

Cements<br />

4.1 Blast-Furnace Slag and other Slag Types<br />

4.1.1 General<br />

Slags are by-products of different metallurgical industries such as the ferrous industry or nonferrous<br />

industries like copper, lead, or nickel production.<br />

When metal is extracted from the ore or refined by a further process, slag of various qualities<br />

are obtained. The slags basically form by fusion of silica, alumina and alkali earth<br />

compounds from ore and flux and combustion residues of the fuel. These reactions take<br />

place at temperatures between 1300 and 1600°C.<br />

The chemical and mineralogical composition of the slag varies considerably according to the<br />

raw materials and the refining process Table 4.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 83 of 220


Cement<br />

Concrete<br />

Table 4<br />

Chemical Composition of Metallurgical Slags - Examples.<br />

Lead<br />

Zinc<br />

U.K.<br />

Nickel<br />

Canada<br />

Non-ferrous Slags<br />

Copper<br />

S. Africa<br />

Phosphorus<br />

Furnace<br />

USA<br />

Ferrous Slags<br />

Steel BOF<br />

Slag<br />

Germany<br />

Iron Blast<br />

Furnace<br />

Europe<br />

SiO 2 18 29 34 41 13 34<br />

CaO 20 4 9 44 47 41<br />

MgO 1 2 4 1 1 7<br />

Al 2 O 3 6 1 6 9 1 13<br />

FeO x + MnO x 38 53 41 1 31 1<br />

CaO/SiO 2 1.1 0.1 0.3 1.1 3.6 1.2<br />

Owing to the favourable chemical composition (no heavy metals, no unstable mineral<br />

phases), blast-furnace slag is the major slag type used in cementitious binders.<br />

Therefore the following chapters are related to blast-furnace slags.<br />

The amount of other slag types than blast-furnace slag for the production of Portland blended<br />

cements is limited. EN standards, for example, define a maximum content of 5%.<br />

Steel furnace slags are widely used as aggregates in road construction as pavement material<br />

(fill, subbase, base) and in asphalt and thin bituminous surfacings.<br />

4.1.2 Definition of Blast-Furnace Slag<br />

Blast-furnace slag is the non-metallic product, consisting essentially of silicates and<br />

alumosilicates of calcium and other bases that is developed in a molten condition<br />

simultaneously with iron in a blast furnace.<br />

4.1.3 Production of Blast-Furnace Slag<br />

4.1.3.1 ORIGIN AND CLASSIFICATION<br />

Blast-furnace slag (bfs) is a by-product in the manufacture of pig iron in the blastfurnace<br />

Figure 11. The raw materials (iron ore, flux, steel slag) and coke introduced at the top of the<br />

furnace move down and become heated from below. Due to air injection near the bottom of<br />

the furnace, the oxidation of the coke supplies enough energy to melt the burden. The oxidic<br />

iron ore is reduced to pig-iron.<br />

The blast-furnace slag forms by fusion of the gangue material of the iron ore (mainly silica<br />

and alumina compounds) with the remaining calcium and magnesium oxides of the thermally<br />

decomposed carbonatic flux (limestone, dolomite) and combustion residues of the coke.<br />

These reactions take place at temperatures between 1300 and 1600°C.<br />

The slag floats on the top of the liquid iron and is drawn off at regular intervals. In<br />

dependence of the ore composition and efficiency and size of the blastfurnace, the amount of<br />

slag per ton of pig iron varies. For an economic iron production the ratio of slag to pig iron<br />

weight should be as low as possible. In modern blast-furnaces processing high quality ore<br />

the slag/pig iron ratio is ~250-300 kg/t.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 84 of 220


Cement<br />

Concrete<br />

Figure 11<br />

Schematic Section of a Blast Furnace.<br />

The main constituents of bfs are lime-silica-alumina and magnesia compounds. The<br />

chemical composition of blastfurnace slags depends on the burden and the fuel used in the<br />

blastfurnace. Thus, the composition of slags from different sources varies within certain<br />

limits. Slags originating from the same blastfurnace exhibit a relatively constant composition,<br />

because to ensure a constant quality of the iron, significant changes of the feed have to be<br />

avoided.<br />

Other properties of the slag such as mineralogical composition, density and porosity are<br />

mainly influenced by the cooling procedure. With decreasing cooling rate the crystalline<br />

proportion of the slag rises. Basically two types of slag can be distinguished – air-cooled and<br />

quenched slag (Table 5).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 85 of 220


Cement<br />

Concrete<br />

Air-cooled slag (Figure 12) is obtained by slow cooling of the molten slag in open pits. It is<br />

mainly crystalline and hardly exhibits any cementitious properties. Crushed and sieved, the<br />

material is used as aggregate and filler in road construction as pavement material (fill, base,<br />

subbase) and also in concrete manufacture.<br />

The specific gravity of crystalline slag lies between 2.38 and 2.76, the bulk density between<br />

1150 and 1440 kg/m 3 .<br />

In the past 50 years the percentage of air-cooled slag has gradually decreased in favour of<br />

quenched slag, which owing to its reactivity is widely used in cementitious applications.<br />

Quenched slag, a predominantly vitreous product, is obtained by rapid cooling (quenching)<br />

of bfs. Depending on the quenching procedure, slag wool, expanded, pelletised, and<br />

granulated bfs can be distinguished. Due to their glass content, quenched slags exhibit<br />

latent hydraulic properties, i.e. upon activation, they show cementitious behaviour. Unground<br />

quenched slags can be used as aggregate or even lightweight aggregate. In ground form<br />

quenched slags are used in the production of blended cements as clinker substitute and in<br />

concrete manufacture as cement substitute. Granulated slag (Figure 12) is the most<br />

important slag type for cementitious applications. Therefore in the following, emphasis is laid<br />

on production and properties of granulated bfs. The production of pelletized slag (Figure<br />

12) is mentioned shortly as well.<br />

Table 5 gives an overview on the different types of blast-furnace slag.<br />

Figure 12<br />

Different Types of Blast-furnace Slag.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 86 of 220


Cement<br />

Concrete<br />

Table 5 Different Types of blast-furnace slag.<br />

Blast-Furnace Slag<br />

quenched<br />

air-cooled<br />

granulated pelletized foamed/ expanded<br />

mainly vitreous partially crystalline partially crystalline crystalline<br />

dense or porous porous porous dense<br />

mainly ground in<br />

cement or concrete<br />

lightweight aggregate<br />

or ground in<br />

cement/concrete<br />

lightweight aggregate<br />

road or concrete<br />

construction<br />

4.1.3.2 QUENCHING TECHNIQUES<br />

4.1.3.2.1 Granulation<br />

Wet granulation (with water) is the most common and well-established quenching method.<br />

Granulated blast-furnace slag (gbfs) is produced by breaking-up and quenching the stream<br />

of molten slag by water jets to particles with a maximum size around 5 mm. Figure 13 shows<br />

as an example a schematic illustration of the Salzgitter granulator, Germany, which is<br />

operated by Alsen.<br />

The formation of gbfs with a glass content >95% requires rapid quenching of the slag melt to<br />

temperatures below the crystallisation temperature of approx. 840°C. Therefore ~5-10 tons<br />

of water are necessary, depending on the granulation equipment.<br />

The bulk density of the granulate usually is between 0.9 and 1.1 kg/l.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 87 of 220


Cement<br />

Concrete<br />

Figure 13<br />

Schematic illustration of a granulator (Source: Alsen).<br />

From the granulation tank the mixture of granulate and water is transported to the dewatering<br />

system, which differs depending on the equipment supplier. The basic dewatering systems<br />

are<br />

- Pit method (granulate settles under gravity in a pond and is clammed out with a crane<br />

or excavator)<br />

- De-watering silos (granulate settles and water flows off the top through a steel mesh<br />

in the sides and discharge cone of the silo)<br />

- Gravel bed filters (similar to silos, only with a bed of gravel at the<br />

base of the cells acting as a filter. After drainage, the granulate is removed by an<br />

overhead crane)<br />

- INBA drum (Paul Wurth) (most widely used unit ; slurry flows under gravity to a<br />

distributor box located along the axis of a horizontal cylindrical mesh filter. The drum<br />

contains lifters that elevate the dewatering slag until it falls on a conveyor belt for<br />

drum exit)<br />

- Dewatering wheel (AJO) (Single or double wheels rotate slowly in a secondary slurry<br />

reservoir, wheels reclaim the gbfs via pockets with steel mesh to allow the water to<br />

flow back into the reservoir as the pocket rises above water level, gbfs falls out of the<br />

pocket onto a belt conveyor)<br />

- Screw conveyors (RASA) (screw conveyor rotates inclined at an angle in the reservoir<br />

elevating the granulate, while the water runs back into the reservoir under gravity,<br />

gbfs falls off the screw on a belt conveyor)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 88 of 220


Cement<br />

Concrete<br />

After dewatering the granulate is stored in silos or stockpiles, where it continues to drain. The<br />

residual moisture content is normally around 10 wt.-%. Depending on the grinding technique<br />

(compound or separate grinding) and equipment (mill type) drying of the gbfs before grinding<br />

can be necessary.<br />

4.1.3.2.2 Pelletizing<br />

The slag pelletiser was patented by National Slag Ltd., Hamilton, Canada in the late 1960’s.<br />

At present only few units are still working due to environmental and availability problems.<br />

Pelletised slag is formed by the expansion of molten bfs under water sprays. The pyroplastic<br />

material passes a spinning drum with fins, which break the slag and flinge it in the air, where<br />

due to surface tension, the pellets form (Figure 14).<br />

Figure 14<br />

Schematic illustration of a slag pelletizer<br />

Advantages are the lower water consumption compared with granulation (about 0.5 tons of<br />

water per ton of slag), and the low residual moisture of the pellets of ~5.5 wt.-%. Also gas<br />

emissions are lower owing to the entrapment of sulfide gases in the pores of the pellets.<br />

However, the emissions occur at ground level and cause more occupational health problems.<br />

Additionally, pelletisers create noise levels in excess of most industrial standards. Also the<br />

formation of considerable amounts of slag wool may create problems.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 89 of 220


Cement<br />

Concrete<br />

The glass content of pelletised bfs can be controlled by fractionation, because it increases<br />

with decreasing grain size. For the use as light weight aggregate, the demand for low water<br />

sorption and good thermal conductivity requires a quite crystalline structure. For this<br />

application bigger grains are used. The smaller fraction is used for cementitious applications,<br />

where higher glass contents are favourable.<br />

The grindability and also the hydraulic activity of slag pellets do not differ significantly from<br />

that of granulated slag.<br />

4.1.4 Processing of Blast-Furnace Slag<br />

4.1.4.1 GRANULATED SLAG<br />

The granulate is mostly stockpiled. During the storage dewatering continues. The granulate<br />

can be used in the unground form as sand substitute, e.g. in road construction.<br />

For the production of binders, the granulate has to be ground. This can be done either<br />

- separately or<br />

- together with the clinker.<br />

When gbfs and clinker are interground it is normally difficult to achieve the necessary slag<br />

fineness for optimum strangth development without overgrinding the clinker, which is easier<br />

to grind. Therefore separate grinding is recommended. Separate grinding offers<br />

furthermore a higher flexibility regarding the design of different cements with defined clinker<br />

and slag fineness according to the market demand.<br />

Depending on the grinding system and the moisture of the granulate drying of the slag prior<br />

to grinding can be necessary.<br />

Major factors influencing the grinding energy consumption of a gbfs are<br />

- bulk density<br />

- chemical composition, especially TiO 2 content<br />

- glass content<br />

- Degree of hydration / loss on ignition<br />

- Grinding system (see Chapter 12).<br />

Grinding aids, e.g. diethylenglycol, offer the possibility to decrease the grinding energy<br />

consumption in the range of 5-10%.<br />

Basically, the grinding energy consumption of gbfs increases exponetially with<br />

fineness (Figure 15). Therefore it is beneficial to process slags of high quality, which do not<br />

require too high finenesses for a satisfactory performance in cement or concrete.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 90 of 220


Cement<br />

Concrete<br />

Figure 15<br />

Grinding energy consumption of different slags (HMC Laboratory trials<br />

with ball mill).<br />

The ground slag is stored in silos. It can be added directly into the concrete or mixed with<br />

a cement to produce a blended cement. This is done in multiple chamber silos with batch or<br />

continuous mixing units.<br />

4.1.4.2 AIR-COOLED SLAG<br />

After cooling of the surface of the slag in the open pits it is sprayed with water to accelerate<br />

cooling and enhance crack formation by thermal shock. After entire cooling the slag can be<br />

digged out more easily. It is then transported to a crushing and sieving plant. The different<br />

grain size fractions are used a s aggregates in road construction and concrete.<br />

4.1.5 Properties of Granulated Blast-Furnace Slag and Influence on Processing<br />

and Performance<br />

4.1.5.1 CHEMICAL COMPOSITION<br />

The main constituents of bfs are lime-silica-alumina and magnesia compounds. In<br />

dependence on the composition of the burden and the fuel used, the chemical composition of<br />

slags from different sources varies within certain limits. Typical chemical compositions of bfs<br />

are listed in Table 6.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 91 of 220


Cement<br />

Concrete<br />

Table 6 Chemical composition of bfs worldwide (Locher, 2000).<br />

Western<br />

Europe<br />

Russia /<br />

Ukraine<br />

USA,<br />

Canada<br />

Rep. South<br />

Africa India Japan Australia<br />

SiO 2 30 … 39 30 … 40 33 … 42 30. … 36 27 … 39 31 … 40 33 … 38<br />

Al 2 O 3 9 … 18 5 … 17 6 … 16 9 … 16 17 … 33 13 … 17 15 … 19<br />

CaO 33 … 48 30 … 50 36 … 47 30 … 40 30 … 40 38 … 45 39 … 44<br />

MgO 2 … 13 2 … 14 1 … 16 8 … 21 0 … 17 2 … 8 1 … 4<br />

FeO 0,1 … 1 0,2 … 0,9 1,3 … 4,5 - < 0,5 < 0.7 -<br />

MnO 0,2 … 3 0,2 … 1,4 - - < 1 0,3 … 1,5 -<br />

Na 2 O 0,2 … 1,2 - - 0,2 … 0,9 - - < 0,2<br />

K 2 O 0,4 … 1,3 - - 0,5 … 1,4 - - < 0,5<br />

SO 3 0 … 0,2 - - - - - -<br />

S 2- 0,5 … 1,8 - - 1,0 … 1,6 < 1 - 0,6 … 0,8<br />

The chemical composition has a very strong influence on the reactivity of a slag, because it<br />

controls the solubility and hence the reactivity of the slag during hydration. In general, the<br />

hydraulic activity of a slag increases with increasing contents of CaO, (MgO), Na 2 O, and<br />

Al 2 O 3 and with decreasing contents of SiO 2 , FeO, TiO 2 , MnO, and MnS.<br />

For the prediction of slag reactivity, numerous compositional moduli were developed, for<br />

example<br />

♦ I - Basicity Index 1.0 ≤ CaO/SiO 2 ≤ 1.4<br />

♦ II - according to DIN 1164<br />

♦ III - EN Standard<br />

♦ III - Keil ("F-value")<br />

CaO + MgO + Al O<br />

SiO<br />

CaO + MgO<br />

≥ 1<br />

SiO<br />

2<br />

2<br />

2 3<br />

≥ 1<br />

(CaO+<br />

0.5MgO+<br />

Al2O3<br />

+ 2.25S<br />

F =<br />

(SiO + 1.29Mn)<br />

15 . ≤F ≤ 2.<br />

0 good hydraulic properties 10 . ≤ F ≤ 15 . poor hydralic properties<br />

2<br />

2−<br />

)<br />

♦ IV - Langavant Index i = 20 + CaO + Al 2 O 3 + ½MgO - SiO 2<br />

i ≤ 12<br />

poor hydraulic properties 12 ≤ i ≤ 16 good hydralic properties<br />

i>16 Very good hydraulic properties (but difficult to granulate)<br />

SiO<br />

♦ VI - Silica-alumina Ratio 18 . ≤<br />

2 ≤19<br />

.<br />

Al O<br />

2 3<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 92 of 220


Cement<br />

Concrete<br />

These moduli are convenient tools for rapid quality control of slag originating from the same<br />

furnace, but their applicability for a general prediction of strength development of blended<br />

cements is questionable.<br />

Strength testing remains the most reliable method to access slag reactivity.<br />

As a guideline, for optimum reactivity in cementitious applications the gbfs should meet the<br />

following chemical criteria (composition of granulate, without impurities):<br />

Desirable<br />

Optimum<br />

CaO/SiO 2 >1.0 1.3<br />

Al 2 O 3 >10% 13%<br />

MgO 5-10%<br />

TiO 2<br />

as low as possible ≤0.5%<br />

The CaO/SiO 2 ratio influences the solubility of the slag glass and therefore is a decisive<br />

parameter for slag reactivity. To a limited degree low CaO/SiO 2 ratios can be compensated<br />

by elevated Al 2 O 3 or MgO contents.<br />

Elevated Al 2 O 3 contents in the slag increase the strength development of composite<br />

cements, especially at early age owing to enhanced ettringite formation.<br />

TiO 2 in bfs is very undesired, because it strongly reduces reactivity and increases grinding<br />

energy consumption of the granulate. TiO 2 is added to the blast furnace in order to increase<br />

the service life of the refractory.<br />

The composition of the slag is adjusted by the iron producer in order to optimise the quality of<br />

the iron and to maximize the efficiency of the blast furnace. Subsequent changes in slag<br />

composition mostly are not advantageous for the cement producer. An adjustment of the slag<br />

composition according to the demands for cementitious applications is possible only to a<br />

limited degree.<br />

Because the composition of the bfs can change over longer time periods the cement<br />

producer must ensure to obtain granulate of a constant quality by specification of relevant<br />

slag properties in the supply contract he signs with the steel mill.<br />

4.1.5.2 MINERALOGICAL COMPOSITION AND GLASS CONTENT<br />

The reactivity of a gbfs cannot be evaluated based on the chemical composition only. Also<br />

the mineralogical composition, especially the glass content, influence reactivity.<br />

According to EN standard gbfs for cement should exhibit a glass content of at least 2/3 by<br />

mass standard. With state-of-the-art granulation equipment in most cases glass contents ><br />

90% are obtained, which from a reactivity point of view are satisfactory.<br />

Perfect vitrification is no criterion for optimum reactivity. Often an increase of the<br />

compressive strength of slag cements is observed when the slag contained small amounts of<br />

finely distributed crystals, which may act as hydration nuclei.<br />

The presence of crystal phases in gbfs reduces grinding energy consumption, however at<br />

higher amounts (>10-15%) at cost of reactivity.<br />

The main crystalline constituents in gbfs are calcium silicates and silico-aluminates (Table<br />

7). Melilite (solid solutions in the composition range from Gehlenite (C 2 AS) to Akermanite<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 93 of 220


Cement<br />

Concrete<br />

(C 2 MS 2 )) and Merwinite are the most common mineral phases in gbfs. Slag granulate can<br />

also contain some metallic iron. The iron should be removed (magnetic separator) prior to<br />

grinding and within the grinding circuit to avoid a concentration of iron in the mill, which can<br />

damage the mill or cause higher grinding time.<br />

Table 7 Some crystalline constituents of blast-furnace slag.<br />

Mineral name mineralogical formula simplified formula<br />

Silicates<br />

Melilite<br />

Gehlenite 2 CaO . Al 2 O . 3 SiO 2 C 2 AS<br />

Akermanite 2CaO . MgO . 2SiO 2 C 2 MS 2<br />

Dicalcium Silicate* 2 CaO . SiO 2 C 2 S<br />

Rankinite 3 CaO . 2 SiO 2 C 3 S 2<br />

Wollastonite CaO . SiO 2 CS<br />

Forsterite 2 MgO . SiO 2 M 2 S<br />

Enstatite** MgO . SiO 2 MS<br />

Merwinite* 3 CaO . MgO . 2 SiO 2 C 3 MS 2<br />

Monticellite* CaO . MgO . SiO 2 CMS<br />

Anorthite** CaO . Al 2 O . 3 2 SiO 2 CAS 2<br />

Diopside** CaO . MgO . 2 SiO 2 CMS 2<br />

Oxides<br />

Spinels MgO . Al 2 O 3 MA<br />

Sulfurous compounds CaS, MnS, FeS<br />

Others Carbonaceous, sulphurous, nitrogenous compounds, alkali thiosulphates etc<br />

* only in basic slags; ** only in acidic slags<br />

In slags with high lime contents dicalcium silicate (C 2 S) can form. It occurs in three different<br />

crystalline forms. Below 675°C β-C 2 S is stable, which at atmospheric temperatures<br />

transforms into γ -C 2 S, accompanied by a volume increase of ~10%. This can lead to a<br />

spontaneous disintegration of the slag matrix. Therefore, in some standard specifications the<br />

CaO/SiO 2 ratio is limited (e.g. BS 6699 and 1047 CaO/SiO 2 max = 1.4).<br />

The mineralogical composition and glass content can be measured by X-ray diffraction<br />

(XRD) or by microscopy, the latter being often coupled with an image analysis system.<br />

By differential thermal analysis, vitreous slag shows irreversible exothermic peaks at about<br />

860°C, when heated between 800 and 900°C. These peaks are mainly due to the heat of<br />

devitrification of the glassy part of the slag. Some authors tried to find a correlation between<br />

the areas under the peaks and the slag content.<br />

4.1.5.3 AGE / LOSS ON IGNITION (LOI)<br />

As gbfs is latent hydraulic, it hydrates also in absence of an activator, for instance during<br />

being stockpiled. The loss on ignition is a measure for the degree of hydration of a slag<br />

glass.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 94 of 220


Cement<br />

Concrete<br />

The grinding energy required to grind a gbfs with high LOI (e.g. in the range of 2 wt.%) to the<br />

same fineness like a fresh gbfs (LOI normally below 0.5%) is considerably lower. The reason<br />

are the soft hydrates at the surface of the granules, which are very easy to grind and<br />

excessively contribute to the fineness measurement ("false Blaine"). The comparably lower<br />

fineness of the reactive slag glass causes a lower strength development compared with fresh<br />

ggbfs of same fineness. The lower reactivity can be compensated by higher fineness of the<br />

hydrated granulate.<br />

4.1.5.4 BULK DENSITY<br />

The bulk density of granulated slag influences<br />

- its moisture content<br />

- grinding energy consumption.<br />

A slag with low bulk density is easier to grind, but exhibits elevated moisture contents, which<br />

can create transportation problems, can reduce the stockpile life of the granulate and may<br />

require drying of the granulate prior to grinding. Granulate of high bulk density has low<br />

moisture contents, but requires a higher grinding energy and in some cases may also be less<br />

reactive. An optimum bulk density range would be 0.9 - 1.1 kg/l.<br />

Factors influencing bulk density are slag temperature, slag composition, temperature of the<br />

granulation water, and the granulation process.<br />

4.1.5.5 FINENESS AND GRAIN SIZE DISTRIBUTION<br />

As for most cementitious materials, the hydraulic activity of ggbfs increases with its surface<br />

area and thus, with the fineness. The effect is especially strong for the late strength<br />

development.<br />

The influence of the fineness on strength development can increase with the CaO/SiO 2 ratio<br />

of the slag.<br />

Finer grinding of gbfs is a common countermeasure for compensation of poor slag reactivity.<br />

The limiting factors for slag fineness are performance indicators such as shrinking and<br />

economical considerations (grinding costs).<br />

The experiences regarding the influence of the grain size distribution of the ggbfs on strength<br />

development are conflicting. A shallow grain size distribution can reduce water demand and<br />

thus favour strength development. But also opposite experiences were described.<br />

4.1.6 Application of Ground Granulated Blast-Furnace Slag<br />

Concrete containing ggbfs offer numerous advantages compared with concrete with pure<br />

OPC such as higher long-term strength, durability, and resistance to chemical attack,<br />

sea water, sulphates, lower heat of hydration, low sensitivity towards alkali silicate<br />

reaction, bright color. Also the properties of the fresh concrete are improved (lower w/c,<br />

improved workability). The hydration rate and early strength development is lower.<br />

Therefore curing is of paramount importance.<br />

Ggbfs is widely used in concrete construction, especially where strength and durability<br />

aspects are important:<br />

- Industrial construction (waste gas desulphurization plants, coking plants, silos, sewage<br />

treatment plants, towers, chimneys etc.)<br />

- Transportation (bridges, tunnels, parking houses and areas, air ports, etc.)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 95 of 220


Cement<br />

Concrete<br />

- Hydraulic / marine construction<br />

- Mining, etc.<br />

In mass concrete construction the low heat of hydration is beneficial for rapid construction<br />

progress.<br />

Ggbfs is used in precast products, in prestressed concrete, shotcrete, ultra-high<br />

performance concrete, and self compacting concrete.<br />

Granulated and ground granulated bfs is also used in road construction (e.g. in road<br />

binders).<br />

The following figures illustrate some applications of ggbfs.<br />

Figure 16<br />

TV Tower and administrative building, Duesseldorf, Germany.<br />

TV-Tower/Admin. Building: 350/300 kg/m3 slag cement, w/c 0.50/0.57, aggregate 1860<br />

kg/m3, Fly ash 0/50 kg/m3 (Source: Weber et al., 1998).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 96 of 220


Cement<br />

Concrete<br />

Figure 17<br />

Flood gates in the Philips-Dam, Netherlands.<br />

Prestressed concrete, impermeable, 300 kg/m3 slag cement, w/c ~0.5, aggregate 1895<br />

kg/m3, superplasticizer, air entraining agent (Source: Weber et al., 1998).<br />

Figure 18<br />

Coking plant in Duisburg-Huckingen, Germany.<br />

Heat and chemical resistant concrete with air cooled bfs as aggregate (no strength losses at<br />

elevated temperature (400°C), 350-440 kg/m3 slag cement, w/c ~0.5, aggregate 1660-1829<br />

kg/m3, partially superplasticizer (Source: Weber et al., 1998).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 97 of 220


Cement<br />

Concrete<br />

Figure 19 Concrete sculpture Beethoven in Bonn, Germany (Weber et al., 1998).<br />

Low heat of hydration, satisfactory strength, bright color, 380 kg/m3 slag cement, w/c 0.45,<br />

aggregate max. size 8 mm, retarder (Source: Weber et al., 1998).<br />

Figure 20<br />

Mixing of road binder with soil.<br />

Example for composition of a road binder for subbase and base layers: 76% slag, 10%<br />

clinker, 5% CKD, 9% gypsum, fineness ~3500 cm2/g.<br />

More details on slag processing and reactivity are given in the HMC report MIC 00/5005/E<br />

"Properties, Grindability, and Reactivity of Granulated blast-furnace slag: A Literature<br />

Review".<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 98 of 220


Cement<br />

Concrete<br />

4.2 Fly Ash<br />

4.2.1 Definition of Fly Ash<br />

Fly ash is a fine powder of mainly spherical glassy particles derived from burning of<br />

pulverized coal, with pozzolanic properties consisting mainly of SiO2 and Al2O3. Fly ash is<br />

formed mainly by the unburnable part of the coal, which derives from inorganic material in<br />

the coal and bedrock material mined with the coal.<br />

4.2.2 Production of Fly Ash<br />

Fly ashes are obtained by electrostatic or mechanical precipitation of dust-like particles from<br />

the flue gases from furnaces fired with pulverized coal.<br />

The pulverized coal is injected into the furnace, ignites in the burner and burns while moving<br />

upwards through the boiler. Retention time of the coal particles in the boiler is around 10<br />

seconds, while the coal and ash particles see different temperature ranges as well as<br />

partially oxidizing and reducing atmospheres. Larger particles fall at the bottom of the boiler<br />

and are discharged wet or dry forming so called bottom ash.<br />

The non-burnable particles and partially unburned carbon is carried with the exhaust gases<br />

to the dust filters (electrostatic or baghouse filters), where they are separated from the flue<br />

gas and collected (Figures 21, 22). The electrostatic filters consist of a number of filter<br />

stages, usualy 3, where different fractions of the fly ashes are separated. The coarser<br />

particles are separated at the first stage, finer partcles at the following stages 2 and 3<br />

(Figure 23).<br />

The ratio between bottom ash and fly ash of the total ash content is around 20:80. Volume<br />

ratio of the fly ash at the filter stages is 80:16:4 from filter stage 1 to 3.<br />

Depending on the coal fired resp. the ash composition of the coal fly ashes are mainly<br />

characterized as siliceous or calcereous. The former have pozzolanic properties, and the<br />

latter have, in addition, hydraulic properties. There are no strict limits between these<br />

properties, as the types of coal are manifold.<br />

Fly ash is a mixture of mineral matters which have undergone thermal transformations and<br />

which contain still unburnt materials. The main components of the fly ash are complex glass<br />

phases consisting of Si, Al, Fe and Ca in various compositions. Quartz is the main cristaline<br />

phase.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 99 of 220


Cement<br />

Concrete<br />

<strong>Technical</strong>ly fly ashes can be categorized according to the burning temperature in the boiler.<br />

Three types of fly ash may be distinguished:<br />

♦ Type I: Fluidized Bed combustion<br />

• Burning temperature 850 to 1000°C<br />

• crystal structure remains<br />

• no melted material<br />

• pozzolanic properties due to sintered clays (high water demand)<br />

♦ Type II: Dry bottom furnace<br />

• Burning temperature 1100 to 1500°C<br />

• most minerals are melted (about 50-80%)<br />

• slow reaction with Ca(OH), only medium pozzolanic activity<br />

• grain size distribution similar to cement<br />

♦ Type III: Wet bottom or slag tap furnace<br />

• Burning temperature 1500 to 1700°C<br />

• all minerals are melted<br />

• grain size distribution similar to cement<br />

• rapid cooling of the fly ash induces the production of 60-80% of glassy particles<br />

• development of good pozzolanic properties<br />

It is generally known that the volume of fly ash produced in most countries exceeds largely<br />

the volume that can be utilized, but specific statistics are not available.<br />

Worldwide utilization rate of fly ash in all kinds of applications is estimated at 30%. Leaders<br />

in utilization rate and level of sophistication are Germany and France (100%), Japan (70%)<br />

and the US (30%). Free fly ash volumes are estimated at about 300 mio t not indicating<br />

suitable volumes regarding quality and geographical availability.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 100 of 220


Cement<br />

Concrete<br />

Figure 21<br />

Coal Fired Power Station - schematic.<br />

Figure 22<br />

Coal Fired Power Station.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 101 of 220


Cement<br />

Concrete<br />

4.2.3 Properties of Fly Ash and Influence on Processing and Performance<br />

The ash quality depends greatly on the type of coal and its bedrock material, type of boiler<br />

and its operation. The operating conditions as well as the coal quality strongly influence the<br />

variability of the ash with respect to residual carbon content, fineness, chemical composition<br />

and pozzolanic activity. Environmental measures to reduce SO2- and NOx-emissions can<br />

influence the fly ash properties negatively (increased carbon and SO3, ammonia smell).<br />

Key parameters influencing fly ash quality and quantity are:<br />

Coal<br />

- Coal type<br />

- Ash-content<br />

- Bed rock material<br />

Boiler<br />

- Boiler Type, Temperature<br />

Dedusting equipment<br />

- Electrostatic precipitator<br />

- Baghouse<br />

Environmental installation<br />

- High dust equipment for DeSOx and DeNOx<br />

4.2.3.1 CHEMICAL AND PHYSICAL PROPERTIES<br />

The main chemical components of ashes are SiO2, Al2O3, Fe2O3 and CaO. Substantial<br />

amounts of alkalis and sulfate may also be present. In the case of lignite or brown coal, the<br />

observed amount of free lime as well as of anhydrite is generally high (Table 8). Significant<br />

amounts of periclase can also be noticed in lignitic fly ash. The typical aspect of fly ash<br />

observed by microscopy is given in Phototable 1. The photos demonstrate the different<br />

morphology of the fly ash depending on the type of boiler. Figure 24 shows the grain size of<br />

the fly ashes from the different filter stages of the electrostatic precipitator obtained by laser<br />

granulometry.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 102 of 220


Cement<br />

Concrete<br />

Figure 23<br />

Granulometry of fly ash from different filter stages of the electrostatic<br />

precipitator (ESP)<br />

Figure 24<br />

Morphology of fly ashes from different boiler types (Scanning Electron<br />

Micrography)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 103 of 220


Cement<br />

Concrete<br />

4.2.3.2 MINERALOGICAL PROPERTIES<br />

The mineralogical composition of fly ashes depends on the amount of glassy material. The<br />

following minerals can be observed (Table 9), depending on the operating conditions and<br />

type of coal. Moreover, calcareous fly ash possesses hydraulic properties, due to a more or<br />

less greater amount of dicalcium silicate CS.<br />

Table 8<br />

Chemical and Physical Composition of Fly Ash<br />

Table 9<br />

Mineralogical Composition of Fly Ash<br />

Minerals<br />

Quantity % in weight<br />

Glass 70 – 98<br />

Quartz 1 – 15<br />

Mullite 1 – 10<br />

Magnetite 1 – 8<br />

Hematite 1 – 5<br />

Calcite < 2<br />

Plagioclase < 2<br />

Merwinite < 2<br />

Melilite < 2<br />

Anhydrite 1 – 8<br />

Free lime 1 – 8<br />

Periclase 1 – 6<br />

Wollastonite 1 – 5<br />

Dicalcium silicate 1 – 15<br />

4.2.3.3<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 104 of 220


Cement<br />

Concrete<br />

4.2.3.4 CARBON CONTENT<br />

The carbon content is an important criterion to assess quality of fly ash. Residual carbon is<br />

due to the non-burnt organic matter and lies generally in the range of:<br />

♦ 5 to 12% in anthracitic fly ash<br />

♦ 2 to 8% in bituminous fly ash<br />

♦ 1 to 5% in sub bituminous fly ash<br />

♦ 1 to 4% in lignite fly ash<br />

The different C-contents are due to differences in burnability and grindability of the coal<br />

types. The variations in C-content of each coal are mainly due to instabilities of the burning<br />

and grinding system, to variations in burning intensity or to changes in coal supply. The<br />

installation of Low-NOx burners to reduce NOx emissions can lead to an increase of the<br />

carbon content.<br />

Types of unburnt carbon are either not burnt coal or coke. The latter is very porous and has a<br />

very high specific surface.<br />

A high carbon content (> 5%) is considered detrimental for the quality of concrete since it<br />

might<br />

• impair the air-entrainment in freeze/thaw resistant concrete<br />

• have a negative effect on workability and strength development of concrete<br />

From the manufacturing point of view, it is worth while mentioning that most of the unburnt<br />

material is agglomerated and concentrated in the coarse fraction of fly ashes. Recent<br />

developments in the technology of fly ash benefication allow to improve quality by<br />

mechanical or electrical separation of the organic matter, making it possible to use highcarbon<br />

fly ashes (separated unburnt coal can be used as additional fuel for clinker burning).<br />

The main beneficiaiton technologies are air classifiers, triboelectric separation and burn-out<br />

systems.<br />

4.2.3.5 CLASSIFICATION<br />

Fly ash can be classified according to its chemical and physical properties as well as to its<br />

performance. The main standards used world-wide are EN 197, EN 450 and ASTM C 618.<br />

Fly ashesa re in the following classes of Table 10.<br />

Table 10<br />

Classification of Ashes<br />

Characteristics ASTM EN<br />

Fly ash from anthracite or bituminous coal Class F<br />

Siliceous fly ash<br />

Class V<br />

Fly ash from lignite or subbituminous coal Class C<br />

Calcareous fly ash<br />

Class W<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 105 of 220


Cement<br />

Concrete<br />

4.2.4 Standard Specifications<br />

The requirements do not significantly differ between ASTM and EN standards. Tables 11<br />

and 12 summarizes the ASTM 618 and EN 450 requirements for fly ash used in concrete,<br />

whereas those for EN 197 for cement additions are given in Table 13.<br />

Table 11<br />

EN 450 and ASTM 618 Chemical Requirements<br />

Table 12<br />

EN 450 and ASTM 618 Physical Requirements<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 106 of 220


Cement<br />

Concrete<br />

Table 13<br />

EN 197-1 Requirements<br />

4.2.5 Optimum Fly Ash Properties<br />

The key quality parameters of fly ashes are fineness and unburnt carbon and additionally<br />

free lime content for lignite fly ashes. The following parameters give an indication what a<br />

good quality fly ash would look like:<br />

♦ Hard coal fly ash in Europe (EN 450, EN 197-1 Type V)<br />

♦ Sub bituminous fly ash in US (ASTM 618 Class C)<br />

♦ High glass content (> 80 %)<br />

♦ Spherical particles (> 50 %)<br />

♦ High total CaO content (> 20%)<br />

♦ Low CaO free content (< 1 % or highly reactive)<br />

♦ Low LOI (Europe: < 5%, US: < 2%)<br />

♦ High fineness (> 2500 cm 2 /g Blaine/ residue on 45µm < 10 %)<br />

Fly ash properties can be improved either by upgrading (carbon reduction, classification),<br />

activation (e.g. alkali activators) or grinding.<br />

4.2.6 Applications<br />

Fly ashes can be used in most building materials and construction applications. It is used as<br />

binder replacement as well as aggregate or conditioning material in soils. In countries where<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 107 of 220


Cement<br />

Concrete<br />

fly ash is used it is a standard constituent of most of the concretes used and a part of the<br />

specifications.<br />

The following list gives an indication of the variety of applications and can also be seen an a<br />

value ranking of the fly ash by application:<br />

♦ Composite Cements<br />

♦ Concrete addition in all types of concrete<br />

♦ Autoclaved Aerated Concrete<br />

♦ Non-aerated concrete blocks<br />

♦ Sand lime brick<br />

♦ Bricks + ceramics<br />

♦ Lightweight aggregate<br />

♦ Cement raw material<br />

♦ Asphalt filler<br />

♦ Road construction<br />

♦ Grouting<br />

Actual developments are the use of fly ash in Self Compacting Concrete as well as in high<br />

performance concrete without silica fume.<br />

First applications date back in the 40ties, where fly ash was used to reduce hydration heat in<br />

mass concrete in dams (Figure 25). Over the last 60 years today sophistication of fly ash<br />

applications has increased considerably. Fly ash is used due to improved workability and<br />

durability in high performace concrete with up to 105 Mpa (Figure 26).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 108 of 220


Cement<br />

Concrete<br />

Figure 25<br />

First large scale application of fly ash in concrete<br />

Figure 26<br />

High performance application of fly ash in concrete<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 109 of 220


Cement<br />

Concrete<br />

4.3 Pozzolans<br />

Pozzolans are classified according to their genesis into the two types ‘natural’ (Table 14) and<br />

‘artificial’ (Table 15).<br />

Table 14<br />

Natural Pozzolans<br />

Genetic process<br />

Examples<br />

1 Explosive volcanic activity Italian Pozzolan earth (‘Pozzolana’),<br />

Greek Santorin earth, volcanic ashes and<br />

pumicites<br />

2 Explosive volcanic activity + zeolitic<br />

diagenesis<br />

Italian Neapolitan Yellow Tuff, German<br />

Rhenish Trass<br />

3 Meteorite impact German Bavarian Trass<br />

4 Building of skeletons of siliceous Diatomaceous earth, Moler earth<br />

organisms<br />

5 Ultrafine weathering of siliceous rocks Gaize, Tripel, opaline cherts<br />

Table 15<br />

Artificial Pozzolans<br />

Genetic process<br />

Examples<br />

6 Fast cooling of silicate melts in Non-ferrous slags, HSR-granulate<br />

metallurgical processes<br />

7 Flue gas cleaning of power stations Coal fly ash (class F)<br />

8 Oxidation and condensation of SiO-gas Silica fume<br />

in metallurgical Si- and Ti-processes<br />

9 Thermal activation of clay minerals and<br />

rocks<br />

10 Controlled burning of agricultural, SiO 2 -<br />

rich wastes<br />

Metakaolin, oil schist, crushed bricks,<br />

shale, phonolite<br />

Rice husk ash, crop ash<br />

All examples have in common that they consist of fine glassy particles with a CaO : SiO 2 ratio<br />

below 0.5. Obviously this is prerequisite to develop a pozzolanic reaction.<br />

In the following sections focus will be on pozzolans 1, 2, 8 and 9 of Tables 14 and 15. More<br />

detailed information about pozzolans is given in HMC report “Pozzolan Survey 2001” by P.<br />

Kruspan/CPD. Due to their importance Fly Ashes class F (pozzolan 7 in Table 15) are<br />

discussed separately in the preceding chapter 4.2.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 110 of 220


Cement<br />

Concrete<br />

4.3.1 Definition of Pozzolans<br />

Pozzolans are siliceous or siliceous and aluminous materials which in themselves possess<br />

little or no cementitious properties. When finely ground, they react in the presence of water<br />

at ambient temperatures with dissolved calcium hydroxide (Portlandite Ca(OH) 2 ) from<br />

lime or Portland cement clinker to form strength developing calcium silicate and calcium<br />

aluminate compounds.<br />

Brought into a simplified chemical notation the pozzolanic reaction can be written as follows:<br />

x Ca(OH) 2 + y SiO 2 (amorphous) + z H 2 O → x CaO ⋅ y SiO 2 ⋅ (x+z) H 2 O<br />

Portlandite + Pozzolan + Water → Calcium Silicate (Hydrate)<br />

It becomes obvious that only with a constant supply of dissolved Portlandite the pozzolanic<br />

reaction can occur. Furthermore a fast pozzolanic reaction is directly controlled by the<br />

dissolution rate of the pozzolan. Therefore a ‘good’ pozzolan has the following<br />

characteristics:<br />

• small particle size<br />

• high specific surface/porosity<br />

• glass with high amount of (earth)alkalies<br />

• zeolites<br />

4.3.2 Natural Pozzolans<br />

The vast majority of natural pozzolans used today are of volcanic origin (pozzolans 1 and 2<br />

in Table 14). Therefore the discussion in this section is restricted to this type.<br />

The pozzolans were named after the Italian town Pozzuoli near Naples (Italy) where deposits<br />

were formed by the explosive eruption of Campi Flegrei caldera 12’000 years ago. The<br />

Roman used these pozzolans together with burnt lime as building materials, hardening either<br />

in air or even under water. Many remainders from that time prove the good quality of these<br />

materials. When Portland cement was invented in the 19 th century, pozzolans fell into<br />

oblivion, but today regained their importance.<br />

4.3.2.1 GENESIS AND ITS INFLUENCE ON PROPERTIES AND<br />

PERFORMANCE OF NATURAL POZZOLANS<br />

The genesis of pozzolans of volcanic origin follows three steps (processes):<br />

1) melting in the magma chamber<br />

2) fragmentation and cooling of the melt during the volcanic eruption<br />

3) alteration, weathering and diagenesis of the deposited volcanic material<br />

For every single process a certain amount of parameters can be defined which control the<br />

course of the process. Furthermore every process is influenced strongly by the preceding<br />

one. And to make it more complicated, even a single volcanic event does not produce<br />

materials with constant quality. Pozzolan deposits are therefore heterogeneous by their very<br />

nature. For this reason the main challenge is to find high-quality pozzolans by means of<br />

specific geologic prospecting.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 111 of 220


Cement<br />

Concrete<br />

Helpful for successful prospecting in a volcanic area or in a single deposit is the distinction<br />

between volcanic materials with different geologic histories (geneses). In Table 16 and<br />

Figure 27 two opposing histories are listed, one leading to high-quality pozzolans (‘Plinian’<br />

genesis), the other one leading to low-quality pozzolans (‘Hawaiian’ genesis).<br />

The most important parameter for the melting, fragmentation and indirectly even the<br />

diagenesis process is the viscosity of the magma. In other words the whole genesis of<br />

pozzolans is mainly controlled by the viscosity. The viscosity on its part depends on<br />

temperature, chemical composition (SiO 2 , Fe 2 O 3 , FeO, CaO, Na 2 O, K 2 O), amount of volatiles<br />

(H 2 O, CO 2 , SO 2 ) and amount of crystallized phases.<br />

Table 16<br />

Pozzolans with different qualities depending on their genesis<br />

High quality pozzolans<br />

Low quality pozzolans<br />

Type of eruption Explosive (‘Plinian’) Non-explosive - effusive<br />

(‘Hawaiian’)<br />

Viscosity of magma High Low<br />

Degree of magma<br />

fragmentation<br />

High<br />

Low<br />

Cooling rate of magma High Low<br />

Eruption phenomena Pyroclastic flow, ash fall Fire fountain, lava flow<br />

Deposited material<br />

Examples<br />

Ash, tuff, pumice.<br />

Fine-grained, high glass<br />

content, high to medium<br />

porosity<br />

Mt. St. Helens, Vesuvius, Mt.<br />

Pinatubo<br />

Scoria, spatter, lava.<br />

Coarse-grained, low glass<br />

content or even fully<br />

crystallized, medium to low<br />

porosity<br />

Hawaii, Stromboli, Etna<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 112 of 220


Cement<br />

Concrete<br />

Figure 27<br />

Two opposing types of volcanic eruptions: Plinian (left) and Hawaiian<br />

(right); (Table 16). Not to scale.<br />

In Figure 28 the pozzolanicity of samples with different geologic histories was tested<br />

according to EN 196-5. This test determines the calcium binding capacity of the pozzolan in<br />

an OPC-pozzolan paste (cf. chapter 4.3.1). The lower the CaO-concentration in the solution<br />

(y-axis), the higher the pozzolanicity of the sample. It can be seen that high pozzolanicities<br />

are favoured by pyroclastic flows generated exclusively in explosive Plinian eruptions. If such<br />

deposits of pyroclastic flows are additionally solidified by zeolitisation during diagenesis the<br />

pozzolanicity is risen significantly.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 113 of 220


Cement<br />

Concrete<br />

Figure 28<br />

Pozzolanicity (EN 196-5) of samples generated by volcanoes with Plinian<br />

and Hawaiian type of eruption, respectively.<br />

From Figure 27 it becomes clear that an evaluation of natural pozzolans starts in the field<br />

with an assignment of specifically collected samples to their correct geologic history. This<br />

knowledge will provide the quarry manager with the ability to separate high-quality pozzolans<br />

from low-quality material in order to premix different qualities as early as possible in the<br />

production chain.<br />

Ignoring the geologic context (genesis) and considering only certain properties of an<br />

unspecified pozzolan sample (e.g. chemical composition) will give results hardly useable for<br />

an evaluation.<br />

Therefore self-standing prescriptions of some pozzolan characteristics like specification<br />

ASTM C 618 (Table 17) should be used only as a very rough guideline. It is impossible<br />

to predict the performance of a pozzolan in cement or concrete on the basis of such a<br />

prescriptive specification.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 114 of 220


Cement<br />

Concrete<br />

Table 17<br />

Specification ASTM C 618 (chemical requirements for natural pozzolans)<br />

Specifications<br />

Class N<br />

- Silicon dioxide (SiO 2 ) + aluminium oxide (Al 2 O 3 ) + iron oxide (Fe 2 O 3 ), min % 70.0<br />

- Sulfur trioxide (SO 3 ), max. % 4.0<br />

- Moisture content, max. % 3.0<br />

- Loss on ignition, max. % 10.0<br />

Supplementary optional chemical requirements. These optional requirements apply only<br />

when specifically requested.<br />

- Magnesium oxide (MgO), max. % 5.0<br />

- Available alkalies, as Na 2 O equiv., max. % 1.5<br />

For the time being the cement producer should control the quality of a natural pozzolan by<br />

applying a combination of following methods:<br />

• chemical composition<br />

• mineralogical composition<br />

• physical properties<br />

• strength development of a blended cement (ASTM C 1157) or of a concrete mix<br />

4.3.3 Artificial Pozzolans<br />

4.3.3.1 SILICA FUME<br />

Silica fume is a by-product of the manufacture of silicon metal (Figure 29). Silicon metal is<br />

used for the production of ferrosilicon, which is required by the steel industry for the<br />

improvement of its products.<br />

SiO-gas produced in an electric arc furnace is oxidized, condensed to ultra fine SiO 2 -particles<br />

and subsequently separated from the flue gases in a baghouse filter where it is collected as<br />

silica fume.<br />

Figure 29<br />

Production of Silica Fume<br />

CO + O ⎯⎯→ CO 2<br />

SiO + O ⎯⎯→ SiO 2 (silica fume)<br />

↑<br />

QUARTZ + COAL<br />

(SiO 2 + 2C)<br />

ELECTRIC ARC FURNACE<br />

(2000°C)<br />

⎯⎯→ Si (silicon metal)<br />

(98% purity)<br />

Condensed silica fume consists of very fine spherical particles with a high content of<br />

amorphous silica. It is an extremely fine powder, very much finer than Portland cement or fly<br />

ash, its fineness being roughly comparable with the finest particles in tobacco smoke (Table<br />

18).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 115 of 220


Cement<br />

Concrete<br />

Table 18<br />

Fineness of Different Materials<br />

Type of Material<br />

Fineness (cm 2 /g)<br />

Silica fume approx. 200’000<br />

Tobacco smoke approx. 100’000<br />

Fly ash 2’000 to 5’000<br />

Normal Portland cement approx. 3’000<br />

The grain size distribution is shown in Figure 30. The extreme fineness gives rise to a<br />

number of problems in handling and transporting this material. Its bulk weight is not higher<br />

than 300 kg/m 3 . Exposed piles are stirred by the least breath of wind. Silica fume is either<br />

packed in bags as such, or densified and packed in bags, as well as in the form of a 50%<br />

water slurry. Transport and transshipment facilities clearly necessitate efficient dedusting<br />

systems.<br />

The chemical composition (Table 19) may vary somehow depending on the impurities of<br />

quartz, the features of the manufacturing process, the dust precipitation facilities and the coal<br />

used. The effect of different compositions can become quite significant depending on the<br />

particular type of application.<br />

Table 19<br />

Chemical Composition of Silica Fume<br />

Elements<br />

Range<br />

Min.<br />

Max.<br />

Loss on Ignition 0.7 2.5<br />

SiO2 90.0 96.0<br />

Al2O3 0.3 3.0<br />

Fe2O3 0.2 0.8<br />

CaO 0.1 0.5<br />

MgO 0.5 1.5<br />

NaO 0.2 0.7<br />

HO 0.4 1.0<br />

C 0.5 1.5<br />

S 0.1 0.4<br />

Only silica fume that meets the following requirements shall be used:<br />

- amorphous silica SiO2 ≥ 85% by mass<br />

- loss on ignition ≤ 4% by mass<br />

- specific surface (untreated) ≥ 15 m 2 /g (BET)<br />

Owing to the great fineness, its completely amorphous state and its high content of SiO2,<br />

silica fume exhibits properties closely resembling those of pozzolan. However, the usual<br />

methods for characterizing the pozzolanic activity are not suitable and they had to be first<br />

adapted to suit its unusual properties.<br />

The high pozzolanic activity made it seem obvious to add the silica fume directly to cement<br />

or concrete. Tests performed in a great number of laboratories proved that there are a<br />

number of problems connected with the handling properties of the silica dust, however, they<br />

can be overcome when suitable modifications to the process technology are made.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 116 of 220


Cement<br />

Concrete<br />

The level of replacement cement by silica fume is limited and lies generally in the range of 8<br />

2%. The following hardening characteristics (Table 20) have been observed for an ISO<br />

standard mortar with a pure Portland cement compared to a 5% silica fume substitution.<br />

Table 20<br />

Characteristics of an ISO Standard Mortar<br />

Strength after ... Days Pure Portland Cement 95% PC<br />

+ 5% Silica Fume<br />

2 20.3 N/mm 2 21.4 N/mm 2<br />

28 42.8 54.1<br />

90 53.3 61.2<br />

The amount of water required by the binder is higher and the setting time is generally<br />

somewhat shorter compared to pure PC. The increase of strength is remarkable and conduct<br />

naturally to the use of silica fume for the production of high performance concrete. In order to<br />

reduce the higher water requirement, it is recommended to add superplasticizer in the<br />

concrete mix. In spite of its remarkably high price, this material is chosen in most cases<br />

where production of a high performance concrete is required.<br />

Figure 30<br />

Grain Size Distribution<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 117 of 220


Cement<br />

Concrete<br />

4.3.3.2 THERMALLY ACTIVATED MINERALS AND ROCKS<br />

Artificial Pozzolans of this type are produced by means of thermal treatment from nonpozzolanic<br />

materials (e.g. marl, shale) or from materials with very low pozzolanic properties<br />

(e.g. phonolite). They are activated by a rise of temperature which lies between 150°C and<br />

about 1500°C depending on the nature of the raw material. The materials develop pozzolanic<br />

properties by a transformation of their crystalline structure into a glass.<br />

Clays, shales and rocks with very low pozzolanic properties need a calcination at<br />

temperatures between 500 and 1000°C in order to be activated.<br />

Burnt clays<br />

After calcination at a relatively low temperature (usually below 800 °C), some clays can<br />

be converted to pozzolanic materials, after production of reactive phases by alteration of<br />

the original crystalline structure of the clay. The required heat treatment is generally<br />

variable, since it depends on the nature and fineness of clays. Both the calcination<br />

temperature and time have to be adjusted to get best enhancement of pozzolanic<br />

activity.<br />

Burnt Shale, Calcined Marls and ‘KALSIN’<br />

Burnt shale, in special cases burnt oil shale, calcined marls and Kalsin (a patented product of<br />

the ‘Holderbank’ Group) are produced in conventional rotary kilns as well as fluidized beds at<br />

a temperature of approximately 800°C. Owing to the composition of the natural materials and<br />

the manufacturing process, these products contain hydraulic mineral phases identical to<br />

those produced during the clinkering operation. These phases are mainly dicalcium silicate<br />

and monoaluminate as well as reactive gehlenite and calco-spurrite. Due to the specific<br />

burning process (generally under reducing atmosphere) only small amounts of free lime is<br />

produced whereas calcium sulfate proportion depends essentially on the sulfur present in the<br />

raw material. Beside these hydraulic minerals, large proportions of pozzolanic reacting<br />

oxides, especially silica and alumina are present.<br />

In a very finely ground state burnt shale, calcined marls and Kalsin show like Portland<br />

cement pronounced hydraulic and in addition pozzolanic properties.<br />

According to EN 196-1 these burnt products must develop a compressive strength of at least<br />

25.0 N/mm 2 after 28 days. If not they are classified in the category of artificial pozzolan.<br />

The expansion shall be less than 10 mm (Le Châtelier test) in accordance with EN 196-3<br />

using a mixture of 30% by mass burnt material plus 70% by mass reference cement.<br />

4.3.4 Applications<br />

As discussed in chapter 3.3 natural and artificial pozzolans are preferably used wherever a<br />

high durability, high long-term strength or a low heat development is needed. This is<br />

especially the case in structures severely exposed to water of different chemical quality and<br />

in mass concrete applications such as:<br />

• Spillways<br />

• Stilling basins<br />

• Water canals<br />

• Water and waste-water treatment facilities<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 118 of 220


Cement<br />

Concrete<br />

• Landfill construction purposes<br />

• Underwater constructions and repairs<br />

• Dams<br />

• Bridges (Figure 31)<br />

• Offshore platform construction<br />

• Bank vault construction<br />

• High-pressure concrete pipes<br />

• Concrete in high-rise structures<br />

• Shotcrete operations<br />

Figure 31<br />

Construction of Golden Gate Bridge (San Francisco, USA) in 1932 using<br />

a Portland pozzolan cement with 25% interground calcined siliceous<br />

shales<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 119 of 220


Cement<br />

Concrete<br />

5. References<br />

Locher, W. (2000) Zement, Grundlagen der Herstellung und Verwendung, Verlag Bau +<br />

Technik, 522 p.<br />

Matthes, W. (2000) Properties, Grindability, and Reactivity of Granulated blast-furnace slag:<br />

A Literature Review, HMC report MIC 00/5005/E, 91 p.<br />

Weber, R., Bilgeri, P., Kollo, H, Vissmann, H.-W (1998) Hochofenzement -<br />

Eigenschaften und Anwendung im Beton, Verlag Bau und Technik, 2nd edition, 56 p.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 120 of 220


Cement<br />

Concrete<br />

Cement Admixtures<br />

1. CEMENT ADMIXTURES / PROCESSING ADDITIONS................................................. 122<br />

2. GRINDING AIDS AND PERFORMANCE ENHANCERS ............................................... 122<br />

2.1 Functions and effects of cement admixtures.......................................................... 122<br />

3. ADDITIVES FOR THE MANUFACTURE OF MASONRY CEMENT .............................. 123<br />

4. FUNCTIONAL ADDITIONS FOR USE IN HYDRAULIC CEMENTS .............................. 124<br />

5. THE MOST FREQUENT SUPPLIERS OF CEMENT ADMIXTURES ARE: ................... 124<br />

6. SERVICES OF CEMENT ADMIXTURE SUPPLIERS: ................................................... 124<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 121 of 220


Cement<br />

Concrete<br />

1. Cement Admixtures / Processing Additions<br />

Definition: Cement Admixtures are added to the cement raw mix during or before the<br />

grinding process in amounts of less than 1% of cement weight. Typical<br />

dosages are 200 - 500 g admixture per ton of cement. Cement Admixtures<br />

are based on organic and inorganic compounds.<br />

An inquiry on application of cement admixtures within the "Holcim"-group was answered by<br />

80 plants in 1998. We received the following results:<br />

• with 57% Cost saving is the most-used argument for<br />

admixture application.<br />

• Some 41% of the "Holcim" plants answered that<br />

admixture application is a must for receiving the desired cement quality (pack-set,<br />

early strength).<br />

• On the other hand special characteristics are with 2% of<br />

applications real minor in quantity. But with cement admixtures some niche<br />

products are produced as well, with good additional customer values (and<br />

interesting margins).<br />

Cement admixtures can be divided into the following groups, depending on their purpose<br />

of use:<br />

1. Grinding Aids and Performance Enhancer.<br />

2. Additives for the Manufacture of Masonry Cement.<br />

3. Functional Additions for Use in Hydraulic Cements.<br />

2. Grinding Aids and Performance Enhancers<br />

Grinding Aids and Performance Enhancers are traditionally most widespread additions in<br />

cement manufacture. In 1998, 77% of the "Holcim" plants used such cement admixtures. In<br />

Europe 100%, in North America 96%, and in Latin America 83% of the cements with a<br />

fineness above 400 m 2 /kg were produced with admixtures. One or several of the following<br />

effects are achieved.<br />

2.1 Functions and effects of cement admixtures<br />

Grinding aides - typically organic compounds based on alcohol and amines - do not really<br />

change the engineering properties of cement. The action of the grinding aids is based on the<br />

reduction of the adhesive forces between the cement particles. They may, however, facilitate<br />

the handling of the cement due to the resulting improvement in flowability.<br />

Grinding aids acts on:<br />

• Coating reduction of grinding bodies<br />

(Reduction of specific energy consumption / improvement of specific mill output)<br />

• Improvement of flow properties / "pack-set" reduction<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 122 of 220


Cement<br />

Concrete<br />

Performance modifiers influencing significantly the cement quality, in particular water<br />

requirement and strength development.The performance modifiers are of similar nature as<br />

the products used in the concrete mix. Such admixtures are generally based on accelerators<br />

and water reducers and thus improve the workability and strength development of the<br />

cement. The performance enhancer allows:<br />

• Optimized particle size distribution (lower average diameter at same Blaine) →<br />

higher strength<br />

• Activation of clinker hydration<br />

• Activation of mineral components (GBFS, Puzzolan)<br />

• Water reduction in case ASTM - testing (constant flow)<br />

Cement admixtures are placed in the first compartment of the mill or on the clinker on the<br />

conveyer belt at mill feed. Cement admixture has shown positive effects on all mill types.<br />

The dosage of cement admixtures is a function of the cement fineness and the desired<br />

effect. Admixtures used as processing additions are added within a dosage range of 100 -<br />

200g/t of active materials. As performance enhancer a minimum content of 200g/t is required<br />

and the dosage may be as high as 3kg /t. Limits by EN197/1 are 0.5% for active organic<br />

material and 1% in general (except for pigments).<br />

As a quality target, periodically the flow properties of the cements should be measured with<br />

an adequate test. We recommend the "HGRS-cement flow test". This measurement is<br />

specially recommended during new application tests with processing admixtures.<br />

To produce ASTM standard cements with cement admixtures testing of these products is<br />

required. Cement admixtures have to fulfill requirements according to ASTM C 465<br />

(Standard Specification for Processing Additions for Use in the Manufacture of Hydraulic<br />

Cements).<br />

As a quality target it is recommended to use ASTM 465 requirements for all cement<br />

admixtures used by "Holcim" plants.<br />

3. Additives for the manufacture of masonry cement<br />

Most of the masonry cements are produced in America and in Europe. There are three<br />

mortar types recognized under ASTM C270 "Standard Specification for Mortar for Unit<br />

Masonry". They are type N (750 psi min.), type S (1800 psi min.) and type S (2500 psi min.).<br />

Masonry cements are produced with limestone addition and an integrated air entrainer to get<br />

an air content of up to 18% in the fresh mortar. In the masonry cement production, various<br />

types of admixtures are used:<br />

• Air Entrainers (e.g. resin soaps, surfactants)<br />

• Retarding Admixtures (e.g. phosphates, saccharids)<br />

• Water Retaining Agents (e.g. modified cellulose)<br />

• Hydrophobing Agents (e.g. salts of fatty acids)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 123 of 220


Cement<br />

Concrete<br />

4. Functional additions for use in Hydraulic Cements<br />

Functional additions for use in hydraulic cements are described in ASTM C 688. On<br />

purchasers request, hydraulic cements may be modified by functional additions out of the<br />

following classes:<br />

• Accelerating Additions<br />

• Retarding Additions<br />

• Set-Control Additions<br />

• Water-Reducing and Accelerating Additions<br />

• Water-Reducing and Retarding Additions<br />

Cements produced according to this standard are always speciality products for special<br />

markets. In certain regions special "low water binders" for high performance concrete<br />

production were developed. Cements with addition of water-reducing admixtures are one<br />

solution.<br />

Addition of accelerator to cement can be used to produce speciality cements for spayed<br />

concrete applications (dry shotcrete).<br />

5. The most frequent suppliers of Cement Admixtures are:<br />

‣ Grace<br />

‣ Holderchem Building Chemicals, Lebanon (Société Ciments Libanais)<br />

‣ Westvaco<br />

‣ CP-Cervices<br />

‣ Roan Industries<br />

‣ Tecnochem<br />

‣ Chryso (Lafarge.Group)<br />

‣ Addiment (Heidelberger-group)<br />

‣ Chemical companies (Hoechst, BASF, Shell, Union Carbid, … ) for base materials<br />

like Diethylene glycol, Dipropylene glycol and Triethanolamine<br />

6. Services of Cement admixture suppliers:<br />

‣ Consulting in optimum cement mix design<br />

‣ <strong>Support</strong> in process optimization<br />

‣ Providing of equipment (Dosage pumps, storage tanks)<br />

‣ Performing of pre-test with local components to evaluate optimum admixtures.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 124 of 220


Cement<br />

Concrete<br />

Cement Grinding<br />

1. GENERAL....................................................................................................................... 126<br />

1.1 Grinding of Composite Cement.............................................................................. 126<br />

1.2 Particle size distribution in the different grinding systems...................................... 127<br />

2. CONCEPTS FOR THE PRODUCTION OF COMPOSITE CEMENTS .......................... 128<br />

2.2 Temperature and moisture conditions.................................................................... 133<br />

3. GRINDING TECHNOLOGY FOR CLINKER AND MIC'S............................................... 136<br />

3.1 The Ball Mill option................................................................................................. 137<br />

3.2 The Vertical Roller Mill option ................................................................................ 137<br />

3.3 The Roller Press option.......................................................................................... 138<br />

3.4 The Horizontal Roller Mill option ............................................................................ 139<br />

3.5 System Design and equipment Selection .............................................................. 140<br />

4. FINAL COMMENTS........................................................................................................ 145<br />

5. MESSAGES .................................................................................................................... 145<br />

6. LITERATURE.................................................................................................................. 146<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 125 of 220


Cement<br />

Concrete<br />

1. General<br />

The cement components have to be ground to fine particles, in order to attain the required<br />

cementitious properties. The fineness after grinding is usually characterized by the specific<br />

surface area or by the particle size distribution (PSD). The type of cement mill used can have<br />

a considerable effect on the PSD.<br />

During the grinding process with the traditional systems (ball mills), only a small portion of<br />

the introduced energy is consumed for the comminution of the cement particles. A large<br />

quantity of heat is set free and the temperature of ground cement increases appreciably. In<br />

the modern grinding systems, less heat is produced, resulting in lower cement temperature<br />

during grinding.<br />

Both, the fineness and the temperature of grinding are principal factors in determining the<br />

cement properties.<br />

1.1 Grinding of Composite Cement<br />

The properties of composite cements are decisively influenced by the fineness of the cement<br />

and its components respectively. Composite cements must generally be ground to a higher<br />

overall fineness than Portland cements to maintain a similar strength development.<br />

The grinding behaviour of the different components in composite cements may vary quite<br />

significantly as illustrated in Figure 1. At constant fineness, the softer materials like limestone<br />

and natural pozzolan yield a wider PSD than the clinker and blast furnace slag. Despite its<br />

worse grindability, the grain size distribution of blast furnace slag does, however, not differ<br />

too much from the one of clinker. The mentioned differences in grindability are of great<br />

significance in the grinding of composite cements.<br />

1<br />

Steepness of Rosin-Rammler distribution<br />

0.8<br />

0.6<br />

0.4<br />

Characteristic diameter<br />

of Rosin-Rammler-distribution<br />

d' = 16 µm<br />

0.2<br />

0 20 40 60 80 100 120<br />

Grindability index in cm²/(g·s)<br />

Blast furnace slag Clinker Pozzolan Limestone<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 126 of 220


Cement<br />

Concrete<br />

Figure 1: Steepness of the RRS-distribution of ground blast furnace slag, clinker, pozzolan<br />

and limestone at same characteristic diameter d' in function of the grindability index<br />

1.2 Particle size distribution in the different grinding systems<br />

The grinding in the modern cement mills goes together with a narrower PSD of the produced<br />

cements. Due to the more efficient grinding process, less under- and over-size particles are<br />

produced, which obviously leads to a shorter particle range and to higher steepness of the<br />

PSD.<br />

The steepness n as expressed by the RRS-distribution ranges from 0.8 for an open circuit<br />

ball mill up to 1.2 for the newest grinding systems like vertical mill. Typical n values of<br />

cements ground in various industrial mill systems are indicated in Table 1.<br />

The characteristic diameter d' of commercial cements produced in the different grinding<br />

systems varies typically between 10 and 30 µm. For identical specific surface, the d' values<br />

are lower in the systems which give a narrower PSD, that means that the overall fineness of<br />

the cement at same Blaine will be higher. At same d', the Blaine will be lower when the PSD<br />

gets narrower.<br />

Table 1: Typical n values for PSD of cements ground in various industrial mill systems<br />

Mill type Steepness n of RRS-distribution<br />

Ball mill (open circuit) 0.8 - 0.9<br />

Ball mill (closed circuit) 0.9 - 1.0<br />

Ball mill (high efficiency separator) 1.0 - 1.1<br />

Vertical mill, roller press, Horomill 1.1 - 1.2<br />

As mentioned before, the different PSD of the various grinding systems are also reflected in<br />

the relationship between Blaine and sieve residues (e.g. on 45 µm), which are the usual<br />

fineness measures applied in practice. The corresponding trends observed in the modern<br />

grinding systems compared to the traditional ball mills are as follows:<br />

♦ lower Blaine at same sieve residue or<br />

♦ lower sieve residue at same Blaine<br />

It is important to mention that, in view of certain quality problems experienced with a too<br />

narrow PSD, the actual tendency for the new grinding systems is to adjust the PSD to a<br />

somewhat wider distribution.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 127 of 220


Cement<br />

Concrete<br />

2. Concepts for the production of composite cements<br />

Composite cements are produced according to one of three basically different grinding<br />

concepts:<br />

<br />

<br />

<br />

by compound grinding of the constituent cement components,<br />

by separate grinding oft the cement components, or<br />

by semi-compound grinding of the components.<br />

Each of these solutions has its merits and disadvantages.<br />

2.1.1 Compound grinding<br />

The compound grinding of clinker, gypsum and mineral component(s) is still the most<br />

common practice for the production of composite cements.<br />

The combined grinding with mineral components softer than the clinker like limestone will<br />

widen the grain size distribution of the resulting composite cement, whereas the mixture of<br />

clinker with a harder material like blast furnace slag will give a somewhat steeper PSD than<br />

the ground clinker alone. The different PSD can be explained by the fact that for compound<br />

grinding, the harder material is enriched in the coarser fraction and the softer material in the<br />

finer fractions of the cement.<br />

In compound grinding, the different cement components can accordingly not be ground<br />

individually or independently from each other. For a given grinding system, the fineness of<br />

the components is pre-determined by their respective grindabilities; it is thus not possible to<br />

adjust freely their fineness. The inevitable enrichments of certain components in the fine or<br />

coarse fraction of the composite cement are, however, reduced with the modern grinding<br />

technologies.<br />

This lack of flexibility in compound grinding may limit the optimisation of the properties of<br />

composite cements. With soft mineral components, the clinker will always remain rather<br />

coarse, in particular at higher replacement levels, so that its hydraulic potential cannot be<br />

fully exploited. On the other hand, there might be an overgrinding of the mineral component<br />

like in the case of the natural pozzolans leading to an increase in water demand. This is not<br />

optimum solution also with regard to the energy efficiency of the grinding process.<br />

In case of slag cement, the clinker will get indeed finer and contribute as desired to the<br />

strength development. The slag may, however, not be adequately refined and activated.<br />

Compound grinding of the cement components is a strong solution in cases<br />

<br />

<br />

<br />

that sufficient grinding capacity is available anyway,<br />

that the portion of mineral components added to the cement is less than 30%, thus<br />

rather low, and<br />

that the envisaged annual production rates are low.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 128 of 220


Cement<br />

Concrete<br />

2.1.2 Semi-compound Grinding<br />

Semi-compound grinding is a solution that has been selected sporadically only for production<br />

of composite cements (e.g. for the Tecoman and the Nobsa works). The concept allows for<br />

separate pre-grinding of the harder component, preventing by doing so that the soft<br />

component is overground. It is evident that energy efficiency of the grinding process is better<br />

compared to compound grinding. Disadvantage is the additional grinding unit required, above<br />

all in the case of low capacity installations. Despite this disadvantage semi-compound<br />

grinding is an option to be considered in case that a works grinding capacity needs to be<br />

increased as a prerequisite for the production of composite cements.<br />

2.1.3 Separate grinding<br />

Separate grinding of composite cements gives more flexibility in the design and optimisation<br />

of the cement quality than compound grinding, since it permits free choice of the fineness of<br />

the cement components. Nevertheless, the opinions on the real benefits of separate griding<br />

are still controversial: It is evident that energy efficiency of such a concept is high.<br />

Disadvantage again is the additional grinding unit required.<br />

Separate grinding of the cement components is an option to be considered in cases:<br />

<br />

that introduction of composite cement ask for the installation of additional grinding<br />

capacity,<br />

that the portion of mineral components added to the cement is high (i.e. exceeding 40%),<br />

<br />

<br />

that production of client specific products, thus a high product variety is required, and<br />

that high annual production rates of composite cements are required.<br />

In the following, the experience with separate grinding for the most relevant composite<br />

cements containing blast furnace slag, fly ash, natural pozzolan and limestone is discussed.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 129 of 220


Cement<br />

Concrete<br />

2.1.3.1 SLAG CEMENTS<br />

The studies on separate grinding of slag cement revealed that the fineness of the clinker and<br />

slag influence the cement quality in the following way:<br />

♦ the clinker fineness is mainly related to early strength. In cements with low slag content<br />

(


Cement<br />

Concrete<br />

2.1.3.2 FLY ASH CEMENTS<br />

The separate grinding of fly ash cement as such has hardly been investigated. It appears<br />

that at constant grinding energy separate grinding gives certain possibilities to fine tune the<br />

final strength development of the fly ash cement.<br />

Presently the most appropriate solution to produce fly ash cements is to add the fly ash to<br />

the separator of the grinding system. The example in Figure 3 of a cement containing 16% fly<br />

ash demonstrates that in terms of quality and consumption of grinding energy, this seems to<br />

be the best solution, also compared to intergrinding.<br />

60<br />

50<br />

OPC 30<br />

OPC 40<br />

28 d<br />

MPa (Rilem)<br />

40<br />

30<br />

20<br />

addition to separator<br />

intergrinding<br />

mixing<br />

OPC 40<br />

10<br />

OPC 30<br />

2 d<br />

0<br />

0 10 20 30 40 50<br />

kWh/t cem<br />

Figure 3: Influence of treatment undergone by the fly ash on the strength of cement with 16%<br />

fly ash<br />

2.1.3.3 POZZOLANIC CEMENTS<br />

Separate grinding of cements with natural pozzolan gives a somewhat higher early strength<br />

and lower final strength than intergrinding at constant energy input. This relationship seems<br />

quite logical as in intergrinding the clinker responsible for the early strength remains rather<br />

coarse and the pozzolan contributing to the final strength is refined.<br />

The studies on separate grinding of pozzolanic cements carried out at Holcim showed a<br />

certain potential to increase the clinker factor at same cement quality, though at the expense<br />

of a higher grinding energy. A typical increase might be in the order of 5% as it is illustrated<br />

by the comparison of intergrinding and separate grinding for a pozzolanic cement from<br />

Mexico in Table 2.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 131 of 220


Cement<br />

Concrete<br />

Table 2: Comparison of intergrinding and separate grinding for pozzolanic cement from<br />

Mexico<br />

Physical and mechanical<br />

properties<br />

Compound grinding (ball mill)<br />

Separate grinding<br />

(ball mill and vertical mill)<br />

Cement<br />

Pozzolan (%) 20 25<br />

Blaine (cm 2 /g) 4170 4120<br />

n (-) 1.0 x)<br />

R 45 µm (%) 3.3 7.5<br />

Paste ASTM<br />

Water demand (%) 28.3 27.0<br />

Setting time (min.)<br />

- initial 160 145<br />

- final 225 175<br />

Mortar ASTM<br />

w/c-ratio 0.52 0.53<br />

Compressive strength (MPa)<br />

- at 1 day 11.0 11.6<br />

- at 3 days 20.9 20.7<br />

- at 7 days 27.8 26.0<br />

- at 28 days 35.0 34.8<br />

x) Pozzolan: n = 0.95 R 45 µm = 20.9% Blaine = 2170 cm 2 /g<br />

Clinker/gypsum: n = 0.95 R 45 µm = 3.1% Blaine = 4300 cm 2 /g<br />

2.1.3.4 LIMESTONE CEMENTS<br />

Studies on separately ground limestone showed that the limestone fineness as such has<br />

virtually no influence on the strength development. The PSD of the limestone powder can,<br />

however, play a role with regard to the workability characteristics of the limestone cement: a<br />

wide distribution is in this respect more favourable than a narrow one.<br />

In combined grinding with clinker, the limestone is automatically ground to the favourable<br />

wide PSD. This quite advantageous behaviour in intergrinding, at least at lower limestone<br />

dosages (up to 20%) may also explain the fact that separate grinding of limestone cement to<br />

improve cement quality has usually not been applied in practice.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 132 of 220


Cement<br />

Concrete<br />

2.2 Temperature and moisture conditions<br />

The grinding of cement influences the properties of cement not only through an increase in<br />

fineness, but also through the reactions taking place in the cement mill. Depending on the<br />

temperature and moisture conditions prevailing in the mill, dehydration and hydration occur<br />

which influence the grinding process, flowability, lump formation in silos, setting and<br />

hardening of cement.<br />

Due to the heat liberated during the grinding process, the temperature in the traditional ball<br />

mills rises to temperatures above 100°C. In the modern grinding systems, the grinding<br />

temperatures are significantly lower (down to 50°C - 60°C). In a particular mill, the exit<br />

temperature of cement can vary in a wide range, in function of the inlet temperature of<br />

clinker, cooling conditions and fineness of grinding.<br />

The effect of the grinding temperature on the cement properties is mainly related to the<br />

dehydration of gypsum. With increasing temperature, gypsum (CaSO 4 2H 2 O) gets unstable<br />

and transforms to hemihydrate and anyhdrite III under the release of water. The dehydration<br />

of the gypsum will not only depend on the temperature, but also on the time the gypsum is<br />

exposed to this temperature (Figure 4). Another factor of less importance for gypsum<br />

dehydration is the humidity in the mill atmosphere (Figure 5).<br />

Figure 4: Influence of temperature on the dehydration of gypsum<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 133 of 220


Cement<br />

Concrete<br />

Figure 5: Influence of humidity on the gypsum dehydration<br />

According to the degree of dehydration, the gypsum will exert a different influence on the<br />

cement properties (see also paper on cement hydration). If great part of the gypsum is<br />

converted to the more easily soluble hemihydrate and anhydrite III, there may be some<br />

problems with false setting in case a clinker of low reactivity is used. On the other hand, a too<br />

low degree of dehydration may lead to flash setting tendency with reactive clinkers.<br />

The dehydration of the gypsum may also have an impact on the storage stability of the<br />

cement. If the cement enters the silo with a high temperature (80 - 90°C), further water can<br />

be released form the still not dehydrated gypsum and lead to the hydration of the cement.<br />

These hydration reactions can cause lump formation and affect the strength of the cement<br />

(Figures 6 and 7). Clinkers with high C 3 A and alkali content are particularly subject to<br />

hydration reactions during storage.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 134 of 220


Cement<br />

Concrete<br />

Figure 6: Lump formation in a storage sensitive cement after one-week storage at various<br />

temperatures<br />

Figure 7: Compressive strength of a storage sensitive cement after one week Storage at<br />

various temperatures<br />

Some measures to ensure the storage stability of the cement are:<br />

♦ low cement temperature in the silo<br />

♦ short storage time<br />

♦ reduction of gypsum content in the cement<br />

♦ substitution of gypsum by natural anhydrite<br />

The storage stability is usually less problematic with the new grinding technologies, where<br />

the temperatures of the cement coming from the miss are generally low. The cement<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 135 of 220


Cement<br />

Concrete<br />

temperatures in ball mills can be lowered by cooling the cement during the grinding process<br />

(e.g. water injection) or by installing a cement cooler after the mill. For the cooling by means<br />

of water injection, the temperature in the mill should always be kept above 100°C. Otherwise,<br />

prehydration of the cement and strength losses may occur.<br />

3. Grinding technology For clinker and mic's<br />

As seen before, the response of the different cement constituents to grinding can be quite<br />

different:<br />

<br />

<br />

<br />

Grindability is good for limestone but gets worse for clinker and blast furnace slag.<br />

Clinker is at about two times harder to grind compared to limestone and granulated blast<br />

furnace slag even up to three times.<br />

The slope of product particle size distribution too varies in a wide range. It is flat for fly<br />

ashes but steeper for clinker and ground granulated blast furnace slag. Particle size<br />

distribution of the composite cement affects sensitive characteristics as rheological<br />

properties and strength performance.<br />

The effect on (hydraulic) activity is different for the different mineral components.<br />

Limestone in contrast to slag do not contribute to the (hydraulic) activity and by this the<br />

strength performance of a concrete based on composite cement.<br />

It is evident that the nature of the mineral component that is selected for a project affects<br />

system design of as well as equipment selection for the grinding installation.<br />

The question to which fineness the components of composite cements shall be ground to<br />

obtain optimum cement properties is often debated. The general concept of Holcim Group<br />

<strong>Support</strong> is that the hydraulic potential of the clinker should be used as much as possible by<br />

grinding it to a sufficiently high fineness. The mineral components may be ground coarser,<br />

but latent hydraulic and pozzolanic materials must still have a sufficient fineness to be<br />

suitably activated to provide good final strength.<br />

Selection of the appropriate grinding technology is out of a variety of options. Prerequisite for<br />

finding the optimum solution for a given project is the knowledge of all process technological<br />

possibilities.<br />

Conceptual system design follows two lines:<br />

single stage grinding, and<br />

two stage grinding.<br />

Speaking of single stage grinding, equipment selection is basically out of four options<br />

(Figures 8a to 8d):<br />

<br />

<br />

<br />

<br />

the ball mill system,<br />

the vertical roller mill system,<br />

the roller press system and<br />

the horizontal roller mill system.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 136 of 220


Cement<br />

Concrete<br />

Figure 8b: Vertical roller mill<br />

Figure 8a: Ball mill<br />

Figure 8c: Roller press system<br />

Figure 8d: Horizontal roller mill<br />

3.1 The Ball Mill option<br />

The ball mill is typically operated in a closed circuit with a dynamic separator. This concept<br />

represents the conventional solution in cement and mineral component grinding. Operation<br />

of ball mill systems is easy, not asking for highly skilled staff, although optimisation for<br />

efficient performance requires quite some effort. Common product fineness levels, which<br />

may be as high as 5700 cm 2 /g (e.g. for ground granulated blast furnace slag of the Grade<br />

120 class) are not a limiting factor for the use of a ball mills. Product drying can take place<br />

within the mill, but the mill can handle pre-dried components without any problems as well.<br />

Product handling within the system is not problematic.<br />

But energy efficiency of the ball mill is poor. This limits the system's production rate despite<br />

the permitted high drive power of up to 6500 kW, in rare cases of up to 8000 kW.<br />

3.2 The Vertical Roller Mill option<br />

In vertical roller mill systems (Fig. 9) product collection is by means of a bag type dust<br />

collector. The concept represents the standard solution in raw and coal grinding.<br />

Although the application for grinding clinker and mineral components is proven since many<br />

years, the market was sceptic regarding product quality. Major concern was the steeper<br />

slope of the product grain size distribution that affects the cements rheological properties as<br />

well as its strength performance. Recent analysis of cements and fine slag produced on<br />

vertical mill systems in the Far East proved that the vertical roller mill products are<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 137 of 220


Cement<br />

Concrete<br />

comparable with ball mill products with regard to particle size distribution, rheological<br />

properties and strength performance.<br />

The vertical roller mill is the optimum concept regarding product drying. It is fit for both,<br />

compound and separate grinding applications. Mill energy efficiency is significantly higher<br />

compared to a ball mill (about 2.5 times for slag grinding; about 2 times for clinker grinding).<br />

The largest mill sizes available allow for a drive power rating of about 4000 kW and as a<br />

consequence for high production rates.<br />

But the application of the vertical mill concept is limited by the maximum product fineness<br />

levels that safely can be achieved in a single stage, namely 5000 cm 2 /g for granulated blast<br />

furnace slag and 4500 cm 2 /g for ordinary Portland cement.<br />

Product<br />

FIGURE 9: Vertical mill<br />

3.3 The Roller Press option<br />

The roller press (Fig. 10) as a single stage grinding system is typically operated in a closed<br />

circuit with a disagglomerator and a dynamic separator. The product is collected in a bag<br />

type dust collector. The market was sceptic with regard to product quality for the same<br />

reasons as mentioned for the vertical roller mill concept.<br />

Energy efficiency of the roller press is slightly better than for the vertical roller mill, thus<br />

significantly better compared to a ball mill (about 2.5 times for slag grinding; up to 2 times for<br />

clinker grinding). The largest press sizes available allow for a drive power rating of up to 2 x<br />

2200 kW and as a consequence for high production rates. Product fineness levels that safely<br />

can be achieved in a single stage installation, namely 5500 cm 2 /g for granulated blast<br />

furnace slag and 4500 cm 2 /g for ordinary Portland cement hardly limit the system's<br />

application.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 138 of 220


Cement<br />

Concrete<br />

As drying is possible in the separator only feeding roller presses is critical in that respect that<br />

hot and dry recycle material is mixed with moist fresh feed. This may result in local<br />

condensation of the generated vapours and by this in material build-up, e.g. in the shaft of<br />

the recycle bucket elevator and auxiliaries dust collectors. As a consequence heating and<br />

insulation of the auxiliary equipment is indispensable as to prevent that system temperature<br />

fall locally below dew point.<br />

Furthermore roller presses are sensitive to damages of the roller surfaces by pieces of<br />

tramp-metal and excessive feed granulometry of the fresh feed (the maximum size of the<br />

feed grains should as a rule of thumb not exceed two times the operational gap width).<br />

Product<br />

FIGURE 10: Roller press mill<br />

3.4 The Horizontal Roller Mill option<br />

The horizontal roller mill (Fig. 11) represents the most recently developed grinding<br />

technology. In a single stage configuration it is operated, similar as the roller press, in closed<br />

circuit with a dynamic separator but does not require the installation of a disagglomerator.<br />

The product is collected in a bag type dust collector.<br />

Energy efficiency is comparable to that of a vertical roller mill, thus significantly better<br />

compared to a ball mill (about 2.5 times for slag grinding; about 2 times for clinker grinding).<br />

The limits of the concept regarding production rate, achievable product fineness levels, ease<br />

of operation, etc. are not established yet. The concept is fit for compound as well as for<br />

separate grinding applications.<br />

But horizontal roller mill systems are limited in drying, as drying is possible in the separator<br />

only. Similar as for the roller press systems a complex dedusting system is required including<br />

heating and insulation of the auxiliary equipment as to prevent local condensation due to<br />

high dew points. The first horizontal roller mill installations suffered from mechanical reliability<br />

and operational problems, which seem to be solved in the meantime according to the<br />

information received from the supplier and some of the users.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 139 of 220


Cement<br />

Concrete<br />

The further development of the horizontal roller mill technology needs to be followed<br />

carefully. For the time being an industrial application would ask for careful analysis and<br />

system design.<br />

FIGURE 11: Horizontal mill<br />

3.5 System Design and equipment Selection<br />

A decision for a ball mill system is a low risk, but conservative selection.<br />

Energy efficiency of the available modern grinding technology based either on a vertical roller<br />

mill, a roller press or a horizontal roller mill is significantly better compared to that of the ball<br />

mill technology. This, as comminution is achieved by compression in a bed of material and<br />

no longer by friction.<br />

The vertical roller mill as well as the roller press can not only be used in a single-stage<br />

grinding arrangement but also as a pregrinder in a two-stage configuration with a ball mill<br />

(Figures 12a and b) for separate, semi-compound and compound grinding applications.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 140 of 220


Cement<br />

Concrete<br />

FIGURE 12a: Vertical roll mill + ball mill<br />

FIGURE 12b: Roller press mill+ ball mill<br />

The vertical roller mill system as a pregrinder to a ball mill may be considered in cases that<br />

high production rates and/or high fineness requirements must be met. Drying will exclusively<br />

take place in the vertical roller mill.<br />

The roller press will be considered above all for the modifications of ball mill systems with<br />

regard to a capacity increase or the introduction of composite cement production. Addition of<br />

a closed circuit roller press system for pre-grinding purposes results in a complex, difficult to<br />

operate system. Addition of an open circuit roller press system with slab recirculation as a<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 141 of 220


Cement<br />

Concrete<br />

pregrinding system is worth to be considered above all for semi-compound grinding<br />

applications.<br />

To facilitate selection of the appropriate grinding technology based on process technological<br />

criteria decision-trees have been developed<br />

<br />

<br />

for the production of composite cements based on compound grinding of the constituent<br />

components (Figure 13),<br />

for separate grinding of granulated blast furnace slag (Figure 14) and<br />

for the production of ordinary Portland cement (Figure 15).<br />

1) BASED ON ENERGY CONSUMPTION<br />

VRM, RP OR HM<br />

Fineness<br />

≤4500<br />

No, >4500cm 2 /g<br />

2) BASED ON ENERGY CONSUMPTION<br />

BM<br />

3) BASED ON TOTAL MILL FEED<br />

1)<br />

Yes<br />

Abs. Power Mill (FI)<br />

t/h x kWh/t<br />

No, ≤4000kW<br />

>4000kW<br />

Yes<br />

3)<br />

Feed moisture<br />

Yes<br />

Feed moisture<br />

2)<br />

Abs. Power Mill (BM)<br />

t/h x kWh/t<br />

≤2%<br />

Yes<br />

No, ≤X%<br />

No<br />

(Comparable BM)<br />

Particle Size Distr.<br />

Steep<br />

Yes<br />

≤2% Yes<br />

Particle Size Distr.<br />

No, ≤X%<br />

No<br />

(Comparable BM)<br />

Steep<br />

Yes<br />

>8000kW<br />

Yes<br />

Feed moisture<br />

No, ≤8000kW<br />

% Min. Component<br />

% Min. Component<br />

≤2%<br />

Yes<br />

high/<br />

low<br />

high<br />

low<br />

high/<br />

low<br />

high<br />

low<br />

No, ≤X%<br />

VRM<br />

HM<br />

RP<br />

2 VRM<br />

2 HM 2 RP VRM+BM RP+BM BM<br />

FI , ONE GRINDING UNIT FI , TWO GRINDING UNITS SF BM<br />

FIGURE 13: Decision-tree for the production of composite cements based on compound<br />

grinding of the constituents.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 142 of 220


Cement<br />

Concrete<br />

Fineness<br />

5300cm 2 /g<br />

1) BASED ON ENERGY<br />

CONSUMPTION<br />

VRM, RP OR HM<br />

1)<br />

Yes<br />

Abs. Power Mill (FI)<br />

t/h x kWh/t<br />

2) BASED ON ENERGY<br />

CONSUMPTION BM<br />

No, ≤4000kW<br />

>4000kW<br />

Yes<br />

Complexity of System<br />

Yes<br />

Complexity of System<br />

2)<br />

Abs. Power Mill (BM)<br />

t/h x kWh/t<br />

simple /<br />

complex<br />

complex<br />

simple /<br />

complex<br />

complex<br />

>8000kW<br />

No, ≤8000kW<br />

simple<br />

Particle Size Distr.<br />

simple<br />

Particle Size Distr.<br />

Yes<br />

medium<br />

Steep<br />

steep<br />

Yes<br />

No<br />

experience<br />

medium<br />

steep<br />

No<br />

experience<br />

Steep<br />

Yes<br />

VRM<br />

HM<br />

RP 2 VRM 2 HM 2 RP<br />

VRM+BM RP+BM BM<br />

FI , ONE GRINDING UNIT FI , TWO GRINDING UNITS SF BM<br />

FIGURE 14: Decision-tree for separate grinding of granulated blast furnace slag<br />

Fineness<br />

4000kW<br />

Yes<br />

Complexity of System<br />

Yes<br />

Complexity of System<br />

2)<br />

Abs. Power Mill (BM)<br />

t/h x kWh/t<br />

simple /<br />

complex<br />

complex<br />

simple /<br />

complex<br />

complex<br />

>8000kW<br />

No, ≤8000kW<br />

simple<br />

Particle Size Distr.<br />

simple<br />

Particle Size Distr.<br />

Yes<br />

medium<br />

steep<br />

Steep<br />

Yes<br />

medium<br />

steep<br />

Steep<br />

Yes<br />

VRM<br />

HRM<br />

RP 2 VRM 2 HRM 2 RP<br />

VRM+BM<br />

RP+BM<br />

2 BM<br />

BM<br />

SINGLE STAGE GRINDING<br />

ONE UNIT<br />

SINGLE STAGE GRINDING<br />

TWO UNITS<br />

TWO STAGE<br />

GRINDING<br />

SINGLE STAGE<br />

GRINDING<br />

FIGURE 15: Decision-tree for the production of ordinary Portland cement<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 143 of 220


Cement<br />

Concrete<br />

From these decision-trees the process technological superiority of the modern grinding<br />

technologies, above all the vertical roller mills becomes evident. This applies above all for<br />

granulated blast furnace applications with its high energy substitution factor (Figure 16).<br />

[cm2/g]<br />

SINGLE STAGE GRINDING<br />

ONE UNIT<br />

SINGLE STAGE GRINDING<br />

TWO UNITS<br />

TWO STAGE<br />

GRINDING<br />

SINGLE STAGE<br />

GRINDING<br />

VRM HM RP 2 VRM 2 HM 2 RP VRM+BM RP+BM 2 BM BM<br />

50% 47% 47% 50% 47% 47%<br />

79%<br />

82%<br />

100<br />

%<br />

100<br />

%<br />

≤ 4000 [kW] > 4000 [kW] > 4000 [kW]<br />

48% 46% 47%<br />

48% 46% 47%<br />

75%<br />

84%<br />

100<br />

%<br />

100<br />

%<br />

≤ 4000 [kW]<br />

> 4000 [kW] > 4000 [kW]<br />

72%<br />

87%<br />

100<br />

%<br />

100<br />

%<br />

FIGURE 16: Comparison between energy consumption for blast furnace slag grinding<br />

Factors that have influence on the selection of the grinding technology for composite cement<br />

applications are, besides process technological criteria:<br />

<br />

<br />

<br />

<br />

<br />

<br />

the geographical location,<br />

requirements and conditions of the local market,<br />

investment versus operation cost,<br />

flexibility in system operation,<br />

flexibility in product quality,<br />

etc.<br />

As to keep the decision trees simple these factors could not be considered.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 144 of 220


Cement<br />

Concrete<br />

4. Final comments<br />

‣ For "Holcim" group CTS/PT observes a trend towards compound grinding<br />

solutions for the production of pozzolana, limestone and fly ash based<br />

composite cements. Exceptions: Morning Star Cement, Vietnam; LIMED<br />

(CIOR), Morocco.<br />

‣ For composite cements based on granulated blast furnace slag separate<br />

grinding is the preferred concept.<br />

‣ With regard to fly ash addition to the cements CTS/PT observes that the<br />

concepts are followed in Europe and the US are quite different. In Europe fly<br />

ash is typically added to the separator feed preventing by this over-grinding<br />

of the already sufficiently fine portion while in the US fly ash is typically feed<br />

to the mill.<br />

‣ Regarding equipment selection still more weight is given to the criteria<br />

investment cost and technological risk, outweighing the advantage of energy<br />

savings and by this the selection of modern solutions. This despite industrial<br />

experience has confirmed the lower energy requirement, the reliability of<br />

modern grinding technologies. Reservations regarding product quality based<br />

on a too steep product particle size distribution can hardly be maintained any<br />

longer.<br />

5. Messages<br />

‣ The modern grinding systems have proven superior energy efficiency as well<br />

as operational reliability in grinding of clinker and mineral components.<br />

‣ For the selection of the appropriate technology system for new composite<br />

cement production facilities efficiency characteristics should be considered<br />

as well and not by principal lowest investment exclusively. Reservations<br />

regarding product quality based on product particle size distribution can<br />

hardly be maintained any longer.<br />

‣ Give modern, energy efficient grinding technology a chance. It should have a<br />

predominant place in the "Holcim" group.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 145 of 220


Cement<br />

Concrete<br />

6. Literature<br />

Grinding and cement properties<br />

Bapat, J.D., Higher qualities from modern finish grinding processes, International Cement<br />

Review, January 1998, pp. 54 - 56<br />

Montani, S., Influence of grinding on the properties of composite cements, 34th <strong>Technical</strong><br />

Meeting, Davos, 1996, PT 96/14'096/E<br />

Albeck, J., Kirchner, G., Influence of process technology on the production of marketorientated<br />

cements, Cement-Lime-Gypsum, No. 10, 1993, pp. 615 - 626<br />

Gebauer, J:, Cement grinding and quality problems, 32nd <strong>Technical</strong> Meeting, Montreux,<br />

1992, VA 92/5972/E<br />

Schiller, B., Ellerbrock, H.-G., The grinding and the properties of cements with several main<br />

constituents, Cement-Lime-Gypsum, No. 7, 1992, pp. 325 - 334<br />

Ellerbrock, H.-G., Deckers, R., Mill temperature and cement properties, dto., No. 1, 1988, pp.<br />

1 - 12<br />

Sprung, S., Kuhlmann, K., Ellerbrock, H.-G., Particle size distribution and properties of<br />

cement, dto., Part I: No. 4, 1985, pp. 169 - 178, Part II: No. 9, 1985, pp. 528 - 534<br />

Sprung, S., Influence of process technology on cement properties, dto., No. 10, 1985, pp.<br />

577 - 585.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 146 of 220


Cement<br />

Concrete<br />

Properties and Application of Composite Cements<br />

1. COMPOSITE (BLENDED) CEMENTS.............................................................................. 149<br />

1.1 Definition................................................................................................................. 149<br />

1.1.1 Abbreviations ........................................................................................... 149<br />

1.2 A brief history of composite cements...................................................................... 150<br />

2. GENERAL PROPERTIES OF COMPOSITE CEMENTS.................................................. 151<br />

2.1 Bleeding.................................................................................................................. 151<br />

2.2 Workability .............................................................................................................. 152<br />

2.3 Setting time............................................................................................................. 154<br />

2.4 Strength .................................................................................................................. 155<br />

2.5 Shrinkage................................................................................................................ 157<br />

2.6 Creep...................................................................................................................... 159<br />

2.7 Heat of hydration .................................................................................................... 160<br />

2.8 Porosity................................................................................................................... 162<br />

2.9 Durability................................................................................................................. 163<br />

2.10 Lime leaching.......................................................................................................... 164<br />

2.11 Carbonation ............................................................................................................ 165<br />

2.12 Sulphate attack....................................................................................................... 165<br />

2.13 Chloride attack........................................................................................................ 166<br />

2.14 Sea water attack..................................................................................................... 167<br />

2.15 Alkali Silica Reaction .............................................................................................. 168<br />

2.16 Frost resistance ...................................................................................................... 169<br />

3. ASPECTS OF MANUFACTURE OF COMPOSITE CEMENTS ....................................... 171<br />

3.1 Pretreatment of additions........................................................................................ 171<br />

3.2 Properties of the addition and type of grinding ....................................................... 171<br />

3.2.1 Grindability ............................................................................................... 171<br />

3.2.2 Strength activity........................................................................................ 172<br />

3.2.3 Type of grinding ....................................................................................... 172<br />

3.2.4 Use of chemical admixtures ..................................................................... 173<br />

3.3 Proportioning .......................................................................................................... 174<br />

4. TRENDS OF PRODUCTION AND MARKETING OF COMPOSITE<br />

CEMENTS....................................................................................................................... 175<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 147 of 220


Cement<br />

Concrete<br />

5. CONCLUSIONS ................................................................................................................ 176<br />

6. LITERATURE REFERENCES .......................................................................................... 176<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 148 of 220


Cement<br />

Concrete<br />

1. Composite (blended) cements<br />

1.1 Definition<br />

A Composite or Blended Cement is a binder that contains<br />

Portland clinker<br />

gypsum or other set regulators<br />

and one of, or a combination of, the following materials in relevant amount (normally<br />

>5%)<br />

a latent hydraulic component<br />

a pozzolanic component<br />

an inert component (filler)<br />

and is produced by grinding (separate or compound) or blending of the constituents.<br />

Typical materials for the production of composite cements are therefore<br />

Latent hydraulic<br />

Having natural hydraulic potential (they harden when mixed with water). A suitable activator<br />

is needed to accelerate the reaction with water (examples: blast furnace slag, class C fly<br />

ash)<br />

Pozzolanic<br />

Having no hydraulic properties. They react in presence of water and Ca(OH) 2 from clinker<br />

hydration, to form compounds contributing to strength (examples: natural pozzolana, class F<br />

fly ash)<br />

Inert<br />

Having no real participation in the chemical hydration process (examples: limestone, sand).<br />

Important to mention is that clinker hydration, slag hydration and reaction of pozzolana with<br />

lime cause the formation of the same type of compounds, i.e. calcium-silicate-hydrates<br />

possessing binding properties.<br />

For specific aspects regarding composition and properties of mineral components, please<br />

refer to the related chapter.<br />

1.1.1 Abbreviations<br />

In the following text and in relevant literature, acronyms and abbreviations may be used to<br />

indicate specific materials. These are the most important.<br />

mic<br />

mineral component<br />

(g)(g)bfs (ground)(granulated) blast-furnace slag<br />

(p)fa<br />

(c)sf<br />

(pulverised) fuel ash or fly ash<br />

(condensed) silica fume<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 149 of 220


Cement<br />

Concrete<br />

1.2 A brief history of composite cements<br />

The hydraulic properties of blends of pozzolana and lime are known from thousands of<br />

years, from the times of the ancient Greeks and Romans, whose constructions (“opus<br />

cementitium”) survived the centuries and can be still seen nowadays. At those times, natural<br />

pozzolana or crushed bricks were used as cement constituents together with lime, but the<br />

association of pozzolana with Portland cement only started at the beginning of the 1900.<br />

The practice of replacing a hydraulic material for Portland cement dates back to the end of<br />

the 19 th and the beginning of 20 th century.<br />

The discovery of the hydraulic properties of blast furnace slag was made in 1862 and the first<br />

use in Germany dates back to 1882.<br />

During the two World Wars the clinker shortage suggested the addition of finely ground<br />

limestone to Portland cement to increase the availability of cement. At that time, the<br />

replacement was considered not legal or at least questionable; however, the use of this<br />

cement did not give rise to any claims from the users owing to the good results obtained in<br />

terms of both strength and durability. In the sixties, at the end of long and accurate studies,<br />

limestone cement was definitively accepted among standard cements.<br />

An example of the many different composite cements is given in the following table, where<br />

the cement types according the European standard EN 197/1 are listed.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 150 of 220


Cement<br />

Concrete<br />

2. General properties of Composite cements<br />

Blended cements have properties very similar to those of Portland cements for two main<br />

reasons:<br />

all of them contain Portland cement clinker<br />

their products of hydration are identical to those occurring in Portland cement pastes both<br />

in terms of composition and microstructure.<br />

Increasing amounts of blending components gradually change any original property of the<br />

plain Portland cement. This replacement may either improve or worsen the properties of the<br />

parent Portland cement (clinker) as a consequence of the effect of mineral components and<br />

additions on chemical, physical and mechanical properties of cement, mortar and concrete.<br />

Typically, addition of most mineral components in relevant quantities reduces the strength,<br />

but increases the chemical resistance and the related durability of the cementitious products.<br />

Proper cement design may contribute to limit the undesired effects to the minimum and<br />

enhance the advantages of mineral components.<br />

In the following chapters, the impact of mineral components on the most important properties<br />

of blended cements is considered.<br />

2.1 Bleeding<br />

Bleeding occurs when the various constituents of a mix start to separate so letting the water<br />

rise to the surface of the fresh concrete. Whereas a certain amount of bleeding is favourable<br />

for adequate curing of concrete, an excess of bleeding water can cause undesired surface<br />

effects like efflorescence (crystallisation of calcium carbonate and alkali salts) and weaker<br />

surface strength.<br />

Bleeding depends on the grain size distribution and the specific surface of the cement and of<br />

the individual cement components. Therefore, natural pozzolanas, burnt clay, silica fume and<br />

limestone reduce bleeding, while fly ash and ground granulated blastfurnace slag tend to<br />

increase it.<br />

On the other hand, an increase of water demand in concrete due to the presence of highspecific<br />

surface materials, may have the contrary effect of increasing the porosity in the<br />

hardened dry structure, reducing the strength and increasing the permeability.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 151 of 220


Cement<br />

Concrete<br />

2.2 Workability<br />

As a general rule, workability of concrete decreases with the increase in water demand of the<br />

cement and therefore with the specific surface area of the components.<br />

As a consequence, workability of cements containing slag, fly ashes and limestone on<br />

average is higher than that of the parent Portland cement, whereas that of cements<br />

containing natural pozzolanas and silica fume is lower.<br />

However, within each cement type and cement class, the values of workability (slump test)<br />

are widely scattered owing to the large variability of both chemical and mineral composition<br />

and fineness of cements.<br />

Example: slump frequency distribution<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 152 of 220


Cement<br />

Concrete<br />

The different performance of fly ash and slag cements can also be explained by the following<br />

reasons<br />

little water absorption of the gbfs/fly ash particles<br />

smooth surface of the gbfs and fly ash particles.<br />

Problems may arise with very soft and porous pozzolanas, which tend to reduce workability.<br />

Example: compaction time for OPC/BFS cements<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 153 of 220


Cement<br />

Concrete<br />

2.3 Setting time<br />

The degree to which the setting time is affected depends on:<br />

the grinding fineness<br />

the type of clinker and mic used<br />

the mic content of the composite cement<br />

the initial curing temperature of the concrete<br />

the water/cement ratio of concrete<br />

Typically, the initial set is extended by half-an-hour to one hour at temperatures of 23° C /<br />

73° F. Above 29° C / 85° F no change in setting time occurs compared to OPC.<br />

Example: data from Products Handbook<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 154 of 220


Cement<br />

Concrete<br />

2.4 Strength<br />

On replacing a Portland cement with other fine materials, the strength development curve<br />

changes depending on the type and the amount of the addition.<br />

Clinker is the main responsible for early and medium term strength development, and<br />

therefore its substitution with less active constituents may lead to reduced strength. The<br />

difference from the plain Portland cement tends to disappear after about 28 days of curing,<br />

owing to the reaction of the mic’s and depending on their activity. In the case of silica fume, a<br />

very reactive material, strength recovery takes place earlier.<br />

However, when clinker replacement does not exceed 10%, early strength of cement may<br />

increase since fine additions tend to accelerate hydration.<br />

The ultimate strength can be higher than that of OPC when the<br />

cement contains other hydraulic components. In this case, after 28<br />

days, the strength development curve can intersect and surpass the<br />

curve of the parent Portland cement.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 155 of 220


Cement<br />

Concrete<br />

Strength and strength gain of concrete with blended cement depend on following factors:<br />

Fineness of the cement<br />

Mic and clinker reactivity<br />

Curing temperature<br />

Content of mic in the blended cement<br />

The drop in early strength should not always be considered as a negative effect since it is<br />

often associated with the improvement in other properties.<br />

In any case, the cement factories can remedy the strength loss by taking appropriate action.<br />

Such corrective actions may increase production costs, so the advantage of any solution<br />

must be assessed on a case-by-case basis.<br />

As a consequence, blended cements predominate in the cement classes having lower<br />

strength.<br />

This prevalence does not mean that it is impossible to produce rapid hardening and highearly<br />

strength composite cements, but that they are not often produced, at least in large<br />

quantities, since they are more expensive than Portland cement.<br />

Concrete strength depends on a number of factors, the most significant of which are:<br />

♦ cement strength,<br />

♦<br />

♦<br />

♦<br />

cement content,<br />

water/cement ratio,<br />

curing degree.<br />

The type of cement is not a primary factor provided that cements with the same<br />

strength and the same w/c ratio are compared.<br />

On the contrary, if concrete is compared on the basis of equal workability, some<br />

differences could occur depending on the different water demand.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 156 of 220


Cement<br />

Concrete<br />

2.5 Shrinkage<br />

Shrinkage of cementitious products depends on the initial water content of the mix.<br />

For this reason, shrinkage of standard mortars made with the same w/c ratio is not<br />

influenced by the cement type.<br />

In the practice, differences in shrinkage occur if comparison of different cement types is<br />

made on mortars or concrete having the same workability but different w/c ratio.<br />

Differences among cements, if any, decrease passing from paste to mortar and concrete, the<br />

shrinkage of concrete being about 25% of that of a mortar with the same w/c ratio.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 157 of 220


Cement<br />

Concrete<br />

In any case, the influence of the type and strength class of cement on shrinkage is noticeably<br />

less than that of the water/cement ratio of concrete and of the type of aggregate used<br />

therein.<br />

Beneficial effects from the use of fly ash and blast furnace slag have been noticed, in<br />

combination with adequate curing of the structure.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 158 of 220


Cement<br />

Concrete<br />

2.6 Creep<br />

Creep of concrete represents the deformation of concrete under load. It.depends on the<br />

water content of the mix and the modulus of elasticity of the aggregate.<br />

The creep value is linked to the strength of the concrete at the time of loading and so, all<br />

other conditions being equal, it depends on:<br />

♦<br />

♦<br />

♦<br />

the strength class of the cement,<br />

the water/cement ratio,<br />

the curing degree.<br />

Thus, if the same load is applied on concrete having the same strength, the cement type is<br />

not significant.<br />

Creep in blended cement concrete is higher than in Portland cement only when concrete is<br />

loaded too early.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 159 of 220


Cement<br />

Concrete<br />

High early creep may have a positive effect since, as an example, it allows massive<br />

structures to settle before the concrete becomes dangerously rigid.<br />

2.7 Heat of hydration<br />

Blending Portland cement with natural and artificial pozzolana, ground slag and calcareous<br />

filler of similar fineness lowers strength and heat of hydration.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 160 of 220


Cement<br />

Concrete<br />

However, if the composite cement attains the same strength as the plain Portland cement,<br />

thanks to finer grinding or because it contains particularly active components such as silica<br />

fume, its heat of hydration is the same as that of the plain Portland cement.<br />

In fact, there is a significant correlation specially between the early compressive strength of<br />

mortars and the heat of hydration of cement pastes after 24 and 48 hours of curing.<br />

An increasing interest for crack prevention in voluminous concrete elements leads to an<br />

increasing demand of blended cements in many countries. Applying these cements for mass<br />

concrete leads to economical benefits for the concrete producers as well as for the<br />

construction companies (no cooling system or chemical admixtures are needed).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 161 of 220


Cement<br />

Concrete<br />

2.8 Porosity<br />

Porosity is an intrinsic property of concrete, which may be limited but not completely<br />

eliminated.<br />

Porosity mostly results from the porosity in the cement paste and the paste/aggregate<br />

interface. It is responsible for many properties of concrete, from its mechanical features to its<br />

durability and directly affects strength, since pores reduce the cross sectional area of the<br />

structure.<br />

Several experimental works have assessed that hardened plain cement without pores would<br />

attain compressive strength higher than 700 MPa.<br />

Hardened Portland cement paste has a lower porosity than pozzolanic and blastfurnace slag<br />

cements and thus it should have higher strength. However, differences are little and they<br />

become negligible in concrete where the interfacial porosity is definitely higher than that of a<br />

bulk paste.<br />

Moreover, the interfacial zone of concrete made with blended cements is generally less<br />

porous, less permeable and poor in portlandite crystals. For this reason the interfacial bond<br />

is stronger, as is shown by the higher flexural strength.<br />

Porosity is also related to the following transport properties of concrete:<br />

permeability (entry of fluids under pressure),<br />

sorptivity (entry of fluids by capillary suction),<br />

diffusion (entry of gas and ions dissolved in the water).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 162 of 220


Cement<br />

Concrete<br />

These properties are not synonymous with porosity since they depend on the structure of the<br />

pores rather than on the total porosity of cementitious products<br />

As a matter of fact, hardened cement pastes containing ground blastfurnace slag or<br />

pozzolana have greater porosity but lower permeability and sorptivity than Portland cement<br />

pastes, due to the difference in pore diameters.<br />

2.9 Durability<br />

The improvement achieved with the use of active mineral components is direct consequence<br />

of:<br />

♦<br />

♦<br />

♦<br />

lower permeability of cement stone<br />

lower content of calcium hydroxide<br />

reduced leaching of calcium hydroxide<br />

Permeability of hardened concrete with blended cement is decreased due to the reduction of:<br />

♦<br />

♦<br />

the pore size in the hydrated cement gel<br />

the filling of capillaries by further hydration products<br />

The decreased permeability prevents the penetration and migration of damaging chemicals<br />

such as chloride and sulfates and leads to high durability of the concrete<br />

Actually, pores of pozzolanic cement pastes - which are larger than Portland cement ones -<br />

are connected by more segmented or finer pores.<br />

The differences in porosity and permeability of cement pastes as a result of the type of<br />

cement have, in any case, very little impact on the actual concrete where the porosity of the<br />

aggregate/cement paste interface is much higher than that of the bulk paste.<br />

When concrete is dense, i.e. strong, any difference in permeability becomes statistically very<br />

little and only low strength concrete shows a wide variability in permeability<br />

In the practice, when 28-day strength exceeds about 35 N/mm2, permeability does no longer<br />

significantly decrease.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 163 of 220


Cement<br />

Concrete<br />

2.10 Lime leaching<br />

Portlandite makes up some 20-22% of hardened Portland cement paste and is quite soluble<br />

in water. So, concrete that is permanently kept in contact with water will lose lime.<br />

Leaching of Ca(OH)2 makes porosity and permeability increase and strength and durability<br />

decrease.<br />

Cement pastes containing pozzolanic materials, as well as granulated blast furnace slag,<br />

release less lime than Portland cement. This is because they have a smaller Ca(OH) 2<br />

content and a greater C-S-H gel content that make paste generally less permeable.<br />

The effectiveness of hydraulic additions increases as their % content increases.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 164 of 220


Cement<br />

Concrete<br />

Leaching rate increases when waters contain aggressive carbon dioxide but blended<br />

cements resist better than Portland cement.<br />

Reduced hydrated lime availability decreases susceptibility to acids.<br />

2.11 Carbonation<br />

For some time, it had been generally held that pozzolanic cements and cements with a high<br />

slag content were less resistant to carbonation than Portland cements due to the lower<br />

Portlandite content in their pastes.<br />

However, a number of studies have established that there is no connection between the<br />

depth of carbonation and the type of cement, but that the depth of carbonation decreases<br />

with increasing strength, i.e. with decreasing porosity and permeability.<br />

2.12 Sulphate attack<br />

Waters and soils containing sulphate ions attack concrete through the formation of ettringite<br />

and, occasionally, gypsum.<br />

Preventing or reducing the consequences of the attack requires the following factors to be<br />

minimized:<br />

♦<br />

♦<br />

♦<br />

tricalcium aluminate (C3A) content in the clinker,<br />

Ca(OH) 2 content in the paste,<br />

permeability of concrete.<br />

Cements rich in ground granulated blastfurnace slag and pozzolanic cements satisfy the<br />

three criteria and thus resist well to sulphate attack.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 165 of 220


Cement<br />

Concrete<br />

Blended cements decrease the permeability of concrete but their high resistance to the<br />

sulphate attack is due also to other factors.<br />

In fact, by comparing Portland cement and fly ash cement pastes, both stored in a 0.7M<br />

Na 2 SO 4 solution for 12 weeks, it was found that, while their permeability was roughly the<br />

same, the moment corresponding to the onset of the first crack differed by one order of<br />

magnitude.<br />

Replacement of part of the sand in concrete with ground limestone reduces sulphate<br />

expansion in mortars. However, this result cannot be attributed to chemical reasons. It is due<br />

to the decrease in permeability caused by an increase in the fine particles of the mix<br />

2.13 Chloride attack<br />

Waters containing chlorides are detrimental to the concrete structure because they may<br />

cause or facilitate:<br />

♦<br />

♦<br />

♦<br />

leaching of free lime,<br />

break-up of concrete,<br />

corrosion of the reinforcement.<br />

Blended cements containing high percentages of natural pozzolanas, fly ash, silica fume and<br />

slag significantly reduce these risks since the resulting concrete shows:<br />

♦<br />

♦<br />

a lower portlandite content and<br />

a lower diffusion coefficient of the dissolved ions<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 166 of 220


Cement<br />

Concrete<br />

The compressive strength of pozzolanic cement or blastfurnace cement pastes is little<br />

affected by prolonged immersion in concentrated CaCl 2 solution, whereas the strength of the<br />

parent Portland cement already begins to drop after 7 days’ immersion.<br />

2.14 Sea water attack<br />

Seawater attacks concrete as a result of the penetration of chloride and sulphate. No<br />

concrete is strictly impermeable, thus any structure exposed to seawater for many years<br />

shows:<br />

♦<br />

♦<br />

decrease in CaO content,<br />

increase in MgO and SO 3 content.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 167 of 220


Cement<br />

Concrete<br />

The chemical attack is associated with:<br />

♦<br />

♦<br />

mass loss and<br />

expansion.<br />

Apart from the general requirement of high compactness, concrete attack can be reduced by<br />

using pozzolanic and blastfurnace cement since the resulting pastes have:<br />

♦<br />

♦<br />

a lower content of soluble CaO,<br />

a higher C-S-H content<br />

and this helps reduce:<br />

♦<br />

♦<br />

the water permeability and<br />

the diffusion coefficient of the aggressive ions.<br />

Marine structure built by the Romans with pozzolana concrete have kept a large part of<br />

original strength after about two thousand years.<br />

2.15 Alkali Silica Reaction<br />

Certain constituents of the aggregates can react with the alkali ions present in the pore<br />

solution of the cement paste giving rise to expansion and cracking of concrete and<br />

consequent loss of strength and durability of the structure.<br />

The majority of research studies have concerned the “alkali-silica reaction” resulting from<br />

certain forms of reactive silica in the aggregate.<br />

Harmful expansion can be prevented by using Portland cements with an equivalent Na 2 O<br />

content of less than 0.6%. However, modern cement technology and pollution-preventive<br />

regulations make this target too expensive.<br />

Blended cements containing considerable amounts of natural pozzolana, fly ashes, silica<br />

fume or blastfurnace slag give the same result. Alkalis react quickly with the amorphous<br />

silica of the fine ground mineral component and are no longer available for further reactions<br />

with the aggregates.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 168 of 220


Cement<br />

Concrete<br />

Limestone decreases expansion not because particular reactions occur but only because it<br />

makes the total alkali content in cement diminish.<br />

2.16 Frost resistance<br />

In many countries frost is a serious cause of damage of concrete. Freezing of free water<br />

present in the hardened cement paste results in an increase of volume, which causes the<br />

formation of stresses inside the concrete. If these exceed the flexural strength of concrete,<br />

cracks occur.<br />

High strength concrete resists better than low strength concrete, but the sole system that has<br />

proved to be effective in preventing the effects of frost is the entrainment of microscopic air<br />

bubbles inside concrete.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 169 of 220


Cement<br />

Concrete<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 170 of 220


Cement<br />

Concrete<br />

3. Aspects of Manufacture of Blended Cements<br />

For optimum production of blended cements, some important aspects related to the grinding<br />

process should be carefully evaluated<br />

♦<br />

♦<br />

♦<br />

♦<br />

pretreatment of additions<br />

<br />

<br />

drying<br />

preblending<br />

properties of the addition<br />

<br />

<br />

grindability<br />

strength activity<br />

type of grinding<br />

<br />

<br />

compound<br />

separate<br />

dosing of components.<br />

3.1 Pretreatment of additions<br />

The additive may have a considerable moisture content, and drying has to be done prior to or<br />

during the grinding process. The dosing facilities should be adequate to allow precise and<br />

trouble free operation even in case of very moist and sticky materials.<br />

Depending on the nature and the number of the additions used, pre-blending of the materials<br />

would be necessary to achieve sufficient homogeneity of the component and of the cement<br />

itself. Some mic’s are industrial by-products and therefore subject to changes in composition<br />

and physical state.<br />

3.2 Properties of the addition and type of grinding<br />

3.2.1 Grindability<br />

The grindability of the components is generally different, even in materials of the same type.<br />

Big differences might therefore be justified between different plants, and generally a specific<br />

solution has to be found for each plant.<br />

The following table summarizes some data, taken out of the annual report, giving information<br />

on physical properties and the grinding energy needed for the production of blended cements<br />

containing slag, pozzolan or fly ash compared to Portland cement.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 171 of 220


Cement<br />

Concrete<br />

Average Physical Properties and Grinding Energy for Blended Cements<br />

Material Density<br />

[g/cm 3 ]<br />

Blaine<br />

[cm 2 /g]<br />

Grinding<br />

Energy<br />

[kWh/t]<br />

Clinker 3.18 3135 32.43<br />

Slag 2.90 2980 34.17<br />

Nat. Pozzolan 2.55 6150 20.92<br />

Fly ash 2.45 4405 11.92<br />

This table shows that slag is usually harder to grind than Portland cement clinker, while<br />

pozzolanic materials are mostly softer. The advantage of fly ash is that it can be added<br />

directly to the separator feed, since most of it is already of the required cement fineness.<br />

This also implies that during grinding of a slag cement, clinker will be ground finer, thus<br />

contributing to early stregth development, while the slag remains coarser and will react lately.<br />

Thanks to its glassy nature, slag does not greatly affect the water demand and the<br />

workability characteristics and can be used at very high addition rates.<br />

In the case of pozzolanic cements, the very soft pozzolana will concentrate in the fine<br />

fraction of cement, thus affecting water demand, workability and early strength development.<br />

Proper selection of the grinding equipment (central discharge mills, vertical roller mills,<br />

separate grinding) will prevent overgrinding of pozzolanas and improve cement properties.<br />

High fineness of silica fume and/or important variability of the fineness of fly ash may be<br />

responsible for severe problems of handling, transporting and proportioning. This type of<br />

material is generally introduced after the mill, in the output bucklett or directly in the<br />

separator, in order to provide a good mixing, avoid quality fluctuations and improve mill<br />

operation.<br />

When designing a new grinding plant for blended cement manufacturing, the process can be<br />

optimized to the requirements. Unfortunately, in most plants blended cements have to be<br />

produced in old grinding equipment designed for Portland cement production.<br />

3.2.2 Strength activity<br />

A mineral component is clearly expected to contribute to strength development of cement. Its<br />

strength activity will reflect on cement design and on the proportioning system. Typical<br />

dosage of a mineral component ranges between 25 and 40%; other contents can be chosen<br />

depending on standards and target performance.<br />

3.2.3 Type of grinding<br />

As already mentioned, grindability of the various components can influence the choice of the<br />

grinding system, especially in new plants.<br />

Intergrinding of cement is easier from the point of view of plant design and operation, but<br />

requires accurate dosing and might be considerably disturbed by quality fluctuations of the<br />

different constituents, also in terms of grindability.<br />

The advantage of separate grinding is that each component can be ground at the optimum<br />

fineness, thus enhancing its peculiarities and reducing drawbacks usually associated with<br />

unproper grinding.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 172 of 220


Cement<br />

Concrete<br />

3.2.4 Use of chemical admixtures<br />

Some of the drawbacks of blended cements can be compensated by the use of selected<br />

chemical admixtures.<br />

Besides the simple addition of grinding aids especially in the case of high replacement rates<br />

with hard to grind materials, more sophisticated quality improvers can be applied<br />

♦<br />

♦<br />

♦<br />

water reducers/plasticizers<br />

early strength improvers<br />

multipurpose admixtures.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 173 of 220


Cement<br />

Concrete<br />

3.3 Proportioning<br />

The most important quality requirement of blended cements are usually related to their<br />

performance in the different applications.<br />

Some of these properties can be summarized as follows:<br />

♦<br />

♦<br />

♦<br />

♦<br />

♦<br />

♦<br />

♦<br />

Satisfactory rheological properties for good workability in concrete and mortar<br />

Acceptable setting behavior<br />

Adequate strength development and sufficiently high early and 28-day strength<br />

Satisfactory expectations for durability of the resulting concrete, as regards<br />

residual porosity and permeability<br />

resistance to carbonation<br />

freeze-thaw resistance<br />

resistance to chemical attack (sulfates, pure water, sea water, weak acids)<br />

Acceptance of resulting concrete color<br />

Reduction of Alkali Aggregate Reaction (AAR) risks<br />

Lower heat of hydration<br />

and all specific properties providing better service or higher performances compared to<br />

ordinary Portland cement.<br />

The design of composite cements requires first of all the establishment of clear objectives<br />

with respect to product characteristics and performances. Both characteristics and<br />

performances must comply to cement standard specifications and test methods as well as for<br />

concrete production.<br />

The optimum proportion of mineral additions in blended cements is then to be found as a<br />

consequence of the available materials and the related cement standards.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 174 of 220


Cement<br />

Concrete<br />

In general terms, slags are usually characterised by a good hydraulic potential and a low<br />

water requirement in cement, derived from their glassy nature; these features allow for the<br />

production of slag cements with high contents of BFS (up to 95%!).<br />

On the other hand, the amount of pozzolana is a compromise between the minimum amount<br />

necessary to combine the hydration lime and produce additional silicate hydrates and the<br />

maximum acceptable content due to the softness and high porosity of pozzolanas that yield<br />

particles with high water demand, thus affecting cement workability. In this case, a normal<br />

range for natural pozzolanas or activated clays in cement is 20-45%.<br />

As a rule of thumb, one shall consider that pure replacement of OPC with a mineral<br />

component will cause a loss in early strength of 1.5 times the amount of the addition. That is<br />

to say that 20% replacement will drop early strength by approximately 30%. This strength<br />

loss can be compensated to a great extent by finer grinding, either the cement as such in<br />

case of intergrinding, or the clinker fraction in case of separate grinding and following<br />

blending.<br />

Increase of grinding fineness will anyway reflect in higher energy consumption of the mill and<br />

could impair some of the cement characteristics, especially water demand. In this respect,<br />

the technology of chemical admixtures is capable to provide a selection of suitable products<br />

for the various needs. Through the proper use of additions, one can improve mill<br />

performance, increase early strength, reduce water demand, improve workability.<br />

4. Trends of production and marketing of composite cements<br />

The present and future challenges for the cement industry relate to its positioning with<br />

respect to the various environmental matters.<br />

On one side, it can help solve the problems related to an optimum reuse of industrial byproducts,<br />

as alternatives for raw meal, fuel and cement materials. On the other hand it will be<br />

pushed to reduce CO 2 emissions mostly deriving from the production of clinker.<br />

The consequences of these trends are<br />

• the limited ability to comply with specific requirements by pure OPC type cements,<br />

requiring clinkers of specific composition, which are not suitable in the production of other<br />

types of binders (example – low-alkali, low C 3 A, low-heat clinkers and cements)<br />

• the increased versatility in providing different performing cements, produced with the<br />

same clinker, but with different type and content of additions.<br />

The production of composite cements is therefore going to gain increasing importance in the<br />

next years, as well as the drafting of new standards where blended cement properties are<br />

adequately considered and specified.<br />

Rather than talking about cements, it will be necessary in the next years to consider binding<br />

agents and concrete mixes, in which Portland cement clinker will still be the principal active<br />

ingredient. Already today, concretes are being produced in which the clinker represents only<br />

the minor part of the binding agent, the rest of the constituents such as fly ash, silica fume,<br />

ground slag, limestone filler, etc. being added directly at the concrete batching plant.<br />

If we want to maintain the same quality of concrete and assure the future of our industry, we<br />

must increase drastically the proportion of additions in cement and master the production of<br />

binders with a maximised content of clinker substitutes. The possibilities for substitution are<br />

numerous and the optimum cement design will derive from the optimisation of<br />

♦ clinker composition and production<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 175 of 220


Cement<br />

Concrete<br />

♦<br />

♦<br />

♦<br />

♦<br />

optimum gypsum content<br />

mic type and content<br />

grinding process<br />

use of chemical admixtures.<br />

5. Conclusions<br />

The production of composite cements is based on technical, economical and environmental<br />

reasons.<br />

The use of ground granulated blastfurnace slag cements as well as pozzolanic cements has<br />

long been confined to those applications where resistance to chemical attack and lower heat<br />

of hydration were considered more important than strength.<br />

Later on, their production has been seen as an advantageous way to find a use for waste<br />

materials from certain industrial processes or as a way to save on energy.<br />

Also the Portland limestone cement, largely used in the past only in specialised fields, has<br />

proved to have useful properties like other more known cements.<br />

All these historical reasons have, however, been swept away by years of laboratory and field<br />

research which have shown that blended cements are interchangeable with Portland<br />

cements in the overwhelming majority of building applications.<br />

Only in special cases some type and classes of cement are specifically recommended.<br />

Typically, Portland cement is preferred whenever a rapid hardening is required, whereas<br />

blended cements, with high content of complementary hydraulic constituents, perform best in<br />

aggressive environment conditions.<br />

The real degree of replacement for these types of substitute materials will vary as a function<br />

of their individual reactivity. They can be mixed with other types of possible additions (triple<br />

blends or multiple blends or ‘cocktail’ cements), the clinker (or the activator) proportion<br />

depending on the quantity of CaO needed for the activation of the additions.<br />

6. Literature References<br />

1) Lea’s Chemistry of Cement and Concrete, 4th Edition - edited by P.C. Hewlett - Arnold<br />

2) Mineral Admixtures ic Cement and Concrete, vol. 4 - Sarkar, Ghosh - abi New Delhi<br />

3) Fly Ash in Concrete, 2nd edition - Malhotra, Rameziananpour - Canmet Canada<br />

4) Proceedings of the 9th Int. Congress on the Chemistry of Cement, vol. 1 - New Delhi<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 176 of 220


Cement<br />

Concrete<br />

Special Cements<br />

1. DEFINITION ...................................................................................................................... 178<br />

2. TYPES OF SPECIAL CEMENTS...................................................................................... 178<br />

2.1 ...Special Portland Cements for Durability ................................................................ 179<br />

2.1.1 Sulfate resisting cements ......................................................................... 179<br />

2.1.2 Low heat cements .................................................................................... 179<br />

2.1.3 Leaching resistant cements...................................................................... 180<br />

2.1.4 Low alkali cements................................................................................... 180<br />

2.1.5 Sea-water resisting cements.................................................................... 180<br />

2.1.6 Freeze-thaw resisting cements ................................................................ 180<br />

2.2 ...Composite Cements for Durability ......................................................................... 180<br />

2.3 ...High Early Strength Cement, Rapid Hardening Cement........................................ 181<br />

2.4 ...Fast Setting Cement (Vicat Cement) .................................................................... 181<br />

2.5 ...Sulfo-aluminate Cement......................................................................................... 182<br />

2.6 ...Regulated Set (Regset) Cement ............................................................................ 183<br />

2.7 ...Oil-well cements..................................................................................................... 183<br />

2.8 ...White cement ......................................................................................................... 186<br />

2.9 ...Hydrophobic cement .............................................................................................. 186<br />

2.10 Masonry Cements................................................................................................... 186<br />

2.11 Ultrafine Cements (Microcements) ......................................................................... 187<br />

2.12 High Alumina Cement............................................................................................. 188<br />

2.13 Phosphate cements ................................................................................................ 190<br />

2.14 Other non-Portland cements................................................................................... 191<br />

2.14.1 Alkali-activated cements........................................................................... 191<br />

2.14.2 Alkali-activated slags................................................................................ 191<br />

3. OVERVIEW OF PRODUCTION IMPLICATIONS IN SPECIAL CEMENTS ..................... 192<br />

4. LITERATURE.................................................................................................................... 193<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 177 of 220


Cement<br />

Concrete<br />

1. Definition<br />

Special cements are cements with special properties that meet particular requirements which<br />

are not fulfilled by ordinary cements.<br />

These properties refer either to the performance of cement in fresh and hardened concrete<br />

as well as in other cementitious blends, or to a special field of application. They can be<br />

produced with appropriate selection of clinker raw materials and/or cement constituents,<br />

adoption of special measures in manufacturing, tailored cement compositions.<br />

As a consequence of these actions, special cements can still comply with existing standards<br />

on common cements, but very often they are better described in appropriate specifications,<br />

or otherwise they are produced on the basis of specific agreements between producer and<br />

user.<br />

In spite of the several types listed in the following, their application is still limited in quantity<br />

and the total amount of marketed special cements can be estimated in the order of 10 to<br />

15% of production. Nevertheless the additional possibilities given by the recently issued new<br />

cement standards and the increased severity of the environments where concrete is put into<br />

service are going to make this quantity increase in the near future.<br />

2. Types of Special Cements<br />

Special cements are usually developed and produced to meet performance and durability<br />

requirements, in particular<br />

♦ improved strength development<br />

♦ increased resistance to chemical attack<br />

♦ improved compatibility with reactive aggregates<br />

♦ suitability for use at elevated temperatures and pressures<br />

♦ suitability for use in special applications<br />

♦ applicability in architectural purposes<br />

As already discussed in the chapter on special clinkers, special cements may belong to three<br />

main categories<br />

• Portland cement or composite Portland cement<br />

• Modified Portland cement<br />

• non-Portland cement.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 178 of 220


Cement<br />

Concrete<br />

2.1 Special Portland Cements for Durability<br />

Their main hydraulic constituent is generally Portland cement clinker, very often tailored to<br />

obtain special characteristics that yield the desired properties to the cement. For this reason,<br />

particular measures need to be taken in the production of clinker, expecially from the point of<br />

view of the selected raw materials to employ. In some cases this could not be enough or the<br />

involved costs are much higher than the obtained benefit.<br />

So, the same or even better final properties can be achieved by blending clinker with<br />

appropriate mineral components and/or additions, as described in chapter , dealing with<br />

design and properties of composite cements.<br />

Main durability characteristics for Portland type special cements are<br />

• resistance to sulfate attack<br />

• low heat of hydration<br />

• resistance to pure water attack<br />

• low alkali content or low reactivity with amorphous silica<br />

• resistance to freeze-thaw cycles.<br />

Before dealing with the different durability aspects that can require use of special cements, a<br />

general remark must be done. The main factor influencing durability is the proper design,<br />

production, compaction and curing of concrete. Any special cement, designed for the<br />

enhancement of durability will fail when used in the production of a poor concrete. A well<br />

compacted, high strength, low porosity concrete will be by far less prone to be attacked by<br />

external agents, even when produced with ordinary cements.<br />

2.1.1 Sulfate resisting cements<br />

Sulfates can be found in natural and industrial waters, as well as in soils. Soluble sulfates<br />

can react with lime and aluminates present in the hardened concrete and form respectively<br />

gypsum and ettringite. Both reactions entail expansion and consequent concrete<br />

deterioration. Sulfate resisting cements are characterized by a low C3A content, to minimize<br />

the risk of ettringite formation. This type of binders set and harden normally and in fully<br />

compacted concrete are not attacked by sulfates in a wide range of concentration. At the<br />

same time they possess low-heat properties.<br />

2.1.2 Low heat cements<br />

Hydration reactions of cement develop heat. When cement is used in the production of mass<br />

concrete, temperature gradients generate between core and surface of the conglomerate<br />

cause strains and may eventually lead to cracking. Low heat cements have a reduced heat<br />

of hydration, obtained by altering the chemical composition (low C3S and C3A contents). The<br />

use of low-heat cement is recommended for mass concrete production or for large structural<br />

sections. They usually set and harden at a lower rate than for ordinary cements, especially<br />

in cold weather, but ultimate strengths may be higher. They possess also good sulfate<br />

resisting properties.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 179 of 220


Cement<br />

Concrete<br />

2.1.3 Leaching resistant cements<br />

Waters with low salinity or a high content of carbon dioxide (CO2) are capable to dissolve<br />

hydration lime present in hardened concrete structures and may subsequently also subtract<br />

lime from silicate hydrates, thus causing damage to the hardened cement paste. Cements<br />

resisting to leaching have a low development of calcium hydroxide (lime) after hydration of<br />

clinker silicates, as a consequence of their low C3S content. Their use is suggested in<br />

hydraulic works like basins, river barriers and sides. At the same time they possess rather<br />

low heat evolution characteristics.<br />

2.1.4 Low alkali cements<br />

Some aggregates may contain forms of reactive amorphous silica. In the presence of water<br />

this can react with the soluble alkali of cement and form locally an expansive gel that can<br />

deteriorate concrete. In such cases, when the use of reactive aggregates is absolutely<br />

unavoidable, the use of a cement proven to counteract alkali-silica reaction (ASR) is<br />

suggested. Such cements can either be low-alkali or composite cements, since the<br />

pozzolanic or slag material can immediately react with alkali and prevent subsequent<br />

reaction with aggregates in the hardened paste.<br />

2.1.5 Sea-water resisting cements<br />

Deterioration of concrete which is in contact to sea water takes place as a consequence of<br />

some of the already described phenomena. Concurring factors are chemical attack by<br />

MgSO4, mechanical stress caused by tydal waves, crystallisation pressure due to deposition<br />

of salts in the wind and water line. Cold climates add freeze-thaw effects, while in warm<br />

areas some reactions are accelerated. Sea-water resisting cements are standardised in<br />

some countries, being moderate C3A content or composite cements.<br />

2.1.6 Freeze-thaw resisting cements<br />

In harsh climates with high temperature differences and repeated freeze-thaw cycles,<br />

freezing of water contained in gel and aggregate pores will cause a volume increase of ab.<br />

10% and the consequent development of internal pressure; repeated actions of this type<br />

deteriorate the concrete. Entrainment of air in form of microbubbles will produce a closed<br />

artificial porosity acting as expansion chambers for the ice generated. Some standards such<br />

as ASTM provide for cement types with air entrainment.<br />

2.2 Composite Cements for Durability<br />

In the chapters 2.1.1.to 2.1.5 specifications set limits for clinker components, i.e. maximum<br />

C3A and C3S content, maximum Na20 equivalent, as reflected in the ASTM C150<br />

specification for Portland cement.<br />

In the case of Composite cements, it would be more correct to refer to these values<br />

expressed as total content in cement, as the clinker fraction may vary from case to case. As<br />

a matter of fact, using less clinker to produce a composite cement is a good starting point to<br />

reduce the amount of the harmful clinker components in that cement.<br />

Additionally, the use of a mineral component like slag, pozzolan or fly ash will improve the<br />

durability and the pore structure of hardened concrete by lowering its lime content and<br />

porosity as a consequence of the enhanced development of calcium silicate hydrates.<br />

Then, the ASTM C595 and C1157 deal with composite cements and set limits referred to the<br />

performance of cement with respect of sulfate resistance and heat of hydration, without any<br />

prescription on the basis of cement composition.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 180 of 220


Cement<br />

Concrete<br />

If we look at these durability aspects in the context of the European standard EN 197-1, we<br />

can immediately realize that most of the Portland composite cements of the II/B type, with<br />

addition of 21 up to 35% of non-clinker materials, as well as type III (slag), type IV (pozzolan)<br />

and type V (composite) cements can be considered as “special” with respect to many<br />

properties mainly related to durability.<br />

Nevertheless, a CEN working group is now drafting new ENV 197 parts specifically dealing<br />

with special cements, in particular low heat- and sulfate resisting cements.<br />

2.3 High Early Strength Cement, Rapid Hardening Cement<br />

High early strength cement (HES) and rapid hardening cement (RHC) may either be finer<br />

than ordinary cement or have a special clinker composition.<br />

Especially in the past, production of HES/RHC cements often required the production of a<br />

special clinker, with high C3S and C3A contents.<br />

In present times, the improvements gained in grinding technology and in the use of quality<br />

enhancers as grinding aids, combined with the need of rationalization for storage and<br />

transports in cement plants, allow for the production of high early strength cements based on<br />

the same clinker used for ordinary binders. The higher grinding fineness and the related<br />

enhanced reactivity usually require increased gypsum dosage.<br />

Physical properties of HES/RHC cements in comparison with ordinary Portland cements are<br />

shown in table 1. It should be noted that an increase in fineness would often also reflect in<br />

higher late strengths.<br />

Table 1<br />

Physical Properties of High Early Strength and Ordinary Portland<br />

Cement<br />

Type Blaine (cm²/g) Initial setting Compr. strength MPa<br />

time (minutes) 2 days 28 days<br />

OPC 3000 ÷ 3500 150 ÷ 240 10 ÷ 20 38 ÷ 43<br />

High early strength 3800 ÷ 4500 120 ÷ 180 20 ÷ 30 48 ÷ 55<br />

Rapid hardening 4800 ÷ 5500 90 ÷ 150 30 ÷ 40 58 ÷ 67<br />

(ISO)<br />

The use of HES/RHC cements is indicated when a rapid strength development is required,<br />

e.g. if formwork has to be removed or re-used after a short time (precast elements<br />

production) or when sufficient strength is required for further construction (slipforming).<br />

However, since rapid strength gain is usually associated to high rate of heat devlopment,<br />

HES/RHC cements should not be used in mass concrete or in large structural sections. On<br />

the other hand, concreting at low temperatures would profit from the use of a high heat<br />

cement as a safeguard against early frost damage.<br />

2.4 Fast Setting Cement (Vicat Cement)<br />

A particular type of fast setting cement is the so called Vicat cement from the name of its<br />

inventor and main producer. It is obtained by burning at low temperature (1200 to 1250 °C)<br />

selected natural marls with rather high alumina contents. The clinker produced is mainly<br />

composed of belite and aluminates; after grinding it is capable to set in a few minutes,<br />

developing sufficient strength to fulfil requirements for easy and fast repair or small<br />

construction jobs.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 181 of 220


Cement<br />

Concrete<br />

“Artificial” fast setting cement can be produced by blending OPC and High Alumina cement<br />

(see § 2.12) approximately at a 9:1 ratio.<br />

2.5 Sulfo-aluminate Cement<br />

Cements having calcium sulfo-aluminate (C4A3S*) as main component. Examples of this<br />

binder type are type K cement produced in USA and the so called third cement series (TCS)<br />

in India and China. Typical composition ranges are reported in Table 4.<br />

Characteristics of these cements are<br />

∗ reduced setting times<br />

∗ high workability, low water demand<br />

∗ rapid strength development<br />

∗ low shrinkage (even expansive)<br />

Hydration of C4A3S* leads to the formation of ettringite, which is mainly responsible for early<br />

strength development. Microstructure of ettringite depends on the presence of lime. When<br />

ettringite forms in presence of lime it provokes expansion and this property is used to<br />

produce no-shrinkage cements. In the absence of lime, ettringite is not expansive and mainly<br />

contributes to strength development.<br />

Binders of this type can then be used, according to mineralogy, to prevent shrinkage in<br />

concrete or to develop high early strength for special applications.<br />

The blend of TCS cements with OPC yields to flash setting.<br />

Table 4<br />

Chemical and mineralogical composition of Sulfo-aluminate Cements<br />

Sulfo-aluminate<br />

clinker (type K)<br />

TCS sulfo-aluminate<br />

cement (SAC)<br />

SiO2 2 - 4 3 - 10 6 - 12<br />

Al2O3 45 - 49 28 - 40 25 - 30<br />

Fe2O3 1 - 2 1 - 3 5 - 12<br />

CaO 37 - 39 36 - 43 43 - 46<br />

SO3 7 - 10 8 - 15 5 - 10<br />

CaO free 0 - 0.3 0 - 0.3 0 - 0.3<br />

mineralogical composition<br />

C4A3S* 55 - 70 55 - 75 35 - 55<br />

CA 15 - 20 - -<br />

C2AS 15 - 20 - -<br />

C4AF 0 - 5 3 - 6 15 - 30<br />

C2S - 15 - 30 15 - 35<br />

TCS ferro-aluminate<br />

cement (FAC)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 182 of 220


Cement<br />

Concrete<br />

2.6 Regulated Set (Regset) Cement<br />

Modified Portland cement type with rapid setting and hardening characteristics. The active<br />

component is a calcium fluoroaluminate (11 CaO. 7Al2O3. CaF2).<br />

The regset cement is characterized by a setting time of 15 to 60 minutes. The setting is<br />

controlled by the addition of retarders (citric acid). Additional characteristic property is its<br />

unusually high early strength. A comparison between strengths achieved with Regset and<br />

HES cement is reported in Table 2.<br />

Table 2<br />

Properties of Regset and High Early Strength Cement<br />

EN method Regset Cement H.E.S. Cement<br />

Compr. Strength MPa<br />

2 hours 6 0<br />

8 hours 8 2<br />

16 hours 9 15<br />

2 days 10 35<br />

7 days 18 47<br />

28 days 35 60<br />

Setting time minutes<br />

Initial 45 120<br />

Final 75 180<br />

The very early strength gain makes this cement useful for all applications where there is a<br />

vital need for short times between placing and hardening of concrete, i.e. shotcrete, repairs<br />

on airport runways or highways, pavements.<br />

Regset cements are batched, handled and mixed in pretty well the same way as Portland<br />

cements, making however the necessary provisions for rapid handling and placing of<br />

concrete. Although production is as cheap and easy as for Portland cement, the fast and<br />

often unpredictable setting behaviour puts serious limits to the use in constructions and the<br />

total world production is estimated in the order of 10000 tons/year.<br />

2.7 Oil-well cements<br />

Oil-well cements are developed for use in oil and gas wells and are designed to set and cure<br />

at high temperatures and pressures in well grouting. They can also be used for sealing water<br />

wells, waste disposal wells and geothermal wells. Cement plays an important part in the<br />

successful drilling of a well. It is used primarily to seal the annulus between the walls of the<br />

borehole and the steel casing, to isolate the pressured or weak zones encountered whilst<br />

drilling.<br />

The oil-well cement must possess the following properties:<br />

∗ low permeability<br />

∗ form a good bond between rock and casing<br />

∗ maintain these properties under downhole temperature and pressure conditions<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 183 of 220


Cement<br />

Concrete<br />

∗<br />

protect the casing against corrosion and collapse<br />

To achieve these aims, the cement slurries must stay pumpable for sufficient time to permit<br />

placement, give stable suspensions, harden rapidly once in place and retain high strength<br />

and low permeability during well lifetime.<br />

API specification 10 from American Petroleum Institute classifies eight different oil-well<br />

cements. Other national standards exist, e.g. in Russia, China, India, most Eastern European<br />

countries, but so far no EN standard has been drafted. In Table 3 an overview of API 10<br />

cements is compiled.<br />

Table 3<br />

Class<br />

A<br />

B<br />

C<br />

D<br />

E<br />

F<br />

G and H<br />

J<br />

Overview of oil-well cements<br />

Typical use<br />

from surface to 1830 m, special properties not required, ASTM type I<br />

from surface to 1830 m, moderate or high sulfate resistance<br />

from surface to 1830 m, high early strength<br />

between 1830 and 3050 m, moderately high T and p conditions, moderate<br />

or high sulfate resistance<br />

between 3050 and 4270 m, high T and p conditions, moderate or high<br />

sulfate resistance<br />

between 3050 and 4880 m, extremely high T and p conditions, moderate<br />

or high sulfate resistance<br />

from surface to 2440 m, for use with accelerators or retarders, moderate<br />

or high sulfate resistance, class H coarser than G<br />

between 3660 and 4880 m, pure OPC, extremely high T and p conditions<br />

Such cements need special test procedures for suitable characterisation and additional care<br />

during manufacture so as to ensure consistency of quality between batches of the same<br />

plant.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 184 of 220


Cement<br />

Concrete<br />

In use, they are frequently mixed with additives in various proportions to produce satisfactory<br />

slurry performance for given well conditions, so they should additionally be compatible and<br />

responding to these additives (Table 4).<br />

Table 4<br />

Common types of additives for Oil-well Cements<br />

Classification Function Example<br />

Accelerator Reduces thickening time of cement CaCl2, sea-water<br />

Retarder Lengthens thickening time Lignosulfonates, sugars<br />

Dispersant Improves flow properties of cement Superplasticizers<br />

Lightweight extender Improves stability of suspension Bentonite, various clays<br />

Weighing agent Improves density of slurry Haematite, barite, sand<br />

Lost<br />

controller<br />

circulation<br />

Prevents cement losses through strata<br />

Walnut shells, cellophane<br />

flakes, expanded clay<br />

Inhibitor of strength<br />

retrogression<br />

Prevents loss of strength and<br />

formation of low strength silicate hydr.<br />

Silica flour, silica sand<br />

Fluid loss controller Controls rate of water loss polymers, cellulose<br />

Defoamers Removes foaming during mixing Lauryl alcohol, glycols<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 185 of 220


Cement<br />

Concrete<br />

2.8 White cement<br />

One of the most important special Portland cements. It is characterized by a white colour,<br />

obtained with suitable selection of raw materials, in which the colouring elements iron,<br />

chromium and manganese must be kept at the lowest possible level. Reduced formation of<br />

melt phase during burning is sometimes compensated by addition of fluoride as mineralizer.<br />

Mechanical properties of white cements are comparable to those of OPC. Lower strength<br />

binders are produced intergrinding (up to 35%) high purity limestone. To improve whiteness,<br />

use of “optical whiteners” such as methylene blue in cement grinding can be applied.<br />

White cement is used for architectural purposes in white or coloured concrete. To achieve<br />

best results, a properly coloured aggregate has to be used.<br />

2.9 Hydrophobic cement<br />

A small amount of a water-repellent agent (stearic acid, oleic acid, a.s.o.) is added to<br />

Portland cement during grinding. This forms a protective coating around each cement<br />

particle that retards hydration until cement is mixed with water and prevents deterioration<br />

during storage, especially in humid countries.<br />

2.10 Masonry Cements<br />

Masonry cements are successfully used in the production of mortars and plasters for most<br />

non-structural building purposes.<br />

Their main characteristics are: low early- and late strengths, low shrinkage, limited water<br />

permeability, high water retention, excellent plasticity and cohesiveness in the fresh state,<br />

frequently also air entrainment for better workability and freeze-thaw resistance.<br />

These properties are achieved by a low to medium clinker content, use of a suitable<br />

limestone and/or other mineral additions, use of air-entraining agents.<br />

MC’s are mostly produced by intergrinding Portland cement clinker, gypsum, precrushed<br />

limestone and/or other mineral additions. When the grindability of the constituent materials is<br />

much different, it might be opportune to grind them separately, especially if the following<br />

blending facilities are available.<br />

The main advantage from the production point of view is the limited clinker content (may be<br />

as low as 25%) and the availability of the limestone in the plant. Production<br />

25 Holcim Group plants produce masonry cements in an extremely wide range of<br />

compositions and properties, reported in the following table 5 (based on 1996 Annual<br />

<strong>Technical</strong> Report).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 186 of 220


Cement<br />

Concrete<br />

Table 5<br />

Characteristics of Masonry Cements in the Holcim Group<br />

Masonry Cement Characteristics avg. min. max.<br />

Clinker content, %<br />

1)<br />

54.2 24.0 81.5<br />

Specific surface acc. to Blaine, cm²/g 6470 3765 9420<br />

Residue on a 45-µm sieve, % 8.9 1.7 25.7<br />

Residue on a 90-µm sieve, % 5.5 1.1 9.5<br />

Air entraining agent, dosage g/t 1210 91 4000<br />

Sulfate content, % SO3 1.93 0.62 3.00<br />

Water demand, %<br />

2)<br />

26.3 21.5 30.2<br />

Initial setting time, minutes 2) 190 113 383<br />

Compressive strength 7 days, MPa<br />

Compressive strength 28 days, MPa<br />

2)<br />

2)<br />

12.8 3.7 25.6<br />

15.2 3.4 29.5<br />

Specific milling energy consumption, kWh/t 65.6 34.3 132<br />

1)<br />

main other constituent is limestone, but also fly ash and pozzolana are used<br />

2)<br />

figures related to both ASTM and EN test methods are considered<br />

2.11 Ultrafine Cements (Microcements)<br />

These binders are characterized by a narrow and steep particle size distribution (see figure<br />

1), specially designed for injection grouts used for sealing and improving mechanical<br />

properties of porous systems (rocks, damaged concretes, soils).<br />

Due to the reduced particle size, water-microcement suspensions are highly stable and<br />

penetrating than OPC grouts. Results obtained using microcement injections are comparable<br />

to those of chemical products such as resins and soluble silicates, with the advantage of<br />

being more environment friendly.<br />

Microcements are usually Portland or blast furnace slag cement based and are produced by<br />

ultrafine grinding or subsequent efficient separation of finer fractions from common cements.<br />

Particle size distribution may range between 0 and 10 -15 µm. Due to this, they also develop<br />

much higher early strength than the corresponding common cement , but usually require<br />

higher water demand. This is not a drawback in most applications, since for injection grouts a<br />

W/C ratio of 1.5 to 3 is quite in the normal range.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 187 of 220


Cement<br />

Concrete<br />

Figure 1<br />

Particle size distribution of various cement types<br />

100<br />

Particle size distribution of various cement types<br />

%<br />

pa<br />

ssi<br />

ns<br />

80<br />

60<br />

40<br />

Microcement<br />

HES<br />

OPC<br />

20<br />

0<br />

1 1.5 2 3 5 10 15 30 50 75 100 150<br />

sieve opening (µm)<br />

As mentioned, their main use is in fields where common cements cannot be used due to their<br />

limited fineness characteristics<br />

• agglomeration of loose soils<br />

• sealing of microcracked rocks<br />

• restoration of foundations, tunnels, leaching dams, historical buildings.<br />

2.12 High Alumina Cement<br />

Typical non-Portland cement, high-alumina cement (HAC) produces a concrete which has an<br />

exceptionally fast rate of hardening and is resistant to attacks by most sulfate solutions. It<br />

also has a higher resistance against acidic solutions than Portland cement based concretes,<br />

but it does not resist the attack of caustic alkali.The strength development of HAC is<br />

demonstrated in Figure 2. About 80% of its ultimate strength is achieved after only 24 hours.<br />

The high rate of strength gain of HAC is due to the rapid hydration of the anhydrous calcium<br />

aluminate (CA).<br />

Rapid hardening is not accompanied by fast setting. Initial setting takes place between two<br />

and six hours after mixing, becuase of the slow setting pattern of hydration of the main<br />

compound CA.<br />

Setting time can be shortened by addition of OPC. HAC/OPC blends are used in applications<br />

where rapid setting is compulsory, but lower ultimate strengths are then obtained.<br />

Inversely, the same effect on setting is achieved by replacement of 10 to 20% HAC to OPC.<br />

The high rate of heat evolution (Figure 3) of high-alumina cement (ab. 40 J/g.hour in the first<br />

24 hours) makes it necessary for HAC concrete to be placed in thin sections and never in<br />

large mass. The rise in temperature causes cracking and adversely affects strength. HAC<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 188 of 220


Cement<br />

Concrete<br />

concrete needs also more water for hardening and better curing during the first two days<br />

than OPC concrete.<br />

The main drawback of high-alumina cement concrete is the loss of strength associated with<br />

expansion due to the conversion of the aluminate hydrates under moist conditions at<br />

elevated temperatures. The hexagonal hydrate CAH10 is converted into the cubic aluminate<br />

C3AH6 as from the reaction<br />

3 CAH10 → C3AH6 + 2 AH3 + 18 H<br />

This means that concrete which is properly placed and has developed a high strength will<br />

lose a considerable proportion of its strength upon exposure to temperatures over 30 °C and<br />

moisture.<br />

Figure 2<br />

Strength development of different cements<br />

Compressive strength [N/mm²]<br />

50<br />

40<br />

30<br />

20<br />

10<br />

HAC (High alumina cement)<br />

Type III<br />

Regset<br />

Type I<br />

0<br />

15 1 5 1 3 7<br />

28<br />

minutes<br />

hours<br />

Age<br />

days<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 189 of 220


Cement<br />

Concrete<br />

Figure 3<br />

Heat evolution of different cements<br />

50<br />

Aluminous cement<br />

Rise in temperature [°C]<br />

40<br />

30<br />

20<br />

10<br />

Rapid hardening<br />

Portland cement<br />

Normal Portland cement<br />

0<br />

0 10 20 30 40 50 60 70<br />

Time [hours]<br />

2.13 Phosphate cements<br />

A mixture of tetracalcium phosphate and dicalcium phosphate, when mixed with a dilute<br />

phosphoric acid or other aqueous solutions will harden like a cement-producing<br />

hydroxyapatite, Ca10(PO4)6HOH, as the final product. The setting property, combined with<br />

biocompatibility, makes calcium phosphate cements useful in many applications in dentistry<br />

and medicine.<br />

Pastes prepared from diammonium orthophosphate, as well as NaH2PO4 or Napolyphosphate,<br />

and calcined MgO exhibit a fast setting and hardening at room temperature<br />

associated with NH3 liberation. MgO-phosphate binders may be used for high temperature<br />

applications, since strength remains preserved up to 1000 °C.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 190 of 220


Cement<br />

Concrete<br />

2.14 Other non-Portland cements<br />

2.14.1 Alkali-activated cements<br />

“Pyrament “ (developed in USA) belongs to this type of binders; they are mainly composed of<br />

pozzolans and possibly other active silica-based materials such as silica fume, fly-ash, slag<br />

or Portland cement, and alkali compounds (alkali silicate, hydroxide or carbonate); water<br />

reducers and retarders are also added to set regulation. They display excellent early strength<br />

properties as well as high density and strength at later age, but the complicated<br />

multicomponent formulation indicates a very sensitive system, difficult to control under field<br />

conditions.<br />

Development of “Geopolymers” is based on the work of Davidovits. The main reaction takes<br />

place between sodium hydroxide and kaolinite to form hydrosodalite at room temperatures or<br />

preferably between 150 and 180 °C. This reaction is the typical hydrothermal synthesis of<br />

zeolites.<br />

2.14.2 Alkali-activated slags<br />

The ability of activating blast furnace slags (BFS) by the addition of alkalis has been known<br />

for many decades. Hydration of slag requires the breaking of bonds and dissolution of the<br />

three-dimensional structure of glass and this is easily achieved in the high pH environments<br />

produced by alkali. Research on slag activation dates back to the early 50’s; interest in<br />

alkaline activation has grown markedly and in recent years alkali-activated cement and<br />

concrete have received greater attention worldwide.<br />

Compared with ordinary Portland cement and interground slag cement, alkali activated slag<br />

cement has some advantageous properties, including rapid and high strength development,<br />

good durability and high resistance to chemical attack.<br />

Finely ground, well granulated BFS can be utilized in the production of cements suitable for<br />

the precast industry, after addition of alkaline activators and superplasticizers, as for the “F”<br />

cement (Finland). When used with thermal curing, these cements can develop strengths 30%<br />

higher than normal Portland cements.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 191 of 220


Cement<br />

Concrete<br />

3. Overview of Production Implications in Special Cements<br />

In the following table 5, the implications related to the production of special cements are<br />

summarized. In particular, after highlighting the main properties and applications of each<br />

special cement type, a list of special measures that need to be taken at the plant to achieve<br />

the expected properties is compiled.<br />

Table 4<br />

Properties , application and measures for proper production of special<br />

cements<br />

Type Properties Application<br />

Raw<br />

mat<br />

Man<br />

ufact<br />

Addi<br />

tions<br />

App<br />

licat<br />

ion<br />

OPC normal workability<br />

and stength<br />

general application in<br />

building and construction<br />

- - - -<br />

ASTM II /<br />

SRC/LHC<br />

moderate sulfate res.<br />

low heat of hydration<br />

drainage, large piers and<br />

retaining walls, sea water<br />

x - (x) -<br />

ASTM III /<br />

HES-RHC<br />

high early strength,<br />

rapid hardening<br />

precast concrete, repairs,<br />

cold weather concreting<br />

- x<br />

finer<br />

- -<br />

ASTM IV /<br />

LHC<br />

low heat of hydration mass concrete, large dams x - x x<br />

ASTM V /<br />

SRC<br />

sulfate resistant<br />

severe sulfate action in<br />

(soils, ground water)<br />

x (x) x x<br />

Air entrain. frost resistance roads, freeze-thaw action - - x -<br />

Low alkali low reactivity with<br />

amorphous silica<br />

in concrete with reactive<br />

aggregates<br />

x x x -<br />

Leaching<br />

resisting<br />

low content of hydrat.<br />

lime<br />

in presence of pure waters<br />

or high CO2 solutions<br />

x - x -<br />

Oil well retarded setting,<br />

moderate sulfate res.<br />

in oil, gas and other types<br />

of wells<br />

x - (x) x<br />

Regset fast setting time road repairs, dry mortars x x - x<br />

White cem white colour architectural concrete x xx (x) x<br />

Sulfoalum.<br />

cement<br />

fast setting, high<br />

workab., (expansive)<br />

precast concrete, repairs,<br />

no-shrinkage concrete<br />

x x - x<br />

Ultrafine max size < 15 µm injection grouts - xx (x) xx<br />

High alum.<br />

cement<br />

high early strength,<br />

rapid strength devel.<br />

demand of very high early<br />

strength, refractories<br />

x xx - xx<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 192 of 220


Cement<br />

Concrete<br />

4. Literature<br />

1) Special and New Cements - A K Chatterjee Proceedings of the 9th ICCC, New Delhi<br />

1992<br />

2) Development in non-Portland Cements - Nuzhen, Kurdowski, Sorrentino as above<br />

3) Mineral Admixtures in Cement and Concrete - Sarkar, Ghosh Progress in Cement and<br />

Concrete vol. 4 - Akademia Books International, New Delhi 1993<br />

4) Specialty Cements with Advanced Properties - Scheetz, Landers, Odler, Jennings<br />

Symposium Proceedings Materials Research Society 1989<br />

5) International Development Trends in Low-energy Cements - Stark, Müller ZKG 4/88<br />

6) J. Bensted - Oilwell cements World Cement 10/89<br />

7) Structure and Performance of Cements - ed.Barnes Applied Science Publishers 1983<br />

8) New Ultra-rapid Cements - Costa Proceedings of the FAST Congress, Milan 1997<br />

9) Special Inorganic Cements - I. Odler E&FN SPON, London 2000<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 193 of 220


Cement<br />

Concrete<br />

Cement Standards<br />

1. INTRODUCTION ............................................................................................................... 195<br />

1.1 ...Importance of Standards........................................................................................ 195<br />

1.2 ...Development of Standards..................................................................................... 195<br />

2. CEMENT STANDARDS.................................................................................................... 196<br />

2.1 ...Product Standards ................................................................................................. 196<br />

2.2 ...Significance of Specifications................................................................................. 197<br />

2.3 ...Testing Methods..................................................................................................... 198<br />

3. INTERNATIONAL CEMENT STANDARDS ..................................................................... 199<br />

3.1 ...The ASTM Standards............................................................................................. 199<br />

3.2 ...THE EUROPEAN STANDARD ENV 197-1 ........................................................... 204<br />

3.3 ...THE AUSTRALIAN STANDARD AS 3972-1991.................................................... 209<br />

3.4 ...Conclusion ............................................................................................................. 211<br />

4. STANDARDS FOR MINERAL COMPONENTS ............................................................... 211<br />

5. CONCRETE STANDARDS............................................................................................... 212<br />

5.1 ...Purpose.................................................................................................................. 212<br />

5.2 ...Content of the Standards and Specifications ......................................................... 212<br />

5.2.1 Quality of the Concrete Components ....................................................... 212<br />

5.2.2 Quality of Fresh Concrete ........................................................................ 213<br />

5.2.3 Strength of Hardened Concrete ............................................................... 213<br />

5.2.4 Durability of Concrete............................................................................... 214<br />

5.3 ...The European Standard ENV 206 ......................................................................... 215<br />

5.3.1 Cement Content and Durability ................................................................ 215<br />

5.3.2 Chloride Content ...................................................................................... 219<br />

5.3.3 Alkali Aggregate Reaction........................................................................ 219<br />

5.3.4 Maintenance............................................................................................. 219<br />

5.4 ...Referenced documents .......................................................................................... 220<br />

5.4.1 Relevant ASTM Standards for Concrete Components............................. 220<br />

5.4.2 Relevant EN Standards for Concrete Components.................................. 220<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 194 of 220


Cement<br />

Concrete<br />

1. Introduction<br />

1.1 Importance of Standards<br />

In many of our daily activities we are unconsciously confronted with standards.<br />

When we go to see a football game at 3.30 p.m. we seldom realize that if it really begins at<br />

the scheduled time it is simply because we have standardized our time according to fixed<br />

rules. Then, it is up to us to decide buying an expensive Swiss watch in case we look for a<br />

reliable product, proven to keep its conformity to the time standard long enough.<br />

Many other less trivial examples could be reported to witness to what extent our life is<br />

defined and conditioned by standards of any type (mainly moral...).<br />

We need standards because they are the way we have to fix generally valid and accepted<br />

rules to measure and compare things with one another and so define the relationship<br />

between “producer” and “consumer”.<br />

Producing according to an official standard means then putting on the market a product of<br />

well defined, measurable and recognizable characteristics.<br />

This will positively affect the producer’s position with respect to own customers in terms of<br />

reliability, confidence and commercial relationship and will protect him against unfair<br />

competition and undue complaints.<br />

The consumer is then assured that by using a standardized product he can rely on a marketconforming,<br />

well defined and reasonably constant quality product, that will reflect in a<br />

consistent and cost-effective production.<br />

The designer and contractor will be provided with confidence about the essential material<br />

characteristics they need to design and execute the desired structures.<br />

It is therefore of vital importance that cement and concrete producers be aware of the<br />

expected developments of own standards and, if needed, take the necessary actions to<br />

influence development in the desired direction, be it through either direct or own<br />

association’s involvement.<br />

1.2 Development of Standards<br />

Developing a standard is not an easy deal. From previous comments, it should be clear that<br />

a product standard involves both the producer’ and the consumer’s interests and this is not<br />

an ideal starting point to find rapidly an overall agreement on the standard objectives.<br />

On the other hand, standards can meet easy and widespread acceptance only when all parts<br />

involved participated in its development.<br />

Standards shall then be a common elaborate of<br />

• the producer<br />

• the seller<br />

• the consumer<br />

• the designer<br />

• the official testing institute<br />

• the universities or research laboratories<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 195 of 220


Cement<br />

Concrete<br />

• the government.<br />

They should also be conceived (and periodically revised) taking into account<br />

• the present technical situation of the industry<br />

• the need of progress for the country<br />

• an effective environmental compatibility (energy saving and by-products recycling)<br />

• all technical improvements and economic changes that took place from the last<br />

edition<br />

• modified market exigences and demands<br />

• last but not least, the effort to pull down trade barriers in large State federations.<br />

2. Cement Standards<br />

2.1 Product Standards<br />

Product standards, which we are mostly interested in, are usually meant to define<br />

♦ the quality of a product, in terms of composition, performance and application (i.e. the<br />

European ENV 197-1 standard on common cements)<br />

♦ the common rules for uniformly testing and assessing the prescribed product quality (i.e.<br />

the associated EN 196 series of testing methods)<br />

♦ the conditions for the use of the product in particular applications (i.e. recommendations<br />

for durability of concrete structures, a.s.o.).<br />

They include classification, composition, specification and compliance criteria. As regards<br />

specification, that is a statement of one or a set of requirements the material shall comply<br />

with, they can be prescriptive- or performance oriented.<br />

The prescriptive specification defines a product according to its composition, the<br />

performance specification defines a product according to its function in application, thus<br />

products of different composition can comply.<br />

It is rather difficult to say which is better. On one hand a performance standard is supposed<br />

to be tightly related to the product’s field performance and should better represent its ability<br />

to fulfill the customer’s expectations and needs. On the other hand, assessing particular<br />

properties can involve use of complicated and time consuming testing methods; sometimes<br />

even the correlation between lab testing and field performance is still to be demonstrated.<br />

Then, in many cases, an easy prescriptive standard (i.e. maximum C 3 A for sulfate<br />

resistance) is preferred.<br />

We should not forget anyway, that some of the specifications for cement rely on performance<br />

testing, such as setting time and strength.<br />

In the case of cement, product specifications take into account<br />

♦ production aspects, as regards raw materials availability, plant design, environmental<br />

needs<br />

♦ application aspects, as regards experience of use, available technology, needs of<br />

differentiation, customer expectations<br />

♦ marketing aspects, as regards cost effectiveness of products, principles for fair<br />

competition, appropriate price/quality differentiation<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 196 of 220


Cement<br />

Concrete<br />

♦ quality aspects, by setting minimum and/or maximum acceptable values for most of the<br />

standardized properties, as a protection against dangerous behaviour of non-conforming<br />

products.<br />

Some of the most recent international standards (e.g. ENV 197-1 on common cements) now<br />

include also a chapter on conformity evaluation.<br />

This is a very innovative and clever approach to product standardization, since it introduces<br />

statistical tools to evaluate quality, not simply in terms of fulfilling absolute limit values, that<br />

are still considered as threshold values for conformity, but mainly taking into account the<br />

characteristic values, that include calculation of mean value and standard deviation of the<br />

measured property. An example will be given in the following pages.<br />

After highlighting the merits of standards, we cannot help mentioning their main drawback.<br />

Product standards are capable to depict the present situation and technology only, without<br />

being able to forecast, and open paths for, technical development. This sometimes reflects in<br />

a serious hurdle to innovation, since a real new product will hardly find a corresponding<br />

standard to be referred to.<br />

This is a strong argument for those people who are in favour of performance standards, that<br />

hinder technical progress much less than prescriptive ones.<br />

Standardization bodies should therefore find the right balance and make their mind more<br />

progress-oriented, so to walk aside innovation and not run after it.<br />

2.2 Significance of Specifications<br />

We have seen that a specification is a requirement the product shall comply with. But why<br />

are these requirements specified?<br />

The main concern in a cement standard is that the binder has to be used in constructions<br />

where safety, comfort and durability must be safeguarded. Additionally, special applications<br />

need special cement properties that have to be adequately described in specifications.<br />

In the following, a series of examples indicate properties that are specified for cements and<br />

also the inconveniences they must prevent; as mentioned, requirements can be related<br />

either to cement composition or to performance.<br />

• minimum setting time - to prevent early or flash setting of cement, that would<br />

excessively shorten the life time of the fresh and workable blend;<br />

• soundness (Le Chatelier or autoclave test) - maximum expansion is specified to<br />

prevent the risk of cracking in hardened structures;<br />

• maximum SO3 content - same as above;<br />

• maximum C3A content - to limit reactivity to sulfate attack and heat development of<br />

cement;<br />

• maximum C3S content - to limit reactivity to pure water attack and heat development<br />

of cement;<br />

• maximum heat of hydration - to prevent excessive heat development;<br />

• compressive strength - to guarantee a minimum strength development in properly<br />

designed concretes and cementitious blends;<br />

and so on.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 197 of 220


Cement<br />

Concrete<br />

2.3 Testing Methods<br />

Product standards usually deal with the characteristics of a product, making reference to the<br />

proper testing methods to be used for the assessment of these characteristics.<br />

Standards about testing methods aim at defining uniform conditions under which the product<br />

is tested.<br />

They fix the product’s testing conditions, in the case of cement if the measurement of the<br />

property will be carried out on<br />

• powder<br />

• paste<br />

• mortar<br />

• concrete;<br />

they define<br />

• scope of testing<br />

• referenced documents<br />

• testing apparatus<br />

• reactants<br />

• calibration of apparatus<br />

• testing procedure<br />

• calculation of results<br />

• reporting<br />

• precision and bias of the method.<br />

As for product specifications, the purpose is to have a common tool to evaluate the product<br />

and reproduce the results obtained by different testers with reasonably narrow deviation. So,<br />

a product with a wide (even transnational) market distribution can be checked in different<br />

places with sufficient confidence on the accuracy and reliability of results.<br />

This is clearly assuming a growing importance in the view of free trade activities, but the<br />

most practical advantage for a plant technologist is to use the same language and the same<br />

criteria to identify products and properties.<br />

Drafting testing methods is usually easier than for product specifications, because it does not<br />

involve commercial and political issues, nevertheless some main principles should be<br />

observed. A testing method should be<br />

♦ reliable (repeatable and reproducible) and meaningful<br />

♦ able to allow easy correlations between results and product characteristics and/or<br />

properties<br />

♦ easy to be implemented in any official as well as plant laboratory<br />

♦ as far as possible not hazardous to workers or environment.<br />

Most of the main testing methods for cement are well established standards. If a new<br />

property needs to be measured, then a working group is usually formed and a draft test is<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 198 of 220


Cement<br />

Concrete<br />

issued; the following round-robin testing will prove the effectiveness of the method to<br />

transform it into an official standard.<br />

3. International Cement Standards<br />

Standards are usually issued on a National basis by the local standardization institute.<br />

Such institutes are ASTM for USA, BSI for UK, DIN for Germany, AFNOR for France, and so<br />

on.<br />

Recent formation of large State federations (i.e. the European Community) have brought to a<br />

centralized management of standardization, to make it compatible with the free trade of<br />

products among countries. The European Committee for standards (CEN) has then rapidly<br />

become the main technical reference for all member and affiliated countries in Europe and<br />

also in some other areas of the world.<br />

In the context of international standardization of testing methods, even ISO, the International<br />

Standards Organization, has simply and directly adopted some of the new documents<br />

developed by CEN technical Committees.<br />

As a consequence, the trend all over the world is towards a reduction of the number of<br />

standards to a selected minimum, even maintaining some national peculiarities. Therefore<br />

we are going to mention here the cement standards that gained worldwide recognition and<br />

application, the ASTM and the EN specifications, as well as the Australian standard, an<br />

example of a very simple and effective way to draft a technical specification for cement.<br />

People that are interested in a complete compilation of the world cement standards can<br />

consult the publication “Cement Standards of the World”, issued by Cembureau, Brussels.<br />

3.1 The ASTM Standards<br />

ASTM standards are widely adopted not only in the United States of America, but also in<br />

many other countries of Latin America, as well as Asia and Africa.<br />

In the following pages, the specifications contained in the relevant ASTM cement standards<br />

are reported<br />

♦ C 150 - “Standard Specification for Portland Cement”<br />

♦ C 595M - “Standard Specification for Blended Hydraulic Cements”<br />

♦ C 1157M - “Standard Performance Specification for Blended Hydraulic Cements”.<br />

a) Cements defined by C 150 are divided in five different types.<br />

• Type I For use in general concrete construction<br />

• Type II For use when exposed to moderate sulfate action, or where<br />

moderate heat of hydration is required<br />

• Type III For use when high early strength is required<br />

• Type IV For use when low heat of hydration is required<br />

• Type V For use when high sulfate resistance is required<br />

(Types I to III can also be produced as air-entrained cements, when air entrainment is<br />

desired; in this case they bear the suffix “-A”).<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 199 of 220


Cement<br />

Concrete<br />

Table 1<br />

ASTM C150 - Standard Specification for Portland Cement<br />

Cement type → I I A II II A III III A IV V<br />

SiO2, min % - 20.0 - - -<br />

Al2O3, max % - 6.0 - - -<br />

Fe2O3, max % - 6.0 - 6.5 -<br />

MgO, max % 6.0 6.0 6.0 6.0 6.0<br />

SO3, max %<br />

when C3A8% 3.5 not applicable 4.5 n.ap. n.ap.<br />

Loss on ignition 3.0 3.0 3.0 2.5 3.0<br />

Insoluble resid. 0.75 0.75 0.75 0.75 0.75<br />

C3S, max % - - - 35 -<br />

C2S, min % - - - 40 -<br />

C3A, max % - 8 - 7 5<br />

C4AF+C3A, - - - - 25<br />

max %<br />

% air in mortar<br />

min - 16 - 16 - 16 - -<br />

max 12 22 12 22 12 22 12 12<br />

Fineness, m²/g<br />

turbidimeter 160 160 160 160 - - 160 160<br />

air permeability 280 280 280 280 - - 280 280<br />

Autoclave exp. 0.80 0.80 0.80 0.80 0.80 0.80 0.80 0.80<br />

Compr.<br />

strength<br />

MPa 1 day - - - - 12.0 10.0 - -<br />

3 days 12.0 10.0 10.0 8.0 24.0 19.0 - 8.0<br />

7 days 19.0 16.0 17.0 14.0 - - 7.0 15.0<br />

28 days - - - - - - 17.0 21.0<br />

Gillmore setting<br />

minutes init.set 60 60 60 60 60 60 60 60<br />

final set 600 600 600 600 600 600 600 600<br />

Vicat setting<br />

minutes init.set 45 45 45 45 45 45 45 45<br />

final set 375 375 375 375 375 375 375 375<br />

For special applications, additional chemical and physical requirements are set, e.g.<br />

• Limits on C3A content for cement types III and III A for moderate (


Cement<br />

Concrete<br />

b) Provisions for blended cements in C 595 are as follows<br />

• Type IS - Portland Blast-Furnace Slag Cement (slag content 25÷70%)<br />

• Type IP - Portland-Pozzolan Cement (pozzolan content 15÷40%)<br />

• Type P - Portland-Pozzolan Cement with low early strength<br />

• Type S - Slag Cement (slag content > 70%)<br />

• Type I(PM) - Pozzolan-modified Portland Cement (pozzolan content < 15%)<br />

• Type I(SM) - Slag-modified Portland Cement (slag content < 25%)<br />

these types can be MS (moderate sulfate resistent), A (air entrained) or MH (moderate heat<br />

of hydration).<br />

Table 2a ASTM C595M - Specification for Blended Hydraulic Cement -<br />

Composition<br />

Cement type<br />

Clinker<br />

+ gypsum<br />

BF slag<br />

Pozzolan<br />

/ fly ash<br />

IS (ev. MS, A, MH) 30 - 75 25 - 70<br />

S (ev. A) < 30 > 70<br />

IP (ev. MS, A, MH) 60 - 85 * 15 - 40<br />

P (ev. MS, A, MH) 60 - 85 * 15 - 40<br />

I (PM) (ev. MS, A, MH) > 85 * < 15<br />

I (SM) (ev. MS, A, MH) > 75 < 25<br />

* These cements can be produced by blending pozzolan either with ordinary Portland cement<br />

or with an IS type cement.<br />

Table 2b<br />

ASTM C595M - Specification for Blended Hydraulic Cement - Chemical<br />

Requirements<br />

Cement type →<br />

I(SM), I(SM)-A,<br />

IS, IS-A<br />

S. S-A I(PM), I(PM)-A,<br />

P, PA, IP, IP-A<br />

MgO, max % - - 6.0<br />

SO3 , max % 3.0 4.0 4.0<br />

Sulfide sulfur, max % 2.0 2.0 -<br />

Insoluble resid., max % 1.0 1.0 -<br />

Loss on ignit., max % 3.0 4.0 5.0<br />

Water sol. alkali, max % - 0.03 -<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 201 of 220


Cement<br />

Concrete<br />

Table 2c<br />

ASTM C595M - Specification for Blended Hydraulic Cement - Physical<br />

Requirements<br />

Cement type<br />

→<br />

I(SM)<br />

I(PM)<br />

IS, IP<br />

same<br />

with<br />

air en.<br />

IS(MS)<br />

IP(MS)<br />

same<br />

with<br />

air en.<br />

S SA P PA<br />

Autoclave exp. 0.80 0.80 0.80 0.80 0.80 0.80 0.80 0.80<br />

A. contraction 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.20<br />

Vicat setting<br />

minutes initial 45 45 45 45 45 45 45 45<br />

hours final 7 7 7 7 7 7 7 7<br />

% air of mortar 12 19±3 12 19±3 12 19±3 12 19±3<br />

Compr.strengt<br />

h<br />

min. MPA 3 d 13.0 10.0 11.0 9.0 - - - -<br />

7 d 20.0 16.0 18.0 14.0 5.0 4.0 11.0 9.0<br />

28 d 25.0 20.0 25.0 20.0 11.0 9.0 21.0 18.0<br />

Heat of hydrat.<br />

kJ/kg max 7d 290 290 290 290 - - 250 250<br />

28 d 330 330 330 330 - - 290 290<br />

Water req. % - - - - - - 64 56<br />

Drying shrink. - - - - - - 0.15 0.15<br />

Mortar<br />

expans.<br />

max % 14 0.02 0.02 0.02 0.02 0.02 0.02 0.02 0.02<br />

d<br />

8 weeks 0.06 0.06 0.06 0.06 0.06 0.06 0.06 0.06<br />

Sulfate resist.<br />

max exp.180 d - - 0.10 0.10 - - - -<br />

c) An interesting example of a performance oriented standard is ASTM C 1157M. Here<br />

blended cement are classified according to their performance, with no restrictions on the<br />

composition of the cement or its constituents.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 202 of 220


Cement<br />

Concrete<br />

The specified cement types are based on specific requirements for general use (GU), high<br />

early strength (HE), resistance to sulfate attack (MS and HS), heat of hydration (MH and LH),<br />

low reactivity with alkali-reactive aggregates (R).<br />

Table 3<br />

ASTM C1157M - Performance Spec. for Blended Cement<br />

Cement type → GU HE MS HS MH LH<br />

Autoclave length change, max % 0.80 0.80 0.80 0.80 0.80 0.80<br />

Vicat setting time, minutes<br />

initial 45 45 45 45 45 45<br />

final 420 420 420 420 420 420<br />

% air of mortar to be specified on documents to customer<br />

Compressive strength, min MPa<br />

1 day - 10.0 - - - -<br />

3 days 10.0 17.0 10.0 5.0 5.0 -<br />

7 days 17.0 - 17.0 10.0 10.0 5.0<br />

28 days - - - 17.0 - 17.0<br />

Heat of hydration, max kJ/kg<br />

7 days - - - - 290 250<br />

28 days - - - - - 290<br />

Mortar bar exp. 14 days, max % 0.020 0.020 0.020 0.020 0.020 0.020<br />

Sulfate exp., max % 6 months - - 0.10 0.05 - -<br />

1 year - - - 0.10 - -<br />

Option R, low ASR reactivity<br />

expansion 14 days, max % 0.020 0.020 0.020 0.020 0.020 0.020<br />

expansion 56 days, max % 0.060 0.060 0.060 0.060 0.060 0.060<br />

Optional physical requirements<br />

Early stiffening, final penetration<br />

min % 50 50 50 50 50 50<br />

Compressive strength, min. MPa<br />

28 days 28.0 - 28.0 - 22.0 -<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 203 of 220


Cement<br />

Concrete<br />

3.2 THE EUROPEAN STANDARD ENV 197-1<br />

This standard was issued in the early 90’s after some 25 years of drafting, with the aim at<br />

including in one and only specification all the cements produced all over Europe. It is clear<br />

that the extremely variable climatic, industrial and economic conditions of the continent were<br />

the main reason of the huge diversity of binders produced and consequently of the very<br />

tough job for the members of CEN to reach a reasonable compromise among member<br />

countries.<br />

The ENV 197-1 standard is divided in nine chapters, the most important are<br />

♦ chapter 4 “Constituents”, dealing with cement components; here some main<br />

characteristics of the materials to be used in cement production are set<br />

♦ chapter 5 “Cement types, composition and designation”, where the official names and<br />

abbreviations are described for every type of cement, according to its composition<br />

♦ chapter 6, 7, 8 “Mechanical-, Phisical-, Chemical requirements” setting limits for the main<br />

cement properties<br />

♦ chapter 9 “Conformity Criteria”, dealing with the procedures to apply for the autocontrol<br />

testing of cement and the subsequent evaluation of its conformity to the specifications.<br />

Description of cement components (constituents) in Chapter 4 is very accurate, complete<br />

and rigorous, including additional specifications (Table 5) that define the main component<br />

characteristics in order to assure their performance in cement.<br />

The ENV 197-1 standard recognizes and legitimates the use of some constituents that have<br />

a long tradition of use in many countries, but at the same time are not available in other<br />

areas (i.e. natural pozzolan, fly ash, burnt shale). There is absolute freedom, with the<br />

exception of Portland cement clinker as the only obligatory constituent, in the choice of<br />

materials to produce cements according to ENV 197-1, provided they comply with the<br />

requirements of Chapter 4.<br />

This freedom is reflected in Chapter 5 (Table 4), where cement compositions are reported.<br />

As many as 25 different types of cements can be produced, using one or more of the<br />

indicated constituents.<br />

It is worthwile noting that the relative amounts are in mass-% on the cement nucleus, that is<br />

without considering gypsum (since its amount in cement can vary according to purity and<br />

optimum gypsum content). Then, for each type of cement the industrial composition shall be<br />

calculated with the actual gypsum content.<br />

Mechanical, physical and chemical requirements are listed in Table 6 and 7. A significant<br />

difference in the strength classification is the subdivision of every main strength class into<br />

two subclasses, to differentiate rapid hardening cement from those with normal strength<br />

development. This makes the theoretical number of producible cements as high as 150 (25<br />

types x 6 classes)!<br />

Additionally, the upper limit for 28 day strength is also set. This is somewhat a marketing<br />

besides technical requirement, it has the advantage to force all producers to keep cement<br />

strengths (and quality) within a certain interval and at the same time protects against unfair<br />

competition. The drawback is, for some blended cements (high slag content, mainly), the<br />

serious difficulty to comply with both the minimum 2-day and the maximum 28-day strength<br />

limit, that is often exceeded.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 204 of 220


Cement<br />

Concrete<br />

The last chapter introduces a new criterion to evaluate conformity of cement, based on a<br />

continuous statistical control (autocontrol).<br />

Cements must fulfill<br />

• conformity criteria on single absolute values<br />

• statistical conformity criteria (according to variables or attributes) on characteristic<br />

limits.<br />

This means that for the main parameters, cement must conform to a minimum (or maximum)<br />

limit value, as it was previously provided by the standards, and also to a characteristic limit.<br />

All results over a six- or twelve month period are evaluated according to the mean value and<br />

standard deviation and calculated to yield a consumer’s risk lower than 5% (that is the risk<br />

that a defective lot is accepted as conforming the standard) for lower strength limits and 10%<br />

for the upper one.<br />

To make it simple, let’s consider the conformity for 28 day strength. According to ENV 197-1,<br />

the following equation shall apply<br />

x - k A s ≥ L (lower limit) as well as x + k A s ≥ U (upper limit)<br />

where L is the lower (resp. upper) strength limit at 28 days, s is the calculated standard<br />

deviation, k A is the acceptability constant (depending on the number of tested samples).<br />

Now if a cement of mean 28 day strength = 38.0 MPa is produced with s = 2.0 and tested<br />

over 110 samples/year (k A = 1.93 for L and 1.53 for U), the conformity equation for a 32.5<br />

MPa characteristic value will yield a result of<br />

x - k A s = 38.0 - 2.0 x 1.93 = 34.1 MPa, which is higher than 32.5 as well as<br />

x + k A s = 41.1 MPa which is lower than 52.5 then the cement is conforming the standard.<br />

But a value of s = 3.2 will yield<br />

x - k A s = 38.0 - 3.2 x 1.93 = 31.8 which is lower than 32.5; in this case the cement does not<br />

fulfill the standard specification set for minimum late strength.<br />

The following graphs will help to visualize the situation. The first one shows that a “good”<br />

cement, with low strength variability (s = 2) fulfill the requirements with an average 28-day<br />

strength of 36.4 MPa, while the other refers to a “bad” cement (s = 4) and the necessary<br />

average strength to fulfill specs is 40.2 MPa (3.8 MPa higher!).<br />

This approach is quite interesting and innovative, because it comes clear from the proposed<br />

example, that a consistent production, with low deviation, can allow for the production of a<br />

complying cement with lower mean strength values, that cannot be sufficient if the deviation<br />

is higher, thus forcing to raise the average strength to higher and more costly levels.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 205 of 220


Cement<br />

Concrete<br />

Figure 1 Conformity of a “good” cement (s = 2)<br />

60<br />

Compr. strength (MPa)<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

52.5<br />

32.5<br />

x - k . s > L<br />

a<br />

x + k . s < U<br />

a<br />

20<br />

0 1 2 3 4 5 6 7<br />

Standard deviation (s)<br />

Figure 2 Conformity of a “bad” cement (s = 4)<br />

60<br />

Compr. strength (MPa)<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

52.5<br />

32.5<br />

x - k . s > L<br />

a<br />

x + k . s < U<br />

a<br />

20<br />

0 1 2 3 4 5 6 7<br />

Standard deviation (s)<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 206 of 220


Cement<br />

Concrete<br />

Table 4<br />

The ENV 197-1 specification on common cements - Cement types and<br />

composition<br />

Cement<br />

type<br />

I<br />

Designation Abbrev. Clinker GGBF slag Micro<br />

silica<br />

Portland<br />

Cement<br />

Natural<br />

pozzol.<br />

Industr.<br />

pozzol.<br />

Silic.<br />

fly ash<br />

Calcar.<br />

fly ash<br />

Burnt<br />

shale<br />

I 95-100 - - - - - - -<br />

Portland Slag II/A-S 80-94 6-20 - - - - - -<br />

Cement II/B-S 65-79 21-35 - - - - - -<br />

Portland MS<br />

Cem<br />

Portland<br />

Pozzol.<br />

II/A-D 90-94 - 6-10 - - - - -<br />

II/A-P 80-94 - - 6-20 - - - -<br />

II/B-P 65-79 - - 21-35 - - - -<br />

Cement II/A-Q 80-94 - - - 6-20 - - -<br />

II/B-Q 65-79 - - - 21-35 - - -<br />

II II/A-V 80-94 - - - - 6-20 - -<br />

Portland<br />

Ash<br />

Fly<br />

II/B-V 65-79 - - - - 21-35 - -<br />

Cement II/A-W 80-94 - - - - - 6-20 -<br />

Portland<br />

Burnt<br />

Shale<br />

Cement<br />

Portland<br />

Limest.<br />

II/B-W 65-79 - - - - - 21-35 -<br />

II/A-T 80-94 - - - - - - 6-20<br />

II/B-T 65-79 - - - - - - 21-35<br />

II/A-L 80-94 - - - - - - -<br />

Cement II/B-L 65-79 - - - - - - -<br />

Portland<br />

Compos<br />

Limest. Other<br />

constit<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

6-20 0-5<br />

21-35 0-5<br />

II/A-M 80-94 <br />

Cement II/B-M 65-79 <br />

III/A 35-64 36-65 - - - - - -<br />

III Slag Cement III/B 20-34 66-80 - - - - - -<br />

III/C 5-19 81-95 - - - - - -<br />

- 0-5<br />

- 0-5<br />

- 0-5<br />

IV Pozzolanic IV/A 65-89 - - - - 0-5<br />

Cement IV/B 45-64 - - - - 0-5<br />

V Composite V/A 40-64 18-30 - - - - 0-5<br />

Cement V/B 20-39 31-50 - - - - 0-5<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 207 of 220


Cement<br />

Concrete<br />

Table 5<br />

ENV 197-1 - Requirements for cement constituents<br />

Type of constituent Property Conformity Limit<br />

Clinker sum of silicates C3S + C2S > 2/3 of total<br />

ratio CaO / SiO2 > 2.0 %<br />

MgO content < 5.0 %<br />

Blast Furnace Slag glass content > 2/3 of total<br />

CaO + MgO + SiO2<br />

> 2/3 of total<br />

CaO + MgO / SiO2 > 1.0<br />

Fly ash loss on ignition < 5.0 %<br />

Siliceous fly ash reactive CaO < 5.0 %<br />

reactive SiO2 > 25 %<br />

Calcareous fly ash reactive SiO2 > 5 %<br />

28 d strength of ground material > 10 MPa<br />

expansion on a 30% blend < 10 mm<br />

Pozzolans reactive SiO2 > 25 %<br />

Burnt shale 28 d strength of ground material > 25 MPa<br />

expansion on a 30% blend < 10 mm<br />

Limestone CaCO3 content > 75 %<br />

clay content (methilene blue test) < 1.2 %<br />

total organic carbon (TOC) < 0.2 %<br />

Table 6<br />

ENV 197-1 - Mechanical and Physical Requirements<br />

Strength Compressive strength (N/mm² - MPa) Initial Expans.<br />

class Early strength Standard strength setting (mm)<br />

2 days 7 days 28 days time(min)<br />

32,5 - > 16 > 32,5 < 52,5<br />

32,5 R > 10 - > 60<br />

42,5 > 10 - > 42,5 < 62,5 < 10<br />

42,5 R > 20 -<br />

52,5 > 20 - > 52,5 - > 45<br />

52,5 R > 30 -<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 208 of 220


Cement<br />

Concrete<br />

Table 7<br />

ENV 197-1 - Chemical Requirements<br />

Property Test refer. Cement type Strength class Requirement<br />

Loss on ignit. EN 196-2 CEM I All < 5.0 %<br />

CEM III<br />

Insol. residue EN 196-2 CEM I<br />

CEM III<br />

CEM I<br />

CEM II<br />

Sulfates SO3 EN 196-2 CEM IV<br />

CEM V<br />

CEM III<br />

All < 5.0 %<br />

32,5 - 32,5 R<br />

42,5<br />

< 3.5 %<br />

42,5 R<br />

52,5 - 52,5 R < 4.0 %<br />

All<br />

Chlorides EN 196-21 All All < 0.10 %<br />

Pozzolanicity EN 196-5 CEM IV All positive test<br />

After drafting for common cements, CEN is also preparing new specifications for “special<br />

cements”, such as sulfate resisting and low heat cements. The drafts are in an advanced<br />

stage and are likely to be distributed for approval by member States in 1998.<br />

3.3 THE AUSTRALIAN STANDARD AS 3972-1991<br />

This is a good example of how a standard can be drafted in a simple, clear and effective<br />

way.<br />

It consists of two pages only, where reference documents, materials and cements are listed<br />

and defined, as well as requirements, properties and dispatching conditions.<br />

All one need to know is contained in this astonishingly essential document.<br />

Cement types are defined as follows<br />

• type GP - general purpose Portland cement<br />

• type GB - general purpose blended cement<br />

while special purpose cements are<br />

• type HE - high early strength cement<br />

• type LH - low heat cement<br />

• type SR - sulfate resisting cement.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 209 of 220


Cement<br />

Concrete<br />

A blended cement should contain more than 5 % of mineral addition (fly ash, BF slag, or<br />

both).<br />

Specified properties are compiled in Table 8.<br />

Table 8<br />

AS 3972-1991 - Portland and blended cements<br />

Property<br />

Type Setting time Exp. SO3 Compr. strength Heat hydrat. C3A<br />

min. max. max max minimum MPa max. J/g max<br />

min h mm % 3 d 7 d 28 d 7 d 28 d %<br />

GP 45 10 5 3.5 - 25 40 - - -<br />

GB 45 10 5 3.5 - 15 30 - - -<br />

HE 45 10 5 3.5 20 30 - - - -<br />

LH 45 10 5 3.5 - 10 30 280 320 -<br />

SR 45 10 5 3.5 - 20 30 - - 5.0<br />

Additional properties such as<br />

• loss on ignition<br />

• fineness or fineness index<br />

• nature and proportion of materials in the cement<br />

• major oxide composition of the cement<br />

• chloride content, if exceeding 0.05 %<br />

can be provided by the manufacturer upon specific request from the purchaser.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 210 of 220


Cement<br />

Concrete<br />

3.4 Conclusion<br />

The importance of standards in our technical life and the present situation for cement have<br />

been highlighted. The expected trend for cement standards is represented in table 9.<br />

Table 9<br />

World trends in cement standards<br />

Object Past/present situation Trend<br />

Number of standardized<br />

cements<br />

Mineral components or<br />

additions in cement<br />

Few Increasing number of<br />

common and special<br />

cements<br />

Not allowed in OPC, limited<br />

use of blended cements<br />

Allowed in OPC, increased<br />

number of blended cements<br />

Strength Compressive and flexural Only compressive str.<br />

Compr. strength limit Minimum strength requir. Min. and max. stength<br />

Age of tests 1, 2, 3, 7, 28 days 2 and 28 days<br />

Standard mortar constant consistency, constant W/C ratio<br />

constant W/C ratio<br />

W/C ratio different in each country 0.5<br />

Setting time initial and final settiong t. initial setting time<br />

4. Standards for Mineral Components<br />

Mineral components or additions have always been used in the production of building<br />

materials.<br />

In ancient Greece and Rome blends of burnt lime and pozzolan were used for brickworking<br />

and building purposes.<br />

Nowadays the use of mineral additions has improved in quantity and quality.<br />

On one hand the available mineral components are quite extensively evaluated and utilized<br />

in cement manufacture, for production, cost-effectiveness and environmental purposes, on<br />

the other hand the advancement in technology of high performance concrete, in terms of<br />

ultra-high early strengths but also high durability, have benefited by the availability of these<br />

materials.<br />

Then the more sophisticated uses of mineral components require adequate standards to<br />

describe their properties and performance in cement and concrete.<br />

As a matter of fact, mineral components (and other mineral additions) can be mainly used as<br />

♦ cement constituents, in the production of blended or Portland modified cements<br />

♦ raw materials in concrete production, as value added products which impart additional<br />

characteristics to the cementitious conglomerate.<br />

Recent standardization has covered both of these aspects, and more is expected in the<br />

future.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 211 of 220


Cement<br />

Concrete<br />

Requirements set forth by standards dealing with cement production (see as an example<br />

Table 7 in previous chapter) are usually enclosed in the standard itself; they consider<br />

aspects of the chemical composition of mineral components and put sometimes also limits<br />

with respect to the hydraulic activity, the main responsibility on the effectiveness of the<br />

addition being left on the cement producer’s shoulders since he has to comply with the final<br />

product requirements.<br />

Standards on materials to be used as main concrete constituents are separate documents.<br />

They deal about the same parameters as for cement production, but they are mainly<br />

focussed on the impact that these characteristics will have on concrete.<br />

Some examples related to mineral components are reported<br />

♦ ASTM C 311 “Sampling and Testing Fly Ash or Natural Pozzolans for Use as a Mineral<br />

Admixture in Portland-Cement Concrete”<br />

♦ EN 450 “Fly Ash for Concrete”<br />

♦ ASTM C 989 “Specification for Ground Granulated Blast Furnace Slag for Use in<br />

Concrete and Mortar<br />

♦ ASTM C 1240 “Specification for Silica Fume for Use in Hydraulic Cement Concrete and<br />

Mortar.<br />

5. Concrete Standards<br />

5.1 Purpose<br />

The purpose of the concrete standards and specifications is to regulate the relations between<br />

the different parties involved in the activity of concrete construction.<br />

Besides the interest to regulate the interaction between the different parties, there is a<br />

supreme interest of the society to ensure that the constructions are stable, aesthetic and<br />

durable. A typical example are the antiseismic codes, which, although they lead to more<br />

costly structures, prevent enormous losses of human lives and goods in case of earthquakes<br />

compared to structures built in a normal way.<br />

Similarly to cement standards, the responsibility for the emission of the concrete standards<br />

varies a lot from country to country. For instance in Germany and Great Britain, this is the<br />

responsibility of the national standard entity (DIN and British Standards respectively),<br />

whereas in the United States and Switzerland, it is the task of a professional organisation<br />

(ACI (American Concrete Institute) and SIA (Society of Engineers and Architects)<br />

respectively). In other countries like Italy, it can also be the responsibility of the government<br />

through the Ministry of public works.<br />

5.2 Content of the Standards and Specifications<br />

It will be referred here exclusively to the aspects of concrete technology within the wider field<br />

of codes, standards and specifications, with a special emphasis on the corresponding<br />

prescriptions in use in the United States (ASTM, ACI) and in Europe (ENV 206).<br />

5.2.1 Quality of the Concrete Components<br />

In general, the concrete standards make reference to the respective standards for each of<br />

the ingredients (cement, aggregates, water, mineral and chemical admixtures). That is to say<br />

it is simply indicated that the components must satisfy the quality requirements established in<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 212 of 220


Cement<br />

Concrete<br />

the mentioned standards. The relevant ASTM and EN standards for the concrete<br />

components cement, admixtures and aggregates are listed in the annex.<br />

5.2.2 Quality of Fresh Concrete<br />

The following properties of the fresh concrete are specified:<br />

• consistency/workability (always)<br />

• air content (when air is intentionally entrained)<br />

• temperature (in extreme climates)<br />

• density (for light weight or heavy concrete)<br />

• stiffening rate (in extreme climates or for slip forming)<br />

In general, the constructor specifies the consistency of the concrete adequate for the<br />

structural element to be concreted and for the available means of placing and compacting.<br />

The concrete producer has to deliver a material which complies with this consistency within a<br />

certain tolerance. For instance the ASTM standard C 94 for ready-mix concrete prescribes:<br />

Specified Slump<br />

Tolerance<br />

< 50 mm ± 15 mm<br />

50-100 mm ± 25 mm<br />

> 100 mm ± 40 mm<br />

With respect to the content of intentionally entrained air, the ACI recommends percentages<br />

which depend on the severity of exposure of the concrete and the maximum aggregate size.<br />

A higher severity and a lower maximum aggregate size correspond to higher percentages of<br />

air. The ASTM standard C 94 establishes a tolerance of ± 1.5% for the specified air content.<br />

5.2.3 Strength of Hardened Concrete<br />

The strength is the most important characteristic to be specified for the hardened concrete. In<br />

the past, it was common to specify the mean strength of the concrete, so that, if the average<br />

of the results of the specimens was equal or greater than the specified values, the strength<br />

requirement was considered as fulfilled. Based on this criterion, the following two series were<br />

equivalent, since both presented the same average (30 MPa):<br />

Series "A": 32, 28, 34, 29, 28, 30, 32, 28, 29, 30<br />

Series "B": 40, 27, 48, 30, 15, 30, 36, 20, 28, 26<br />

At simple view, it results that the series "A" is better than "B", because its values are more<br />

uniform (s = 2 MPa against s = 9 of the series "B"). That is to say, although both comply with<br />

the specified mean strength, a structure built with concrete "A" will be safer from the<br />

structural point of view than if it is built with concrete "B", as the latter presents a strength<br />

result which is only the half of the specified value.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 213 of 220


Cement<br />

Concrete<br />

As already mentioned for cement, standards and specifications take into account the<br />

variability of concrete and apply the criterion of fractiles to specify the concrete strength. The<br />

specified strength corresponds to a fractile related to a certain probability, which means that<br />

the designer of the structure accepts that a certain percentage of the produced concrete is<br />

"defective". The allowed percentage of "defectives" varies according to the country (it has to<br />

be mentioned that the percentage has not necessarily to be related to the safety of the<br />

structures):<br />

Percentage p<br />

Country<br />

2 % Switzerland<br />

5 % Europe<br />

10 % United States<br />

20 % United States (massive concrete)<br />

Assuming that the distribution of strength values follows the law of Gauss, the fractile x P is<br />

calculated as<br />

x P = x M - z P . s<br />

where z P is a function of the percentage of "defectives" p. We can see that the mean x M as<br />

well as the standard deviation s participate in this criterion, which explains its universal<br />

acceptation.<br />

This principle has now to be applied for the design of the concrete mixes. If we have<br />

specified a certain strength that we will denote f' C (which corresponds to the fractile x P for a<br />

certain percentage of "defectives"), we have to design our mix with a certain margin taking<br />

into account the expected variability s E in the production process. That is to say we have to<br />

design the mix for a certain mean design strength f' D which will be equal to:<br />

f' D = f' C + margin = f' C + z P<br />

. sE<br />

It is evident that the producer which elaborates the concrete in a more uniform manner (lower<br />

s E ) will be able to design his mix for a lower strength and therefore more economically.<br />

5.2.4 Durability of Concrete<br />

The traditional criterion to guarantee the durability of the structures against attack of the<br />

environment has been the establishment of limits with respect to concrete composition. For<br />

instance ENV 206 and ACI code 318 for reinforced concrete establish limits regarding the<br />

w/c-ratio of the concrete in function of the exposure condition to which the structure will be<br />

submitted, trying to control concrete permeability to improve its durability against aggressive<br />

agents.<br />

The fundamental problem of such type of specification is that it is often omitted at the<br />

moment of ordering the concrete and anyway its interpretation is unclear (are they mean<br />

values, absolute maximums, fractiles?). Moreover, its compliance is very difficult to verify in<br />

practice (how is the w/c-ratio of a concrete determined?). The same criticism can be made<br />

with respect to the project of the European Standards, where, apart from the maximum w/cratio,<br />

also a minimum cement content is specified.<br />

A more practical criteria is the one established in the Australian Standards, where, instead of<br />

fixing limits for the concrete composition, a minimum strength is specified according to the<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 214 of 220


Cement<br />

Concrete<br />

type of exposure to which the structure will be submitted. The advantage of this criterion is<br />

that it is more easy to interpret and to verify its compliance in practice.<br />

5.3 The European Standard ENV 206<br />

Most discussions on concrete standards are concentrated on the issue of concrete durability,<br />

and in this context the specifications on cement content, water cement ratio, chloride content<br />

and alkali aggregate reaction.<br />

The durability aspects are here discussed in connection with the new ENV 206 draft<br />

„Concrete. Performance, production, placing and compliance criteria“, which was issued as a<br />

prestandard in 1989. The development of this standard has a similar history to the EN<br />

cement standard; the CEN Committee TC 94 started work in 1981, and the present draft is<br />

far from being the final version. Nevertheless, it is a good review of the present state of<br />

knowledge on the topic, comprising a consensus of expert opinions in Europe.<br />

ENV 206 should cover concrete in general; for all types of concrete - site, precast and ready<br />

mixed concrete, plain, reinforced and prestressed concrete. For special purposes, however,<br />

additional specifications may be needed, and for these reference is made to the individual<br />

national standards. The standard also makes reference to other European standards as<br />

regards types of cements, aggregates, mineral and chemical admixtures.<br />

5.3.1 Cement Content and Durability<br />

The minimum cement content in steel reinforced concrete is specified according to<br />

environmental conditions and type of concrete structure; it varies between 260 to 300 kg/m 3 ,<br />

as shown in tables 10 and 11.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 215 of 220


Cement<br />

Concrete<br />

Table 10<br />

Exposure Classes related to Environmental Conditions<br />

Exposure Class Examples of Environmental<br />

Conditions<br />

1 Dry environment interior of dwellings or offices 1)<br />

2 Humid<br />

environment<br />

a<br />

without frost<br />

b<br />

with frost<br />

3 Humid environment with frost<br />

and de-icing agents<br />

4 Seawater<br />

environment<br />

a<br />

without frost<br />

b<br />

with frost<br />

- interior of buildings where humidity is<br />

high (e.g. laundries)<br />

- exterior components<br />

- components in non-aggressive soil<br />

and/or water<br />

- exterior components exposed to frost<br />

- components in non-aggressive soil<br />

and/or water and exposed to frost<br />

- interior components where the<br />

humidity is high and exposed to frost<br />

interior and exterior components<br />

exposed to frost and de-icing agents<br />

- components completely or partially<br />

submerged in seawater or in the<br />

splash zone<br />

- components in saturated salt air<br />

(coastal area)<br />

- components partially submerged in<br />

seawater or in the splash zone and<br />

exposed to frost<br />

- components in saturated salt air and<br />

exposed to frost<br />

The following classes may occur alone or in combination with the above<br />

classes:<br />

5 Aggressive a - slightly aggressive chemical<br />

environment 2) - aggressive industrial atmosphere<br />

chemical<br />

environment (gas, liquid or solid)<br />

b moderately aggressive chemical<br />

environment (gas liquid or solid)<br />

c<br />

highly aggressive chemical environment<br />

(gas, liquid or solid)<br />

1) This exposure class is valid only as long as during construction the structure or some of<br />

its components is not exposed to more severe conditions over a prolonged period of time<br />

2) Chemically aggressive environments are classified in ISO 9690. The following equivalent<br />

exposure conditions may be used:<br />

∗<br />

∗<br />

∗<br />

Exposure class 5a: ISO classification A1G, A1L, A1S<br />

Exposure class 5b: ISO classification A2G, A2L, A2S<br />

Exposure class 5c: ISO classification A3G, A3L, A3S<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 216 of 220


Cement<br />

Concrete<br />

Table 11<br />

Durability Requirements related to Environmental Exposure<br />

Requirements Exposure Class according to Table 2<br />

Max. w/c ratio for 2)<br />

- plain concrete - 0.70<br />

1 2a 2b 3 4a 4b 5a 5b 5c 1)<br />

- reinforced concrete 0.65 0.60 0.55 0.50 0.55 0.50 0.55 0.50 0.45<br />

- prestressed concrete 0.60 0.60<br />

Min. cement content 2)<br />

in kg/m 3 for<br />

- plain concrete 150 200 200 200<br />

- reinforced concrete 260 280 280 300 300 300 280 300 300<br />

4) 4) 4)<br />

- - yes yes - yes - - -<br />

- prestressed concrete 300 300 300 300<br />

Min. air content of<br />

aggregate size of 3)<br />

fresh concrete in % for<br />

nominal max<br />

- 32 mm - - 4 4 - 4 - - -<br />

- 16 mm - - 5 5 - 5 - - -<br />

- 8 mm - - 6 6 - 6 - - -<br />

Frost resistant<br />

aggregates 6)<br />

Impermeable concrete<br />

according to clause<br />

7.3.1.5<br />

Types of cement for<br />

plain and reinforced<br />

concrete according to<br />

EN 197<br />

- - yes yes yes yes yes yes yes<br />

sulfate resisting<br />

cement 5) for sulfate<br />

contents<br />

> 500 mg/kg in water<br />

> 3000 mg/kg in soil<br />

These values of w/c ratio and cement content are based on cement<br />

where there is long experience in many countries. However at the<br />

time of drafting this pre-standard experience with some conditions in<br />

some countries. Therefore during the life of this pre-standard,<br />

particularly for exposure classes 2b, 3, 4b the choice of the type of<br />

cement and its composition should follow the national standard or<br />

regulations valid in the place of use of the concrete. Alternatively the<br />

suitability for use of the cements may be proved by testing the<br />

concrete under the intended conditions of use.<br />

Additionally cement CEI may be used generally for prestressed<br />

concrete. Other types of cement may be used if experience with<br />

these types is available and the application is allowed by the national<br />

standards or regulations valid in the place of use of the concrete.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 217 of 220


Cement<br />

Concrete<br />

1) In addition, the concrete shall be protected against direct contact with the aggressive<br />

media by coatings unless for particular cases such protection is considered unnecessary.<br />

2) For minimum cement content and maximum water/cement ratio laid down in this standard<br />

only cement listed in clause 4.1 shall be taken into account. When pozzolanic or latent<br />

hydraulic additions are added to the mix, national standards or regulations, valid in the<br />

place of use of the concrete, may state if and how the minimum or maximum values<br />

respectively are allowed to be modified.<br />

3) With a spacing factor of the entrained air void system < 0.20 mm measured on the<br />

hardened concrete.<br />

4) In cases where the degree of saturation is high for prolonged periods of time. Other<br />

values or measures may apply if the concrete is tested and documented to have<br />

adequate frost resistance according to the national standards or regulations valid in the<br />

place of use of the concrete.<br />

5) The sulphate resistance of the cement shall be judged on the basis of national standards<br />

or regulations valid in the place of use of the concrete.<br />

6) Assessed against the national standards or regulations valid in the place of use of the<br />

concrete.<br />

These requirements reflect the generally accepted opinion that the durability of steel<br />

reinforced concrete is governed mainly by its porosity and thus strongly affected by the w/c<br />

ratio and cement content.<br />

An important feature of the standard is the differentiation of the cement content according to<br />

the severity of its exposure. This allows the concrete producer to use their skills and<br />

knowledge in order to design concrete mixes more economically. A prerequisite of this<br />

differentiation is, of course, that information on the different conditions to which the specific<br />

concrete job is subjected is available.<br />

The prescription of cement content and w/c ratio has some problems: It is difficult to<br />

determine both of them in fresh and hardened concrete for control purposes. Furthermore,<br />

concrete durability depends not only on cement content as a w/c ratio, it is affected as well<br />

by the care taken in its transport, placing, compacting and curing. Therefore, it is<br />

indispensable that the quality of concrete in the construction after placing and curing is<br />

tested. We know that the porosity of a concrete governs nearly all aspects of durability.<br />

Therefore, investigations are under way to introduce a test method enabling the<br />

measurement of the porosity of the concrete from specimens taken from the hardened<br />

concrete structure. It consists of measuring the amount of gas which flows through the pore<br />

system of the concrete specimen in a unity of time. By so measuring the porosity, not only is<br />

care taken by the concrete producer, but also that of the contractor placing, compacting and<br />

curing the concrete is checked.<br />

An alternative way of testing the porosity and additional aspects of the concrete microstructure<br />

is the microscopical examination of very small specimens. It is very useful for a<br />

qualitative inspection of concrete quality, however, it seems difficult to apply this method for a<br />

quantitative assessment in the framework of a standard.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 218 of 220


Cement<br />

Concrete<br />

5.3.2 Chloride Content<br />

The maximum chloride content in concrete permitted in plain, reinforced and prestressed<br />

concrete is shown in table 12.<br />

Table 12<br />

Maximum Chloride Content of Concrete<br />

Concrete<br />

C1- by Mass of Cement<br />

Plain concrete 1%<br />

Reinforced concrete 0.4%<br />

Prestressed concrete 0.2%<br />

The relatively high content of chloride permitted in steel reinforced concrete, reflects the<br />

results of recent investigation, which show that no corrosion of steel is expected at a chloride<br />

content of 0.4% by mass of cement. Similar soft limits are prescribed in the ACI standards as<br />

shown in table 13.<br />

Table 13 Chloride Content in Concrete, ACI 318-83<br />

Maximum water-soluble chloride ion concentrations in hardened<br />

concrete at an age of 28 days contributed from the ingredients<br />

including water, aggregates, cementitious materials and admixtures<br />

Type of Member<br />

Limit<br />

(by mass<br />

of concrete)<br />

Prestressed concrete 0.06<br />

Reinforced concrete exposed to chloride in service 0.15<br />

Reinforced concrete that will be dry or protected from moisture 1.00<br />

in service<br />

Other reinforced concrete construction 0.30<br />

5.3.3 Alkali Aggregate Reaction<br />

To prevent deterioration of concrete due to alkali aggregate reaction, it is recommended to<br />

apply one or more of the following measures:<br />

♦ Limit the total alkali content of the concrete mix<br />

♦ Use a cement with a low effective alkali content<br />

♦ Change the aggregates<br />

♦ Limit the degree of saturation of the concrete e.g. by impermeable membranes<br />

Unfortunately no mention is made of the effective measures of using blended cements or<br />

mineral admixtures.<br />

5.3.4 Maintenance<br />

Finally, the author would like to point out one serious deficiency of concrete standards; it is<br />

the lack of specifications on maintenance of concrete. This is becoming recognized as a gap<br />

and attempts are being made to bridge it. Of course this means turning away from the<br />

opinion the cement producers have frequently promoted in the past, that concrete need no<br />

maintenance whatsoever.<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 219 of 220


Cement<br />

Concrete<br />

5.4 Referenced documents<br />

5.4.1 Relevant ASTM Standards for Concrete Components<br />

5.4.1.1 CEMENT<br />

C 150<br />

C 595<br />

Specification for Portland Cement<br />

Specification for Blended Hydraulic Cements<br />

C 1157 Performance Specifications for blended Hydraulic Cements<br />

5.4.1.2 ADMIXTURES<br />

C 260<br />

C 494<br />

Specification for Air-Entraining Admixtures for Concrete<br />

Specification for Chemical Admixtures for Concrete<br />

C 618 Specification for Fly Ash and raw or Calcined Natural Pozzolan for Use as Mineral<br />

Admixture in Portland Cement Concrete<br />

5.4.1.3 AGGREGATE<br />

C 33 Specification for Concrete Aggregates<br />

C 330<br />

C 331<br />

C 332<br />

C 637<br />

Specification for Lightweight Aggregates for Structural Concrete<br />

Specification for Lightweight Aggregates for Concrete Masonry Units<br />

Specification for Lightweight Aggregates for Insulating Concrete<br />

Specification for Aggregates for Radiation-Shielding Concrete<br />

5.4.2 Relevant EN Standards for Concrete Components<br />

5.4.2.1 CEMENT<br />

ENV 197-1 Common Cement<br />

ENV 197-X Low-heat Cement<br />

ENV 197-Y Sulfate resisting Cement<br />

5.4.2.2 ADMIXTURES<br />

EN 480 Admixtures for Concrete, Mortar and Grout. Test Methods<br />

EN 934 Admixtures for Concrete, Mortar and Grout<br />

5.4.2.3 AGGREGATES<br />

EN 932 Tests for General Properties of Aggregates<br />

EN 933 Tests for Geometrical Properties of Aggregates<br />

EN 1097 Tests for Mechanical and Physical Properties of Aggregates<br />

EN 1367 Tests for Thermal and Weathering Properties of Aggregates<br />

EN 1744 Tests for Chemical Properties of Aggregates<br />

Course for Cement Applications - <strong>2005</strong> Cement Section - Page 220 of 220


Cement<br />

Concrete<br />

Concrete - Contents<br />

Chapter No.<br />

1. CONCRETE MAIN COMPONENTS................................................................................. 2<br />

2. FRESH AND HARDENED CONCRETE ........................................................................ 33<br />

3. CONCRETE MIX DESIGN ............................................................................................. 68<br />

4. REDUCTION IN MATERIALS COSTS........................................................................... 85<br />

5. CONCRETE ADMIXTURES........................................................................................... 99<br />

6. CONCRETE PRODUCTION PRACTICES................................................................... 114<br />

7. CONCRETE SHRINKAGE ........................................................................................... 135<br />

8. CONCRETE DURABILITY........................................................................................... 157<br />

9. ALKALI-AGGREGATE REACTIONS IN CONCRETE ................................................ 168<br />

10. SELF-COMPACTING CONCRETE.............................................................................. 180<br />

Additional Papers<br />

CONCRETE TECHNOLOGY ............................................................................................. 180<br />

DURABLE HIGH-PERFORMANCE CONCRETE.............................................................. 216<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 1 of 235


Cement<br />

Concrete<br />

Concrete Main Components<br />

1. MAIN COMPONENTS OF CONCRETE ............................................................................... 3<br />

1.1 Definition of concrete.................................................................................................. 3<br />

1.2 Cement ....................................................................................................................... 3<br />

1.3 Aggregates ................................................................................................................. 4<br />

1.3.1 Aggregates from natural sources ................................................................. 4<br />

1.3.2 Classification ................................................................................................ 6<br />

1.3.3 Chemical properties ..................................................................................... 9<br />

1.3.4 Physical properties....................................................................................... 9<br />

1.3.5 The grading ................................................................................................ 12<br />

1.3.6 Alternative aggregates ............................................................................... 16<br />

1.4 Admixtures................................................................................................................ 18<br />

1.4.1 Workability agents...................................................................................... 19<br />

1.4.2 Air-entraining agents .................................................................................. 23<br />

1.4.3 Agents affecting setting and hardening...................................................... 25<br />

1.4.4 Other admixtures........................................................................................ 27<br />

1.5 Mixing and curing water for concrete........................................................................ 28<br />

1.5.1 Introduction ................................................................................................ 28<br />

1.5.2 Effect of impurities...................................................................................... 29<br />

1.5.3 Effects of algae on air content and strength............................................... 30<br />

1.5.4 Curing water............................................................................................... 30<br />

2. LITERATURE FOR ADMIXTURES AND MIXING WATER................................................ 31<br />

3. APPENDICES ..................................................................................................................... 31<br />

3.1 REFERENCES ......................................................................................................... 31<br />

3.2 Relevant ASTM Standards ....................................................................................... 32<br />

3.2.1 Aggregate................................................................................................... 32<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 2 of 235


Cement<br />

Concrete<br />

1 Main Components of Concrete<br />

1.1 Definition of concrete<br />

Concrete is a composite material<br />

that consists of aggregate (gravel + sand),<br />

cement and water<br />

and frequently admixtures and/ or additives.<br />

The aggregate consists of a conglomerate of generally different large grains, mostly natural<br />

rocks. In particular cases also of waste wood, polystyrene or metals (waste steel or other).<br />

As binders hydraulic materials are used, generally ordinary Portland cement. In special<br />

cases asphaltic materials or artificial resin will be applied.<br />

Aggregates are generally designated as either fine (sand: 0.025 to app. 5 mm) or coarse<br />

(gravel: from app. 5 to 32 mm or larger). A 'concrete' with sand only is named mortar.<br />

The cement paste coats the particles of aggregate and fills the spaces between them. While<br />

fresh, the cement paste also provides the lubrication that reduces the friction between the<br />

aggregate particles and imparts workability to the fresh mix. When hardened, the paste (now<br />

called cement stone) binds together the particles of aggregate.<br />

The following figure shows the<br />

SYSTEM: Binder + Water + Aggregate = Concrete<br />

General Case<br />

Special Case<br />

Concrete = Filler + Binder<br />

OPC concrete = (Gravel + Sand)<br />

aggregate<br />

+ OPC paste<br />

Mortar = Sand + OPC paste<br />

OPC paste = OPC + Water<br />

1.2 Cement<br />

The most important component of concrete, the cement, has been treated in detail in a<br />

separate chapter.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 3 of 235


Cement<br />

Concrete<br />

1.3 Aggregates<br />

1.3.1 Aggregates from natural sources<br />

Definition<br />

Natural aggregates are a mixture of uncrushed and/or crushed rocks and minerals. It<br />

consists of particles of different (or sometimes approx. equal) sizes and are used with a<br />

binder (cement) and water to produce concrete or mortar [1].<br />

General<br />

The mineralogical, physical and chemical characteristics of rocks are determined mainly by<br />

the events of their geological history. A knowledge of the ways in which rocks are formed,<br />

and of the various natural processes (whereby their original characteristics are altered) may<br />

therefore lead to a better understanding of those intrinsic properties which determine the<br />

suitability of a rock as a source of concrete aggregate.<br />

Rock classification according to their formation:<br />

♦ Igneous rocks<br />

♦ Sedimentary rocks<br />

♦ Metamorphic rocks<br />

Igneous rocks<br />

Igneous rocks are those which are formed by the solidification of molten masses, and many<br />

of their characteristics are determined by the rate and condition of cooling. Extruded volcanic<br />

rocks, i.e. those which have been ejected on the earth's surface, are cooled very rapidly, so<br />

that crystallisation of the component minerals is generally only partly effected, and the<br />

resulting rock consists of a mixture of crystalline ingredients and glassy matter.<br />

On the other hand rocks of the intruded igneous type, i.e. those which are formed from<br />

molten rock which has been intruded into overlying rock masses to form sills (a tabular<br />

intrusion of the surrounding rock), laccoliths (a concordant intrusion that has domed the<br />

overlying rocks) and dikes (rock solidification as a tabular body in a vertical fissure) etc. have<br />

been allowed to cool slowly, so that all the several compounds of the rock mass have had<br />

time to become thoroughly crystallised [8, 9].<br />

The igneous rocks are the primary rocks and from these all the other rocks are ultimately<br />

derived. The mineral constituents of igneous rocks have been classified as follows:<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 4 of 235


Cement<br />

Concrete<br />

Table 1:<br />

Minerals of igneous rocks<br />

Feldapatic silicates<br />

Orthoclase<br />

Plagioclase<br />

Leucite<br />

Ferromagnesian silicates<br />

Pyroxene<br />

Amphibole<br />

Biotite<br />

Olivine<br />

Free silica<br />

Quartz<br />

Tridymite<br />

Opal<br />

Accessory minerals<br />

Magnetite<br />

Ilmenite<br />

Haematite<br />

Apatite<br />

Sedimentary rocks<br />

Sedimentary rocks are derived from the chemical or mechanical breakdown of older rocks.<br />

The fragments resulting from such disintegration accumulate in deposits, for the most part<br />

under water, and the particles may be cemented together by the deposition of other<br />

siliceous, calcareous, argillaceous or ferruginous materials to form a dense mass.<br />

Chemical precipitates are crystalline, but by far the greater proportion of sedimentary rocks<br />

are made up of fragments of all sizes of earlier rocks. As in the case of sand or gravel beds,<br />

the fragments may lie loosely together, or may be firmly cemented as a compact material.<br />

The fragments of the older rocks may be sharply angular, but as they are often transported<br />

by wind or by water or by glacier for considerable distances, attrition may cause them to<br />

become smoothly rounded in their passage to the final deposit.<br />

As a result of their mode of deposition, the structure of sedimentary rocks is almost invariably<br />

stratified, and is essentially different from that of igneous rocks. Glacial deposits or coral<br />

reefs are not stratified and in this respect differs from other rocks of this class.<br />

Due to the geological and climatological changes, sudden radical changes often occur in the<br />

nature of the material being deposited. For this reason limestone are frequently interbedded<br />

with shale, sandstone with siltstones and quartzite with dolomite [3].<br />

Metamorphic rocks<br />

Metamorphic rocks are those resulting from the alteration in place of pre-existing<br />

sedimentary or igneous rocks. Such alteration is brought about by the application of high<br />

temperatures and/ or pressures to the rock mass.<br />

A large group of rocks in this class is characterised by foliated structure, which is not to be<br />

confused with the stratification of sedimentary rocks nor with the flow banding of some lava,<br />

but which is the result of the segregation of one or more of the constituents minerals and<br />

their reorientation in form of parallel plates.<br />

Metamorphic rocks vary greatly both in structure (the attitude and relative position of the rock<br />

masses of an area) and in texture (the general appearance of a rock, e.g. the size, shape,<br />

and arrangement of the constituent elements). Where the rock has been formed under high<br />

temperature and pressure, an equigranular texture and a massive internal structure results<br />

and lend great strength and toughness to the rock. Such rocks include many hornfelses and<br />

quartzites.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 5 of 235


Cement<br />

Concrete<br />

1.3.2 Classification<br />

From the petrologic view of point the aggregates, whether crushed or naturally reduced in<br />

size, can be divided into several groups of rocks having common characteristics. The group<br />

classification does not imply suitability of any aggregate for concrete-making: unsuitable<br />

material can be found in any group, although some groups tend to have a better record than<br />

others.<br />

Table 2: Classification of natural aggregates according to rock type [4].<br />

Basalt Group Flint Group Gabbro Group<br />

Andesite Chert Basic Diorite<br />

Basalt Flint Basic Gneiss<br />

Basic Porphyrites Hornstone Gabbro<br />

Dolerites<br />

Hornblende-rock<br />

Spilite<br />

Norite<br />

Serpentine<br />

Granite group Gritstone group Hornfels group<br />

Gneiss Arkose Contact-altered rocks<br />

Granite Greywacke (all kinds except marble)<br />

Granulite<br />

Grit<br />

Pegmatite<br />

Sandstone<br />

Syenite<br />

Tuff<br />

Pophyry Group Limestone group Quartzite group<br />

Aplite Dolomite Ganister<br />

Felsite Limestone Quartzitic sandstone<br />

Granophyre Marble Re-crystallized quartzite<br />

Porphyry<br />

Rhyolite<br />

Trachyte<br />

Schist group<br />

Phyllite<br />

Schist<br />

Slate<br />

All severely shared rocks<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 6 of 235


Cement<br />

Concrete<br />

Figure 1:<br />

Classification according other parameters:<br />

According to type of particle<br />

Naturally granular<br />

Crushed<br />

According to the origin of the<br />

Igneous rock<br />

Sedimentary rock<br />

Metamorphic rock<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 7 of 235


Cement<br />

Concrete<br />

According to particle size<br />

Fine Sand Gravel Coarse<br />

The particle dispersion limits are<br />

indicated in national standards<br />

According to particle characteristics<br />

Surface texture<br />

General characteristics<br />

Smooth<br />

Moderately<br />

rough<br />

Rough<br />

Weathered<br />

Flaky<br />

Healthy<br />

Shape<br />

Compact<br />

Flat<br />

Long<br />

Rounded or angular<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 8 of 235


Cement<br />

Concrete<br />

1.3.3 Chemical properties<br />

While it is desirable that the aggregate should as far as possible be chemically inert, many<br />

natural aggregates contain substances which are deleterious in concrete. Substances<br />

considered chemically deleterious may be broadly classified into five groups:<br />

Group 1:<br />

Substances soluble in water, which may be leached out of the aggregate thereby<br />

weakening it or promoting efflorescence in the concrete, e.g. common salt.<br />

Group 2:<br />

Soluble substances, or substances which become soluble in the cement matrix, which may<br />

interfere with the normal hydration of the cement, e.g. humid acid.<br />

Group 3:<br />

Substances which react with the cement destroying its properties, e.g. sodium sulphate.<br />

Group 4:<br />

Substances which may react with the alkali constituents of the cement, e.g. opal (AAR,<br />

Alkali-Aggregate Reaction).<br />

Group 5:<br />

Substances which may cause corrosion of reinforcing steel, e.g. common salt.<br />

1.3.4 Physical properties<br />

Perhaps the most important concept developed from research is that the aggregate must be<br />

studied not only in its relation to the hardened cement paste, i.e. as a component of<br />

concrete, but in relation to the environment of the concrete during its service life. With this<br />

concept in mind, the physical properties of natural stone aggregates may be considered in<br />

terms of following factors:<br />

Table 3:<br />

Properties of rocks<br />

Rock Type<br />

Relative<br />

density<br />

[kg/m 3 ]<br />

Various properties<br />

Water<br />

Absorption<br />

[Vol. %]<br />

Compres<br />

-sion str.<br />

[MPa]<br />

10%-<br />

Crushing<br />

value [kN]<br />

Strength Properties<br />

Modulus<br />

of rupture<br />

[MPa]<br />

Modulus of<br />

elasticity<br />

[GPa]<br />

Shear<br />

strength<br />

[MPa]<br />

Andesite 2,7 - 2,9 0,4 - 0,5 500 - 540 450 -- 80 - 110 --<br />

Basalt 2,8 - 3,0 0,1 - 0,3 190 - 470 340 14 - 24 49 - 98 --<br />

Dolerite 2,8 - 3,1 0,1 - 0,7 160 - 375 180 - 340 14 -24 60 - 105 --<br />

Dolomite 2,7 - 2,9 0,1 - 0,3 200 - 380 130 - 180 -- 100 - 130 --<br />

Felsite -- 0,2 - 1,5 360 - 450 230 - 380 -- 70 - 90 --<br />

Granite 2,6 - 2,8 0,2 - ,5 70 - 325 120 - 220 0 - 20 14 - 70 14 - 30<br />

Hornfels 2,6 - 2,8 -- -- 140 - 260 -- -- --<br />

Limestone 2,6 - 2,9 0,2 - 0,6 20 - 240 180 4 - 15 21 - 71 8 - 21<br />

Marble 2,6 - 2,9 0,2 - 0,6 20 - 240 -- 4 - 24 48 - 96 9 - 45<br />

Quartzite 2,6 - 2,8 0,2 - 0,6 105 - 480 160 - 280 12 - 24 64 - 86 --<br />

Rhyolite 2,6 - 2,8 -- 180 - 530 210 - 290 -- -- --<br />

Sandstone 2,5 - 2,7 0,2 - 0,9 10 - 255 50 - 280 3 - 14 14 - 55 2 - 21<br />

Syenite 2,6 - 2,7 0,4 - 1.5 95 - 445 -- -- 55 - 72 --<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 9 of 235


Cement<br />

Concrete<br />

Strength<br />

The compressive strength of cubical specimens of natural stone from which acceptable<br />

aggregate is derived, normally varies from about 70 to 400 MPa, while individual results as<br />

high as 540 MPa have been recorded [5].<br />

With a wide range there appears to be only a poor correlation between the compressive<br />

strength of the aggregate and the flexural or compressive strength of the concrete.<br />

As an index of overall quality the "10 % fines aggregate crushing value" or 10 % FACT value<br />

of coarse aggregates is useful (BS 812). However, it must be repeated that the usefulness of<br />

these tests is confined to a general assessment of quality and to the establishment of<br />

acceptance limits. New researches by Davis and Alexander have, however, shown a<br />

relationship between the aggregate type and the properties of concrete [6].<br />

Elasticity<br />

The modulus of Elasticity of concrete depends to a considerable degree on that of the<br />

aggregate from which it is made. The flexural strength of concrete is also depending on this<br />

property of the aggregate, and the use of an aggregate having a high elastic modulus will<br />

usually result in a concrete of high flexural strength, other factors being equal [7].<br />

The drying shrinkage of concrete is reduced by the use of aggregate having a high elastic<br />

modulus. Low shrinkage, induced by the use of rigid aggregates may be undesirable as the<br />

restraint provided by the aggregate increases the cracking tendency of the paste, thereby<br />

reducing the durability of the concrete.<br />

It would appear on balance that the reduction in durability resulting from the use of aggregate<br />

of high elastic modulus would in most cases more than offset any advantages which may be<br />

gained thereby.<br />

Porosity<br />

Many authorities consider that the size, abundance and continuity of pores in a rock particle<br />

are its most important physical properties. It is considered that the size and nature of the<br />

pores affect the physical strength of the aggregate, and control water absorption and<br />

permeability, thereby determining the durability of the aggregate in regard to freezing and<br />

thawing, and its resistance to chemical attack.<br />

It has been previously pointed out that the strength of the aggregate over a wide range<br />

cannot be considered as an important factor. Furthermore, in tropic countries a high<br />

resistance of freezing and thawing is not often an important quality of concrete. This property<br />

should, however, be considered where impermeability or high resistance of chemical attack<br />

are important features of the construction. Sandstone and shale show an average porosity of<br />

18 to 19 %, some limestone types 8 % and quality rocks 1 to 3 %. Individual results may be<br />

much higher [3].<br />

Thermal expansion<br />

In recent years it has come to be appreciated that the deterioration of concrete structures<br />

may be significantly affected by differences between the coefficients of thermal expansion of<br />

aggregate and the cement matrix in which it is embedded. In particular, the use of an<br />

aggregate of very low coefficient of expansion may lead to disintegration of the concrete. As<br />

the temperature of concrete, made with such an aggregate is lowered, the cement paste<br />

tends to contract more than the aggregate, with the result that tensile stresses are set up in<br />

the former, which may result in cracking.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 10 of 235


Cement<br />

Concrete<br />

Bond characteristics<br />

Perhaps one of the most important attributes of a concrete aggregate is its capacity for<br />

bonding strongly with cement paste. This factor has a significant influence on the flexural<br />

strength [7].<br />

In a wider field the word "surface texture" should be replaced by "bond characteristics". The<br />

latter term embraces not only surface texture, but the extent to which cement paste can<br />

penetrate into pores are surface depressions, the angularity of planes of fracture, the friable<br />

of the surface, and the presence or absence of loosely bonded or friable coatings.<br />

The effect of bond characteristics, while sufficiently significant in the medium strength range<br />

of mixes, becomes much more pronounced in the case of high-strength concretes.<br />

Particle shape and surface texture<br />

The shape of aggregate particles in both sand and stone, is one of the most significant<br />

factors affecting the behaviour of a concrete mix. Concrete water demand and water<br />

requirement on the design of concrete mixes, are strongly influenced by the shape of<br />

aggregate. Spherical, cubical or chunky shapes produce concrete having lower water<br />

demand than particles that are elongated or flaky.<br />

Surface texture of the grains also affects water demand, although to a lesser extent than<br />

shape. Rough textures increase water demand due to greater surface area and to increase<br />

friction and mechanical interlock of the particles.<br />

It seems that the shape and surface texture of aggregate influence considerably the strength<br />

of concrete. The flexural strength is more affected than the compressive strength Some data<br />

of Kaplan's research [7] are reproduced in the following table, but this gives more than an<br />

indication of the type of influence.<br />

Table 4:<br />

Effect of aggregate properties<br />

Relative effect of aggregate properties<br />

Property of concrete Shape Surface Texture Modulus of<br />

Elasticity<br />

Flexural strength 31 % 26 % 43 %<br />

Compressive strength 22 % 44 % 34 %<br />

Thermal properties<br />

Thermal properties which may be of significance are:<br />

• specific heat (or heat capacity)<br />

• thermal conductivity<br />

• thermal diffusivity<br />

All these properties are significant in the design of radiation shielding for nuclear power<br />

plants and in estimating loads on air-conditioning plant for buildings.<br />

Resistance to abrasion<br />

This is of significance in the choice of aggregate for pavements, industrial floors, channels,<br />

conveying abrasive materials and for certain type of silos.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 11 of 235


Cement<br />

Concrete<br />

Soundness<br />

The sulphate soundness test (ASTM, C 88 and SABS 1083) is not highly significant but<br />

provides a rough indication of the durability of an aggregate when subjected to expected<br />

environmental conditions. There is no evidence available to suggest that the mechanism of<br />

disruption of the test can be directly correlated with performance of an aggregate when<br />

subjected to these conditions. For this reason it is important that when an aggregate is to be<br />

evaluated on the basis of this test, it should be done by comparison with an aggregate of<br />

similar mineralogical composition and geological history and which has proved to be<br />

satisfactory in practice.<br />

1.3.5 The grading<br />

The grading of an aggregate refers to the distribution of particles of various sizes, and is<br />

determined by passing a representative sample through a series of standard sieves.<br />

Concrete aggregates are divided into two categories of fine aggregate or sand and coarse<br />

aggregate or stone. The boundary between these two is the 4 mm sieve but a fine aggregate<br />

may contain a small proportion of oversized particles and coarse aggregates may have some<br />

undersized material and still comply with the definition.<br />

The German standard DIN 1045 sieves for fine and coarse aggregates are:<br />

0.25 - 0.50 - 1.0 - 2.0 - 4.0<br />

8.0 - 16.0 - 31.5 - 63.0 - mm<br />

The sizes of the openings in the consecutive sieves are related by a constant ratio, the clear<br />

opening of each sieve size being twice or nearly twice that of the next smaller size.<br />

Fineness modulus<br />

The grading analysis provides data from which the fineness or coarseness of aggregate can<br />

be judged. The measure of fineness or coarseness is expressed in terms of an index known<br />

as the fineness modulus or FM. This is defined as an empirical factor obtained by adding the<br />

total percentages of material retained on each of the standard sieve sizes, except this<br />

amount smaller than 0,25 mm, and dividing the sum by 100. An example is given in the<br />

following table:<br />

Table 5:<br />

Determination of fineness modulus<br />

Sieve size<br />

[mm]<br />

Mass retained on<br />

sieve [g]<br />

Mass retained on<br />

sieve [wt-%]<br />

Total retained on<br />

sieve [wt-%]*<br />

Total passing on<br />

sieve [wt-%]<br />

31,5


Cement<br />

Concrete<br />

The FM is an index of the material, that is, a measure of the average particle size based on<br />

purely empirical classification. But the index in itself does not describe the grading. The FM<br />

indicates whether a material is fine, medium or coarse. A very fine aggregate grading with a<br />

maximum grain size of 31,5 mm is one having a modulus of 3,30 or less but in general run<br />

up to about 4.0. Medium aggregates are those with FM from 4.0 to 5.0 while coarse<br />

aggregates have higher fineness modulus (> 5.5).<br />

Grading of sand<br />

While the grading of a sand has a relatively minor effect on water demand, it has a major<br />

influence on the workability, cohesiveness and bleeding properties of concrete in its plastic<br />

state. Experience gained over many years has shown that the proportions of sand passing<br />

the 0.090, 0.125 and 0.25-mm sieves have the greatest effect on the properties of the mix<br />

while the contribution on the remaining fractions is less significant. Real values are shown in<br />

the following table:<br />

Table 6:<br />

A guide to sand grading<br />

Sieve Size [mm]<br />

Percentage passing<br />

Suggested outer limits Preferred limits *)<br />

4,0 80 - 100 90 - 100<br />

2,8 68 - 100 80 - 100<br />

2,0 55 - 100 75 - 95<br />

1,4 43 - 92 70 - 85<br />

1,0 32 - 85 55 - 70<br />

0,5 16 - 65 40 - 60<br />

0,25 5 - 43 20 - 40<br />

0,125 2 - 22 10 - 20<br />

0,09 0 - 11 3 - 6<br />

*) limits are suggested for pump concrete and for concrete used in sliding formwork<br />

The dramatic increase in surface area as the particle size decrease is illustrated by a few<br />

examples. The effect of grading on concrete behaviour is not constant but depends on the<br />

cement content and workability of the mix. Generally it may be accepted that a lower cement<br />

content calls for a finer grading of the aggregates: in other words, the leaner the mix, the<br />

higher should be the proportion of fineness (the total sum of sand < 0,1 mm plus cement<br />

should be approx. 350 kg per m 3 concrete).<br />

Table 7:<br />

Specific surface of various materials<br />

MATERIAL<br />

Specific Surface [m 2 /kg]<br />

32-mm concrete aggregate 0,07<br />

Average grading (B32) of aggregates 2,7<br />

Average concrete sand 3,5<br />

Silt 35<br />

Ordinary Portland cement 300<br />

Kaolinite 5'000<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 13 of 235


Cement<br />

Concrete<br />

Blending of sands<br />

The grading of sand has a more pronounced effect on workability than has that of the coarse<br />

aggregate. While sand from a single source may not meet the grading requirements, it is<br />

often possible to overcome this difficulty by using a combination of sands.<br />

In the following table the grading is tabulated for the mixture of aggregates of 67 % of the<br />

crusher sand and 33 % of the pit sand:<br />

Table 8:<br />

Grading of mixture of two sands<br />

Sieve size [mm]<br />

Crusher Sand<br />

A<br />

Percent passing<br />

Pit Sand<br />

B<br />

Mixture<br />

67 % A + 33 % B<br />

4,0 82 100 88<br />

2,0 56 100 71<br />

1,0 45 100 63<br />

0,5 33 97 54<br />

0,25 19 58 32<br />

0,125 7 31 15<br />

0,090 3 18 8<br />

FM 3.58 1.14 2.77<br />

A comparison of concrete with crushed aggregates and concrete with rounded aggregates is<br />

shown in the following table:<br />

Table 9:<br />

Properties of fresh concrete with crushed aggregates<br />

Fresh concrete<br />

Crushed aggregate has in comparison with rounded aggregate:<br />

• Better stability of stiff concrete after immediate demoulding<br />

• Comparing flowing properties during vibration<br />

• Comparable properties of pumped concrete<br />

• Less bleeding<br />

• Comparable or better concrete surface<br />

• Consistency (on DIN A flow table) less favourable<br />

Table 10:<br />

Properties of hardened concrete with crushed aggregates<br />

Hardened concrete<br />

Crushed aggregate has in comparison with rounded aggregate:<br />

• Better flexural strength<br />

• Higher grip resistance<br />

• Less depth of water penetration<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 14 of 235


Cement<br />

Concrete<br />

Grading of stone<br />

The grading of coarse material has a smaller influence on the workability of a mix than that of<br />

the sand but haphazard grading cannot be permitted where the quality of the concrete is of<br />

importance. For this reason the material is screened into its various size fractions and<br />

recombined to conform with specified or suitable grading requirements. The following figure<br />

shows the grading limits according to DIN 1045.<br />

Figure 2:<br />

Examples of particle size distribution for different gradings<br />

Pa<br />

ss<br />

in<br />

g<br />

[%<br />

100<br />

80<br />

60<br />

40<br />

20<br />

A - B favorable<br />

B - C usable<br />

U gap<br />

Grading acc. DIN 1045<br />

C<br />

B<br />

A<br />

U<br />

0<br />

0 0.125 0.25 0.5 1.0 2.0 4.0 8.0 16.0 31.5<br />

Sieve size<br />

Gap-graded and continuously-graded concrete<br />

The relative merits of concrete made with gap-graded and continuously-graded stone have<br />

been much debated [10, 11,12]. The technical advantages of each are listed below but, in<br />

comparing merits, it must be stated that economic considerations may outweigh the<br />

technical.<br />

Table 11:<br />

Properties of gap-graded concrete<br />

Gap-graded concrete [2,12]<br />

compared with continuous-graded concrete<br />

• Less danger of particle interference.<br />

• Grater sensivity of consistence to change in water content.<br />

This makes for more accurate control mixing of water which in<br />

turn ensures more consistent strength results.<br />

• Stiffer mixes are more responsive to vibration.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 15 of 235


Cement<br />

Concrete<br />

Table 12:<br />

Properties of continuously-graded concrete<br />

Continuously-graded concrete<br />

compared with gap-graded concrete<br />

• Wetter mixes are less prone to segregation.<br />

• Less sensitive to slight changes in water content. This is an advantage<br />

where uniform workability is important.<br />

• Improve pumpability especially at higher pressure.<br />

• Improve flexural strength due to the increased surface area of graded stone.<br />

1.3.6 Alternative aggregates<br />

Aggregates from natural sources make up the bulk of the aggregates used in concrete.<br />

However, there are a number of alternative materials which can be used as aggregate and<br />

concrete technologists should be encouraged to explore their potential performance and<br />

economy for use in concrete.<br />

Materials used for a greater or lesser extent of natural aggregate are:<br />

♦ Metallurgical slag<br />

Metallurgical slags that have been found suitable for use in concrete include blastfurnace,<br />

ferromanganese, ferro-silicon-manganite, phosphate, chrome, copper and platinum slag<br />

[14].<br />

♦ Clinker<br />

Well-burnt furnace residues from furnaces fired with pulverised fuel (BS 1165).<br />

♦ Expanded Clays (shale and slate)<br />

When certain clays or shale are rapidly heated to the point of incipient in fusion, the<br />

material softens, becomes plastic and tends to entrap gases which are generated within<br />

its mass. Trade names are: Haydite, Cel Seal, Aglite, Gravelite. Leca and Güllät.<br />

♦ Sintered fly ash<br />

In one of the manufacturing processes of low-density aggregate from fly ash, the ash is<br />

first palletised and then formed noodles are sintered at 1'000 to 1'200 °C.<br />

♦ Burnt-clay bricks<br />

Burnt-clay aggregate made by crushing broken bricks is being used more and more to<br />

some extent in cast-in-situ concrete, but mainly for making precast concrete panels and<br />

concrete masonry units.<br />

♦ Colliery Spoil<br />

In the UK in 1972, two plants were operating on colliery waste and producing an<br />

aggregate known as Aglite. It is an expanded low-density aggregate.<br />

♦ Exfoliated Vermiculite<br />

Vermiculite is the geological name given for a group of micaceous minerals. It has the<br />

unique property that when heated to ground 1'200 °C, the flakes expanded (exfoliated)<br />

up to 15 times their original volume.<br />

♦ Perlite<br />

Perlite and certain other volcanic glasses expand when heated to the point of incipient<br />

fusion. Expanded perlite is usually produced only in sand size. It has a maximum bulk<br />

density of 240 kg/m 3 .<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 16 of 235


Cement<br />

Concrete<br />

♦ Glass<br />

Waste glass is a potential aggregate for concrete and has been used in precast concrete<br />

elements. Waste glass is susceptible to alkali-aggregate-reaction and may produce<br />

concrete with a high expansion if used in combination with a high alkali content.<br />

♦ Recycled concrete as aggregate<br />

Depletion of normal aggregate sources and waste disposal problems have made<br />

concrete reclaimed from the demolition of concrete structures an attractive proposition as<br />

aggregate. Useful data have been published by Buck [13].<br />

♦ Sawdust and Wood Wool<br />

Concrete made from mixtures of Portland cement and sawdust is generally a rather<br />

unreliable material and its properties cannot be easily predicted.<br />

♦ Expanded polystyrene<br />

Expanded polystyrene beads have a closed cellular structure and can easily be used to<br />

produce cellular concrete. Expanded polystyrene beads are extremely light (density 12 to<br />

16 kg/m 3 ) so that they can segregate from the mix and are hydrophobic.<br />

♦ Remarks<br />

Some of the alternative aggregates, such as air-cooled metallurgical slag are used<br />

merely because they are more economical than natural aggregates and can be used<br />

without detriment. In other cases, certain technical properties are required (e.g. low<br />

density, superior thermal insulation) and for these purposes aggregates are often<br />

"purpose-made" for example low-density aggregates such as bloated clay and shale,<br />

sintered fly ash and exfoliated vermiculite. Many of these materials are only used to make<br />

precast concrete products such as concrete masonry units.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 17 of 235


Cement<br />

Concrete<br />

1.4 Admixtures<br />

Admixtures are chemicals which are added in relatively small quantities to the basic<br />

constituents of concrete. The quantity of an admixture is usually measured against the<br />

quantity of cement, expressed as its percentage by mass.<br />

It is convenient to make a distinction between admixtures and additives. Admixtures are used<br />

in quantities of maximum 5 wt-% (usually 0.5 to 2.0%). Additives such as fly ash, or other<br />

pozzolanas and blast furnace slag, are applied in quantities of more than 5% by mass of<br />

cement (usually 15 to 35%). Additives should therefore be considered as additional<br />

constituents of the concrete mix.<br />

Admixtures are usually categorised according to their principal uses or according to their<br />

primary effects as follows:<br />

• Improvement of workability of fresh concrete<br />

• Control of setting time and early hardening<br />

• Air-entrainment<br />

• Other effects (stability, high cohesion, colouring, etc.)<br />

The official definition of an admixture according to ASTM C125:<br />

♦ An admixture is a material other than water, aggregates, hydraulic cement, and fibre<br />

reinforcement used as an ingredient of concrete or mortar and added to the batch<br />

immediately before or during its mixing.<br />

This definition encompasses a wide range of materials that are utilised in modern<br />

concrete technology. Subsections of admixtures are:<br />

• Accelerating admixtures<br />

Admixtures that accelerates the setting and early strength development of concrete.<br />

• Air-entraining admixtures<br />

Admixtures that causes the development of a system of microscopic air bubbles in<br />

concrete or mortar during mixing.<br />

• Retarding admixtures<br />

Admixtures that retards the setting of concrete.<br />

• Water-reducing admixtures<br />

Admixtures that either increases the slump of freshly mixed mortar or concrete<br />

without, increasing the water content, or that maintains the slump with a reduced<br />

amount of water due to factors other than air entrainment.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 18 of 235


Cement<br />

Concrete<br />

1.4.1 Workability agents<br />

Admixtures in this category are either called<br />

• plasticizers<br />

• superplasticizers<br />

• water-reducing agents<br />

• high range water-reducing agents.<br />

The primary effect of the admixtures is an improvement of workability.<br />

The following diagram shows the different purposes of using plasticizers in a typical concrete<br />

mix:<br />

Figure 3:<br />

1 constant workability (flow); reduced w/c ratio and increased strength<br />

2 constant w/c ratio and strength; increased workability (flow)<br />

3 mix of both cases<br />

70<br />

65<br />

60<br />

with SP<br />

without SP<br />

Flow acc. DIN [cm]<br />

55<br />

50<br />

45<br />

40<br />

1<br />

3 2<br />

35<br />

30<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75<br />

w/c ratio<br />

Admixtures in this group are all based on chemical compounds which affect the forces<br />

between solid particles suspended in water and reduces the surface tension of water. The<br />

admixtures contain surfactants which is being absorb on the surface of the particles. The<br />

strength of the adsorption depends on the type of particle. In case of Portland cement the<br />

C 3 A compound appears to provide the strongest attraction. The dispersing action is also<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 19 of 235


Cement<br />

Concrete<br />

enhanced by the development of a layer of adsorbed molecules of the plasticizers which<br />

separate the particles of cement; see following figure:<br />

Figure 4:<br />

Cement particle<br />

floc<br />

Particle repulsion<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

_<br />

Flocculated particles of cement<br />

before the addition of a superplasticizer<br />

The negative charges of the 'tails' of<br />

the molecules of the admixture<br />

adsorbed on to the cement particles<br />

generate repulsive forces and<br />

disperse the particles<br />

The superplasticizers are broadly classified into four groups, namely:<br />

• SMF:<br />

• SNF:<br />

• MLS:<br />

• Others:<br />

Sulphonated Melamine-Formaldehyde condensate<br />

Sulphonated Naphthalene-Formaldehyde condensate<br />

Modified Ligno Sulphonates<br />

Polyacrylates, sulphonic-acid-esters, carbohydrate esters etc. A very new<br />

invention is based on derivatives of maleic acid and vinyl monomers.<br />

Variations exist in each of these classes and some formulations may content a second<br />

ingredient.<br />

Ordinary plasticizers<br />

Ordinary plasticizers are mainly based on lignosulphonate salts - a minority use salts of<br />

hydroxy-carboxylic acid. Calcium lignosulphonate and the better water soluble sodium<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 20 of 235


Cement<br />

Concrete<br />

lignosulphonate normally used as admixtures for concrete. The lignosulphonate molecules<br />

are in the form of polymers with molecular weights varying from approx. 20 to 30'000.<br />

Dosage of lignosulphonate type plasticizers varies between 0.2 to 0.6% of the cement in the<br />

mix. The effectiveness of the lignosulphonate plasticizers depends on the composition of the<br />

particular Portland cement used. Cements with high content of C 3 A are likely to require<br />

increase dosage of admixtures. The lignosulphonate admixtures appear to be less effective<br />

for cements with moderate high cement content than for mixes with low cement contents,<br />

below approx. 270 kg/m 3 .<br />

Lignosulphonate based plasticizers have the tendency to produce significant side effects,<br />

namely:<br />

(1) Air-entrainment<br />

The additional percentage of entrained air which can be generated by normal doses<br />

varies between 0.5 to 2.5 vol-% (slump approx. 6 cm, cement content approx. 300<br />

kg/m 3 ).<br />

(2) Retardation<br />

The lignosulphonates themselves interfere with the hydration of cement and cause<br />

some retardation. This effect is more pronounced when the calcium lignosulphonate<br />

admixture is used. High dosages of such admixtures can lead to unacceptable<br />

retardation.<br />

The hydroxy-carbolic plasticizers (mostly sodium salts of citric, tatratic, gluconic, and maleic<br />

acid) are not very common, and hydroxylated polymers are rarely found in concrete<br />

construction practice.<br />

Overdosing the concrete with an ordinary plasticizer increases workability. The increase in<br />

non-linear and the effect varies with the cement content and the initial workability (slump) of<br />

the mix. The following graphic shows the increase of slump as a function of the dosage of an<br />

ordinary Portland cement:<br />

Figure 5:<br />

20<br />

18<br />

16<br />

14<br />

Slump [cm]<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5<br />

Dosage rate of lignosulphonate [wt-%]<br />

Influence of the dosage rate of an ordinary lignosulphonate plasticizer on slump<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 21 of 235


Cement<br />

Concrete<br />

Superplasticizers<br />

Superplasticizers are based mainly on two types of polymers, namely the salts of<br />

formaldehyde naphthalene sulphonate and formaldehyde melamine sulphonate. The term<br />

superplasticizer indicates the much greater potential for increasing workability of concrete<br />

without undesirable side effects when compared to ordinary plasticizers.<br />

The chemical composition of the superplasticizers differ from that of the ordinary plasticizers<br />

in that they do not delay the setting times and hardening of fresh concrete. On the contrary,<br />

some acceleration of the setting and hardening is usually observed. It is not entirely clear if<br />

the acceleration is due simply to a better dispersion of the cement particles or if other<br />

processes are also involved.<br />

The accelerating effect of the superplasticizers appears to be primarily responsible for the<br />

relatively short periods of effectiveness of the admixture. Small additions of purified<br />

lignosulphonates are sometimes added to the admixture to moderate the rapid early loss of<br />

workability.<br />

The periods of effectiveness of the superplasticizers have been quoted to vary between 15<br />

minutes to 40 minutes. The period of effectiveness of the superplasticizer is usually<br />

measured as the time it takes for the workability of the mix to decrease to the level of the<br />

initial slump before the admixtures had been added.<br />

It has been common practice in the ready-mixed concrete industry to delay the first dose of<br />

the superplasticizing admixture until the time the concrete reaches the delivery point. It is<br />

then possible to re-mix the concrete more than once, each time adding another dose of the<br />

superplasticizer and thus maintaining or even increasing workability without detrimental<br />

effect on the properties of hardened concrete.<br />

The chemical nature of the superplasticizer determines its effectiveness in increasing the<br />

workability (slump, flow table spread, etc.). For e.g. to obtain a slump of about 25 cm from an<br />

initial value of 5 cm, it may be necessary to add 0.6% SMF or MLS-based superplasticizer<br />

whereas this could be accomplished with only 0.4% SNF-based admixture. In the following<br />

diagram the loss of workability is shown in a control concrete (containing no admixtures)<br />

compared with that containing MLS (Modified Ligno Sulphonate), SMF (Sulphonated<br />

Melamine-Formaldehyde) and SNF (Sulphonated Naphthalene-Formaldehyde):<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 22 of 235


Cement<br />

Concrete<br />

Figure 6:<br />

Flow table spread [cm]<br />

70<br />

65<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

0 0.25 0.5 0.75 1 1.25 1.5<br />

Time [hours]<br />

MLS<br />

SMF<br />

SNF<br />

Ref<br />

1.4.2 Air-entraining agents<br />

Effect of MLS, SMF and SNF on loss of flow table spread<br />

The air-entraining admixtures are invariably organic substances which helps to generate<br />

microscopic bubbles of air in the fresh concrete. The majority of the commercially available<br />

ones are based on neutralised wood resins (vinsol) in which salts of abietic and pimeric<br />

acids are usual active chemical ingredients. A few air-entrained admixtures are based on the<br />

salts of fatty acids or other organic chemical compounds. It is also possible to use some<br />

types of the lignosulphonate based admixtures in higher dosages to produce simultaneous<br />

improvement of workability and generation of air-entrainment.<br />

The technique of air-entrainment was originally developed in the USA primarily to increase<br />

the resistance of concrete to freezing and thawing and to the scaling effects produced by deicing<br />

salts (mostly sodium or calcium chloride). The entrainment of air in concrete has been<br />

defined as the introduction into fresh concrete of air in controlled amounts and in the form of<br />

properly-sized bubbles (preferably within the size range 0.05 to 0.3 mm diameter).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 23 of 235


Cement<br />

Concrete<br />

The entrained air in concrete should be clearly distinguished from accidentally entrapped air.<br />

The two differ in magnitude, amount, and properties of the air bubbles:<br />

♦ Entrained air<br />

This air is intentionally incorporated by means of chemical admixtures, usually between 3<br />

to 7% by volume and a maximum size of about 0.5 mm; entrained air produces discrete<br />

cavities in the cement paste so that no channels for the passage of water are formed and<br />

the permeability of the concrete is not increased. The voids never become filled with the<br />

products of hydration of cement as gel can form only in water.<br />

♦ Entrapped air<br />

Accidentally incorporated air is called entrapped air. This air usually forms very much<br />

larger bubbles, some as large as the familiar pockmarks on the surface of the concrete.<br />

The amount is in general between 0.5 to 1.5% by volume.<br />

The air-entrained admixtures contain surfactants which are adsorbed on to the surfaces of<br />

the cement grains. As the microscopic bubbles form during mixing of the concrete, they<br />

stabilise in their small sizes and are attracted to the layers of surfactant on the cement<br />

particles. A simplified diagram of the cement-water-air structure is shown in the following:<br />

Figure 7:<br />

-<br />

- - + - -<br />

water<br />

a<br />

-<br />

- c + - -<br />

- - + -<br />

- - - -<br />

+<br />

+ +<br />

-<br />

- -<br />

+<br />

a<br />

- - c<br />

- -<br />

aggregate +<br />

-<br />

a +<br />

--<br />

-<br />

- + +<br />

+<br />

+<br />

-<br />

--<br />

-<br />

entrapped<br />

-<br />

+ -<br />

- -<br />

a<br />

air<br />

a<br />

- -<br />

- -<br />

- +<br />

-<br />

+<br />

+<br />

- -<br />

- -<br />

+ +<br />

- + +<br />

--<br />

+ - +<br />

c<br />

- - -<br />

+ - - +<br />

- + -<br />

- + +<br />

+ c<br />

+ + -<br />

- +<br />

-<br />

- + - a<br />

-<br />

-<br />

-<br />

- a - - -<br />

-<br />

-<br />

- -<br />

- - -<br />

- -<br />

- - -<br />

a<br />

a<br />

- +<br />

-<br />

- +<br />

- -<br />

-<br />

entrapped<br />

+ +<br />

-<br />

aggregate<br />

c<br />

-<br />

air - - +<br />

-+<br />

-<br />

+<br />

-<br />

- - -<br />

- a -<br />

+ + +<br />

a<br />

-<br />

- +<br />

- - - -<br />

c<br />

-<br />

a<br />

bubbles of entrained air<br />

c<br />

particles of cement<br />

Wo-concprop.pre-95<br />

A simplified structure of the cement - water - aggregate - entrained air system<br />

in an air-entrained concrete mix<br />

Air entraining agents should be added as solution, dissolved in the mixing of the concrete. If<br />

other admixtures are also used, the air entraining agent should be added separately rather<br />

than mixed with the other admixtures, because sometimes there are reactions between<br />

materials that result in a decrease in the effectiveness of the air entraining agents.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 24 of 235


Cement<br />

Concrete<br />

The usual dosage rate of these materials is between 0.3 to 1.0 ml (density ≤ 1.1 g/cm 3 ) per<br />

kg of cement. This rate is roughly equivalent to 0.01%, solid admixtures substance to<br />

cement. The dosage rate can vary according to the composition of the mix.<br />

Air entraining agents can be used with cements other than Portland. When used with<br />

blended cements, a larger amount of agent may be required to obtain the desired air content<br />

of the concrete.<br />

1.4.3 Agents affecting setting and hardening<br />

Accelerators<br />

Accelerators have their primary application in cold weather concreting where they may be<br />

used to permit earlier starting of finishing operations. They reduce the time required for<br />

curing, and permit earlier removal of forms or loading of the concrete. Accelerators cannot be<br />

used as antifreeze agents since, at the allowable dosages, the freezing point will be lowered<br />

less than 2 °C.<br />

Accelerating admixtures can be divided in three groups:<br />

(1) Soluble inorganic salts<br />

Most soluble inorganic salts will accelerate the setting and hardening of concrete to<br />

some degree; calcium salts generally being the most effective. Calcium chloride is<br />

the most popular choice because it gives more acceleration at a particular rate of<br />

addition than other accelerators and is also reasonably inexpensive. Admittedly it is<br />

very corrosive (as all water soluble chlorides) against the reinforcement of the<br />

concrete and therefore in many countries forbidden or limited at a particular level.<br />

Soluble carbonates, aluminates, fluorides, and ferric salts have quick-setting properties.<br />

Sodium carbonate and sodium aluminate are the most common ingredients of<br />

shotcreting admixture formulations used to promote quick setting. Calcium fluoroaluminate<br />

can be used as admixtures to obtain rapid-hardening characteristics.<br />

(2) Soluble organic compounds<br />

A variety of organic compound have accelerating properties (although many more act<br />

as retarders), but triethanolamine, calcium formiate, and calcium acetate account for<br />

most commercial uses. They are commonly used in formulations of water-reducing<br />

admixtures to offset their retarding action.<br />

Although triethanolamine is listed as an accelerator, recent research shows that its<br />

reaction with Portland cement is rather complex. It can cause retarding or flash setting,<br />

depending on the amount used.<br />

(3) Miscellaneous solid materials<br />

Retarders<br />

Solid materials are not often used for accelerating. Addition of calcium aluminate cements<br />

cause Portland cements to set rapidly, but strength development is poor. Concrete<br />

can be "seeded" by adding fully hydrated cement that has been finely ground<br />

during mixing to cause more rapid hydration. Finely divided carbonates (calcium or<br />

magnesium), silicate minerals, and silicas are reported to decrease setting time.<br />

Retarders can be used whenever it is desirable to set off the effects of high temperatures<br />

which decrease setting times. Prolonging the plasticity of fresh concrete can be used to<br />

advantage in placing mass concrete. Successive lifts can be blended together by vibration,<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 25 of 235


Cement<br />

Concrete<br />

with the elimination of cold joints that would occur if the first lift were to harden before the<br />

next were placed.<br />

Retarders can also be used to resist cracking due to form deflection that can occur when<br />

horizontal slabs are placed in sections. Concrete that has set but has acquired little strength<br />

is liable to microcrack when subsequent pouring alters the amount of form deflection. If the<br />

plastic period is prolonged, the concrete can adjust to form deflection without cracking.<br />

The composition of retarders can be divided into several categories, based on their chemical<br />

composition:<br />

Lignosulphonic acids and their salts (1)<br />

Hydroxycarboxylic acids and their salts (2)<br />

Sugar and their derivatives (3)<br />

Inorganic salts (4)<br />

It will be noted that categories (1) and (2) also possess water-reducing properties, and these<br />

admixtures can be classified under both groups. Lignosulphonate-based admixtures are<br />

prepared from pulp and paper industrial wastes, and studies have indicated that most of the<br />

retarding properties of these admixtures may be due to compounds that belong to category<br />

(2) or (3). Some inorganic salts (borates, phosphates, and zinc and lead salts) can act as<br />

retarders but are not used commercially.<br />

The influence of an admixture on air entrainment should be considered, particularly if the<br />

admixture also has water-reducing properties. Retarders may also increase the rate of loss<br />

of workability in fresh concrete (slump loss), even when abnormal setting behaviour does not<br />

occur.<br />

Whenever a retarding admixture is used, some reduction in the 1-day strength of the<br />

concrete should be anticipated. Within 7 days, the strength should approach that of an<br />

unretarded concrete unless an overdose has been used (see following diagram):<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 26 of 235


Cement<br />

Concrete<br />

Figure 8:<br />

Compressive strength [MPa]<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Accelerated No admixture (Over-)retarded<br />

2 4 6 8 10 12 14 16 18 20 22 24 26 28<br />

Time [days]<br />

Effect of retarders and accelerators on strength development of concrete<br />

Retarding admixtures have been reported to increase ultimate compressive strength and, to<br />

a lesser extent, flexural strength. Although set-controlling admixtures are reported to<br />

increase drying shrinkage and creep, the effects depend on changes in mix design, time of<br />

hydration, and time of drying or loading.<br />

1.4.4 Other admixtures<br />

There are many other types of admixtures that are commercially available. The consumption<br />

of these various materials added together is less than the amount used of any single type so<br />

far mentioned. Some brief discussion of the more important kinds is warranted, however.<br />

Bonding (or polymer) admixtures<br />

Polymer latex emulsions are used to improve the bonding properties of concrete. This can be<br />

bonding between old and new concrete in repair work, or bonding between concrete and<br />

other materials, such as steel.<br />

The brittle nature of concrete is an inherent property of the material and one that is overcome<br />

by the use of reinforcing materials. The very high porosity of concrete is also a disadvantage.<br />

Several approaches have been taken to improve concrete properties, resulting in quite<br />

different materials. Three different kinds of materials are often used:<br />

PIC: Polymer-impregnated concrete<br />

PIC involves filling the capillary pores of hardened concrete<br />

with a polymer<br />

LMC: Latex-modified concrete<br />

LMC is made by incorporating a polymer latex with fresh<br />

concrete, which im-proves the tensile properties of concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 27 of 235


Cement<br />

Concrete<br />

Corrosion inhibitors<br />

The incorporation of compounds that will prevent or reduce corrosion has been suggested for<br />

concretes in which corrosion of reinforcement can be a problem. Generally, these will be<br />

salts that contain an oxidizable ion, such as nitrites, thiosulphates, benzoates, stannous<br />

salts, and ferrous salts.<br />

It is however, doubtful whether the use of inhibitors is really warranted, since they are not<br />

considered to provide protection in the presence of chloride ion, which is precisely the<br />

situation in which protection is desired.<br />

Dampproofing admixtures<br />

The term 'dampproofing' implies prevention of water penetration into dry concrete or the<br />

transmission of water through concrete. No admixture can actually prevent such movement<br />

of water, although it may reduce the rate at which it occurs.<br />

Certain formulations based on salts of fatty acids (soaps) or petroleum products (mineral<br />

oils and asphalt emulsion) may give the concrete a water-repellent effect. Such materials<br />

have been used to prevent the penetration of rain into porous concrete blocks but are not<br />

likely to affect the performance of dense, well-cured concrete.<br />

Expansive Cements<br />

One of the major disadvantages of Portland cement concrete is its high drying shrinkage and<br />

its susceptibility to tensile strength when volume contraction is wholly or partially restrained.<br />

The development of expansive cements dates back about 60 years. Commercial production<br />

began in the US in the late 1960s, although total production remains quite small, about<br />

500'000 tons annually. A standard specification, ASTM C845-90, covers expansive cements.<br />

All three variants of present-day expansive cements (Type K, M, and S) are based on the<br />

formation of ettringite (hydrated calcium sulphoaluminate) in considerable quantities during<br />

the first few days of hydration. The material from which ettringite is formed differ substantially<br />

in each cement, but all require a source of calcium aluminate and sulphate ions. Expansion<br />

has also been achieved by using clinker with high content of free lime as admixture to cause<br />

expansion of a superplasticized concrete mix.<br />

Successful use of expansive cements depends upon proper control of the expansion of the<br />

cement during hydration and is sensitive to a number of variables, for example:<br />

Reinforcement -- Mix design -- Handling and curing<br />

Grouting admixtures<br />

Cement-based grout used for speciality applications such as cementing oil well contain<br />

different admixtures. Flocculating admixtures, thickeners, and mineral admixtures are used to<br />

prevent bleeding and segregation and to increase cohesion and retention of water during<br />

pumping. Retarders are commonly used to extend pumping times.<br />

1.5 Mixing and curing water for concrete<br />

1.5.1 Introduction<br />

The quality of water is important because poor-quality water may affect the time of setting,<br />

strength development, or cause staining. Almost all natural water, fresh water, and water<br />

treated for municipal use are satisfactory as mixing water for concrete if they have no<br />

pronounced odour or taste.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 28 of 235


Cement<br />

Concrete<br />

Because of this, very little attention is usually given to the water used in concrete, a practice<br />

that is in contrast to the frequent checking of admixture, cement, and aggregate components<br />

of the concrete mixture.<br />

1.5.2 Effect of impurities<br />

A popular criterion as to the suitability of water for mixing concrete is the classical<br />

expression:<br />

If water is fit to drink<br />

it is all right for making concrete<br />

This does not appear to be the best basis for evaluation, since water containing small<br />

amounts of sugar or citrate flavouring would be suitable for drinking but not mixing concrete,<br />

and, conversely water suitable for making concrete may not necessarily be fit for drinking.<br />

The most extensive series of tests on the subject "Impurities in mixing water" was conducted<br />

by Abrams (ACI Proceedings, Vol 20, 1924, 442-486). Approximately 6000 mortar and<br />

concrete specimens representing 68 different water samples were tested in this investigation.<br />

Among the water tested were sea and alkali water, bog water, mine and mineral water, and<br />

water containing sewage and solutions of salt. Tests with fresh and distilled water were<br />

included for comparative purposes. Some of the more significant conclusions based on these<br />

data are as follows:<br />

The time of setting of Portland cement mixtures containing impure mixing water was about<br />

the same as those observed with clean fresh water with only few exceptions. In most<br />

instances, the water giving low relative compressive strength of concrete caused slow<br />

setting.<br />

♦ Non of the waters caused unsoundness of the neat Portland cement paste when tested<br />

over boiling water.<br />

♦ In spite of the wide variation in the origin and type of water used, most of the samples<br />

gave good results in concrete due to the fact that the quantities of injurious impurities<br />

present were quite small.<br />

♦ The quality of mixing water is best measured by the ratio of the 28-day concrete or mortar<br />

strength to that of similar mixtures made with pure water. Water giving strength ratios that<br />

are below 85% should be considered unsatisfactory.<br />

♦ Neither odour nor colour is an indication of quality of water for mixing concrete. Water<br />

that was most unpromising in appearance gave good results. Distilled waters gave<br />

concrete strength essentially the same as other fresh waters.<br />

♦ Based on a minimum strength-ratio of 85% as compared to that observed with pure<br />

water, following samples were found to be unsuitable for mixing concrete:<br />

• acid water<br />

• lime soak water from tannery waste<br />

• carbonated mineral water discharged from galvanising plants<br />

• water containing over 3% sodium chloride or sulphate<br />

• water containing sugar or similar compounds<br />

The concentration of total dissolved solids in these waters was over 6'000 ppm<br />

except for the highly carbonated water that contained 2'100 ppm total solids. Very few<br />

natural waters other than sea water, contain more than 5'000 ppm of dissolved solids.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 29 of 235


Cement<br />

Concrete<br />

♦ Based on the minimum strength-ratio of 85%, the following waters were found to be<br />

suitable for mixing concrete:<br />

• bog and marsh water<br />

• waters with a maximum concentration of 1% sulphate<br />

• sea water (but nor for reinforced concrete)<br />

• alkali water with a maximum of 0.15% Na 2 SO 4 or NaCl<br />

• water from coal and gypsum mines<br />

• waste water from slaughterhouses, breweries, gas plants,<br />

and paint and soap factories.<br />

1.5.3 Effects of algae on air content and strength<br />

A rather extensive series of laboratory tests showed that the use of water containing algae<br />

had the unusual effect of entraining considerable quantities of air in concrete mixtures with<br />

an accompanying decrease in strength. The data in the following table were based on tests<br />

with 19-mm maximum-size aggregate concrete having a w/c ratio of 0.5 and a slump of 4.0<br />

to 7.5 cm, with a constant ratio of coarse to fine aggregate:<br />

Table 13:<br />

Algae in mixing water [%] Air in concrete [%] Compr. strength at 28 days [MPa]<br />

none (control mix) 2.2 33.3<br />

0.03 2.6 33.4<br />

0.09 6.0 27.9<br />

0.15 7.9 22.8<br />

0.23 10.6 17.8<br />

In addition to the detrimental effect of strength, one of the important aspects of these data is<br />

that considerable quantities of air can be entrained in concrete by the use of mixing water<br />

containing algae.<br />

1.5.4 Curing water<br />

There are two primary considerations with regard to the suitability of water for curing<br />

concrete:<br />

♦ One is the possibility that it might contain impurities that would cause staining, and the<br />

other<br />

♦ that it might contain aggressive impurities that would be capable of attacking or causing<br />

deterioration of concrete. The latter possibility is unlikely, especially if water satisfactory<br />

for use in mixing concrete is employed.<br />

In some instances the staining or discoloration of the surface of concrete from curing water<br />

would not be objectionable.<br />

The most common cause of staining is usually a relatively high concentration of iron or<br />

organic matter in the water; however, relatively low concentrations of these impurities may<br />

cause staining, especially if the concrete is subjected to prolonged wetting by run off of<br />

curing water from portions of the structure.<br />

Test data show that there is no consistent relationship between dissolved iron content and<br />

degree of staining. In some cases, 0.08 ppm of iron resulted in only a slight discoloration and<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 30 of 235


Cement<br />

Concrete<br />

in other cases, waters with 0.06 ppm of iron gave a moderate rust-coloured stain, while 0.04<br />

ppm produced considerable brownish-black stain.<br />

2 Literature for admixtures and mixing water<br />

♦ P. Klinger, J. Lamond: Concrete and Concrete-Making Materials; ASTM, STP 169C, Aug.<br />

1994<br />

♦ P. Bartos: Fresh Concrete, Properties and Tests; Elsevier, Amsterdam, 1992<br />

♦ K. Wesche: Baustoffe für tragende Bauteile, Teil 2: Beton, Mauerwerk; Bauverlag,<br />

Wiesbaden, 1992<br />

♦ A. Neville: Properties of Concrete; Longman Scientific & <strong>Technical</strong>, New York, 1991<br />

♦ B. Addis: Fulton's Concrete Technology: Portland Cement Institute, Midland, South<br />

Africa, 1986<br />

♦ S. Mindess: Concrete; Prentice-Hall, Englewood Cliffs, New Jersey, 1981<br />

3 Appendices<br />

3.1 References<br />

[1] DIN 4226 (1983): Zuschlagstoff für Beton: Deutsches lnstitut für Normung e.v.; Beuth<br />

Verlag GmbH, Berlin<br />

[2] B. Addis et al.: Fulton's Concrete Technology, chapter 3, PCI, Midland, South Africa,<br />

1986; (All extracts mainly from this paper)<br />

[3] R. Rhoades et al.: Petrographic and mineralogic characteristics of aggregates; ASTM<br />

symp. on Mineral Aggregates, Philadelphia, 1948, p. 20 - 48<br />

[4] British Standard 812 (1967): Mineral, Aggregates, Sand & Fillers, Brit. Stand. House,<br />

2 Park Street, London, W1Y 4AA<br />

[5] J. Phemister et al.: Roadstone, geol. aspects and phys. tests; London, Dept. of Sc.<br />

and And. Res. Road Spec. Report No. 3,1946<br />

[6] D. Davis, M. Alexander Properties of aggregate in concrete; Hippor Quarries Techn.<br />

Publ. (1992), 94 Rivonia Road, Sandton 2199, South Africa<br />

[7] M. Kaplan: The flexural and compressive strength of concrete as affected by the<br />

properties of coarse aggregates; Proc. America Concrete Institute v. 55, May 1959,<br />

p. 1193 ff<br />

[8] L. Roberts et al.: Dictionary of Geological Terms; American Geological institute,<br />

Anchor Books, New York (1983)<br />

[9] Webster's New World Dictionary of Amedcan English, Third College Edition,<br />

Cleveland, USA (1988)<br />

[10] L. Mercer et al.: The law of grading for concrete aggregates; Melbourne, Techn. Coll.<br />

Press, 1951, Research Bull. no. 1<br />

[11] V. Barnher Gap-graded concrete; London Cement & Concrete Association, 1952, C<br />

& CA Library Translation no. 42<br />

[12] H. Schäffer Beton mit Ausfallkörnungen; Betonwerk + Fertigteil-Technik, Heft 6/1979<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 31 of 235


Cement<br />

Concrete<br />

[13] A. D. Buck: Recycled concrete as source of aggregate; Proc. American Concrete<br />

Institute, v. 74, 1977, p. 212-219<br />

[14] F. Rossouw: Report on the suitability of some metallurgical slags as aggregate for<br />

concrete; Pretoria, NBRI, 1981, NBRI Special Report Bou 56, p. 1 - 25<br />

3.2 Relevant ASTM Standards<br />

3.2.1 Aggregate<br />

C 33-90<br />

C 40-84<br />

C 70-79<br />

C 88-90<br />

C 123-90<br />

C 131-89<br />

Specification for Concrete Aggregates<br />

Test for Organic Impurities in Sands for Concrete<br />

Test for Surface Moisture in Fine Aggregate<br />

Test for Soundness of Aggregates by Use of Sodium Sulphate or Magnesium<br />

Sulphate<br />

Test for Lightweight Pieces in Aggregates<br />

Test for Resistance to Degradation of Small-Size Coarse Aggregate by<br />

Abrasion and impact the Los Angeles Machine<br />

C 227-87 Test method for Potential Alkali Reactivity of Cement-Aggregate<br />

Combinations (Mortar-Bar Method)<br />

C 289-87<br />

C 294-86<br />

C 330-89<br />

Test method for Potential Reactivity of Aggregates (Chemical Method)<br />

Descriptive Nomenclature of Constituents of Natural Mineral Aggregates<br />

Specification for Lightweight Aggregates for Structural Concrete<br />

C331-89 Specification for Lightweight Aggregates for Concrete Masonry Units<br />

C 332-87<br />

C 586-86<br />

C 637-84<br />

C 638-84<br />

E 11-87<br />

Specification for Lightweight Aggregates for Insulating Concrete<br />

Test method for Potential Alkali Reactivity of Carbonate Rocks for Concrete<br />

Aggregates (Rock Cylinder Method)<br />

Specification for Aggregates for Radiation-Shielding Concrete<br />

Descriptive Nomenclature of Constituents of Aggregates for Radiation<br />

Shielding Concrete<br />

Specification for Wire Cloth Sieves for Testing Purposes<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 32 of 235


Cement<br />

Concrete<br />

Fresh and Hardened Concrete<br />

1. GENERAL.........................................................................................................................35<br />

2. DEFINITION ......................................................................................................................35<br />

2.1 Defintion of concrete ................................................................................................ 35<br />

2.2 Composition of concrete .......................................................................................... 36<br />

2.3 Green and young concrete, workability and resistance to loading of<br />

concrete ................................................................................................................... 37<br />

2.4 Strength of concrete................................................................................................. 37<br />

3. FRESH CONCRETE ......................................................................................................... 38<br />

3.1 Workability................................................................................................................ 38<br />

3.2 Influence on the Workability ..................................................................................... 40<br />

3.2.1 Effect of water ............................................................................................ 40<br />

3.2.2 Effect of solid constituents ......................................................................... 40<br />

3.2.3 Effects of admixtures.................................................................................. 41<br />

3.2.4 Effect of temperature.................................................................................. 41<br />

3.3 Other properties of fresh concrete ........................................................................... 41<br />

3.3.1 Bulk density (unit weight) ........................................................................... 41<br />

3.3.2 Air content .................................................................................................. 42<br />

3.4 From fresh to hardened concrete - Stages of development of concrete .................. 42<br />

3.5 Properties of ‘Green’ and ‘Young’ Concrete............................................................. 43<br />

4. HARDENED CONCRETE............................................................................................... 181<br />

4.1 General .................................................................................................................. 181<br />

4.2 Strength.................................................................................................................. 183<br />

4.3 Deformation under load: modulus of elasticity and creep ...................................... 184<br />

4.4 Density (weight of volume unit) .............................................................................. 184<br />

4.5 The durability of concrete....................................................................................... 184<br />

4.5.1 Volume changes ...................................................................................... 185<br />

4.5.2 Volume changes due to physical stress................................................... 187<br />

4.5.3 Volume changes due to alterations of the substance............................... 187<br />

5. FACTORS INFLUENCING THE STRENGTH OF CONCRETE ..................................... 190<br />

5.1 Influence of the constituents .................................................................................. 191<br />

5.1.1 Aggregates............................................................................................... 191<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 33 of 235


Cement<br />

Concrete<br />

5.1.2 Water........................................................................................................ 192<br />

5.1.3 Cement..................................................................................................... 115<br />

5.2 Influence of mix proportions ................................................................................... 115<br />

5.2.1 Water / Cement ratio ................................................................................ 116<br />

5.2.2 Cement content........................................................................................ 116<br />

5.2.3 Aggregate/Cement ratio ........................................................................... 117<br />

5.3 Influence of handling .............................................................................................. 117<br />

5.4 Influences of curing ................................................................................................ 117<br />

5.4.1 Moisture ................................................................................................... 118<br />

5.4.2 Temperature............................................................................................. 119<br />

5.5 Influences of testing methods ................................................................................ 119<br />

6. CONCLUSIONS.............................................................................................................. 120<br />

7. SUPPLEMENTARY LITERATURE (BOOKS):............................................................... 120<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 34 of 235


Cement<br />

Concrete<br />

1 General<br />

The properties of a building material are of great importance with respect to the function for<br />

which the building was intended. Besides the required load capacity and durability, which are<br />

the main conditions, a building must protect against cold, heat, rain, wind, etc. A comfortable<br />

and pleasant living atmosphere should be created, which cannot be achieved simply by the<br />

design of the building or the application of the specific material, but by the favorable<br />

combination of several materials. The correct choice of material, however, can only be made<br />

if everybody involved with the construction has good knowledge of the material properties.<br />

The properties of building materials such as bricks and steel will remain unchanged<br />

throughout the building process, while the final properties of concrete will only become<br />

established after placing. Cement is only an intermediate product of a building material.<br />

New developments in concrete technology and building technique, based on many years of<br />

research, have made it possible to build such constructions as the 550 m tall TV tower in<br />

Toronto, prestressed concrete bridges with a span of up to 240 m and drilling platforms in the<br />

ocean. Flowing concrete has changed the method of placing. So far, concrete has been able,<br />

in most cases, to compete with other building materials.<br />

All of these constructions and techniques have been made possible by the improvement of<br />

cement quality and on the methods of its application. Should further material requirements<br />

arise, would it be possible to adapt the concrete properties? Can the cement manufacturer<br />

contribute? He should at least make an attempt to meet any new requirements, because<br />

each successful concrete construction will favorably affect the cement production.<br />

On the other hand, the cement manufacturer must be able to defend himself if defects in<br />

concrete are unjustly attributed to the cement. He can only be persuasive if he has thorough<br />

knowledge of concrete properties and concrete technology.<br />

2 Definition<br />

2.1 Defintion of concrete<br />

ACI has defined concrete as follows:<br />

Concrete is a composite material<br />

consisting essentially of a binding medium,<br />

within which particles or fragments of aggregate are<br />

embedded.<br />

In Portland cement concrete, the binder is a mixture of<br />

Portland cement and water.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 35 of 235


Cement<br />

Concrete<br />

2.2 Composition of concrete<br />

Fig. 1: Range of Proportions of materials used in concrete (by weight)<br />

Stages of concrete<br />

Coarse aggregate<br />

Fine aggregate<br />

Cement<br />

Water<br />

Air<br />

Admixture<br />

0 5 10 15 20 25 30 35 40 45 50 55 60<br />

Minima/maxima proportions (wt-%)<br />

The properties of concrete depend very much on its age. Roughly three stages can be<br />

distinguished.<br />

Fresh<br />

(a workable mass)<br />

<br />

concrete<br />

Transition<br />

through intermediate stages<br />

(sometimes called ‘green’ or ‘young’ concrete)<br />

<br />

Hardened<br />

concrete<br />

(an artificial stone which has reached the required<br />

properties for a specific structure)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 36 of 235


Cement<br />

Concrete<br />

The transition from fresh to hardened concrete is a continuous process during which it<br />

changes from a workable mass to an artificial stone. The performance of the concrete at<br />

certain ages has a great influence on its applicability, and, therefore, it is of interest to know<br />

the concrete properties not only when placed (fresh) or used (hardened), but sometimes also<br />

during the entire development.<br />

2.3 Green and young concrete, workability and resistance to loading<br />

of concrete<br />

The concrete is called ‘green’ as soon as it is compacted in the framework until its<br />

solidification by setting. When the concrete turns solid, it is called ‘young’ until it reaches a<br />

certain degree of strength permitting the removal of the form.<br />

For the practical behaviour of concrete in service, its mechanical properties are decisive. Its<br />

resistance to loading is essential in the hardened stage. It is, however, insignificant in fresh<br />

concrete where the workability is of main concern.<br />

Workability is not well defined and, therefore, not directly measurable, because it includes a<br />

certain number of fresh properties. As an example, two definitions of workability from the<br />

same country (USA) follow:<br />

♦ WORKABILITY is that property of freshly mixed concrete or mortar which determines the<br />

ease and homogeneity with which it can be mixed, placed and compacted and finished.<br />

(ACI definition)<br />

♦ WORKABILITY is that property of concrete which determines the effort required to<br />

manipulate a freshly mixed quantity of concrete with minimum loss of homogeneity.<br />

(ASTM definition)<br />

Resistance to loading is the ability of the hardened concrete to bear the service load (dead<br />

and live load); (dead load = constant load in structures due to the mass of the members, the<br />

supported structure and permanent attachments or accessories; live load = any load that is<br />

not permanently applied to the structure).<br />

Resistance to loading can be specified and measured exactly for different types of loads.<br />

Usually, it is expressed as strength.<br />

2.4 Strength of concrete<br />

Strength is defined as maximum resistance to load that a member or structure is capable to<br />

develop before failure occurs. It is measured with reference to the cross section of the<br />

structure member.<br />

F ⎡ N ⎤<br />

Strength : σ = ⎢ orMPa<br />

2<br />

A<br />

⎥<br />

⎣mm<br />

⎦<br />

where: F = applied load<br />

A = cross section on which the load is applied<br />

We distinguish different types of strength according to the type of load exerted:<br />

Compressive, tensile, flexural splitting,<br />

shear, torsion adhesive, and impact<br />

strength.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 37 of 235


Cement<br />

Concrete<br />

The loading resistance of various structural members results from the combination of<br />

different types of strength, according to the actual stresses, and from concrete quality.<br />

3 Fresh concrete<br />

Fresh Concrete is the product that is obtained immediately after mixing the components.<br />

3.1 Workability<br />

The homogenized concrete mix, after it is taken from the mixer, must be suited to be<br />

transported to the destination point and to be placed into the moulds or formwork. The fresh<br />

concrete must completely fill the mould, even if the shape of the elements is very<br />

complicated and the interstices between the reinforcement are very small. A good<br />

compaction must be achieved and through all these steps the mix must remain<br />

homogeneous without segregation.<br />

The ability of fresh concrete to meet all these requirements is called workability which<br />

includes a number of more or less defined properties (see following figure).<br />

Fig. 2: Properties of fresh concrete related to workability<br />

Workability<br />

ease of the mixing<br />

transportability<br />

compactibility<br />

consistency (flow)<br />

stability<br />

mobility<br />

compactibility<br />

finishability<br />

viscosity of paste<br />

no segregation<br />

bleeding<br />

water retention<br />

no honeycombing<br />

friction of aggregate<br />

air content<br />

The workability influences not only all operations of placing and consolidating fresh concrete<br />

but also to a large extent, the quality of the hardened concrete. A dense concrete must<br />

contain a high amount of solid matter and very few voids filled with water, vapor or air. Very<br />

important concrete properties depend on the density.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 38 of 235


Cement<br />

Concrete<br />

Fig. 3: Effect of the workability on fresh and hardened concrete properties<br />

WORKABILITY<br />

FRESH CONCRETE<br />

Easy mixing<br />

Easy transportation<br />

(no segregation)<br />

Easy placing<br />

(no segregation)<br />

Easy compaction<br />

(minimum of voids)<br />

HARDENED CONCRETE<br />

Density<br />

Mechanical properties<br />

Durability<br />

Impermeability<br />

Appearance<br />

The workability requirements vary from one country to another. They depend on the level of<br />

the building industry, on the quality requirements and on the quality of the available<br />

materials. Furthermore, tradition, economical and subjective factors influence the technical<br />

demands.<br />

To characterize the workability of concrete mixes, the following terms are usually used in<br />

Central Europe:<br />

Terms for workability:<br />

very stiff<br />

stiff<br />

plastic<br />

soft (wet)<br />

flowing (liquid)<br />

However, those terms are not clearly defined. In England for instance, completely different<br />

terms are used:<br />

♦ Low, medium, and high consistency.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 39 of 235


Cement<br />

Concrete<br />

3.2 Influence on the Workability<br />

The workability of concrete is influenced in various ways by its solid and liquid components<br />

and environmental conditions.<br />

3.2.1 Effect of water<br />

The workability is related to the water content that is available for the lubrication of a mix. By<br />

increasing the water content in the mix up to a certain limit, the workability of the mix can be<br />

improved, i.e. the mix will be more wet. If the water content exceeds a certain amount, there<br />

is the danger of segregation.<br />

The water content of the mix is expressed in liters or kilograms of water per cubic meter of<br />

concrete (pound per cubic yard).<br />

3.2.2 Effect of solid constituents<br />

Solid matter (aggregate + cement) shows a double effect on the workability:<br />

♦ The shape of the solid grains influences the mobility of concrete in the following manner:<br />

• Coarse particles impede the mobility by their angularity and friction;<br />

• The fine grains, on the other hand, improve it because they act almost like ballbearings<br />

between the coarse grains. This action is distinctly noticeable when lime or<br />

fly ash are added to concrete: their spherical particles improve concrete consistency.<br />

♦ By absorption of the surface, eventually also by chemical bond, part of the mixing water<br />

is fixed. Only the remaining ‘mobile’ part influences concrete consistency.<br />

The effect of aggregate properties on workability is considerable.<br />

Note:<br />

Grain size distribution, angularity and surface texture of the<br />

aggregates significantly affect concrete consistency.<br />

The effect of cement properties on workability is very small<br />

The effect must be attributed to the water requirement of cement which can be determined<br />

with standard methods on the cement paste. The chemical-mineralogical composition has a<br />

stronger influence on the water requirement than fineness and grain size distribution. Grain<br />

size distribution shows only little variation in industrial cements.<br />

If the amount of cement added to a concrete mix is increased at a constant water content, a<br />

stiffening action is the result because more water is absorbed. If, however, both cement and<br />

water are increased, so that the w/c ratio remains unchanged, the lubricating action of the<br />

additional cement paste increases the fluidity.<br />

Thus, the workability of a given aggregate mix can be improved in two ways: by increasing<br />

the water content only, or by increasing the dosage of both, water and cement. In the first<br />

case the quality of hardened concrete will be inferior; in the second case the quality is<br />

maintained but at higher costs.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 40 of 235


Cement<br />

Concrete<br />

3.2.3 Effects of admixtures<br />

Some admixtures such as plastifiers, water reducers and superplastifiers, can modify the<br />

concrete consistency even if added in small quantities (see chapter ‘Concrete Main<br />

Components’, paragraph 3.).<br />

3.2.4 Effect of temperature<br />

Increased temperature of the concrete and surrounding air contributes towards a stiffer<br />

consistency.<br />

Fig. 4: Effect of temperature on the consistency (slump)<br />

12<br />

11<br />

10<br />

Slump (cm)<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

0 5 10 15 20 25 30 35 40<br />

Temperature (°C)<br />

The consistency in the above mentioned figure was tested immediately after the mixing of<br />

the concretes according to the same mix design. Only the temperature of the materials<br />

(cement + aggregate + water) and the ambient temperature were changed.<br />

3.3 Other properties of fresh concrete<br />

3.3.1 Bulk density (unit weight)<br />

The unit weight of fresh concrete gives some indication about the final void content which is<br />

responsible for final concrete properties. It is also a measure of the yield; i.e. it indicates the<br />

concrete volume produced with a specific amount of cement. The bulk density is determined<br />

by weighing a defined concrete volume. It depends very much on the water and less on the<br />

cement dosage. After compaction, the unit weight of ordinary concrete should be higher than<br />

2300 kg/m 3 .<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 41 of 235


Cement<br />

Concrete<br />

3.3.2 Air content<br />

Ordinary concrete which is well compacted, has an air content of about 0.5 to 1.5%. The<br />

measuring of the air content is important if entrained air is used. Air-entrained concrete is<br />

produced by using either an air-entrained cement or an air-entraining admixture during the<br />

mixing of concrete. Air-entrained concrete contains 3 to 7 Vol. % pores. - Entrained air<br />

improves the workability of fresh concrete.<br />

3.4 From fresh to hardened concrete - Stages of development of<br />

concrete<br />

‘Green’ and ‘young’ concrete are the terms for specific intermediary stages during the<br />

transition of concrete from a workable, more or less plastic mass, to an artificial stone.<br />

With the development of modern concrete technology and building methods, the properties<br />

of concrete at these intermediate stages have become important as well. The continuous<br />

development of hardened concrete is a sequence of periods of:<br />

Concrete is a process:<br />

Stiffening → Setting → Hardening<br />

These periods are characterized by different growth rates of the various mechanical<br />

properties of the concrete (see following figure).<br />

Fig. 5: Stages of development of concrete<br />

Stiffening<br />

of concrete is a change in the workability (sometimes also called slump loss) and begins<br />

immediately after mixing, sometimes even during mixing.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 42 of 235


Cement<br />

Concrete<br />

This can have unfavourable consequences if fresh concrete is not placed properly after<br />

mixing, or, especially if ready mix concrete is used, in precast manufacturing, or on large<br />

building sites and placing has to be delayed.<br />

Concrete should be placed as soon as possible after mixing, but often there may be a time<br />

lapse of up to one hour or even more between mixing and placing. In these cases, too rapid<br />

stiffening can impede placing.<br />

The time during which concrete is still workable - at the right<br />

consistency - is very important for transporting and placing.<br />

This time depends on the purpose for which concrete is used and on<br />

the method of placing.<br />

Setting<br />

is the beginning of the hydration process and is indicated by the transition from the plastic to<br />

the solid state within a relatively short period of time.<br />

The resistance against deformation during setting increases rapidly, strength less rapidly.<br />

Note:<br />

The time of setting of concrete is not identical with that of cement<br />

measured on paste according to standards.<br />

The reason is that the setting of concrete is influenced not only by the setting behavior of the<br />

cement, but also by varying cement and water contents, by temperature and type of<br />

construction.<br />

Until setting, deformations of concrete are almost totally irreversible; the elastic part of<br />

deformations becomes predominant only after setting.<br />

Hardening<br />

is the subsequent improvement of mechanical properties of concrete after setting. Soon after<br />

setting, the strength begins to increase more rapidly than the elasticity (Young-modulus). The<br />

rate of increase reaches its maximum between 5 and 20 hours after the addition of water.<br />

Green concrete:<br />

3.5 Properties of ‘Green’ and ‘Young’ Concrete<br />

The mechanical properties of green concrete are important in practice, especially in the<br />

precast industry, where blocks, tubes and other concrete products are moulded by pressing.<br />

The moulded pieces have to be removed from the press as soon as possible in order to<br />

make room for the next series. However, this is only possible if the cohesion of the concrete<br />

is high enough, even before setting, to bear its own weight without deformation. Sometimes<br />

the moulded pieces (e.g. blocks) are stored in several layers. In this case, the strength after<br />

demoulding will have to be higher, so as to carry the whole load.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 43 of 235


Cement<br />

Concrete<br />

Young concrete:<br />

Since young concrete at early age very often has to support a mechanical stress, it must<br />

have a minimum strength. Thus, the strength at this stage (after some hours) is decisive in<br />

the precast industry, where it determines the intervals at which precast elements can be<br />

demoulded, transported and piled up. It is also important in other building processes, for<br />

instance with prestressed concrete. The strength requirements vary according to the size<br />

and shape of the concrete element and the building method.<br />

Early strength:<br />

The compressive strength during the first few hours after placing is<br />

called early strength.<br />

It must reach approx. 10 to 20 MPa for demoulding in reinforced<br />

concrete and<br />

35 MPa for the release of the prestressing wires in the pretensioned,<br />

prestressed concrete.<br />

4 Hardened concrete<br />

4.1 General<br />

Hardened concrete is the final building material as it is obtained after stiffening, setting and<br />

hardening of fresh concrete. Hardening, however, continues for many years, slowing down<br />

after a certain strength has been reached. Therefore, it is generally accepted to consider the<br />

properties of hardened concrete at the age of 28 days as characteristic. They are commonly<br />

used for the design of concrete structures.<br />

Hardened concrete must meet various requirements and maintain its properties during a very<br />

long time. Most important are the load-bearing properties and the durability of concrete.<br />

The resistance to loading of concrete depends on its strength. But a building can never be<br />

loaded to the strength limit. Structural members are designed in such a way that the<br />

calculated stresses in concrete do not exceed certain permissible working values. According<br />

to the function of the building, other special requirements may be of importance.<br />

A large variety of concrete properties can be obtained by the choice of the components and<br />

the mix design. However, the nature of concrete, the availability of the materials and<br />

economical considerations limit these possibilities.<br />

4.2 Strength<br />

According to the type and direction of load and stress, various types of strength react in<br />

concrete:<br />

♦ Compressive strength:<br />

Compressive strength is the most important characteristic of hardened concrete. High<br />

compressive strength is, in most cases, accompanied also by an improvement of the<br />

other properties. compressive strength determinations show the best reproducibility.<br />

Therefore, compressive strength is considered as a general measure of concrete quality.<br />

The compressive strength of commonly used concrete after 28 days is in the range of 10<br />

to 70 N/mm 2 . The low values (∼ 10 to 20 N/mm 2 ) are used for plain concrete, the medium<br />

values (∼ 20 to 45 N/mm 2 ) for reinforced concrete and the high values (∼ 50 to 70 N/mm 2 )<br />

for prestressed concrete and precast elements.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 44 of 235


Cement<br />

Concrete<br />

♦ Tensile strength:<br />

Tensile strength is relatively low and amount to only approx. 1/10 of the compressive<br />

strength of hardened concrete. Because the tensile strength is low, it is practically not<br />

taken into consideration for structural design. Tensile stresses in the construction are<br />

carried by steel reinforcements.<br />

♦ Flexural strength:<br />

It is difficult to determine the tensile strength of concrete and therefore bending or flexural<br />

strength is measured.<br />

♦ Impact strength;<br />

Impact strength plays a significant role in special applications, e.g. for piles to be driven<br />

into the ground.<br />

4.3 Deformation under load: modulus of elasticity and creep<br />

Even momentary loads cause deformation of concrete. If they exceed certain limits, there is<br />

a risk of cracking.<br />

Note:<br />

Concrete is a plastic-elastic material<br />

Its deformation is always composed of the two components: elastic and plastic.<br />

The elastic deformation disappears when the load is removed, the plastic deformation<br />

remains.<br />

Deformation<br />

Elastic deformation<br />

REVERSIBLE<br />

Plastic deformation<br />

IRREVERSIBLE<br />

In hardened concrete, the main part of deformation is elastic.<br />

A quantitative measure of elasticity is the ratio between stress and corresponding strain. This<br />

ratio is termed modulus of elasticity E (Young modulus = initial tangent modulus).<br />

σ ∆L<br />

E= ε =<br />

ε<br />

[ GPa] [] 1<br />

L 0<br />

The modulus of elasticity may be measured in tension, compression or shear. The modulus<br />

in tension is usually equal to the modulus in compression and is frequently referred to as<br />

Young Modulus of elasticity (Table).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 45 of 235


Cement<br />

Concrete<br />

Table 1: Modulus of Elasticity of Different Building Materials<br />

Material<br />

Modulus of elasticity<br />

[GPa]<br />

Steel 200 to 230<br />

Aluminium 74<br />

Copper 130<br />

Natural stone 12 to 80<br />

Concrete 20 to 50<br />

Mortar 5 to 20<br />

Timber 6 to 15<br />

Due to the fact that concrete is neither ideally plastic nor ideally elastic material, the manner<br />

in which its modulus of elasticity is defined is somewhat arbitrary. Various forms of the<br />

modulus, which are used, are illustrated on the stress-strain curve in following diagram:<br />

Fig. 6: Typical stress-strain curve for concrete<br />

Initial tangent (= Young)<br />

modulus<br />

Stress (MPa)<br />

Secant modulus<br />

Unloading<br />

Tangent<br />

modulus<br />

Strain (%)<br />

It also has to be distinguished between the ‘dynamic E-modulus’ and the ‘static E-modulus’.<br />

The dynamic E-modulus is applied when concrete is exposed to oscillation. It can be<br />

calculated, for instance, from the rate of propagation of ultrasonic impulses and is a measure<br />

for the progression of deformation resistance. Using certain formulas, the compressive<br />

strength can be estimated from these results.<br />

Creep:<br />

Plastic deformation becomes more pronounced with longtime loading. The irreversible<br />

deformation by longtime loading is called creep.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 46 of 235


Cement<br />

Concrete<br />

Definition of plastic deformation:<br />

The modulus of deformation is the ratio between the applied load and<br />

the irreversible deformation, expressed as function of time.<br />

The creep is strongly influenced by the stress/strength ratio. With higher stress/strength ratio<br />

the creep increases. It also depends on other factors, such as temperature, humidity, etc.<br />

Creep attains approx. 1.5‰ per 1 MPa load during 1 year. Besides being a disadvantage,<br />

creep is also useful because it diminishes internal stresses. On the other hand, it reduces the<br />

effect of prestressing.<br />

The deformability of concrete has certain limits:<br />

♦ If the stress grows beyond these limits, the concrete cracks or breaks.<br />

♦ Strength and resistance against deformation dictate the limits:<br />

• The higher the strength and the lower the resistance against deformation, the higher<br />

the limits of deformation.<br />

σ<br />

lim<br />

Deformation limit: ε<br />

lim<br />

= ⋅1000(‰<br />

)<br />

E<br />

Where: σ = strength, E = Young modulus<br />

Fig. 7: Time-dependent deformations in concrete subjected to a sustained load<br />

Deformation<br />

Creep<br />

Total deformation<br />

Shrinkage<br />

Elastic strain<br />

Time since application of load<br />

The deformability of hardened concrete is in the range of about 1‰. Another characteristic of<br />

the deformability is the ratio of the transverse strain to the longitudinal compression or of the<br />

transverse contraction to the longitudinal strain.<br />

Poisson’s ration = ε ε<br />

transverse<br />

longitudinal<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 47 of 235


Cement<br />

Concrete<br />

While the modulus of elasticity indicates only the alteration of size in one direction, the<br />

Poisson’s ratio indicates to what extent the load modifies the shape and the volume of the<br />

concrete (compressibility). It decreases with time and is about 0.2 in hardened concrete.<br />

Poisson’s ratio and modulus of elasticity describe completely the deformation behavior of a<br />

material.<br />

4.4 Density (weight of volume unit)<br />

The dry density of the dry solid mass and the bulk density of the concrete have to be<br />

distinguished.<br />

Density:<br />

The density (specific gravity) - the weight per unit volume of a dry and pore-free substance<br />

(mass) - depends above all on the mix design as the density of cement (ρ = 3.0 to 3.15<br />

g/cm 3 ), aggregate (ρ = 2.6 to 2.7 g/cm 3 ) and water ((ρ = 1.0 g/cm 3 ) vary little. The specific<br />

gravity of concrete is in the order of 2.3 to 2.5 g/cm 3 .<br />

Bulk density:<br />

The bulk density (unit weight) - the weight of concrete per unit volume including voids -<br />

(about 2.2 to 2.4 t/m 3 for ordinary concrete) depends on the compaction of the concrete.<br />

The ratio<br />

Specificgravity−Bulkdensity<br />

Specificgravity<br />

indicates the amount of voids in the bulk volume which are filled with water or air.<br />

Voids weaken all mechanical properties of concrete and are, furthermore, decisive for the<br />

impermeability and thus the durability of concrete. Therefore, it is most important to compact<br />

the concrete as firmly as possible. Well compacted concrete contains not more than 1 to 2%<br />

voids of its volume.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 48 of 235


Cement<br />

Concrete<br />

4.5 The durability of concrete<br />

It is essential that concrete keeps its shape and size and does not suffer any deterioration of<br />

its substance which could also cause volume changes (soundness).<br />

As was shown in previous chapters, the mechanical properties of concrete generally<br />

continue to improve with time. Concrete resists moisture and putrefaction as well as high<br />

temperatures (between 200 and 300°C). It does not burn. These are very important<br />

advantages that concrete has over other building materials.<br />

In this paragraph, some factors influencing the durability are discussed.<br />

Fig. 8: Durability of concrete<br />

4.5.1 Volume changes<br />

Concrete is always subject to changes of volume which can be detrimental if they exceed<br />

certain limits.<br />

4.5.2 Volume changes due to physical stress<br />

There are several types of stress which may cause volume changes:<br />

♦ Temperature Changes<br />

Temperature changes cause expansion and contraction in concrete as in all other<br />

materials. These changes are in the same order of magnitude as in steel (approx. 10 -5 for<br />

1°C). If these volume changes are impeded, high stresses result, causing cracks.<br />

Changes due to the freezing - thawing cycle are very hazardous. At freezing<br />

temperatures, the water volume in concrete increases by about 9%. Due to this<br />

expansion and thermal stress, concrete can disrupt. The risk is heightened if de-icing<br />

agents (salts) are employed, because of additional complicated, detrimental effects. The<br />

best way to avoid this reaction is to created voids allowing the water to expand. This is<br />

done by enlarging pore volume by about 3 to 7 Vol.% with the aid of air-entraining agents<br />

(air-entrained concrete), (see chapter ‘Concrete Main Components’, paragraph 3.4).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 49 of 235


Cement<br />

Concrete<br />

♦ Moisture Changes:<br />

Moisture changes - desiccation or absorption of water in hardened concrete - cause<br />

shrinkage or swelling. These changes continue until an equilibrium of moisture with the<br />

surroundings is reached which usually takes several years. Maximum final shrinkage<br />

(determined as a change in length) of the hardened concrete is 0.2 to 0.65 mm/m,<br />

maximum swelling through storage in water at 20°C, 0.1 to 0.3 mm/m depending on<br />

atmospheric conditions.<br />

Fig. 9: Moisture movements in concrete - swelling and shrinkage<br />

Specimen A<br />

0.2<br />

Swelling (%)<br />

Shrinkage (%)<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

-0.15<br />

-0.2<br />

-0.25<br />

-0.3<br />

0<br />

Specimen B<br />

Stored in water<br />

Stored in air<br />

Drying<br />

Specimen A<br />

Alternative<br />

wetting and drying<br />

Time (days)<br />

Much higher and also more dangerous is shrinkage before the setting of concrete - plastic<br />

shrinkage. At intensive desiccation it can reach several per mill (‰). If a lot of water is<br />

allowed to evaporate during the critical period, namely when concrete has only low strength<br />

and is not prevented from shrinking, e.g. through reinforcement, plastic shrinkage-cracks<br />

may result.<br />

To avoid the early shrinkage-cracks, concrete must be protected against desiccation. This<br />

can be achieved by covering the concrete with burlap or plastic sheets or by spraying it with<br />

water.<br />

4.5.3 Volume changes due to alterations of the substance<br />

Chemical alterations of a substance result in volume changes, causing stresses and possibly<br />

cracks.<br />

In concrete mainly the cement stone or the reinforcement (corrosion) are attacked while<br />

aggregates are generally more resistant.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 50 of 235


Cement<br />

Concrete<br />

The volume of cement stone decreases during hydration because the volume of the<br />

hydration product is about 3 to 11% smaller than the original volume of cement and water.<br />

This chemical shrinkage has to be distinguished from drying shrinkage. Chemical shrinkage<br />

enlarges the gel pores and generally does not cause cracks.<br />

A higher risk for concrete is the expansion of the cement stone caused by the reaction of free<br />

lime, periclase and sulfate. These compositions produce unsoundness.<br />

Subsequent reactions between cement and reactive aggregates in hardened concrete can<br />

have a deteriorating effect on concrete, especially if alkalis in the cement react with reactive<br />

silica in the aggregates.<br />

From the external chemical attacks the most important reactions are those with CO 2 in the<br />

air, with acids, sulfates and ammonium salts.<br />

5 Factors influencing the strength of concrete<br />

The concrete properties are the result of a combination of many factors. Some can be<br />

attributed to the material itself, but just as important are the external influences. As has been<br />

mentioned previously, the most important and most characteristic property of concrete is its<br />

compressive strength after 28 days which can also serve as an indication for other<br />

properties. Therefore, this paragraph will be mainly concerned with concrete strength.<br />

The main factors influencing the strength are (see also Fig. 10):<br />

Strength influencing factors:<br />

♦ Properties of the constituents<br />

♦ Mix proportions (mix design, formula)<br />

♦ Way of handling fresh concrete<br />

♦ Conditions under which the concrete hardens<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 51 of 235


Cement<br />

Concrete<br />

Fig. 10: Factors influencing the strength of concrete<br />

Water<br />

Aggregates<br />

Mix Desgin<br />

Cement<br />

Admixtures<br />

Quality of Components<br />

Properties<br />

Testing<br />

of<br />

Concrete<br />

Curing<br />

Time<br />

Humidity<br />

Temperature<br />

Batching<br />

Homogenity<br />

Placing<br />

Mixing<br />

Degree of<br />

Compaction<br />

Air Content<br />

external influences influence of materials<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 52 of 235


Cement<br />

Concrete<br />

One and the same concrete may show various values when tested with different methods.<br />

Thus, the method of testing can influence the test results but not the actual properties of<br />

concrete.<br />

The following figure gives an indication of the range in which strength values can vary; it<br />

shows the results of tests carried out on specimens of the most common cements produced<br />

in ready-mix plants which are affiliated with Holcim.<br />

Fig. 11: Compressive strength after 28 days of samples of 44 Ready-Mix Plants<br />

5.1 Influence of the constituents<br />

5.1.1 Aggregates<br />

Aggregates represent the largest part of concrete in volume (60 to 75%) as well as in weight.<br />

Compressive strength of natural aggregates in normal weight concrete exceeds 2 to 3 times<br />

the strength of cement stone. Thus, it is the strength of the cement stone or the bond that<br />

limits the concrete strength, the cement stone being the weaker constituent.<br />

Aggregates of various origin - if they are clean and free of deleterious substances - develop<br />

practically equal concrete strength if used with the same cement and mix formula (see<br />

following figure).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 53 of 235


Cement<br />

Concrete<br />

Fig. 12: Effect of aggregate on compressive strength concrete with 350 kg/m 3 cement<br />

and w/c ratio of 0.55<br />

Note:<br />

The strength of aggregates does not have a influence on<br />

the compressive strength of ordinary concrete if they meet<br />

basic quality requirements.<br />

The concrete made of various aggregates with equal strength show very different<br />

consistencies This is obviously due to the grading, angularity and surface roughness of the<br />

aggregates, requiring various amounts of water to obtain concretes of the same workability.<br />

The water requirement is mostly influenced by the amount and properties of sand (fine<br />

aggregates). Grading, the maximum size and shape of the aggregates determine the volume<br />

of voids in a compacted aggregate mix as well. The percentage of void volume (varying<br />

between 20 to 40%) indicates the volume of cement paste (cement plus water) required to<br />

obtain a good concrete mix.<br />

Note:<br />

Grading and shape of the aggregates determine the required<br />

amount of water and cement and thus strongly influence the<br />

strength of concretes of equal consistency.<br />

Other properties of aggregates such as: porosity, unsoundness or lack of chemical<br />

resistance affect the concrete durability.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 54 of 235


Cement<br />

Concrete<br />

5.1.2 Water<br />

Water represents the second largest portion of the volume of fresh concrete. The quality of<br />

water only exerts influence on concrete strength if it contains deleterious impurities.<br />

The required water content of a concrete mix depends on the other constituents (cement,<br />

aggregate, admixture).<br />

5.1.3 Cement<br />

Cement has a primary and direct effect on concrete strength by its material properties:<br />

Composition Water requirement<br />

= Cement properties Stiffening and hardening<br />

Fineness Final strength<br />

The water requirement of cement measured on paste of normal consistency has only little<br />

influence on the water requirement of concrete (Fig. 13), much less than the aggregate.<br />

Table 6: Water requirement of standard cement paste and concretes of different<br />

Consistency<br />

OPC Cement A OPC Cement B OPC Cement C<br />

Consistency<br />

Ref.: Standard<br />

cement paste<br />

Water<br />

Requirement<br />

(%)<br />

Vebe<br />

Consistency<br />

(sec)<br />

Water<br />

Requirement<br />

(%)<br />

Vebe<br />

Consistency<br />

(sec)<br />

Water<br />

Requirement<br />

(%)<br />

24 -- 29 -- 29.75 --<br />

Vebe<br />

Consistency<br />

(sec)<br />

Concrete: stiff 52 14.8 53.5 15.0<br />

Stiff-plastic 53.5 10.0 55 10.5<br />

Plastic 55 5.0 57.5 4.5 58 5.6<br />

Concluding from the strength test results on standard cement mortar, only a rough estimate<br />

of the concrete strength can be made.<br />

With increasing standard mortar strength, concrete strength (with the same formula and<br />

aggregate mix) generally tends to increase slightly, especially after 28 days (see diagram<br />

below). The correlation between strength of concrete and mortar is poor; the ratio of both<br />

varies strongly with each individual cement. Cements with higher mortar strength can have<br />

even lower concrete strength.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 55 of 235


Cement<br />

Concrete<br />

Fig. 13: Relation between concrete and standard mortar strength<br />

(determined in Concrete Laboratory of the TC-MD on 37 OPC)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 56 of 235


Cement<br />

Concrete<br />

Note:<br />

It is not possible to deduce from the standard mortar<br />

strength of various cements the strength of concretes of<br />

equal composition.<br />

Important statement:<br />

The hardening rate of concrete is much faster than that of<br />

standard mortar, as can be seen in the following figure.<br />

This is of utmost practical importance because on the job<br />

site under normal conditions (t = 15 to 25°C) the concrete<br />

reaches a certain percentage of its final strength much<br />

sooner than standard mortar.<br />

Fig. 14: Maturity Percentage of Mortar and Concrete<br />

Due too different rates of hardening, the effect of cement properties on early<br />

strength is much stronger than on final strength. Finer grinding results in only slightly higher<br />

final strength but considerably higher early strength. With different cement compositions<br />

(above all blended cements) a concrete showing lower initial strength may have the same or<br />

even higher final strength than a concrete with higher initial strength. For more information<br />

regarding the influence of the cement properties, see paper ‘Cement’.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 57 of 235


Cement<br />

Concrete<br />

The influence of admixtures has already been described in paper ‘Concrete Main<br />

Components’.<br />

5.2 Influence of mix proportions<br />

5.2.1 Water / Cement ratio<br />

As already mentioned, the concrete strength depends essentially on the strength of the<br />

cement stone, and is only partially determined by the cement strength as measured on<br />

standard mortar. There must be other factors as well. The strength of mortar made with a<br />

given cement depends on its compaction, i.e. on the part of volume occupied by solid<br />

material and by voids (gel pores and capillaries).<br />

The water requirement of cement for complete hydration is about 25%; another 15% is<br />

contained in the gel pores. If concrete contains less than 40% water, cement does not<br />

hydrate completely. Furthermore, since an extremely dry concrete cannot be compacted<br />

adequately, a deficiency of water causes inferior strength. For reasons regarding the<br />

workability, however, the concrete usually contains more water than the cement requires.<br />

The following figure shows how the compaction of mortar depends on the ratio of the water<br />

content to the cement content:<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 58 of 235


Cement<br />

Concrete<br />

Figure 15: Structure of fresh cement mortar<br />

This weight proportion is usually termed water/cement ratio (w/c ratio). While<br />

standard mortar has always the same prescribed w/c ratio, the water content of concrete<br />

varies according to the desired consistency and the water requirement of the aggregate and<br />

of the cement to obtain this consistency. The next two diagrams show the effect of the w/c<br />

ratio on the concrete strength.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 59 of 235


Cement<br />

Concrete<br />

Fig. 16: Relation between compressive strength and water/cement ratio of concrete<br />

Important statement:<br />

For the best concrete quality it is important to choose the<br />

lowest possible w/c ratio that still enables the concrete to<br />

be perfectly compacted.<br />

In the concrete standards of different countries the maximum permissible w/c ratio for<br />

different types of structure is fixed (see tables later: Requirements for concrete acc. to DIN).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 60 of 235


Cement<br />

Concrete<br />

One of the most common causes of poor quality of concrete is the addition of too much<br />

water. Following figure shows how an inaccurate measurement of the quantity of mixing<br />

water or an uneven moisture content of aggregates may impair the strength of concrete. The<br />

moisture content of the aggregates must be deduced from the amount of water to be added.<br />

Fig. 17: Influence of change in water content on compressive strength<br />

Fig. 18: Relation between calculated and observed concrete strength<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 61 of 235


Cement<br />

Concrete<br />

The calculated concrete strength is only obtained, however, if the concrete is perfectly<br />

compacted (porosity < 1%).<br />

5.2.2 Cement content<br />

The content of cement or cement paste in concrete must be sufficient to envelop the<br />

aggregate grains in order to reduce their friction, to glue them, to fill the voids between them<br />

and also to protect the reinforcement against corrosion. Thus, an insufficient amount of<br />

cement impairs the workability and compaction of concrete and as a consequence, its<br />

strength and durability.<br />

After all voids are filled, a surplus of cement paste does not increase the concrete strength<br />

any further since cement stone has a lower strength than aggregate (Fig. 20a). By increasing<br />

the cement content above an optimum, the ratio between compressive strength and cement<br />

content decreases (see figure). If the cement content in concrete is sufficient for full<br />

compaction, the strength of fully compacted concrete is given by the strength of the cement<br />

stone. It depends only on its w/c ratio and on the standard cement strength.<br />

Fig. 19: Utilization of cement in concrete at constant workability<br />

5.2.3 Aggregate/Cement ratio<br />

The optimum ‘aggregate/cement’ ratio is not only considered from a technical but also from<br />

an economical point of view since aggregate is stronger and cheaper than cement. Thus, the<br />

grading of aggregate is most important, not only for workability but also for the minimum<br />

volume of voids. It makes it possible to obtain the required strength with less cement.<br />

5.3 Influence of handling<br />

It is most important that the consistency of concrete permits full compaction with the tools<br />

that are available on the job site. The need for compaction becomes apparent in the following<br />

figure:<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 62 of 235


Cement<br />

Concrete<br />

Fig. 20: Relation between strength and compaction<br />

Another prerequisite for the optimum utilization of the material properties is the homogeneity<br />

of the mix. To maintain uniformity within one batch and between several batches, accurate<br />

weighing, thorough mixing and careful transporting and placing is necessary. These<br />

measures also prevent segregation.<br />

5.4 Influences of curing<br />

5.4.1 Moisture<br />

After placing, the subsequent conditions strongly influence the development of the concrete<br />

properties. As for the hydration of cement water is required, loss of moisture must be<br />

prevented or even additional moisture supplied. The next figure shows how concrete strength<br />

is impaired by the lack of moisture, especially during the first 3 days.<br />

Fig. 21: Strength of concrete increase as long as moisture is present forhydration of<br />

cement<br />

Concrete that hardens submerged in water has the highest strength. Once it is hardened, dry<br />

concrete has higher strength than moist concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 63 of 235


Cement<br />

Concrete<br />

5.4.2 Temperature<br />

Higher temperatures accelerate setting and hardening of concrete, like other chemical<br />

processes(see following figures). Seasonal and even daily temperature fluctuations alter the<br />

hardening characteristics of concrete.<br />

Fig. 22: Effect of temperatures on the compressive strength of concrete at various<br />

ages (T ≥ 20 °C)<br />

Fig. 23: Effect of temperatures on the compressive strength of concrete at various<br />

ages (T ≤ 20 °C)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 64 of 235


Cement<br />

Concrete<br />

Note:<br />

♦ The temperature during the first 24 hours approximately strongly affects the initial<br />

strength development of concrete and it predetermines the later hardening and even the<br />

final strength. Thus, it is futile to maintain high curing temperatures after the first 24<br />

hours.<br />

♦ Concrete cured at low temperature achieves a slightly higher final strength than the<br />

concrete cured at 30 °C or more and vice-versa.<br />

Accelerated hardening obviously creates structures of hydration products that are less<br />

favorable for later hardening.<br />

The effects of temperature depend not only on external curing conditions, but also on the<br />

temperature within the concrete created by accumulation of hydration heat.<br />

The development of concrete temperature depends on cement properties, the composition of<br />

concrete and the size and shape of the concrete element. This fact is illustrated the following<br />

figure which compares the temperatures of mortar and concrete and of prisms and cubes. If<br />

concrete is protected against heat-losses and ‘autocuring’ takes place; some mass-concrete,<br />

the internal temperature of which may rise up to 90 °C, has to be cooled to prevent cracks<br />

caused by internal stress.<br />

Figure 24: Temperature of mortar and concrete specimen<br />

It is important to note that sooner or later the temperature or curing effects may activate the<br />

potential for hardening of various cements; the potential as such, however, cannot be<br />

increased by these measures.<br />

5.5 Influences of testing methods<br />

The influence of the testing method can sometimes be decisive for the final test result.<br />

Specimen size, moulding, curing and testing procedure are some of the influencing factors.<br />

The above mentioned figure demonstrates how test results of temperature evolution<br />

measurements depend on specimen size and concrete composition. The differences<br />

between results on test specimens and concrete in construction can be very significant.<br />

Thus, results of specimen tests are never identical to those obtained from tests on concrete<br />

in situ. They merely give some indications. (See separate paper: ‘Testing’.)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 65 of 235


Cement<br />

Concrete<br />

6 Conclusions<br />

There are many factors which influence the properties of hardened concrete. The next figure<br />

summarizes the effect of these various factors.<br />

Fig. 25<br />

Factors Influencing the Properties of Concrete<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 66 of 235


Cement<br />

Concrete<br />

7 Supplementary Literature (Books):<br />

• P. Hewlett, LEA's Chemistry of Cement & Concrete, 1997<br />

• J. Skalny, S. Mindess, Materials Science of Concrete II, 1997<br />

• ACI Manual of Concrete Practice, 1996<br />

• A.M. Neville, Properties of Concrete, 1996<br />

• Behavior of Fresh Concrete During Vibration / Aci, 1993<br />

• P.K. Mehta, P.J.M. Monteiro, Concrete: Structure, Properties and Materials (Prentice<br />

Hall International Series in Civil Engineering and Engineering Mechanics), 1992<br />

• P. Bartos, Fresh Concrete. Properties and Tests, 1992<br />

♦ A.M. Neville, Hardened concrete: physical and mechanical aspects.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 67 of 235


Cement<br />

Concrete<br />

Concrete Mix Design<br />

1. GENERAL.........................................................................................................................69<br />

2. BASIC DESIGN CONSIDERATIONS............................................................................... 69<br />

2.1 Costs ........................................................................................................................ 69<br />

2.2 Workability................................................................................................................ 72<br />

2.3 Strength and durability ............................................................................................. 72<br />

3. SPECIFICATIONS FOR CONCRETE MIXES .................................................................. 73<br />

4. PROCESS OF MIX DESIGN FOR NORMAL-WEIGHT CONCRETE............................... 75<br />

5. EXAMPLE OF MIX DESIGN CALCULATION.................................................................. 79<br />

6. CONCLUDING REMARKS............................................................................................... 83<br />

7. LITERATURE.................................................................................................................... 84<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 68 of 235


Cement<br />

Concrete<br />

1 General<br />

Mix design is a process that consists of two interrelated steps:<br />

(1) Selection of the suitable ingredients (cement, aggregate, water and admixture) of<br />

concrete and<br />

(2) Determination of their relative quantities to produce as economically as possible<br />

concrete of appropriate workability, strength and durability<br />

Although many concrete properties are important, most design procedures are aimed at<br />

achieving a specified compressive strength at a given workability; it is assumed that if this is<br />

done, the other properties will also be satisfactory. An exception represent the resistance to<br />

freeze-thaw and other durability problems (for instance sulphate resistance), which require<br />

special attention in the mix design.<br />

2 Basic Design considerations<br />

2.1 Costs<br />

The cost of concreting is made up of material costs, plant and labour expenses. Except for<br />

some special concretes, the costs of labour and equipment are, however, to a large extent<br />

independent of the type and quality of concrete produced. It is therefore the material costs<br />

that are most important in determining the relative costs of different mix designs. Since<br />

cement is much more expensive than aggregate (see Table 1), it is clear that minimising the<br />

cement content is the most important single factor in reducing concrete costs (see also<br />

Figure 1).<br />

Table 1: Prices of concrete constituents in Switzerland per ton (approx.)<br />

Concrete constituent<br />

Price (CHF)<br />

OPC cement silo 110.--<br />

OPC cement bag 125.--<br />

Sand fraction 0/4 mm 20.20<br />

Gravel fraction 4/8 mm 16.90<br />

Gravel fraction 8/16 mm 12.--<br />

Gravel fraction 16/32 mm 9.40<br />

To economise on material costs, the proportioning should minimise the cement content<br />

without scarifying concrete quality. Since the quality depends primarily upon the w/c ratio, the<br />

water content should be reduced to lower the cement content. Some of the steps to minimise<br />

water and cement contents are to use:<br />

♦ the stiffest practicable mixture<br />

♦ the largest possible maximum size of aggregate<br />

♦ the optimum ratio of fine and coarse aggregates<br />

The cement reduction is, however, often restricted by specifications stating the minimum<br />

cement content (see Table 2).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 69 of 235


Cement<br />

Concrete<br />

Table 2:<br />

Minimum cement content in concrete for different national standards<br />

(kg/m 3 )<br />

Type of concrete<br />

Germany<br />

DIN 1045<br />

Switzerland<br />

SIA 162<br />

France*<br />

NF P-18-305<br />

USA<br />

England<br />

Plain concrete ≥ 100 -- -- -- --<br />

Reinforced/prestressed<br />

concrete<br />

280 (OPC<br />

25)<br />

240 (OPC<br />

35)<br />

300<br />

for class 250<br />

and higher:<br />

(250 +<br />

B)/D 1/5 none none<br />

* B = compr. strength in bar; D = max. aggregate size<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 70 of 235


Cement<br />

Concrete<br />

Figure 1:<br />

Costs of concrete: material costs<br />

Cement<br />

17.4<br />

Rich mix<br />

Lean Mix<br />

Cement<br />

9.3<br />

W: 6.7<br />

Water<br />

7.8<br />

Cement<br />

60.2<br />

Fine<br />

aggreg.<br />

Cement<br />

46.5<br />

Fine<br />

Aggreg.<br />

28.0<br />

37.4<br />

Fine<br />

aggreg.<br />

Coarse<br />

Aggreg.<br />

37.4<br />

Fine<br />

aggreg.<br />

22.8<br />

Coarse<br />

aggreg.<br />

Coarse<br />

aggreg.<br />

56.0<br />

21.4<br />

Coarse<br />

aggreg.<br />

32.1<br />

17.0<br />

Mix proportions<br />

by weight %<br />

Costs<br />

%<br />

Mix proportions<br />

by weight %<br />

Costs<br />

%<br />

It should be noted here that in addition to cost, there are other benefits by using a low<br />

cement content; shrinkage will in general be reduced and there will be less heat of hydration.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 71 of 235


Cement<br />

Concrete<br />

However, if the cement contents are too low, they will diminish the early strength and<br />

durability of concrete and will make uniformity of the concrete a more critical consideration.<br />

Besides the cement reduction as such, there is also an interesting potential to reduce costs<br />

through the replacement of cement by mineral components (e.g. fly ash or ground blast<br />

furnace slag). Prerequisite is of course that mineral components of good quality are available<br />

at the concrete plant at convenient prices.<br />

To determine the most economical mix proportions, the relative costs of fine and coarse<br />

aggregates should also be taken into consideration. Since admixtures to reduce the water<br />

requirement will increase material costs, it should be assessed in each case whether their<br />

use is justified by the savings in labour costs and cement cost eventually.<br />

The economy of a particular mix design should also be related to the degree of quality<br />

control that can be expected on a job. At least on small jobs, it may be cheaper to overdesign<br />

the concrete than to provide extensive quality control that would be required with a more<br />

cost-efficient concrete.<br />

2.2 Workability<br />

Clearly, a properly designed mix must be capable of being placed and compacted properly<br />

with the equipment available. Finishability must be adequate, and segregation and bleeding<br />

should be minimised. As a general rule, the concrete should be supplied at the minimum<br />

workability that will permit adequate placement. The water requirement for workability<br />

depends mostly on the characteristics of the aggregate rather than of those of the cement.<br />

Where necessary, workability should be improved by increasing the mortar content rather<br />

than by simply adding more water or more fine material. Thus, co-operation between the mix<br />

designer and the contractor is essential to ensure a good concrete mix. In some cases, a<br />

less economical mix may be the best solution.<br />

2.3 Strength and durability<br />

In general, concrete specifications will require a minimum compressive strength. They may<br />

also impose limitations on the permissible w/c-ratios and minimum cement contents. It is<br />

important to ensure that these requirements are not mutually incompatible. It is not<br />

necessarily the 28 day strength that is most important; strength at other ages may control the<br />

design.<br />

Specifications may also require that the concrete meet certain durability requirements, such<br />

as resistance to freezing and thawing, or chemical attack. These considerations may provide<br />

further limitations on the w/c-ratio or cement content and in addition may require the use of<br />

admixtures.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 72 of 235


Cement<br />

Concrete<br />

3 Specifications for Concrete Mixes<br />

There are three systems that can be applied to specify structural concrete. These differ with<br />

respect to the basis of specification, responsibility for the mix design and the parameter by<br />

which the concrete is judged for compliance, as indicated in Table 3.<br />

Table 3:<br />

Types of concrete mixes and their method of specification<br />

Type of mix Designed mix Codified mix Prescribed mix<br />

Design<br />

stage<br />

Production<br />

stage<br />

Compliance<br />

stage<br />

Basis of<br />

specification<br />

Strength<br />

classes<br />

Responsibility<br />

for mix design<br />

Permitted<br />

materials<br />

Strength testing<br />

for information<br />

Basic<br />

parameters of<br />

concrete for<br />

compliance<br />

testing<br />

Strength Mix Proportions Mix Proportions<br />

all ≤ 25 N/mm 2 all<br />

Producer<br />

Any materials<br />

complying with<br />

National<br />

Standards<br />

Basis of mix<br />

Strength<br />

National code<br />

or specification<br />

Restricted<br />

range of<br />

materials<br />

complying with<br />

Nat. Standards<br />

Not usually<br />

necessary<br />

Purchaser<br />

any<br />

Desirable<br />

especially for<br />

higher strength<br />

classes<br />

Mix proportions Mix proportions<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 73 of 235


Cement<br />

Concrete<br />

Designed mixes<br />

The mix should be designed to have an adequate margin of strength between the specified<br />

characteristic strength or the designed mean strength based on the standard deviation<br />

obtained through experience. If there are no previous data, or if there are less than 30 results<br />

obtained under equal conditions (same plant, source of materials, and supervision), the<br />

proposed minimum value for standard deviation as per curve A (Figure 2) should be<br />

considered.<br />

Figure 2:<br />

Proposed minimum values for the standard deviation<br />

If there are between 30 and 100 results, the standard deviation should be the value obtained<br />

but not less than that given by curve B, i.e. for characteristic strength equal to or greater than<br />

20 N/mm 2 , the minimum standard deviation is 4 N/mm 2 . For more than 100 results curve C<br />

gives the minimum standard deviation.<br />

These values of standard deviation will be used for the calculation of characteristic strength.<br />

Designed mixes are the most common types of mixes used in concrete practice.<br />

Codified mixes<br />

Codified mixes are recommended in National Standards or Codes and are based on the<br />

characteristics of the cement and aggregates available in each country. The adequacy of the<br />

structure is generally assured since the cement content will usually be greater than that of<br />

the corresponding designed mixes.<br />

Codified mixes are restricted to the lower strength classes of concrete and may be made<br />

only with restricted types of materials. They are sometimes used as approximate guidelines<br />

for designed mixes.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 74 of 235


Cement<br />

Concrete<br />

Prescribed mixes<br />

Prescribed mixes are applied when the purchaser wishes to specify the mix proportions to be<br />

used. There are a number of circumstances which call for such mixes; for example, if the<br />

purchaser knows that with local materials certain mix proportions will produce concrete of the<br />

required proportions, or if he requests the use of specific mixes because he wants a concrete<br />

with special characteristics (high or low density, etc.). Since the purchaser designs the mix, it<br />

is also his responsibility to ensure that the strength and other requirements for the safety of<br />

the construction are met.<br />

4 Process of mix design for normal-weight concrete<br />

Before a concrete mix can be designed and proportioned, certain points of information<br />

should be known:<br />

1) Size and shape of structural members<br />

2) Spacing and diameter of reinforcement<br />

3) Required strength<br />

4) Exposure conditions (requirements for durability and chemical attack)<br />

5) Placing and compaction methods<br />

6) Basic material quality data on cement and aggregate<br />

It is up to the design engineer and the architect to provide the information for points 1), 2)<br />

and 3). The information for points 5) and 6) can be obtained from the contractor firm and<br />

material supplier. As to point 4), the requirements for exposure conditions are normally<br />

specified in the standards.<br />

The basic factors that have to be considered in determining the mix proportions are<br />

represented schematically in Figure 3.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 75 of 235


Cement<br />

Concrete<br />

Figure 3:<br />

Basic factors in the process of mix design<br />

Liability to chemical<br />

attack, or size of<br />

concrete mass<br />

Method<br />

of compaction<br />

Size of section and<br />

spacing of reinforcement<br />

Quality<br />

control<br />

Minimum<br />

strength<br />

Required<br />

workability<br />

Maximum size<br />

of aggregate<br />

Aggregate shape<br />

and texture<br />

Mean<br />

strength<br />

Type of cement<br />

Durability<br />

w/c Ratio<br />

Aggreg./Cement<br />

Ratio<br />

Grading of<br />

aggregate<br />

Proportion<br />

of each size<br />

fraction<br />

Mix proportions<br />

The American Concrete Institute ACI recommends a maximum permissible w/c ratio for<br />

different types of structures and degree of exposure. The European Norm EN 206 for<br />

concrete describes as well the maximum permissible w/c ratio, minimum cement content, the<br />

minimum air content and range of aggregate grading for the different exposure conditions.<br />

It should be explained that a design in the strict sense of the word is not possible: many of<br />

the material properties and effects in the concreting process cannot be truly assessed<br />

quantitatively, so that we are really making no more than an intelligent guess at the optimum<br />

combinations of the ingredients on the basis of the relationships established empirically.<br />

Therefore, we not only have to calculate or estimate the proportions of the available<br />

materials, but must also make trial mixes. Properties of trial mixes are checked and<br />

adjustments in the proportioning are made until a fully satisfactory mix is attained.<br />

Trial mixes are usually relatively small batches made with laboratory precision so that a<br />

certain security factor should be calculated when using laboratory results on large job-site<br />

batches.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 76 of 235


Cement<br />

Concrete<br />

The calculation of material quantities for mix design of ordinary normal-weight concrete is<br />

based on the absolute volume of the ingredients:<br />

Volumes in 1m 3 compacted fresh concrete:<br />

3<br />

[ ]<br />

C A W<br />

+ + + P=<br />

1000 dm<br />

δ δ δ<br />

c<br />

A<br />

w<br />

C = cement content in kg/m 3<br />

A = aggregate content in kg/m 3<br />

W = water content in kg/m 3<br />

P = air pores;<br />

in compacted non-air-entrained concrete:10 ÷ 20 dm 3 /m 3 ;<br />

in air-entrained concrete: 30 ÷ 70 dm 3 /m 3<br />

δ c = bulk density of cement kg/dm 3<br />

δ A = bulk density of aggregate kg/dm 3<br />

δ w = bulk density of water = 1.0 kg/dm 3<br />

Guide values for the above bulk densities are given in Table 4<br />

Table 4: Guide values for bulk densities δ in kg/dm 3<br />

Cements<br />

Aggregates<br />

Portland cement 3.05 - 3.15 kg/dm 3 Sand, gravel 2.6 - 2.65 kg/dm 3<br />

Blast furnace<br />

3.0 kg/dm 3 Limestone 2.7 kg/dm 3<br />

slag cement<br />

Pozzolanic<br />

2.9 kg/dm 3 Basalt 2.9 kg/dm 3<br />

cement<br />

Volcanic gravel + 2.1 - 2.4 kg/dm 3<br />

sand<br />

Lightweight<br />

aggregate<br />

for structural<br />

lightweight<br />

concrete<br />

0.4 ÷ 1.9 kg/dm 3<br />

There are various methods to calculate a concrete mix. The difference between those<br />

methods is not to be attributed to the basic factors but rather to the mode of calculation and<br />

adjustment applied. As example, three methods to calculate the mix design of normal weight<br />

concrete are compared in Table 5.<br />

Here, only the general principle of the mix design procedure shall be briefly presented (see<br />

also example of mix design in the next chapter):<br />

First of all, the maximum permissible w/c-ratio for the required concrete strength has to be<br />

fixed; this value can be taken from the relationship w/c-ratio - concrete strength in function of<br />

the strength class of the cement.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 77 of 235


Cement<br />

Concrete<br />

Then, the required water content w (mixing water plus surface humidity of aggregate) has to<br />

be defined. The water content w can be estimated based on the needed consistency and the<br />

aggregate grading and modulus respectively.<br />

Table 5:<br />

Example of methods to calculate a concrete mix<br />

Step<br />

No.<br />

United States<br />

ACI<br />

Country & Standard (Code)<br />

England<br />

Dep. of Environment<br />

1. workability = slump mean strength (w/cratio)<br />

2. max. size of aggregate water content<br />

(workability)<br />

3. mixing water + (air cement content<br />

content)<br />

W. Germany<br />

DIN<br />

cement quality<br />

consistency<br />

grading of<br />

aggregate<br />

4. w/c ratio aggregate content water content<br />

5. cement content fine/coarse aggregate concrete strength<br />

6. coarse aggregate content trial mixes w/c ratio<br />

7. fine aggregate content cement content<br />

8. adjustment for aggregate<br />

moisture<br />

weight of<br />

ingredients<br />

9. trial batch adjustments trial mix<br />

For<br />

details<br />

consult:<br />

ACI 211.1-91, ACI<br />

Manual of Concrete<br />

Practice, American<br />

Concrete Institute, USA<br />

A. Neville: Properties of<br />

concrete. Chapter 14.<br />

Mix Design., Longman,<br />

1995<br />

K. Walz:<br />

Herstellung von<br />

Beton nach DIN<br />

1045. Beton-Verlag<br />

GmbH, Düsseldorf,<br />

1971.<br />

The necessary cement content c in the concrete mix can accordingly be calculated with the<br />

w/c-ratio and the water content w by means of the following formula:<br />

c = w/(w/c)<br />

Knowing the cement and water content, we can calculate the required amount of aggregate<br />

using the equation on the absolute volume of the ingredients (assumption on air pores p).<br />

Thus, the weight of oven dry aggregate A for 1 m 3 of concrete is:<br />

A = δ A (1000 - c/δ c - w - p)<br />

If humid aggregate is used, then the amount of surface humidity during concrete preparation<br />

has to be determined and to be included for the proportioning of the mixing water and the<br />

aggregate. The mixing water is calculated by subtracting the content on surface humidity<br />

from the water content w. For the proportioning of the aggregate, the amount A calculated<br />

according to the above equation has to be increased by the content of surface humidity.<br />

If plasticising concrete admixtures (e.g. water reducers or air entrainer) are added to the<br />

concrete, the water content w for a given consistency is reduced. The fine air pores produced<br />

by the addition of air entrainer improve also indeed the workability, but they reduce at the<br />

same time the concrete strength; in case of air entrained concrete, the real water content has<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 78 of 235


Cement<br />

Concrete<br />

therefore to be adjusted for both the plasticising and strength reducing effect of the air<br />

entrainer.<br />

5 Example of Mix Design Calculation<br />

For illustration, a mix design shall be calculated here with the aid of Table 6 and Figures 4 to<br />

6 from DIN standard 1045.<br />

In our example, the concrete requirements shall be as follows::<br />

♦ consistency: compaction factor 1.20<br />

♦ strength class: B 35<br />

♦ exposure condition: interior of building (thus no further limitations on w/c-ratio and cement<br />

content)<br />

♦ max aggregate size: 32 mm<br />

♦ aggregate grading: between sieve curves A and B (see Figure 6)<br />

The available concrete components are:<br />

♦ cement class Z 35 (density 3.10 kg/dm 3 )<br />

♦ aggregates:<br />

sand 0/2 mm (density 2.52 kg/dm 3 )<br />

natural aggregate 2/8 mm (density 2.62 kg/dm 3 )<br />

natural aggregate 8/32 mm (density 2.72 kg/dm 3 )<br />

with the following grading:<br />

Passing sieves (%) 0.25mm 0.5 mm 1 mm 2 mm 4 mm 8 mm 16 mm 32 mm<br />

sand 6 50 80 97 100 100 100 100<br />

aggr. 2/8mm 3 8 10 10 55 95 100 100<br />

aggr. 8/32 mm 2 2 3 3 4 6 50 100<br />

The average compressive strength tested on three cubes in one trial batch for the concrete<br />

strength class B 35 should be at least 45 N/mm 2 (see Table 6). Accordingly, the<br />

water/cement-ratio can be estimated from Figure 4; it amounts to 0.47.<br />

Aiming at a grading of the aggregate mixture between A and B sieve curves, the estimated<br />

amount of aggregate fractions is:<br />

♦ sand:<br />

25 vol.%<br />

♦ 2/8 mm: 23 vol. %<br />

♦ 8/32 mm: 52 vol. %<br />

The resulting grading of the aggregate mixture is thus:<br />

Passing sieves (%) 0.25 mm 0.5 mm 1 mm 2 mm 4 mm 8 mm 16 mm 32 mm<br />

3 15 24 28 39 49 74 100<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 79 of 235


Cement<br />

Concrete<br />

With the corresponding grading modulus 4.68 and the concrete compacting factor of 1.20,<br />

the water content w estimated from Figure 5 is 153 kg/m 3 .<br />

The calculated cement content c is therefore:<br />

153 : 0.47 = 325 kg/m 3 of concrete<br />

The volume proportion of the oven dry aggregate mixture in 1 m 3 of concrete can then be<br />

calculated by means of the absolute volume of the ingredients (assuming an air content P of<br />

15 dm 3 ):<br />

consisting of:<br />

A = 1000 - (c/3.10 - w - P) = 1000 - (325/3.10 - 153 - 15) = 727 dm 3<br />

727 x 0.25 x 2.62 = 476 kg of sand<br />

727 x 0.23 x 2.62 = 438 kg of fraction 2/8 mm<br />

727 x 0.52 x 2.72 = 1028 kg of fraction 8/32 mm<br />

That means 1942 kg of aggregate mix in 1 m 3 of concrete.<br />

The mix design is thus:<br />

aggregate 1942 kg/m 3<br />

cement 325 kg/m 3<br />

water 153 kg/m 3<br />

resulting in 2420 kg/m 3 of fresh compacted concrete.<br />

Table 6: Requirements for concrete in preliminary test<br />

Required Group of concrete Strength class Required compr.<br />

strength<br />

(N/mm 2 1), 2)<br />

)<br />

consistency<br />

values 2)<br />

B I B 5<br />

B 10<br />

B 15<br />

B 25<br />

≥ 11<br />

≥ 20<br />

≥ 25<br />

≥ 35<br />

K 1: v = 1.30...1.26<br />

K 2: v = 1.15...1.11<br />

a = 39...40 cm<br />

K3: v = 1.06...1.04<br />

a = 48...50 cm<br />

B II B 35<br />

B 45<br />

B 55<br />

40 + m 3) according to<br />

60 + m 3) site, incl. margin<br />

50 + m 3) requirements of the<br />

1)<br />

2)<br />

average compressive strength of three cubes of one batch<br />

site-mixed and ready-mixed concrete, not for concrete in concrete works<br />

3) choose margin m according to experience, otherwise m ≥ 5 MPa are to be used<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 80 of 235


Cement<br />

Concrete<br />

Figure 4:<br />

Relationship strength versus w/c ratio<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 81 of 235


Cement<br />

Concrete<br />

Figure 5: Relationship between consistency, grading modulus and water content of<br />

concrete<br />

Correction of water content:<br />

- by using crushed + 5 ÷ 10%<br />

aggregate<br />

- by using plasticisers - 5%<br />

- by using air-entraining<br />

agents<br />

- 1% per 1% of air<br />

content<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 82 of 235


Cement<br />

Concrete<br />

Figure 6:<br />

Standard grading curves<br />

6 Concluding Remarks<br />

There are several well established methods to calculate concrete mixes. Given the variability<br />

of the properties of ingredients and the difficulty in describing them, the results of these<br />

calculations are, however, really only guesses. On the other hand, such methods are<br />

concerned basically with the technical aspects of mix design without proper consideration of<br />

the economical side of the problem.<br />

The establishment of a proper concrete mix design will always involve the preparation and<br />

testing of trial mixes and consecutive adjustments. The experience and knowledge on the<br />

influence of the various factors upon the properties of concrete can of course help to improve<br />

the first guess and limit the number of trial mixes to be tested.<br />

To facilitate the selection of the optimum concrete mix design (also from the economic point<br />

of view), more and more computer based tools are used nowadays in practice. Such tools do<br />

not replace experience, but allow to explore more quickly the different alternatives and to<br />

arrive faster at the optimum solution.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 83 of 235


Cement<br />

Concrete<br />

7 Literature<br />

ACI 211.1-91, Standard practice for selecting proportions for normal, heavyweight, and mass<br />

concrete, in: ACI Manual of Concrete Practice 1996, Part 1, American Concrete Institute,<br />

Farmington Hills, USA, 38 pp.<br />

Neville, A.M., Selection of concrete mix proportions (mix design), in: Properties of Concrete,<br />

Longman, England, 1995, pp. 724 - 772<br />

Day, K.W., Concrete mix design, quality control and specifications, E & F.N. Spon, London,<br />

1995, 350 pp.<br />

Kosmataka, S.H, and Panarese, W.C., Design and control of concrete mixtures, PCA,<br />

Skokie, 1994, 205 pp.<br />

Daly, D.D., Concrete mix design, in: Fulton's Concrete Technology, PCI, Midrand, 1994, pp.<br />

209 - 217<br />

Torrent, R., and Honerkamp, M., Computer aided optimised concrete mix-design, 33rd<br />

<strong>Technical</strong> Meeting/10th Aggregate and Ready-Mixed Concrete Conference, Basle, 1994,<br />

HMC Report MA 94/3157/E<br />

Bai, Y., and Amirkhanian, S.N., Knowledge-based expert system for concrete mix design,<br />

Journal of Construction Engineering and Management, Vol. 120, No. 2, 1994, pp. 357 - 373<br />

Foo, H.C., and Akhras, G., Expert systems and design of concrete mixtures, Concrete<br />

International, July 1993, pp. 42 – 46<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 84 of 235


Cement<br />

Concrete<br />

Reduction in Materials Costs<br />

1. INTRODUCTION............................................................................................................... 86<br />

2. DESCRIPTION OF THE PROBLEM................................................................................. 86<br />

3. OPTIMISATION THROUGH PROPER SELECTION OF THE<br />

COMPONENTS................................................................................................................. 87<br />

3.1 Cement..................................................................................................................... 87<br />

3.2 Mineral Additions...................................................................................................... 89<br />

3.3 Aggregates............................................................................................................... 90<br />

3.4 Chemical Admixtures ............................................................................................... 93<br />

4. OPTIMISATION TROUGH REFINEMENT IN STRENGTH LEVEL.................................. 93<br />

5. CONCLUSIONS................................................................................................................ 98<br />

6. REFERENCES.................................................................................................................. 98<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 85 of 235


Cement<br />

Concrete<br />

1 Introduction<br />

We are convinced that, even though most Ready-Mixed Concrete (RMC) companies within<br />

the "Holderbank" Group operate at high levels of rationalisation in terms of materials and<br />

processes, there is always room to achieve further reductions in materials costs .<br />

The objective of this contribution is to discuss the basic aspects of the selection of raw<br />

materials for concrete, as well as their proportioning, in view of possible cost reductions in<br />

the production of RMC. It is expected that the discussion of those aspects serves as a<br />

motivation and incentive to our companies in their search of optimising their costs (BCM<br />

Project). Although they will not be dealt with here, it must be mentioned that aspects of<br />

production process and quality control must always be considered in the optimisation, since<br />

they are very important as well.<br />

2 Description of the Problem<br />

The basic problem consists in minimising the cost of concrete ($h):<br />

where:<br />

$h = C.$c + A.$a + W.$w + M.$m + Q.$q (1)<br />

$h = materials cost of concrete ($/m3)<br />

C ; $c = content (kg/m3) and price of cement ($/kg)<br />

A ; $a = idem for aggregates<br />

W ; $w = idem for mixing water<br />

M ; $m = idem for mineral additions<br />

Q ; $q = idem for chemical admixtures<br />

Clearly, the solution of the problem does not consist in choosing the cheapest components,<br />

because the mix design will depend on their quality. Moreover, the terms in (1) are not<br />

independent, since the sum of the volumes of the components must always add to 1 m3:<br />

C A W M Q<br />

--- + --- + --- + --- + --- + P = 1 m3 (2)<br />

dc da dw dm dq<br />

where di means the density of component i en kg/m3 and P is the volume, in m3, occupied<br />

by entrapped or entrained air.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 86 of 235


Cement<br />

Concrete<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Figure 1<br />

Switzerland<br />

Argentina<br />

Cement Gravel Sand Plasticizer<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Fig. 1 shows the relative unit prices of the main concrete components in Switzerland and<br />

Argentina (Cement =1). As is well known, cement price is several times higher than that of<br />

the aggregates and, hence, to a large extent cost reductions in concrete have to be achieved<br />

by reducing its cement content. Moreover, the relatively high unit price of plasticizers either<br />

increase the cost of the mix or has to be compensated by larger savings in cement costs to<br />

make their use economical.<br />

The following analyisis of cost optimisation will be made on the basis of mixes that are<br />

technically equivalent (competitive). That means that they must fulfil the characteristics that<br />

are normally specified by the customers, i.e. they must possess similar<br />

consistency/workability and compressive strength.<br />

3 Optimisation through proper Selection of the<br />

Components<br />

3.1 Cement<br />

The type and physical/chemical characteristics of the cement may have a considerable<br />

influence on their consumption in the concrete mixes.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 87 of 235


Cement<br />

Concrete<br />

Figure 2<br />

Binder Content (kg/m 3 )<br />

for concrete with Slump= 70 mm and f’c cyl@28<br />

= 25 MPa<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

Blends with UNT2<br />

RO GA UNT5 UNT2 Fly Ash<br />

(20%)<br />

Slag<br />

(35%)<br />

Filler<br />

(15%)<br />

Binder Type<br />

Fig. 2 shows the cement content required to produce concretes of similar performance<br />

(consistency, air content and 28-day strength), for various Portland and blended cements [1].<br />

It can be seen that there is a considerable difference in their consumption. The most efficient<br />

cement (GADOR, Type II, export-quality Spanish cement) requires some 225 kg/m3 whilst<br />

the lest efficient (ROCHE Special, Switzerland) requires 340 kg/m3 to produce equivalent<br />

concretes. It must be mentioned that the latter is not a commercial cement, because it was<br />

purposely ground rather coarse (Blaine 2470 g/cm²) and, besides, it possesses a low C3S<br />

content. In any case, GADOR cement demands some 50 kg/m3 less than the other<br />

commercial cements, produced at UNTERVAZ, Switzerland, that have average<br />

characteristics of the market. The consumption of blended cements is somewhat higher than<br />

that of their Portland component (UNTERVAZ M.2), especially of that containing limestone<br />

filler, which reflects the scarce hydraulic contribution of this addition.<br />

The convenience of selecting one of the different solutions varies from case to case. Clearly,<br />

it depends on the relative prices of each cement type (and of the mineral additions when they<br />

are added directly into the mixer). But even for a given plant it depends on the characteristics<br />

of the mix being designed. For instance, a more efficient, albeit more expensive cement,<br />

might be more convenient for concretes of relatively high strength, but not so for concretes of<br />

lower strengths or when a minimum amount of cement has to be added anyway due to<br />

particular specifications.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 88 of 235


Cement<br />

Concrete<br />

Cement Content in Concrete (kg/m³)<br />

Figure 3<br />

350<br />

RO<br />

GA<br />

U5<br />

U2<br />

300<br />

FAsh<br />

Slag<br />

Filler<br />

250<br />

200<br />

40 45 50 55 60<br />

28-day ISO Mortar Strength (MPa)<br />

Fig. 3 shows that the cement strength, measured on standard mortar, is not a very accurate<br />

indicator to predict the efficiency of the cement in concrete mixes.<br />

3.2 Mineral Additions<br />

Mineral additions can be incorporated into concrete either with the cement (blended<br />

cements) or directly, batched as a component, at the concrete plant.<br />

The convenience of one or the other procedure is a matter of arguments between cement<br />

and concrete producers [2]. For obvious commercial reasons, the former support the<br />

introduction of the addition at the cement plan, on the following grounds:<br />

a) Better control of the quality and quantity of the addition at the cement plant<br />

b) Avoidance of handling, storing and batching another component at the concrete plant<br />

c) Better uniformity in the distribution of the addition when it is incorporated into the<br />

cement<br />

The first two reasons are generally valid, whilst practical experience and some investigations<br />

[3] have shown that it is possible to introduce the addition at the concrete plant without<br />

negative effects on its performance.<br />

The concrete producer, as long as he can find mineral additions of good and uniform quality<br />

and at a reasonable price, will prefer in general to add them at his plant. Indeed, this<br />

procedure will enable him to vary the dosage of the addition according to the characteristics<br />

of the concrete to be produced, thus widening his range of products and optimising costs.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 89 of 235


Cement<br />

Concrete<br />

The use of a mineral addition might lead to cost reductions, depending on its hydraulic<br />

contribution and price; it will be profitable if:<br />

$m dC<br />

---- £ ---- (3)<br />

$c M<br />

where dC is the reduction in Portland cement content achieved by using a quantity M of<br />

mineral addition.<br />

There is an interesting potential to reduce costs through the judicious use of mineral<br />

additions, when such materials of good quality are available at the concrete plant at<br />

convenient prices [4,5].<br />

3.3 Aggregates<br />

In 1976/77 HMC carried out a comparative study, COMBET, among 44 RMC plants of the<br />

'Holderbank' Group [6]. Among other aspects, the quality of the aggregates used in each<br />

plant was compared. Representative samples were taken from each plant, observing the<br />

proportion of the different fractions used, with which mixes with the same composition were<br />

prepared at HMC laboratories (350 kg/m3 of the same cement and a w/c ratio of 0.55).<br />

Figure 4<br />

Compressive Strength (MPa)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Reference concrete - Cement: 350 kg/m 3 , w/c= 0.55<br />

f’cm = 30 ± 3.1 MPa<br />

Slump < 50 mm<br />

Slump = 50 - 250 mm<br />

Aggregates used in 38 RMC GCs<br />

Fig. 4 shows the results of cube strength at 28 days of such mixes; it can be seen that the<br />

direct effect of aggregate quality on the compressive strength is not very significant. It must<br />

be mentioned, however, that those mixes are not strictly comparable because, as indicated<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 90 of 235


Cement<br />

Concrete<br />

in the graph, the mixes had different consistencies. From the consistencies actually<br />

measured it is possible to calculate the consumption of cement required to achieve concretes<br />

of similar consistency and w/c ratio (strength).<br />

Figure 5<br />

Reference concrete – Slump = 70 mm, w/c= 0.55<br />

600<br />

No. of aggregate fractions in use:<br />

2 3 = 4<br />

Cement Consumption (kg/m 3 )<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Aggregates used in 38 RMC GCs<br />

Fig. 5 shows the cement consumption calculated for the 33 aggregates studied, to produce<br />

mixes of slump = 70 mm and w/c = 0.55; they vary along a very large range: 310 to 500<br />

kg/m3. Moreover, it can be observed that the aggregates composed by just two fractions<br />

tend to give the worse results, whilst the cement content tends to decrease when more<br />

fractions are used. The explanation is that, with more fractions available, a better design of<br />

the aggregate grading is possible.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 91 of 235


Cement<br />

Concrete<br />

Figure 6a<br />

Reference concrete – Slump = 70 mm, w/c= 0.55<br />

600<br />

DIN Grading Zone:<br />

< C B-C A-C A-B<br />

Cement Consumption (kg/m 3 )<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Aggregates used in 38 RMC GCs<br />

Figure 6b<br />

100<br />

% Passing<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

C<br />

B<br />

A<br />

too fine<br />

usable<br />

optimum<br />

too coarse<br />

0<br />

0.15<br />

0.3<br />

0.6<br />

2.4<br />

1.2<br />

4.8<br />

Sieve Opening (mm)<br />

9.6<br />

12.5<br />

19<br />

25<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 92 of 235


Cement<br />

Concrete<br />

Fig. 6 shows the calculated cement consumption of the equivalent mixes, indicating the<br />

location of the aggregate grading with respect to the limits established in DIN Standards. In<br />

general, aggregates laying within the more favourable zones require less cement.<br />

One of the conclusions of COMBET study was that in about 30% of the plants investigated, it<br />

would have been possible to improve the grading by a better proportioning of the fractions<br />

actually available in them; in general, these cases corresponded to mixes with rather high<br />

cement consumption.<br />

3.4 Chemical Admixtures<br />

Chemical admixtures have become a common component of RMC mixes. The more<br />

widespread ones are the water reducers/plasticizers (conventional or of high<br />

range/superplasticizers), retarders, air entraining agents, and combinations of them.<br />

Given their high unit price, its utilisation is often confined to the purpose of conferring special<br />

characteristics to the mixes, rather than in reducing costs. In principle, the admixtures with<br />

highest potential for cost reductions are the plasticizers. In a first analysis, a saving in costs<br />

will be feasible if :<br />

$q 1 1<br />

---- £ ---------------- . ----- (4)<br />

$c 1 - dW/W0 Q/C<br />

where dW/W0 is the reduction in water content of the mix generated by the use of the<br />

plasticizer in the doses Q/C. According to [6] a superplasticizer that, used in doses of 1% of<br />

the weight of cement, allows a water reduction of 10%, will be economical only if its price is<br />

less than 11 times that of the cement. In many countries the price of the admixture exceeds<br />

that relation and, thus, they do not allow a reduction in costs.<br />

As the water reduction capability of the plasticizers is very variable and depends also on the<br />

characteristics of the other components and of the mix itself [7] and also as the price varies<br />

widely from country to country, it is difficult to generalise on their potential for cost<br />

optimisation.<br />

4 Optimisation trough Refinement in Strength Level<br />

The strength of concrete is specified on the basis of statistical criteria. This means that, when<br />

a customer is specifying a given compressive strength, he is prepared (or should be) to<br />

accept that a certain proportion p of the concrete delivered is 'defective', i.e., of a strength<br />

below the specified value (Fig. 7). That proportion p varies according to the codes and<br />

standards of every country:<br />

p = 2 % in Switzerland<br />

p = 5 % in the rest of Europe, Australia, New Zealand, Brazil and Argentina<br />

p = 10 % in North-America , most of Latin-America and partially in France<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 93 of 235


Cement<br />

Concrete<br />

Figure 7<br />

Mean<br />

Frequency<br />

Specified<br />

“defective”<br />

Concrete<br />

p<br />

Margin<br />

15 20 25 30 35 40 45<br />

Compressive Strength (MPa)<br />

This means that the producer must design his mixes to reach a mean strength exceeding the<br />

specified strength by a certain Margin. This Margin must compensate the unavoidable<br />

variations of quality during production. As shown in Fig. 8, the higher his variability, the<br />

higher the Margin the producer has to adopt and, consequently, the higher the cost of his<br />

mixes.<br />

Figure 8<br />

Means<br />

Low Variability<br />

Frequency<br />

Specified<br />

High Variability<br />

M<br />

15 20 25 30 35 40 45 50 55<br />

Compressive Strength (MPa)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 94 of 235


Cement<br />

Concrete<br />

Approximately, a reduction of 1 MPa in the standard deviation s of the strength will allow a<br />

reduction in cement content of:<br />

Percentage of 'Defectives'<br />

p<br />

Cement Savings through<br />

1 MPa reduction in s<br />

2% ~ 14 kg/m3<br />

5% ~ 11 kg/m3<br />

10% ~ 9 kg/m3<br />

50% (mean strength) 0<br />

Therefore, the potential for cost reductions by increasing the production controls so as to<br />

diminish the variability is higher in countries where the proportion of 'defectives' accepted is<br />

lower. Obviously, a reduction in the variability often has a cost, due to more controls, better<br />

installations, more components to process, etc.<br />

100<br />

90<br />

Figure 9<br />

Swiss RMC Plant, 1980<br />

% of Total Variability<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Operation<br />

Cement<br />

Aggregate<br />

Testing<br />

0<br />

Consistency<br />

Strength<br />

The question that arises is: how can we diminish the variability in concrete quality?. An<br />

investigation carried out at HMC in 1980 [8], allowed an evaluation of the contribution of<br />

different factors to the variability in consistency and strength in RMC production. Fig. 9<br />

summarises the results obtained; it can be seen that the variability of the components<br />

(cement + aggregates) contributed some 10% of the total variability of strength. It is obvious<br />

that the main source of variability in concrete quality lies in the production process and, in<br />

particular, in the uncompensated variation in moisture of the aggregates.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 95 of 235


Cement<br />

Concrete<br />

Figure 10<br />

50<br />

Mean of 6 tests conducted in 39 Holcim RMC Plants<br />

45<br />

Actual Strength (MPa)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

15 20 25 30 35 40<br />

Required Strength (MPa)<br />

Moreover, the comparative study COMBET (1976/77) also showed that many of the<br />

investigated mixes were overdesigned in strength (Fig. 10), indicating that there is an<br />

interesting potential for cost reduction there.<br />

Probability of accepting the lot<br />

(%)<br />

100<br />

80<br />

60<br />

Figure 11<br />

Mean strength for fck= 30 MPa and s= 4 MPa<br />

40.3 38.2 36.6 35.1 33.4 30.0<br />

AFNOR P18-305<br />

ASTM C94<br />

40<br />

20<br />

0<br />

0.1 0.5 1 2 5 10 20 50<br />

0.1 1 10 100<br />

Actual % of 'defectives' in the lot<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 96 of 235


Cement<br />

Concrete<br />

In defining the right Margin, it is useful to know the O-C Curves of the particular acceptance<br />

criterion by which the concrete supplied is to be judged. More information on the concept of<br />

the O-C Curve can be found in [9]. As an example, Fig. 11 presents the O-C curve for<br />

criterion of AFNOR P 18-305 (10% of defectives) [10] which is based on the fulfilment of the<br />

following two conditions (for non-certified production, fck£ 30 MPa):<br />

where:<br />

fm3 ³ fck + 4 MPa (1)<br />

fi ³ fck - 1 MPa (2)<br />

fm3 : moving average of 3 test results<br />

fck<br />

fi<br />

: specified strength (characteristic strength)<br />

: each individual test result<br />

The O-C curve must be understood as follows. Along the x-axis the actual percentage of<br />

'defectives' (% of concrete with strength below the specified strength fck) is plotted and in the<br />

y-axis the probability of accepting a lot with such amount of 'defectives' is indicated. To make<br />

the graph more meaningful, the mean strength of the lots is plotted on the top x-axis,<br />

assuming an fck= 30 MPa and a standard deviation of 4 MPa.<br />

It can be seen that, if the supplier designs the concrete with exactly the right Margin for 10%<br />

"defectives" (Margin= 1.28*s ) he will still get a relatively high proportion of his concrete<br />

rejected (about 20%). If he wants to ensure a lower proportion of rejections , say 5%, he has<br />

to overdesign his mix to a Margin= 1.75*s. For s=4 MPa this means about 2 MPa more of<br />

mean strength and some extra 15 kg/m3 of cement in his mix. If plants are automated, one<br />

can imagine a scenario where the Margin can be adapted to the characteristics of the client.<br />

A client that carries out a strict quality control will receive concrete designed with a higher<br />

margin to avoid rejections, whilst another one that does not, will receive the strictly minimum<br />

margin of 1.28*s (going below this margin will constitute a dishonest practice and must not<br />

be done).<br />

As a comparison, Fig. 11 shows also the acceptance criterion of ASTM C94, which is based<br />

on the fulfilment of the following two conditions [11]:<br />

fm3 ³ fck (3)<br />

fi ³ fck - 3.4 MPa (4)<br />

It can be seen that the ASTM C94 criterion is more favourable to the producer (less risk of<br />

rejection of lots with up to 10% 'defectives') but is a bit unfair to the customer (higher risk of<br />

acceptance of lots with more than 10% 'defectives'). This illustrates the consequences that<br />

the acceptance criterion has on the risks for both the producer and the consumer of<br />

concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 97 of 235


Cement<br />

Concrete<br />

5 CONCLUSIONS<br />

Several factors known for their potential in achieving cost reductions in ready-mix concrete<br />

mixes have been discussed.<br />

It has been seen that the cement type may have a significant influence on the cement<br />

consumption and, therefore, on the cost of the mixes.<br />

Certain potential also exists in adding mineral additions at the concrete plant (when they are<br />

suitable and available), which also opens possibilities in products differentiation.<br />

It is considered that the aggregates, given their enormous incidence in the cement<br />

consumption of the mixes, offer the best potential for mix optimisation and cost reductions.<br />

The potential of achieving cost reductions through the utilisation of chemical admixtures<br />

depends strongly on their efficiency and price.<br />

Important savings can be achieved by fine-tuning the mix design to strict margins that cover<br />

the production variability and certain unforeseeables. A good knowledge of the risks implied<br />

in the compliance criteria applicable to RMC is essential to that end.<br />

Appendix A shows the result of a BCM (Better Cost Management) initiative undertaken by St.<br />

Lawrence Group in Canada, with the support of HMC, to find ways to reduce costs in Ready-<br />

Mixed Concrete production. Huge potential for cost savings was detected in many of the<br />

areas discussed in this paper.<br />

6 REFERENCES<br />

[1] Torrent, R. and Jornet, A., "Covercrete Study, Part I", MA Report 90/3815/E, 1990.<br />

[2] Braun, H., "Composite cements and hydraulic additives for ready-mixed concrete",<br />

5th Stone & Ready-Mixed Concrete Conference, paper No. 14, Lucerne, 1983.<br />

[3] Longo, A. and Torrent, R., "Methods of addition of blast-furnace slag: their effect on<br />

the compressive strength of mortars and concretes", ACI SP-91, 1986, v.2, pp.1381-1400.<br />

[4] Frenzer, G., "Reduction of Portland cement in concrete - how far can we go?", 11th<br />

Stone & Ready-Mixed Concrete Conference, San José de Costa Rica, Nov. 1995.<br />

[5] Ruiz Rico, Enrique, "Substitution of cement by fly ash in concrete and mortars",11th<br />

Stone & Ready-Mixed Concrete Conference, San José de Costa Rica, Nov. 1995.<br />

[6] "Comparative study on ready-mixed concrete 1976/77", MA Report 77/2486/E, 1977.<br />

[7] Gebauer, J., "Chemical admixtures - Essential constituents of concrete. Prospects<br />

and limitations", 6th Stone & Ready-Mixed Concrete Conference, paper No.2, Mons, 1985.<br />

[8] Dratva, T., Gebauer J. et al., "Study on material and operating factors in the readymixed<br />

concrete production", MA Report 81/2810/E, 1981.<br />

[9] Torrent, R., "Concrete quality control: compressive strength. Producer's and<br />

consumer's approach and risks", Workshop on Concrete Technology, St. Lawrence Cement,<br />

Canada, February 8-12, 1988. MA Report 88/3524/E.<br />

[10] AFNOR P 18-305: "Béton prêt a l'emploi", Normalisation Française, Dec. 1994.<br />

[11] ASTM C94: "Standard Specification of Ready-Mixed Concrete", May 1994.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 98 of 235


Cement<br />

Concrete<br />

Concrete Admixtures<br />

KEY COMPONENTS OF MODERN CONCRETE<br />

P.-C. AITCIN<br />

Université de Sherbrooke, Sherbrooke, Canada<br />

M. BAALBAKI<br />

Holcim Group <strong>Support</strong> - Corporate Product Development<br />

1. ABSTRACT..................................................................................................................... 100<br />

2. INTRODUCTION............................................................................................................. 100<br />

3. AN ARTIFICIALLY COMPLICATED TERMINOLOGY................................................... 101<br />

4. DISPERSING ADMIXTURES KEY COMPONENTS FOR CONCRETE<br />

DURABILITY................................................................................................................... 103<br />

4.1 Durable concrete.................................................................................................... 103<br />

4.2 Why cement needs to be dispersed?..................................................................... 105<br />

5. CEMENT/ADMIXTURE COMPATIBILITY...................................................................... 106<br />

6. ENTRAINED AIR ............................................................................................................ 109<br />

7. CONCLUSIONS.............................................................................................................. 110<br />

8. REFERENCES................................................................................................................ 111<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 99 of 235


Cement<br />

Concrete<br />

1 Abstract<br />

Concrete admixtures are still too often perceived as mysterious ingredients for making<br />

concrete. Their unseen chemical nature, their mysterious modes of action too often cast<br />

them as the results of alchemy rather than high-tech chemistry. While some admixtures<br />

were originally industrial by-products and only marginally beneficial, more and more modern<br />

products are specially prepared or synthecized for the concrete industry. Their action is<br />

more specific and more engineered, making them more efficient. We can now state that the<br />

most recent technological developments in concrete technology rely on enhanced admixture<br />

efficiency rather than on improvements in cement manufacturing.<br />

Amongst all the admixtures currently used by the concrete industry, superplasticizers play a<br />

major role: it is now possible to make workable or flowing concretes with reduced<br />

water/cement ratios with the mix containing just enough, or even less, water needed to fully<br />

hydrate all the Portland cement particles. The microstructure of the resulting hydrated<br />

cement paste is very compact, which does away with the weak transition surrounding the<br />

aggregates, increases significantly the compressive strength, and drastically reduces<br />

concrete permeability to outside aggressive agents.<br />

However, it should be emphasized that these improvements in the quality of concrete are<br />

only achieved if admixtures are correctly used and if the concrete is placed and cured<br />

properly. Modern admixtures will not transform a poor mix, that has been badly placed and<br />

not cured, into a high-performance concrete. There are no miraculous products, only<br />

chemical ingredients that must be used judiciously to transform an already good concrete<br />

into a better one.<br />

Keywords: admixture, cement/superplasticizer compatibility, concrete, curing, dispersing<br />

admixtures, durability, flowing concrete, microstructure, permeability, superplasticizer<br />

2 Introduction<br />

The idea to introduce chemicals when making concrete is not new in itself; Latin authors<br />

reported that a hundred years before Christ, blood or eggs were used to improve concrete<br />

properties [1]. This practice was still in use during the Middle Ages since Villard de<br />

Honnecourt, a cathedral builder, was recommending ca 1230 to moisten cement with linseed<br />

oil to make a vessel that could hold water [2]. Closer to the present day, it was just before<br />

the Second World War that the beneficial use of chemical admixtures was re-emphasized<br />

accidentally in the USA: a dispersant used to thoroughly disperse some carbon black to<br />

differentiate the color of the concrete of the center line of a 3-lane concrete highway,<br />

entrained as a secondary effect some tiny air bubbles, which made the concrete easier to<br />

place and more durable to freezing and thawing. This fortuitous discovery led to the birth of<br />

a successful business in cement and concrete technology, the admixture business.<br />

In spite of the fact that admixture business is a multimillion dollar business, it is then<br />

surprising to find very few comprehensive books about them [3][4][5][6][7], similarly there are<br />

very few international conferences on admixtures [8][9] and even in 1994 a great number of<br />

engineers and scientists are still questioning the beneficial use of achnixtures in cement and<br />

concrete technology.<br />

Is there actually a science of admixtures? Would it not be better to simply ban the use of<br />

admixtures in modem concrete technology rather than to face all the compatibility problems<br />

that field engineers increasingly have to deal with? Would it not be better to simply add more<br />

cement and more water in modern concrete rather than trying to complicate the<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 100 of 235


Cement<br />

Concrete<br />

manufacturing and placing of such a simple material as concrete by introducing in it "minute"<br />

amounts of mysterious chemicals?<br />

However, when the role played by admixtures on concrete properties is well understood, it is<br />

better appreciated that admixtures are essential components of modern concrete. Of course,<br />

the chemical system that results from the simultaneous use of a very complex mineral<br />

system (Portland cement) and several organic polymers is very complex but this system can<br />

be mastered when the mode of cement/admixtures and admixtures/admixtures interaction is<br />

well understood.<br />

Concrete technology recently evolved more rapidly than cement, concrete, and admixtures<br />

standards, so that it is now urgent to correct the situation if problems are not likely to be<br />

created artificially and if we want to benefit from the opportunities to drastically improve<br />

concrete technology and concrete durability by adequately using admixtures. Admixtures are<br />

no longer made from cheap by-products but rather composed of elaborate polymers. No<br />

longer is it simply a matter of business; it is a matter of technology and a matter of science.<br />

3 An artificially complicated terminology<br />

A major source of confusion in the field of admixture is the artificially complicated terminology<br />

that has developed through the years not only in the commercial literature but also in<br />

standards. It can be agreed that this abundance of terms to describe the multitude of<br />

admixtures offered to the concrete industry is perhaps being artificially created to fulfill<br />

marketing strategies. For example, in the booklet Design of Concrete Mixtures edited by the<br />

Portland Cement Association in chapter 6: Admixtures for concrete [10], the following types<br />

of admixtures are listed: accelerators, air detrainers, air entraining admixtures, alkalireactivity<br />

reducers, bonding admixtures, coloring agents, corrosion inhibitors, dampproofing<br />

admixtures, finely divided mineral admixtures (cementitious, pozzolans, pozzolanic and<br />

cementitious and nominally inert) fungicides, germicides, insecticides, gas formers, grouting<br />

agents, permeability reducers, pumping aids, retarders, superplasticizers, water reducers,<br />

and workability agents.<br />

French technical literature in the admixture field is also very rich. In his book on Concrete<br />

Admixtures VENUAT [6] lists: plasticizers, water reducers, fluidizers, air entraining agents,<br />

mass dampproofing admixtures, set accelerators, hardening accelerators, set retarders,<br />

foam formers, gas formers, admixtures for grouting, coloring agents, stiffeners for shotcrete.<br />

This very rich terminology is also found in standards, for example ASTM C494 [11]<br />

recognized seven different types of water reducers identified by specific letters from A to G<br />

as shown in Table 1, and the new European standards EN934-2 recognize nine types of<br />

admixtures listed in Table 2.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 101 of 235


Cement<br />

Concrete<br />

Table 1 — Different types of water reducers according to ASTM C494 standard<br />

Type Function Effect<br />

A<br />

B<br />

C<br />

D<br />

E<br />

F<br />

G<br />

Water reducer<br />

Retarder<br />

Accelerator<br />

Water reducer and retarder<br />

Water reducer and accelerator<br />

Water reducer - high range<br />

Water reducer - high range -<br />

and retarder<br />

Reduce water demand at least 5%<br />

Retard setting time<br />

Accelerate setting and early-strength<br />

development<br />

Reduce water (minimum 5%) and retard<br />

set<br />

Reduce water (minimum 5%) and<br />

accelerate<br />

Reduce water demand (minimum 12%)<br />

Reduce water demand (minimum 12%)<br />

and retard set<br />

This very rich terminology creates confusion rather than clarification so that for many people<br />

admixture terminology is still more related to alchemy rather than physico-chemistry.<br />

This situation is the result of different factors: the secrecy of admixtures formulation brought<br />

to a paroxysm by admixture companies, and the use of industrial by-products as the main<br />

source of chemicals admixtures. As these industrial by-products most of the time contain<br />

several active chemical components with respect to cement hydration, their action is a<br />

multifunctional one rather than a specific one. Finally, it should be recognized that such a<br />

situation prevails due to a lack of scientific understanding of the mode of action of these<br />

admixtures.<br />

If this very complicated terminology is scrutinized more closely it is very easy to radically<br />

simplify, if admixtures are classified according to the main physicochemical effects they<br />

produce on the reaction of hydration and the properties of fresh and hardened concrete. For<br />

example, DODSON [5] proposes to group all concrete admixtures into four different<br />

categories.<br />

Admixtures can function by several mechanisms.<br />

1. Dispersion of the cement in the aqueous phase of concrete.<br />

2. Alteration of the normal rate of hydration of the cement, in particular the tricalcium<br />

silicate phase.<br />

3. Reaction with the by-products of the hydrating cement, such as alkalies and calcium<br />

hydroxide.<br />

4. No reaction with either the cement or its by-products."<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 102 of 235


Cement<br />

Concrete<br />

This presentation will only deal with dispersing and air entraining admixtures that play a key<br />

role on concrete durability.<br />

Type<br />

Table 2 — Different types of admixtures proposed by EN 934-2 standard<br />

Action<br />

Water reducing/plasticizing Admixture which without affecting the consistence, permits<br />

a reduction in the water content of a given concrete mix, or<br />

which, without affecting the water content increases the<br />

slump/flow or which produces both effects simultaneously.<br />

High range water<br />

reducing/<br />

superplasticizing<br />

admixture<br />

Water reducing admixture<br />

Air entraining admixture<br />

Admixture which, without affecting the consistence, permits<br />

a high reduction in the water content of a given concrete<br />

mix, or which, without affecting the water content increases<br />

the slump/flow considerably, or which produces both effects<br />

simultaneously.<br />

Admixture which reduces the loss of water by reduction of<br />

bleeding<br />

Admixture that allows a controlled quantity of small<br />

uniformly...<br />

Set accelerating admixture Admixture which decreases the time to commencement of<br />

transition of the mix from plastic to rigid state<br />

Hardening<br />

admixture<br />

accelerating<br />

Admixture which increases the rate of development of early<br />

strength in the concrete, with or without affecting the setting<br />

time<br />

Set retarding admixture<br />

Water resisting admixture<br />

Multifunction admixture<br />

Admixture which extends the time to commencement of<br />

transition of the mix from the plastic to rigid state<br />

Admixture which reduces the capillarity absorption of<br />

hardened concrete<br />

Admixture that affects several properties of fresh and/or<br />

hardened concrete by performing more than one of the<br />

main functions<br />

4 Dispersing admixtures key components for concrete<br />

durability<br />

4.1 Durable concrete<br />

When concrete is subjected to external aggression the most effective way to decrease the<br />

intensity of this aggression is to reduce its porosity and permeability [12]. This is why the<br />

water/cement ratio has always been the controlling factor of concrete durability. The<br />

water/cement ratio, and not strength has to remain the pivotal controlling criterion for<br />

designing durable concrete structures [13].<br />

Concrete durability is becoming a subject of major concern in many countries, a great<br />

number of international conferences are held regularly on this subject, a great number of<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 103 of 235


Cement<br />

Concrete<br />

papers analyze the causes of these repeated failures but there are very few papers dealing<br />

with the crucial question — how to make a durable concrete that could be used either to<br />

repair or rehabilitate damaged structures or to build new durable ones.<br />

A durable concrete structure is built with workable low water/cement concrete. A workable<br />

low water/cement concrete is made using appropriately an efficient dispersing admixture.<br />

Experience shows that, for most of the time, it is impossible to make a workable concrete<br />

having a slump higher than 100 mm by simply mixing water and cement at a water/cement<br />

ratio lower than 0.40, while it is possible to make an almost fluid concrete having a slump<br />

higher than 200 mm having a water/cement ratio below 0.30 when using a superplasticizer.<br />

When the microstructure of a high and low W/C concretes are observed under a scanning<br />

electron microscope it is possible to visualize the drastic consequence of the reduction of the<br />

water/cement ratio on concrete microstructure and consequently on concrete durability [14].<br />

Figure 1 depicts schematically the microstructure of 0.65 and 0.25 W/C cement pastes. In<br />

this schematic representation, the surface representing the water (pale grey) and the surface<br />

representing the cement (dark grey) are in the same proportions as the water/cement ratio by<br />

mass.<br />

Fresh Cement Paste<br />

Anhydrous<br />

cement<br />

grains<br />

Water<br />

C-S-H<br />

0.65 0.25<br />

Ettringite<br />

Lime<br />

Hydrated Cement Paste<br />

Fig. 1 — The impact of water-cement ratio on concrete paste microstructure<br />

The cement particles are far from each other in the 0.65 W/C paste and very close to each<br />

other in the 0.25 one. In the latter paste hydration products fill the gap between cement<br />

particles very rapidly and achieve a strong and dense microstructure. On the other hand, in<br />

the case of the 0.65 water/cement ratio paste, hydration products have to develop over long<br />

distance before reaching the hydration products from another cement particle. Moreover, as<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 104 of 235


Cement<br />

Concrete<br />

there is more water than necessary to fully hydrate all the cement particles, this remaining<br />

water constitutes a continuous medium through which aggressive agents will penetrate and<br />

attack concrete. If for any reason this 0.65 concrete dries, the network of empty pores<br />

constitute a pathway for the penetration of gas, bacterias, fluids, etc.<br />

This difference in microstructure has two very important consequences on strength and<br />

permeability: concrete strength increases greatly and permeability decreases when the W/C<br />

decreases. However, a 0.25 water/cement ratio concrete is not totally impervious: when<br />

such a concrete is tested for rapid chloride ion permeability in accordance with AASHTO T-<br />

277 (Standard Method of Test for Rapid Determination of the Chloride Permeability of<br />

Concrete) it is found that chloride ions can pass through such low water/cement concrete but<br />

at a very low rate when compared to 0.65 W/C concrete. Figure 2 gives a correlation<br />

between W/C and chloride ion permeability for some concretes [15].<br />

4.2 Why cement needs to be dispersed?<br />

Cement particles have a surface containing a great number of unsaturated electrical charges<br />

so that they have a strong tendency to flocculate when they are in contact with water [16].<br />

Cement flocs trap a part of the mixing water which is no longer available to fluidify the mix<br />

with the practical result that it is not possible to make a workable concrete having a<br />

water/cement ratio lower than 0.40.<br />

Rapid chloride permeability<br />

(coulomb)<br />

8000<br />

6000<br />

4000<br />

2000<br />

1d cure<br />

7d cure<br />

0<br />

0.20 0.30 0.40 0.50 0.60 0.70 0.80<br />

Water/cement ratio<br />

Fig. 2 — Rapid chloride permeability of concrete as a function of w/c and the length of<br />

curing<br />

Water dispersing admixtures are polymers that reduce this flocculation, so that more or all<br />

the water introduced initially in the mix is available to make the concrete workable [5]. Water<br />

dispersing molecules, as it will be seen, also contributes to lubrication process of the<br />

concrete. The most efficient polymer, presently used as water dispersing admixtures are the<br />

so-called superplasticizers or high range water reducers [17].<br />

Of course, the superplasticizer/cement interaction depends on the physicochemical<br />

parameters of the superplasticizer and of the cement. In general terms, this interaction can<br />

be divided into two main effects: a physical effect that is also found when a superplasticizer<br />

is used to disperse a non cementitious fine powder, and a chemical effect related to the<br />

reaction of the superplasticizer with the most reactive sites of the hydrating cement particles.<br />

The physical effect is a three fold one [18]:<br />

1. adsorption of the superplasticizer molecule by Van der Waals and electrostatic forces on<br />

cement particles.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 105 of 235


Cement<br />

Concrete<br />

2. surface charging which induces long range interparticle repulsive forces<br />

3. steric hindrance between adsorbed polymer molecules and neighboring particles.<br />

All these physical actions tend to defloculate and disperse cement particles [18].<br />

However, the action of superplasticizer molecules is not only physical. Their sulfonate sites<br />

can react with C 3 A. It has been found that the shorter polynaphthalene molecules have a<br />

greater tendency to react with C 3 A so that an efficient superplasticizer must contain the<br />

minimum amount of short polymers (di and trimers).<br />

This double action of a superplasticizer on cement particles reveals itself when observing the<br />

shift of heat development curves under increasing dosages of superplasticizers. Upon first<br />

addition of a superplasticizer the heat flux peak is shifted substantially to shorter times due to<br />

improved water/cement contact upon deflocculation and a more homogeneous dispersion of<br />

the cement particle occurs. Upon further addition of superplasticizers some retardation is<br />

observed; retardation that increases with superplasticizer dosage. This retardation is<br />

observed for both the beginning of the accelerated phase and the maximum of the thermal<br />

peak. This is another point which makes low molecular weight superplasticizer not at all<br />

interesting from a practical point of view. This retardation is the result of the adsorption of<br />

superplasticizer molecule on C3S sites as the superplasticizer dosage increases. It has<br />

been found that low molecular weight molecules induce more pronounced retardation [18].<br />

From this brief discussion about the mode of action of superplasticizer molecules it can be<br />

concluded that an efficient superplasticizer must have a high degree of sulfonation to offer<br />

many electrically active sites in order to disperse all the cement particles. Moreover, it has<br />

been found that among the two positions that the SO3 -- sites can occupy in the naphthalene<br />

rings, the so-called β site is more efficient to disperse cement particles. Secondly, a<br />

superplasticizer should contain as few short polymers (di and trimers) as possible and as<br />

many high molecular weight chains as possible without too much crossliriking in order to<br />

confer to the polymeric chain as a large covering effect as possible [19].<br />

In order to control the quality of commercial superplasticizers, standards have been<br />

developed. Most of these standards are in fact extensions of standards developed for water<br />

reducers, there are not always best suited to evaluate objectively the actual efficiency of a<br />

superplasticizer.<br />

5 Cement/admixture compatibility<br />

It is somewhat surprising that cement/admixture compatibility problems are becoming more<br />

and more frequent, especially in the field of superplasticizers. In the past, few cases of poor<br />

compatibility between a cement and a water reducer were reported. This kind of problem<br />

was quite rare. In the few cases reported in the literature [20][21][22] it was found that the<br />

rapid slump loss observed in the field were most of the time associated with the presence of<br />

anhydrite as a source of calcium sulfate to control Portland cement setting. Where such<br />

cements were used in conjunction with a lignosulfonate based water reducer the anhydrite<br />

did not pass into solution fast enough so that not enough SO4 -- ions were available to react<br />

with the C3A to form ettringite. In such a case, the slump loss was due to a flash set<br />

situation i.e. the rapid formation of hydrogarnet, in spite of the fact that the total SO3 content<br />

of the cement complied to cement standards.<br />

The first documented cement/superplasticizer compatibility problems reported in the<br />

literature have been explained in the same way [23], when it was not a specific deficiency of<br />

the commercial superplasticizer that was the principal cause of the compatibility problem (low<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 106 of 235


Cement<br />

Concrete<br />

degree of sulfonation, high percentage of cc site sulfonated, high percentage of di or trimers,<br />

low degree of crosslinking).<br />

It has been also observed that cement/superplasdcizer compatibility problems seem to be<br />

more frequent in low water/cement ratio concretes [24]. This type of concrete contains per<br />

unit volume more C3A and less water where SO4 -- ions from the calcium sulfate bearing<br />

minerals can dissolve. In some cases the initial hydrating conditions are such that there are<br />

not enough SO4 -- ions dissolved in water, or the calcium sulfate bearing minerals do not<br />

dissolve fast enough to neutralize the available C3A. Such a situation can occur because<br />

the optimization of the calcium sulfate content of Portland cement is done at the cement plant<br />

in conditions that are quite different from the ones that prevail in low W/C concrete. The<br />

calcium sulfate content is presently optimized on a paste or a mortar having a W/C ratio of<br />

about 0.50 using only pure water [25].<br />

In such a system there is less C3A and more water by unit volume than in a low W/C<br />

concrete and the admixture could interfere with the solubility rate of the different sources of<br />

SO4 -- .<br />

Moreover, a system having a W/C = 0.50 is rich enough in water so that the water in fact<br />

controls most of the rheological properties of the mix. In such a system hydrating cement<br />

grains are far enough from each other so that they do not interfere very much and affect the<br />

rheological properties of the system.<br />

In some cases, the rapid rate of consumption of water during the formation of ettringite (C3A<br />

. 3 CaSO4 . 32 H2O) decreases drastically the already quite low amount of mixing water. To<br />

be more specific, typical HPC concretes are mixed using 125 to 145 L/m 3 water, that is 20 to<br />

40 liters less than a typical 0.50 W/C ratio air-entrained concrete or 40 to 60 liters less than a<br />

typical 0.50 W/C ratio non air-entrained concrete. Therefore any rapid consumption of water<br />

can lead to a situation where there is not enough water to ensure a proper workability in a<br />

low W/C ratio concrete.<br />

From a practical point of view four typical situations can be found when studying the rheology<br />

of a low W/C grout with a Marsh cone. Figure 3a, b, c and d illustrate these different<br />

situations where the superplasticizer dosage and the Marsh cone flow time have been<br />

reported respectively along x and y axes [26].<br />

Figure 3a represents the case of a perfectly compatible combination of cement and<br />

superplasticizer, the superplasticizer dosage at the point of saturation is low and the 60<br />

minutes curve is close to the 5 minutes one. Figure 3b on the contrary represents a case of<br />

incompatibility, the superplasticizer dosage at the saturation point is quite high and the 60<br />

minutes curve is far up the 5 minutes. Very often in such a case the grout stops to flow very<br />

rapidly, often as fast as 15 minutes after the beginning of the mixing.<br />

Figure 3c and 3d represent intermediate cases. In Fig. 3c the 5 minutes curve is similar to<br />

the 5 minutes curve presented in Fig. 3a but the 60 minutes curve is similar to the 60 minutes<br />

curve presented in Fig. 3b. In Fig. 3d the 5 minutes curve is similar to the 5 minutes curve<br />

presented in Fig. 3b and the 60 minutes curve has a relative position to the 5 minutes curve<br />

similar to the situation prevailing in Fig. 3a.<br />

In another paper [26] each typical case has been explained in greater detail and some<br />

solutions involving mainly the addition of an appropriate amount of retarder have been<br />

proposed so that the cement and superplasticizer combinations presented in Fig. 3b, 3e and<br />

3d behave, in the best cases, as similarly as possible to the combination presented in Fig.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 107 of 235


Cement<br />

Concrete<br />

3a. Unfortunately it is not always possible to adequately correct the situation to achieve the<br />

flow time curves as those presented in Fig. 3a.<br />

Presently it is not possible to develop polynaphthalene or polymelamine sulfonates that work<br />

with any Portland cement at any water/cement ratio due to the too great a variability of the<br />

solubility rate of the different calcium sulfates used to control the setting of the commercial<br />

cements [24]. However, when some attention is given to this important characteristic at low<br />

W/C ratio, it is possible to manufacture Portland cement that is perfectly compatible with<br />

good polynaphthalene or polymelamine superplasticizers. If such an ideal situation does not<br />

exist it is better to develop a blended superplasticizer whose composition is adjusted to the<br />

particular cement [27][28]. The development of such a superplasticizer should be done<br />

through a cooperative development program involving the admixture manufacturer, the<br />

cement producer and the user.<br />

120<br />

100<br />

(a)<br />

W/C = 0.35<br />

T = 22 °C<br />

120<br />

100<br />

(b)<br />

W/C = 0.35<br />

T = 22 °C<br />

Marsh flow time (sec)<br />

80<br />

60<br />

40<br />

0.0<br />

120<br />

100<br />

0.8<br />

1.6<br />

(c)<br />

2.4<br />

60 min<br />

3.2<br />

5 min<br />

W/C = 0.35<br />

T = 22 °C<br />

4.0<br />

80<br />

60<br />

40<br />

0.0<br />

120<br />

100<br />

0.8<br />

1.6<br />

(d)<br />

2.4<br />

60 min<br />

5 min<br />

3.2<br />

W/C = 0.35<br />

T = 22 °C<br />

4.0<br />

80<br />

60 min<br />

80<br />

60 min<br />

60<br />

5 min<br />

60<br />

5 min<br />

40<br />

0.0<br />

0.8<br />

1.6<br />

2.4<br />

3.2<br />

4.0<br />

40<br />

0.0<br />

0.8<br />

1.6<br />

2.4<br />

3.2<br />

4.0<br />

Percent superplasticizer (by wt. solids)<br />

Fig. 3 — Rheological characteristics of different cement/superplasticizer combinations<br />

Up to now, there have been very few incentives to invest time, effort and money to solve this<br />

kind of problem, as HPC still represents a very small part of the concrete market. However,<br />

when the concrete will be designed essentially for durability and not uniquely for strength,<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 108 of 235


Cement<br />

Concrete<br />

this situation may change and it will be possible to see more and more people involved in this<br />

tricky business: trying to optimize a cement/superplasticizer combination at low W/C ratio<br />

values. Cement producers will find that it is not too difficult or too costly to adjust the<br />

solubility rate of the sulfates and to control it [25].<br />

To acheive such an ideal situation, it is only necessary to recognize that admixtures in<br />

general and superplasticizers in particular are, when properly used, essential constituents of<br />

durable concrete as important as cement or water. In fact, due to the very efficient<br />

dispersant properties of polynaphthalene and polymelamine superplasticizers it is possible to<br />

make flowing concretes having very low W/C ratios in the order of 0.30, or even lower, that<br />

are particularly impermeable [14]. These concretes are not penetrated easily by aggressive<br />

agents so that they will last longer than 0.50 to 0.70 W/C ratio concretes that are indirectly<br />

specified when designers are designing their structures with 20 to 40 MPa concretes.<br />

It would be unwise to be over-confident on the durability of low W/C concretes and take<br />

advantage of their low permeability to decrease the thickness of the concrete cover or to<br />

eliminate early curing. Admixtures are not miraculous products, they will never transform a<br />

bad concrete poorly cured into a good one, but they can transform a good concrete properly<br />

cured into a better and more durable one.<br />

It is amazing to see that as the 21 st century approaches, there are still engineers<br />

recommending the use of high W/C ratio concretes which are porous and permeable, and<br />

later they complain that these concretes are subjected to carbonation, that they do not<br />

protect adequately reinforcing steels from corrosion, and that they are not resistant to<br />

freezing and thawing. Would it not have been better to start specifying a impermeable<br />

concrete having a low W/C ratio and try to use efficiently the extra megapascals brought by<br />

this type of concrete? This is the future of concrete.<br />

6 Entrained air<br />

The intentional entrainment of air bubbles in an ordinary concrete is always associated with<br />

the idea to improve its durability to freezing and thawing. However, there are still some who<br />

do not accept this idea. But, generally, it is accepted that when the bubble system has some<br />

well defined characteristics (spacing factor, specific surface, diameter, total volume) it is a<br />

necessary condition but not always sufficient because, and that is often forgotten, these<br />

characteristics have to be developed in a concrete having a W/C ratio lower than a given<br />

value depending on the harshness of the environmental conditions. These maximum W/C<br />

ratios values vary from one country to another and usually they are in the 0.40 to 0.55 range.<br />

Of course, the introduction of a certain amount of air in a concrete results in a decrease in its<br />

compressive strength. It is usually admitted as a rule of thumb that each per cent of air<br />

decreases compressive strength by 5 per cent. This means that a given non air entrained<br />

concrete that contains 2 per cent of entrapped air and that tests 40 MPa will not test more<br />

than 32 MPa if its air content is increased to 6 per cent by the addition of an appropriate<br />

amount of air entraining admixture. This means that as long as concrete structures are<br />

designed on a strength basis, an air entrained concrete must always have a lower W/C ratio<br />

than a non air entrained concrete having the same design strength. This is a valid point for<br />

the purpose of durability.<br />

A second interesting effect of entrained air is that its beneficial effect on concrete rheology<br />

and workability is all too often forgotten. Air entrained concretes are more creamy and less<br />

prone to breeding. Moreover, the entretainment of air bubbles in the concrete results in an<br />

increase of the slump, or in the use of less mixing water if the slump has to remain constant.<br />

For example, ACI standard 211 "Standard Practice for Selecting Proportions for Normal,<br />

Heavyheight and Mass Concrete" [29] recommends to reduce by 15 to 25 liters the amount<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 109 of 235


Cement<br />

Concrete<br />

of water to produce a concrete of a given strength made with a coarse aggregate having a<br />

given maximum size and a sand of a given fineness modulus. This reduction in the amount<br />

of mixing water necessary to obtain a given slump has a positive effect on the water/cement<br />

ratio needed to obtain a given strength.<br />

The presence of a discontinuous air bubble system has also a positive action on the<br />

capillarity of concrete and on the development of microfissures within the hydrated cement<br />

paste. All these secondary effects of entrained air contribute to some extent to make air<br />

entrained concrete of a given strength more durable than a non air entrained concrete of the<br />

same strength. But, is this well established situation still valid when dealing with very low<br />

W/C concretes? Is it appropriate to add entrained air in low W/C concretes to make them<br />

more durable?<br />

Presently it is not clearly accepted by all the scientific community that HPC with low W/C<br />

ratios needs any air at all to be freeze thaw resistant [30]. Of course, it is found when<br />

performing procedure A of ASTM C 666 Standard, that HPC having W/C higher than 0.25<br />

needs entrained air to pass successfully the 300 freezing and thawing cycles under water<br />

[31], but there are more and more people questioning the severity of this test to decide if a<br />

field concrete is freeze thaw resistant [32].<br />

However, field experience developed during the construction of several high performance<br />

concrete bridges have shown an unexpected advantage of entrained air when making HPC<br />

[26][27]. The workability of RPC is considerably improved by the addition of entrained air, as<br />

well as its finishability. Air entrained mixes are sheared much more easily, they are less<br />

sticky, they can be placed much more easily and are less thixotropic, which is translated in<br />

the field by significant gains in productivity during placing. Up to now it has not been<br />

possible to quantify this type of improvement but it is hoped that with the recent development<br />

of different workabilimeters by different researchers it will be possible to designate numbers<br />

on this qualitative appreciation.<br />

The authors are convinced that in the future HPC will increasingly contain some entrained air<br />

(3 to 5 per cent) not to improve their durability, but to improve their workability. This<br />

improvement in the workability of HPC will have beneficial effects, particularly on the quality<br />

of the covercrete and indirectly on HPC durability.<br />

7 Conclusions<br />

Concrete's resistance to exterior aggressive agents is intimately linked to the facility with<br />

which it can be penetrated by these aggressive agents. This facility of penetration depends<br />

among other factors directly on the compactness of the hydrated cement paste. It is also<br />

directly linked to the water/cenient or water/cementitious ratio of the concrete mixture. Due<br />

to the very efficient dispersing properties of polynaphthalene and polymelamine<br />

superplasticizers it is now possible to make durable concretes having W/C equal to 0.30 or<br />

even lower; as such concretes are almost impervious they will be much more durable than<br />

normal concrete. They are, of course, stronger than normal strength concrete and designers<br />

will have to learn how to use efficiently these extra megapascals needed to insure better<br />

durability to concrete structures.<br />

Unfortunately, not all current commercial cements have been optimized in terms of the<br />

solubility rate of their SO4 so that so-called cement/superplasticizer compatibility problems<br />

are more likely to occur in the future as the W/C ratio of concrete used decreases. However,<br />

this situation can be easily corrected when the actual cause of the incompatibility has been<br />

found. Most of the time, it is found that there is a discrepancy between the quantity and the<br />

reactivity of the C3A present in the mixture and the amount of SO4 -- ions going into solution<br />

within the mixing water.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 110 of 235


Cement<br />

Concrete<br />

Air entraining agents will be more and more incorporated intentionally in low W/C concretes<br />

not to specifically improve their freeze thaw durability, but rather to improve their workability.<br />

As soon as a HPC concrete having a W/C ratio of 0.30 or less contains 3 to 5% of entrained<br />

air a marked change in its workability appears. Experience shows that when high quality<br />

admixtures are used with a cement well balanced in terms of C3A and soluble sulfates there<br />

is no problem to entrain the desired amount of air in a superplasticized low W/C ratio<br />

concrete.<br />

Admixtures, and more specifically superplasticizers, are not miraculous products. They will<br />

never transform a bad concrete mixture into a good one, but can be used to transform a<br />

good concrete into a better one; a concrete that can be placed more easily, a more compact<br />

and more impermeable concrete that will offer greater resistance to the penetration of<br />

aggressive agents in as much as this concrete is cured properly.<br />

8 References<br />

1. Mindess, S., Youg, J.F., (1981) "Concrete", Prentice Hall Englenwod Cliffs NJ, pp. 15.<br />

2. Garrison, E.G. (1991) A History of Engineering and Technology, CRC Press, Boca<br />

Raton FA, pp. 106.<br />

3. Ramachandran, V.S. (1984) Concrete Admixture Handbook, Noyes Publications, Park<br />

Ridge, NJ, U.S.A., 626 pp.<br />

4. Rixon, M.R. and Mailvaganam, N.P. (1978) Chemical Admixtures for Concretes, E &<br />

F.N. Spon, New York, U.S.A., 306 pp.<br />

5. Dodson, V. (1990) Concrete Admixtures, Van Nostraud Reinhold, New York, U.S.A., 211<br />

pp.<br />

6. Venuat, M. (1984) Adjuvants et Traitements, Edited by M. Venuat, 66 avenue C.<br />

Perrière, 92320 Chatillon-sous-Bagneux, France, 830 pp.<br />

7. Joisel, A. (1973) Admixture for Cement - Physico Chemistry of Concrete and its<br />

Reinforcement, Published by the author, 3 avenue André, 95230 Soisy, France, 253 pp.<br />

8. Malhotra, V.M. 1 st , 2 nd , 3 rd and 4 th International Conferences on Superplasticizers<br />

organized in 1978 (ACI SP-62), 1981 (ACI SP-68), 1984 (ACI SP- 1 19) and 1994 held<br />

in october 1994.<br />

9. Admixture for Concrete — Improvement of properties (1990) Edited by Vasquez,<br />

Published by Chapman and Hall, London, U.K., 486 pp.<br />

10. Design and Control of Concrete Mixtures (1990) Published by Portland Cement<br />

Association, 5420 Old Orchard Road, Skokie, Ill, U.S.A., pp. 63-76.<br />

11. ASTM C494 (1990) Standard Specification for Chemical Admixtures for Concrete, pp.<br />

250-257.<br />

12. Aïtcin, P.-C. (1994) Durable Concrete — Current Practice and Future Trends. Edited by<br />

P.K. Mehta, ACI SP-144, pp. 85-104.<br />

13. Neville, A. (1981) Properties of Concrete, Pitman, London, U.K., 779 pp.<br />

14. Aïtcin, P.-C. and Neville, A. (1993) High-performance Concrete Demystified. Concrete<br />

International, ACI Journal, Vol. 15, No. 1, January, pp. 21-26.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 111 of 235


Cement<br />

Concrete<br />

15. Gagné, R. and Aïtcin, P.-C. (1993) Superplasticizers for Durable Concrete. Concrete<br />

Durability, Edited by R.R. Villarreal, Universidad Autonoma de Nuevo Leon, Monterrey,<br />

N.L. Mexico, pp. 201-217.<br />

16. Kreijger, P.-C. (1980) Plasticizers and Dispersing Admixtures. Admixtures Concrete<br />

International, the Construction Press, London, U.K., pp. 1-16.<br />

17. Aïtcin, P.-C., Jolicoeur, C., and MacGregor, J.G. (1994) Superplasticizers: How They<br />

Work and Why they Occasionally Don't, ACI Concrete International, Vol. 16, No. 5, May,<br />

pp. 45-52.<br />

18. Jolicoeur, C., Nkinamubanzi, P.-C., Simard, M.-A., and Piotte, M. (1994) Progress in<br />

Understanding the Functional Properties of Superplasticizer in Fresh Concrete, 4 th<br />

CANMET/ACI International Conference on Superplasticizers and Other Chemical<br />

Admixtures in Concrete, Montreal, October.<br />

19. Piotte, M. (1993) Caractérisation du poly (naphtalènesulfonate) et interactions de ce<br />

superplastifiant dans une suspension de ciment. PhD Thesis, Chemical Department,<br />

Université de Sherbrooke, 205 pp.<br />

20. Paillère, A.-M., Alègre, R., Ranc, R., and Buil, M. (1984) Interaction entre les réducteurs<br />

d'eau-plastifiants et les ciments, Lafarge / Laboratoire central des ponts et chaussées,<br />

Bd. Lefebvre, Paris, France, pp. 105-108.<br />

21. Dodson, V.H., and Hayden, T.D. (1989) Another Look at the Portland Cement Chemical<br />

Admixture Incompatibility Problem. Cement, Concrete and Aggregate, Vol. 11, No. 1,<br />

pp. 52-56.<br />

22. Manabe, T. and Kawada, N. (1960) Abnormal Setting of Cement Paste Owing to<br />

Calcium Lignosulfonate. Semento Konkunto, No. 162, pp. 24-27.<br />

23. Ranc, R. (1990) Interactions entre les réducteurs d'eau-plastifiants et les ciments.<br />

Ciments, bétons, plâtres, chaux, No. 782, pp. 19-20.<br />

24. Tagnit-Hamou, A., Baalbaki, M., and Aïtcin, P.-C. (1992) Calcium-sulfate Optimization in<br />

Low Water/Cement Ratio Concretes for Rheological Purposes, 9 th International<br />

Congress on the Chemistry of Cement, New Delhi, India, Vol. 5, pp. 21-25.<br />

25. Tagnit-Hamou, A. and Aïtcin, P.-C. (1993) Cement and Superplasticizer Compatibility.<br />

World Cement, Vol. 24, No. 8, August, pp. 38-42.<br />

26. Lessard, M., Gendreau, M., Baalbaki, M., and Pigeon, M. (1993) Formulation d'un béton<br />

à hautes performances à air entrainé. Bulletin de liaison des laboratoires des ponts et<br />

chaussées, Paris, No. 188, Novernber-December, pp. 41-51.<br />

27. Aïtcin, P.-C. and Lessard, M. Canadian Experience With Air-entrained, Highperformance<br />

Concrete. ACI Concrete International, to be published.<br />

28. Lessard, M., Dallaire, E., Blouin, D., and Aïtcin, P.-C. High-performance Concrete for<br />

MacDonald's. ACI Concrete International, to be published.<br />

29. ACI Standard 211. Standard Practice for Selecting Proportions for Normal, Heavyheight<br />

and Mass Concrete. ACI Manual of Concrete, Part 1, pp.211.1-1 211.1-38.<br />

30. Hammer, T.A. and Sellevold, E.J. (1990) Frost Resistance of High-strength Concrete,<br />

2 nd International Symposium on Utilization of High-strength Concrete, University of<br />

California, Berkeley, ACI SP-121, 1990, pp. 457-487.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 112 of 235


Cement<br />

Concrete<br />

31. Gagné, R., Pigeon, M., and Aïtcin, P.-C. (1991) The Frost Durability of High-performance<br />

Concrete, 2 nd Canada/Japan Meeting on the Effect of Low Temperature on Concrete,<br />

Edited by J.G. Litvan/K. Sakai, Japan, pp. 57-87.<br />

32. Phileo, R.E. (1986) Freezing and Thawing Resistance of High-strength Concrete.<br />

National Cooperative Highway Research Program Synthesis of Highway Practice 129.<br />

T.R.B., Washington DC, December, 31 pp.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 113 of 235


Cement<br />

Concrete<br />

Concrete Production Practices<br />

1. GENERAL....................................................................................................................... 115<br />

2. BATCHING OF MATERIALS.......................................................................................... 115<br />

3. MIXING............................................................................................................................ 116<br />

3.1 Types of mixers...................................................................................................... 116<br />

3.2 Uniformity of mixing................................................................................................ 117<br />

3.3 Mixing time ............................................................................................................. 117<br />

4. HANDLING, TRANSPORTING, PLACING..................................................................... 117<br />

5. COMPACTING................................................................................................................ 118<br />

5.1 Internal vibrator ...................................................................................................... 119<br />

5.2 External vibrator ..................................................................................................... 119<br />

5.3 Vibrating table ........................................................................................................ 120<br />

5.4 Surface vibrator...................................................................................................... 120<br />

6. CURING OF CONCRETE ............................................................................................... 120<br />

6.1 Length of curing period .......................................................................................... 120<br />

6.2 Curing methods...................................................................................................... 121<br />

6.3 Curing Compounds ................................................................................................ 122<br />

6.4 Steam curing .......................................................................................................... 123<br />

7. HOT WEATHER CONCRETING .................................................................................... 123<br />

8. COLD WEATHER CONCRETING.................................................................................. 126<br />

9. READY-MIXED CONCRETE .......................................................................................... 129<br />

10. PUMPED CONCRETE.................................................................................................... 130<br />

11. SPECIAL CONCRETING PROCESSES ........................................................................ 133<br />

12. LITERATURE.................................................................................................................. 134<br />

12.1..Concrete Technology ............................................................................................. 134<br />

12.2..Concrete practice ................................................................................................... 134<br />

12.3..Testing of Concrete................................................................................................ 134<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 114 of 235


Cement<br />

Concrete<br />

1 General<br />

Characteristics and properties of fresh and hardened concrete depend to a great extent upon<br />

mix design and quality of the constituent materials.<br />

However, the importance of mixing, transporting, placing and curing techniques should not<br />

be neglected.<br />

Improved production practices and techniques will then contribute considerably to achieving<br />

a good concrete. Each stage of concrete production is important and has an influence on the<br />

final concrete serviceability.<br />

2 Batching of Materials<br />

To produce concrete of uniform quality, the ingredients must be measured accurately for<br />

each batch. Most new specifications require that batching be done by weight rather than by<br />

volume because of the inaccuracies in measuring solid materials (especially damp sand) by<br />

volume. The weight system for batching provides greater accuracy, simplicity and flexibility.<br />

Flexibility is necessary because changes in the aggregate moisture content require frequent<br />

adjustments in batch quantities of water and aggregates. Water can be measured accurately<br />

by either volume or weight.<br />

ACI specifications generally require that materials be measured within this percentage of<br />

accuracy:<br />

cement ± 1%<br />

aggregates ± 2%<br />

water ± 1%<br />

admixtures ± 3%<br />

The ERMCO (European Ready Mixed Concrete Organisation) Code of good practice for<br />

Ready-Mixed Concrete gives recommendations for batching tolerance as listed in Table 1.<br />

Table 1: Accuracy of batching acc. ERMCO<br />

Material<br />

Tolerance<br />

Cement ± 3%<br />

Coarse aggregate ± 3%<br />

Fine aggregate ± 3%<br />

Admixtures ± 5%<br />

Water ± 3%<br />

Equipment should be capable of measuring quantities within these tolerances for the<br />

smallest batch regularly used, as well as for larger batches.<br />

The accuracy of batching equipment should be checked periodically and adjusted when<br />

necessary. Admixture dispensers should be checked daily since errors in admixture<br />

batching, particularly overdosing, can lead to serious problems in both fresh and hardened<br />

concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 115 of 235


Cement<br />

Concrete<br />

3 Mixing<br />

The objective of mixing is:<br />

♦ to coat the surface of all aggregate particles with cement paste<br />

♦ to blend all the ingredients of concrete into a uniform mass<br />

♦ to maintain uniformity of concrete at the discharging from the mixer<br />

All concrete should be mixed thoroughly until it is uniform in appearance, with all ingredients<br />

evenly distributed. Mixers should not be loaded above their rated capacities and should be<br />

operated at approximately the speeds for which they were designed. If the blades of the<br />

mixer become worn or coated with hardened concrete, the mixing action will be less efficient.<br />

Badly worn blades should be replaced and hardened concrete should be removed<br />

periodically, preferably after each day’s run of concrete.<br />

3.1 Types of mixers<br />

The method of discharging is one of the criteria for the classification of concrete mixers.<br />

Tilting Mixer: Mixing chamber (drum) it tilted for discharging (see Figure 1).<br />

Figure 1:<br />

Tilting mixer<br />

Drum (non-tilting) mixer: The axis of the mixer is always in a horizontal position, and<br />

discharge is effected by inserting a chute into the drum or by reversing the direction of<br />

rotation of the drum, as applied in the truck mixer used in the ready-mixed concrete<br />

production (see Figure 2).<br />

Figure 2:<br />

Drum (non-tilting) mixer<br />

Pan mixer: consists of a circular pan rotating about its axis, with one or two stars of paddles<br />

rotating about a vertical axis not coincident with the axis of the pan (see Figure 3).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 116 of 235


Cement<br />

Concrete<br />

Figure 3:<br />

Pan mixer<br />

The capacity of tilting and drum mixers is usually large (up to 6 m 3 ), whereas the pan mixer is<br />

smaller (up to 2 m 3 ) and is particularly efficient for stiff and cohesive mixes often used in<br />

precast concrete industry.<br />

Bowl-and-stirrer Mixer (cake mixer): works according to the same principle as the pan mixer<br />

and is sometimes used for mixing of mortar.<br />

3.2 Uniformity of mixing<br />

The efficiency of the mixer can be determined by the variability of the mix discharged into a<br />

number of samples without interrupting the flow of concrete.<br />

The values given in Table 2 are the highest acceptable values for a ‘satisfactory’ mixer.<br />

Table 2:<br />

Variability of concrete in a ‘satisfactory’ mixer<br />

Compressive strength 4 - 6%<br />

Percentage of coarse aggregate 6 - 8%<br />

Percentage of fine aggregate or 5 - 8%<br />

cement<br />

3.3 Mixing time<br />

The mixing time varies with the type of mixer and, strictly speaking, it is not the mixing time<br />

but the number of revolutions of the mixer that is the criterion of adequate mixing. Generally<br />

about 50 revolutions are sufficient.<br />

4 Handling, transporting, placing<br />

Each step in handling, transporting and placing concrete should be carefully controlled to<br />

maintain uniformity within the batch and from batch to batch so that the completed work is<br />

consistent throughout. It is essential to avoid separation of the coarse aggregate from the<br />

mortar or of water from the other ingredients.<br />

Concrete is handled and transported by many methods. These include the use of chutes,<br />

buggies operated over runways, buckets handled by cranes or cabinways, small railroad<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 117 of 235


Cement<br />

Concrete<br />

cars, trucks, pumping through pipelines, and pneumatically forcing the concrete or dry<br />

concrete materials through hoses.<br />

Preparation prior to placing includes compacting, trimming and moistening the subgrade;<br />

erecting the forms; and setting the reinforcing steel. A moist subgrade is especially important<br />

in hot weather to prevent extraction of water from the concrete.<br />

Forms should be clean, tight, adequately braced, and constructed of materials that will impart<br />

the desired texture to the finished concrete. Sawdust, nails and other debris should be<br />

removed before concrete is placed. Wood forms should be moistened before placing<br />

concrete; otherwise they will absorb water from the concrete and swell. Forms also should be<br />

treated with a releasing agent such as oil or lacquer to facilitate their removal. For<br />

architectural concrete, lacquer or emulsified stearates are used since they are non-staining.<br />

Reinforcing steel should be clean and free of loose rust or mill scale at the time concrete is<br />

placed. Any coatings of hardened mortar should be removed from the steel.<br />

5 Compacting<br />

The process of compacting concrete consists essentially of the elimination of air voids.<br />

The oldest means of achieving this is by ramming or punning the surface of concrete. The<br />

most common method for compacting is by vibration. The use of vibration makes it possible<br />

to work with drier mixes, so that concrete with a lower water/cement ratio and a lower cement<br />

content can be manufactured. Whether or not vibrators can be employed is determined by<br />

the consistency of the mix. Mixes which are very wet should not be vibrated as separation<br />

may result. Each type of vibrator requires a different consistency.<br />

Various types of vibrators have been developed (see also Figure 4):<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 118 of 235


Cement<br />

Concrete<br />

Figure 4:<br />

Various types of vibrators<br />

5.1 Internal vibrator<br />

Of the several types of vibrators this is perhaps the most common one. It consists essentially<br />

of a poker, housing an eccentric shaft driven through a flexible drive from a motor. The poker<br />

is immersed in concrete and thus applies approximately harmonic forces to it; hence, the<br />

alternative names of poker- or immersion vibrator.<br />

The frequency of vibration varies up to 12’000 cycles of vibration per minute.<br />

The poker is easily moved from place to place, and is applied at 50 to 70 cm centers for 5 to<br />

30 sec., depending on the consistency of the mix.<br />

5.2 External vibrator<br />

This type of vibrator is rigidly clamped to the formwork resting on an elastic support, so that<br />

both the form and the concrete are vibrated. As a result, a considerable proportion of the<br />

work done is used in vibrating the formwork, which also has to be strong and tight so as to<br />

prevent distortion and leakage of grout.<br />

The principle of an external vibrator is the same as that of an internal one, but the frequency<br />

is usually between 3000 and 6000 cycles.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 119 of 235


Cement<br />

Concrete<br />

5.3 Vibrating table<br />

This can be considered as a case of formwork clamped to the vibrator instead of the other<br />

way round, but the principle of vibrating the concrete and formwork together is unaltered.<br />

The source of vibration, too, is similar. Generally a rapidly rotating eccentric weight makes<br />

the table vibrate with a circular motion. With two shafts rotating in opposite directions the<br />

horizontal component of vibration can be neutralized so that the table is subjected to a<br />

simple harmonic motion in the vertical direction only.<br />

5.4 Surface vibrator<br />

A surface vibrator applies vibration through a flat plate direct to the top surface of the<br />

concrete. In this manner the concrete is restrained in all directions so that the tendency to<br />

segregate is limited; for this reason a more intense vibration can be used.<br />

6 Curing of concrete<br />

Properties of concrete such as resistance to freezing and thawing, strength, watertightness,<br />

wear resistance, and volume stability improve with age as long as conditions are favorable<br />

for continued hydration of the cement. The improvement is rapid at early ages but continues<br />

more slowly for an indefinite period. Two conditions for such improvement in quality are<br />

required:<br />

♦ the presence of moisture<br />

♦ a favorable temperature<br />

Excessive evaporation of water from newly placed concrete can significantly retard the<br />

cement hydration process at an early age. Loss of water also causes concrete to shrink, thus<br />

creating tensile stresses at the drying surface. It these stresses develop before the concrete<br />

has attained adequate strength, surface cracking may result. All exposed surfaces, including<br />

exposed edges and joints, must be protected against moisture evaporation.<br />

Hydration proceeds at a much slower rate when the concrete temperature is low; from a<br />

practical standpoint there is little chemical action between cement and water when the<br />

concrete temperature is near or below freezing. It follows that concrete should be protected<br />

so that moisture is not lost during the early hardening period and the concrete temperature is<br />

kept favorable for hydration.<br />

6.1 Length of curing period<br />

Since all the desirable properties of concrete are improved by curing, the curing period<br />

should be as long as practicable in all cases.<br />

The length of time that concrete should be protected against loss of moisture is dependent<br />

upon the type of cement, mix proportions, required strength, size and shape of the concrete<br />

mass, weather and future exposure conditions. This period may be a month or longer for<br />

lean concrete mixtures used in structures such as dams; conversely, it may be only a few<br />

days for richer mixes, especially if Type III of high-early-strength cement is used. Steamcuring<br />

periods are normally much shorter.<br />

For must structural uses, the curing period for cast-in-place concrete is usually 3 days to 2<br />

weeks, depending on such conditions as temperature, cement type, mix proportions, etc.<br />

More extended curing periods are desirable for bridge decks and other slabs exposed to<br />

weather and chemical attack.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 120 of 235


Cement<br />

Concrete<br />

6.2 Curing methods<br />

There are different methods to cure concrete:<br />

1) Methods that supply additional moisture to the surface of the concrete during the early<br />

hardening period. These include sprinkling and using wet coverings. Such methods afford<br />

some cooling through evaporation, which is beneficial in hot weather.<br />

2) Methods that prevent loss of moisture from the concrete by sealing the surface. This may<br />

be done by means of waterproof paper, plastic sheets, liquid membrane-forming<br />

compounds, and forms left in place.<br />

3) Methods that accelerate strength gain by supplying heat and moisture to the concrete.<br />

This is usually accomplished with live steam or heating coils.<br />

Sprinkling (see Figure 5)<br />

Continuous sprinkling with water is an excellent method of curing . If sprinkling is done at<br />

intervals, care must be taken to prevent the concrete from drying between application of<br />

water. A disadvantage of sprinkling may be its cost. The method requires an adequate<br />

supply of water and careful supervision.<br />

Figure 5:<br />

Sprinkling<br />

Wet coverings<br />

Burlap, cotton mats, or other moisture-retaining fabrics are used for curing. Treated burlaps<br />

that reflect light and are resistant to rot and fire are available. Coverings should be placed as<br />

soon as the concrete has hardened sufficiently to prevent surface damage. Care should be<br />

taken to cover the entire surface, including the edges of slabs such as pavements and<br />

sidewalks. The coverings should be kept continuously moist so that a film of water remains<br />

on the concrete surface throughout the curing period.<br />

Waterproof paper (see Figure 6)<br />

Waterproof curing paper is an efficient means of curing horizontal surfaces and structural<br />

concrete of relatively simple shapes. One important advantage of this method is that periodic<br />

additions of water are not required.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 121 of 235


Cement<br />

Concrete<br />

Figure 6:<br />

Waterproof paper<br />

Plastic sheets (see Figure 7)<br />

Certain plastic sheet materials (polyethylene films) are used to cure concrete. They are<br />

lightweight, effective moisture barriers and easily applied to simple as well as complex<br />

shapes.<br />

In some cases, the use of thin plastic sheets for curing may discolor the hardened concrete.<br />

This may be especially true when the concrete surface has been steel-troweled to a hard<br />

finish. When such discoloration is objectionable, some other curing method is advisable.<br />

Figure 7:<br />

Plastic sheets<br />

6.3 Curing Compounds<br />

Liquid membrane-forming curing compounds retard or prevent evaporation of moisture from<br />

the concrete. They are suitable not only for curing fresh concrete, but also for further curing<br />

of concrete after removal of forms or after initial moist curing.<br />

The concrete surface should be moist when the coating is applied. Normally only one<br />

smooth, even coat is applied, but two coat may be necessary to ensure complete coverage.<br />

A second coat, when used, should be applied at right angles to the first.<br />

Curing compounds used in hot weather should be white colored.<br />

Curing compounds are applied by hand-operated or power-driven spray equipment (see<br />

Figure8) immediately after the disappearance of the water sheen and the final finishing of the<br />

concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 122 of 235


Cement<br />

Concrete<br />

Figure 8:<br />

Curing compounds<br />

6.4 Steam curing<br />

Of the methods in Group 3, accelerating the strength development, the most common is<br />

steam curing. It can be used to advantage where early strength gain in concrete is important<br />

or where additional heat is required to accomplish hydration, as in cold-weather concreting.<br />

Two methods of steam curing for early strength gain are used today:<br />

♦ curing in live steam and atmospheric pressure (for enclosed cast-in-place structures and<br />

manufactured precast concrete units)<br />

♦ curing in high-pressure steam autoclaves (for small manufactured units).<br />

A steam-curing cycle consists of:<br />

1) an initial delay prior to steaming<br />

2) a period for increasing temperature<br />

3) a period for holding the maximum temperature constant<br />

4) a period for decreasing temperature<br />

A typical atmospheric steam-curing cycle is shown in Fig. 11, chapter ‘Concrete Categories’,<br />

paragraph 3.6.<br />

Steam curing at atmospheric pressure is generally done in a steam chamber or other<br />

enclosure to minimize moisture and heat losses.<br />

High-pressure steam curing in autoclaves takes advantage of temperatures in the range of<br />

160 to 190°C and corresponding pressures. Hydration is greatly accelerated and the<br />

elevated temperatures and pressures may produce additional beneficial chemical reactions<br />

between the aggregates and/or cementitious materials that do not occur under normal steam<br />

curing.<br />

7 Hot weather concreting<br />

Hot weather for concreting is defined as any combination of high air temperature, low relative<br />

humidity and wind velocity, tending to impair the quality of fresh or hardened concrete or<br />

otherwise resulting in abnormal properties. Hot weather can adversely affect the properties<br />

and serviceability of concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 123 of 235


Cement<br />

Concrete<br />

Concrete mixed, placed and cured at elevated temperatures normally develops higher early<br />

strength than concrete produced at normal temperatures, but the 28 day or later strength is<br />

generally lower (Figure 9).<br />

Figure 9:<br />

Influence of curing temperatures on strength at day and 28 days<br />

Plastic shrinkage cracking is usually associated with hot weather concreting. High concrete<br />

temperature, high air temperature, high wind velocity and low humidity, or combinations<br />

thereof, cause rapid evaporation which significantly increases the probability that plastic<br />

shrinkage cracking will occur.<br />

As shown in Figure 10, the amount of mixing water required to make a concrete of a certain<br />

consistency, increases considerably as the temperature of fresh concrete increases.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 124 of 235


Cement<br />

Concrete<br />

Figure 10:<br />

Water requirement of a concrete mix increases with an increase in<br />

temperature<br />

For the common, non-massive types of structure a maximum concrete temperature of 32°C<br />

is recommended by ACI as a reasonable upper limit. For the massive types of construction a<br />

temperature of 16°C or even lower would be desirable.<br />

The most practical method of maintaining low concrete temperatures is to control the<br />

temperature of the materials.<br />

The equation given below shows that, for concrete of conventional properties, a reduction of<br />

the concrete temperature by 1°C requires reducing either the cement temperature by about<br />

8°C or the water temperature by about 4°C or the aggregate temperature by about 2°C.<br />

Equation:<br />

T = temperature of freshly mixed concrete (deg. C)<br />

T a , T c , T w , T wa = temperature of aggregate, cement, mixing water<br />

and water on aggregate respectively (deg. C)<br />

W a , W c , W w , W wa = weight of aggregate, cement, added mixing water<br />

and water on aggregate, respectively (kg)<br />

Of the materials contained in concrete, water is the easiest to cool. It can be cooled by<br />

refrigeration or by adding ice which is used as part of the mixing water, provided that it is<br />

completely melted by the time mixing is completed.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 125 of 235


Cement<br />

Concrete<br />

Cement temperature has only a minor effect on the temperature of the freshly mixed<br />

concrete because of the low specific heat of cement and the relatively small amount of<br />

cement contained in the mix. Fresh hot cement from the plant can cause difficulties.<br />

Aggregates have a pronounced effect on the fresh concrete temperature because they<br />

represent 60 to 80% of the total weight of concrete. Aggregate stockpiles should be shaded<br />

from the sun and kept moist by sprinkling. Before concrete is placed, the forms, the<br />

reinforcing steel, and the subgrade should be cooled by sprinkling as well.<br />

Transporting and placing of concrete should be done as quickly as practicable. During<br />

extremely hot periods improved results may be obtained by restricting the placing of concrete<br />

to the early morning or evening hours.<br />

In order to prevent concrete from drying out, curing should be started as soon as the<br />

concrete surface is finished. In special cases of concreting during hot weather a retarding<br />

admixture may be used to delay the setting time.<br />

8 Cold Weather Concreting<br />

Concreting in cold weather requires special precautions. Concrete sets slowly in cold<br />

weather and development of strength is delayed. At low temperatures (at +5°C or lower<br />

during placing and the early curing period), one or more of these recognized protective<br />

measures should be used:<br />

♦ Heating the area where concrete is placed<br />

• The presence of ice and the possibility of ice formation during concreting must be<br />

avoided. Temperature of the concrete forms should be raised to above freezing, as<br />

well as that of adjacent concrete and subgrade, through the use of heated<br />

enclosures.<br />

♦ Heating the water and concrete materials<br />

• In cold weather, freshly placed concrete should be at least 10°C and not more than<br />

32°C when poured in the forms. In addition to heating water, it may be necessary to<br />

heat aggregates.<br />

♦ Use of chemical accelerators<br />

• During low temperatures, the use of chemical accelerators can speed up the set of<br />

concrete. Calcium chloride is the most commonly used and may be used up to 2% by<br />

weight of cement.<br />

♦ Calcium chloride or admixtures containing soluble chlorides must not be used:<br />

• in reinforced and prestressed concrete<br />

• in concrete containing embedded aluminium<br />

• in lightweight insulating concrete place over metal decks<br />

• in concrete that will be in contact with soils, or water containing sulfates<br />

♦ Maintaining concrete temperatures<br />

• Concrete slabs lose heat and moisture rapidly in cold weather. They need protection<br />

against wind and cold. A heated enclosure or insulation should be provided to keep<br />

concrete temperature above 10°C. The following Time-Temperature Chart shows<br />

minimum periods in which concrete temperatures should be maintained:<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 126 of 235


Cement<br />

Concrete<br />

20°C 10°C<br />

plain concrete 3 days 7 days<br />

plain concrete with<br />

calcium chloride<br />

2 day 3 days<br />

• Note: After the periods shown above, concrete temperature should be maintained<br />

above 5°C for at least four days. Concrete should not be allowed to dry out.<br />

♦ Special provisions for curing<br />

• Cold weather air acts like a sponge to draw moisture from concrete in both its fresh<br />

and hardened state. This rapid drying-out process must be avoided. Curing and<br />

protection from start to finish should be continuous and uninterrupted until the<br />

concrete attains its designed strength. At the end of the curing period (see Time-<br />

Temperature Chart above), protection should be removed in such a way that the<br />

temperature of the concrete will not drop faster than 5°C in 24 hours.<br />

The following figure 11 illustrates the precautions that should be taken for cold weather<br />

concreting.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 127 of 235


Cement<br />

Concrete<br />

Figure 11:<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 128 of 235


Cement<br />

Concrete<br />

9 Ready-Mixed Concrete<br />

If, instead of being batched and mixed on the job site, concrete is delivered ready for placing<br />

from a central plant, it is referred to as ready-mixed concrete.<br />

This type of concrete is used extensively in many countries as it offers numerous advantages<br />

as compared to the job-site-mixed concrete (Table 3).<br />

Table 3:<br />

Production of ready-mixed concrete in W. Europe (from ERMCO Annual<br />

Report 1976/77 and 1995/96)<br />

Country<br />

Cement consumption<br />

in ready-mixed<br />

concrete in % of total<br />

cement consumption<br />

Ready-mixed concrete<br />

production in mio m 3<br />

1976/77 1995/96 1976/77 1995/96<br />

Austria 19.5 44.0 4.3 8.6<br />

Belgium 30.9 39.0 6.34 9.1<br />

W. Germany 44.6 55.0 49.8 68.3<br />

Finland 57.7 40.0 3.4 1.6<br />

France 26.3 40.6 25.2 29.7<br />

Great Britain 42.1 50.0 24.8 22.3<br />

Italy 23.1 43.8 31.7 54.0<br />

Spain 19.0 36.6 15.6 36.2<br />

Holland 36.4 52.0 6.8 8.0<br />

Sweden 48.4 61.7 5.2 2.4<br />

Switzerland 40.7 57.0 5.6 9.0<br />

USA 60.0 71.5 130.0 192.0<br />

Ready-mix concrete plants use precision scales to weigh the ingredients as the producer is<br />

responsible for delivering concrete of the required quality.<br />

There are three types of mixing:<br />

♦ Transit mixed concrete is mixed completely in a truck mixer. The batching of ingredients<br />

is carried out in a dry-batch plant.<br />

♦ Central mixed concrete is batched and mixed completely in a stationary batching and<br />

mixing plant (premix-plant) and is delivered in a special dump truck (tipper) without<br />

mixing or agitating, or in a truck mixer at agitating speed.<br />

♦ Shrink-mixed concrete is mixed partially in a stationary mixer and then completed in a<br />

truck mixer.<br />

Because of the many advantages offered by ready-mixed concrete, this industry has shown<br />

a phenomenal growth in recent years. These advantages can be summarized as follows:<br />

♦ Use of modern precision batching equipment assures an accurately proportioned mix and<br />

allows the use and modification of different mix designs.<br />

♦ Thorough mixing of each batch helps to produce uniform concrete.<br />

♦ More efficient operation on the job site (no space problems with respect to storage of<br />

ingredients, etc.).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 129 of 235


Cement<br />

Concrete<br />

♦ Delivery at the time required.<br />

♦ Delivery in the exact quantity desired, thus eliminating waste.<br />

♦ Equipment on the building site not necessary.<br />

♦ Possibility to use different chemical admixtures.<br />

♦ Possibility to use special concretes.<br />

♦ Clear and easy base for delivery contracts; no problems with estimates of concrete cost.<br />

The basis for the purchase of ready-mixed concrete is the concrete volume (m 3 , cubic yard)<br />

and required concrete quality. To attain the desired concrete quality, different types of<br />

concrete mixes (mix design) can be used (see also paper on "Concrete Mix Design"):<br />

♦ Designed mix (performance mix): concrete is designed by the ready-mix producer and<br />

should fulfill all requirements as stated by the purchaser.<br />

♦ Codified mix: mix proportions are established in a National Code and are usually suitable<br />

for a restricted range of concrete strength classes and applications (for example in CEB-<br />

Code only for mixes with strength ≤ 25 N/mm 2 ).<br />

♦ Prescribed mix: the purchaser is responsible for designing the concrete mix and specifies<br />

the mix proportions and the materials to be used by the ready-mix producer.<br />

In recent years the ready-mixed concrete industry has gained considerable influence on the<br />

development of concrete standards and specifications. Several new developments in the<br />

concrete technology have been introduced by the ready-mixed concrete industry, e.g.<br />

♦ flowing concrete<br />

♦ ready-mixed mortar<br />

♦ developments of new accelerated testing methods, etc.<br />

10 Pumped Concrete<br />

Pumped concrete may be defined as concrete conveyed by pressure through either a rigid<br />

pipe or flexible hose and discharged directly into the desired area.<br />

Pressure is applied by (see Figures 12 to 14):<br />

♦ Piston pumps<br />

♦ Pneumatic pumps (compressed air)<br />

♦ Squeeze pressure pumps<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 130 of 235


Cement<br />

Concrete<br />

Figure 12: Schematic drawings of the various pump types -<br />

piston type concrete pump<br />

Inlet valve opens while outlet valve is closed and concrete is drawn into cylinder by gravity<br />

and piston suction. As piston moves forward inlet valve closes, outlet valve opens, and<br />

concrete is pushed into pump line.<br />

Figure 13:<br />

Pneumatic type concrete pump<br />

Compressor builds up air pressure in tank, which forces concrete in placer through the line.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 131 of 235


Cement<br />

Concrete<br />

Figure 14:<br />

Squeeze Pressure Type Concrete Pump<br />

Pumping may be used for most concrete constructions, but it is especially useful where<br />

space for construction equipment is limited.<br />

Very often the ready-mixed concrete companies include a pumping division. The portion of<br />

pumped concrete as compared to the total ready-mixed concrete production in various<br />

countries is shown in Table 4.<br />

Table 4:<br />

Pumped concrete in proportion to the total ready-mixed concrete<br />

production (1996)<br />

(Ready-mix concrete companies affiliated with the Holcim Group)<br />

Country No. of Plants Pumped concrete<br />

in % of the total<br />

production<br />

Belgium 55 30<br />

France 116 7<br />

Switzerland 18 15<br />

W. Germany 38 49<br />

Greece 3 74<br />

Colombia 29 37<br />

Costa Rica 13 33<br />

Mexico 62 32<br />

Brazil 40 30<br />

Ecuador 7 70<br />

South Africa 38 16<br />

The concrete pumps can be:<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 132 of 235


Cement<br />

Concrete<br />

♦ Truck-mounted or<br />

♦ Stationary<br />

Ease of operation, safety and economy is required of a good pump. Today the most widely<br />

used diameter of a pipe for pumping is 125 mm; the usual maximum grain size is 32 mm, the<br />

minimum cement content about 300 kg and the consistency should be plastic (slump 7 to 12<br />

cm).<br />

Pumping jobs over heights of more than 200 m and with volumes of up to 110 m 3 /hour are<br />

quite common in today’s concrete practice.<br />

11 Special Concreting Processes<br />

The various special concreting processes will not be discussed in detail. However, some of<br />

them are briefly explained in the following:<br />

♦ Vacuum-processed concrete: concrete from which, after compaction and before<br />

hardening, water is extracted by a vacuum process. The w/c ratio is thus reduced so that<br />

concrete with higher strength, higher density, lower permeability and better durability can<br />

be obtained.<br />

♦ Preplaced-aggregate concrete: concrete produced by placing coarse aggregate into a<br />

mould and later injecting a cement-sand grout, usually with admixtures to fill the voids.<br />

♦ Dry-packed concrete: concrete mixture sufficiently dry to be consolidated by heavy<br />

ramming.<br />

♦ Spun concrete: concrete compacted by centrifugal action, e.g. in the manufacture of<br />

pipes (centrifugal process, roller-suspension process).<br />

♦ Tremie concrete: concrete placed under water through a pipe or tube.<br />

♦ Shotcrete: mortar or concrete conveyed through a hose and projected at high velocity<br />

onto a surface; also known as air-blown mortar, pneumatically applied mortar or<br />

concrete, sprayed mortar and gunned concrete (gunite).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 133 of 235


Cement<br />

Concrete<br />

12 LITERATURE<br />

12.1 Concrete Technology<br />

1) A.M. Neville - Properties of Concrete, 4th Edition - Longman Group Ltd.<br />

2) Lea’s Chemistry of Cement and Concrete, 4th Edition - Arnold Ed.<br />

3) Ghosh - Cement and Concrete Science and Technology - abi Books Private Ltd.<br />

12.2 Concrete practice<br />

1) W.H. Taylor - Concrete Technology and Practice<br />

2) Ramachandran - Concrete Admixtures Handbook - Noyes Publications<br />

3) ACI Manual of Concrete Practice - aci Publications<br />

4) Properties of Fresh Concrete, Proceedings of the Rilem Colloquium - Chapman & Hall<br />

12.3 Testing of Concrete<br />

1) Bartos - Fresh Concrete, Properties and Tests - Elsevier Ed.<br />

2) Concrete Test Methods.<br />

RILEM, Secrétariat Général, 12 rue Brancion, 75737 Paris CEDEX 15<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 134 of 235


Cement<br />

Concrete<br />

Concrete Shrinkage<br />

1. DEFINITION AND DESCRIPTION OF SHRINKAGE ............................................136<br />

2. TYPES OF SHRINKAGE AND SIGNIFICANCE....................................................136<br />

2.1 General .........................................................................................................136<br />

2.2 Plastic shrinkage ...........................................................................................137<br />

2.3 Drying shrinkage ...........................................................................................139<br />

2.4 Autogenous shrinkage ..................................................................................141<br />

2.5 Carbonation shrinkage ..................................................................................143<br />

3. FACTORS INFLUENCING CONCRETE SHRINKAGE.........................................145<br />

3.6 Overview .......................................................................................................145<br />

3.7 Concrete composition ...................................................................................145<br />

3.8 Type and Grading of Aggregate....................................................................147<br />

3.9 Cement characteristics..................................................................................148<br />

3.10 Mineral components......................................................................................149<br />

3.11 Chemical Admixtures ....................................................................................150<br />

3.12 Curing and Storage Conditions .....................................................................150<br />

3.13 Dimension of Structural Member...................................................................151<br />

3.14 Reinforcement...............................................................................................152<br />

3.15 Combined Effect on Individual Factors .........................................................152<br />

4. CONTROL OF CRACKING DUE TO SHRINKAGE ..............................................153<br />

4.16 Introduction ...................................................................................................153<br />

4.17 Reduction of Cracking Tendency ..................................................................154<br />

4.18 Reinforcement...............................................................................................154<br />

4.19 Joints.............................................................................................................155<br />

4.20 Shrinkage-Compensating Concrete ..............................................................156<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 135 of 235


Cement<br />

Concrete<br />

1 Definition and description of shrinkage<br />

Shrinkage and the contrary effect of swelling, is a seemingly simple phenomenon of<br />

contraction which is defined as the reduction in the outer volume without the influence of an<br />

external force. Strictly speaking, shrinkage is a three-dimensional deformation but it is<br />

usually expressed as a linear strain because in the majority of exposed concrete elements<br />

one or two dimensions are much smaller than the third dimension, and the effects of<br />

shrinkage are greatest in the largest dimension.<br />

In common usage, the term shrinkage is a shorthand expression for drying shrinkage<br />

of hardened concrete but there are several other types of shrinkage deformation in<br />

concrete which, depending on circumstances, may or may not occur simultaneously<br />

or be independent of one another [A6].<br />

Volume changes due to changes in temperature are not to be covered by the terms<br />

shrinkage and swelling, although stresses created through the release and flowing off of heat<br />

of hydration are often called thermal shrinkage or thermal contraction.<br />

For practical purposes, shrinkage is usually described by the amount of shrinkage in one<br />

dimension as expressed by the following formula:<br />

ε (t) = (l t - I 0 )/I 0 = ∆l/l 0<br />

being:<br />

I 0 : initial length<br />

l t :length at time t<br />

The dependence of time shows that shrinkage deformations vary with time. For constant<br />

ambient conditions, the amounts of shrinkage tend towards a corresponding final state<br />

(designated as final shrinkage value ε s,00 ) the amount of shrinkage depends mainly on the<br />

relative humidity and temperature of the ambient air, the dimensions of the structural element<br />

and the concrete’s composition. In case of water storage, the shrinkage will show positive<br />

values (swelling).<br />

2 Types of shrinkage and significance<br />

2.1 General<br />

All shrinkage phenomena occurring in concrete systems are basically caused by changes in<br />

the water balance within the porous matrix of the concrete. According to the time of<br />

occurrence and the origin of the change in this water balance, four different types of<br />

shrinkage for concrete can be distinguished.<br />

• Shrinkage due to capillary forces in the fresh and still plastic concrete. This kind<br />

of shrinkage is called “plastic shrinkage” or early shrinkage.<br />

• Shrinkage due to drying out of the hardened concrete (“drying shrinkage”).<br />

• Shrinkage in connection with the chemical volume reduction during cement hydration<br />

called “autogenous shrinkage”.<br />

• Shrinkage due to carbonation of the hardened concrete called “carbonation shrinkage”<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 136 of 235


Cement<br />

Concrete<br />

2.2 Plastic shrinkage<br />

Plastic shrinkage cracking is a constant source of concern in the concrete industry. It is a<br />

frequent cause of anxiety, yes conflict, between the concrete supplier and the client when<br />

cracks are observed on a freshly placed concrete surface. It also causes concern to the<br />

designers as “long-term durability” comes into question.<br />

After placement and compaction of the fresh concrete, the concrete surface is usually<br />

covered by a film of water. The thickness depends mainly on the bleeding tendency of the<br />

concrete. At this stage the concrete is still in a plastic state and hence reference is made to<br />

plastic shrinkage.<br />

The most important factor influencing plastic shrinkage and plastic shrinkage cracking is the<br />

rate of evaporation of water from the concrete surface. A similar loss can occur by suction by<br />

the underlying dry concrete or soil (placing concrete on a dry subgrade should be avoided).<br />

If the rate of evaporation exceeds the rate at which bleeding water rises to the surface, then<br />

plastic shrinkage and cracking are likely to occur .<br />

The state is easily visible by the disappearance of the sheen from the surface of concrete.<br />

But it has been observed that cracks also form under a layer of water and merely become<br />

apparent on drying.<br />

Another type of cracking on the surface of fresh concrete is caused by differential settlement<br />

of fresh concrete due to some obstruction to settlement, such as large particles of aggregate<br />

or reinforcing bars. This plastic settlement cracking can be avoided by the use of a dry mix,<br />

good compaction and by not allowing too fast a rate of build-up of concrete. Plastic<br />

shrinkage cracking and plastic settlement cracking are sometimes confused with one<br />

another.<br />

The temperature, relative humidity and wind velocity of the air and the temperature of the<br />

concrete influence the rate of evaporation. It should be remembered that evaporation is<br />

increased when the temperature of the concrete is much higher than the ambient<br />

temperature; under such circumstances, plastic shrinkage can occur even if the relative<br />

humidity of the air is high.<br />

The ACI 305R-91 nomograph (Figure 2.1) is still the preferred method for predicting the<br />

evaporation rate of bleed water from the surface of freshly placed concrete. Precautionary<br />

should be taken when the rate evaporation is expected to approach 0.2 [psf/h] (1.0 kg/m²/hr).<br />

The Canadian code [C1] nominates 0.75 kg/m 2 /hr as the critical value while Australian<br />

references quote 0.5 kg/m 2 /hr as a value at which precautions should be taken. Such<br />

conditions mainly occur during hot weather concreting, but any atmospheric condition<br />

resulting in an excessive evaporation rate can lead to plastic shrinkage cracks. According to<br />

ACI 305R-91, the risk of plastic cracking is the same at the following combinations of<br />

temperature and relative humidity:<br />

• 41°C and 90%<br />

• 35°C and 70%<br />

• 24°C and 30%<br />

Recently Paul J. Uno has developed a “single operation” equation which can be used on a<br />

simple hand-held calculator and is very appropriate for on-site quick checks to see if<br />

evaporation is going to be a critical factor in plastic shrinkage cracking.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 137 of 235


Cement<br />

Concrete<br />

E = 5[(T c +18) 2.5 – r (T a +18) 2.5 ][V+4]x10 -6<br />

Where<br />

E = evaporation rate, kg/m 2 /hr<br />

T c = concrete temperature, °C<br />

T a = air temperature, °C<br />

r = (RH percent)/100<br />

V = wind velocity, kph<br />

Under favourable climatic conditions, the amount of plastic shrinkage for concrete can reach<br />

up to 400 µm/m. According to experience, the width of the individual crack amounts to about<br />

0.1 to 3 mm and the cracks can extend in thick concrete members down to a depth of 10 cm.<br />

A typical feature of the plastic shrinkage cracks is that they are relatively wide at the surface<br />

and then rapidly decrease in width towards the interior of the concrete. The pattern of cracks<br />

is generally irregular (“map” pattern). For concrete slabs, also patterns with a series of<br />

parallel lines at very approximately 45° to the edge of the slab are observed.<br />

The most notable characteristic of plastic shrinkage cracks in slabs is that they normally do<br />

not extend to the edge of the slab, because the edge can shrink without restraint.<br />

Plastic shrinkage and the above described crack formation is most likely to take place are the<br />

hours after compaction and finishing of the concrete up to the beginning of the hardening,<br />

which is typically the period of 2 to 8 hours after concrete placement. According to a study<br />

carried out at HGRS on restrained concrete slabs, a very critical period for cracking seems to<br />

be 2 to 4 hours after mixing and placing. Retarders, which maintain the concrete workable for<br />

a longer time, can intensify or occasionally even initiate the plastic shrinkage cracking.<br />

To prevent the cracks, the site personnel must be aware of the factors that will cause<br />

cracking and be prepared for adverse conditions. In some case, not pouring at all could be<br />

the best choice.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 138 of 235


Cement<br />

Concrete<br />

Figure 2.1 – Effect of air temperature, relative humidity, wind velocity and concrete<br />

temperature on rate of evaporation of surface moisture from concrete.<br />

To effectively avoid these cracks from forming, the golden rule is not to let the concrete dry<br />

out too rapidly. This means that it is crucial to finish the concrete as quickly as possible and<br />

to start curing immediately. In severe conditions, such as very hot, dry and windy, it may be<br />

necessary to apply an evaporation retarder to the fresh concrete. These compounds are<br />

alcohol-based and are not substitutes for normal curing compounds. They are specifically<br />

designed to be applied to fresh concrete and are very effective in preventing excessive water<br />

loss during the initial stage of strength gain.<br />

Concrete composition is of importance insofar as water rich fresh concrete with high water<br />

retention ability is more susceptible to shrinkage and cracking than other concrete, so that<br />

certain improvement is also possible by adaptations in concrete characteristics.<br />

2.3 Drying shrinkage<br />

As a result of a disequilibrium in the relative humidity between concrete and its environment,<br />

any concrete loses water when exposed to an environment whose relative humidity is lower<br />

than that existing in the capillary network in the concrete. As a result of this water loss,<br />

concrete shrinks. The magnitude of drying shrinkage depends on many factors, including the<br />

properties of the materials, temperature, relative humidity of the environment, the age at<br />

which concrete is first exposed to a drying environment and the size of the concrete element<br />

that is drying.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 139 of 235


Cement<br />

Concrete<br />

The driving force for drying shrinkage is the evaporation of water from the capillary network<br />

in the concrete at the menisci which are exposed to air with a relative humidity lower than<br />

that within the capillary pores. As the free water contained in the capillary pores is held by<br />

forces that are inversely proportional to the diameter of the pores, the loss of water is<br />

progressive and proceeds at a decreasing rate. Researchers still disagree about the<br />

fundamental mechanism responsible for drying shrinkage. These mechanisms include the<br />

capillary tension, the surface adsorption, and the interlayer water removal.<br />

Drying shrinkage occurs first at the surface of concrete, and indeed only at a surface<br />

exposed to dry air; this could be a single surface or all external surfaces of a concrete<br />

element. Drying shrinkage then progresses toward the core of the element so that the loss<br />

of water expressed as a percentage of the apparent volume of concrete is smaller for<br />

concrete elements with a lower surface/volume ratio. Factors influencing the magnitude of<br />

the loss of water are the porosity of the concrete and the characteristics of the capillary pore<br />

network such as the size and shape of the pores and their continuity.<br />

Typical basic shrinkage values of structural members according to the location and climatic<br />

condition respectively are shown in Table 2.1. As hardened concrete has a very low tensile<br />

elongation at rupture of 0.1 to 0.2 x 10 -3 [G1], structural members are susceptible to cracking<br />

if the deformation is hindered by external restraint. If the tensile stresses, caused by the<br />

external restraint, exceed the tensile strength of the concrete, continuous cracks are formed<br />

(Figure 2.2). These cracks can affect the function of the corresponding structural members.<br />

Table 2.1 – Basic shrinkage value in dependence of location of structural member<br />

Location of structural member Average rel.<br />

humidity about (%)<br />

Basic shrinkage<br />

value (µm/m)<br />

IN WATER 100 +100<br />

In very humid air, e.g. directly above<br />

water<br />

90 -130<br />

Generally in the open air 70 -320<br />

In dry air, e.g. in dry interior rooms 50 -460<br />

A further consequence of the drying shrinkage is loss of the pretension in pre-stressed<br />

structural members. Also warping of concrete slabs is directly related to the amount of drying<br />

shrinkage. Finally, significant time-dependent bending of structural members is possible due<br />

to drying shrinkage, similar to effects observed under load by means of creep.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 140 of 235


Cement<br />

Concrete<br />

Figure 2.2 – Cracking of concrete due to drying shrinkage under restraint<br />

To summarise, it can be stated that drying shrinkage of concrete can not be avoided, but the<br />

amount of shrinkage can be controlled to a certain extent by adjusting some influencing<br />

factors in a favourable manner. The shrinkage deformation depends mainly on the time, the<br />

dimensions of the structural member, the storage conditions and the concrete composition.<br />

2.4 Autogenous shrinkage<br />

As far as concrete with low water/binder ratios is concerned, high levels of autogenous<br />

shrinkage are generally associated with self-desiccation. For ordinary concrete, autogenous<br />

shrinkage is negligible (typical value are 40 µm/m at the age of one month and 100 µm/m<br />

after five years). This difference in the behaviour of these two types of concrete can be<br />

explained in terms of differences in the size distribution of the pore and capillary networks.<br />

According to Aïtcin the expression autogenous shrinkage was first mentioned in the literature<br />

in 1934 and positively defined by Davis as a macroscopic volume change caused by the<br />

hydration of cement. The expression chemical shrinkage is also found in the literature<br />

because autogenous shrinkage is a consequence of the decrease in the absolute volume of<br />

the hydrated cement paste during its hydration.<br />

It may be worth noting that, because the products of hydration can form solely in water-filled<br />

space, only a part of the water in the capillary system can be used in hydration. Thus, for<br />

hydration to take place there must be enough water present both for the chemical reactions<br />

and for filling of gel pores. These gel pores are formed by the reactions of hydration of<br />

cement. They are much finer than capillary pores so that they drain water from the coarsest<br />

capillary pores. As a consequence in the absence of any external water supply, the coarsest<br />

capillary pores start to dry in a manner similar to drying by evaporation. This is why this<br />

phenomenon is called self-desiccation.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 141 of 235


Cement<br />

Concrete<br />

The causes of autogenous shrinkage are the same as the causes of drying shrinkage<br />

because it is the same physical phenomenon that is developing within concrete: the creation<br />

of menisci within the capillary system and the resulting tensile stresses. Nevertheless, there<br />

are some major differences between autogenous shrinkage and drying shrinkage:<br />

• autogenous shrinkage develops without any mass loss, unlike drying shrinkage;<br />

• autogenous shrinkage develops isotropically within the concrete while drying shrinkage<br />

progresses from the drying surface to the core of the element;<br />

• autogenous shrinkage does not develop any relative humidity gradients, drying shrinkage<br />

does;<br />

Autogenous shrinkage is linked directly to cement hydration and starts to develop uniformly<br />

and isotropically only few hours after the casting of concrete (practically always before 24<br />

hours), whereas drying shrinkage starts to develop a little slower and at the surface when<br />

hardened concrete is exposed to a dry environment which is usually a matter of days rather<br />

than hours (Figure 2.3).<br />

Less than one day<br />

Several days<br />

Figure 2.3 – Typical plot of change in temperature versus concrete shrinkage.<br />

Autogenous shrinkage can be avoided if the coarse capillaries are refilled by an external<br />

supply of water (water curing). As long as the gel pore system and capillary pore systems<br />

are connected to this external supply of water, no menisci are formed within the capillary<br />

system, so that there are no induced tensile stresses and no autogenous shrinkage.<br />

The ingress of external water favours further hydration of the still unhydrated parts of cement<br />

particles and at this stage, hydration results in an increase in the absolute volume because<br />

the added water from outside the concrete. As a consequence, the gel pore system and the<br />

capillary pore system may become disconnected, depending on the local pore size<br />

distribution in the hardening concrete. As soon as the water phase has ceased to be<br />

continuous, water ingress from outside stops and conditions for the development of<br />

autogenous shrinkage are created.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 142 of 235


Cement<br />

Concrete<br />

Hydration<br />

Volumetric contraction<br />

Self desiccation<br />

Menisci<br />

Autogenous<br />

shrinkage<br />

No<br />

External<br />

supply of<br />

water<br />

No autogenous<br />

shrinkage<br />

No<br />

Yes<br />

Pores<br />

and capillaries<br />

connected?<br />

Yes<br />

No menisci<br />

Autogenous shrinkage tends to increase at higher temperatures, with a higher cement<br />

content, and possibly with finer cements and with cements which have a high C 3 A and C 4 AF<br />

content. At a constant content of blended cement, a higher content of fly ash leads to lower<br />

autogenous shrinkage .<br />

Research at CSIRO shows that during the first 24 hours after casting, some high strength<br />

concrete exhibits total shrinkage values of 600-800 µm/m under standard temperature<br />

conditions. Autogenous strains of this magnitude translate to a total deformation of between<br />

6 and 8 mm for every 10 m length of concrete element. This amount of autogenous<br />

shrinkage in the first 24 hours is equivalent to the total drying shrinkage measured over a<br />

period of 1 year for the same concrete. According to Mak these autogenous deformations<br />

may cause a real risk of early age cracking in in-situ, precast and composite members.<br />

2.5 Carbonation shrinkage<br />

The absorption of CO 2 from the air significantly changes the chemical and physical<br />

characteristics of the cement stone. Not only the free Ca(OH) 2 is transformed to CaCO 3<br />

under release of water, but also the hydration products of the cement are decomposed under<br />

the influence of CO 2 . This carbonation process is accompanied by an increase in weight of<br />

the concrete and by an increase in the irreversible shrinkage. The reversible shrinkage,<br />

however, is reduced considerably, which can be explained by the formation of relatively<br />

coarse crystals of calcium carbonate.<br />

The mechanisms of carbonation shrinkage are not very clear and can not be explained by<br />

the liberation of previously bound water alone. Carbonation shrinkage is possibly caused by<br />

the dissolution of crystals of Ca(OH) 2 while under compressive stress (imposed by the drying<br />

shrinkage) and depositing of CaCO 3 in spaces free from stress; the compressibility of the<br />

paste is thus temporarily increased. Carbonation of the hydrates present in the gel does not<br />

contribute to shrinkage, as the reaction does not involve dissolution and re-precipitation.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 143 of 235


Cement<br />

Concrete<br />

The effect of carbonation depends a lot on the relative humidity and drying state of the<br />

concrete. Carbonation increases shrinkage mostly at the intermediate humidities of about<br />

50%. Water saturated or very dry concrete absorbs practically no CO 2 and accordingly do not<br />

have higher shrinkage values. An additional factor playing an important role is the sequence<br />

of drying and carbonation. Simultaneous shrinkage and carbonation (which usually happens<br />

in practice) produces lower total shrinkage than when drying is followed by carbonation<br />

(Figure 2.4).<br />

For the amount of shrinkage of whole structural members, the carbonation shrinkage is not<br />

significant, since in general it only affects a very small part of the total cross-section.<br />

Conditions under which carbonation shrinkage can be noticed are encountered on vertical<br />

surfaces of exposed structural members, which are seldom exposed to rain-showers and<br />

which often get sunshine. In the edge zones of these members, the carbonation shrinkage<br />

can lead to internal (tensile) stresses in the surface area, creating a dense net of fine<br />

unevenly distributed cracks.<br />

Figure 2.4 - Influence of the sequence of drying and carbonation of shrinkage<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 144 of 235


Cement<br />

Concrete<br />

3 Factors Influencing Concrete Shrinkage<br />

3.1 Overview<br />

Among the four types of shrinkage described in the previous section, drying shrinkage is<br />

most common and important in the present concrete construction practice. As stated before,<br />

autogenous shrinkage is relatively important only for high performance concrete (HPC)<br />

having a low water/cement ratio (< 0.40) and in large mass concrete structure. HPC<br />

represents only a small volume of concrete that is used in the present market (but is<br />

expected to be more important in the future for reasons of durability).<br />

Plastic shrinkage can usually be controlled effectively by an adequate curing procedure, so<br />

that cracking of concrete in its plastic state can be avoided.<br />

The literature lists many factors that have an influence on shrinkage of concrete (mostly<br />

drying shrinkage). The following list gives an overview of the most commonly mentioned<br />

variables:<br />

• Concrete composition (aggregate, water and cement content, w/c-ratio)<br />

• Cement characteristics (composition, fineness)<br />

• Type and grading of aggregate<br />

• Mineral admixtures<br />

• Chemical admixtures<br />

• Curing and storage conditions<br />

• Dimensions of structural member (size and shape)<br />

• Reinforcement<br />

3.2 Concrete composition<br />

The most important influence on the shrinkage of concrete is exerted by the aggregate,<br />

which restrains the amount of shrinkage that can actually be realised. Cement paste in<br />

concrete, if unrestrained, would shrink from 5 to 15 times as much as concrete. The ratio of<br />

shrinkage of concrete (S c ) to shrinkage of neat paste (S p ) depends on the aggregate content<br />

in the concrete, a, and is<br />

S c = S p (1 – a) n<br />

The experimental values of n vary between 1.2 and 1.7. The paste content of concrete is<br />

thus one of the main factors governing shrinkage. It is controlled by the total amount of<br />

cement and water in the concrete. In practical terms, at a constant water/cement ratio, drying<br />

shrinkage increases with an increase in the cement content because this result in a larger<br />

volume of hydrated cement paste that is liable to shrinkage. However, at a given workability,<br />

which roughly means a constant water content, shrinkage is unaffected by an increase in the<br />

cement content, or may even decrease, because the water/cement ratio is reduced and the<br />

concrete is, therefore, better able to resist shrinkage. The overall pattern of these influences<br />

on drying shrinkage is shown in Figure 2.5.<br />

The dual influences of water/cement ratio and aggregate content are shown in Figure 2.6,<br />

but it must be remembered that the shrinkage values given are not more than typical for<br />

drying in a temperate climate.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 145 of 235


Cement<br />

Concrete<br />

The water content of concrete affects shrinkage in so far as it reduces the volume of<br />

restraining aggregate. Thus, in general, the water content of a mix would indicate the order<br />

of shrinkage to be expected, following the general pattern of Figure 2.7, but the water content<br />

per se is not a primary factor. In consequence, mixes having the same water content, but<br />

widely varying in composition, may exhibit different values of shrinkage. An important factor,<br />

which influences the water requirement of the concrete at given consistency, is the<br />

temperature of the fresh concrete. Particularly in hot climates, measures have to be adopted<br />

to keep this temperature at a reasonably low level.<br />

Figure 2.5 - Shrinkage as a function of cement content, water content, and w/c ratio;<br />

concrete moist-cured for 28 days, thereafter dried for 450 days.<br />

Figure 2.6 - Influence of w/c ratio and aggregate content on shrinkage<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 146 of 235


Cement<br />

Concrete<br />

Figure 2.7 - Relation between the water content and drying shrinkage<br />

3.3 Type and Grading of Aggregate<br />

Considerable variation in concrete shrinkage can result from the type of aggregate used,<br />

even within the range of ordinary aggregates (Figure 2.8). This is mainly attributed to the<br />

different elastic properties of the aggregates, which determine the degree of restraint<br />

provided. Aggregates with a higher modulus of elasticity and lower compressibility<br />

respectively give a lower shrinkage. The presence of clay in aggregate lowers its restraining<br />

effect on shrinkage, and since clay itself is subject to shrinkage, clay coatings on aggregate<br />

can increase shrinkage by up to 70%.<br />

Figure 2.8 - Shrinkage of concrete of fixed mix proportions but made with different<br />

aggregates, and stored in air at 21° and a relative humidity of 50 %.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 147 of 235


Cement<br />

Concrete<br />

Beside their compressibility, the shrinkage of the aggregates themselves may be of<br />

considerable importance in determining the shrinkage of the concrete. Shrinking rocks have<br />

usually high absorption (see Table 2.2) rates, so that this can be treated as a warning sign<br />

that the aggregate should be carefully investigated for its shrinkage properties. Sandstone<br />

and slate are typical examples for aggregates, which show already on their own considerable<br />

volume changes upon wetting and drying.<br />

Table 2.2 – Effect of type of aggregate of a single size on concrete shrinkage *[C3]<br />

Lightweight aggregate usually leads to higher shrinkage, largely because the aggregate,<br />

having a lower modulus of elasticity, offers less restraint to the potential shrinkage of the<br />

cement paste.<br />

Size and grading of aggregate per se do not influence the magnitude of shrinkage, unless<br />

water requirement is affected to a greater extent (e.g. in sand rich aggregate mixes). A larger<br />

aggregate, however, permits the use of a leaner mix and hence results in lower shrinkage.<br />

The larger aggregate also tends to restraint the concrete more, and as this may result in<br />

internal microcracking, such internal cracking may be harmful. The particle shape of<br />

aggregate has little effect on concrete shrinkage except if it determines the amount of mixing<br />

water that must be used to obtain a certain workability.<br />

3.4 Cement characteristics<br />

Finer Portland cements show a tendency towards greater concrete shrinkage, but the<br />

increase is normally not very large. In some cases, finer cements produce even concrete<br />

with lower shrinkage. Fineness of cement seems to be a factor as it can change the water<br />

demand of the concrete and because they contain less very coarse cement particles, which<br />

hydrate comparatively little, and can function as restraining bodies along with the aggregate<br />

particles.<br />

Regarding Portland cement composition, there are certain indications that shrinkage is<br />

increased for aluminate and alkali rich cements. Reduced drying shrinkage has been<br />

observed for alite rich cement, which is probably due to a coarser structure of the hydration<br />

product. Carlson found that among the potential compounds in Portland cement their relative<br />

contribution to shrinkage increases in the order C 3 S, C 2 S and C 3 A.<br />

The gypsum dosage in the Portland cement has a major effect on shrinkage. According to<br />

Lerch, drying shrinkage values can be up to 50% higher for cement deficient in gypsum. The<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 148 of 235


Cement<br />

Concrete<br />

optimum gypsum content for minimum shrinkage is more or less equal to the one for<br />

adequate retarding action and strength development, thus primarily depends on C 3 A and<br />

alkali content of the cement. The importance of gypsum dosage to minimise contraction of<br />

concrete was also confirmed by Pickett.<br />

In practice, the shrinkage of concrete of same consistency will vary but little for a wide range<br />

of cement contents, because a richer mix will have a lower water/cement ratio and these<br />

effects offset each other. At constant water content, the amount of shrinkage is more or less<br />

in the same order of magnitude for mixes which are rich or poor in cement (see again Figure<br />

2.5), independent of the water/cement-ratio applied. This phenomenon is explained by the<br />

fact that mixtures with high w/c ratio give off the water more easily, but under lower<br />

development of shrinkage forces, whereas at low w/c-ratio less water evaporates under the<br />

release of higher shrinkage forces.<br />

3.5 Mineral components<br />

Relatively little data is available on the influence of mineral components on the shrinkage<br />

behaviour of concrete. It is believed that the effect of mineral components on drying<br />

shrinkage depends mainly on their influence on the concrete water requirement. In general<br />

those mineral components which increase the water requirement, will increase drying<br />

shrinkage and those, which decrease the water requirement, will decrease it.<br />

Due to the higher water demand, natural pozzolans in the cement will therefore usually<br />

increase the drying shrinkage of concrete. The replacement of clinker by limestone, fly ash<br />

and blast furnace slag does apparently not influence the drying shrinkage behaviour to an<br />

appreciable degree, since the water demand for same workability remains virtually<br />

unchanged compared to Portland cement concrete. Silica fume, when used together with a<br />

superplasticizer to maintain the workability, appears to have also no major effect on drying<br />

shrinkage.<br />

For HPC, where relatively high binder contents and low water/binder ratios are adopted, the<br />

rate of hydration may be so high as to cause severe shrinkage and cracking specially at an<br />

early age. Some work has illustrated that the use of mineral components such as silica fume<br />

and pulverised fuel ash results in increased autogenous shrinkage. Research workers have<br />

also observed that using metakaolin or silica fume at certain water/binder ratios as partial<br />

cement replacement materials may lead to reduced autogenous shrinkage. The authors<br />

attributed the reduction in shrinkage to the resulting pore refinement and improved particle<br />

size distribution of the blend, together with early strength development.<br />

Recently it has been shown that autogenous shrinkage may be negligible by using triple<br />

blend cement with 10% metakaolin and 20% fly ash. Such compositions could have<br />

significant commercial application in controlling and eliminating cracking.<br />

This dual role of mineral components on autogenous shrinkage probably explains why there<br />

is some disagreement between researchers on the influence of certain mineral components<br />

on shrinkage in general.<br />

The situation is further complicated by variations in environmental conditions, curing<br />

procedure and period, cement type and content, use of superplasticizers, mineral<br />

components type and content, and specimen size and shape. It is also possible that a<br />

particular mineral component may reduce one form of shrinkage whilst increasing another.<br />

Within the framework of ESCOCEM-low shrink cement it has been decided to focus on the<br />

use of mineral components for such a development.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 149 of 235


Cement<br />

Concrete<br />

3.6 Chemical Admixtures<br />

The effect of water reducing admixtures on shrinkage can not be correlated with the<br />

reduction in the water content of the concrete. Actually, the influence of plasticizers or<br />

superplasticizers on drying shrinkage is in general negligible despite the substantial water<br />

reduction achieved. The use of such admixtures is thus no guarantee that concrete<br />

shrinkage will be reduced. The only way to know if a certain water-reducing admixture will<br />

result in lower shrinkage is to test it with the corresponding concrete mix.<br />

The entrainment of air by means of corresponding chemical admixtures has been found to<br />

have virtually no influence on shrinkage. Set-retarding admixtures can have a varied effect,<br />

but do generally not influence the shrinkage values to a greater extent. Special attention has<br />

to be given to the accelerators calcium chloride and triethanolamine. The increase of<br />

shrinkage is disproportionally large with small amounts of these admixtures.<br />

Several chemical admixtures designed for the purpose of drying shrinkage reduction are<br />

available in the market (shrinkage reducing admixture). These admixtures are claimed to be<br />

effective measures to minimise shrinkage, but the exact mechanisms are not known.<br />

Possible explanations for the reduced shrinkage could be the weakening of the capillary<br />

tensions or improved water-retaining properties of the concrete.<br />

To keep a watching brief and be aware of this HGRS carried out several studies on concrete<br />

to establish the shrinkage reducing potential of such admixtures. The results of these<br />

different studies are compiled in reports RMP 96/13135/E, PDA 97/17026/E and HC<br />

98/18009. It has been revealed that the potential for shrinkage reduction admixture seems to<br />

be quite limited (30%) with a negative side effect on strength development. It should also be<br />

noted that the use of these admixtures increases the cost of concrete.<br />

3.7 Curing and Storage Conditions<br />

The length of moist curing is relatively unimportant in the control of shrinkage. After a curing<br />

period to achieve adequate strength development of the concrete, additional moist curing will<br />

not help to reduce shrinkage. Prolonged moist curing only delays the onset of shrinkage, but<br />

the final shrinkage values reached remain virtually the same. The amount of shrinkage may<br />

be slightly reduced by the swelling experienced in course of the curing process.<br />

The relative humidity of the medium surrounding the concrete after curing greatly affects the<br />

magnitude of shrinkage (Table 2.1 and Figure 2.9). The smaller the relative humidity of air to<br />

which the concrete is exposed, the higher is the final shrinkage of the corresponding<br />

structural member. Figure 2.9 also illustrates the greater absolute magnitude of shrinkage<br />

compared to swelling in water.<br />

The effect of wind on drying shrinkage of concrete is not significant. As the moisture<br />

conductivity of hardened concrete is so low that only a small rate of evaporation is possible,<br />

this rate cannot be increased by movement of air.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 150 of 235


Cement<br />

Concrete<br />

Figure 2.9 - Relation between shrinkage and time for concrete stored at different<br />

relative humidities.<br />

3.8 Dimension of Structural Member<br />

The dimension of the structural member is a very important factor regulating drying shrinkage<br />

of concrete as the magnitude of shrinkage varies considerably with the size and shape of the<br />

specimen. A part of the size effect may also be due to the pronounced carbonation of small<br />

specimens. For practical purposes, shrinkage thus cannot be considered as purely an<br />

inherent property of the concrete and has always to be related to the size of the concrete<br />

member.<br />

Both the rate and the final values of shrinkage decreases as the concrete member becomes<br />

larger, but above some value the size effect are no longer apparent. Although the<br />

volume/surface ratio does not perfectly reflect variations of both size and shape, there is a<br />

satisfactory correlation with the shrinkage for design purposes. There appears to be a linear<br />

relation between this ratio and the logarithm of shrinkage (Figure 2.10).<br />

Figure 2.10 - Relation between ultimate shrinkage and volume/surface ratio.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 151 of 235


Cement<br />

Concrete<br />

Accordingly it can be concluded that drying shrinkage in mass concrete is not of major<br />

importance, unless the surface drying contributes to the beginning cracks that may otherwise<br />

not form. Thin structural members and structures with large exposed surfaces (e.g. slabs and<br />

walls) will always have a higher shrinkage and are thus more prone to cracking, particularly if<br />

they are not free to move.<br />

3.9 Reinforcement<br />

In reinforced concrete drying shrinkage results in a compressive stress in the bars, which<br />

themselves resist and reduce the overall drying shrinkage in the reinforced concrete<br />

member. The actual value of shrinkage reduction is obviously dependent on the amount of<br />

reinforcement used.<br />

The addition of discrete fibres such as polypropylene, glass and steel to the concrete can<br />

reduce drying shrinkage to a certain extent. More than for the purpose of shrinkage reduction<br />

itself these fibres are, however, added to prevent cracking of the concrete due to shrinkage.<br />

3.10 Combined Effect on Individual Factors<br />

It is believed that the combined effect of various influencing factors on concrete shrinkage is<br />

the product and not the sum of the individual effects of the concerned factors. This<br />

cumulative effect on shrinkage of making poor choices in the selection of materials and<br />

construction practices has been emphasised by Powers and Tremper.<br />

To illustrate the effect of combination of influencing factors, the shrinkage results of Powers<br />

are presented in Table 2.3. It shows the cumulative and individual effects of the most<br />

unfavourable versus the most favourable choices with regard to six factors influencing the<br />

amount of shrinkage. He concluded from his data that wrong choices of alternatives with<br />

respect to volume change could result in about seven time as much shrinkage as would<br />

result from the best choices.<br />

It is of course not very likely that all negative factors occur at the same time. Nevertheless, it<br />

is worthwhile to note the magnitude of increase in shrinkage that can occur if several<br />

unfavourable choices are made.<br />

Table 2.3 – Individual and cumulative effects of various factors on<br />

shrinkage<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 152 of 235


Cement<br />

Concrete<br />

4 Control of Cracking due to Shrinkage<br />

4.1 Introduction<br />

The reason why cracks can be formed by drying shrinkage has already been illustrated in<br />

Figure 2.2. If hardened concrete could shrink without any restraint, it would not crack.<br />

However, in a real structure the concrete is always subject to some degree of restraint either<br />

by the foundation or another part of the structure or by the reinforcing steel embedded in the<br />

concrete. The combination of shrinkage and restraint develops tensile stresses and if these<br />

exceed the tensile strength of the concrete, cracks will be formed.<br />

The magnitude of tensile stress developed during drying of hardened concrete is not only<br />

influenced by the amount of shrinkage, but also by other factors such as:<br />

• degree of restraint<br />

• modulus of elasticity of the concrete<br />

• creep or relaxation of the concrete<br />

As far as cracking is concerned, a high extensibility of concrete (low modulus of elasticity and<br />

high creep) is generally desirable, because it permits concrete to withstand greater volume<br />

changes.<br />

The importance of cracking due to drying shrinkage and the minimum width at which a crack<br />

is considered significant depend on the conditions of exposure of the concrete. A general<br />

guide for the tolerable crack width in reinforced concrete is given in Table 2.4. Since under<br />

given physical conditions, the total crack width per unit length of concrete is constant and the<br />

cracks should be as fine as possible, it is desirable to have more cracks. Accordingly, the<br />

restraint to cracking should be uniform along the length of the member.<br />

The measures adopted for the control of cracking due to drying shrinkage comprise the<br />

reduction of the cracking tendency of the concrete to a minimum, the use of adequate and<br />

properly positioned reinforcement and the use of control joints. Cracking can also be<br />

minimised by the application of expansive cements to produce shrinkage compensating<br />

concrete.<br />

Table 2.4 – Tolerable crack width in reinforced concrete according to exposure<br />

condition<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 153 of 235


Cement<br />

Concrete<br />

4.2 Reduction of Cracking Tendency<br />

As mentioned previously, the cracking tendency is due not only to the amount of shrinkage,<br />

but also to the degree of restraint, the modulus of elasticity and the creep or relaxation of the<br />

concrete. Some factors, which reduce the shrinkage at the same time may decrease the<br />

creep and relaxation and increase the modulus of elasticity, so that they offer only little help<br />

to reduce cracking tendency. Nevertheless, in a general way, the measures that can be<br />

taken to reduce the shrinkage of concrete should also decrease the cracking tendency.<br />

In accordance with the influencing factors described in section 2.3, corresponding measures<br />

can be derived to reduce the amount of drying shrinkage of concrete. For least shrinkage,<br />

the concrete mix proportioning should incorporate those factors that contribute to the lowest<br />

paste and water content respectively, which means:<br />

• largest practical maximum size of aggregate<br />

• lowest practical sand content<br />

• lowest practical slump<br />

• lowest practical temperature of fresh concrete<br />

Regarding the materials used for concreting, the following selections have to be preferred:<br />

• cement with low C 3 A and alkali and high C 3 S content, optimum gypsum dosage and not<br />

too high fineness (preferably no ASTM type III cement or equivalent)<br />

• aggregate of low compressibility and absorption, free of clay, dirt and excessive fines.<br />

The use of calcium chloride and triethanolamine containing admixtures should be avoided. In<br />

how far the reduction in water content by means of water reducers will result in a net<br />

shrinkage reduction must be tested on trial mixes. The use of shrinkage reducing chemical<br />

admixtures may also be taken into consideration.<br />

Another way to reduce the cracking tendency, which is seldom used, is to apply a surface<br />

coating to the concrete, that will prevent the rapid loss of moisture. Effective coatings are for<br />

instance chlorinated rubber and waxy or resinous materials. The slowing down of the rate of<br />

evaporation and thus the rate of shrinkage will be beneficial, because concrete has a<br />

remarkable quality of relaxing under sustained stress. Concrete may be able to withstand<br />

without cracking two or three times as much slowly applied shrinkage as it can rapid<br />

shrinkage.<br />

4.3 Reinforcement<br />

Properly placed reinforcement in adequate amounts not only reduces the total amount of<br />

shrinkage, but prevents also unsightly cracking. By the distribution of the shrinkage strains<br />

along the reinforcement through bond stresses, a larger number of very fine cracks will occur<br />

instead of few wide cracks.<br />

The placement of reinforcement to limit crack width can be mainly foreseen for relatively thin<br />

structural members. In massive structures, reinforcement may be placed in the 20 to 40 cm<br />

thick surface zone, since in the interior massive concrete will shrink only to a little extent.<br />

The addition of discrete fibres to the concrete matrix is a further alternative to reduce<br />

cracking tendency. The fibres increase the tensile strength of the concrete, so that it can<br />

withstand higher shrinkage without cracking.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 154 of 235


Cement<br />

Concrete<br />

4.4 Joints<br />

The use of joints is the most effective method of preventing formation of unsightly cracking. If<br />

no joints are provided in large surface structures such as walls, slabs or pavements, the<br />

concrete will make its own “joints” by cracking.<br />

Contraction joints in walls are made, for example, by fastening wood or rubber strips to the<br />

forms, which leave narrow vertical grooves in the concrete on the inside and outside of the<br />

wall. Cracking of the wall due to shrinkage should occur at the grooves, relieving the stresses<br />

and preventing formation of unsightly cracks. The grooves should be sealed on the outside of<br />

the wall to prevent penetration of moisture. Sawed joints are commonly used in pavements,<br />

slabs and floors.<br />

Some guide values for the distances of contraction joints in structural concrete members are<br />

given in Table 2.5. For further indications on the placing and formation of joints, it is referred<br />

to the relevant literature.<br />

Table 2.5 – Guide values for the maximum permissible distance an of contraction<br />

joints in some structural concrete members. The values vary<br />

considerably according to insulation, heat insulation, reinforcement or<br />

anchorage.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 155 of 235


Cement<br />

Concrete<br />

4.5 Shrinkage-Compensating Concrete<br />

A possible way to minimise or reduce shrinkage cracking is shrinkage compensation by the<br />

use of expansive cements. General guidelines for the use of such shrinkage-compensating<br />

concrete have been established in the ACI Manual of Concrete Practice.<br />

The basic concept of shrinkage compensation by means of expansive cements is to increase<br />

the concrete volume before the commencement of shrinkage, so that the concrete will attain<br />

after the shrinkage process its original volume. The use of expansive cement in shrinkagecompensating<br />

concrete does not accordingly prevent the development of shrinkage; the early<br />

expansion just balances the subsequent normal shrinkage.<br />

There exist several types of expansive cements to be used in shrinkage compensating<br />

concrete. The specific features and characteristics of these cements and their application will<br />

be described in the subsequent section.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 156 of 235


Cement<br />

Concrete<br />

Concrete Durability<br />

1. INTRODUCTION ............................................................................................................... 158<br />

2. DEFINITION ...................................................................................................................... 158<br />

3. DELETERIOUS PROCESSES AFFECTING CONCRETE............................................... 158<br />

3.1 Sulfate Attack......................................................................................................... 158<br />

3.2 Chloride Attack ....................................................................................................... 159<br />

3.3 Acid Attacks............................................................................................................ 159<br />

3.4 Carbonation ............................................................................................................ 159<br />

3.5 Alkali-aggregate Expansive Reactions ................................................................... 161<br />

3.6 Abrasion.................................................................................................................. 161<br />

3.7 Biodeterioration....................................................................................................... 161<br />

3.8 Freeze-Thaw........................................................................................................... 161<br />

4. CHOOSING THE MATERIALS AND MIX DESIGN PARAMETERS TO<br />

GUARANTEE CONCRETE DURABILITY...................................................................... 162<br />

4.1 Cement ................................................................................................................... 162<br />

4.2 Aggregates ............................................................................................................. 162<br />

4.3 Water ...................................................................................................................... 162<br />

5. MIX DESIGN ..................................................................................................................... 163<br />

5.1 Water Cement Ratio ............................................................................................... 163<br />

5.2 Compressive Strength ............................................................................................ 166<br />

6. CONCLUSIONS ................................................................................................................ 166<br />

7. REFERENCES.................................................................................................................. 167<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 157 of 235


Cement<br />

Concrete<br />

1 Introduction<br />

During many years the strength was considered the most important property of the concrete<br />

and the parameter to access its quality. However, nowadays the durability of the concrete is<br />

gaining importance, particularly due to the high costs involved on repairs and maintenance of<br />

structures of concrete incorrectly proportioned or applied. In UK, for example, around 40% of<br />

the total invested on civil construction is destined to repairs and maintenance (Neville 1997)<br />

and in Europe the values spent on repairs of bridges grew approximately 65% between 1985<br />

and 1990. According to the Federal Highway Administration (USA) (in Nmai 2000), 42% of<br />

the 575600 American bridges can be considered structurally deficient or unsafe due to the<br />

corrosion of their reinforcements. It is also estimated that, from 1981 to 2000, 102 billions<br />

dollars were spent on repairs of these bridges and that, from this total, 1 billion dollars could<br />

have been saved only with the improvement of concrete quality.<br />

Studies carried on all over the world show that the majority of the durability problems are<br />

related to bad execution, but the decline of concrete quality is also related to the use of nonappropriated<br />

materials and mix design, with water to cement ratio higher and higher.<br />

According to Wischers (1984) to produce a concrete with compressive strength of 30MPa, at<br />

28 days, on the 40's in England, it was necessary to work with w/c equal to 0.47 and cement<br />

content of 300kg/m 3 . Thirty years later, that means in the 70's, this same concrete could be<br />

produced with w/c equal 0.72 and cement content of 250kg/m 3 . What is apparently an<br />

advantage, i. e., the decrease of cement content and the increase of water to cement ratio,<br />

is, in fact, one of the main causes of the augmentation of durability problems and of losses<br />

due to repairs and maintenance of concrete structures.<br />

2 Definition<br />

Durability is the capacity that the concrete has to resist the weathering (freeze-thaw),<br />

abrasion or chemical processes, such as sulfate and chloride attack, acid attack, carbonation<br />

and reactions involving alkalis and aggregates. These processes may lead to the formation<br />

of expansive products with consequent concrete cracking or to the decomposition of the<br />

hydrated components of the cement paste.<br />

3 Deleterious Processes Affecting Concrete<br />

3.1 Sulfate Attack<br />

The sulfate attack occurs through the reaction between the sulfates, in general from external<br />

sources, such as polluted water or soil, and the C 3 A presents on the Portland clinker and/or<br />

the Ca(OH) 2 generated by cement hydration. It results on ettringite and gypsum formation,<br />

with consequent expansion and fissures.<br />

The sulfate attack also promotes deterioration of the hydrated cement products, causing<br />

losses of concrete mass and strength.<br />

This type of pathology can be prevented by using sulfate resistant cements (normally with<br />

low C 3 A) or cements with high amounts of active mineral components (slag, fly ash or<br />

pozzolan).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 158 of 235


Cement<br />

Concrete<br />

3.2 Chloride Attack<br />

Chloride ions can be introduced into the concrete by its mixing water, by the use of<br />

contaminated aggregates, de-icing salts or chloride-based accelerators (whose usage is<br />

forbidden in several countries).<br />

In alkaline environments the concrete reinforcement is covered by a thin, impermeable and<br />

adherent film of iron oxide that protects it from corrosion. In this case, the reinforcement is<br />

"passivated". The chloride ions can destroy this layer de-passivating the rebars. This process<br />

that can be described according to the reactions 1 and 2 generates iron hydroxide. As this<br />

compound has volume 600% higher than the reagents, its formation is accompanied by<br />

expansion.<br />

(1) Fe 2+ + 2Cl - ⇒ FeCl 2<br />

(2) FeCl 2 + 2H 2 O ⇒ Fe(OH) 2 + 2HCl<br />

The effects of chloride attack are twofold - losses of rebar section due to the iron<br />

decomposition and fissures caused by the expansion.<br />

The chloride attack is minimized by decreasing the permeability of the cement paste, a<br />

characteristic that can be achieved by the adoption of low w/c and by the use of cements<br />

with high amount of active mineral components.<br />

3.3 Acid Attacks<br />

The cement paste constituents are compounds stable in alkaline environment, with low<br />

resistance to acids. Thus, when the concrete pH decreases reaching values lowers than 6.5,<br />

the calcium hydroxide produced during cement hydration reacts with the acids, generating<br />

water-soluble compounds that are then lixiviated, increasing the concrete permeability.<br />

These acids have diverse origin, being derived, for example, from the combustion of fuels<br />

with sulfur (sulfuric acid), from the metabolism of microorganisms and animals (organic<br />

acids) and industrial residues.<br />

3.4 Carbonation<br />

In alkaline environment the concrete rebars are protect from corrosion. However when the<br />

pH changes in direction to more acid values, this protection is destroyed and corrosion can<br />

take place.<br />

The carbonation - a chemical process that involves the reaction of the calcium hydroxide<br />

from the cement paste and the CO 2 from the atmosphere - modifies the pH of the concrete,<br />

favoring the corrosion of its reinforcement.<br />

The carbonation is accelerated in concretes exposed to environments with relative humidity<br />

between 50% and 70% also increasing with the increase of CO 2 supplied and the w/c.<br />

Although the carbonation is an inevitable process, it can be reduced with the use of cements<br />

with low amount of mineral components, with adequate cement content and with the increase<br />

of concrete strength and curing period (Figures 1 and 2).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 159 of 235


Cement<br />

Concrete<br />

10<br />

9<br />

8<br />

Depth of Carbonation (mm)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

humid cure<br />

dry cure<br />

2<br />

1<br />

0<br />

35 40 45 50 55 60<br />

Compressive Strength (MPa)<br />

FIGURE 1 - Influence of curing on concrete compressive strength and carbonation<br />

depth of concrete (after Neville 1995).<br />

140<br />

TIME OF CORROSION INITIATION (days)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 5 10 15 20 25 30<br />

CURING PERIOD (days)<br />

FIGURE 2 – Influence of curing duration on corrosion initiation (after Neville 1995)<br />

Corrosion can also be prevented by protecting the reinforcement with a layer of concrete<br />

thick enough to guarantee that the carbonation front will not reach the rebar. This thickness<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 160 of 235


Cement<br />

Concrete<br />

of concrete cover is a parameter included in many national concrete standards. The ACI<br />

Manual of Concrete Practice 2002, Part 1 (201.2R-19) consider that for concrete exposed to<br />

moderate-to-severe corrosion environments, the concrete covering the rebars should be at<br />

least 38mm and preferably, 50mm thick.<br />

3.5 Alkali-aggregate Expansive Reactions<br />

The alkali-aggregate reactions are chemical processes that involve some types of<br />

aggregates and sodium or potassium ions. They generate expansive products that filling up<br />

porous and discontinuities of concrete paste, cause fissures and its disintegration.<br />

Once started, these reactions cannot be stopped, but they can be prevented by using inert<br />

aggregates or cements with high amount of active mineral components.<br />

3.6 Abrasion<br />

The concrete abrasion leads to its disintegration. Abrasion effects can be minimized by the<br />

densification of the cement paste, which is achieved by adopting low w/c. Experience shows<br />

that the resistance to abrasion of a concrete with w/c around 0.30 is comparable to the one<br />

of good quality rocks.<br />

3.7 Biodeterioration<br />

The biodeterioration process is, essentially, an acid attack process to the concrete. Some<br />

bacteria and fungi can grow on concrete surface and through the acids released during their<br />

metabolism, decompose the cement hydrated compounds.<br />

The reduction of the w/c results in hardened cement pastes with low porosity, making difficult<br />

the proliferation of microorganisms and concrete degradation.<br />

3.8 Freeze-Thaw<br />

When the temperature of a saturated concrete decreases, the water present in its porous can<br />

freeze and the augment of volume that accompanies this phenomenon causes concrete<br />

expansion.<br />

The repetition of this process due to successive cycles of freeze-thaw can lead to<br />

disintegration of the concrete.<br />

The resistance of the concrete to freeze-thaw can be improved by decreasing its<br />

permeability, which is obtained by using low water to cement ratio, and also by using air<br />

entraining agents: the voids they create in the hardened cement paste allow the water<br />

expansion without damaging the concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 161 of 235


Cement<br />

Concrete<br />

4 Choosing the Materials and Mix Design Parameters to<br />

Guarantee Concrete Durability<br />

The processes that lead to concrete deterioration can be avoided or minimized with the right<br />

choice of its constituents, rational mix design and careful preparation and application.<br />

4.1 Cement<br />

In many markets it is possible to find several cement types, whose difference may be due to<br />

the kind of additions (MIC) their contain. In general, it could be said that the cements can be<br />

divided in 2 groups, namely: OPC, which is composed essentially by Portland clinker and<br />

gypsum, composite cements, which can have active mineral components such as slag, fly<br />

ash and pozzolan as addition, or limestone.<br />

The addition of active mineral components has as main consequence, the improvement of<br />

the durability of the cement products. This is achieved through the reaction between the MIC<br />

and the Ca(OH) 2 , with formation of C-S-H, and by the dilution of the C 3 A of the clinker,<br />

minimizing the potentially deleterious actions of these two compounds in the hardened<br />

concrete. As seen, the C 3 A can, for example, react with sulfates and chlorides generating<br />

expansive products, being also responsible for expressive amount of the heat released<br />

during cement hydration.<br />

The Ca(OH) 2 can also react with sulfates, resulting in expansion, can suffer carbonation,<br />

favoring rebar corrosion, being lixiviated by water, generating efflorescence and increasing<br />

concrete permeability, as well acting as a catalytic for alkali-aggregate reaction.<br />

The active mineral components decrease or avoid the deleterious effect of these processes,<br />

producing concretes with lower heat of hydration and permeability and therefore, more<br />

durables.<br />

4.2 Aggregates<br />

The coarse and fine aggregates should have characteristics compatible with the standards<br />

requirements and free of harmful substances or compounds that can take part on deleterious<br />

reactions. The most common harmful substances present in aggregates are organic material,<br />

like leaves for example, organic substances such as sugar and high content of very fine<br />

material or clay. Additionally, the aggregate should be preferably inert face to alkalis, in order<br />

to prevent expansive reactions.<br />

4.3 Water<br />

The concrete mixing water should be potable and free of substances that may be harmful to<br />

the concrete, as for example, sugar, acids or citrates. It should also have limited sulfate,<br />

chloride and alkali content.<br />

However, as important as the quality is the quantity of water added to the concrete. If in<br />

excess it creates voids that decreases its strength, increases its permeability and facilitates<br />

deleterious attack driven by the ingress of aggressive agents into its cement paste.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 162 of 235


Cement<br />

Concrete<br />

5 Mix Design<br />

5.1 Water Cement Ratio<br />

The w/c is one of the most important parameters regarding concrete quality. Since Abrams'<br />

studies in 1919, it is known that there is an inversely proportional relation between w/c and<br />

concrete compressive strength (Figure 3). Nowadays, it is also known that the w/c has a<br />

fundamental role on concrete quality and durability.<br />

45<br />

CONCRETE COMPRESSIVE STRENGTH (MPa)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0.4 0.5 0.6 0.7 0.8 0.9 1 1.1<br />

FIGURE 3 - Relation between w/c and concrete compressive strength<br />

The quantity of water required for the complete hydration of the cement added to the<br />

concrete is equivalent to a w/c around 0.30. Therefore, higher values indicate water excess<br />

and available to create voids and channels of percolation on the hardened cement paste, as<br />

can be seen on Photo 1. Naturally, these voids and discontinuities cause decrease of<br />

concrete strength and as they are interconnected, favor the ingress of deleterious agents into<br />

the concrete.<br />

Aiming guaranteeing a long service-life for concrete structures, the European standard EN<br />

206/1 - Concrete - Part 1: Specification, performance, production and conformity, establishes<br />

maximum w/c according to the environment to what concrete will be exposed. It is<br />

summarized on Table 1.<br />

w/c<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 163 of 235


Cement<br />

Concrete<br />

PHOTO 1 - Concrete observed under petrographic microscope, showing interconected<br />

voids created by water in excess<br />

TABLE 1 - Exposure conditions and requirements of EN 206/1 (simplified)<br />

Type of attack<br />

Chloride-induced<br />

corrosion<br />

Carbonatio<br />

n-induced<br />

corrosion<br />

Sea water<br />

Chloride<br />

other than<br />

from sea<br />

water<br />

Freezethaw<br />

attack<br />

Aggressive<br />

chemical<br />

environment<br />

Exposure class XC1 XC4 XS1 XS3 XD1 XD3 XF1 XF4 XA1 XA3<br />

Maximum w/c 0.65 0.50 0.50 0.45 0.55 0.45 0.55 0.45 0.55 0.45<br />

Minimum cement<br />

content (kg/m 3 )<br />

260 300 300 340 300 320 300 340 300 360<br />

The maximum value permitted, equal to 0.65 is justified by studies that show that in<br />

hardened cement pastes with w/c higher than 0.70 all porous are connected,<br />

increasing significantly its permeability and decreasing its durability.<br />

It means that the higher the aggressiveness of the environment, the lower has to be the w/c.<br />

Photos 2 and 3 show concretes with w/c equal to 0.30 and 0.50 and highlight the<br />

microstructural differences due to the decrease of the water content in concrete. With w/c<br />

equal to 0.30 the cement paste is more compact and therefore, more resistant and durable.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 164 of 235


Cement<br />

Concrete<br />

1.1<br />

1.1<br />

PHOTO 2 – Concrete with w/c = 0,50 at 28 days, showing the aggregate<br />

(G) and the cement paste (P) with high porosity. Scanning electron<br />

microscope; magnification 1500x (in Vieira 1998).<br />

1.1<br />

1.1<br />

PHOTO 3 – Concrete with w/c = 0,30 at 28 days, showing aggregate (G)<br />

and cement paste (P) compact, without porous. Scanning electron<br />

microscope; magnification 1500x (in Vieira 1998).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 165 of 235


Cement<br />

Concrete<br />

5.2 Compressive Strength<br />

Although in many countries minimum compressive strength values are not a required in the<br />

concrete standards, nowadays there is a trend in the sense of including it in the<br />

specifications and normally values around 20MPa - 25MPa are being adopted.<br />

The explanation for this resolution is that concretes prepared with the maximum w/c ratio and<br />

minimum cement content determined by the standards (for example the European one - EN<br />

206/1), will have compressive strength higher than 20MPa - 25MPa. Therefore the<br />

compressive strength can be used as a indirect mean of verification the compliance with the<br />

standards, concerning durability.<br />

6 Conclusions<br />

The durability of the concrete is one of the fundamental aspects that should be considered<br />

since the project phase and materials selection, until the final stages of its application.<br />

Regarding materials, the quality of the concrete depends on the quality of the<br />

aggregates and mixing water, on the right choice of the cement type and when using<br />

chemical admixtures, on the selection of products compatibles with the cement<br />

selected.<br />

The proportioning of theses materials has to be rationally defined: particularly important in<br />

this phase, are the definitions of the optimums cement content and w/c.<br />

The compressive strength specified in the project has to be compatible with the solicitations<br />

to what the concrete will be subject to and in case of reinforced concrete the concrete layer<br />

that covers the rebars have to have the thickness required by the standards in order to<br />

guarantee their protection against corrosion.<br />

Finally, the execution, application and cure of the concrete have to be carefully made.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 166 of 235


Cement<br />

Concrete<br />

7 References<br />

NEVILLE, A. M. (1995) Properties of Concrete. Longman Group Limited. 4 th Ed. 844p.<br />

NEVILLE, A. M. (1997) Maintenance and durability of structures. Concrete International<br />

v. 19, n. 11, p: 52-56.<br />

NMAI, C.K. Recent developments in the design of reinforced concrete structures for<br />

long service lives from a corrosion perspective. In: CONGRESSO BRASILEIRO DO<br />

CONCRETO, 42, 13-18 de agosto 2000, Ceará. Anais...<br />

VIEIRA, S. R. S. S. Concretos comum e de alto desempenho: análise do<br />

comportamento através da microscopia eletrônica de varredura. (Ordinary concrete and<br />

high performance concrete: A behaviour analysis through scanning electron microscopy). In:<br />

CONGRESSO BRASILEIRO DO CONCRETO, 40, Rio de Janeiro, 1998, Rio de Janeiro.<br />

Anais... (in Portuguese).<br />

VIEIRA, S. R. S. S. Vida longa às estruturas de concreto: o concreto de alto<br />

desempenho como garantia da durabilidade. (Long life to the concrete structures: the<br />

high performance concrete as guarantee of durability). Revista Engenharia, Ciencia &<br />

Tecnologia, Vitória, ES, v. 4, n.1, p.51-56, jan/fev. 2001 (in Portuguese).<br />

WISCHERS, G. (1984) The Impact of the Quality of Concrete Constructions on the<br />

Cement Market. "Holderbank" Group Meeting 1984.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 167 of 235


Cement<br />

Concrete<br />

Alkali-Aggregate Reactions in Concrete<br />

1. INTRODUCTION AND DEFINITIONS .............................................................................. 169<br />

2. ALKALI-AGGREGATE REACTION: REQUIREMENTS .................................................. 169<br />

3. TYPICAL FEATURES FROM ALKALI-AGGREGATE REACTION................................. 169<br />

4. HISTORY........................................................................................................................... 174<br />

5. TYPES OF ALKALI-AGGREGATE REACTION .............................................................. 175<br />

5.1.1 Alkali-silica reaction.................................................................................. 175<br />

5.1.2 Alkali-silicate reaction............................................................................... 175<br />

5.1.3 Alkali-carbonate reaction.......................................................................... 175<br />

6. PREVENTING THE ALKALI-AGGREGATE REACTION................................................. 175<br />

7. AGGREGATE REACTIVITY ............................................................................................. 176<br />

8. TESTING METHODS ........................................................................................................ 178<br />

8.1 Evaluation of the potential reactivity of the aggregate ............................................ 178<br />

8.1.1 Petrography.............................................................................................. 178<br />

8.1.2 Accelerated Test (ASTM C 1260/94) ....................................................... 178<br />

8.2 Evaluation of the efficiency of a cement or MIC to inhibit the alkaliaggregate<br />

reaction ................................................................................................. 179<br />

8.2.1 Accelerated Test (ASTM C 441) .............................................................. 179<br />

8.3 Evaluation of the potential reactivity of a certain cement-aggregate<br />

combination............................................................................................................ 179<br />

8.3.1 Accelerated method (ASTM C227) .......................................................... 179<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 168 of 235


Cement<br />

Concrete<br />

1 Introduction and Definitions<br />

Alkali-aggregate reactions are complex chemical processes that occur in concrete, involving<br />

soluble alkali, namely sodium and potassium, and reactive aggregates. The result of the<br />

alkali aggregate reactions are expansive products - amorphous or crystallised - which filling<br />

up discontinuities of the hydrated paste, can cause fissures (Photo 1) and structural<br />

displacements.<br />

The source of alkalis is, normally, the cement, but they can also be derived from some<br />

chemical admixtures used in concrete, from aggregate contamination, concrete mixing water,<br />

etc.<br />

The reactivity of the aggregates is determined by the presence of some specific compounds<br />

or by textural characteristics, which denote stress.<br />

2 Alkali-Aggregate Reaction: Requirements<br />

Besides the soluble alkalis and reactive aggregates, the occurrence of the alkali-aggregate<br />

reactions also depends on the availability of water, even in very small amounts, percolating<br />

the hardened cement paste. Thus, the permeability of the concrete is also an important factor<br />

in the process: the more permeable the concrete, the more susceptible to damages due to<br />

these expansive reactions.<br />

The presence of construction defects, shrinkage fissures or other discontinuities increases<br />

the permeability of the concrete, favouring the alkali-aggregate reactions.<br />

3 Typical Features from Alkali-Aggregate Reaction<br />

The alkali-aggregate reactions can be recognized by cracks with a very typical "map" pattern<br />

(Photos 1 to ) and by the presence of gel, as a white, glassy or porcelain-like material around<br />

the aggregates or filling up porous or cracks in the cement paste. .<br />

In a microscopic level, crystallized products from alkali-aggregate reactions can also be<br />

identified under different shapes: filaments, flowers or leaves-like (Photos ).<br />

Both, the gel and the crystallized products are compounds of silica, calcium and alkalis, most<br />

commonly, potassium (Figure 1).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 169 of 235


Cement<br />

Concrete<br />

Map cracks<br />

PHOTO 1: Cracks due to alkali-aggregate reaction<br />

F<br />

PHOTO 2: Crack in concrete, due to alkali-aggregate reaction, partially filled up with<br />

expansive product (white).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 170 of 235


Cement<br />

Concrete<br />

PHOTO 3: Gel white-bluish disperses on the cement paste and around aggregate<br />

evidencing alkali-aggregate reaction. Stereoscopic microscope; magnification 6x.<br />

PHOTO 4: White-bluish gel contouring aggregate in concrete affected by alkaliaggregate<br />

reaction. Stereoscopic microscope; magnification 12x.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 171 of 235


Cement<br />

Concrete<br />

PHOTO 5: Micro-crack in concrete, generated by expansions due alkali-aggregate<br />

reaction. Concrete observed under transmitted light microscope; magnification 50x.<br />

PHOTO 6: Crystallized product from alkali-aggregate reaction, growing on the<br />

aggregate surface. Concrete observed under scanning electron microscope;<br />

magnification 750x.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 172 of 235


Cement<br />

Concrete<br />

PHOTO 7 – Filament-like alkali-aggregate reaction product, identified in concrete.<br />

Scanning electron microscope; magnification 500x.<br />

PHOTO 8 – Alkali-aggregate reaction product identified in concrete. Scanning electron<br />

microscope; magnification 750x.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 173 of 235


Cement<br />

Concrete<br />

PHOTO 9 – Gel produced by alkali-aggregate reaction, observed under scanning<br />

electron microscope (magnification 125x). In the box, it is represented<br />

its chemical composition.<br />

4 History<br />

The alkali aggregate reaction was first described in the early 40's by Stanton, in California,<br />

and since then, several occurrences have been reported all over the world. Dams are the<br />

most common structure affected by alkali-aggregate reactions that have been also found in<br />

concrete pavements, bridges, railway sleepers, etc.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 174 of 235


Cement<br />

Concrete<br />

5 Types of Alkali-Aggregate Reaction<br />

The alkali-aggregate reactions can be divided in three groups, according to the mineral or<br />

compound responsible for the expansion. There are:<br />

5.1.1 Alkali-silica reaction<br />

The fastest and better known reaction.<br />

Occurs between alkalis and some varieties of amorphous or cryptocrystalline silica, such as,<br />

volcanic glass, found in some basalts (photo 1), opal and chalcedony.<br />

(1) SiO 2 + Na+ or K+ + H 2 O → (Na or K) 2 SiO 3 .2H 2 O<br />

Expansive gel<br />

Derived from the aggregate<br />

Derived from the cement, chemical<br />

admixtures, water, etc<br />

5.1.2 Alkali-silicate reaction<br />

This reaction is similar to the alkali-silica one, occurring between some silicates, specially<br />

deformed or microcrystalline quartz, present in rocks as gneiss, granites, milonites, etc.<br />

Some textural and structural features of the rocks are indicative of its potential reactivity face<br />

to alkalis. The most important are:<br />

Presence of quartz with undulatory extinction, presence of minerals in fine or very fine grains,<br />

and minerals strongly oriented according to preferential directions, due to shear.<br />

5.1.3 Alkali-carbonate reaction<br />

The mechanism that leads to alkali-carbonate reaction is not very well understood yet, but it<br />

seems to be consensus that it involves calcareous rocks with both calcite and dolomite<br />

(calcite-to-dolomite ratio = 1) and 5% to 25% of clay minerals. One of the mechanisms<br />

proposed to explain the alkali-carbonate reaction is the dolomite decomposition, with<br />

formation of brucite, according to the reaction 2. Nevertheless some authors believe that the<br />

expansion is due to the clay constituents of the rock, and, therefore, an alkali-silicate<br />

reaction, in fact.<br />

(2) CaMg(CO 3 ) 2 + 2(K or Na)OH → CaCO 3 + Mg(OH) 2 + (Na or K) 2 CO 3<br />

6 Preventing the Alkali-Aggregate Reaction<br />

The alkali-aggregate reactions are process that, once started cannot be paralysed, and<br />

therefore their prevention is fundamental. It can be done using inert aggregates, cement with<br />

low alkali content (Na eq. < 0.60%) or high amount of active mineral components or limiting<br />

the amount of soluble alkalis in concrete, which according to the general recommendations<br />

should not exceed 3.0kg/m 3 .<br />

The Ca(OH) 2 presents in the hardened cement paste contributes, as a catalyst, to the alkaliaggregate<br />

reactions; the pozzolans, fly ashes, slags and silica fume consume the calcium<br />

hydroxide, thus inhibiting the process.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 175 of 235


Cement<br />

Concrete<br />

7 Aggregate reactivity<br />

The aggregate reactivity depends on the presence of some specific constituents and<br />

structural/textural aspects derived from the rock genesis. Some reactive compounds are:<br />

Volcanic glass, present in some basalts (Photo 10)<br />

<br />

<br />

Strained, deformed quartz, a common constituent of granites, gneiss and<br />

quartzites (Photo 11)<br />

Varieties of amorphous or micro-crystalline silica, such as opal and chalcedony,<br />

found e.g., in gravels.<br />

Some textural and structural aspects are also indicative of rock reactivity. They are:<br />

Very fine silicates, as fine-grained quartz,<br />

Very pronounced foliation produced by shear processes (Photo 12).<br />

V<br />

PHOTO 10: Volcanic glass (V) as a constituent of a basalt rock. Its presence indicates the<br />

reactivity of the aggregate to alkalis. Optical microscope, magnification 25x.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 176 of 235


Cement<br />

Concrete<br />

Q<br />

PHOTO 11: Quartz (Q) with undulatory extinction evidencing stress and the potential<br />

reactivity of this aggregate. Optical microscope, magnification 25x.<br />

PHOTO 12: Aggregate showing minerals strongly oriented due to shear. This feature<br />

suggests its reactivity when in contact with alkalis. Optical microscope; magnification 25x.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 177 of 235


Cement<br />

Concrete<br />

8 Testing Methods<br />

Several methods have been proposed to access the reactivity of aggregate and the<br />

performance of a cement or cement-aggregate combination with respect to alkali-aggregate<br />

reaction. Here only some of them - the most commonly used ones are presented.<br />

8.1 Evaluation of the potential reactivity of the aggregate<br />

8.1.1 Petrography<br />

Petrography is a technique that allows the identification of the mineral composition of the<br />

aggregate and the way these compounds are distributed (structure/texture). In order to<br />

evaluate its potential reactivity to soluble alkalis, a thin section of the aggregate is analysed<br />

under the petrographical microscope, where features such as:<br />

<br />

<br />

<br />

Presence of volcanic glass, opal, chalcedony, etc,<br />

Deformed quartz with undulatory extinction,<br />

Very fine to fine grained silicates,<br />

Minerals strongly oriented, as result of shear<br />

can be recognized.<br />

8.1.2 Accelerated Test (ASTM C 1260/94)<br />

The method prescribed in the ASTM C 1260/94 access the potential reactivity of the<br />

aggregate through the measurement of the expansions showed by mortar bar prepared with<br />

the studied aggregate and a no-inhibitor cement, after curing for 14 days in a solution of<br />

NaOH, at 80 o C.<br />

According to this procedure, if at 14 days curing, the expansions are:<br />

Lower than 0.10%, the aggregate is considered inert,<br />

Higher than 0.20%, the aggregate is considered reactive and<br />

<br />

Between 0.10% and 0.20%, the aggregate is considered potentially reactive, but<br />

additional tests are necessary.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 178 of 235


Cement<br />

Concrete<br />

8.2 Evaluation of the efficiency of a cement or MIC to inhibit the alkaliaggregate<br />

reaction<br />

8.2.1 Accelerated Test (ASTM C 441)<br />

In this method the expansion of mortar prepared with the studied cement or MIC and Pyrex<br />

glass used as aggregate, cured in water at 38 o C, during 16 days is measured.<br />

Expansions lower than 0.02% indicate that the cement or MIC is able to prevent the<br />

occurrence of alkali-aggregate reaction, even when using reactive aggregates.<br />

8.3 Evaluation of the potential reactivity of a certain cement-aggregate<br />

combination<br />

8.3.1 Accelerated method (ASTM C227)<br />

Mortar bars prepared with a certain cement-aggregate pair is stored in water at 37.8 o C, and<br />

eventual expansions after 3 and 6 months can be used as indicators of its reactivity.<br />

Although the limit between a reactive and non-reactive combination is not clearly defined,<br />

expansions higher than 0.05% after 3 months and 0.10% after 6 months, can be considered<br />

indicatives of potential reactivity.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 179 of 235


Cement<br />

Concrete<br />

Self-compacting Concrete<br />

1. DEFINITION .................................................................................................................... 181<br />

2. SELF-COMPACTING CONCRETE DEVELOPMENT .................................................... 181<br />

3. SELF-COMPACTING REQUIREMENTS........................................................................ 183<br />

4. MATERIALS FOR MANUFACTURING SELF-COMPACTING CONCRETE................. 184<br />

4.1....Powder 184<br />

4.2....Superplasticizers.................................................................................................... 184<br />

4.3....Viscosity agents ..................................................................................................... 185<br />

5. MIX PROPORTION DESIGN METHOD.......................................................................... 187<br />

5.1....JSCE Method ......................................................................................................... 187<br />

5.2....Okamura’s Method................................................................................................. 190<br />

6. TEST METHODS FOR SELF-COMPACTABILITY EVALUATION................................ 191<br />

6.1....Slump flow.............................................................................................................. 191<br />

6.2....Test Method for passability through spaces using U or Box-shaped<br />

apparatus ............................................................................................................... 192<br />

6.3....Flow-through test method using V-funnel .............................................................. 115<br />

6.4....L-Box test ............................................................................................................... 115<br />

6.5....Ring test ................................................................................................................. 116<br />

6.6....Acceptance test at jobsite ...................................................................................... 116<br />

7. PRODUCTION AND PLACEMENT OF SELF-COMPACTING CONCRETE ................. 117<br />

7.1....Storage of aggregate ............................................................................................. 117<br />

7.2....Plant facilities ......................................................................................................... 117<br />

7.3....Transportation ........................................................................................................ 118<br />

7.4....Placing 119<br />

7.5....Finishing and curing ............................................................................................... 119<br />

7.6....Formwork and supports ......................................................................................... 120<br />

8. QUALITY CONTROL AND INSPECTION DURING CONSTRUCTION ......................... 120<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 180 of 235


Cement<br />

Concrete<br />

1 Definition<br />

It is not possible to make only one definition of Self-Compacting Concrete. According to the<br />

most widely definition SCC is:<br />

High fluidity concrete with significant increase in fluidity that can be placed under its own<br />

weight and can fill the form work without vibration and achieve good consolidation without<br />

exhibiting any segregation nor bleeding. A SCC should present high deformability and great<br />

stability, without risk of blockage. See 3.3. [A1]<br />

2 Self-compacting concrete development<br />

The first prototype of SCC was developed by Ozawa et. al. [1988] by using granulated blast<br />

furnace slag and fly ash together with superplasticizer according to three states:<br />

‣ Fresh state: self-compactable with high resistance to segregation<br />

‣ Early age: avoidance from initial defects caused by heat generation of<br />

hydration, hardening or drying shrinkage and settlement<br />

‣ Hardened: protection against external factors (low permeability and frost<br />

resistance when required)<br />

Due to heat generation (limitation with the type and content of cement) and hardened<br />

concrete properties (low porosity) it was very difficult to achieve a flowable and at the same<br />

time stable concrete without the use of mineral components as well as chemical admixtures.<br />

Further development continues in Japan, especially in the research institute of large<br />

construction companies. As a result, Self-compacting has been applied into many practical<br />

structures over the last few years for the main following reasons:<br />

‣ To shorten construction period [Kashima et. al.1998]<br />

‣ To ensure compaction in confined zone which makes vibration difficult or<br />

impossible<br />

‣ To reduce noise due to vibrating compaction especially at pre-fabrication plants<br />

[Uno, 1998]<br />

The construction of the two anchorages of Akashi-Kaikyo bridge, one of the longest<br />

suspension (1’991 meters) in the world, is one of the typical applications where SCC was<br />

used. The use of SCC shortened the anchorage construction period by 20%, from 2.5 to 2<br />

years [Kashima et. al.1998].<br />

In another application, high strength (60 MPa) self-compacting concrete was used to<br />

fabricate Osaka Gas’ new prestressed concrete outer tank for liquefied natural gas storage<br />

(12’000 m 3 of concrete), which accommodates the world’s largest capacity of 180’000 kl. As<br />

a result, the construction period was reduced from 15 to 11 months. [Kitamura et. Al. 1998].<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 181 of 235


Cement<br />

Concrete<br />

Figure 1 – Application to anchorage of Akashi-Kaikyo<br />

Figure 2 – Application to LNG tank<br />

Nowadays, the interest for Self-compacting concrete has rapidly increased world-wide<br />

following the pioneering work done in Japan. We should note that outside Japan, that has<br />

successfully experienced this technology mainly in huge and sophisticated construction<br />

projects, self-compacting concrete is still in a development phase with very few experimental<br />

projects.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 182 of 235


Cement<br />

Concrete<br />

3 Self-compacting requirements<br />

How do requirements differ between conventional and self-compacting concrete? The<br />

essential difference in requirement concerns the fresh properties of SCC which must display<br />

a substantial flowability (filling capacity) between small gaps of reinforcements and achieve<br />

full compaction without vibration, while possessing high resistance to segregation during<br />

transportation, pumping and placing into formwork. When a shorter period of concreting is<br />

required, SCC must flow with a lower inclination even if the depositing speed is high.<br />

Apart from fresh concrete requirements all other requirements for conventional concrete such<br />

as curing must be fulfilled.<br />

Figure 3 – Essential requirements for self-compacting concrete<br />

The level of filling capacity required for SCC depends on the conditions not only of the<br />

structural detailing of formwork and reinforcements but also of the concreting practices such<br />

as speed and volume of concrete deposited. Self-compacting concrete with the same mix<br />

proportion and the same rheological properties may show the different filling capacities under<br />

the different conditions, therefore SCC should be designed to satisfy the level of filling<br />

capacity required according to the condition [Noguchi et. al. 1998].<br />

The following three ranks has been established as levels of self-compactability by the Japan<br />

Society of Civil Engineers [Concrete engineering series 31]:<br />

‣ Rank 1: Self-compactability into members or portions having complicated<br />

shapes and/or small cross-sectional areas with a minimum steel clearance in<br />

the range of 35 to 60 mm.<br />

‣ Rank 2: Self-compactability into reinforced concrete structures or members with<br />

a minimum steel clearance in the range of 60 to 200 mm. This normally<br />

corresponds to a steel content of 350 to 100 kg/m 3 .<br />

‣ Rank 3: Self-compactability into members or portions having large crosssectional<br />

areas and a small amount of reinforcement (less than 100 kg/m 3 ) with<br />

a minimum steel clearance of more than 200 mm.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 183 of 235


Cement<br />

Concrete<br />

4 Materials for manufacturing self-compacting concrete<br />

The self-compacting characteristic of concrete requires a high fluidity with sufficient<br />

segregation resistance. The two features, high fluidity and sufficient segregation resistance,<br />

are in general contrary to each other. In order to attain the both contradictory factors at the<br />

optimal level, a large quantity of powder is required with help of superplasticizer. It is not<br />

clear, however, what kind of powder is the most suitable and what of physical properties of<br />

powder control the amount of water. In any case a trial an error technique is in fact applied<br />

to find out a satisfactory mix proportion. Following are examples of materials for SCC for<br />

which sufficient field experience has been acquired in Japan.<br />

4.1 Powder<br />

Powder is a generic term for cement and other solid concrete materials having a fineness<br />

equal to or higher than cement. These include cement, ground granulated blast-furnace<br />

slag, fly ash, silica fume, and limestone powder.<br />

1) Belite-rich Portland cement: this is a type of Portland cement developed for<br />

concrete proportioned to have a high binder content. The belite content is 40 to<br />

70%, which is higher than moderate-heat cement. It is used where high<br />

strength and/or low heat are required.<br />

2) Binary/ternary system low-heat cement: this type of cement is a normal or<br />

moderate-heat Portland cement mixed with ground granulated blast-furnace<br />

slag and/or fly ash. It is used where low heat is required without reducing the<br />

powder content<br />

3) High-fineness ground granulated blast-furnace slag: it has been reported<br />

[Nawa et. al., 1998] that the thixotropicity of paste incorporating fine-fineness<br />

blast furnace slag powder (greater than 10'000 cm 2 /g) decreases compared to<br />

ordinary Portland cement.<br />

4) Fly ash: concrete flowability is enhanced by the “ball-bearing” effect of spherical<br />

form of fly ash.<br />

5) Limestone powder: consists primarily of CaCo 3 (calcite) pulverised to a specific<br />

surface area by Blaine of 2500 to 8000 cm 2 /g. It is used where desired low<br />

heat and/or strength be reduced without reducing the powder content.<br />

4.2 Superplasticizers<br />

Superplasticizers used in SCC are of two types. High range water reducing agent (HRWR)<br />

and high range AE water reducing agent (HRAE). As show in figure 4, currently used<br />

superplasticizers include naphthalene sulfonate, polycarboxylate, melamine sulfonate, and<br />

amino sulfonate-based agents. The type of superplasticizer to be used in SCC is decided<br />

considering the level of retention characteristics required and the existence of interaction with<br />

cement and admixture (viscosity agent).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 184 of 235


Cement<br />

Concrete<br />

Figure 4 – Classification of superplasticizers<br />

4.3 Viscosity agents<br />

The viscosity agents used in SCC may be broadly divided into four groups:<br />

1) Cellulose-based water-soluble polymers<br />

2) Acrylic-based water-soluble polymers bio-polymers<br />

3) Glycol-based water-soluble polymers<br />

4) Inorganic viscosity agents<br />

The standard addition quantity, characteristics, and interaction effects of viscosity agents<br />

vary depending on the type. For instance, cellulose-based water-soluble polymers and<br />

acrylic-based water-soluble polymers are used in viscosity-agent type SCC and combined<br />

type SCC. On the other hand, bio-polymers and inorganic viscosity agents are used only in<br />

the combined type SCC. In addition to the above, admixtures possessing both reducing<br />

action and viscosity action, have also been developed.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 185 of 235


Cement<br />

Concrete<br />

The action of mechanisms in viscosity agents that impart ingredient segregation resistance<br />

varies largely depending on the type. As shown figure 5, mechanisms in viscosity agents<br />

that impart ingredient segregation resistance may be broadly divided into two types [Nawa et.<br />

al.,1998]:<br />

Mechanisms that act in powder grains as cement<br />

1) Mechanisms that act in water in the concrete<br />

Figure 5 – Mechanisms imparting material segregation resistance<br />

Almost, all viscosity agents used in concrete have one or more functional groups in their<br />

molecular structures and are consequently adsorptive. In the case of adsorptive viscosity<br />

agents, the decline of the flow value is presumed to form the bridge structure by adsorbing<br />

on the surface of cement particles. As the dosage of the non-adsorptive viscosity agents<br />

increases, apparent plastic viscosity of mortar increases, yet the flow value does not change.<br />

Therefore, it becomes possible to control the viscosity keeping the flowability constant by<br />

using non-adsorptive agents. Also, non-adsorptive viscosity agents do not contest with<br />

superplasticizer for adsorption sites and the amount of adsorbed superplasticizer stays<br />

same. Therefore, non-adsorptive agents can lead to an adequate fluidity and viscosity of<br />

mortar. It is expected that the unique property of non-adsorptive viscosity agents is good for<br />

self-compacting concrete.<br />

Viscosity agents that act in cement grains include cellulose-based water-soluble polymers<br />

and acrylic-based water-soluble polymers. Glycol-based water-soluble polymer does not<br />

adsorb at all on to the surface of cement. From the bio-polymers, polysaccharide polymers,<br />

micro-organisms, and inorganic viscosity agents do not dissolve in water, but the polymers<br />

themselves adsorb the water, swell, and impart viscosity.<br />

Glycol type and water-soluble amide type, liquid viscosity agents are beginning to be used in<br />

place of conventional powder types, such as the cellulose type, acrylic type, polysaccharide<br />

polymers, biopolymers and inorganic viscosity agents.<br />

There is affinity between viscosity agents and superplasticizers. If the affinity becomes poor,<br />

a conspicuous slump loss or coagulation occurs, and the flowability of concrete degrades.<br />

This affinity is well known from experience, but the mechanism between them is still not<br />

clear. For instance, it has been reported that if cellulose-based water-soluble polymer, which<br />

has a poor affinity, is combined with naphthalene sulfonate, the interaction between the two<br />

becomes extremely weak because of the conversion of sodium naphthalene sulfonate to<br />

calcium naphthalene sulfonate.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 186 of 235


Cement<br />

Concrete<br />

5 Mix proportion design method<br />

Self-compacting concrete is not only a concrete but a whole family of concrete types exactly<br />

as the situation is today for conventional concrete. With SCC two normally incompatible<br />

properties of deformability and segregation resistance are both realised to achieve selfcompactability.<br />

A typical mix proportion design method for SCC consists of three stages:<br />

1. Proportioning condition examination: The performance requirements of<br />

concrete are selected on the basis of the structural, construction and<br />

environmental conditions, and the target performance requirements then are<br />

determined.<br />

2. Proportioning examination: The concrete materials are selected and a tentative<br />

mix proportion is determined.<br />

3. Proportioning verification: The initial proportion is checked by trial mixing if it<br />

satisfies the performance requirements. If it fails in any area, the initial mix<br />

proportion is modified and tested again.<br />

A wide variety of formulas are possible for self-compacting concrete satisfying the selfcompactability<br />

requirements as well as performance requirements. SCC is therefore,<br />

proportioned by selecting an adequate combination of materials in consideration of the<br />

restricting conditions at production plants and economical efficiency including their availability<br />

and transportation cost. Two major methods – one by the Japanese Society of Civil<br />

Engineers (JSCE) and another described by Prof. Okamura et. al. – are introduced here<br />

[Nawa et. al., 1998].<br />

5.1 JSCE Method<br />

Although various classification systems are possible for SCC, this method classifies it<br />

roughly into three categories, i.e., the powder type, viscosity agent type and combination<br />

type. The mix proportioning procedure is illustrated in figure 6.<br />

Figure 6 – JSCE mix design method for SCC<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 187 of 235


Cement<br />

Concrete<br />

5.1.1 Powder-type self-compacting concrete<br />

This type of SCC is proportioned to provide the required self-compactability not by using a<br />

viscosity agent but primarily by reducing the water/powder ratio (in effect increasing the<br />

powder content) to impart adequate segregation resistance and using air-entraining and<br />

high-range water-reducing admixture or superplasticizer to impart high deformability. The<br />

type of powder to be used should be adequately selected in consideration of the purpose of<br />

use of the concrete. This type of SCC has generally high strength and good durability.<br />

a) Maximum size of coarse aggregate and coarse aggregate content: the<br />

standard maximum size of coarse aggregate should be 20 mm or 25 mm. The<br />

content should be selected to provide the required self-compactability ranks are<br />

as follows:<br />

Self-compactability<br />

Unit absolute volume of coarse aggregate<br />

Rank 1 280 – 300 L/m 3<br />

Rank 2 300 – 330 L/m 3<br />

Rank 3 320 – 350 L/m 3<br />

b) Unit water content: It is desirable that the unit water content be<br />

minimised, as it affects the qualities of hardened concrete. It should<br />

normally be in the range of 155 to 175 kg/m 3 .<br />

c) Water/powder ratio: It should be in the range of 28% to 37% by mass of<br />

cement or 0.85 to 1.15 by volumetric ratio to cement, though dependent<br />

on the type and composition of powders.<br />

d) Powder content: It should normally be in the range of 160 to 190 L/m 3 ,<br />

though dependent on the type and composition of powders.<br />

e) Air content: The standard air content of fresh concrete requiring high<br />

resistance to frost damage should normally be 4.5%.<br />

f) Unit fine aggregate content: Should be determined from the unit coarse<br />

aggregate content, unit water content, unit powder content and air<br />

content.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 188 of 235


Cement<br />

Concrete<br />

5.1.2 Viscosity agent-type self-compacting concrete<br />

This is a type of SCC considered to be an extension of the anti-washout underwater<br />

concrete. Segregation resistance is increased by the use of viscosity agent with the powder<br />

content being low compared with the powder-type. The mixture proportions of such concrete<br />

are relatively similar to those of conventional concrete except that a viscosity agent and an<br />

air-entraining and high-range water-reducing admixture or superplasticizer are used. It may<br />

be manufactured even at the construction site by using truck agitators.<br />

a) Maximum size of coarse aggregate and coarse aggregate content: the<br />

standard maximum size of coarse aggregate should be 20 mm or 25 mm. The<br />

content should be selected to provide the required self-compactability ranks are<br />

as follows:<br />

Self-compactability<br />

Unit absolute volume of coarse aggregate<br />

Rank 1 280 – 310 L/m 3<br />

Rank 2 300 – 330 L/m 3<br />

Rank 3 300 – 360 L/m 3<br />

b) Unit water content: The unit water content should be the minimum<br />

required to attain the specified self-compactability, as it produces strong<br />

effects on the qualities of hardened concrete.<br />

c) Water/binder ratio and water/powder ratio:<br />

‣ It is recommended that the minimum water/binder ratio be selected<br />

from among those determined by the performance items required of<br />

the concrete, such as strength, durability, water-tightness and steelprotecting<br />

capacity.<br />

‣ When the required self-compactability cannot be obtained by the<br />

water/binder ratio determined in above, it is recommended that the<br />

water/binder ratio be reduced, by increasing the binder content or that<br />

the water/powder ratio be reduced, by using a non-binder powder.<br />

d) Air content: The standard air content of fresh concrete should generally<br />

be 4.5% of the volume of concrete. However, a value higher by<br />

approximately 1% point may be recommended for certain proportioning<br />

conditions.<br />

e) Chemical admixture dosage: The dosages must be determined in the<br />

ranges of not adversely affecting the qualities, such as the specified timerelated<br />

quality changes of fresh concrete, setting time and strength<br />

development.<br />

f) Unit fine aggregate content: Should be determined from the unit coarse<br />

aggregate content, unit water content, unit powder content and air<br />

content.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 189 of 235


Cement<br />

Concrete<br />

5.1.3 Combination-type self-compacting concrete<br />

This type of SCC has a reduced water/powder ratio and contains a viscosity agent. The<br />

difference from the viscosity agent-type is that the purpose of using a viscosity agent is to<br />

alleviate the fluctuation of fresh concrete properties due to the fluctuation of surface moisture<br />

content, aggregate grading, etc. The addition of small amount of specified viscosity agent to<br />

powder-based SCC minimises variation in flowability not only due to the changes in surface<br />

moisture ratio or grading of fine aggregate, but also the changes in concrete temperature.<br />

The mixture proportion of this type of SCC is similar to the powder-type except for the use of<br />

a viscosity agent.<br />

5.2 Okamura’s Method<br />

This method (Fig. 7) is based on the assumptions that moderate-heat Portland cement or<br />

belite-rich Portland cement is the only source of powdered materials. With this method,<br />

therefore, if the requirement for self-compactability is satisfied, the required performance of<br />

the hardened concrete is generally achieved. Considering the variability that arises during<br />

manufacturing, such as variations in material quality and weighing errors, the target selfcompactability<br />

is determined in advance at a higher level than required. The method is<br />

explained below.<br />

Figure 7 – Okamura’s mix design test method for SCC<br />

a) Air content: this should be determined at 4 – 7%, dependent on the<br />

environment to which the content structure will be exposed.<br />

b) Unit coarse aggregate content: the relative volumetric coarse aggregate ratio<br />

(G/Glim, a ratio of coarse aggregate volume to solid volume of coarse<br />

aggregate in concrete volume excluding air) should be 0.50. At levels higher<br />

than 0.50, self-compactability suddenly falls.<br />

c) Unit fine aggregate content: the volumetric fine aggregate ratio (V s /V m ,<br />

volumetric ratio of fine aggregate with particles larger than 90 µm to mortar<br />

volume excluding air) should be 0.40. This value is determined on the safe side<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 190 of 235


Cement<br />

Concrete<br />

one for self-compactability considering the variability in material quality level<br />

during manufacturing.<br />

d) Volumetric water/powder ratio: value at which the relative flow area (T m ) is 5<br />

and the relative funnel speed (R m ) is 1 in mortar tests (flow test and V-funnel<br />

test).<br />

e) Superplasticizer dosage: dosage at which the relative flow area (T m ) is 5 and<br />

the relative funnel speed (R m ) is 1 in mortar tests (flow test and V-funnel test).<br />

6 Test methods for self-compactability evaluation<br />

Relevant test methods are essential during mix design development and at the actual site<br />

delivery. Many different test methods for SCC are reported all over the world.<br />

6.1 Slump flow<br />

The slump flow is one of the most popular methods of evaluating the consistency of<br />

concrete, both in the laboratory and the construction sites due to its ease of operation. This<br />

is a conventional method widely used for evaluating the flowability of concrete §(Fig.8). The<br />

flowability is characterised by the final diameter and the T50-value. The final diameter, the<br />

slump-flow value, describes the yield stress and the T50-value describes the plastic<br />

viscosity. Normally a slump-flow of approximately 600 – 700 mm is found appropriate, with<br />

stability evaluated by observing coarse aggregate distribution at the very periphery of the<br />

concrete, while no segregation border of water, paste or fine mortar should be evident. Until<br />

today no recommendations for appropriate T50-values exist, a too low value always indicates<br />

too low plastic viscosity and thus a segregating mix [Billberg, 1999].<br />

Slump-flow being a very simple method is suitable both for plant or laboratory development<br />

as well as site delivery. The only property slump-flow does not measure is the blocking<br />

behaviour.<br />

Figure 8 – Slump-flow test<br />

Note: A mortar flow test (Fig. 9) is used to evaluate materials for mix proportioning.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 191 of 235


Cement<br />

Concrete<br />

Figure 9 – Mortar flow test<br />

6.2 Test Method for passability through spaces using U or Boxshaped<br />

apparatus<br />

The requirements for self-compactability can be different from one another depending on the<br />

structure to be compacted. From the point of view of rationalising quality control, Okamura<br />

and Ozawa proposed the so-called “Box-type test” (Fig. 10), modified version of U-shaped<br />

test (Fig. 11), whose bottom part is flat. Box-test is more sensitive to concrete with low<br />

segregation resistance and can therefore, detect such concrete easily.<br />

Concrete with the height of over 300 mm can be judged as self-compactable to practical<br />

structures. To stiff concrete cannot fill enough in the opposite tower due to its low<br />

deformability and too flowable concrete also cannot fill due to segregation between coarse<br />

aggregate and mortar in front of the obstacle.<br />

According to JSCE, when Rank 2 self-compactability is required, the concrete is judged as<br />

satisfying the requirement, if its filling height is not less than 300 mm by a filling tester with<br />

obstacle R2. When Rank 1 self-compactability is required, the concrete is judged as<br />

satisfying the requirement, if its filling height is not less than 300 mm using obstacle R1. The<br />

methods for concrete of Rank 3 include testing by a filling tester without any obstacle<br />

(obstacle R1 with five D10 deformed bars and obstacle R2 with three D13 deformed bars).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 192 of 235


Cement<br />

Concrete<br />

Figure 10 – Box-shaped test<br />

Figure 11 – U-shaped test<br />

6.3 Flow-through test method using V-funnel<br />

The viscosity of mortar and concrete are proposed to be evaluated through V-funnel tests<br />

(Figure 12). It is known that the apparent flow-through time is elongated after a certain rest<br />

period. Concrete with a low segregation resistance can lead to a large difference in flowthrough<br />

time between two tests. The degree of segregation can be judged by this test to a<br />

certain extent.<br />

Mortar<br />

Figure 12 – V-funnel tests<br />

Concrete<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 193 of 235


Cement<br />

Concrete<br />

6.4 L-Box test<br />

The L-box test provides information regarding flowability, blocking and segregation (Figure<br />

13). The following parameters are measured [A2]:<br />

a) The time for the front to reach 200 mm marking<br />

b) The time for the front to reach 200 mm marking<br />

c) When the concrete stops, the distances H1 and H2 are measured. Acceptable<br />

values of the so-called blocking ratio, H2/H1, can be >0.80.<br />

Figure 13 – L-box test [A2]<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 194 of 235


Cement<br />

Concrete<br />

6.5 Ring test<br />

The Ring test has a diameter of 300 mm with a fixed spacing of the vertical bars (Fig. 14).<br />

The blocking is assessed as a ratio of the average height of the concrete retained inside the<br />

ring at an approximately 200 mm diameter to the average height of the concrete outside of<br />

the ring [A3].<br />

Figure 14 – Ring test<br />

6.6 Acceptance test at jobsite<br />

Since the degree of compaction into the structure mainly depends on the self-compactability<br />

of concrete and poor self-compactability cannot be compensated by the construction work,<br />

self-compactability have to be checked for all concrete just before casting at the jobsite.<br />

However, conventional testing methods for self-compactability require sampling and then this<br />

can be laborious if acceptance test is to be carried out for all the concrete. Testing method<br />

was developed by [Ouchi et. al., 1996]:<br />

1) Testing apparatus is installed between agitator truck and pump at the jobsite.<br />

All the concrete is poured into the apparatus (Fig. 15)<br />

2) If the concrete flows through the apparatus, the concrete is considered as selfcompactable<br />

for the structure. If not the mix-proportion have to be adjusted.<br />

This apparatus was successfully used at the construction site of LNG tank of Osaka Gas<br />

and saved labour for the acceptance test (Fig. 16)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 195 of 235


Cement<br />

Concrete<br />

Figure 15 – Outline of acceptance testing apparatus at jobsite<br />

Figure 16 – Acceptance test at construction site<br />

7 Production and placement of self-compacting concrete<br />

There are some significant differences in production of SCC compared to conventional<br />

concrete. The most obvious is that the rheological properties of fresh SCC can be allowed to<br />

vary only to a very limited extent. For conventional concrete, any quality variation can be<br />

handled by varying the compaction work. Therefore, the production of SCC should be<br />

carried out in plants where the equipment, operation and materials are properly controlled.<br />

7.1 Storage of aggregate<br />

SCC is generally more susceptible to fluctuation of unit water content than normal concrete.<br />

For this reason, it is necessary to control fluctuation of the surface moisture ratio of<br />

aggregate, particularly fine aggregate. The surface moisture should be measured by an<br />

adequate method and its correction should be made accurately and promptly.<br />

Where the fluctuation of aggregate grading can have adverse effects on the fluctuation of<br />

SCC, aggregates with different grading should be stored separately. Thus, variation in<br />

aggregate grading is more difficult to handle during production.<br />

7.2 Plant facilities<br />

Both pan-mixers as well as free-fall mixers can be used to produce SCC (worn-down and old<br />

types of free-fall mixers are excluded). The type and status of mixer though has an impact<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 196 of 235


Cement<br />

Concrete<br />

on mixing efficiency, thus affecting mixing time. Therefore, the method of mixing SCC should<br />

be adequately established on the basis of field experience. The mixing time should be as a<br />

rule not less than 90 seconds.<br />

For instance, the mixing sequence is very important. SCC that should be frost resistance<br />

and thus have a controlled air-pore system, an optimum addition time of air entraining agent<br />

during mixing and a slightly prolonged mixing time is needed, but of course this also depends<br />

on mixer efficiency.<br />

7.3 Transportation<br />

SCC should be transported to the site within an adequate time in consideration of time<br />

required for placing so that placing can be completed while the required self-compactability is<br />

retained. The time limit from mixing to the end of placing should be as a rule 120 minutes.<br />

7.4 Placing<br />

The maximum distance of free drop should be as a rule not more than 5 meters. The<br />

maximum lateral flow distance should be as a rule 8 meters and must not exceed 15 meters.<br />

The point of depositing shall be shifted to avoid segregation due to lateral flow of concrete.<br />

When more than one layer is placed, the upper layer must be placed within the time during<br />

which the previous layer retains its fluidity, so that both layers can be monolithically<br />

integrated. If the second layer of SCC is placed on the existing layer which, has begun to<br />

set, cold joints or colour difference may occur (Fig. 17).<br />

Figure 17 - Colour difference on surface of SCC<br />

More or less perfectly continuous casting results in a completely defect free surface which is<br />

necessary to attain perfect concrete surfaces. In the case of architectural concrete, it is<br />

better to lower the free fall height to reduce generation of air bubbles on the surface of<br />

concrete (Figure 18). In vertical element a better finishing has been obtained when the<br />

concrete were placed from bottom to top.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 197 of 235


Cement<br />

Concrete<br />

Figure 18 – Air bubbles on the surface of SCC<br />

7.5 Finishing and curing<br />

In comparison with ordinary concrete, the setting time of SCC tends to be delayed according<br />

to the type of powder and superplasticizer and time of concrete placing. Therefore, care<br />

should be exercised when finishing the surface.<br />

The viscosity of SCC is high and bleeding is almost non-existent. When monolithic surface<br />

finish is required, measures to spray water, etc. must be taken during finish work. Since<br />

there is almost no bleeding in SCC, the surface tends to dry earlier than ordinary concrete,<br />

easily causing plastic cracks. Therefore, the top of concrete slabs, etc. must be covered or<br />

sprayed with water after placing so as to prevent drying.<br />

7.6 Formwork and supports<br />

a) Although some results show smaller lateral pressure than liquid pressure<br />

[Billberg, 1999 and Orsa Béton, 1999], the lateral pressure of concrete should<br />

as a rule be regarded as liquid pressure when designing formwork and<br />

supports. Quantitative calculation of lateral pressure based on the properties of<br />

fresh SCC, frame conditions, and placing conditions has not been made yet.<br />

When formwork is broken and SCC begins to flow out, it is almost impossible to<br />

stop it and a lot of labour is necessary to handle concrete that has flowed out.<br />

There will be no alternative but to suspend concrete placing scheduled for the<br />

day.<br />

b) In confined spaces, vent holes should be provided at adequate positions in the<br />

top forms.<br />

c) Formwork should be accurately constructed to prevent loss of concrete. When<br />

there is a gap in shuttering, paste or mortar in SCC flows out easily, and the<br />

outflow may continue for a long time (Fig. 19)<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 198 of 235


Cement<br />

Concrete<br />

d) Where the aesthetic appearance of concrete surface is of particular<br />

importance, the materials for sheathing and type of form remover should be<br />

suitably selected to minimise air-voids on formed surfaces.<br />

Figure 19 – Outflow of paste or mortar<br />

8 Quality control and inspection during construction<br />

The performance of produced and transported SCC to the construction site can differ from<br />

that designed and tested at the laboratory. It is, therefore, important to control before placing<br />

that produced and transported concrete attains the established self-compactability for<br />

ensuring reliability of the structure. Since the changes in the temperature and transportation<br />

time during the construction can alter the self-compactability from the designed level, it has<br />

to be confirmed by a suitable method by the time the placing begins.<br />

The deformability, segregation resistance and self-compactability should be controlled at the<br />

point of unloading taking into account the quality changes during conveyance from the<br />

unloading point to the placing point. Self-compactability may be inspected by passability<br />

tests with a filling tester (see section 2.5.2) used at the verification stage. The obstacle<br />

conditions corresponds to the established self-compactability rank should be used for testing.<br />

The inspection shall be carried out at least once for each 50 m 3 .<br />

Whether the concrete attains the required self-compactability at the point of unloading can<br />

also be judged by placing a passability apparatus (see section 2.5.6) between the agitator<br />

truck and the mobile pump, with which the passability of the whole truckload can be<br />

automatically tested.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 199 of 235


Cement<br />

Concrete<br />

Concrete Technology: Current Situation and Future Trends<br />

Pierre-Claude Aïtcin<br />

Université de Sherbrooke, Québec, Canada<br />

Moussa Baalbaki<br />

Holcim Group <strong>Support</strong> - CPD<br />

1. CONCRETE THE MOST WIDELY USED CONSTRUCTION MATERIAL ..................... 201<br />

2. THE DEVELOPMENT OF CONCRETE IN A SUSTAINABLE<br />

DEVELOPMENT PERSPECTIVE................................................................................... 202<br />

2.1 Sustainable development....................................................................................... 202<br />

2.2 Concrete technology status.................................................................................... 202<br />

2.3 Modern concrete technology is still ignored in several parts of the<br />

world....................................................................................................................... 203<br />

2.4 Life cycle of concrete structures............................................................................. 203<br />

3. NEW PROMISING TECHNOLOGIES IN THE DOMAIN OF CONCRETE ..................... 205<br />

3.1 Cheap concrete or expensive concrete.................................................................. 205<br />

3.2 Some recent technological break though in the domain of concrete<br />

materials................................................................................................................. 205<br />

4. HIGH PERFORMANCE CONCRETE ............................................................................. 207<br />

5. SELF LEVELING CONCRETE ....................................................................................... 208<br />

6. ROLLER COMPACTED HIGH PERFORMANCE CONCRETE ..................................... 209<br />

7. REACTIVE POWDER CONCRETE................................................................................ 210<br />

7.1 Increase of the homogeneity.................................................................................. 210<br />

7.2 Increasing of the compactness .............................................................................. 210<br />

7.3 Improvement of the microstructure by thermal treatment ...................................... 211<br />

7.4 Ductility of reactive powder concrete ..................................................................... 211<br />

8. THE CHALLENGES OF THE CEMENT AND CONCRETE INDUSTRY IN<br />

THE 21 ST CENTURY....................................................................................................... 212<br />

9. REFERENCES................................................................................................................ 214<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 200 of 235


Cement<br />

Concrete<br />

1 Concrete the most widely used construction material<br />

During the twentieth century, concrete became the most widely used construction material.<br />

Total concrete production in 1997 exceeded 7 billion cubic meters, which is just over 1 cubic<br />

meter or 3 tons per capita.<br />

How did concrete attain this status in 100 years? There is no single reason that accounts for<br />

its success, but rather a conjunction of facts :<br />

• concrete is not expensive;<br />

• concrete technology is simple;<br />

• concrete is primarily made from local materials;<br />

• concrete materials are widespread all over the globe;<br />

• concrete has high compressive strength;<br />

• concrete stays plastic long enough to be transported and placed;<br />

• concrete is a versatile construction material;<br />

• concrete is durable;<br />

• concrete is environmentally friendly.<br />

During the same period of time, most countries experienced general urbanization. As history<br />

teaches us, urbanization has consistently been the source of wealth since the times of the<br />

Sumerians and the Egyptians or the Incas and the Aztecs. It is not our intention to develop<br />

the socioeconomic reasons behind this fact; we only call attention to it and show that the use<br />

of cement and concrete has increased as the urbanization process has developed during the<br />

twentieth century.<br />

Any construction of streets, buildings, houses, schools, hospitals, water supply systems, and<br />

wastewater treatment plants in modern cities involves the use of concrete. The land, water,<br />

air, and even airwave links between cities needed to promote trade between urban areas<br />

and create the wealth of a nation or a continent have been built on concrete. We aren't<br />

overly optimistic or enthusiastic in stating that concrete is one of the foundations of wealth in<br />

modern civilization.<br />

While the urbanization process has plateaued in most industrialized nations, it will continue<br />

to develop during the first part of the 21 st century in many countries and result in an overall<br />

increase in the consumption of cement and concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 201 of 235


Cement<br />

Concrete<br />

2 The development of concrete in a sustainable<br />

development perspective<br />

2.1 Sustainable development<br />

Sustainable development is presently more than reality, it is a serious commitment towards<br />

the future. We must understand it and deal with it specially in relation to its globalization.<br />

The highest socioeconomic output at the lowest environmental impact must be searched for<br />

in any human activity. Compatibility between production, consumption, national and<br />

international trade and environmental issues must be permanently looked for. The linkage<br />

between all the parameters that may affect sustainability is one of the greatest challenges of<br />

our times; it is more than simply looking for decreasing industrial polluting emissions or<br />

energy saving, it is optimizing any activity so that its environmental impact will be the lowest<br />

possible.<br />

Sustainability is largely influenced by : economy and evergrowing social demands and<br />

welfare, energy availability such as fossil fuel, renewable resources, technological<br />

development, existing infrastructure, educational level, country size/population density,<br />

urbanization level, standard of living, new geopolitical and commercial barriers, among other<br />

parameters.<br />

In spite of the fact that concrete is well accepted to be for its own nature a material having a<br />

low energy content causing a low impact on the global environment, it is important to see<br />

what can be done to improve the situation in a sustainable development approach because<br />

great amounts of concrete are used every year all around the world.<br />

Production processes must be energy saving and should not generate harmful by-products<br />

or emissions in the production. Adaptation of a community to the principles of sustainable<br />

development leads to a need of a remarkable increase in the lifespan and quality of different<br />

products. Durability will be a key factor in the future from economical and environmental<br />

point of views in building as well as in road programs. Maintenance is expensive from any<br />

point of view.<br />

2.2 Concrete technology status<br />

The author is deeply convinced that presently :<br />

• concrete is not used at its best not only from a sustainable development point of view but<br />

also from an economical one;<br />

• putting into practice on a worldwide basis what is already a well established technology in<br />

some parts of the world would result in a more efficient use of cement and aggregates;<br />

• the impact of concree use on the environment is very limited when compared to other<br />

human activities.<br />

There are presently too many countries making too much concrete with a very low<br />

compressive strength of about 15 MPa to 20 MPa which results not only in a waste of<br />

cement but also in an increase of the rate of aggregate depletion in many urban areas.<br />

It is very easy to show that ecologically and economically the use of low strength concrete is<br />

equally harmful. A lot of engineers, contractors and owners are only concerned with the unit<br />

price of 1 m 3 of concrete when they should rather look at the cost of the Megapascal they<br />

need for structural purpose or at the life cycle cost of a concrete structure.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 202 of 235


Cement<br />

Concrete<br />

In order to illustrate this point let us consider the concrete columns of the structural<br />

laboratory of the University of Sherbrooke which have been built 30 years ago with a 25 MPa<br />

concrete, that should have contained about 300 kg of cement per cubic meter of concrete. In<br />

1998 these columns could have easily been built with a 75 MPa HPC containing 450 kg of<br />

cement per cubic meter of concrete, corresponding to a 50% increase in the cement dosage<br />

of the concrete. However, if we consider the amount of concrete per m 3 , the cross section<br />

area of the high strength columns would be 1/3 of that of the present columns. Therefore, 3<br />

times less HPC would be used for each column and the quantity of cement necessary to<br />

achieve the same structural performance would be half the one used 30 years ago (C × 1/3 ×<br />

1.5 = 0.5 C). This significant saving in cement is also concomitant with an even greater<br />

saving in the quantity of aggregates used to build these columns. Less than 1/3 of the<br />

aggregates used 30 years ago would be necessary to build in 1998 high performance<br />

concrete columns having the same structural performance.<br />

Moreover, if we look further in the future it is easy to realize that when the present 25 MPa<br />

concrete will be recycled, it could only be used as a road base aggregate, while a 75 MPa<br />

concrete could be used once or twice as a source of recycled aggregate to make concrete of<br />

a lower compressive strength before being used as a road aggregate.<br />

Of course the saving in cement and aggregate is not as significant in structural member<br />

made of high performance concrete working in flexure.<br />

2.3 Modern concrete technology is still ignored in several parts of the<br />

world<br />

First, for too many people concrete is still a mixture of cement, water and aggregates. The<br />

use of admixtures that improves so much the efficient use of cement in concrete is ignored<br />

so that large amounts of cement are presently wasted.<br />

Portland cement particles tendency to flocculate when mixed with water resulting in the<br />

addition of an unnecessary increase of mixing water and cement to achieve a workable<br />

mixture, can be overcome by the use of organic molecules known as dispersants, water<br />

reducers or superplasticizers. Unfortunately the use of water reducers is not the general rule<br />

so that at the dawn of the twenty first century it can be estimated that the same amount of<br />

concrete having the same compressive strength and the same durability could be made in a<br />

worldwide basis using about 10 percent less cement, which means that between 100 and<br />

200 million tonnes of cement are wasted by ignorance every year.<br />

Second, the selection of the type of concrete to be used to build a structure is too often made<br />

only by taking into account structural aspects without considering the environmental<br />

conditions to which this concrete will be exposed during its service life. The result is that too<br />

many concrete structures cannot perform their socioeconomic function for a very long period.<br />

The premature failure of 20 MPa to 30 MPa external parking garages in Canada is a good<br />

example of such an inadequate use of concrete. When an external parking garage has to be<br />

built the selection of concrete must not be done based only on structural criteria, but rather<br />

on durability criteria. A similar situation has been faced in Midle-East because the peculiar<br />

environmental conditions that concrete would have to face were not given enough<br />

consideration. Even in Europe too many concrete structures suffer a premature failure due<br />

to carbonation : the skin of a low strength concrete is more prone to carbonate than that of a<br />

higher strength concrete.<br />

2.4 Life cycle of concrete structures<br />

Increasing the life cycle of concrete structures is essential in a sustainable development<br />

approach but this analysis of the cycle life cast must include the evaluation of the impact in<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 203 of 235


Cement<br />

Concrete<br />

the environment of a system that includes the whole activities associated to a product or<br />

service from raw meal extraction until the waste elimination. This evaluation must be<br />

performed all over the duration of the infrastructure.<br />

Moreover, presently, the construction industry is too fractionated which is very unfortunate,<br />

and very few people have an integrated view of the whole construction process. The<br />

decisions taken by each partners are taken on an individual basis without taking into account<br />

the final cost of the project. Too often, in order to save a penny at one step of the<br />

construction a subcontractor will oblige one of the next subcontractors to spend a dollar to<br />

overcome a difficulty that could have been solved for much less than one dollar, if the<br />

solution of the problem had been more wisely shared between the two subcontractors. As<br />

an example let us consider the increase in the C 3 S content in modern Portland cement. This<br />

has been the response of the cement industry to the demand of contractors who wanted to<br />

speed up the construction process. Not only the 24-hours compressive strength of the<br />

concrete has been increased but also the 28-day one, but with the result that after 28 days<br />

such concretes do not increase their compressive strength very much while former concretes<br />

made with a cement having a lower C 3 S content were still developing a significant amount of<br />

strength after 28 days, years ago. Therefore, the durability rules established for these former<br />

concretes made with a low C 3 S cement are no more valid for concretes made with high C 3 S<br />

cements because the concrete is practically not gaining anymore strength after 28 days.<br />

Concrete durability does not depend on compressive strength but rather on the water/cement<br />

ratio which dictates the porosity of the hydrated cement paste and the progression rate of<br />

aggressive agents within concrete. On that respect a given 28-d compressive strength can<br />

be achieved with a modern cement at a much higher water/cement ratio than some years<br />

ago, which is not so good from the durability point of view.<br />

These are only a few examples of what can be done to improve the performance of concrete<br />

in a sustainable development approach.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 204 of 235


Cement<br />

Concrete<br />

3 New promising technologies in the domain of concrete<br />

3.1 Cheap concrete or expensive concrete<br />

At the dawn of the 21 st century there are some promising concrete technologies that are<br />

already confirmed or emerging in some countries that should be used increasingly in the first<br />

part of the 21 st century, but, given the limitated allocated time, it is my intention to limitate my<br />

presentation to four of them with which I am familiar and confident to see spreading all over<br />

the world.<br />

Before presenting these technologies I think that is important to analyze why in the last<br />

twenty years, concrete has found an increasing use in some niche markets instead of staying<br />

uniquely a commodity product. A few recent technological developments in the domain of<br />

admixtures and supplementary cementitious materials or filler, have resulted in the<br />

development of new types of efficient concretes that are already competing steel on some of<br />

its traditional construction markets.<br />

Concrete efficiency has to be considered in a broader view then simply the decrease of its<br />

unit price. The future of the concrete industry is not linked to the development of less<br />

expensive concretes but rather in the development of more expensive concretes, the use of<br />

such expensive concrete will result in a decrease of the final cost of concrete constructions,<br />

because the cost of concrete represents always only a small part of whole cost of a concrete<br />

structure, and, great savings can be realized in the other steps of the construction process, if<br />

concrete composition is optimized according to its final use. What is important is not the<br />

price of 1 m 3 of concrete but rather the cost of for example 1 MPa or of 1 more year in the life<br />

cycle of a concrete structure. As we will see later on in the three of the four new concrete<br />

technologies that will be presented the use of a concrete having a more expensive unit cost<br />

resulted in significant decrease of the initial cost of some concrete structures.<br />

It is hoped that a new philosophy in the use of concrete will emerge in the first part of the 21 st<br />

century as part of the growing tendency to deliver built key projects or BOOT (Build-Own-<br />

Operate-Transfer) infrastructures. For having been involved in a BOOT project, I can assure<br />

you that not only the working philosophy of the contractor is becoming more integrated but<br />

also that his decisions are taken in the perspective of a minimization of maintenance work.<br />

The philosophy of the lowest bidder has been and still is catastrophic in too many cases for<br />

the concrete industry, with the result that concrete present world wide a poor image. As long<br />

as maintenance and replacement costs are not integrated in tenders, concrete will be abused<br />

in the sake of making a fast and easy dollar. Our societies are not rich enough to base their<br />

development on such an attitude.<br />

3.2 Some recent technological break though in the domain of concrete<br />

materials<br />

In too many parts of the world, oncrete is still essentially a mixture of cement, water and<br />

aggregates but modern concrete always contains one or several admixtures. It is not wrong<br />

to say that most of the recent advances in concrete technology are rather due to an<br />

innovative use of admixtures rather than in an improvement of cement manufacturing.<br />

Portland cement is a wonderful building material, but from a physico chemical point of view it<br />

suffers from one intrinsic weakness : during its final grinding too many electric charges are<br />

developed at the surface of cement particles, with the result that, when concrete is mixed<br />

with water, cement particles have a strong tendency to flocculate. The poor dispersion of<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 205 of 235


Cement<br />

Concrete<br />

portland cement particles in water is definitely limiting the binding potential of portland<br />

cement, because, in order to improve concrete workability, non-stochiometric water has to be<br />

added in concrete in order to obtain a sufficient workability to be able to place concrete<br />

without having to develop too much vibration energy. This extra water succeeds in<br />

dispersing cement particles, but at the same time pull apart individual cement particles so<br />

that hydrates must bridge larger gaps to create binding links. The resulting hardened<br />

microstructure is therefore very porous and weak and prone to favor the ingress of<br />

aggressive agents within the structure.<br />

Fortunately, very few amounts (usually less than 1 percent of cement mass) of so-called<br />

water reducers or superplasticizer can be used to neutralize these electrical charges and<br />

increase the dispersibility of portland cement particles in water. Therefore, presently, it is<br />

possible to make flowing concrete having a slump of 200 to 250 mm and containing less<br />

water that the amount necessary to hydrate all the cement particles present in the mix.<br />

Moreover, the characteristicts of these polymeric materials have been optimized so that the<br />

retention of such a high slump during 1 or 2 hours does not cause any problems and does<br />

not result any more in a decrease of the 24-h strength due to some hydration retardation, as<br />

it was observed with the first generation of superplasticizers.<br />

A second technological break through in the domain of concrete technology, that occurred in<br />

the last twenty year, has been the use of silica fume, a by-product of the silicon and<br />

ferrosilicon industry. In fact, silica fume is composed of very fine spherical particles (one<br />

hundred time finer than a cement particle) of amorphous silica. Silica fume can be<br />

considered as a superpozzolan. In spite of the fact that limited amounts of silica fume are<br />

presently available, the discovery of the particular properties of this material has given a<br />

momentum to the research of the effects of the use of very fine powders in concrete. In that<br />

domain, the last word has not yet been said and in the first part of the 21 st century, it will not<br />

be surprising to see all kind of very fine powders being used when making concrete, rice<br />

husk ash and metakaolin can be such products.<br />

Finally, colloïdal agent, initially developed for underwater concreting, are finding new uses in<br />

self leveling concrete as it will be seen later.<br />

Due to time limitation it will not be possible to cover corrosion inhibitors and shrinkage<br />

reducing admixtures and some other admixtures that begin to be used.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 206 of 235


Cement<br />

Concrete<br />

4 High performance concrete<br />

High performance concrete should be rather called low water/cement ratio (W/C) or low<br />

water/binder ratio (W/B) concretes. In fact, the high performance W/C or W/B ratio is usually<br />

comprised, at present time, between 0.25 and 0.35, but, concretes having a lower W/C or<br />

W/B have been used. These low W/C or W/B ratios traduce in 28-d compressive strengths<br />

comprised between 50 and 100 MPa, but, concretes with a higher 28-d compressive strength<br />

have been made.<br />

The compressive strength of high performance concrete, among other factors, is a function of:<br />

• the WC or W/B, of the amount of entrained air it contains,<br />

• the efficiency of the selected cementitious combination used,<br />

• the "strength of the coarse aggregates", of the efficiency of the superplasticizer<br />

used among other factors.<br />

When making high performance concrete one of the most important factor is to find a cement<br />

or a cementitious combination which is compatible with one of the commercially available<br />

superplasticizer. A cement or cementitious combination is said to be compatible when a low<br />

W/C or W/B ratio concrete having an initial slump higher than 200 mm is able to maintain this<br />

slump during 1 hour or 1 hour and half and still have a high 1-d compressive strength.<br />

The compatibility between a cement and a superplasticizer is governed by quite complex<br />

physico-chemical parameters and is still more or less well understood from a scientifical point<br />

of view. However, several empirical methods can be used to find if a particular<br />

cement/superplasticizer is compatible or not. When a particular combination is found to be<br />

non compatible or poorly compatible several tricks can be used to try to improve the<br />

compatibility like : adding a slight amount of retarder, a slight amount of alkali sulfate, using<br />

the so-called double introduction technique.<br />

The use of silica fume is not mandatory to make high performance concrete, but when it is<br />

available at an economical price it is suggested to use it because it facilitates the fabrication<br />

of high performance concrete. Several cementitious combinations have been used to make<br />

high performance concrete depending on the local availability of supplementary cementitious<br />

materials.<br />

When making high performance concrete it is important to select a "strong" coarse aggregate<br />

because in some cases it has been found that the coarse aggregate was in fact limiting the<br />

compressive strength that could be achieved. Moreover, it is found that the elastic properties<br />

of the coarse aggregate are influencing very strongly the elastic properties of high<br />

performance concrete so that the usual empirical formulas used to predict the elastic<br />

modulus from the compressive strength should be used with great care.<br />

Generally speaking it is admitted that the durability of high performance concrete is greater<br />

than that of ordinary concrete. There is still one controversial point concerning the durability<br />

of high performance concrete, it is its freeze-thaw durability. Should high performance<br />

concrete contains entrained air like ordinary concrete to be freeze thaw resistant or not?<br />

Presently, in Canada, the rule is that high performance concrete must be air entrained to be<br />

freeze thaw resistant, however the usual spacing factor requirement for ordinary concrete<br />

has been relaxed by some institutions. For example, the department of transportation of the<br />

Province of Quebec accept that the spacing factor of high performance concrete be lower<br />

than 350 m and not lower than 230 m like for ordinary concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 207 of 235


Cement<br />

Concrete<br />

One major difference between the behavior of high performance concrete and ordinary<br />

concrete is their shrinkage pattern. High performance concrete develops a very rapid and<br />

strong autogenous shrinkag, while autogenous shrinkage is perfectly negligible in ordinary<br />

concrete. Therefore, in order to minimize the development of the potential autogenous<br />

shrinkage of a high performance concrete it is important and necessary to water-cured high<br />

performance concrete before hydration reaction begins, that is 3 to 6 hours after placing. If a<br />

high performance concrete is not water-cured as early as that, it can be badly cracked, which<br />

in turn will affect the durability of the concrete structure.<br />

Depending on the compressive strength that is loked for 1 m 3 of high performance concrete<br />

can be 30 to 100% more expensive than a 25 to 30 MPa concrete. But, when properly used<br />

high performance concrete can be used to build less expensive concrete structures because<br />

the savings in the amount of concrete used, in the form work, in the reinforcing steel, in the<br />

footings etc. result in a an initial cost decrease. If the maintenance savings, and the initial<br />

socio-economical costs are taken into account it is found that building with high performance<br />

concrete is building cheaper.<br />

High performance concrete has been used to build outstanding structures but also to solve<br />

day-to-day problems like the reconstruction of a MacDonald's restaurant entrance. High<br />

performance concretes will be used more and more used in the 21 st century because of their<br />

durability or their rapid hardening rather than their final strength.<br />

5 Self leveling concrete<br />

Self leveling concretes have been developed in Japan, but start to be used in several<br />

countries. Self leveling concrete is a type of expensive concrete which use can result in a<br />

significant saving of the initial cost of some specific structures where placing cost of ordinary<br />

concrete could be very high. This is particularly the case of highly reinforced concrete<br />

elements in seismic zones.<br />

But, it has been found that self leveling concretes can be used also for another reason : it is<br />

a noiseless concrete that don't need to be vibrated to be placed, therefore, in urban areas<br />

self leveling concrete can be placed any time without disturbing the neighborhood :<br />

particularly late in the evening or during the night or early in the morning. For example, in an<br />

urban area, a contractor can wait till 6 o'clock to close his form instead of rushing to close<br />

them at 4 o'clock, therefore, every day if beneficiate of 2 more working hours from his crew.<br />

The self leveling concrete could be placed after 6 o'clock without any disturbance for the<br />

neighborhood under the supervision of a foreman assisted by a helper.<br />

Self leveling concretes have a composition that is altered from that of an ordinary concrete,<br />

they are made with a smaller size coarse aggregate, they contain more sand, more<br />

cementitious material, more superplasticizer and a colloïdal agent to control their viscosity,<br />

their segregation and their bleeding.<br />

Self leveling concretes have been used to build some huge structures but also to repair small<br />

unaccessible structural element.<br />

Without any doubt there is a niche market for self leveling concrete in the 21 st century.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 208 of 235


Cement<br />

Concrete<br />

6 Roller compacted high performance concrete<br />

Roller compacted concrete is not per se a new type of concrete, roller compacted concretes<br />

have been extensively used for dam construction. When used to build a dam roller<br />

compacted concrete in a very lean mix which compressive strength is quite low, 15 to 20<br />

MPa, in order to develop the lowest amount of heat. Roller compacted concrete increases<br />

the construction rate of a dam and decreases significantly the amount of material to be used<br />

to build it because of the steeper slopes that can be used.<br />

This placing technique is now used in richer mixes that can achieve a 50 to 60 MPa 28d<br />

compressive strength. Roller compacted high performance concretes are dry mixes which<br />

workability is measured in term of so-called Vebe time. The placing is done using standard<br />

asphalt placing equipments, with the great difference for the working crew that it is a cold mix<br />

rather than a hot one that has to be placed.<br />

The batching of high performance roller compacted concrete can be done either in any<br />

concrete batching plant or in an asphalt plant which has been slightly modified. Instead of<br />

adding the hot liquid asphalt as a binder, portland cement is added.<br />

High performance roller compacted concrete are quite economical mixes, it has been<br />

possible to achieve a 50 MPa compressive strength with 290 kg of cement per m 3 .<br />

Moreover, recently it was found that high performance roller compacted concrete can be air<br />

entrained using a new type of air entraining agent.<br />

Roller compacted high performance concrete has already found a use in a certain number of<br />

industries to build unreinforced large working areas (metallurgical plant, pulp and paper mills,<br />

warehouses, lumber mills, saw mills, scrap yards etc). What makes roller compacted high<br />

performance concrete attractive for such applications is the simplicity of use the rapidily of<br />

the construction time frame, the high wear resistance of the concrete surface, and also<br />

economics if the life cycle of such a strongs lab is taken into account. Interestingly, most of<br />

the projects have involving the use of high performance roller compacted concrete been<br />

done by asphalt companies which have found a very interesting new market for their placing<br />

equipment.<br />

There is no need for joints, usually roller compacted high performance concrete is let free to<br />

crack, the presence of fine cracks being of no concern for the users. In the case of the<br />

Domtar paper mill near Sherbrooke 86 000 m 2 of two 150 mm layers of high performance<br />

roller compacted concrete were cast in less than 6 weeks (such a surface represents the<br />

surface of 16 adjacent soccer fields). Very fine transversal cracks were observed every 30 m<br />

and compressive strength measured on cores was found to be about 50 MPa.<br />

It is sure that this placing technique will be widely used during the 21 st century in all kind of<br />

industrial applications. For example, a roller compacted high performance concrete was sold<br />

to increase the cleanliness of a log yard in order to increase the afternoon output of a saw<br />

mill because the logs could be washed and be less sandy before being sawed. As the wear<br />

of the saws was significantly decreased, the afternoon output was greatly improved, every<br />

evening the sharpening of the saws took less time and the saws lasted longer.<br />

It will be the rôle of the marketing people to find markets for high performance roller<br />

compacted concrete for economical reasons that very often will be very indirectly related to<br />

the traditional qualities of concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 209 of 235


Cement<br />

Concrete<br />

7 Reactive powder concrete<br />

Reactive powder concrete is a new type of cement based material. Reactive powder<br />

concrete does not contain any coarse aggregate, the maximum size of particles used to<br />

make it is usually comprised between 300 and 600 m. Therefore, some people argue that<br />

reactive powder concrete should not be called concrete since from a particle size point of<br />

view it is rather a mortar. However, when a reactive powder is reinforced with long and fine<br />

steel fibers it is easy to demonstrate, when considering the scale effect, that these fibres are<br />

acting in the reactive powder concrete as reinforcing bars in an ordinary concrete.<br />

Three principles are at the base of the development of reactive powder concrete :<br />

• an increase of the homogeneity of the material by eliminating the coarse aggregate,<br />

limiting the sand content, improving the mechanical properties of the paste and<br />

suppressing the transition zone around the aggregate;<br />

• an increase of the compactness of the mix by optimizing the grain size distribution of the<br />

different constituents and by pressing the mix before and during the setting (optional);<br />

• a refinement of the microstructure by a special heat treatment.<br />

7.1 Increase of the homogeneity<br />

The elimination of the coarse aggregate and its replacement by a fine sand tends to reduce<br />

considerably the size of the microfissures of mechanical, thermal and chemical origins that<br />

are linked, in an ordinary concrete to the presence of an inclusion (aggregate) in a matrix that<br />

is homogeneous (the hydrated cement paste). This reduction of the size of the<br />

micrrofissures is accompanied by an improvement of the properties of the RPC on their<br />

compressive and tensile strength. The sand content is reduced in order to pass from a rigid<br />

skeleton of aggregates that are adjacent and which present a large area in contact which<br />

each other, to a set of inclusions within a continuous matrix. This modification of the<br />

organization of the material results in a reduction of the negative effects on the porosity that<br />

can result from a partially restrained shrinkage when the aggregate skeleton is too stiff.<br />

The increase of the mechanical properties of the paste, more particularly of its elastic<br />

modulus, and, at the same time the reduction of the gap between the rigidity of the paste and<br />

that of the aggregates tend to reduce the effects related to the modification of the stress field<br />

around the aggregates.<br />

The increase of the compactness of the aggregate skeleton by the use of constituents having<br />

complementary grand size distribution and by the use of a high dosage of a very fine material<br />

like silica fume, result in the elimination of the transition zone between the paste and the<br />

sand particles increasing the transfer of the mechanical properties between the aggregates<br />

and the paste.<br />

7.2 Increasing of the compactness<br />

The increase of the compactness of an aggregate mix tend to minimize the voids between<br />

the particles and to reduce the amount of mixing water necessary to achieve a certain<br />

degree of workability. A theoretical approach associated with laboratory experiments has<br />

been used to select the best commercial components of the RPC and to fix their mixing<br />

proportions in order to increase its final compactness. Table 1 gives a typical composition of<br />

a RPC.<br />

The compactness of a reactive powder concrete can be increased even more if during<br />

setting a load is applied on the fresh concrete. The application of such a load results in the<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 210 of 235


Cement<br />

Concrete<br />

elimination of entrapped air that can developed during mixing, in the extraction of a part of<br />

the mixing water, in reducing the amount of water occupied in the fresh reactive powder<br />

concrete, in reducing the water/binder ratio and finally in compensating for a part of the<br />

chemical shrinkage. The pressing of a fresh reactive powder concrete can result in a gain of<br />

more than 5% on the relative density.<br />

7.3 Improvement of the microstructure by thermal treatment<br />

The heating in water of reactive powder concrete at a temperature comprised between 70<br />

and 90°C after its setting results in an acceleration of the pozzolanic reaction of the silica<br />

fume while modifying the structure of the hydrates already developed. This heat treatment<br />

results in a mechanical strength that is much higher than that of untreated samples. A heat<br />

treatment at 400°C can even alter further the microstructure of the reactive powder concrete<br />

and result in compressive strength gains that are much more significant (over 500 MPa, and<br />

even over 800 MPa have been obtained when doing this). In this last case RPC was<br />

confined with a pressure of 50 MPa and contained metallic aggregates. This type of curing is<br />

more difficult to realize, particularly for large structural elements.<br />

The curing in water between 90 to 100°C over a period of 48 hours is quite easy to do, this is<br />

the type of curing that has been used during the experimental works done at the Université of<br />

Sherbrooke. Reactive powder concretes having a 200-MPa compressive strength are<br />

fabricated on a daily basis in this way by technicians and graduate students. However, such<br />

reactive powder concrete is not a ductile material. The improvement of the ductility of this<br />

type of concrete is essential in order to make its use more interesting.<br />

7.4 Ductility of reactive powder concrete<br />

Reactive powder concrete matrices have a purely linear elastic behavior and their rupture is<br />

brittle. Their ductility is greatly improved by the addition of fibers or by their confinement in a<br />

steel tube.<br />

As it can be seen in the Figure 1, the addition of 140 kg/m 3 (1.8% volumetric fraction) of steel<br />

fibers having a length of 12 mm gives to reactive powder concrete some ductility in<br />

compression. The positive effect of the fibers on the ductility is more evident in the case of<br />

the flexural strength as it can be seen in Figure 2. The flexural behavior of a reactive powder<br />

concrete containing steel fibers is similar to that of reinforced concrete. In fact, adding 12mm<br />

long fibers having a diameter of 0.2 mm in a material where the maximum size of the<br />

coarsest sand grain is about 200 m correspond at ordinary concrete scale to add<br />

reinforcing 20 mm bars 1.2m long in a concrete made with a coarse aggregate having a<br />

maximum size of 20 mm<br />

It is known for a long time that the confinement of concrete in a steel tube result in an<br />

increase of the ultimate strength and of the ductility of the concrete when it is submitted to a<br />

compressive load. This strength increase is very interesting when the basic compressive<br />

strength of concrete is high. In the case of reactive powder concrete the improvement<br />

brought by the confinement is even more significant as it can be seen in Figure 1. In this<br />

figure, it is seen that according to the thickness of the steel tube, the fiber dosage and even<br />

the pressure applied on the concrete during its setting it is possible to obtain compressive<br />

strength varying from 250 to more than 350 MPa, but, what is more impressive is the ductility<br />

obtained when confining the RPC. During a study done at the Université de Sherbrooke on<br />

100-mm diameter specimens, the most interesting results were obtained with a reactive<br />

powder concrete containing steel fibers pressed during the setting and confined in a steel<br />

3mm thick tube. The ultimate compressive strength that was measured on such specimens<br />

was over 350 MPa.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 211 of 235


Cement<br />

Concrete<br />

Reactive powder concretes have been used, up to now, in a limited number of structural<br />

application where its high strength, its light weight high ductility and great durability and low<br />

maintenance cost can be used advantageously. Reactive powder concrete can be perceived<br />

as a very expensive concrete when considering its unit price per m 3 , which is comprised<br />

between 750 to 1000$ <strong>CD</strong>N, but, as reactive powder concrete in most its present<br />

applications compete with steel due to its very high compressive strength and pseudo plastic<br />

behaviour its price should be considered by the tonne rather than by the m 3 . The price of<br />

reactive concrete relate to its mass is 300 to 400 $ <strong>CD</strong>N/tonne in comparison to 1 000 to 1<br />

500 $<strong>CD</strong>N for structural steel.<br />

For sure reactive powder concrete market will always be a very specialized and lucrative<br />

markets. In the author's opinion the pig iron market should be considered as a potential<br />

market for reactive powder concrete, because as Pierre Richard, the inventor of the reactive<br />

powder concrete concept, says reactive powder concrete can be seen as a pseudo plastic<br />

cold ceramic.<br />

These four examples of new concrete technologies that were presented are only a few<br />

examples of successful uses or of new promising technologies developed recently in the<br />

domain of concrete. Another author could have selected other examples to show that<br />

concrete is a versatile material that will continue to be widely used in the 21 st century, in its<br />

traditional markets, but also, in unforeseen new markets. But, it is also sure that successful<br />

concrete technologies of the 21 st century will have to integrate a sustainable development<br />

perspecttive.<br />

8 The challenges of the cement and concrete industry in<br />

the 21 st century<br />

The main challenge of the cement and concrete industry in the 21 st century is to obtain the<br />

highest socioeconomic out put of each cubic meter of concrete at the lowest environmental<br />

cost.<br />

The author is deeply convinced that education is one of the biggest challenge of the concrete<br />

industry in the 21 st century. For to long concrete has been perceived as a low tech<br />

commodity product, and, competition based on price. This attitude was encouraged by the<br />

quite universal lowest bidder philosophy and has definitely tarnish the image of concrete in<br />

the public.<br />

Moreover by tradition in most parts of the world concrete production is fragmented in<br />

numerous small organizations operating locally in an artisanal mode rather than as a modern<br />

and efficient industry. Concrete making is still considered as a low tech activity that does not<br />

demand any particular technical knowledge, so that anyone who has some entrepreneurship<br />

can be in business with a small investment. Therefore, it will be difficult to have a rapid and<br />

significant impact on the concrete industry development.<br />

Therefore, the real chalenge is not what has to be done, it is rather how to do it and how fast<br />

it can be done.<br />

It will be a pity if the development of the concrete industry in developing countries during the<br />

first half of the 21 st century does not take advantage of the present state of the art technology<br />

in cement and concrete manufacturing developed in industrialized countries but rather<br />

repeats the errors of the past.<br />

However, it is not a hopeless situation, because the technology to obtain the most efficient<br />

technological and environmental output of each cubic meter of concrete exists already; it is<br />

only a matter of education to put it into application. When considering the existing<br />

communication means that are presently available to develop a worldwide educational<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 212 of 235


Cement<br />

Concrete<br />

system, it is time to take advantage of this situation to develop a worldwide concrete industry<br />

that will direct its efforts to improve its performance in a sustainable development approach.<br />

Following a significant research effort concrete image has been revamped and now in some<br />

applications concrete is directly competing steel : high rise buildings, offfshore platforms,<br />

large bridges, etc.<br />

Well advanced concrete technologies is already available, the challenge is to make these<br />

technologies available all over the world.<br />

Efficient use of newly developed admixtures should be strongly and rapidly promoted in order<br />

to be able to make more durable concrete out of the same amount of cement.<br />

The use of supplementary cementitious materials, natural pozzolans, artificial pozzolans and<br />

fillers should be promoted rapidly in order also too to make more concrete out of the same<br />

amount of cement.<br />

Increasing the life cycle of concrete structure is another goal that must be rapidly<br />

implemented, concrete structure should be designed based on durability criteria and not<br />

uniquely on structural criteria.<br />

There is still a great future for concrete as a construction material in the commodity market in<br />

the 21 st century but it is sure that a great number of niche markets will erode gradually the<br />

usual commodity market. Concrete in the 21 st century will be a sophisticated mixture of<br />

minerals and organic polymers which performances will atonish us.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 213 of 235


Cement<br />

Concrete<br />

9 References<br />

1. Aïtcin, P.-C., “Concrete the Most Widely Used Construction Material”, Adam Neville<br />

Symposium on Concrete Technology, Edited by V.M. Malhotra, Second CANMET/ACI<br />

International Symposium on Concrete Technology, Las Vegas, June 12, 1995, pp. 257 –<br />

266.<br />

2. Aïtcin, P.-C., “Cement and Concrete Development from an Environmental Perspective”,<br />

International Workshop on Concrete Technology for a Sustainable Development in the<br />

21 st Century, Solvacr, Lofoten Island, Norway, June 1998, to be published by E und FN<br />

SPON in 1999.<br />

3. Aïtcin, P.-C., Marciano, E., “The Future of the Cement and Concrete Industry According<br />

to a Sustainable Development Approach”, Wabe International Symposium on Cement<br />

and Concrete, Montréal, Canada, May 1998, 10 p.<br />

4. Aïtcin, P.-C., Baalbaki, M., “Chemical Admixtures. Key Components for Durable<br />

Concretes, Wabe International Symposium on Cement and Concrete, Montréal, Canada,<br />

May 1998, 9 p.<br />

5. Aïtcin, P.-C., Jolicoeur, C., MacGregor, J.G., “Superplasticizers: How They Work and<br />

Why They Occasionally Don’t”, Concrete International, Vol. 16, No. 5, May 1994, pp. 45 –<br />

52 (available in Spanish and Portuguese).<br />

6. Aïtcin, P.-C., “Durable Concrete – Cement Practice and Future Trends”, ACI-SP 144,<br />

1994, pp. 85 – 104.<br />

7. Aïtcin, P.-C., “From Gigapascals to Nanometres”, Engineering Foundation Conference,<br />

Potosi, M.D., USA, August 1988, 28 p.<br />

8. Aïtcin, P.-C., “Condensed Silica Fume”, Edited by P.-C. Aïtcin, Les Editions des<br />

l’Université de Sherbrooke, ISBN 2-7622-0016-4, 1983, 52 p.<br />

9. Khayat, K.H., Aïtcin, P.-C., “Silica Fume in Concrete” – An Overview”, ACI SP-132, Vol.<br />

2, 1992, pp. 835 – 872.<br />

10. Aïtcin, P.-C., Neville, A.M., “High-performance Concrete Demystified”, Concrete<br />

International, Vol. 15, No. 1, Jan. 1993, pp. 21 – 26.<br />

11. Aïtcin, P.-C., “The Art and The Science of High-performance Concrete”, Mario Collepardi<br />

Symposium and Advances in Concrete and Science and Technology, Edited P.K. Mehta,<br />

Rome 1997, pp. 107 – 126 (Translation in French, Portuguese and Italian are available).<br />

12. Aïtcin, P.-C., “High-performance Concrete”, E and FN SPON, London, U.K., 1998, 591 p.<br />

(Translation in French and Portuguese in progress).<br />

13. Aïtcin, P.-C., “The Influence of the Spacing Factor on the Freeze-Thaw Durability of<br />

HPC”, Sherbrooke 98 Symposium on HPC and RPC, Sherbrooke, August 1998, Vol. 4,<br />

pp. 419 – 434.<br />

14. Aïtcin, P.-C., Pigeon, M., Pleau, R., Gagné, R., “Freezing and Thawing Durability of Highperformance<br />

Concrete”, Sherbrooke 98, Symposium on HPC and RPC, Sherbrooke,<br />

August 1998, Vol. 4, pp. 383 – 392.<br />

15. Aïtcin, P.-C., “Non-shrinking Concrete”, CANMET/ACI/JCI, 4 th International Conference<br />

on Recent Advances in Concrete Technology, Supplementary papers, Tokushima,<br />

Japan, June 1998, pp. 215 – 226.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 214 of 235


Cement<br />

Concrete<br />

16. Aïtcin, P.-C., “Autogenous Shrinkage Measurment”, Autoshrink ’98, International<br />

Workshop on Autogenous Shrinkage of Concrete, Hiroshima, Japan, June 1998, pp. 245<br />

– 256.<br />

17. Lessard, M., Dallaire, E., Blouin, D., Aïtcin, P.-C., “High-performance Concrete Speeds<br />

Reconstruction for McDonald’s”, Concrete International, Vol. 16, No. 9, Sept. 1994, pp.<br />

47 – 50.<br />

18. Khayat, K.H., Aïtcin, P.-C., “Use of Self Compacting Concrete in Canada – Present<br />

Situation and Perspectives”, International Conference on SCC, Kochi, Japan, August<br />

1998, 10 p.<br />

19. Richard, P., Cheyrezy, M.H., “Reactive Powder Concrete with High Ductility and 200-800<br />

MPa Compressive Strength”, ACI SP-144, pp. 507 – 518.<br />

20. Bonneau, O., Poulin, C., Dugat, J., Richard, P., Aïtcin, P.-C., “Reactive Powder<br />

Concretes: From Theory to Practice”, Concrete International, Vol. 18, No. 4, pp. 47 – 49.<br />

21. Bonneau, O., Lachemi, M., Dallaire, E., Dugat, J., Aïtcin, P.-C., “Mechanical Properties<br />

and Durability of Two Industrial Reactive Powder Concretes”, ACI Materials Journal, Vol.<br />

94, No. 4, 1997, pp. 286 – 290.<br />

Table 1. Typical composition of a RPC mixture<br />

Materials kg/m 3<br />

Cement 705<br />

Silica fume 230<br />

Crush quartz 210<br />

Sand 1010<br />

Superplasticizer 17<br />

Steel fibers 140<br />

Water 185<br />

Water/binder 0.26<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 215 of 235


Cement<br />

Concrete<br />

The Art and Science of Durable High-Performance Concrete<br />

P.-C. Aïtcin<br />

Synopsis: High-performance concrete can be defined as a low water/binder ratio concrete<br />

with an optimized aggregate binder ratio to control its dimensional stability. Highperformance<br />

concrete acts as a true composite material in which coarse aggregates<br />

contribute to the mechanical properties, therefore, the relationship between elastic modulus<br />

and strength found in most codes no longer have predictive values for high-performance<br />

concrete.<br />

Achieving high performance is not done with a casual approach, all ingredients must be<br />

carefully selected. It is found that quite often the most difficult thing is to control concrete<br />

rheology long enough to place a 200-230 mm slump high-performance concrete one hour or<br />

more after mixing due to compatibility problems between the cement and the<br />

superplasticizer.<br />

High-performance concretes are very sensitive to plastic and autogenous shrinkage, so that<br />

their use demands an immediate water curing. The use of a curing compound which is<br />

perfectly adequate to cure a concrete having W/B ratio greater than 0.50 is absolutely<br />

inadequate with high-performance concrete because it does not prevent the development of<br />

autogenous shrinkage. When a 0.30 high-performance concrete is not water cured before<br />

setting, it can develop a 200 to 300 microstrains autogenous shrinkage during the first 24<br />

hours, that will be added to its drying shrinkage. On the contrary when such a highperformance<br />

concrete is water cured during the first 24 hours, its swells slightly.<br />

High-performance concrete is definitely more durable than usual concrete. Its increased use<br />

will be more often linked to its durability than its high strength. Durability will become a key<br />

issue because we will become more and more concerned with sustainable development. In<br />

that respect the use of high-performance concrete is more ecological than the use of a usual<br />

concrete: less cement and less aggregates are needed to sustain a certain load, the life<br />

cycle of the concrete structure is increased due to the greater intrinsic durability of highperformance<br />

concrete and, when high-performance concrete will have to be recycled at the<br />

end of its life, it will be recycled one or two times more than usual concrete because of its<br />

higher strength.<br />

Keywords:<br />

Autogenous shrinkage, durability, fire resistance, heat of hydration, high-performance<br />

concrete, plastic shrinkage, recycling, sustanable development, water curing, water/binder<br />

ratio.<br />

P.-C. Aïtcin, Full Professor, Université de Sherbrooke, Sherbrooke (QC) Canada<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 216 of 235


Cement<br />

Concrete<br />

1.0 INTRODUCTION<br />

The recent developments in the field of high-performance concrete (HPC) have been a giant<br />

step in making concrete a high-tech material with enhanced characteristics and durability.<br />

They have even led to it being a more ecological material in the sense that the component<br />

admixtures, aggregates, and water are fully used to produce a material with a longer life<br />

cycle. Be that as it may, we know that concrete will never be a durable material when<br />

measured against a geological time frame. Any concrete, if we look far enough into the<br />

future, will end its life cycle as limestone, clay, and silica sand, which are the most stable<br />

mineral forms of calcium, silica, iron, and aluminium in the earth's environment. Therefore,<br />

all we can do is to extend the life cycle of this artificial rock as much as possible.<br />

2.0 WHAT IS HIGH-PERFORMANCE CONCRETE?<br />

The concrete that was known as high-strength concrete in the late 1970s is now referred to<br />

as high-performance concrete because it has been found to be much more than simply<br />

stronger: it displays enhanced performance in such areas as durability and abrasion<br />

resistance. Although widely used, the expression "high-performance concrete" is very often<br />

criticised as being too vague, even as having no meaning at all. And what's more, there is<br />

no simple test for measuring the performance of concrete.<br />

High-performance concrete can be defined as an engineered concrete in which one or more<br />

specific characteristics have been enhanced through the selection and proportioning of its<br />

constituents. This definition is admittedly vague, but it has the advantage of indicating that<br />

there is no one single type of high-performance concrete, but rather a family of new types of<br />

high-tech concrete whose properties can be tailored to specific industrial needs.<br />

While we could devise a slightly more technically rigorous definition of high-performance<br />

concrete that would still be quite simple, such as it being a concrete with a low w/c (or rather<br />

a low water/binder ratio), in the 0.30 to 0.40 range. But this would still be inexact because<br />

high-performance concretes with a water/binder ratio in the 0.20 to 0.30 range have been<br />

used.<br />

This definition can be technically refined by stating that a high-performance concrete is a<br />

concrete in which autogenous shrinkage can develop due to a phenomenon called self<br />

desiccation when the concrete is not water cured. This technical jargon, however, does little<br />

to clarify things because very few people are familiar with the terms self-desiccation and<br />

autogenous shrinkage.<br />

Since there is no single best definition for the material that is called high-performance<br />

concrete, I prefer to define it as a low water/binder concrete with an optimised<br />

aggregate/binder ratio to control its dimensional stability and which receives an adequate<br />

water curing.<br />

3.0 WATER/CEMENT OR WATER/BINDER RATIO?<br />

Both expressions were deliberately used above, either singly or together, to reflect the fact<br />

that the cementitious component of high-performance concrete can be cement alone or any<br />

combination of cement with so-called mineral components, such as: slag, fly ash, silica fume,<br />

metakaolin, rice husk ash, and fillers such as limestone. Ternary systems are increasingly<br />

used to take advantage of the synergy of some mineral components to improve concrete<br />

properties in the fresh and hardened states and to make high performance concrete more<br />

economical.<br />

In spite of the fact that most high-performance concrete mixtures contain at least one mineral<br />

component, which should favour the use of the more general expression water/binder ratio,<br />

the water/binder and water/cement ratios should be used alongside one another. This is<br />

because most of the mineral components that go into high-performance concrete are not as<br />

reactive as Portland cement, which means that most of the early properties of high-<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 217 of 235


Cement<br />

Concrete<br />

performance concrete can be linked to its water/cement ratio while its long-term properties<br />

are rather linked to its water/binder ratio.<br />

It must be emphasised that the development of high-performance concrete technology has<br />

taught us what Feret expressed in his original formula giving the compressive strength of a<br />

concrete mixture: concrete compressive strength is closely related to the compacity of the<br />

hardened matrix. High-performance concrete has also taught us that the coarse aggregate<br />

can be the weakest link in concrete when the strength of hydrated cement paste is drastically<br />

increased by lowering the water/binder ratio. In such cases, concrete failure can start to<br />

develop within the coarse aggregate itself. As a consequence, there can be exceptions to<br />

the water/binder ratio law when dealing with high-performance concrete. In some areas,<br />

decreasing the water/binder ratio below a certain level is not practical because the strength<br />

of the high-performance concrete will not significantly exceed the aggregate's compressive<br />

strength. When the compressive strength is limited by the coarse aggregate, the only way to<br />

get higher strength is to use a stronger aggregate.<br />

4.0 A COMPOSITE MATERIAL<br />

Standard concrete can be characterised solely by its compressive strength because that can<br />

be directly linked to the water/cement ratio, which still is the best indicator of paste porosity.<br />

Most of the useful mechanical characteristics of concrete can be linked to compressive<br />

strength with simple empirical formulas. This is the case with elastic modulus and the<br />

modulus of rupture (flexural strength), because the hydrated cement paste and the transition<br />

zone around coarse-aggregate particles constitute the weakest links in concrete. The<br />

aggregate component (especially the coarse aggregate) contributes little to the mechanical<br />

properties of ordinary concrete. As the strength of the hydrated cement paste increases in<br />

high-performance concrete, the transition zone between the coarse aggregate and the<br />

hydrated cement paste practically disappears (Figures 1 and 2). Since there is proper stress<br />

transfer under these conditions, high-performance concrete behaves like a true composite<br />

material, as shown by Baalbaki et al. [1]. Stress-strain curves of high-performance concrete<br />

are influenced by the stress-strain curve of the coarse aggregate, as seen in Figure 3.<br />

To express this in another way, the elastic modulus (i.e., the rigidity) of a high-performance<br />

concrete can be tailored to the specific need of the designer simply by selecting the<br />

appropriate coarse aggregate. Nilsen and Aïtcin [2] have been able to make 100-MPa<br />

concrete with elastic moduli varying from 27 to 60 MPa. Therefore, the relationships<br />

between modulus and strength found in most codes have no predictive value with respect to<br />

high-performance concrete. Recently, however, Baalbaki [3] proposed two simple models<br />

that take into account the coarse aggregate's elastic characteristics so that the elastic<br />

modulus of any type of concrete can be calculated. One of the Baalbaki models is presented<br />

in Figure 4 and the predictive value of his model is presented in Figure 5.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 218 of 235


Cement<br />

Concrete<br />

(a)<br />

Oriented CH<br />

crystals<br />

Gap<br />

(b)<br />

CH<br />

(c)<br />

Figure 1. Microstructure of high water/cement ratio concrete: (a) high porosity and<br />

heterogeneity of the matrix, (b) orientated crystal of Ca(OH) 2 on aggregate, (c) CH crystal<br />

5.0 MAKING HIGH -PERFORMANCE CONCRETE<br />

High-performance concrete can not be made by a casual approach. All ingredients must be<br />

carefully selected and checked because their individual characteristics significantly affect the<br />

properties of the final product. What has been said about the coarse aggregate also holds<br />

true for the cement, the mineral components, the sand, the superplasticizer, and the other<br />

admixtures.<br />

Once the materials have been carefully selected, their proportions must also be determined<br />

meticulously. Particular attention must be paid to water content. Even seemingly<br />

insignificant volumes of water present in the aggregates or admixtures must be accounted<br />

for. Increasing the performance of concrete is so difficult and expensive; ruining it is as easy<br />

as including a little too much water.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 219 of 235


Cement<br />

Concrete<br />

Aggregate<br />

(a)<br />

C-S-H<br />

Ca(OH) 2<br />

(b)<br />

Figure 2. Microstructure of a high performance concrete: low porosity and homogeneity of<br />

the matrix<br />

Compressive strengths of from 50 to 75 MPa can usually be achieved fairly easily with most<br />

cements. On the other hand, experience has shown that it is more difficult to control the<br />

rheology long enough to place a 200-mm slump high-performance concrete an hour or more<br />

after mixing, due to the potential for incompatibility between the cement and superplasticizer.<br />

6.0 TEMPERATURE RISE<br />

There is a belief firmly rooted in the concrete community that the heat developed in any<br />

concrete is a direct function of its cement content. This does not always hold true because<br />

Portland cement by itself does not develop heat. Since Portland cement develops heat only<br />

as a result of hydration, it should be rather said that the heat developed in a concrete is a<br />

direct function of the amount of cement that is hydrating and not a direct function of the total<br />

amount of cement contained in the concrete.<br />

The amount of cement hydrating during the first hours can be limited by the amount of<br />

cement in the mix, such as in mass concrete. But, cement hydration can also be affected by<br />

the use of a retarder, a high dosage of superplasticizer, or insufficient water to hydrate all the<br />

cement in the mix. The latter describes the usual situation in high-performance concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 220 of 235


Cement<br />

Concrete<br />

45 MPa<br />

40<br />

Sandstone Limestone Quartzite<br />

σ (MPa)<br />

30<br />

20<br />

10<br />

0<br />

40<br />

Rocks<br />

10 -3 ε<br />

Sandstone<br />

σ (MPa)<br />

30<br />

20<br />

Limestone<br />

10<br />

Quartzite<br />

0<br />

10 -4<br />

Concretes<br />

ε<br />

Figure 3. Stress-strain curves of different rocks and of high-performance concretes made<br />

with crushed coars aggregates made out of these rocks<br />

σ1<br />

σ2<br />

σ1<br />

bV1<br />

aV1<br />

V2<br />

aV1<br />

bV1<br />

bV1 hydrated cement paste in the<br />

transition zone<br />

aV1 hydrated cement paste<br />

V2 volume of the coarse aggregate<br />

Figure 4. W. Baalbaki model for concrete<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 221 of 235


Cement<br />

Concrete<br />

Predicted values (GPa)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

De Larrard & Roy [France, 1992]<br />

Nilsen [Norway,1992]<br />

Smeplass [Norway, 1992]<br />

Alfes [Germany, 1992]<br />

Giaccio & al. [Argentina, 1994]<br />

n = 65<br />

Average deviation = 2.2%<br />

20<br />

20<br />

30 40 50 60<br />

Measured values (GPa)<br />

70<br />

Figure 5. Predictive value of W. Baalbaki model for concrete<br />

The temperature variation in a concrete due to this development of heat is usually positive<br />

(temperature rise), but it can be negative (temperature decrease in winter conditions) or<br />

almost zero (when the amount of heat generated within the concrete equals the heat losses<br />

through the surface of the concrete and through the forms).<br />

Consequently, it is not absolutely true that high-performance concrete develops greater heat<br />

of hydration than plain concrete. In fact, it can be totally false in the case of a particular<br />

structural element or under specific conditions. This has been confirmed by the work of<br />

Cook et al. [4], in which three similar columns measuring 1 x 1 x 2.4 m were cast with three<br />

different concretes with compressive strengths of 30, 80, and 120 MPa. The temperature of<br />

the concrete was monitored at several places within the columns. The temperature rise at<br />

the centre of each column, where the temperature was always the highest, was nearly<br />

identical in the three columns, despite the fact that the cement content varied in the three<br />

mixes from 355 kg/m 3 for the 30-MPa concrete up to 540 kg/m 3 for the 120-MPa concrete as<br />

shown in Figure 6.<br />

Faced with such results, many engineers are surprised that to learn a plain concrete can<br />

develop as much heat and temperature rise as high-performance concrete, because this<br />

statement goes against the long-standing myth that the temperature rise in a concrete is<br />

directly proportional to the cement content.<br />

Indeed, in this example the 120-MPa concrete evidenced the lowest temperature rise<br />

because less cement hydrated during the first 30 hours. Less cement had the chance to<br />

hydrate in the 120-MPa concrete for several reasons: 1) less water was used to make the<br />

concrete, 2) more superplasticizer was used (it is well-known that naphthalene<br />

superplasticizers act as retarders when used at high dosages), 3) a retarder was used, and<br />

4) the hydrates formed in the early stage of hydration in high-performance concrete are so<br />

compact that hydration kinetics are controlled by diffusion of water through the hydrates<br />

rather than by chemical processes of dissolution and precipitation, which controls kinetics<br />

when there is plenty of water in the mixture.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 222 of 235


Cement<br />

Concrete<br />

Temperature Rise, °C<br />

50<br />

40<br />

30<br />

20<br />

10<br />

35 MPa column<br />

90 MPa column<br />

120 MPa column<br />

80<br />

60<br />

40<br />

20<br />

Temperature Rise, °F<br />

0<br />

0<br />

0 24 48 72 96 120<br />

Age, hours<br />

Figure 6. Temperature increases at the center of each column<br />

Moreover, as shown by Lachemi et al. [5], temperature rise is not uniform through the<br />

structure and the maximum temperature is not reached at the same time. The temperature<br />

rise depends not only on the amount of heat developed within the concrete, but also on the<br />

thermodynamic conditions at the boundaries. Of course, the more massive the structural<br />

element, the higher the temperature rise; the higher the ratio of exposed surface to the<br />

volume of concrete, the lower the temperature rise. A temperature decrease can be<br />

observed under severe winter conditions. as shown by Lessard et al. [6] in the reconstruction<br />

of the sidewalk entrance and by Lessard [7] during the construction of a bridge.<br />

Using finite-element modelling, Lachemi et al. [8] studied the influence of ambient<br />

temperature and concrete temperature on the temperature rise in some structural elements<br />

they monitored during a field experiment. Their main findings were that the highest<br />

temperature recorded was obtained at the highest ambient temperature when a hot concrete<br />

mixture was cast, but the more critical conditions for the development of high thermal<br />

gradients were achieved when a hot concrete was cast on a cold day and then cools.<br />

7.0 SHRINKAGE<br />

If water curing is essential to develop the potential strength of cement in plain concrete, early<br />

water curing is crucial for high-performance concrete in order to avoid the rapid development<br />

of autogenous shrinkage and to control concrete dimensional stability, as explained below.<br />

Cement paste hydration is accompanied by an absolute volume contraction that creates a<br />

very fine pore network within the hydrated cement paste. This network drains water from<br />

coarse capillaries, which start to dry out. If no external water is added during curing, the<br />

coarse capillaries empty of water as hydration progresses, just as though the concrete were<br />

drying. This phenomenon is called self -desiccation. The difference between drying and<br />

self-desiccation is that when concrete dries water evaporates to the atmosphere, while<br />

during self-desiccation water stays within concrete (it only migrates towards the very fine<br />

pores created by the volumetric contraction of the cement paste) [9].<br />

In ordinary concrete with w/c greater than 0.50, for example, there is more water than<br />

required to fully hydrate the cement particles. A large amount of this water is contained in<br />

well-connected large capillaries so that the menisci created by self-desiccation appear in<br />

large capillaries where they generate only very low tensile stresses. Therefore the hydrated<br />

cement paste barely shrinks when self-desiccation develops.<br />

In the case of high-performance concrete with a w/b of 0.30 or less, significantly more<br />

cement and less mixing water have been used, so that the pore network is essentially<br />

composed of fine capillaries. When self-desiccation starts to develop as soon as hydration<br />

begins, the menisci rapidly develop in small capillaries if no external water is added. Since<br />

many cement grains start to hydrate simultaneously in high-performance concrete, the drying<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 223 of 235


Cement<br />

Concrete<br />

of very fine capillaries can generate high tensile stresses that shrink the hydrated cement<br />

paste. This early shrinkage is referred to as autogenous shrinkage (of course, autogenous<br />

shrinkage is as large as the drying shrinkage observed in ordinary concrete when these two<br />

types of drying develop in capillaries of the same diameter).<br />

But, when there is an external supply of water, the capillaries do not dry out as long as they<br />

are connected to this external source of water. The result is that no menisci, no tensile<br />

stress, and no autogenous shrinkage develop within a high-performance concrete that is<br />

constantly water cured since its setting.<br />

Thus an essential difference between ordinary concrete and high-performance concrete is<br />

that ordinary concrete exhibits practically no autogenous shrinkage, whether it is water-cured<br />

or not, whereas high-performance concrete can experience significant autogenous shrinkage<br />

if it is not water-cured during the hydration process. Autogenous shrinkage does not develop<br />

in high-performance concrete if the capillaries are interconnected and have access to<br />

external water, but, when the continuity of the capillary system is broken, then and only then<br />

does autogenous shrinkage start to develop within the hydrated cement paste of a highperformance<br />

concrete, as shown in Figure 7.<br />

Plastic<br />

shrinkage<br />

Self<br />

dessication<br />

Drying<br />

How to cure concrete to minimize its shrinkage<br />

Curing<br />

membrane<br />

or fog<br />

misting<br />

Water<br />

curing or<br />

fog misting<br />

Impervious<br />

film<br />

Water must be<br />

prevented from<br />

evaporating<br />

Menisci<br />

formation has<br />

to be avoided<br />

Dessication must be<br />

avoided. Self<br />

dessication develops<br />

until hydration stops.<br />

Figure 7. The most appropriate curing regimes during the course of the hydration reaction<br />

Drying shrinkage of the hydrated cement paste begins at the surface of the concrete and<br />

progresses more or less rapidly through concrete, depending on the relative humidity of the<br />

ambient air and the size of capillaries. Drying in ordinary concrete is therefore rapid because<br />

the capillary network is well connected and contains large capillaries. Drying shrinkage in<br />

high-performance concrete is slow because the capillaries are very fine and soon get<br />

disconnected.<br />

A major difference between drying shrinkage and autogenous shrinkage is that drying<br />

shrinkage develops from the surface inwards, while autogenous shrinkage is homogeneous<br />

and isotropic insofar as the cement particles and water are well dispersed within the<br />

concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 224 of 235


Cement<br />

Concrete<br />

Thus there are considerable differences between ordinary and high-performance concrete<br />

with respect to their shrinkage behaviour. The cement paste of ordinary concrete exhibits<br />

rapid drying shrinkage progressing from the surface inwards, whereas high-performance<br />

concrete cement paste can develop a high isotropic autogenous shrinkage when not water<br />

cured. This difference in the shrinkage behaviour of the cement paste has very important<br />

consequences for concrete curing and concrete durability.<br />

Although the shrinkage of the hydrated cement paste is a very important parameter with<br />

respect to concrete volumetric stability, it is not the only one. A key parameter is the amount<br />

of aggregate and, more specifically, the amount of coarse aggregate. Too often it is forgotten<br />

that the aggregates do more than simply act as fillers in concrete. In fact, they actively<br />

participate in the volumetric stability of concrete when they restrain the shrinkage of the<br />

hydrated cement paste: concrete shrinkage is always much lower than that of a cement<br />

paste having the same w/c. It is common knowledge that concrete shrinkage can be easily<br />

reduced by increasing the coarse-aggregate content; the shrinkage of the hydrated cement<br />

paste stays the same, but it is more restrained, so that the volumetric stability of the concrete<br />

is increased. Restraining the shrinkage of hydrated cement paste by modifying the coarseaggregate<br />

skeleton may or may not produce a network of microcracks, depending on the<br />

intensity of the tensile stresses developed by this process with respect to the tensile strength<br />

of the hydrated cement paste.<br />

8.0 CURING<br />

High-performance concrete must be cured quite differently from ordinary concrete because<br />

of the difference in shrinkage behaviour described above. If HPC is not water-cured<br />

immediately following placement or finishing, it is prone to develop severe plastic shrinkage<br />

because it is not protected by bleed water, and later on develops severe autogenous<br />

shrinkage due to its rapid hydration. While curing membranes provide adequate protection<br />

for ordinary concrete (which is insensitive to autogenous shrinkage), they can only help<br />

prevent the development of plastic shrinkage in high-performance concrete but they have no<br />

value in inhibiting autogenous shrinkage.<br />

The critical curing period for any HPC runs from placement or finishing up to 2 or 3 days<br />

later, and the most critical period is usually from 12 to 36 hours, as shown in Figure 7. In<br />

fact, the short time during which efficient water curing must be applied to HPC can be<br />

considered a significant advantage over ordinary concrete. Those who specify and use HPC<br />

must be aware of the dramatic consequences of missing early water curing. Initiating water<br />

curing after 24 hours is too late because, most of the time, a great deal of autogenous<br />

shrinkage has already occurred and, by this time, the microstructure is already so compact<br />

that external water has little chance of penetrating very deep into the concrete.<br />

Water pounding or fogging are the best ways to cure HPC; one of these two methods must<br />

be applied as soon as possible, immediately following placement or finishing. An<br />

evaporation retarder can be applied temporarily to prevent the development of plastic. If, for<br />

any reason, water pounding or fogging cannot be implemented for 7 days, then the concrete<br />

surface should be covered with wet burlap (hessian) or preferably a prewetted geotextile.<br />

The burlap or the geotextile must be kept constantly wet with a soaker hose and protected<br />

from drying by a polyethylene sheet in order to ensure that at no time during the curing<br />

period is the concrete allowed to dry and experience any autogenous shrinkage.<br />

Moreover it is observed that when any concrete is water cured during setting it does not<br />

shrink but rather swell. Figure 8 illustrate the effect of early water curing on the volumetric<br />

change of concrete. The water curing can be stopped after 7 days because most of the<br />

cement at the surface of concrete has hydrated and any further water curing has little effect<br />

on the development of shrinkage. After 7 days of water curing, HPC experiences slow drying<br />

shrinkage due to the compactness of its microstructure, and autogenous shrinkage has<br />

already dried out the coarse capillaries pores. Even then, theoretically the best thing to do is<br />

to paint HPC or to use a sealing agent so that the last remaining water can be retained to<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 225 of 235


Cement<br />

Concrete<br />

contribute to hydration. There is no real advantage to painting or sealing a very porous<br />

concrete because it is impossible to obtain an absolutely impermeable coating; painting or<br />

sealing HPC, however, can be easy and effective. Figure 9 illustrate the curing conditions<br />

that will result or not into the development of autogenous shrinkage.<br />

Swelling<br />

(× 10 -6 )<br />

Shrinkage<br />

(× 10 -6 )<br />

100<br />

0<br />

-100<br />

-200<br />

-300<br />

-400<br />

Water<br />

curing<br />

7<br />

Sealed after 7 days<br />

Age (days) 28<br />

Air drying<br />

after 7 days<br />

2<br />

No curing<br />

1<br />

3<br />

Figure 8. Length changes according to different curing regimes for the 0.35 W/C concrete<br />

Hydration<br />

Volumetric<br />

contraction<br />

Self<br />

desiccation<br />

Menisci<br />

Autogenous<br />

shrinkage<br />

No<br />

External<br />

supply of<br />

water<br />

No<br />

No<br />

autogenous<br />

shrinkage<br />

Yes<br />

Pores<br />

and capillaries<br />

connected?<br />

Yes<br />

No<br />

menisci<br />

Figure 9. Influence of curing conditions on the occurence of autogenous shrinkage<br />

Partial replacement of coarse aggregate by an equivalent volume of saturated lightweight<br />

aggregate has been used to counteract autogenous shrinkage internally. The saturated<br />

lightweight aggregate particles act as small water reservoirs throughout the mass of<br />

concrete; they can be emptied into the very fine pores created by hydration reaction.<br />

Therefore, the water in the lightweight aggregate particles is drained along with that<br />

contained in the fine capillaries of the HPC. The menisci developed within the cement paste<br />

are not as small, which means lower tensile stress and less autogenous shrinkage.<br />

But, it is well known that concrete is never cured properly in the field, if spite of the fact that it<br />

is always written in the specifications that contractor has to cure concrete. Contractors are<br />

not curing concrete for a very simple reason: they are not specificallypaid for, therefore,<br />

concrete curingis always perceived by them as an unprofitable activity or even a source of<br />

expense and therefore a waste of time. But, when contractors are specifically paid to water<br />

cure concrete they do it as they do it for any other item that is paid. Since three years now<br />

the City of Montreal and the Department of Transportation of Quebec are requesting unit<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 226 of 235


Cement<br />

Concrete<br />

prices for each item directly related to an early water curing. Since the initiation of this new<br />

policy on the early water curing of concrete, it is amazing to see how zealous contractors are<br />

in the matter of water curing. For them water curing is now seen as a source of profit. Form<br />

the first experiences in that matter it has been found that the cost of an early water curing is<br />

about one tehth of one percent. A very modest price when considering the improved<br />

durability of the concrete structures that are built like that.<br />

Therefore the best way to be sure that high-performance concretes are properly and<br />

efficiently cured in the field is to specifically pay contractors to cure concrete [10].<br />

9.0 DURABILITY<br />

9.1 General matters<br />

The durability of a material in a particular environment can only be established over time, so<br />

it is difficult to precisely predict the longevity of high-performance concrete because we do<br />

not have a track record for HPC exposed to very harsh environments for more than 5 to 10<br />

years, except perhaps for some North Sea offshore platforms. It must be remembered that<br />

the first uses of high strength concrete in the late sixties and early seventies were indoor<br />

applications, mainly in columns in high-rise buildings, which is not a particularly severe<br />

environment. Outdoor applications of high-performance concrete only date from the late<br />

eighties and early nineties, which means that not enough time has gone by to properly<br />

assess the real service life of any high-performance concrete structures under outdoor<br />

conditions.<br />

Based on years of experience with ordinary concrete, we can safely assume that high<br />

performance concrete is more durable than ordinary concrete. Indeed, the experience<br />

gained with ordinary concrete has taught us that concrete durability is mainly governed by<br />

concrete permeability and the harshness of the environment [11].<br />

It is easy to assess the harshness of any environment with respect to high-performance<br />

concrete because hydrated cement paste is essentially a porous material that contains some<br />

freezable water. Assessment involves simply examining how the environment affects each<br />

of these characteristics.<br />

On the other hand, it is not always simple to assess how easily aggressive agents will<br />

penetrate concrete. For example, water flow through a 0.70 w/c concrete is easy to<br />

measure, but water flow almost stops in a 0.40 w/c concrete, regardless of the thickness of<br />

the sample and the amount of pressure applied. The gas permeability is also difficult to<br />

measure. Sample preparation, particularly drying, significantly influences gas permeability.<br />

Therefore, the critical question remains how to appropriately assess the permeability of a<br />

concrete with a low w/b and a very compact microstructure.<br />

Despite all the criticism levelled at it, the so-called "Rapid chloride-ion permeability test"<br />

(AASHTO T-277) gives a fair idea of the interconnectivity of the fine pores in concrete that<br />

are too fine to allow water flow. Experience has revealed good correlation between the water<br />

permeability and rapid chloride ion permeability for concrete specimens with a w/c greater<br />

than 0.40. Chloride-ion permeability is expressed in coulombs, which corresponds to the total<br />

amount of electrical charge that passes during the 6-hour test through the concrete sample<br />

when subjected to a potential difference of 50 volts.<br />

When the rapid chloride-ion permeability test is performed on concrete samples with lower<br />

w/c, the number of coulombs passing through the sample decreases. It is easy to achieve a<br />

chloride-ion permeability of less than 1000 coulombs for a high performance concrete<br />

containing about 10% silica fume and having a w/b around 0.30. The only other way to<br />

achieve this would be with latex-modified concrete, which would be much more costly. Much<br />

lower chloride-ion permeability values can be achieved if the w/b is reduced below 0.25.<br />

Values as low as 150 coulombs have been reported, far lower than the 5000 to 6000<br />

coulombs reported for ordinary concrete [12].<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 227 of 235


Cement<br />

Concrete<br />

The rapid chloride-ion test also reveals that the connectivity of the pore system decreases<br />

drastically as the w/b decreases, making the migration of aggressive ions or gas more<br />

difficult in high-performance concrete than in its plain counterpart. The author believes that<br />

this is the best indication that the service life of high-performance concrete should exceed<br />

that of ordinary concrete in the same environment. It is difficult to determine the number of<br />

years by which the service life would be extended because the predictive models developed<br />

for ordinary concrete cannot be readily extrapolated to include high-performance concrete.<br />

However, it can be said that some high-performance concrete structures will outlast the<br />

average life span of a human being in developed countries.<br />

9.2 Durability in a marine environment<br />

9.2.1 Nature of the aggressive action<br />

Sea water by itself is not a particularly harsh environment for plain concrete, but a marine<br />

environment can be very harmful to reinforced concrete due to the multiplicity of aggressions<br />

that it can face [13]. In a marine environment, a concrete structure is essentially submitted to<br />

four types of aggressive factors:<br />

• chemical factors related to the presence of various ions dissolved in the sea water or<br />

transported in the wet air;<br />

• geometrical factors related to the fluctuation of the sea level (tides, storms, etc.);<br />

• physical factors such as freezing and thawing, wetting and drying, etc.;<br />

• Mechanical factors such as the kinetic action of the waves, the erosion caused by sand in<br />

suspension in the sea water, floating debris and even floating ice in northern seas.<br />

It is the combination of these different factors that can be harmful to reinforced concrete<br />

structures. In the following, we will review very briefly the nature of each attack in order to<br />

show how high performance concrete is the best concrete fitted to resist not only each of<br />

these particular factors, but also their combined action.<br />

9.2.2 Chemical attack on concrete<br />

Sea water is not particularly harmful to plain concrete. Several submerged plain concrete<br />

blocks and structures exposed for nearly 100 years in different marine environments are still<br />

in relatively good condition. The only chemical limitation usually recommended for a cement<br />

to be used in a marine environment is related to its C3A content, which should not be greater<br />

than 8%.<br />

Figure 10 represents the different successive altered zones found in a concrete exposed to<br />

sea water for several years: carbonation, formation of brucite, of monochloroaluminate, and<br />

sulfatic attack with formation of gypsum, ettringite or even thaumasite . Each of these<br />

chemical mechanisms is well known and explained in specialized books [13, 14, 15].<br />

Zone 1: Carbonated layer<br />

Zone 2: Magnesia attack<br />

Zone 3: Sulfatic attack with<br />

formation of gypsum<br />

Zone 4: Sulfatic attack<br />

Sea<br />

water<br />

Zone 5:<br />

leaching<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 228 of 235


Cement<br />

Concrete<br />

Figure 10. Schematic representation of the different altered layers found in a concrete<br />

marine structure<br />

Sea water is very harmful to reinforced concrete. Once chlorine ions are reach the<br />

reinforcing steel level, resulting in a rapid spalling of the covercrete, it is easier for chlorine<br />

ions to reach the second level of rebars, and so on. The only way to avoid, or to retard as<br />

long as possible, the corrosion of the rebars by chlorine ions is:<br />

• to specify a very compact and impervious concrete, and place and CURE it correctly,<br />

and;<br />

• to increase the concrete cover.<br />

The development of all these mechanisms of aggression is closely related to the facility with<br />

which aggressive ions can penetrate concrete, therefore it is obvious that a very dense and<br />

impervious matrix, like the one found in high performance concrete, constitutes the best<br />

protection that can be presently offered against a marine environment. High performance<br />

concrete has been used very successfully for more than 20 years to build offshore platforms,<br />

and more recently to build two major bridges for which the owner had requested a 100-year<br />

life cycle: the Confederation bridge in Canada and the Tago bridge in Lisboa, Portugal. It is<br />

interesting to point out that these two bridges have been built in a BOOT mode (Build, Own,<br />

Operate and Transfer) by consortia of contractors that will have to maintain these two<br />

bridges during the entire concession time.<br />

In that respect, it is interesting to note that in the case of these two bridges, the concrete<br />

cover has been extended to 75 mm to meet the 100-year life cycle requirement.<br />

It is also very important to point out that it is not sufficient to specify a Type V cement or a<br />

slag cement to obtain a concrete that will resist to a harsh marine environment. The curing<br />

of this concrete is as important as the selection of an appropriate cementitious system. The<br />

rapid deterioration of the precast elements of the Dubaï causeway in the Arabian peninsula is<br />

a good example of what must not be done.<br />

9.2.3 Physical attack<br />

There is still controversy about the necessity of entraining air in high performance concrete to<br />

make it freeze-thaw resistant. In Canada, any exposed high performance concrete must be<br />

air entrained, which is the case of the concrete of the Confederation bridge. In Norway, high<br />

performance concrete can contain a small amount of air, but more to facilitate its placing and<br />

finishing rather than to improve its freeze-thaw resistance. As freeze-thaw resistance of<br />

marine structures is not a crucial issue in New Zealand, it is not my intention to insist more<br />

on this subject which I covered in detail in the original paper from which this one is derived<br />

[16].<br />

9.2.4 Mechanical attack<br />

In this case also the compactness and the high resistance of the matrix of a high<br />

performance concrete offers a good protection to the abrasive action of the sand of the<br />

debris, or even from floating ice. In the case of the Confederation bridge, the concrete used<br />

to build the conical part of the piles that deflects the ice loads in the tidal zone was a 90 MPa<br />

air-entrained high performance concrete. This concrete is thought to be able to resist the<br />

tidal freezing and thawing cycles in winter and the abrasive action of the floating ice which is<br />

particularly severe in the Northumberland Straight, due to the presence of changing currents<br />

associated with the tides and winds.<br />

9.2.5 Conclusion<br />

It is obvious that high performance concrete is a material of choice fitted to resist the harsh<br />

environment that is found in coastal areas particularly well. It would be a great mistake to<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 229 of 235


Cement<br />

Concrete<br />

take advantage of the improved durability of the concrete cover to reduce its thickness, on<br />

the other hand it would be wise to increase it during the next millenium if we want to start to<br />

implement an efficient sustainable development policy in civil engineering. Having been<br />

raised on the shores of the Atlantic Ocean in Biarritz, France, I know how aggressive marine<br />

environments can be for a reinforced concrete structure built with a poor concrete, that is<br />

badly cured. I also know how durable are the German blockhaus built at the end of the last<br />

World War.<br />

10.0 THE FIRE RESISTANCE OF HPC<br />

For many years, the fire resistance of HPC has been a controversial subject, some reports<br />

saying that HPC performed as well as usual concrete, others the reverse [17-26]. Following<br />

the first fire that occured in a HPC structure, the Chunnel fire [27-28], and from different<br />

studies in progress in several countries it is clear that the fire resistance of HPC does not<br />

seem to be as good as that of usual concrete, but that it is not as bad as some alarming<br />

reports have presented it. As is the case for any concrete, HPC is one of the safest<br />

construction materials as far as fire resistance is concerned.<br />

As constructive details, as well as material resistance by itself, could influence greatly the fire<br />

resistance of a structural element, it is presently impossible to give very simple rules that<br />

should govern the design of HPC structures that could be exposed to a more or less severe<br />

fire. Presently, some models have been developed that can be used to forecast the<br />

structural consequence of a fire on the safety of a HPC structure. Moreover several<br />

promising avenues are presently being investigated to improve the fire resistance of HPC<br />

itself.<br />

Instead of reviewing in detail the controversial literature on the fire resistance of HPC three<br />

brief presentations on the actual fire resistance of HPC and usual concrete will be done. The<br />

first one will deal with the violent fire that occured in the Chunnel, based on a report by J.-M.<br />

Demourieux [28], the second one will deal with the big conflagration that occured in<br />

Düsseldorf airport and the third one will present the latest fndings of the Brite-Euram<br />

HITECO BE-1158 research project [29].<br />

10.1 The Chunnel fire<br />

The fire of a truck in the Chunnel did not suprise the safety department of the Chunnel<br />

administration. The occurence of such a fire had been forecast because each day a truck<br />

burns in France and another in England [27]. If a fire was to happen in the Chunnel, the<br />

engine-man of the train was asked to speed up to get the train out of the Chunnel as soon as<br />

possible.<br />

It had also been forecast that some hydraulic jacks used to support the access ramp used by<br />

the trucks to get on the railway platform could be loosened and create a risk of derailment for<br />

the train. In such a case, the driver could be asked to stop the train and tighten the hydraulic<br />

jack. But, what had not been forecast was that the two incidents could happen<br />

simultaneously so that the engine-man could receive two conflicting orders at the same time<br />

for which no priorization had been included in the safety procedures.<br />

Facing two contradictory orders the engine-man decided on November 11, 1996 to stop the<br />

train at 18 km from the French entrance, fortunately in the driest zone of the Chunnel from a<br />

water seepage point of view. In this area, the blue chalk through which the Chunnel was<br />

excavated was the most impervious of all the rock formations. When the moles were<br />

excavating this area they registered their all time speed record and the workers asked for<br />

some water to be pulverized in the excavated rock to get rid of the fine dry chalk dust<br />

generated by the moles.<br />

It is difficult to imagine the damage that could have occured to the Chunnel if the fire had<br />

taken place a few kilometers further in a fault area where it would not have been possible to<br />

take advantage of the imperviousness of the chalk layer.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 230 of 235


Cement<br />

Concrete<br />

The fire was particularly violent as far as its maximum temperature and length are<br />

concerned. The maximum temperature that was reached was in the order of 700 to 1000 °C<br />

and the total duration of the fire was about 10 hours. The concrete lining in the area of the<br />

fire was composed of precast concrete elements having a design strength of 70 to 80 MPa<br />

HPC having a water/binder ratio of 0.32 [28].<br />

In the most intense zone, the concrete lining was severely damaged and would not have<br />

been able to counteract the hydrostatic pressure for which it has been designed. An<br />

extensive survey done after the fire on the unaffected zone has shown that the in-place<br />

concrete had an average compressive strength comprised between 70 and 80 MPa, a<br />

modulus of rupture comprised between 7 and 8 MPa, and an elastic modulus comprised<br />

between 37 and 44 GPa.<br />

As far as the damaged zone is concerned, it is 480 m long, involving about 300 lining rings<br />

400 mm thick. It could be divided into six parts that displayed different degrees of damage.<br />

In the two farthest zones from the centre of the fire, which were 80 and 100 m long on both<br />

sides, the concrete lining had been only slightly damaged and no steel reinforcement was<br />

apparent.<br />

In the next two adjacent zones, which were 70 m long, the concrete lining was more severely<br />

damaged over a thickness comprised between 50 to 100 mm. In this zone, the first<br />

reinforcing level can be seen here and there. The more severely damaged section was<br />

observed on the upper part of the lining.<br />

The central zone where the fire was the most intense could be divided into two parts, a 50 m<br />

long area very severely damaged and another one 90 m long slightly less damaged. In the<br />

most severely damaged zone the concrete lining was completely destroyed all over its<br />

thickness (400 mm), more specifically in the 3H00 to 8H00 zones. In the less damaged part,<br />

the residual concrete had a thickness comprised between 50 and 200 mm except in the<br />

lowest part where its thickness was still 350 mm. In the less damaged zone, the concrete<br />

lining had a residual thickness of about 200 mm in the upper part and of about 350 mm in the<br />

lowest part.<br />

It was observed that in this zone the reinforcing steel arrangement played a very important<br />

role because the concrete, which was prevented from falling down by the reinforcing bars,<br />

protected the subjacent concrete. Therefore in such zones, the steel reinforcing bars played<br />

a key role in protecting the subjacent concrete. In many places in the central zone, the<br />

reinforcing steel was highly deformed due to the severity of the fire. Under the effect of the<br />

heat, the restrained dilatation of the lining generated large transversal and horizontal<br />

stresses in the damaged concrete so that numerous 45° fissures could be observed at the<br />

edges of the precast elements. Concrete was litterally hollowed out in at the centre of the<br />

reinforcing mesh over a more or less thick depth depending on its location from the centre of<br />

the fire.<br />

Concrete spalled in small pieces having an average thickness of about 10 mm. Some of<br />

theses pieces were no bigger than a coin. The concrete lining was always more damaged in<br />

its upper zone than at its bottom end on the track.<br />

In the two less damaged zones, concrete had spalled over greater surfaces in some places.<br />

It was possible to see that it happened frequently where nylon spacers had been used during<br />

precasting to correctly place the reinforcing steel. Most probably the pressure of the gas<br />

generated during the burning of the nylon spacers was responsible for this spalling.<br />

All the tests done in the non-damaged part of the lining in the fire zone have shown that the<br />

residual concrete remaining in the lining was almost intact in the upper part as well as in the<br />

lower part of the section. SEM observations have shown that the residual concrete had not<br />

been altered significantly, no permanent strains could be seen even at the surface of the<br />

residual concrete. In all cases the residual concrete was altered on a very thin layer, and<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 231 of 235


Cement<br />

Concrete<br />

according to petrographical examinations, the maximum temperature reached by the<br />

concrete in that area was lower than 700 °C.<br />

The observations and test results conducted on the concrete of the track and at the lower<br />

part of the lining resulted in the decision to keep them in place during the reconstruction.<br />

Based on all the observations done and the results obtained, it was decided to rebuild the<br />

lining using a wet shotcrete after a careful cleaning of the damaged concrete.<br />

10.2 The Düsseldorf Airport Fire<br />

On April 16th, 1996 a devastating conflagration broke out at the Düsseldorf Airport, killing<br />

17 people by smoke inhalation and injuring several hundred. The fire was attributable to<br />

improper use of combustible insulating material and the plastic cover of the cables laid in the<br />

hollow ceiling space [29].<br />

The fire was triggered by some welding work which had been performed, and developed<br />

unnoticed for a long time so that the smoke had time to spread into all parts of the building in<br />

the hollow ceiling and through the ventilation system.<br />

The highest estimated temperature to which the concrete ceiling was exposed has been<br />

estimated at 1000 °C. At such a temperature the 25 MPa concrete spalled, but from a<br />

structural point of view the building elements were not damaged. Moreover, it was found that<br />

harmful substances such as dioxins did not penetrate into the concrete so that the concrete<br />

structure was still in serviceable condition.<br />

Since owners and operators intended to erect an extended and modernized airport facility,<br />

they decided to demolish the burnt concrete structure.<br />

10.3 Spalling of concrete under fire conditions<br />

It is difficult to make a direct comparison between these two major conflagrations but it can<br />

be pointed out that in both cases, very fortuituously, the initial cause of the fire was the<br />

burning of some polystyrene, that the maximum temperature reached has been estimated at<br />

1000 °C, and that the high-performance concrete of the Chunnel and the usual concrete of<br />

Düsseldorf Airport both spalled.<br />

Of course, the thickness of the spalling is a function of the maximum temperature that is<br />

reached during the fire and of the duration of exposure to this temperature, as it has been<br />

seen in the various zones of the fire area in the Chunnel.<br />

According to J.-M. Demourieux, a standard ISO 834 fire would have to cause a 30 mm thick<br />

spalling in the lining of the Chunnel.<br />

10.4 The Brite-Euram HITECO BE-1158 Research Project<br />

The preliminary conslusions of this research project, financed by the European Community,<br />

were presented on March 9th at a meeting of the French Civil Engineering Association<br />

meeting. All the tests were done in Finland at the V.T.T. Laboratories, one of the best<br />

equipped European laboratories for fire studies [30].<br />

The conclusion of this presentation is: the experimental study undertaken on a 60 MPa highperformance<br />

concrete without silica fume and a 90 MPa high-performance concrete with<br />

silica fume have shown an excellent fire resistance, except for a small column that was<br />

heavily loaded.<br />

In actual structures more favorable conditions are found :<br />

• the columns have a bigger size than the one tested;<br />

• the loading is the service loading and not the maximum loading.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 232 of 235


Cement<br />

Concrete<br />

The FIREXPO software can be used to predict the thermomechanical behavior of any<br />

structural element.<br />

11.0 THE FUTURE<br />

High-performance concrete is not a passing fad. It is here to stay, not only because of its<br />

high strength, but also because of its durability. For example, an outdoor concrete parking<br />

garage can be built with 20-MPa concrete under current codes. Its columns and slabs would<br />

be somewhat bigger than if an 80-MPa concrete had been used. But the life cycle of the<br />

garage will be very short in an environment as severe as that in Eastern Canada or in a sea<br />

coastal area, because 20-MPa concrete cannot adequately protect the reinforcing steel<br />

against corrosion from chlorine ions. The service life would be somewhat longer in a less<br />

severe environment, but would still be short due to carbonation.<br />

Therefore, at the dawn of the twenty-first century, it is not particularly difficult to predict that<br />

the use of high-performance concrete will increase in order to extend the service life of<br />

concrete structures exposed to severe environments [11]. The durability of a concrete<br />

structure depends on several factors, one of which is the durability of the concrete itself. As<br />

the durability of concrete is essentially linked to its permeability, high-performance concrete,<br />

with its compact microstructure and very low permeability, should obviously be more durable<br />

than ordinary concrete. It must be emphasised, however, that good concreting practice,<br />

including good curing, are essential to creating a durable structure. It would be a pity if<br />

improper practice and poor curing resulted in a structure with impervious concrete in<br />

between the cracks.<br />

We still do not know how to make high-performance concrete with low permeability but<br />

without the high strength. Therefore, designers have to learn to take advantage of the extra<br />

strength provided by low w/b concrete. One day, we may be able to make durable concrete<br />

of lower strength.<br />

Another reason that will lead to a greater use of high-performance concrete in the 21 st<br />

century is society's greater interest in ecological concerns [31]. Many others share my<br />

opinion that we cannot continue the wasteful use of our natural resources and energy that<br />

characterised the 19 th and 20 th centuries. High-performance concrete is more ecological<br />

than ordinary concrete. High-performance concrete, with its more compact microstructure,<br />

provides the sought property with less material.<br />

Moreover, the 21 st century will be the century of recycling. Recycling paper, cardboard,<br />

aluminium, steel, and even plastic has already become quite common in many countries.<br />

The trend towards recycling has already reached the concrete field. We must remember,<br />

however, that recycling a material leads to a product of lower quality. This is because the<br />

original purity of the initial materials are lost when different additives are combined in order to<br />

enhance the final performance of the product. Concrete production has not reached the<br />

same degree of sophistication as rubber manufacturing, although concrete is becoming more<br />

complex every day.<br />

Another new development is reactive-powder concrete, which can have a compressive<br />

strength of about 200 MPa in an unconfined state and as high as 350 MPa when confined in<br />

thin steel tubes. Moreover, the addition of steel fibres to reactive-powder concrete produces<br />

a material with modulus of rupture as high as 25 to 35 MPa in addition to high compressive<br />

strength [32, 33]. Reactive-powder concrete will pose no threat to the high-performance<br />

concrete market for years to come. Once we are able to produce and use reactive-powder<br />

concrete and to design with it at an industrial scale, high-performance concrete will have<br />

already won over a significant part of the market of ordinary concrete.<br />

Now that 1000-MPa Portland cement-based materials can be made, only a pessimist could<br />

refuse to see the future of high-performance concrete.<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 233 of 235


Cement<br />

Concrete<br />

12.0 ACKNOWLEDGEMENT<br />

This presentation was made possible due to the support of the Canadian Centres of<br />

Excellence Research Program on High-Performance Concrete.<br />

The following text was originally presented in a symposium held in tribute to Mario Collepardi<br />

in Rome in October 1997. It was edited by Leslie Strubble and included as chapter 8 —<br />

High-Performance Concrete in the book Materials Science of Concrete V published by the<br />

American Ceramic Society. This present version has been updated. Freeze-thaw resistance<br />

is not presented and is replaced by Fire Resistance.<br />

13.0 REFERENCES<br />

1. Baalbaki, W., Benmokrane, B., Chaallal, O., and Aïtcin, P.-C., "Influence of Coarse<br />

Aggregate on Elastic Properties of High-Performance Concrete," ACI Mater. J. 88(5) pp.<br />

499-503 (1991).<br />

2. Nilsen, A.U., and Aïtcin, P.-C., "Properties of High-Strength Concrete Containing Light-<br />

Normal- and Heavyweight Aggregate," Cem. Concr. Aggregates 14(1) 8-12 (1992).<br />

3. Baalbaki, W., PhD. Thesis, Université de Sherbrooke (1997).<br />

4. Cook, W.D., Miao, B., Aïtcin, P.-C., and Mitchell, D., "Thermal Stress in Large High-<br />

Strength Concrete Columns,"ACI Mater. J. 89(1) 61-66 (1992).<br />

5. Lachemi, M., Lessard, M., and Aïtcin, P.-C., "Early-Age Temperature Development in a<br />

High-Performance Concrete Viaduct," Am. Concr Inst. SP-167 (High Strength concrete:<br />

An International Perspective, edited by J.A. Bickley) 149 - 174 (1996).<br />

6. Lessard, M., Dallaire, É, Blouin, D., and Aïtcin, P.-C., "High-Performance Concrete<br />

Speeds Reconstruction of McDonald's," Concr. Int. 16(9) 47-50 (1994).<br />

7. Lessard, M., unpublished results, 1995.<br />

8. Lachemi, M., and Aïtcin, P.-C., "Influence of Ambient and Fresh Concrete Temperature<br />

on the Maximum Temperature and Thermal Gradient in a High-Performance Concrete<br />

Structure," ACI Mater. J. 94(2) 102-110 (1997).<br />

9. Aïtcin, P.-C., Neville, A.M., and Acker, P., "Integrated View of Shrinkage Deformation,"<br />

Concr. Int. 19(9) 35-41 (1997).<br />

10. Standard Specification for High-Performance Concrete HPC (3VM-20), 19 p. (1998).<br />

11. Aïtcin, P.-C., "Durable Concrete-Current Practice and Future Trends," Am. Concr. Inst.<br />

SP-144 (Concrete Technology: Past, Present, and Future, edited by P.K. Mehta) 83-104<br />

(1994).<br />

12. Gagné, R., Lamothe, P., Aïtcin, P.-C., "Chloride-ion Permeability of Different Concretes",<br />

6th Intl Conf. on Durability of Building Materials Components, Omiya, Japan, 1171-1180<br />

(1993).<br />

13. Duval, R., Hornain, H., "La durabilité du béton vis-à-vis des eaix aggressoves", La<br />

Durabilité des Bétons sous la direction de Baron, J., Ollivier, J.-P., Presses de l'École<br />

Nationale des Ponts et Chaussées, Paris, ISBN 2-85978-184.6, 1992, pp. 376-385.<br />

14. Neville, A.M., "Properties of Concrete", 4th Edition, Longman, ISBN 0-582-23070-5,<br />

1995, pp. 514-517.<br />

15. Mehta, P.K., Monteiro, P., "Concrete - Microstructure, Properties, and Materials",<br />

McGraw Hill, ISBN 0-07-041344-4, 1993, pp. 167-176.<br />

16. Aïtcin, P.-C., "The Art and Science of High-Performance Concrete", American Ceramic<br />

Society, Cements Research Progress, 1996, Edited by L. Struble, ISBN 1-57498-081-5,<br />

1998, pp. 239-251<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 234 of 235


Cement<br />

Concrete<br />

17. Felicetti, R., Gambavora, P.G., Rosati, G.P., Corsi, F. and Giannuzzi, G., "Residual<br />

Strength of HSC Structural Elements Damaged by Hydrocarbon Fire or Impact Loading"<br />

in Utilization of High Strength High Performance Concrete, Paris, ISBN 2 85978 2583,<br />

579-588 (1996).<br />

18. Diederich, U., Spitzner, J., Sandvik, M., Kepp, B. and Gillen, M., "The Behavior of High-<br />

Strength Lightweight Aggregate Concrete at Elevated Temperature" in High Strength<br />

Concrete, ISBN 82 91341-00-1, 1046-1053 (1993).<br />

19. Jensen, J.J., Opheim, E. and Aune, R.B., "Residual Strength of HSC Structural<br />

Elements Damaged by Hydrocarbon Fire on Impact Loading" in Utilization of High<br />

Strength / High Performance Concrete, Paris, ISBN 2-85978-2583, 589-598 (1996).<br />

20. Noumowe, A.N., Clastres, P., Delvicki, G. and Costaz, J.-L., "Thermal Stresses and<br />

Water Vapour Pressure of High-Performance Concrete at High Temperature" in<br />

Utilization of High Strength / High Performance Concrete, Paris, ISBN 2-85978-2583,<br />

561-570 (1996).<br />

21. Sanjayan, G. and Stocks, L.J., "Spalling of High-Strength Silica Fume Concrete in Fire",<br />

ACI Mater. J. 90(2), March-April, 170-173 (1993).<br />

22. Chan, S.Y.N., Peng. G.F. and Chan, J.K.W., "Comparison Between High Strength<br />

Concrete and Normal Strength Concrete Subjected to High Temperature" Mat. Struct. 29<br />

616-619 (1996).<br />

23. Khoylou, N. and England, G.L., "The Effect of Elevated Temperature on the Moisture<br />

Migration and Spalling Behaviour of High Strength and Normal Concretes, ACI SP-167,<br />

263-268 (1996).<br />

24. Breitenbücker, R., "High Strength Concrete C105 with Increased Fire Resistance Due to<br />

Propylene Fibers" in Utilization of High Strength / High Performance Concrete, Paris,<br />

ISBN 2-85978-2583, 571-578 (1996).<br />

25. Jensen, B.C., Aarup, B., "Fire Resistance of Fibre Reinforced Silica Fume Based<br />

Concrete" in Utilization of High Strength / High Performance Concrete, Paris, ISBN 2-<br />

8597802583, 551-560 (1996).<br />

26. Phan, L.T., "Fire Performance of High-Strength Concrete: A Report of the State-of-the-<br />

Art", Res. Rep. NISTIR 5934, NIST, Gaithersburg, Maryland, U.S.A. (1997)<br />

27. Acker, P., Ulm, F.-J., Levy, M., "Fire in the Channel Tunnel - Mechanical Analysis of the<br />

Concrete Damage", Concrete Canada - Technology Transfer Day, October 1, Toronto,<br />

Ontario, 17 p. (1997).<br />

28. Demorieux, J.-M., "Le comportement des BHP à hautes températures — État de la<br />

question et résultats expérimentaux", École Française du Béton et le Projet National<br />

BHP 2000, November 24 and 25, Cachan, France, 27 p. (1998)<br />

29. Neck, U., Personal communication, March 1999.<br />

30. Cheyrezy, M., Beloul, M., "Comportement des BHP au feu", French Civil Engineering<br />

Association Meeting, 7 p., March 9, 1999.<br />

31. Kreijger, P.C., "Ecological properties of Building Materials," Mat. Struct. 20 248-254<br />

(1987).<br />

32. Richard, P., and Cheyrezy, M., "Reactive Powder Concrete With High Ductility and 200-<br />

800 MPa Compressive Strength," Am. Concr. Inst. SP-144 (Concrete Technology: Past,<br />

Present, and Future, edited by P.K. Mehta) 507-518 (1993).<br />

33. Bonneau, O., Poulin, C., Dugat, J., Richard, P., and Aïtcin, P.-C., "Reactive Powder<br />

Concrete From Theory to Practice," Concr. Int. 18 47-49 (1996).<br />

(1996).<br />

Course for Cement Applications - <strong>2005</strong> Concrete Section - Page 235 of 235

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!