Sunteți pe pagina 1din 9

Rheol Acta (2009) 48:447455

DOI 10.1007/s00397-008-0345-5

ORIGINAL CONTRIBUTION

Alkali activated fly ash: effect of admixtures


on paste rheology
M. Criado A. Palomo A. Fernndez-Jimnez
P. F. G. Banfill

Received: 27 June 2008 / Accepted: 24 December 2008 / Published online: 12 February 2009
Springer-Verlag 2009

Abstract In this paper, an investigation related to the


rheological behaviour of alkali-activated fly-ash pastes
(AAFA) is described. Those pastes were prepared by
mixing the fly ash with an alkaline dissolution containing 85% of a 12.5 M NaOH solution and 15% of
waterglass and adding some commercial chemical admixtures usually used in the Portland cement concrete
fabrication, like lignosulphonates, melamines (first and
second generation products) and polycarboxylates (latest generation). The fly ash rheological data were determined by rotational viscometry measurements as well
as by the use of the flow table test. Results indicate
that chemicals admixtures used do not work the same
in the Portland cement systems than in alkali-activated
fly ash systems. As a general rule, it seems that the most
efficient admixtures for these new cementitious pastes
(AAFA) are those based in polycarboxylates.

M. Criado (B)
CENIM (CSIC), Avda. Gregorio del Amo,
no 8, 28040, Madrid, Spain
e-mail: mcriado@cenim.csic.es
A. Palomo A. Fernndez-Jimnez
Instituto Eduardo Torroja (CSIC), Serrano Galvache,
no 4, 28033, Madrid, Spain
A. Palomo
e-mail: Palomo@ietcc.csic.es
A. Fernndez-Jimnez
e-mail: anafj@ietcc.csic.es
P. F. G. Banfill
School of the Built Environment, Heriot-Watt University,
Edinburgh, EH14 4AS, UK
e-mail: P.F.G.Banfill@sbe.hw.ac.uk

Keywords Fly ash Alkali activated Rheological


Admixtures

Introduction
Portland cement has been the construction material
par excellence for decades for its mechanical strength
and cost effectiveness, not to mention its properties in
general that make it particularly well suited to building.
That notwithstanding, possible alternative materials
are being sought and studies conducted in that regard,
prompted by environmental concerns in connection
with quarrying for raw materials, the enormous energy consumption involved in limestone decarbonation
and raw mix clinkerisation (very high temperatures,
1,500 C, with a nominal energy demand of 1,700 to
1,800 MJ/tonne clinker (Taylor 1997)) and, of course,
the emission of greenhouse gases into the atmosphere
(essentially CO2 and NOx : the manufacture of 1 tonne
of cement generates approximately 1 tonne of CO2 , at
a cost of 25/30e (Madridmasd 2006)). In light of these
problems, the scientific community has undertaken to
seek new processes, technologies and materials to provide the construction industry with alternative binders
like is cement obtained by alkali-activating fly ash.
The alkali activation of fly ash alters the vitreous
structure characteristic of most of the ash components to form compact cementitious skeletons (Palomo
et al. 1999, 2004; Lee and van Deventer 2002). The
outcome of this process, which takes place at pH >
12 and temperatures ranging from ambient to around
200 C, is an essentially alkaline aluminosilicate gel that
exhibits steadily increasing polymerisation with time
(Duxson et al. 2005; Fernndez-Jimnez and Palomo

448

Rheol Acta (2009) 48:447455

2005; Fernndez-Jimnez et al. 2006). This aluminosilicate exhibits long- and medium-range disorder, making
it amorphous to X-ray diffraction; but at the nanometric level, it is found to have a zeolite-type threedimensional structure (Palomo et al. 2004; Provis et al.
2005). The role of the alkaline element is to compensate
the negative charge resulting from the replacement of
Si4+ for Al3+ in the structural network.
According to the nuclear magnetic resonance data
on this material, silicon atoms occupy different positions in the three-dimensional structure, with a prevalence of Q4 (3Al) and Q4 (2Al) units Fernndez-Jimenez
et al. 2006. Aluminium coordination, in turn, is essentially tetrahedral (Fernndez-Jimnez and Palomo
2005; Fernndez-Jimnez et al. 2006). The product of
these reactions may in fact be considered to be a zeolite precursor (Palomo et al. 1999, 2004).
These materials are able to develop similar or even
greater mechanical strength than Portland cement mortars: compression strength of up to 50 or 60 MPa
just 20 h after mixing (Buchwald and Schulz 2005).
However, sight should not be lost of the enormous
differences between alkali activation of fly ashes and
Portland cement hydration. Such differences, both as
regard reaction mechanisms (Fernndez-Jimnez and
Palomo 2005; Fernndez-Jimnez et al. 2006; Banfill
et al. 1991) and the nature of the reaction products
formed (see Fig. 1), explain the differing behaviours
observed in alkaline cement compared with Portland
cement.

OH
Si

Ca

Ca

Ca

Ca

Ca

Ca

Ca

Ca

Ca

Ca

Ca

Si
O

Si

Si
OH H O

Gel 2
Si/Al 2
D4R y D6R

Si

Ca

Si

OH

Si

Si

Si
O

O
O

Si
O

HO

O
Si

Si

Si
O

HO

OH
Si
O

The rheological behaviour of Portland cement has


been studied by different authors during the last
decades (Tattersall and Banfill 1983; Grzeszczyck and
Kucharska 1987, 1990). It is well known that the Portland cement exhibits a rheological behaviour that can
be described by the Binghams model (Tattersall and
Banfill 1983; Banfill et al. 1991). In the bibliography,
numerous works about the effect of the commercial
admixtures (additives) on the rheological properties of
Portland cement can be found (Ferraris et al. 2001;
Lachemi et al. 2004; Perrot et al. 2007).
On the contrary, very few rheological studies have
been conducted on alkaline cement pastes. Palacios
et al. (2008), for instance, concluded that the nature of
the alkali activator used in slag pastes and mortars had
a significant impact on the rheology of these materials.
When the activating solution was sodium silicate, these
slag pastes and mortars fit a HerschelBulkley model,
with distinct structural breakdown, while when NaOH
was used as the activator, they behaved like Bingham
fluids. These same authors observed that the presence of naphthalene-based admixtures lowered shear
stress dramatically when added to NaOH-activated slag
pastes and mortars.
Working with alkali-activated fly ash systems, in
turn, Palomo et al. (2005) showed that information on
their reaction kinetics could be obtained by processing
the respective rheological data. Since, however, much
is still unknown about the rheological behaviour of
these new cementitious materials, the present study

OH

2-

SiO4

Al
O

HO

Sodium silicate
solution

AlO4

Na+

OH-

Fig. 1 a Schematic representation of a single layer in the crystal structure for a 14 tobermorite (based on Andersen et al. 2004) and
b proposed structure for the zeolite precursor formed during the alkali activation of fly ash (Criado 2007)

Rheol Acta (2009) 48:447455

449

Table 1 Chemical composition of the raw materials determined as described in Spanish standard UNE 80-230-99
Fly ash
Cement

SiO2

Al2 O3

Fe2 O3

CaO

MgO

SO3

Na2 O

K2 O

TiO2

LOI

IR

Total

53.09
20.03

24.80
6.18

8.01
2.26

2.44
62.50

1.94
0.18

0.23
2.81

0.73
0.27

3.78
0.73

1.07

3.59
2.55

0.32
0.74

100.00
98.28

LOI loss on ignition, IR insoluble residue

aims primarily to determine the effect of some commercial admixtures on the rheological properties of alkaliactivated fly-ash pastes and mortars.

these compounds are commercial products commonly


used in the manufacture of Portland cement concretes.
Methods
Flow table spread

Experimental procedure
Materials

Differential volumen (%/m)

A class F (as per ASTM standard C 618-03) fly ash


from the Compostilla steam power plant in Spain and
a commercial cement (CEM I 52.5 N/SR) were used in
the present study. The chemical composition of these
two raw materials is given in Table 1. The particle size
distribution of the raw materials is shown in Fig. 2
(Fernndez-Jimnez and Palomo 2003).
The activating solution used to make the working pastes was a mix containing 12.5 M NaOH solution (85%) + sodium silicate solution (15%) with a
SiO2 /Na2 O ratio of 0.16. The products used to prepare the solutions were laboratory reagents: 98% pure
NaOH pellets and sodium silicate with a density of
1.38 g/cm3 and the following composition: 8.2% Na2 O,
27% SiO2 , and 64.8% H2 O. Both reagents were supplied by Panreac S.A.
Finally, the main characteristics of the admixtures
used in the research are summarised in Table 2. All

Rotational viscometry study of the pastes

Cement
Fly ash

3
2
1
0
0.1

A control fly-ash mortar (without admixtures) was


prepared with the aforementioned activating solution.
The sand/fly ash ratio used in the mortars was 2:1.
Siliceous sand was employed throughout (SiO2 content
of 99%. 66% of particles with size <1 mm and 35% <
0.5 mm). A number of different fly-ash mortars were
subsequently prepared as above, but adding different
admixtures during the mixing process (0.8 g of admixture/100 g of fly ash). Cement samples were normally
hydrated (with no alkalis) but in identical conditions.
The standard flow table spread was determined as
laid down in Spanish standard UNE 80-116-86, i.e.,
setting the truncated cone mould (larger base at the
bottom) with its funnel in the centre of the table. A
first and then a second lift of mortar were placed in the
mould, consolidating each lift with ten rammer strokes.
After the funnel was withdrawn and the top levelled,
the mould was carefully removed, and the shock table was dropped 15 times in 15 s. The diameter of
the spread mortar was then measured (in millimetres)
in two perpendicular directions, averaging the values
found.

10

100

Particle diameter (m)

Fig. 2 Particle size distribution of the raw materials determined


by laser diffraction

Rotational viscometry was used in a second part of this


study on fly-ash paste rheology in highly alkaline media.
The Haake facility used consisted of a Roto Visco
1 viscometer, a Z38 DIN 53018 rotor with serrated
surface (Ri = 19.010 mm; h = 55 mm), a Z43 DIN 53018
flask, a temperature control unit for coaxial cylindrical
systems and a DC 30-B3 circulation thermostat.
The fly-ash pastes were made by mixing the fly ash
with the activating solutiona solution/ash ratio of
0.4 and 0.8 g of admixture per 100 g fly ash. Tests were
also run on control material without admixtures. Also,
Portland cement pastes were used as a reference system

450

Rheol Acta (2009) 48:447455

Table 2 Commercial admixture properties and composition


Name

Main characteristic

Composition

Manufacturer

Lig
Mel
Car

High rate water-reducing or admixture or fluidifier


Superfluidifier, accelerates hardening
Superplasticiser and high rate water-reducing admixture

Purified lignosulphonate
Melamine-derived synthetic polymers
Modified polycarboxylic ethers

GRACE
BETTOR
SIKA

(water/cement ratio = 0.4 and admixture proportion =


0 and 0.8).
The aim of the rheological study of all these pastes
was to observe the effect of the commercial admixtures
listed in Table 2 on the respective flow curves. The
methodology depicted in Fig. 3 was designed for this
purpose. The rotor spindle speeds were maintained for
1-min intervals at 50, 100, 150, 200, 250, 200, 150, 100
and 50 rpm. One hundred grams of fly ash, or cement,
were mixed with alkaline solution, or water, for 2 min.
Each paste was tested twice.

Fernndez-Jimnez et al. 2005) differ from Portland


cement hydration processes and products (Taylor 1993;
Richardson and Groves 1993).

Results

Flow curve

Flow table spread

The flow curves obtained in the tests described in Fig. 3


with an admixture content of 0.8% are reproduced in
Fig. 5. The shape of the curves found for the ash paste
systems tested, similar in all cases on both the ascending
(when the shear rate was increased) and descending
(when it was decreased) sides, indicates that these mortars are unaffected by the so-called structural breakdown typical of Portland cement pastes (Tattersall
and Banfill 1983; Banfill et al. 1991; Lapasin et al. 1983;
Stuble and Lei 1995).
Since the lower part of all the curves is nearly a
straight line (see Fig. 5), these systems may be regarded
to conform to the Bingham model: i.e., the paste flows
when a small initial force is applied.

The spread for fly-ash mortars with a solution/ash ratio


of 0.4, found using the procedure described in Spanish
standard UNE 80-116-86, was 102 mm (it is labelled Ref
in Fig. 4. The spread value was calculated as follows:
Spread = M* 100 (M*mean of the measurements
of two perpendicular diameters of the spread material).
Although the same experimental method was used,
these activated fly-ash systems should not be expected
to behave in exactly the same way as Portland cement
systems, inasmuch as the reaction mechanisms involved
and the nature of the reaction products formed in the
alkali activation of fly ash (Lee and van Deventer 2002;

Speed (r.p.m.)

250
200

200

Effect of admixtures
As the graphs in Fig. 4 clearly show, each type of
admixture affected fly ash differently. Where (latest
generation) admixture Car, for instance, appeared to
be an effective superplasticiser, Lig and Mel (first and
second generation products) did not.

Car

200
Mel

150

150
100

100

100

Lig

50

50

Ref

0
0

3 4

5 6

Time (min.)
Fig. 3 Viscometer test

50

100

Spread (mm)

Fig. 4 Flow table spread in fly-ash mortar

150

Rheol Acta (2009) 48:447455


70

Discussion

Ref
Lig
Mel
Car

60

40

v v

50
v

30

v v
v

Shear stress (Pa)

451

20
10
50

100

150

200

250

Speed (r.p.m.)

Fig. 5 Flow curves

Determination of plastic viscosity and yield stress


As noted, the lower part of the flow curves in Fig. 5 can
be fitted to a straight line corresponding to the equation proposed by Bingham to describe the rheological
behaviour of Portland cement pastes (Eq. 1):
= 0 +

(1)

Where is the shear stress applied to the paste, is


the shear rate, represents the plastic viscosity of the
system at any given time and 0 is the yield stress.
Figure 6a shows the plastic viscosity values for the
systems tested and Fig. 6b the yield stress for those
systems. The plastic viscosity obtained for all the flyash systems with admixtures was smaller than the value
found for the control paste (without admixtures); in
other words, the effect of the admixtures was to make
the pastes more fluid and hence less viscous.
The yield stress, in turn, was lower in the pastes with
Mel and Car admixtures than in the control. Due to
segregation of the paste containing the Lig admixture,
the yield stress values were very small and possibly not
representative of the actual stress.

Fig. 6 a Plastic viscosity; b


yield stress of alkali-activated
fly-ash pastes with and
without admixtures with a
liquid/solid ratio of 0.4

The alkali activation of fly ash yields a material


with excellent cementitious properties (Buchwald and
Schulz 2005; Fernndez-Jimnez and Palomo 2003),
but sight should not be lost of the enormous differences between alkali activation and Portland cement
hydration. Such differences, both as regard reaction
mechanisms (Fernndez-Jimnez and Palomo 2005;
Fernndez-Jimnez et al. 2006; Banfill et al. 1991) and
the nature of the reaction products formed (see Fig. 1),
explain the differing rheological behaviour observed
in fly ash compared with Portland cement mortars;
and they should also explain the variable effects of
the commercial admixtures used in this study, which
were designed for use with Portland cement, on fly ash
systems.
Organic admixtures based on sulphonic acids, for instance (such as Mel and Lig), contain both a hydrophilic
(sulphonic) and a hydrophobic (lignin and the naphthalene rings) group. The electrical charge resulting from
the adsorption of these compounds on the surface of
cement particles generates the repulsive forces that increase paste fluidity (Hanehara and Yamada 1999). The
lateral chains in admixtures based on polycarboxylates
(Car admixture) generate steric repulsion among the
cement grains, causing deflocculation and dispersion
of the particles, which in turn improves workability
by releasing the water trapped in the flocs (Yamada
et al. 2000; Palacios 2006, see Fig. 7). In any event,
most admixtures are designed to form complexes with
the dissolved Ca2+ ions found in the early phases of
Portland cement hydration. Uchikawa et al. (1995),
for instance, concluded that in the presence of admixtures based on aminosulphonic, naphthalenesulphonic
or lignosulphonic acids, dissolved calcium ions near the
surface of clinker particles bond to those compounds
to form calcium sulphonate. The molecules of polycar-

(a)

Car

Mel

Mel

Lig

Lig

Ref

Ref

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Plastic viscosity (Pa.s)

3.5

4.0

(b)

Car

0.0

0.5

1.0

1.5

2.0

2.5

Yield stress (Pa)

3.0

3.5

4.0

452

Rheol Acta (2009) 48:447455

boxylic admixtures also form complexes with Ca2+ ions.


Attention should be drawn in this regard to the near
absence of dissolved Ca2+ ions in fly ash systems.
To better explain the effect of admixtures on the
rheological parameters of fly-ash pastes, then, a parallel
experiment with Portland cement paste was likewise
conducted.
Effect of admixtures on the early phases of Portland
cement and activated fly-ash pastes
Figure 8 compares the basic rheological parameters
(plastic viscosity and yield stress) of fly ash and Portland cement systems.
Plastic viscosity
In all the systems studied (in the absence or presence
of admixtures), plastic viscosity was observed to be
much greater in fly ash than in the Portland cement
pastes (Fig. 8a). Some of the reasons for this difference
may be: (a) the physical difference in the particles
of the starting materials (cement particles are parallelepipeds with many edges whereas fly ash particles
are spherical); (b) the particle size distribution; (c) the
different reaction mechanisms that lead to the formation of different reaction products (Portland cement
reacts by forming hydrates with short silicate chains
(C-S-H gel) that fill voids, whereas alkali activation
results in a cross-linked aluminosilicate gel (N-A-S-H
gel); (d) kinetic considerations (the activating solution
used comprises primarily soluble, monomeric silicates
(Criado et al. 2008) and
 react very quickly
 monomers
released during fly
with the aluminates A1 (OH)
4
ash dissolution. The result is the rapid precipitation of
an initial aluminium-rich alkaline aluminosilicate gel

+
-

Electrostatic
repulsion

+
+ Cement
particle +
+

+
+

+ Cement
particle +

Steric
repulsion

Superplasticisers

Fig. 7 Steric repulsion and electrostatic exercised by the superplasticisers admixtures based on polycarboxylates between
cement grains (based on Palacios 2006)

(Fernndez-Jimnez et al. 2006) that prompts paste


setting and hardening.
The plastic viscosity obtained for all the fly-ash systems with admixtures was slightly smaller than the
value found for the paste without admixtures (see
Fig. 6). All the admixtures appeared to serve their plasticising purpose in the early phases (time < 10 min.),
forming more fluid pastes whose plastic viscosity is
therefore lower.
The plasticising properties of such admixtures (particularly the Lig and Mel type) vanished in a relatively
short period of time due to the instability of these compounds in very basic media. In a study of activated slag
systems, Palacios and Puertas (2004, 2005) reported
alterations in the infrared spectra of a naphthalenederived admixture when stored in a sodium silicate +
sodium hydroxide solution. Such alterations were attributed to certain changes in the sulphonates, the
functional groups responsible for the superplasticising
effect in this sort of admixture. In conclusion, this loss
of stability in the admixture partially explains their nil
or even adverse effect over time (see Fig. 5). Although
these authors failed to provide a detailed description
of the structural changes taking place in the admixture,
the sulphonate groups may well have been modified or
even released. In any event, further study in this area is
required.
Similar results have been observed by other authors
for alkali-activated slag. Bakharev et al. (2000) and
Douglas and Brandstetr (1990) concluded that the presence of a naphthalene-based admixture lowered compressive strength without improving alkaline system
workability.
Yield stress
Figure 8b shows the yield stress developing in alkaliactivated fly ash and Portland cement pastes under identical conditions. The higher values found for
Portland cement pastes without admixtures than for ash
pastes may be attributed to the fact that the former
undergo so-called structural breakdown; in this case,
greater force needs to be applied to the cement paste
particles (which tend to agglomerate) to induce deflocculation.
The inclusion of admixtures in Portland cement
systems prompts different effects on yield stress. The
Car admixture is of particular interest in this regard.
Its differential behaviour may be due to the fact that
polycarboxylate admixtures act in two ways, simultaneously: if adsorbed on to cement particles, the resulting
highly negative electric charge originates inter-particle
repulsion, while at the same time the many lateral

Rheol Acta (2009) 48:447455


Fig. 8 a Plastic viscosity and
b yield stress of fly ash and
Portland cement systems

453

(a)

Car

Mel

Mel

Lig

Lig

Cement
Fly ash

Ref
0

Plastic viscosity (Pa.s)

chains in the admixture may generate steric repulsion.


This dual action makes such compounds generally more
effective than previous generation admixtures (Lig and
Mel).
The yield stress values obtained for all alkali activated fly-ash systems with admixtures were usually
smaller or similar to the value found for the paste
alone (see Fig. 7b). The admixtures appear to serve
their superplasticising purpose initially, when the interactions between paste flocs are weak, the particles
are dispersed and therefore slightly smaller forces are
required for the pastes to flow. Due to the segregation
taking place in pastes containing the Lig admixture,
the particles constituting the paste separate, forming a
denser lower phase and a more liquid upper phase. As a
result, the yield-stress values obtained are not reliable.
Comparison of the rheological data found
with rotational viscometry and the empirical
flow table spread
The correlation between spread and the rheological
parameters for activated ash pastes shown in Fig. 9
Reference axe

Car
plastic viscosity
yield stress
spread

Mel

Lig

no data

decrease

increase

Fig. 9 Correlation between flow and rheological parameters.


Ref: reference, values obtained for ash without admixtures

(b)

Car

Cement
Fly ash

Ref

10

15

20

25

Yield stress (Pa)

necessitated the establishment of an arbitrary scale to


compare the flow table test data (in millimetres) to
the rheological parameters (in Pascalsecond for plastic
viscosity and Pascal for yield stress). The values obtained for plastic viscosity, shear stress and slump in fly
ash with no admixtures were taken as reference or 0
values; the remaining data were obtained on the basis
of these values and expressed in percent.
The literature contains numerous studies on the effect of different admixtures on Portland cement paste
rheology. In the general pattern observed in these systems, the presence of admixtures induces a decrease in
plastic viscosity and yield stress, which normally goes
hand-in-hand with higher flow values (Ferraris et al.
2001; Lachemi et al. 2004). The behaviour observed in
fly-ash pastes does not follow that pattern (see Fig. 9).
A drop in plastic viscosity or yield stress is not always
found in conjunction with greater flow (such behaviour,
rather, depends on the admixture used).
Consequently, a prior knowledge of the information
provided by the techniques used (rotational viscometer
and slump) is requisite to understanding why this system behaved differently from Portland cement pastes.
The viscosity (Pascalsecond) and threshold shear
stress (Pascal) of plastic state pastes can be obtained
with a computer-controlled co-axial cylinder rotational
viscometer. The underlying principle consists in applying a constant spinning speed and measuring the
strength (torque) exerted by the fluid against the rotor.
Slump, however, is found by an empirical test developed as a method for the simple and practical
evaluation of Portland cement paste workability. The
normal slump test consists in measuring two perpendicular diameters of the fluid mass placed on a flow table
(Spanish standard UNE 80-116-86).
Because the slump test is quasi-static, it is related to
the flow resistance (and yield stress) which governs the
static flow behaviour of the material. Different relationships between slump and 0 have been proposed by

454

many researchers (Tattersall and Banfill 1983; Beaupr


and Mindess 1998). However, these authors have also
mentioned that there is no relationship between slump
and viscosity. Theoretical simulations assuming Bingham behaviour have shown that the effect on slump
of a change in yield value is much greater than that
of a comparable change in plastic viscosity: the slump
is affected by the yield, but it is not affected by the
viscosity (Wallevik 2006).
The present findings also reveal the low correlation
between slump and plastic viscosity. Moreover, no linear correlations between slump and threshold shear
such as reported by Wallevik (2006) were observed. In
these fly-ash systems, the relationship between slump
and threshold shear would appear to depend less on
the volume fraction of the material. Nonetheless, the
rheological parameters and slump in alkali-activated fly
ash depend on the admixture used, as described below,
but this dependence differs widely in nature from the
dependence found in Portland cements.
Admixture Lig (lignosulphonate base) exhibits
anomalous behaviour, attributed in this study to the
segregation taking place in the pastes with which it is
mixed. Paste segregation consists in particle separation,
in which a denser phase is overlain by a more liquid
phase.
Admixture Mel does not appear to be very effective, for while its superplasticising function prompts an
instantaneous decrease in plastic viscosity and shear
stress, this has no clear effect on the flow table spread.
This may be due to the instability of such admixtures
in media with a high pH, such as in the present study
(Palacios and Puertas 2004; Puertas et al. 2003, the loss
of their fluidifying properties stemming from structural
alterations).
Admixture Car (with a polycarboxylate base) may
also be chemically unstable, but since the numerous
side chains in its structure generate steric repulsion that
offsets the tendency of particles to form complexes,
they may retain some plasticising properties.

Conclusions
The chief conclusions to be drawn from the present
study are as follows:
1. The superplasticising effectiveness of commercial
admixtures and their impact on the rheological
properties of alkali-activated fly-ash pastes and
mortars depends on their stability in highly alkaline media. Latest generation admixtures, in which

Rheol Acta (2009) 48:447455

inter-particle electrostatic repulsion prevails over


complex formation, are the most effective.
2. Viscometer measurements are a more precise
method for studying the effect of admixtures on
fly-ash paste rheological behaviour than the flow
table test, inasmuch as they provide more detailed
information. The flow table test, in fact, furnishes
information that is practical and useful for Portland
cement paste only.
Acknowledgements Funding for this research was provided by
the Directorate General of Scientific Research under project
BIA2006-28530-E; also, a post-doctoral contract associated with
the study was awarded by the CSIC and co-financed by the
European social fund (REF. I3P-PC2004L).

References
Andersen MD, Jakobsen HJ, Skibsted J (2004) Characterization
of white Portland cement hydration and the C-S-H structure
in presence of sodium aluminate by 27 Al and 29 Si MAS
NMR spectroscopy. Cem Concr Res 34:857868
Bakharev T, Sanjayan JG, Cheng YB (2000) Effect of admixtures
on properties of alkali-activated slag concrete. Cem Concr
Res 30:13671374
Banfill PFG, Carter RE, Weaver PJ (1991) Simultaneous rheological and kinetic measurements on cement pastes. Cem
Concr Res 21:11481154
Beaupr D, Mindess S (1998) Rheology of fresh concrete: principles, measurement, and applications. In: Skalny J (ed) Materials science of concrete, vol V. American Ceramic Society,
Westerville, p 149. ISBN: 1-57498-027-0
Buchwald A, Schulz M (2005) Alkali-activated binders by use of
industrial by-products. Cem Concr Res 35:968973
Criado M (2007) Nuevos materiales cementantes basados en la
activacin alcalina de cenizas volantes. Caracterizacin de
geles N-A-S-H en function del contenido de slice soluble.
Efecto del Na2 SO4 . PhD thesis, Universidad Autnoma de
Madrid, Spain
Criado M, Fernndez-Jimnez A, Palomo A, Sobrados I, Sanz J
(2008) Effect of the SiO2 /Na2 O ratio on the alkali activation
of fly ash. Part II: 29 Si MAS-NMR survey. Microp Mesop
Mat 109:525534
Douglas E, Brandstetr J (1990) A preliminary study on the alkali activation of ground granulated blast-furnace slag. Cem
Concr Res 20:746756
Duxson P, Provis JL, Lukey GC, Separovic F, van Deventer JSJ
(2005) 29 Si NMR study of structural ordering in aluminosilicate geopolymer gels. Langmuir 21:30283036
Fernndez-Jimnez A, Palomo A (2003) Characterisation of
fly ashes. Potential reactivity as alkaline cements. Fuel 82:
22592265
Fernndez-Jimnez A, Palomo A (2005) MID-infrared spectroscopic studies of alkali-activated fly ash structure. Microp
Mesop Mat 86:207214
Fernndez-Jimnez A, Palomo A, Alonso MM (2005) In: Bilek
V, Kersner Z (eds) 2nd inter. symposium non-traditional cement & concrete, Brno University of Technology, p 1. ISBN:
80-214-2853-8
Fernndez-Jimnez A, Palomo A, Sobrados I, Sanz J (2006) The
role played by the reactive alumina content in the alkaline
activation of fly ashes. Microp Mesop Mat 91:111119

Rheol Acta (2009) 48:447455


Ferraris CF, Obla KH, Hill R (2001) The influence of mineral
admixtures on the rheology of cement paste and concrete.
Cem Concr Res 31:245255
Grzeszczyck S, Kucharska L (1987) Rheological properties of
fresh cement pastes and clinker reactivit. Rheol Acta 26:
566569
Grzeszczyck S, Kucharska L (1990) Hydrative reactivity of cement and rheological properties of fresh cement pastes. Cem
Concr Res 20:165174
Hanehara S, Yamada K (1999) Interaction between cement and
chemical admixture from the point of cement hydration, adsorption behaviour of admixture, and paste rheology. Cem
Concr Res 29:11591165
Lachemi M, Hossain KMA, Lambros V, Nkinamubanzi PC,
Bouzouba N (2004) Performance of new viscosity modifying admixtures in enhancing the rheological properties of
cement paste. Cem Concr Res 34:185193
Lapasin R, Papo A, Rajgelj S (1983) The phenomenological description of the tixotropic behaviour of fresh cement pastes.
Rheol Acta 22:410416
Lee WKW, van Deventer JSJ (2002) Effects of anions on the
formation of aluminosilicate gel in geopolymers. Ind Eng
Chem Res 41:45504558
Madridmasd (2006) http://www.madridmasd.org/informationIDI/
noticias
Palacios M (2006) Efecto de aditivos orgnicos en las
propiedades de los cementos y morteros de escoria activada alcalinamente. PhD thesis. Universidad Autnoma de
Madrid, Spain
Palacios M, Puertas F (2004) Stability of superplasticiser and
shrinkage-reducing admixtures in high basic media. Mater
Construcc 54:6586
Palacios M, Puertas F (2005) Effect of superplasticiser and
shrinkage-reducing admixtures on alkali-activated slag
pastes and mortars. Cem Concr Res 35:13581367
Palacios M, Banfill PFG, Puertas F (2008) Rheology and setting
of alkali-activated slag pastes and mortars: effect of organic
admixture. ACI Mater J 105:140148

455
Palomo A, Grutzeck MW, Blanco MT (1999) Alkali-activated fly
ashes. A cement for the future. Cem Concr Res 29:1323
1329
Palomo A, Alonso S, Fernndez-Jimnez A, Sobrados I, Sanz J
(2004) Alkaline activation of fly ashes. A NMR study of the
reaction products. J Am Ceram Soc 87:11411145
Palomo A, Banfill PFG, Fernndez-Jimnez A, Swift DS (2005)
Properties of alkali-activated fly ashes determined from rheological measurements. Adv Cem Res 17:143151
Perrot A, Lanos C, Meligne Y, Estell P (2007) Mortar physical properties evolution in extrusion flow. Rheol. Acta 46:
10651073
Provis JL, Duxson P, Lukey GC, van Deventer JSJ (2005) Statistical thermodynamic model for Si/Al ordering in amorphous
aluminosilicates. Chem Mater 17:29762986
Puertas F, Palomo A, Fernndez-Jimnez A, Izquierdo JD,
Granizo ML (2003) Effect of superplasticisers on the behaviour and properties of alkaline cements. Adv Cem Res
15:2328
Richardson IG, Groves GW (1993) The incorporation of minor
and trace elements into Calcium Silicate Hydrate (C-S-H)
gel in hardened cement pastes. Cem Concr Res 23:131138
Stuble LJ, Lei W-G (1995) Rheological Changes Associated with
Setting of Cement Paste. Adv Cem Bas Mat 2:224230
Tattersall GH, Banfill PFG (1983) The Rheology of Fresh Concrete. Pitman, London
Taylor HFW (1993) Nanostructure of C-S-H: current status. Adv
Cem Bas Mat 1:3846
Taylor HFW (1997) Cement chemistry. Thomas Telford, London
Uchikawa H, Sawaki D, Hanehara S (1995) Influence of kind
and added timing of organic admixture on the composition,
structure and property of fresh cement paste. Cem Concr
Res 25:353364
Yamada K, Takahashi T, Hanehara S, Matsurisa M (2000) Effects of the chemical structure on the properties of polycarboxylate-type superplasticiser. Cem Concr Res 30:197207
Wallevik JE (2006) Relationship between the Bingham parameters and slump. Cem Concr Res 36:12141221

S-ar putea să vă placă și