Вы находитесь на странице: 1из 13

Cement and Concrete Research 52 (2013) 169–181

Contents lists available at ScienceDirect

Cement and Concrete Research


journal homepage: http://ees.elsevier.com/CEMCON/default.asp

Effect of temperature on the hydration of Portland cement blended with


siliceous fly ash
Florian Deschner a,⁎, Barbara Lothenbach a, Frank Winnefeld a, Jürgen Neubauer b
a
Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory for Concrete and Construction Chemistry, Überlandstrasse 129, 8600 Dübendorf, Switzerland
b
GeoZentrum Nordbayern, Mineralogy, University of Erlangen-Nuremberg, 91054 Erlangen, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The effect of temperature on the hydration of Portland cement pastes blended with 50 wt.% of siliceous fly ash is
Received 14 March 2013 investigated within a temperature range of 7 to 80 °C.
Accepted 9 July 2013 The elevation of temperature accelerates both the hydration of OPC and fly ash. Due to the enhanced pozzolanic
reaction of the fly ash, the change of the composition of the C–S–H and the pore solution towards lower Ca and
Keywords:
higher Al and Si concentrations is shifted towards earlier hydration times. Above 50 °C, the reaction of fly ash also
Fly ash (D)
Blended cement (D)
contributes to the formation of siliceous hydrogarnet. At 80 °C, ettringite and AFm are destabilised and the re-
Temperature (A) leased sulphate is partially incorporated into the C–S–H. The observed changes of the phase assemblage in depen-
Hydration (A) dence of the temperature are confirmed by thermodynamic modelling.
Modelling (E) The increasingly heterogeneous microstructure at elevated temperatures shows an increased density of the C–S–H
and a higher coarse porosity.
© 2013 Elsevier Ltd. All rights reserved.

1. Introduction gel-porosity [13–15] and a cement paste with a more even distribution
of the hydrate phases. Hence, the coarse porosity is reduced and the
Cementitious materials are exposed to a vast range of temperatures compressive strength increased [1,7].
due to factors like local climatic conditions, externally applied heat in Furthermore, the chemistry of the pore solution is affected by the
precast concrete or heat due to the exothermic nature of the hydration temperature due to the change of the solubility of the hydrate phases.
process. Several studies investigated the impact of temperature on the For example, the increase of the temperature in neat OPC leads to a
hydration kinetics, the hydrate assemblage, microstructure and durabil- higher solubility of ettringite and thus to an increase of the sulphate
ity of ordinary Portland cement (OPC) in a temperature range between concentrations in the pore solution [3,4,16,17].
5 and 60 °C [1–8]. In accordance with the Arrhenius law, the augmenta- Although the effect of temperature on the hydration of neat OPC is
tion of temperature leads to the increase of the hydration rate of OPC well studied, there are few studies about fly ash blended OPC at differ-
and therefore to increased early compressive strength [1]. The temper- ent temperatures [11,15]. The present study shows a consistent set of
ature sensitivity of the hydration rate is indicated by the apparent acti- data of the hydration of a Portland cement containing 50 wt.% of sili-
vation energy, which was determined for OPC and cements blended ceous fly ash between 7 and 80 °C. The hydrate phase assemblage, the
with fly ash, slag and other supplementary cementing materials [9,10]. composition of the C–S–H, the development of the microstructure and
At later hydration times, the rate of strength development diminishes, the composition of the pore solution at hydration times between 1
so that after a certain time span, the compressive strength at high tem- and 180 days were investigated. The results are compared to the analy-
peratures is lower than for samples hydrated at lower temperatures. sis of reference samples containing 50 wt.% of quartz powder instead of
This effect is related to the microstructure development of the cement fly ash to allow the observation of the effect of temperature on the reac-
paste [1–3,5,8,11]. The rapidly forming initial hydration products at tion of fly ash.
elevated temperatures are distributed heterogeneously, the density of
the C–S–H is increased and at above 47 °C, ettringite and monocar- 2. Materials and methods
bonate are destabilised to monosulphate and calcite. This leads to a
higher coarse porosity and reduced compressive strength [2,3,12,13]. 2.1. Raw materials
At lower temperatures, the hydration rate is reduced allowing the
dissolved ions more time for diffusion before the hydrates precipitate. An OPC (CEM I 42.5 N) and a siliceous fly ash (type V according to EN
This leads to the formation of a less polymerised C–S–H with higher 197-1) have been used as raw materials. The chemical and mineralogi-
cal compositions of these materials were determined by means of X-ray
⁎ Corresponding author. Tel.: +41 58 765 4535; fax: +41 58 765 4035. fluorescence (XRF) and quantitative X-ray diffraction [18] (Table 1,
E-mail address: florian.deschner@gmail.com (F. Deschner). reproduced from [19]). The average composition of the fly ash glass

0008-8846/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.cemconres.2013.07.006
170 F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181

Table 1
Composition and density of the raw materials.

XRF-analysisa [wt.%] Mineralogical phase compositionb [wt.%] Average composition of


the fly ash glass [wt.%]

FA OPC FA OPC FA

SiO2 50.9 19.4 Mullite 8.2 C3S 57.1 SiO2 54.1


Al2O3 24.7 5.2 Quartz 7.0 β C2S 7.9 Al2O3 23.4
Fe2O3 7.3 3.6 Hematite 0.7 α′ C2S 9.3 Fe2O3 7.1
CaO 3.7 62.1 Magnetite 0.8 C4AF 13.0 CaO 4.6
MgO 1.8 1.8 Amorphous 83.3 C3A cubic 2.0 MgO 2.2
K2O 3.9 1.2 C3A orthorh. 2.0 K2O 5.0
Na2O 0.9 0.3 Calcite 0.4 Na2O 1.1
TiO2 1.1 0.2 Periclase 1.0 TiO2 1.4
Mn2O3 0.1 0.1 Bassanite 2.6 Mn2O3 0.1
P2O5 0.8 0.6 Anhydrite 2.8 P2O5 0.9
SO3 0.4c 3.7 Arcanite 0.5 SO3 b0.1
SrO 0.3 Dolomite 0.5
LOI 3.5 1.1 Magnesite 0.3
C 2.7d Siderite 0.2
Quartz 0.2
Ankerite 0.2
Sum 99.1 99.5 100 100 100
a
Standard deviation of the XRF analyses is b0.1 wt.% for all oxides except SiO2 (0.5 wt.%).
b
Standard deviation ranges from 0.1 to 0.6 wt.%.
c
SO3 content determined by combustion analysis.
d
C-content (part of the LOI) determined by combustion analysis.

was calculated by subtracting the crystalline phases and the readily sol- 2.4. X-ray diffraction (XRD) measurements
uble alkalis from the total composition of the fly ash and normalizing
the results to 100 wt.%. As reference material, a quartz powder, which 5 mm thick discs of the hardened paste samples with a diameter of
showed no pozzolanic activity at 23 °C up to 90 days of hydration was 30 mm were sliced from the cast samples, immersed in isopropanol
used [19]. The particle size distribution of the quartz powder is similar for 2 days to stop the hydration and stored under N2-atmosphere
to that of the fly ash (graph shown in [19]). until measurement. Measurements were performed in a PANalytical
X'Pert Pro MPD diffractometer with attached X'Celerator detector.
2.2. Preparation of cement pastes
2.5. Analysis of the pore solution
All raw materials and mixing tools used for the preparation of the ce-
The pore solutions of the hardened samples were extracted by the
ment pastes were equilibrated at the respective temperature for 1 day
steel die method [21] using pressures up to 250 MPa. The solutions
before use. Cement paste samples containing 50 wt.% of OPC and
were filtered immediately with nylon filters with a mesh size of
50 wt.% of fly ash (OPC-FA) or 50 wt.% of quartz powder (OPC-Qz), re-
0.45 μm. The free OH− concentrations of the pore solutions were calcu-
spectively, were prepared according to EN 196-3 with a water-to-
lated from pH measurements with a pH electrode, calibrated against
binder ratio (w/b) of 0.5. The samples were cast in 500 ml and 60 ml
KOH solutions with known concentrations. The K, Na, Ca, Al, Si and sul-
polyethylene flasks, sealed and stored at 7, 23, 40, 50 and 80 °C. The
phur concentrations were measured by means of inductively coupled
500 ml samples were used for the extraction of the pore solution and
plasma optical emission spectroscopy (ICP-OES). The standard devia-
the 60 ml samples for the analysis of the solid phases. The solid phases
tion of the measured elemental concentrations ranges between 5 and
were analysed by means of TGA in N2-atmosphere, XRD, and scanning
10%.
electron microscopy (SEM).
2.6. Scanning electron microscopy (SEM)
2.3. Thermogravimetric analyses (TGA)
The hydration was stopped by cutting the sample into slices of 5 mm
The hydrated samples were crushed, immersed in isopropanol for thickness with a diameter of 30 mm and keeping them for 3 days in
15 min, washed with diethylether, and filtered. Afterwards, the samples isopropanol and drying them for 3 days at 40 °C. Subsequently, the
were dried for 10 min at 40 °C to evaporate the remaining diethylether samples were impregnated with a modified bisphenol-A-epoxy-resin
and analysed in a Mettler Toledo TGA/SDTA 851e device at a heating and polished by polycrystalline diamond suspension at grades from
rate of 20 K/min in N2-atmosphere. The analysed bound water of all 9 μm down to 1/4 μm. Finally, the samples were carbon-coated and in-
hydrate phases (H) and the crystal water within Ca(OH)2 (HCH) were vestigated under high vacuum conditions in a Philips ESEM FEG XL 30.
determined by the weight loss in the temperature intervals 50–500 °C EDX measurements of the hydrate phases were carried out at a beam
and 400–470 °C. The exact boundaries for the temperature interval of voltage of 10 keV and 10 mm working distance. The content of coarse
portlandite were read from the derivative curve (DTG). The results are pores (pore size N 0.6 μm) was determined by means of image analysis
expressed as percentage of the dry sample weight at 500 °C [20]. Triple of 600 images per sample at magnification of 2000. The upper threshold
preparation and measurement of the samples after 1, 2 and 7 days of for the grey level segmentation of pores was determined by the mini-
hydration at 23 °C shows an absolute error of up to 2 wt.% for the mum point between the peak of pores and the brighter hydration prod-
bound water, and up to 0.6 wt.% for the water loss related to the decom- ucts peak in the histogram of the sum of all 600 images [22,23]. The
position of Ca(OH)2. This includes the errors caused by preparation and error of the quantification of the coarse porosity is estimated to be
measurement. ± 3%.
F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181 171

2.7. Compressive strength of mortar samples strength is higher than at lower temperatures. Afterwards however, the
rate of strength increase at these high temperatures is lower, and there-
Mortar prisms (40 × 40 × 160 mm3) were prepared according to fore the late strength values are lower than at 40 °C. At 80 °C, this effect
EN 196-1 and cured at 7, 20, 40, 50 and 80 °C. After demoulding, two is pronounced: the high early strength of 21.8 MPa after 1 day of hydra-
mortar prisms per testing age were stored inside sealed boxes filled tion is followed by a slow increase of strength resulting in 36.8 MPa
with distilled water at the appropriate temperature until testing. The after 180 days. In contrast to OPC-Qz, no relevant increase of the com-
compressive strength was tested according to EN 196‐1. pressive strength is found after 7 days of hydration at 80 °C.
In comparison to the reference sample OPC-Qz (Fig. 1a), the com-
2.8. Thermodynamic modelling pressive strength of the fly ash blended cement pastes at 20 °C is clearly
higher after 28 days. This additional strength is related to the pozzolanic
The stable phase assemblage of the investigated cementitious sys- reaction of the fly ash, while the quartz powder in the reference shows
tems at thermodynamic equilibrium was calculated in dependence of no significant pozzolanic reactivity at this temperature [31]. From 7 to
the temperature. The calculations were performed with the help of 50 °C, the rise of compressive strength in OPC-FA compared to OPC-Qz
the Gibbs free energy minimisation programme GEMS [24,25] together is shifting to earlier hydration times and thereby indicating a faster hy-
with the thermodynamic data from the PSI-GEMS database [26] ex- dration of fly ash at higher temperatures.
panded with additional data for solids that are expected to form under
cementitious conditions [4,27–29]. For the sake of simplicity, the crys- 3.2. Content of portlandite and bound water
talline phases in the fly ash were considered as inert and the glass
phase of the fly ash was assumed to dissolve uniformly. Fig. 2a–e shows the portlandite contents (CH) and bound water of all
GEMS was also used to calculate the activities of the aqueous species hydrate phases other than portlandite (H*), determined by means of
in the investigated pore solutions by the extended Debye–Hueckel TGA. H* is calculated by subtracting the content of crystal water of
equation [30]. From the activities, the saturation indices (SI) of the portlandite in the sample from the amount of total bound water (H).
solid phases were calculated. The SI with respect to the solids are For each temperature CH and H* in OPC-FA is compared to the reference
given by log(IAP/KS0), where KS0 is the theoretical solubility product sample OPC-Qz. The beginning of the portlandite consumption related
of the respective solid, while IAP is the ion activity product calculated to the pozzolanic reaction is indicated by the relative decrease of the
from the measured concentrations. A positive SI implies oversaturation, portlandite content in OPC-FA compared to OPC-Qz. It shows that the
a negative value undersaturation. As the use of SI can be misleading start of the pozzolanic reaction is shifting to earlier hydration times at el-
when comparing phases which dissociate into a different number of evated temperatures. The onset of portlandite consumption in OPC-FA is
ions, effective SI were calculated by dividing the SI by the number of found at 7 °C after 90 days, at 23 °C after 7 days, at 40 °C after 1 day, at
ions participating in the reactions to form the solids, as described by 50 °C after 16 h, and at 80 °C before 16 h of hydration. The difference
Lothenbach et al. [4]. between H* in OPC-FA and H* in OPC-Qz is related to the amount of hy-
dration products formed by the pozzolanic reaction of fly ash. The point
3. Results and discussion at which H* of OPC-FA is starting to be higher than H* in OPC-Qz is
shifting towards earlier hydration times at elevated temperatures and
3.1. Compressive strength correlates well with the beginning decrease of portlandite.
The decrease of the portlandite content in the OPC-Qz samples ob-
Fig. 1a shows a positive influence of elevated temperature on the served at 40, 50 and 80 °C indicates a pozzolanic reaction of the quartz
early compressive strength of the reference sample OPC-Qz. After in the reference samples. The rate of the quartz reaction increases
28 days, the OPC-Qz samples hydrated at 7 to 40 °C show similar with temperature. Surprisingly at 80 °C, the reaction of the quartz con-
strength. At 50 °C, a late increase of strength is observed compared to tinues while the fly ash reacts much slower after the first hours. This
OPC-Qz at 40 °C. And at 80 °C, a strong increase of compressive strength could be related to a hindrance of the further fly ash reaction at the
from 23.4 to 71.9 MPa is detected between 7 and 180 days. This effect is high dissolved Al concentrations observed in these samples.
probably related to the temperature induced pozzolanic reaction of the
quartz powder. 3.3. Characterization of hydration products
The development of the compressive strength of OPC-FA between 1
and 180 days is shown in Fig. 1b. Between 7 and 40 °C, the compressive 3.3.1. X-ray diffraction (XRD)
strength shows a positive correlation with the temperature throughout Figs. 3 and 4 show details of the XRD patterns of the samples after 1
the whole hydration time. After 1 day of hydration at 50 and 80 °C, the and 180 days of hydration. In addition to anhydrous Portland cement

70 a) OPC-Qz b) OPC-FA
Compressive strength [MPa]

60 7 °C 7 °C
20 °C 20 °C
50 40 °C 40 °C
50 °C 50 °C
40 80 °C 80 °C

30

20

10

0
1 10 100 1 10 100
Time [d] Time [d]

Fig. 1. Development of the compressive strength of (a) OPC-Qz and (b) OPC-FA at temperatures ranging from 7 to 80 °C.
172 F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181

a) 7 °C b) 23 °C
30 20 30 20
28 Bound H2O Portlandite 28 Bound H2O Portlandite
OPC-FA OPC-FA 18 OPC-FA OPC-FA 18
26 26
CH [g/100g dry binder]

CH [g/100g dry binder]


24 OPC-Qz OPC-Qz OPC-Qz

H* [g/100g dry binder]


24

H* [g/100g dry binder]


OPC-Qz 16 16
22 22
14 14
20 20
18 12 18 12
16 16
10 10
14 14
12 8 12 8
10 10
6 6
8 8
6 4 6 4
4 4
2 2
2 2
0 0 0 0
1 10 100 1 10 100
Time [d] Time [d]

c) 40 °C d) 50 °C
30 20 30 20
28 Bound H2O Portlandite 28 Bound H2O Portlandite
OPC-FA 18 OPC-FA OPC-FA 18
26 OPC-FA 26

CH [g/100g dry binder]


CH [g/100g dry binder]

24 OPC-Qz 16 24 OPC-Qz OPC-Qz 16

H* [g/100g dry binder]

H* [g/100g dry binder]


OPC-Qz
22 22
14 14
20 20
18 12 18 12
16 16
10 10
14 14
12 8 12 8
10 10
6 6
8 8
6 4 6 4
4 4
2 2
2 2
0 0 0 0
1 10 100 1 10 100
Time [d] Time [d]

e) 80 °C
30 20
28 Bound H2O Portlandite
OPC-FA 18
26 OPC-FA
CH [g/100g dry binder]

H* [g/100g dry binder]

24 OPC-Qz OPC-Qz 16
22
14
20
18 12
16
10
14
12 8
10
6
8
6 4
4
2
2
0 0
1 10 100
Time [d]

Fig. 2. Comparison of the content of portlandite (dashed lines) and bound water (solid lines) in OPC-FA and OPC-Qz at (a) 7 °C, (b) 23 °C, (c) 40 °C, (d) 50 °C and (e) 80 °C.

clinker phases and fly ash minerals, the diffractograms of OPC-FA show minor amounts of AFm phases. Due to the relatively high content of sul-
the formation of portlandite, ettringite, monosulphate and an AFm phate and magnesium in the Portland cement, the majority of Al is
phase, which is most likely a ternary solid solution of AFm phases con- bound within ettringite and possibly hydrotalcite in OPC-Qz. In OPC-
taining sulphate, carbonate and hydroxide [32,33]. FA, the fly ash dissolution leads to the release of Al and the increase of
After 1 day of hydration (Fig. 3) the diffractograms of OPC-Qz and the Al-to-sulphate ratio and thereby promotes the formation of AFm
OPC-FA are similar. At 7 and 23 °C the main hydrate phases are port- phases. At 80 °C, OPC-FA and OPC-Qz do not show any more reflections
landite and ettringite. The increase of the temperature to 40 or 50 °C related to the presence of ettringite and AFm phases.
leads to the formation of monosulphate in addition to ettringite and The presence of siliceous hydrogarnet is found after 180 days of hy-
portlandite. At 80 °C ettringite is not stable and an AFm phase is formed dration in OPC-FA and OPC-Qz. While the diffractograms at 50 °C and
instead. lower temperatures show only minor amounts of siliceous hydrogarnet,
After 180 days of hydration at 7 and 23 °C, portlandite, ettringite, at 80 °C siliceous hydrogarnet is one of the main hydrate phases. This is
monosulphate, a solid solution of AFm phases and monocarbonate are related to the kinetic hindrance of the formation of hydrogrossular
present in both samples (Fig. 4). At 40 and 50 °C, ettringite and below 50 °C.
portlandite persist in both samples. However, while OPC-FA shows the The change of the stable phase assemblage from ettringite and
presence of monosulphate and other AFm phases, OPC-Qz shows only monosulphate to siliceous hydrogarnet is related to the temperature
F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181 173

a) OPC-Qz b) OPC-FA
CH Qz Qz CH CH Qz CH
Ms Ms
AFmss AFmss Mu Mu
E F E E EF E F E E EF
80 °C

50 °C

40 °C

23 °C

7 °C
8 10 12 14 16 18 20 22 24 26 28 30 8 10 12 14 16 18 20 22 24 26 28 30
°2 Theta °2 Theta

Fig. 3. Comparison of XRD data of (a) OPC-Qz and (b) OPC-FA after 1 day of hydration at 7, 23, 40, 50 and 80 °C. E—ettringite, Ms—monosulphate, AFmss—solid solution of AFm phases, CH—
portlandite, F—ferrite, Mu—mullite, Qz—quartz.

dependent solubility of these phases and the availability of Al, Ca, sul- agreement with transmission electron microanalyses of C–S–H in fly
phate and hydroxide [4]. ash blended white Portland cement [37].
The Ca/Si, Al/Si and S/Si atomic ratios of OPC-FA and OPC-Qz after
180 days of hydration at temperatures ranging from 7 to 80 °C are
3.3.2. Analysis of the C–S–H composition by SEM-EDX given in Table 2. The C–S–H composition of OPC-Qz hydrated for
EDX point analyses were targeted at the outer product in the matrix 180 days at 80 °C shows a significant reduction of the Ca/Si ratio com-
of hydrates and the inner product of fly ash or clinker in order to char- pared to lower temperatures. This is on one hand related to the reduced
acterize the composition of the C–S–H phase. Due to the interaction vol- solubility of Ca(OH)2 at elevated temperatures. On the other hand the
ume of electrons with the specimen and the intergrowth of hydrates, quartz powder is being dissolved at high temperatures and the addi-
the detected X-rays of one spot often consist of the signal of a mix of tionally provided Si is probably incorporated into the C–S–H in a similar
two or more phases [34]. To derive the composition of a phase, element way as during the hydration of silica fume blended OPC [15,38,39].
ratio plots are used [35,36]. Compared to the reference sample, the C–S–H of the fly ash blended
The Al/Ca versus Si/Ca plot of OPC-FA hydrated for 180 days at 7, Portland cement is characterised by lower Ca/Si and higher Al/Si ratios.
50 and 80 °C is shown in Fig. 5. The scatter of data points towards This change of the C–S–H composition is related to the additional Si,
high Si/Ca and high Al/Ca ratios is due to the presence of fly ash particles provided by the fly ash dissolution, which is incorporated in the C–S–H
within the excited sample volume of some EDX point analyses. The Al/Si and results in longer silicate chain lengths [37,40,41]. This effect and
ratio of C–S–H is determined by the slope of a line, which is drawn the provision of Al from the dissolution of fly ash leads to an increased
through the points with the lowest Al/Ca ratio and represents mixed uptake of aluminium, which is substituting the Si atoms located in the
analyses of portlandite and C–S–H without AFm or ettringite. The range bridging tetrahedra of the silicate chains [42–44].
of Si/Ca ratios of the C–S–H is represented by the bulk of data points The comparison of the C–S–H compositions of OPC-FA hydrated at
along this line. OPC-FA at 7 °C shows a range of Si/Ca values between different temperatures shows that the Ca/Si ratio decreases with in-
0.5 and 0.7 or Ca/Si ratios between 1.4 and 2.0. This variance is in creasing temperature and the Al/Si ratio of the C–S–H increases from 7

a) OPC-Qz b) OPC-FA
CH Qz Qz CSH CH Qz CSH
Ms Ms
HG HG
AFmss HG HG AFmss HG Mu Mu
E F E E EF E F E E E F
80 °C

50 °C

40 °C

23 °C

7 °C
8 10 12 14 16 18 20 22 24 26 28 30 8 10 12 14 16 18 20 22 24 26 28 30
Mc °2 Theta CH Mc Mu °2 Theta CH

Fig. 4. Comparison of XRD data of (a) OPC-Qz and (b) OPC-FA after 180 days of hydration at 7, 23, 40, 50 and 80 °C. E—ettringite, Ms—monosulphate, AFmss—solid solution of
hemicarbonate and OH− substituted monosulphate, CH—portlandite, HG—hydrogarnet, F—ferrite, Mu—mullite, Qz—quartz.
174 F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181

Fig. 5. Plot of Al/Ca versus Si/Ca atomic ratios of the EDX measurements focussed on hydrate phases in OPC-FA hydrated for 180 days at 7, 50 and 80 °C. Ettr—ettringite, CH—portlandite,
C2ASH8—strätlingite, C3AH6—katoite, C3ASH4—hydrogrossular, Jen—jennite, Tob—tobermorite.

to 50 °C. This is probably related to the enhanced reactivity of fly ash at measured increased concentrations of dissolved Al and S in the pore
higher temperatures, the increased release of Si and Al ions from the fly solution.
ash dissolution and the increased degree of polymerisation of the C–S–H
at higher temperatures [13–15,45].
The uptake of sulphate in the C–S–H has been shown to be depen- 3.4. Analysis of the pore solutions
dent on the temperature and time of hydration [12,46]. The S/Si ratios
of the C–S–H in OPC-FA hydrated at 7 to 50 °C are about 0.04–0.05 The development of the elemental concentrations in the pore solu-
(Table 2). However, at a hydration temperature of 80 °C, the S/Si ratio tion of fly ash blended OPC has been described by various studies
of the C–S–H in OPC-FA is significantly increased to about 0.09. This is [19,20,47,48]. On one hand, the consumption of portlandite by the poz-
related to the increased solubility of ettringite at 80 °C and the thereby zolanic reaction causes a decrease of the Ca concentrations in the pore
increased availability of sulphate in the pore solution [17]. This effect solutions. On the other hand, the release of Al and Si from the dis-
agrees with investigations on the sulphate concentrations of C–S–H in solving fly ash leads to the increase of the Al and Si concentrations.
plain OPC at elevated temperatures [12,46]. In OPC-Qz the S/Si ratio in- Additionally, the hydroxide concentrations tend to decrease due to
creases slightly, in agreement with the less distinct increase of dissolved the enhanced binding of alkalis in the C–S–H formed by the pozzola-
sulphate in the pore solution. A similar observation has been made in an nic reaction [45,49–51].
unpublished study including microanalyses of the C–S–H of Portland The progress of the Ca, Al, Si, alkali and sulphate concentrations in
cement pastes blended with 50 wt.% silica fume hydrated at 80 °C OPC-Qz and OPC-FA at temperatures ranging from 7 to 80 °C is shown
(personal communication by Dr. B. Lothenbach). in Fig. 7. All elemental concentrations of the investigated pore solutions
To show the relation between the amount of Al and S incorporated in are also given in the appendix in Tables A.1 and A.2.
the C–S–H and the concentration of these elements in the pore solution, Fig. 7a shows that the Ca concentrations in the reference sample
the measured Al/Si or S/Si ratios of the C–S–H are plotted against the OPC-Qz are slightly decreased at elevated temperature, which agrees
measured concentrations of the pore solution (Fig. 6). The presented with the decreased solubility of portlandite. In OPC-FA, a significant
data is expanded by published data from other studies dedicated to drop of the Ca concentrations related to the pozzolanic reaction is ob-
the characterisation of synthesized C–S–H [44,46]. It shows that the in- served. The drop of the Ca concentrations is shifting to earlier hydration
creased amount of Al and S incorporated in the C–S–H agrees with the times by increasing the temperature from 7 to 50 °C due to the faster
onset of the pozzolanic reaction and the faster consumption of Ca(OH)2.
At 80 °C, the effect of the faster onset of the pozzolanic reaction is ob-
Table 2 served in the low measured Ca concentration at 1 day of hydration. Af-
Ca/Si, Al/Si and S/Si ratios of the EDX analyses of the C–S–H in OPC-FA and OPC-Qz after terwards the Ca concentration at 80 °C is higher than at 40 and 50 °C
180 days of hydration at 7, 23, 40, 50 and 80 °C. and the Al and Si concentrations are equal to those of OPC-FA at 40 and
Temperature 7 °C 23 °Ca 40 °C 50 °C 80 °C 50 °C. This is related to the fast reaction of fly ash at 80 °C up to 1 day
and a relatively low rate of reaction at later hydration times, as observed
OPC-FA
Ca/Si 1.7 ± 0.3 1.3 ± 0.2 1.4 ± 0.3 1.4 ± 0.3 1.1 ± 0.2 by the rate of portlandite consumption measured by TGA (Section 3.2)
Al/Si 0.14 ± 0.03 0.17 ± 0.03 0.19 ± 0.02 0.19 ± 0.02 0.19 ± 0.02 and the development of compressive strength (Section 3.1).
S/Si 0.05 ± 0.02 0.04 ± 0.02 0.04 ± 0.02 0.04 ± 0.02 0.09 ± 0.02 Fig. 7b shows the Al concentrations of the pore solution in OPC-Qz
OPC-Qz and OPC-FA. Up to 2 days of hydration the Al concentrations in OPC-
Ca/Si 1.8 ± 0.2 1.6 ± 0.3 1.7 ± 0.3 1.7 ± 0.3 1.2 ± 0.2 Qz are higher at 40–80 °C compared to lower temperatures. This is
Al/Si 0.07 ± 0.02 0.08 ± 0.02 0.08 ± 0.02 0.08 ± 0.02 0.05 ± 0.02 due to the higher ettringite solubility at higher temperature [17]. The
S/Si 0.05 ± 0.02 0.05 ± 0.02 0.05 ± 0.02 0.06 ± 0.02 0.06 ± 0.02 decrease at later ages could be related to the formation of siliceous
a
Data at 23 °C acquired from 250 day old samples (reproduced from [19]). hydrogarnet and Al uptake by C–S–H.
F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181 175

a) b)
0.12
0.20 OPC-Qz OPC-Qz

Atomic Al/Si ratio in the C-S-H

Atomic S/Si ratio in the C-S-H


OPC-FA OPC-FA
0.10
Pardal et al. Barbarulo et al.
0.15
0.08

0.06
0.10

0.04
0.05
0.02

0.00 0.00
0.01 0.1 1 10 0.1 1 10 100
Al [mM] S [mM]

Fig. 6. Dependency of the incorporation of Al and S in C–S–H on the pore solution chemistry. (a) Relation between the Al/Si atomic ratio of the C–S–H and the concentration of dissolved Al.
(b) Relation between the S/Si atomic ratio of the C–S–H and the concentration of dissolved S.

In OPC-FA, the increase of the Al concentrations related to the reac- The effective saturation indices of portlandite, ettringite, monosul-
tion of fly ash is shifting towards earlier times at higher temperatures phate and strätlingite, calculated from the elemental concentrations in
due to the faster onset of the pozzolanic reaction, which is expected the pore solutions, are shown and discussed in Appendix B.
to lead to an increased Al uptake in C–S–H. The acceleration of the poz-
zolanic reaction by higher temperature is also observed for the Si con-
centrations (Fig. 7c). 3.5. Microstructure
The progress of the alkali concentrations in OPC-FA is shown in
Fig. 7d. The effect of temperature on the alkali concentrations is not The effect of temperature on the microstructure of the investigated
straight forward compared to the other elements. There are several fac- samples is studied by back-scattered electron (BSE) images of OPC-FA
tors which influence the concentration of alkalis in the pore solution: and OPC-Qz after 180 days of hydration (Fig. 8). The main difference be-
tween the samples hydrated at different temperatures is found regard-
• The rate of the clinker reaction and the related release of alkalis into ing the porosity of the matrix and the grey level of the C–S–H.
the pore solution Fig. 8a and b shows a very low matrix porosity in the samples hy-
• The rate of the fly ash reaction and the related release of alkalis into drated at 7 °C. Most of the visible pores in the microstructure are hollow
the pore solution shell, so called “Hadley” grains [53] or internal fly ash pores. At higher
• The alkali binding capacity of the hydrate phases, which is related to temperatures, the microstructure is more heterogeneous and the poros-
the amount and composition of the C–S–H [50] ity is higher compared to lower temperatures as shown by the example
• The decrease of the volume of pore solution. images of OPC-FA and OPC-Qz at 50 and 80 °C (Fig. 8c–f).
Another temperature dependent parameter of the microstructure
At early hydration times, the enhanced clinker dissolution leads to is the grey level of the C–S–H in BSE images. In all investigated samples
increased alkali concentrations at elevated temperatures. Therefore at 7 °C, lower grey levels of the C–S–H are observed compared to higher
the highest alkali concentrations are found at 80 °C and the lowest at temperatures. The intensity of the grey level in the BSE image of a
7 °C. As long as the fly ash or the quartz powder is not reacting, the alkali polished sample is dependent on the electron density of the analysed
concentrations in the solution are increasing due to the ongoing dissolu- sample volume, which is mainly related to the atomic number of the
tion of clinker minerals and the reduction of the pore solution volume. analysed elements and the density or microporosity, respectively.
This effect is clearly observed in OPC-FA at 7 °C where the reaction of Since the C–S–H at 7 °C has a higher Ca content compared to the C–S–
fly ash is just slowly starting between 90 and 180 days. At higher tem- H of the samples hydrated at higher temperatures (as shown in
peratures, a reduction of the alkali concentrations is observed due to Fig. 5), the decreased grey levels have to be related to a higher micropo-
the enhanced binding of alkalis on the increased amounts of C–S–H dur- rosity or lower density, respectively. Similar findings have been report-
ing the main reaction of the fly ash. Hence, the reduction of alkalis in the ed for the hydration of neat OPC mortars [2,8] and recently also for fly
pore solution is related to the temperature dependent reactivity of fly ash blended Portland cement [11].
ash. A similar effect is observed for OPC-Qz when the quartz powder The relation between the microstructure of the paste samples
starts to react at temperatures of 40 °C and higher. At 80 °C, the concen- and the compressive strength of the mortar samples is not yet fully
tration of alkalis in OPC-Qz after 7 and 28 days of hydration is signifi- understood. Fig. 9 shows the compressive strength of the mortar sam-
cantly lower than in OPC-FA. This may be related to a higher C–S–H ples plotted versus the coarse porosity (pores N 0.6 μm). Although the
content, and hence enhanced binding of alkalis in OPC-Qz compared graph shows a general trend of increased strength with decreased po-
to OPC-FA. rosity, the compressive strength of samples with low coarse porosity
Fig. 7e shows the sulphate concentrations in the pore solutions of varies significantly depending on the temperature and time of hydra-
OPC-Qz and OPC-FA. At elevated temperatures an increase of sulphate tion. This effect can also visually be seen from the comparison of the
is observed due to the increased solubility of ettringite at higher tem- BSE images shown in Fig. 8a and c. Although OPC-FA after 180 days of
perature [17]. A similar increase of sulphate concentrations with tem- hydration at 50 °C has the highest compressive strength, it shows a
perature was also observed in silica fume blended and neat Portland higher coarse porosity than the same sample hydrated at 7 °C. This ef-
cements [3,15,45,52]. The lower sulphate concentrations in OPC-Qz fect is pronounced in OPC-Qz: At 80 °C, the OPC-Qz mortar shows by
compared to OPC-FA agree with the sulphate concentrations found in far the highest compressive strength after 180 days, although the coarse
the C–S–H. porosity is clearly lower than at 7 °C. On one hand, this may indicate,
176 F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181

that in addition to the content of coarse pores other factors such as the simplified model with constant reaction degrees of C3S, C2S, C3A and
interfacial transition zone at the aggregates in mortars, the hydrate C4AF. The assumed reaction degrees after 180 days of hydration (100%
phase assemblage and the mechanic properties of the hydration prod- for C3S, 86% for C2S, 93% for C3A and 90% for C4AF) are based on an exper-
ucts play an important role. On the other hand, the variance of compres- imental study showing only slight variations of the OPC reaction degree
sive strength values at low porosity is similar to other studies and is after 180 days within a temperature range between 5 and 40 °C [11].
probably related to the uncertainty caused by the fact that fine capillary In order to calculate the phase assemblage the following aspects are
pores with a diameter of less than 0.6 μm are not identified by this included in the model:
image analysis procedure [8].
• The variance of the Ca/Si ratio of the C–S–H is accounted for by using a
4. Modelling of the hydrate phase assemblage at solid-solution model with jennite (C1.67–S–H2.1) and tobermorite
different temperatures (C0.83–S–H1.33) as end-members. The Al and sulphate uptake in the
C–S–H is adapted to the measured Al/Si and S/Si ratios (Section 3.3.2).
The effect of temperature on the hydrate phase assemblage of OPC- • The formation of siliceous hydrogarnet is restricted. Significant quan-
Qz and OPC-FA after 180 days of hydration is simulated by using a tities of this phase have been reported only for cements hydrated at

a) Calcium
100
7 °C 7 °C
OPC-Qz 23 °C OPC-FA 23 °C
40 °C 40 °C
50 °C 50 °C
10 80 °C 80 °C
Ca [mM]

0.1
1 10 100 1 10 100
Time [d] Time [d]

b) Aluminium
10
7 °C
23 °C
40 °C
50 °C
1 80 °C
Al [mM]

0.1 7 °C
23 °C
40 °C
OPC-Qz OPC-FA 50 °C
80 °C
0.01
1 10 100 1 10 100
Time [d] Time [d]

c) Silicium
10
7 °C
OPC-Qz 23 °C OPC-FA
40 °C
50 °C
1 80 °C
Si [mM]

0.1 7 °C
23 °C
40 °C
50 °C
80 °C
0.01
1 10 100 1 10 100
Time [d] Time [d]

Fig. 7. Development of the (a) Ca, (b) Al, (c) Si, (d) alkali and (e) sulphate concentrations in OPC-FA and OPC-Qz hydrated at 7, 23, 40, 50 and 80 °C.
F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181 177

d) Sodium and potassium


600
500

400

Na+K [mM] 300

200 7 °C 7 °C
23 °C 23 °C
40 °C 40 °C
OPC-Qz 50 °C OPC-FA 50 °C
80 °C 80 °C
100
1 10 100 1 10 100
Time [d] Time [d]

e) Sulphate
100
SO4 [mM]

10

7 °C 7 °C
1
23 °C 23 °C
40 °C 40 °C
OPC-Qz 50 °C OPC-FA 50 °C
80 °C 80 °C
0.1
1 10 100 1 10 100
Time [d] Time [d]

Fig. 7 (continued).

higher temperatures [4,29,54–56] or in the presence of Fe(OH)3 given temperature is adapted to fit the measured portlandite content
[57,58]. Since the formation of Al siliceous hydrogarnet (hydro- (Section 3.2).
grossular) is kinetically hindered at room temperature, it is exclud- Fig. 11a shows the calculated change of the phase assemblage in
ed from the calculations at temperatures up to 50 °C. However, Fe OPC-FA up to 50 °C. Due to the increased amount of reacted fly ash
siliceous hydrogarnet (hydroandradite) may form also at lower glass, the elevation of temperature leads to the consumption of port-
temperatures. landite and the formation of additional C–S–H. Additionally, aluminate
is provided by the fly ash dissolution and contributes to the forma-
With these implementations, the development of the phase assem- tion of monosulphate starting from 30 °C. This process leads to the
blage after approximately 180 days of hydration is calculated as a func- destabilisation of ettringite at 50 °C.
tion of temperature and expressed as percentage of the original volume For the modelling of the temperature interval between 50 and 90 °C,
of the raw materials and water (Figs. 10 and 11). the formation of hydrogrossular is allowed in addition to hydro-
In the reference system OPC-Qz, the TGA experiments (Section 3.2) andradite (Fig. 11b). The presence of hydrogrossular affects the amount
show the consumption of portlandite by the pozzolanic reaction of the of ettringite and AFm. Due to the high Al content of the siliceous
quartz powder at temperatures between 40 and 80 °C. This is taken hydrogarnet, the ratio of available sulphate to aluminate is increased,
into account in the model by the adaption of the reaction degree of and hence ettringite persists up to 74 °C while no more AFm phases
the quartz powder at a given temperature to fit the experimentally are predicted to form. Hydrotalcite persists throughout the modelled
measured portlandite content. temperature range.
Fig. 10 shows the calculated change of the phase assemblage, exclud- Comparing the modelled development of hydrate phases in the two
ing the formation of hydrogrossular up to 50 °C. At low temperatures, systems shows lower amounts of C–S–H in OPC-FA compared to OPC-Qz
C–S–H, hydrotalcite, ettringite, monocarbonate, hydroandradite and at temperatures above 50 °C. This may be responsible for the observed
portlandite are predicted to form. At higher temperatures the pozzolanic lower compressive strength of OPC-FA at 80 °C compared to OPC-Qz
reaction of the quartz powder leads to the consumption of portlandite (Fig. 1). As a consequence of the lower C-S-H content in OPC-FA, less al-
and the formation of C–S–H. Due to the binding of Al in the C–S–H, the kali are bound within the C-S-H, which agrees well with the observed
content of monocarbonate is reduced with increasing temperature until higher alkali concentrations in the pore solution (Section 3.4). As at
this phase disappears at 50 °C, in agreement with the XRD results 80 °C, in the OPC-FA and the OPC-Qz where neither ettringite nor
shown in Fig. 4. Ettringite is predicted to be stable up to 72 °C. monosulphate is present, the sulphate is distributed between the pore
Hydrotalcite persists throughout the modelled temperature range. solution and the C–S–H. Hence, the lower content of C–S–H in OPC-FA
In OPC-FA, the calculated average composition of the glass phase is responsible for the observed increased S/Si ratio in the C–S–H
is used as input for the model, since the crystalline fraction of the fly (Section 3.3.2) and the higher sulphate concentrations in the pore solu-
ash shows no reaction [19,59,60]. The reaction degree of fly ash at a tion (Section 3.4).
178 F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181

a) OPC-FA 7 °C b) OPC-Qz 7 °C

20 µm 20 µm

c) OPC-FA 50 °C d) OPC-Qz 50 °C

20 µm 20 µm

e) OPC-FA 80 °C f) OPC-Qz 80 °C

20 µm 20 µm

Fig. 8. BSE images of the microstructure of OPC-FA and OPC-Qz after 180 days of hydration at 7, 50 and 80 °C.

60
7 °C
5. Conclusions
Compressive strength [MPa]

20 °C
50
40 °C
50 °C This study shows the effect of temperature on the hydration kinetics,
40 80 °C the change of the phase assemblage and the microstructure in Portland
cement blended with 50 wt.% of siliceous fly ash in comparison to a ref-
30 erence sample with 50 wt.% of quartz powder.
First of all, the temperature is affecting the OPC hydration, which
20 leads to enhanced early compressive strength at elevated temperatures.
Besides the well-known effects of temperature on the OPC hydration, a
10 strong influence of temperature on the reactivity of fly ash is observed.
The beginning pozzolanic reaction of the fly ash, as indicated by the con-
0 sumption of portlandite, and the change of the pore solution chemistry
0 10 20 30 40
is shifting towards earlier hydration times at elevated temperatures.
Coarse porosity [%]
The increase of the temperature from 7 to 23 °C shifts the start of the
Fig. 9. Compressive strength of OPC-FA plotted versus the coarse porosity. The error of the pozzolanic reaction from 90 days to 7 days. At 50 °C the fly ash reaction
coarse porosity is estimated to be ±3%. starts after 1 day of hydration.
F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181 179

a) b) the interfacial transition zone at the aggregates in mortars, the hy-


100 drate phase assemblage and the mechanic properties of the hydration
products, play an important role for the development of compressive
pore solution
80 strength.
% of original volume

siliceous hydrogarnet

60
Acknowledgements

portlandite
The authors wish to acknowledge the BASF Construction Chemicals
40 ettringite
hydrotalcite
GmbH and Schwenk Zement KG for the analytical and financial support.
C-S-H monocarbonate STEAG Power Minerals GmbH is acknowledged for the provision of the
fly ash. Thanks are extended to Julien Keraudy for the technical support
20
anhydrous Qz and Boris Ingold for the preparation and polishing of the samples for the
anhydrous OPC
SEM investigations.
0
10 20 30 40 50 50 60 70 80 90
Temperature [°C] Appendix A. Elemental concentrations of the pore solutions

Fig. 10. Modelled change of the phase assemblage in OPC-Qz after approximately 180 days The elemental concentrations of the pore solution in OPC-Qz and
as a function of temperature. (a) Calculation excluding the formation of hydrogrossular OPC-FA are shown in Tables A.1 and A.2.
below 50 °C and (b) including hydrogrossular above 50 °C.

Table A.1
At advanced hydration times the effect of temperature on the hydra- Elemental concentrations in the pore solutions of OPC-Qz at 7, 23, 40, 50 and 80 °C.

tion is governed by the enhanced pozzolanic reaction of the fly ash. The Temperature Time [d] K Na Ca Si S Al Fe Cl OH−
typical effects of the pozzolanic reaction on the hydration, like the mM
consumption of portlandite, the release of Al and Si into the pore solu-
7 °C 1 192 26 8.5 0.1 30.4 0.02 0.003 3.66 166
tion and the formation of additional C–S–H are accelerated at elevated
2 184 42 4.4 0.2 3.0 0.14 0.004 1.67 245
temperatures. Al and alkalis are partially incorporated into the C–S–H 7 228 47 3.9 0.2 5.3 0.17 0.002 1.27 303
and the remaining Al contributes to the formation of AFm phases and 28 251 64 3.4 0.2 5.9 0.15 0.005 0.51 296
to the formation of Al containing siliceous hydrogarnet at higher 90 276 93 2.8 0.3 2.0 0.24 0.004 0.33 352
temperatures. 180 295 92 1.5 0.1 7.7 0.15 0.003 0.15 381
23 °C 1 228 45 3.7 0.1 4.8 0.15 0.002 n.a. 283
At 80 °C, Al containing siliceous hydrogarnet forms while ettringite 2 289 74 3.6 0.1 0.9 0.12 0.002 n.a. 311
and AFm phases are destabilised. The released sulphate is partially 7 263 70 2.9 0.1 4.4 0.12 0.002 n.a. 332
enriched in the pore solution and partially incorporated into C–S–H 28 322 97 2.5 0.1 1.8 0.10 0.003 n.a. 323
and siliceous hydrogarnet. 90 303 95 2.2 0.1 3.1 0.11 0.003 n.a. 317
250 205 56 2.1 0.1 7.5 0.13 0.005 n.a. 277
The microstructure of the fly ash blended Portland cement is affect-
40 °C 1 216 53 2.3 0.2 8.9 0.59 0.003 1.33 308
ed in a similar way as OPC by the change of temperature. At 7 °C homo- 2 222 73 2.1 0.2 5.6 0.53 0.007 1.33 296
geneously distributed hydration products and a low content of coarse 7 249 66 2.3 0.1 11.7 0.14 0.002 0.69 327
pores (pores N 0.6 μm) are found. The C–S–H phases show relatively 28 207 59 2.1 0.1 17.5 0.10 0.003 0.42 244
low grey values due to their high microporosity. The increase of temper- 90 186 68 1.9 0.2 20.6 0.11 0.004 0.33 231
180 178 58 1.0 0.2 25.9 0.11 0.002 0.34 214
ature leads to a more heterogeneous distribution of the hydration 50 °C 1 227 58 2.6 0.2 17.4 0.50 0.003 1.23 285
products and the formation of a C–S–H with a lower microporosity, 2 226 76 2.3 0.2 14.8 0.36 0.007 0.95 274
and hence higher grey levels. The amount of coarse porosity increases 7 246 67 2.2 0.2 36.0 0.10 0.002 n.a. 251
with temperature. 28 213 59 2.2 0.1 43.2 0.13 0.004 0.78 216
90 236 86 3.2 0.3 50.8 0.11 0.005 0.38 190
The correlation between the coarse porosity and the compressive
180 172 57 1.4 0.2 42.3 0.09 0.002 0.23 169
strength is not perfectly clear and indicates that other factors, such as 80 °C 1 248 67 2.7 0.2 71.7 0.65 0.006 0.70 88
2 185 65 2.6 0.2 62.1 0.21 0.003 0.40 173
7 209 57 4.0 0.1 83.8 0.06 0.003 0.40 151
a) b)
100 n.a. = not available.

pore solution
80 Table A.2
% of original volume

Elemental concentrations in the pore solutions of OPC-FA at 7, 23, 40, 50 and 80 °C.
monosulphate siliceous hydrogarnet
60 Temperature Time [d] K Na Ca Si S Al Fe Cl OH−
portlandite mM
ettringite 7 °C 1 177 31 10.3 0.1 34.2 0.01 0.004 3.84 194
40
monocarbonate 2 171 47 4.0 0.2 5.2 0.43 0.005 2.30 254
C-S-H
hydrotalcite 7 211 53 3.8 0.2 5.4 0.16 0.003 1.62 303
20 28 221 60 3.6 0.1 6.0 0.12 0.005 0.70 296
glass fraction of FA
anhydrous OPC
90 275 95 1.9 0.2 1.9 0.28 0.004 0.70 352
180 297 98 0.5 0.5 7.8 1.75 0.002 0.30 381
crystalline fraction of FA 23 °C 1 262 68 3.7 0.1 5.8 0.02 0.002 1.24 276
0
10 20 30 40 50 50 60 70 80 90 2 294 86 3.4 0.1 0.9 0.13 0.002 0.75 311
7 322 104 3.0 0.1 1.1 0.13 0.002 0.97 270
Temperature [°C]
28 231 72 1.7 0.3 5.5 1.36 0.004 0.47 285
90 202 84 1.5 0.7 8.8 3.75 0.005 0.74 234
Fig. 11. Modelled change of the phase assemblage in OPC-FA after approximately 180 days 250 254 97 0.5 1.2 13.9 3.57 0.003 0.10 236
as a function of temperature. (a) Calculation excluding the formation of hydrogrossular
below 50 °C and (b) including hydrogrossular above 50 °C. (continued on next page)
180 F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181

Table A.2 (continued) these phases from the XRD patterns (Section 3.3.1). The effective SI of
Temperature Time [d] K Na Ca Si S Al Fe Cl OH− strätlingite indicate that it is not expected to precipitate.
mM In OPC-FA, the effective SI of portlandite (Table B.1) changes from
oversaturation to undersaturation with ongoing hydration at 7–50 °C.
40 °C 1 235 64 2.4 0.2 8.4 0.49 0.005 1.82 308
2 180 73 1.1 0.3 5.7 0.54 0.004 1.77 308
This effect is due to the consumption of portlandite by the pozzolanic re-
7 199 62 0.6 1.2 20.8 5.09 0.002 2.81 251 action of fly ash. When the rate of portlandite consumption by the poz-
28 179 59 0.7 1.1 25.5 4.01 0.003 2.67 234 zolanic reaction overcomes the formation rate of portlandite by the
90 200 84 0.4 1.3 23.3 2.30 0.004 2.65 250 reaction of the OPC, the effective SI yields negative values. The time of
180 222 81 0.2 1.5 38.9 3.49 0.004 3.18 241
the decrease of the effective SI of portlandite therefore depends on the
50 °C 1 207 57 1.7 0.3 13.5 0.66 0.003 1.79 275
2 221 84 0.7 0.8 30.0 2.73 0.013 2.71 255 reaction rate of OPC and fly ash, and hence on the temperature during
7 194 61 0.6 1.3 46.6 3.38 0.003 3.38 215 hydration. The data agrees with the portlandite consumption measured
28 212 68 0.7 0.8 50.9 3.37 0.007 5.15 200 by means of TGA (Fig. 2).
90 217 89 0.5 1.4 49.7 2.40 0.004 3.08 213 The effective SI of ettringite and monosulphate in OPC-FA are affect-
180 239 86 0.3 1.5 63.0 3.20 0.002 2.96 214
80 °C 1 268 86 1.4 0.7 154.6 3.89 0.006 3.06 227
ed in a similar way as in OPC-Qz. The reaction of fly ash shows no signif-
2 271 109 1.3 1.0 149.0 2.60 0.006 3.96 104 icant effect on the effective SI of these phases.
7 303 98 1.3 1.1 166.9 3.12 0.003 5.89 124 The effective SI of strätlingite turns positive with ongoing hydration.
28 305 100 1.6 0.6 174.9 2.96 0.005 5.08 98 This effect is typically observed in fly ash blended samples due to the
consumption of Ca(OH)2 and the provision of Al and Si by the reaction
of the fly ash. Due to the enhanced reactivity of fly ash at higher temper-
Appendix B. Effective saturation indices atures, the increase of the effective SI of strätlingite is shifted towards
earlier hydration times. In fact, OPC-FA does not show the formation
The effective saturation indices (SI) of portlandite, ettringite, mono- of strätlingite due to the presence of siliceous hydrogarnet, as shown
sulphate and strätlingite in OPC-Qz and OPC-FA at temperatures ranging by XRD (Section 3.3.1) and thermodynamic modelling (Section 4).
from 7 to 80 °C are shown in Table B.1. The effective SI of portlandite in
the reference sample OPC-Qz at different temperatures shows a minor References
effect of the temperature. At higher temperatures after 180 days of hy-
[1] J.I. Escalante-Garcia, J.H. Sharp, Effect of temperature on the hydration of the main
dration, the effective SI of portlandite is slightly reduced compared to clinker phases in Portland cements: part I, neat cements, Cem. Concr. Res. 28
lower temperatures. Also the SI of ettringite and monosulphate show (1998) 1245–1257.
[2] K.O. Kjellsen, R.J. Detwiler, O.E. Gjørv, Development of microstructures in plain
lower values at elevated temperatures. This is related to a faster hydra-
cement pastes hydrated at different temperatures, Cem. Concr. Res. 21 (1991)
tion kinetic at higher temperatures. The pore solutions are thus nearer to 179–189.
the thermodynamic equilibrium. [3] B. Lothenbach, F. Winnefeld, C. Alder, E. Wieland, P. Lunk, Effect of temperature on
At 80 °C, the pore solution is undersaturated with respect to et- the pore solution, microstructure and hydration products of Portland cement pastes,
Cem. Concr. Res. 37 (2007) 483–491.
tringite and monosulphate, which agrees with the disappearance of [4] B. Lothenbach, T. Matschei, G. Möschner, F.P. Glasser, Thermodynamic modelling of
the effect of temperature on the hydration and porosity of Portland cement, Cem.
Concr. Res. 38 (2008) 1–18.
[5] K.O. Kjellsen, R.J. Detwiler, O.E. Gjørv, Backscattered electron imaging of cement
pastes hydrated at different temperatures, Cem. Concr. Res. 20 (1990) 308–311.
Table B.1
[6] K.O. Kjellsen, R.J. Detwiler, O.E. Gjørv, Pore structure of plain cement pastes hydrated
Effective saturation indices of selected hydrate phases in OPC-Qz and OPC-FA at 7, 23, 40,
at different temperatures, Cem. Concr. Res. 20 (1990) 927–933.
50 and 80 °C calculated from the respective elemental concentrations in the pore solution. [7] K.O. Kjellsen, R.J. Detwiler, Reaction kinetics of Portland cement mortars hydrated at
CH—portlandite, Ettr—ettringite, Ms—monosulphate, Strätl—strätlingite. different temperatures, Cem. Concr. Res. 22 (1992) 112–120.
[8] X. Zhang, Quantitative microstructural characterisation of concrete cured under re-
Time Effective saturation indices
alistic temperature conditions. (PhD thesis) Laboratory of Construction Materials,
[d] École polytechnique fédérale de Lausanne, Lausanne, 2007.
OPC-Qz OPC-FA
[9] J.L. Poole, K.A. Riding, M.C.G. Juenger, K.J. Folliard, A.K. Schindler, Effects of supple-
CH Ettr Ms Strätl CH Ettr Ms Strätl mentary cementing materials on apparent activation energy, J. ASTM Int. 7 (2010).
[10] W. Ma, D. Sample, R. Martin, P.W. Brown, Calorimetric study of cement blends con-
7 °C 1 0.11 0.65 0.23 −0.38 0.12 0.65 0.21 −0.46
taining fly ash, silica fume, and slag at elevated temperatures, Cem. Concr. Aggr. 16
2 0.10 0.50 0.25 −0.04 0.07 0.59 0.34 0.12
(1994) 93–99.
7 0.12 0.52 0.28 −0.05 0.10 0.52 0.27 −0.05 [11] K. De Weerdt, M. Ben Haha, G. Le Saout, K.O. Kjellsen, H. Justnes, B. Lothenbach, The
28 0.12 0.50 0.25 −0.10 0.11 0.50 0.25 −0.13 effect of temperature on the hydration of composite cements containing limestone
90 0.12 0.39 0.22 −0.01 0.07 0.33 0.17 −0.03 powder and fly ash, Mater. Struct. 45 (2012) 1101–1114.
180 0.03 0.36 0.13 −0.19 −0.19 0.24 0.09 0.09 [12] C. Famy, K.L. Scrivener, A. Atkinson, A.R. Brough, Effects of an early or a late heat
23 °C 1 0.12 0.32 0.17 −0.13 0.16 0.22 0.05 −0.42 treatment on the microstructure and composition of inner C–S–H products of
2 0.18 0.14 0.09 −0.21 0.18 0.13 0.10 −0.20 Portland cement mortars, Cem. Concr. Res. 32 (2002) 269–278.
7 0.12 0.25 0.12 −0.20 0.18 0.12 0.08 −0.21 [13] A. Bentur, R.L. Berger, J.H. Kung, N.B. Milestone, J.F. Young, Structural properties
28 0.15 0.12 0.06 −0.29 0.02 0.31 0.22 0.14 of calcium silicate pastes: II, effect of curing temperature, J. Am. Ceram. Soc. 62
90 0.12 0.15 0.06 −0.28 −0.05 0.35 0.26 0.33 (1979) 362–366.
250 0.02 0.25 0.09 −0.22 −0.22 0.14 0.07 0.18 [14] J. Hirljac, Z.Q. Wu, J.F. Young, Silicate polymerization during the hydration of alite,
40 °C 1 0.06 0.19 0.15 −0.03 0.09 0.18 0.14 −0.05 Cem. Concr. Res. 13 (1983) 877–886.
[15] T.T.H. Bach, C.C.D. Coumes, I. Pochard, C. Mercier, B. Revel, A. Nonat, Influence of
2 0.07 0.12 0.11 −0.04 −0.06 0.01 0.00 −0.11
temperature on the hydration products of low pH cements, Cem. Concr. Res. 42
7 0.09 0.12 0.05 −0.27 −0.22 0.09 0.07 0.21
(2012) 805–817.
28 0.03 0.12 0.01 −0.31 −0.23 0.12 0.08 0.21 [16] D. Damidot, F.P. Glasser, Thermodynamic investigation of the CaO–Al2O3–CaSO4–
90 −0.01 0.12 0.00 −0.28 −0.25 0.00 −0.03 0.06 H2O system at 50 °C and 85 °C, Cem. Concr. Res. 22 (1992) 1179–1191.
180 −0.13 0.03 −0.10 −0.36 −0.36 −0.05 −0.09 0.03 [17] R.B. Perkins, C.D. Palmer, Solubility of ettringite Ca6[Al(OH)6]2(SO4)3 ∗ 26H2O at 5–
50 °C 1 0.09 0.15 0.14 −0.07 0.02 0.08 0.08 −0.07 75 degrees C, Geochim. Cosmochim. Acta 63 (1999) 1969–1980.
2 0.09 0.10 0.09 −0.16 −0.13 0.05 0.06 0.08 [18] D. Jansen, C. Stabler, F. Goetz-Neunhoeffer, S. Dittrich, J. Neubauer, Does ordinary
7 0.05 0.08 0.00 −0.35 −0.26 0.05 0.03 0.13 Portland cement contain amorphous phase? A quantitative study using an external
28 0.00 0.12 0.02 −0.31 −0.20 0.10 0.08 0.13 standard method, Powder Diffract. 27 (2011) 31–38.
90 0.08 0.17 0.07 −0.26 −0.25 −0.02 −0.03 0.03 [19] F. Deschner, F. Winnefeld, B. Lothenbach, S. Seufert, P. Schwesig, S. Dittrich, F.
180 −0.12 0.01 −0.11 −0.38 −0.34 −0.08 −0.09 0.00 Goetz-Neunhoeffer, J. Neubauer, Hydration of Portland cement with high replace-
80 °C 1 0.05 0.01 0.08 −0.18 −0.40 −0.03 −0.02 0.07 ment by siliceous fly ash, Cem. Concr. Res. 42 (2012) 1389–1400.
2 −0.01 −0.06 −0.03 −0.32 −0.34 −0.17 −0.16 −0.21 [20] K. De Weerdt, M. Ben Haha, G. Le Saout, K.O. Kjellsen, H. Justnes, B. Lothenbach, Hy-
dration mechanisms of ternary Portland cements containing limestone powder and
7 0.00 −0.04 −0.07 −0.47 −0.31 −0.04 −0.01 0.06
fly ash, Cem. Concr. Res. 41 (2011) 279–291.
F. Deschner et al. / Cement and Concrete Research 52 (2013) 169–181 181

[21] R.S. Barneyback, S. Diamond, Expression and analysis of pore fluids from hardened [39] A.R. Brough, C.M. Dobson, I.G. Richardson, G.W. Groves, A study of the pozzolanic re-
cement pastes and mortars, Cem. Concr. Res. 11 (1981) 279–285. action by solid-state 29Si nuclear magnetic resonance using selective isotopic en-
[22] J.C. Russ, The Image Processing Handbook, 4th edition CRC Press, London, 2002. richment, J. Mater. Sci. 30 (1995) 1671–1678.
[23] F. Deschner, B. Münch, F. Winnefeld, B. Lothenbach, Quantification of fly ash in hy- [40] J.J. Chen, J.J. Thomas, H.F.W. Taylor, H.M. Jennings, Solubility and structure of calcium
drated, blended Portland cement pastes by backscattered electron imaging, J. silicate hydrate, Cem. Concr. Res. 34 (2004) 1499–1519.
Microsc. 251 (2013) 188–204. [41] X.D. Cong, R.J. Kirkpatrick, Si-29 MAS NMR study of the structure of calcium silicate
[24] D. Kulik, T. Wagner, S.V. Dmytrieva, G. Kosakowski, F.F. Hingerl, K.V. Chudnenko, hydrate, Adv. Cem. Based Mater. 3 (1996) 144–156.
U.R. Berner, GEM-Selektor geochemical modeling package: revised algorithm [42] I.G. Richardson, G.W. Groves, The incorporation of minor and trace-elements into
and GEMS3K numerical kernel for coupled simulation codes, Comput. Geosci. calcium silicate hydrate (C–S–H) gel in hardened cement pastes, Cem. Concr. Res.
17 (2013) 1–24. 23 (1993) 131–138.
[25] T. Wagner, D.A. Kulik, F.F. Hingerl, S.V. Dmytrieva, GEM-Selektor geochemical [43] P. Yu, R.J. Kirkpatrick, B. Poe, P.F. McMillan, X.D. Cong, Structure of calcium silicate
modeling package: TSolMod library and data interface for multicomponent phase hydrate (C–S–H): near-, mid-, and far-infrared spectroscopy, J. Am. Ceram. Soc. 82
models, Can. Mineral. 50 (2012) 1173–1195. (1999) 742–748.
[26] W. Hummel, U. Berner, E. Curti, F.J. Pearson, T. Thoenen, Nagra/PSI Chemical Ther- [44] X. Pardal, F. Brunet, T. Charpentier, I. Pochard, A. Nonat, 27Al and 29Si solid-state
modynamic Data Base 01/01, Universal Publishers/u-PUBLISH.com, USA, also pub- NMR characterization of calcium–aluminosilicate–hydrate, Inorg. Chem. 51 (2012)
lished as Nagra Technical Report NTB 02-16, Wettingen, Switzerland, 2002. 1827–1836.
[27] B.Z. Dilnesa, B. Lothenbach, G. Le Saout, G. Renaudin, A. Mesbah, Y. Filinchuk, A. [45] T.T. Ha Bach, Evolution physico-chimique des liants bas pH hydratés: Influence de la
Wichser, E. Wieland, Iron in carbonate containing AFm phases, Cem. Concr. Res. température et mécanisme de rétention des alcalins. (PhD thesis) UFR Sciences et
41 (2011) 311–323. Techniques, Université de Bourgogne, Bourgogne, 2010, p. 238.
[28] B.Z. Dilnesa, B. Lothenbach, G. Renaudin, A. Wichser, E. Wieland, Stability of mono- [46] R. Barbarulo, H. Peycelon, S. Leclercq, Chemical equilibria between C–S–H and
sulfate in the presence of iron, J. Am. Ceram. Soc. 95 (2012) 3305–3316. ettringite, at 20 and 85 °C, Cem. Concr. Res. 37 (2007) 1176–1181.
[29] B.Z. Dilnesa, Fe-containing hydrates and their fate during cement hydration: ther- [47] K. Luke, E. Lachowski, Internal composition of 20-year-old fly ash and slag-blended
modynamic data and experimental study, Laboratory of Construction Materials, ordinary Portland cement pastes, J. Am. Ceram. Soc. 91 (2008) 4084–4092.
École polytechnique fédérale de Lausanne, Lausanne, 2012. [48] S. Diamond, Effects of two Danish fly ashes on alkali contents of pore solutions of
[30] H.C. Helgeson, D.H. Kirkham, G.C. Flowers, Theoretical prediction of the thermody- cement-fly ash pastes, Cem. Concr. Res. 11 (1981) 383–394.
namic behaviour of aqueous electrolytes at high pressures and temperatures: IV. [49] J. Duchesne, M.A. Bérubé, Effect of supplementary cementing materials on the com-
Calculation of activity coefficients, osmotic coefficients, and apparent molal and position of cement hydration products, Adv. Cem. Based Mater. 2 (1995) 43–52.
standard and relative partial molal properties to 600 °C and 5 Kb, Am. J. Sci. 281 [50] S.Y. Hong, F.P. Glasser, Alkali binding in cement pastes part I. The C–S–H phase, Cem.
(1981) 1249–1516. Concr. Res. 29 (1999) 1893–1903.
[31] F. Deschner, B. Lothenbach, F. Winnefeld, P. Schwesig, S. Seufert, S. Dittrich, J. [51] S.Y. Hong, F.P. Glasser, Alkali sorption by C–S–H and C–A–S–H gels—part II. Role of
Neubauer, Investigation of a model system to characterize the pozzolanic reactivity alumina, Cem. Concr. Res. 32 (2002) 1101–1111.
of two low Ca fly ashes and a quartz powder, GDCh Fachtagung Bauchemie, [52] J.J. Thomas, D. Rothstein, H.M. Jennings, B.J. Christensen, Effect of hydration temper-
Hamburg, 2011, pp. 127–132. ature on the solubility behavior of Ca-, S-, Al-, and Si-bearing solid phases in
[32] H. Pöllmann, Syntheses, properties and solid solution of ternary lamellar calcium Portland cement pastes, Cem. Concr. Res. 33 (2003) 2037–2047.
aluminate hydroxi salts (AFm-phases) containing SO2− 2−
4 , CO3 and OH−, N. Jahrb. [53] D.W. Hadley, The Nature of the Paste-aggregate Interface. (PhD thesis) Purdue Uni-
Mineral. Abh. 182 (2006) 173–181. versity, West Lafayette, 1972.
[33] T. Matschei, B. Lothenbach, F.P. Glasser, The AFm phase in Portland cement, Cem. [54] G. Le Saout, E. Lecolier, A. Rivereau, H. Zanni, Chemical structure of cement aged at
Concr. Res. 37 (2007) 118–130. normal and elevated temperatures and pressures Part I. Class G oilwell cement,
[34] K.L. Scrivener, Backscattered electron imaging of cementitious microstructures: un- Cem. Concr. Res. 36 (2006) 71–78.
derstanding and quantification, Cement Concr. Compos. 26 (2004) 935–945. [55] N. Neuville, E. Lecolier, G. Aouad, A. Rivereau, D. Darnidot, Effect of curing conditions
[35] S. Diamond, Identification of hydrated cement constituents using a scanning elec- on oilwell cement paste behaviour during leaching: experimental and modelling ap-
tron microscope—energy dispersive X-ray spectrometer combination, Cem. Concr. proaches, C. R. Chim. 12 (2009) 511–520.
Res. 2 (1972) 617–632. [56] M. Paul, F.P. Glasser, Impact of prolonged warm (85 °C) moist cure on Portland ce-
[36] A.M. Harrison, N.B. Winter, H.F.W. Taylor, Microstructural development during ment paste, Cem. Concr. Res. 30 (2000) 1869–1877.
the hydration of cement, Materials Research Society Symposia, Pittsburgh, 1987, [57] N.C. Collier, N.B. Milestone, J. Hill, I.H. Godfrey, The disposal of radioactive ferric floc,
pp. 213–222. Waste Manag. 26 (2006) 769–775.
[37] A.V. Girao, I.G. Richardson, R. Taylor, R.M.D. Brydson, Composition, morphology and [58] N.C. Collier, N.B. Milestone, J. Hill, I.H. Godfrey, Immobilisation of Fe floc: part 2, en-
nanostructure of C–S–H in 70% white Portland cement-30% fly ash blends hydrated capsulation of floc in composite cement, J. Nucl. Mater. 393 (2009) 92–101.
at 55 degrees C, Cem. Concr. Res. 40 (2010) 1350–1359. [59] A.L.A. Fraay, J.M. Bijen, Y.M. Dehaan, The reaction of fly ash in concrete—a critical ex-
[38] C.M. Dobson, D.G.C. Goberdhan, J.D.F. Ramsay, S.A. Rodger, 29Si MAS NMR study of amination, Cem. Concr. Res. 19 (1989) 235–246.
the hydration of tricalcium silicate in the presence of finely divided silica, J. Mater. [60] S.A. Rodger, G.W. Groves, The microstructure of tricalcium silicate/pulverized-fuel
Sci. 23 (1988) 4108–4114. ash blended cement pastes, Adv. Cem. Res. 1 (1988) 84–91.

Вам также может понравиться