Sei sulla pagina 1di 11

Cement and Concrete Research 86 (2016) 1–11

Contents lists available at ScienceDirect

Cement and Concrete Research

journal homepage: www.elsevier.com/locate/cemconres

A new view on the kinetics of tricalcium silicate hydration


L. Nicoleau a,⁎, A. Nonat b
a
BASF Research Construction Materials & Systems, BASF Construction Solutions GmbH, 83308 Trostberg, Germany
b
Laboratoire Interdisciplinaire Carnot de Bourgogne UMR6303 CNRS, Université de Bourgogne, 9 avenue Alain Savary, BP 47870, 21078 Dijon Cedex, France

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

Article history: C3S hydration is an interesting example of chemical coupling between C3S dissolution, C–S–H and portlandite
Received 18 June 2015 precipitation. It occurs because Ca2+, OH− and silicate ions are present in C3S, in both hydration products and
11 April 2016 in the surrounding solution. Various experimental data sets reveal that the undersaturation with respect to C3S
Accepted 27 April 2016
always increases when C3S hydration enters into the deceleratory phase, leading to the conclusion that C3S dis-
Available online 3 May 2016
solution is at the origin of this deceleration, not C–S–H growth. In addition, as soon as portlandite precipitates, the
Keywords:
dissolution limits the hydration already in the acceleratory hydration step. The evolution of the undersaturation
Hydration cannot account for the hydration peak. Rather, it results from an extension and subsequently a decrease of the C3S
Kinetics reactive surface area. The formation and coalescence of dissolution etch-pits can provide a reasonable explana-
Pore solution tion for the C3S hydration kinetics.
Ca3SiO5 © 2016 Elsevier Ltd. All rights reserved.
Cement

1. Introduction respect to portlandite to lead to the precipitation of portlandite. In


such conditions, the solution gets more and more concentrated in cal-
Tricalcium silicate (C3S1), the main component of ordinary Portland cium and hydroxide ions. If the maximal supersaturation with respect
cement, hydrates into calcium-silicate-hydrates (C–S–H) and then pos- to portlandite is exceeded, the latter precipitates leading to a new ki-
sibly into portlandite (Ca(OH)2), as follows: netic situation, hence resulting in the C3S dissolution and both C–S–H
C3S dissolution: and Ca(OH)2 precipitation. In all cases, the chemical equations impose
the following:
Ca3 SiO5 þ 5 H2 O→3Ca2þ þ 6OH− þ H4 SiO4 ð1Þ

xþyþz¼3
Cx-S-Hy precipitation:

x Ca2þ þ 2x OH− þ H4 SiO4 →ðCaOÞ x −ðSiO2 Þ−ðH2 OÞ b z being the number of moles of calcium in solution for 1 mol of C3S. The
þ ð2 þ x−bÞ H2 O ð2Þ authors chose H4SiO4 as silicate species for balancing the reactions but
are aware that the solution consists of a mix of H4SiO4, H3SiO− 4 and
Ca(OH)2 precipitation: H2SiO2−4 , for which the speciation depends on pH.
The hydration of cement today is firmly recognized to be a dissolu-
y Ca2þ þ 2y OH− →CaðOHÞ 2 ð3Þ tion–precipitation process, and reaction kinetics has been a prominent
research topic for a century. Le Chatelier, being a fervent defender of
with x + y = 3 and x the calcium to silicon molar ratio that is a variable
the “through-solution” theory, was the first to lay the foundations of hy-
parameter of the C–S–H composition that strongly depends on the con-
dration chemistry [3]. Since then, even if our knowledge considerably
centration of calcium hydroxide in solution. b, the amount of water mol-
increased, the overall picture on hydration is still not complete and
ecules in C–S–H is also a variable close to the value of 1.8 [1,2] for C–S–H
some questions are still a matter of debate (see reviews in [4–6]).
with high Ca/Si ratios. C3S hydration consists of at least two of these re-
More recently, our understanding on the dissolution reaction of
actions which are the dissolution of C3S and the precipitation of C–S–H.
tricalcium silicate has been increased [7–9]. Under the new light of
Using the example of dilute suspensions, it is possible to hydrate C3S
these results, some critical points regarding alite hydration kinetics
under conditions undersaturated or insufficiently supersaturated with
are revisited in this paper (the tricalcium silicate phase in cement is
commonly called alite).
⁎ Corresponding author.
E-mail address: luc.nicoleau@basf.com (L. Nicoleau).
The hydration of alite is routinely measured by the well-established
1
In cement notation, C, S and H stand for CaO, SiO2 and H2O respectively. C3S = Ca3SiO5, technique of isothermal conduction calorimetry. Q̇ , the heat flow re-
C–S–H = (CaO)x−(SiO2)−(H2O)y. leased during the hydration of alite is equal to the sum of heat flows

http://dx.doi.org/10.1016/j.cemconres.2016.04.009
0008-8846/© 2016 Elsevier Ltd. All rights reserved.
2 L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11

released or stored by each reaction: grain [18]. Note that in this latter, the consequence is not a decrease
but a constant hydration rate.
Q_ ¼ RC3 S diss: ΔHC3 S diss: þ RC–S–H prec: ΔHC–S–H prec: þ RCH prec: ΔHCH prec: ð4Þ It is worthwhile mentioning that, to the best of the authors' knowl-
edge, these C–S–H growth-based hypotheses have never been experi-
with ΔHi the reaction enthalpy of the reaction i. In paste, the 3 rates be- mentally verified, but rather have been supported by self-consistent
come rapidly equal and then Q̇ is proportional to the hydration rate. The fitting. Indeed, the typical hydration curve (as plotted in Fig. 1) has
characteristic heat evolution curve (Fig. 1) obtained during the hydra- been modeled or fitted using these theories. More generally, many dif-
tion of C3S in paste presents different phases: ferent models (based on C–S–H) can be used to reproduce the sigmoidal
I The dissolution peak: this non-steady state step is difficult to interpret hydration curve [17–21] showing in some cases, excellent agreement
because of the calorimeter inertia and often because of the frictional between experiments and fits. But even a perfect fit does not provide
heat due to the sample introduction. validation of a theory [22]. The C3S hydration curve is the result of dis-
II The so-called “dormant” period or sometimes called the induction pe- solution and precipitation reactions leading to possible dual interpreta-
riod. This period of low activity precedes the acceleration of hydra- tions if the theory used to produce a curve does not account for both
tion. It is generally very short in the case of pure C3S hydration. reactions. In addition to the possible space constraints on C–S–H
III The acceleratory phase: this is a relatively short period of time but im- growth, C3S is depleting over time while C–S–H grows. After a certain
portant for the early mechanical properties. During these first hours, time, and as long as water is still available, C3S will not be able to provide
the rate increases until reaching a maximum value. enough material for the C–S–H growth and the hydration rate will be
IV The deceleratory phase: with the acceleratory period it forms the limited. This has been described by Nonat et al. [23] and appears to be
main hydration peak. The deceleratory phase continues until the indisputable evidence. However, the corresponding hydration time
end of hydration. has never been experimentally identified, but has only been predicted
to occur.
In this article, a simple approach has been used that consists of the
Among the still-debated issues, two of them have received particular examination of ion concentrations during hydration under various ex-
attention during the last two or three decades. The first one is the occur- perimental conditions. The data are all extracted from the literature.
rence of the dormant period. To explain this, two theories are often re- As explained in the next section, the evolution of ion concentrations al-
ported in the literature. One is based on the precipitation of a transient lows for the determination of which reaction drives the kinetics, for ex-
metastable C–S–H phase [10,11] and another one is based on C–S–H nu- ample through the calculation of the undersaturation with respect to
cleation kinetics, which may be hampered by aluminum ions, [12] for C3S and the supersaturation with respect to hydration products. In
instance. An alternative to these theories, recently proposed, suggests spite of the simplicity of the approach to monitor ion concentrations,
that the dormant period results from underlying mechanisms the evolution of the characteristics of the pore solution and the conse-
governing the opening of dissolution etch-pits [9]. quences on hydration kinetics has never been thoroughly studied, ex-
The second most important issue is the reasons leading to the main cept during the very early period of hydration [24] and in a very
hydration peak. The acceleration is generally attributed to the increase recent theoretical study [25]. The relation between the C3S dissolution
of the surface area of C–S–H in a kind of autocatalytic process. The rate and its undersaturation has been recently described [7]. Unfortu-
entry into the deceleratory regime following the maximum rate value nately, the same kind of work on the C–S–H precipitation is still pend-
is more discussed. So far, most of the hypotheses advanced to rationalize ing. The role of the dissolution rate on hydration kinetics will be put
the deceleration also rely upon C–S–H growth. There are two assump- into perspective, and light will be shed on the phenomena taking
tions, both of which postulated that C–S–H growth is constrained in place during the main hydration peak shown by calorimetry. An alter-
space. It is assumed that some constraints limit the growth rate and con- native hydration scheme will be then proposed and meanwhile the rel-
sequently the hydration rate. They are derived from Avrami's model evance of the existing theories based on C–S–H growth will be
[13–15] and have been adapted to C3S (or cement) hydration [16]. discussed.
Some authors considered that the transformation takes place on the
surface of grains [17,18]. The theories arising in these articles are differ- 2. Notions of heterogeneous kinetics applied to C3S hydration
entiated by the origin of space constraints. The first proposes that C–S–H
growth is hindered by existing C–S–H that grows outward from the sur- C3S hydration is an example of transformation of solid A into solid B
rounding cement grains [17], and the second proposes that the hin- through the solution. As a matter of fact, during C3S hydration, C3S dis-
drance on growth is due to adjacent C–S–H growing on the same solves and releases ions into the solution at the flow Φdiss. At the same
time, hydrates, i.e. C–S–H and portlandite, precipitate and consume
ions from the solution at the flow Φprec. Because C3S, C–S–H and
portlandite are composed of the same species, dissolution and precipita-
tion rates, i.e., the appearance rate of C–S–H and portlandite and the dis-
appearance rate of C3S, are linked. Such a connection is the definition of
the so-called coupled reactions. In case of a steady state or a quasi-
steady state, both flows become equal:

φdiss: ðtÞ ¼ φprec: ðtÞ: ð5Þ

In the following, some known concepts about heterogeneous kinet-


ics are described to establish pertinent groundwork for C3S hydration
kinetics and to show that C3S dissolution and precipitation of C–S–H
and portlandite are coupled. Many interesting examples of coupling
can be found in geochemistry. Indeed coupling is typical in diagenesis,
metamorphism, metasomatism, chemical weathering, and other
Fig. 1. Very typical evolution of alite hydration at water-to-alite ratio of 0.5. The usual steps
water-rock reactions [26–28]. In these geochemical transformations,
are represented: I the pure dissolution peak, II the “dormant” period, III the acceleratory mineral A transforms into mineral B through the surrounding fluid.
period and IV the deceleratory period. The dissolution rate of A is equal to the precipitation rate of B and the
L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11 3

Fig. 2. (a) Evolution of the sodium concentration over the transformation of albite into sanidine in 0.1 M KHCO3 solution extracted from the Ref. [34]. According to the Eq. (8) the sodium
concentration is directly proportional to the degree of reaction. (b) Ratios of albite dissolution rates versus sanidine precipitation rates. (c) Evolution of undersaturation with respect to
albite and the supersaturation with respect to sanidine (SI = saturation index). The saturation index is defined as the logarithm of the supersaturation or undersaturation.
(Reproduced from Ref. [34] with permission from Elsevier, Copyright 2010). For (a), (b) and (c), the dots refer to experimental measurements and the lines are calculated from
numerical reaction path model simulation.

role of the fluid in the process is fundamental. Two ways to couple dis- where Rate is the overall rate of reaction, k the rate constant function of
solution–precipitation reactions have been identified. On the one hand, the temperature, Asurf is the surface extent where reaction occurs
coupling is induced by the stress coming from crystallization as in pseu- (called the reactive surface area), ΔG is the difference in Gibbs free en-
domorphic replacement [27,29]. In this transformation, the coupling of ergy between the solid and its molecules in solution and g(ai) a function
dissolution–precipitation reactions allows the conservation of the vol- highlighting all the possible catalyzing and inhibiting effects of the spe-
ume and texture characteristics of solid A during its replacement by cies i. Despite the multiplicity of the function f(ΔG) in nature, all rates
solid B. On the other hand, as during C3S hydration, if some chemical el- monotonically decrease with the deviation from equilibrium: in any
ements are common in A and B, coupling is ensured through the differ- case, the solid A will dissolve more and more slowly when the solution
ence of chemical potentials between the solution and both solids for is less and less undersaturated with respect to A and the solid B will pre-
elements common in A and B. Because hydration of C3S occurs by the cipitate more and more slowly when the solution is less and less super-
second process, chemical coupling will be developed in the next section. saturated with respect to B. From this point, one may conclude that both
The driving force for any chemical reaction is the deviation from rates can be coupled if some chemical elements are common in A and B.
equilibrium [30]: Indeed, the precipitation of solid B will contribute to increase the
undersaturation with respect to solid A, and, the dissolution of A will
μ Ai Nμ solution
i Nμ Bi ð6Þ contribute to increase the supersaturation with respect to B, and vice
versa. It means that during the transformation of solid A into solid B,
with μi the chemical potential of the species i common in A and B. The the composition of the solution is always between the solubility of A
last equation means there is a gradient of chemical potential(s) in the and the solubility of B. The evolution of the solution upon transforma-
solution from the surface of A to the solution and from the solution to tion is called the kinetic pathway and is generally plotted on an activ-
the surface of B, which allows the transport of species from A to B. ity/activity diagram.
When dissolution or precipitation of solid in water is concerned, the A suitable example of coupled reactions known in geochemistry is
equilibrium state is the solubility of solid in solution, and the deviations the transformation of albite into sanidine in KHCO3 solution:
from equilibrium are respectively called undersaturation and supersat-
uration for the dissolution and the precipitation reactions. At constant
NaAlSi 3 O8 þ Kþ →KAlSi 3 O8 þ Naþ : ð8Þ
temperature and pressure, the Gibbs free energy ΔG varies with the de-
viation from equilibrium. There are kinetic laws with different levels of
complexity, which link the dissolution or precipitation rates of minerals Experiments carried out by Alekseyev et al. [32] and further proc-
to ΔG. A very generic law can be expressed by the following equation essed and fitted by Zhu et al. [33,34] give an example of the coupling be-
[26,31]: tween albite dissolution and sanidine precipitation, which results in a
sigmoidal degree of reaction curve (Fig. 2a). Both reactions run at
Rate ¼ k  Asurf  f ðΔGÞ  gðai Þ ð7Þ equal rate from the 200th minute (Fig. 2b) and chemical balance is
4 L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11

Fig. 3. Schematic representation of the evolution of the saturation degrees with respect to C3S and C–S–H from the surface to the bulk with and without stirring (represented, respectively,
by dotted and full lines).

achieved in solution which always remains undersaturated with respect equals the C–S–H precipitation rate as soon as C–S–H nucleates,
to albite and supersaturated with respect to sanidine (Fig. 2c). More i.e., after less than one hour in typical paste conditions [42]. As a further
generally, the position of the kinetic path with respect to both solubility evidence for the coupling of the C3S dissolution and C–S–H precipitation
products, and its evolution over time, are of primary importance to find rates, it is well known that the composition of the solution does not
out what reactions (dissolution or precipitation) drive the overall trans- evolve suddenly during the hydration except within the first minutes
formation kinetics. Indeed, if the kinetic path is close to the solubility [24]. This is called the quasi-steady state; the dissolution and precipita-
product of the dissolving solid A and far away from the solubility prod- tion flows can be therefore considered equal at any time as described by
uct of B, the transformation kinetic is driven by the early precipitation of Eq. (5). The flow ϕ, either of dissolution or of precipitation, depends on
B (as an example, the authors cite the hydration of calcium sulfate hemi- (1) I, the reactive interface extent and (2) the reaction rate at the inter-
hydrate [35–37]). Conversely, if the kinetic path is close to the solubility face, called the interfacial rate Ri which varies with ΔG, the deviation
product of the precipitating solid B and far away from the solubility of A, from equilibrium.
the transformation is driven by the dissolution of A (as an example, the
authors cite the carbonation of wollastonite [38,39]).
The diffusion of reactants and/or products also has an effect upon the φ ¼ I  Ri ðΔGÞ or φ ¼ I  Ri ðβÞ since ΔG ¼ RT  ln β ð9Þ
transformation kinetics and is an active phenomenon in all solid-
solution reactions [40,41]. Nevertheless, diffusion significantly influ-
ences the overall kinetics only when the diffusion of at least one species where β is the degree of undersaturation for dissolution or the degree of
is sensibly slower than the interfacial reaction rate. In this particular supersaturation for precipitation.
case, the concentration of species close to the reactive interface will
drastically differ from the bulk. As an example, when ions released by
dissolution are not expelled quickly enough from the surface, they are
stored close to the surface and the undersaturation and dissolution
rate are accordingly reduced. Both undersaturation and supersaturation
will change compared to the case of rapid diffusion and the position of
the kinetic pathway in the activity/activity diagram as well. Nonethe-
less, the evolution towards one of the solubility products always indi-
cates that the transformation is kinetically driven by one of the
reactions [34].
Interestingly, C3S hydration reveals many features similar to the al-
bite–sanidine transformation. Indeed, the C3S dissolution rate rapidly

Table 1
Solubility products of C3S, C–S–H and portlandite. The reported Ksp corresponds to the
equilibria Eqs. (1), (2) and (3). * from [7] and ** from [53].

C3S⁎ β–C–S–H⁎⁎ γ–C–S–H⁎⁎ Ca(OH)2 Fig. 4. Evolution of alite content determined by QXRD during the hydration at 20 °C of an
ordinary Portland cement mixed at a water to cement ratio of 0.5. The hydration rate of
ln Ksp −50 −32.7 −39.3 −11.8
alite is calculated after smoothing and derivatizing the alite content curve.
L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11 5

ensured by diffusion and in dilute suspensions by diffusion and stirring.


Secondly, the addition of a large amount of seeds prevents C–S–H from
precipitating onto C3S grains. The hydration is accelerated but the hy-
dration curve remains qualitatively the same [45,46]. This observation
will be further discussed in the next section. It seems that the coverage
of the C3S surface by C–S–H does not modify the hydration sequences.
On the basis of these arguments, the composition of the solution at
the interfaces is related to the bulk solution, but not equal (see Fig. 3).
Thus, the degrees of saturation calculated from the measured bulk solu-
tion composition are apparent and the following relations βiC3S b βbulkC3S,
βiC–S–H N βbulkC–S–H and βiCH N βbulkCH hold. Two cases are considered
and depicted in Fig. 3. In (a), no C–S–H layer onto the C3S surface is
formed. This is representative of the very early hydration or when a sig-
nificant amount of seeds is initially added. There is a strong effect of stir-
ring on both effective saturation degrees at the interface but no effect on
the measured one. In (b), a gap exists between the C3S surface and the
Fig. 5. Evolution of the saturation indexes with respect to C3S solubility (undersaturation) C–S–H layer. The stirring has a very limited effect on the effective satu-
and with respect to C–S–H and Ca(OH)2 solubilities (supersaturations) during the
ration degrees at both C3S and inner surfaces, but decreases the effective
hydration of alite at w/c = 0.5 reported in [47] and shown in Fig. 4.
saturation degree at the outer C–S–H surface. Note that in this case, the
absolute apparent lnβC3S values measured in the bulk solution are
Then, the Eq. (5) is combined to the Eq. (9) to yield: higher than the effective value at the surface. On the contrary, the mea-
sured lnβC– S–H values are smaller than the effective ones.
Rhyd: ¼ IC3 S  RiC3 S  βiC3 S ¼ ICSH  RiCSH  βiCSH
3. Results from the literature
¼ ICaðOHÞ2  RiCaðOHÞ2  βiCaðOHÞ2 ð10Þ
The amount of published data that describe the evolution of ion con-
where IC3S, C–S–H, Ca(OH)2 denote the extent of C3S, C–S–H and portlandite centrations during the hydration of cement (or C3S) is scarce, especially
reactive interfaces and RiC3S, C–S–H, Ca(OH)2 the interfacial rates which de- in paste. In the following, a handful of papers has been carefully exam-
pend on βiC3S, CSH, Ca(OH)2 the under- and super-saturations at the re- ined. Overall, this study covers the hydration of cement (or C3S) at a
spective interfaces. Using this last equation could be risky without water to cement ratio ranging from 0.35 to 250. Data related to the hy-
further precaution. Indeed, serious complications may emerge because dration of cement in paste have been extracted from Lothenbach [47],
different possible interfaces could arise from the complex heteroge- Rothstein [48], Deschner [49], Longuet [50], and data related to the hy-
neous hydration process. Some of these difficulties have been reported dration of C3S in stirred diluted suspensions from Nicoleau [51] and
in the literature. First, the precipitation of hydration products can con- Garrault [52]. In this last experiment, the lime concentration in solution
stitute a partial or complete layer that hampers the release of ions into was maintained constant at 11 mM. In order to evaluate how the inter-
solution. In this case, the dissolving reactive interface is modified. The facial rates vary during hydration, the undersaturation index with re-
interface extent between C3S and water is lower in the case of a perfect spect to C3S solubility and the supersaturation indexes with respect to
contact between the anhydrous and hydrated phases. Second, a gap C–S–H and portlandite solubilities have been calculated. The saturation
might be assumed between the dissolving surface and the hydration index, SI, is defined as the logarithm of the ratio between the ion activity
products. In this case, the saturation indexes would be significantly dif- product Π and the solubility product Ksp, as follows:
ferent from those in the bulk solution.
Nevertheless, there are strong arguments that indicate that the pre- Π
SI ¼ ln β ¼ ln ¼ lnΠ− ln Ksp : ð11Þ
cipitation of hydration products or the diffusion of ions does not signif- Ksp
icantly influence how the hydration progresses, or in other words, it
does not change the overall hydration kinetics. As a matter of fact, C3S The logarithm of solubility products calculated according to Eq.(11)
hydration kinetics in paste or in stirred diluted suspensions are compa- at 20 °C can be found in Table 1. The value chosen for C3S is a reasonable
rable [43] or even alike when the hydration is started in saturated estimation taken from the kinetic study reported by Nicoleau et al. [7].
portlandite solution [44]. In paste, the advection of species is only This estimated value held fixed constant, but should vary slightly

Fig. 6. Evolution of the saturation indexes with respect to C3S, C–S–H and Ca(OH)2 calculated from the ion concentrations measured during the hydration of a white cement at w/c = 0.5
(A) and during the hydration of a gray cement at w/c = 0.35 (B) [48].
6 L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11

3.2. Data from Rothstein — hydration of alite in ordinary Portland and


white cement pastes

The second data set was reported from Rothstein et al. [48]. This set
consists of two hydration experiments. The first experiment is the hy-
dration of a gray cement at w/c = 0.35 and the second one is a white ce-
ment hydrated at w/c = 0.5. Unfortunately, the degree of hydration
over time is not available in the paper. The evolution of the saturation
indexes is reported in Fig. 6A for the white cement and Fig. 6B for the
gray cement. Concerning the white cement, the saturation indexes
slowly increase during the first hours and then the ion concentrations
move towards C–S–H solubility and depart from C3S solubility. The
same tendencies can be highlighted during the hydration of the gray ce-
ment except that the solution gets supersaturated with respect to
portlandite after at least 8 h.

Fig. 7. Evolution of the saturation indexes with respect to C3S, C–S–H and Ca(OH)2
3.3. Data from Longuet — hydration of alite in Portland cement paste
calculated from the ion concentrations measured during the hydration of a Portland
cement at w/c = 0.5. Data from [50].
In reference [50], Longuet was the first scientist to extract the pore
solution from cement pastes and to study its composition during the hy-
dration. The experiment, which is shown in Fig. 7, is the evolution of the
saturation indexes of silicate phases and portlandite during the hydra-
according to pH. Yet, at any value of Ksp, the variations of SI and the in-
tion of Portland cement (noted CPA69 in the article) at w/c = 0.5 and
terfacial rate are only determined by the variation of lnΠ.
at 20 °C. The solubility of C3S has been reached at 5 h and from this
point the undersaturation with respect to C3S decreased. This remains
true until one year, the last experimental point of the study.
3.1. Data from Lothenbach — hydration of alite in ordinary Portland ce-
ment, w/c = 0.5
3.4. Data from Deschner — hydration of alite in Portland cement paste
The first data set evaluated in this paper comes from Lothenbach
et al. [47] who kindly made available the numerical values. In this arti- This data set relates the hydration of ordinary Portland cement at a
cle, the concentration of ions has been measured by ICP optical emission water to cement ratio of 0.5 at 23 °C. In Deschner's article [49], a reason-
spectrometry and the degree of alite hydration by QXRD. In Fig. 4, the able estimation of the hydration rate of alite can be made with the cal-
degree of reaction is plotted versus time. After smoothing and orimetry curve. The evolution of the hydration rate is directly
derivatizing the curve, the hydration peak appears and shows a maxi- correlated to the heat flow released and shown in Fig. 8A. The so-
mum at roughly 6 h. The relatively poor resolution of the degree of hy- called hydration peak is characterized by an increase of the
dration curve and the inherent inaccuracy of the QXRD technique do not undersaturation with respect to C3S and a decrease of the saturation
allow the precise resolution of the hydration rate evolution. The calcula- with respect to C–S–H.
tion of the saturation indexes (Fig. 5) shows that the composition of the
solution approaches the solubility equilibrium of C3S during the first 3.5. Nicoleau's data — hydration of C3S at w/c = 250 started in a saturated
four hours and then deviates from this equilibrium until the end of ex- Ca(OH)2 solution
periment. The opposite conclusion can be drawn for C–S–H: during
the first hours the composition of the solution deviates from the C–S– This data set describes the hydration of pure C3S in dilute suspension
H solubility and then subsequently approaches it. From the first hour [51]. The hydration is started in a solution saturated with respect to
until the end of the experiment, the supersaturation with respect to portlandite. During the experiment, the concentration of calcium and
portlandite evolves only slightly. silicate ions (Fig. 9) as well as the conductivity have been recorded

Fig. 8. Calorimetry curve (A) and evolution of the saturation indexes (B) with respect to C3S, C–S–H and portlandite during the hydration of Portland cement (w/c = 0.5, T° = 23 °C), data
from [49].
L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11 7

Fig. 9. Evolution of calcium and silicate concentrations during the hydration of C3S started
in dilute solution (the liquid to solid ratio is 250) and saturated with respect to portlandite. Fig. 11. Evolution of the degree of hydration and hydration rate of C3S. The silicate
The hydration rate is calculated from the conductivity. Portlandite has not precipitated in concentration is also shown. The hydration is carried out in a solution regulated at
the experiment. constant calcium hydroxide concentration of 11 mM. The liquid to solid ratio is 50.

and the concentration of hydroxide ions calculated by charge balance. calcium hydroxide concentration is fixed, the saturation indexes for
The increase of conductivity is proportional to the increase of calcium C3S and C–S–H, exhibited in Fig. 12, vary only with the silicate concen-
hydroxide concentration and thus proportional to the degree of hydra- tration. Thus, the saturation index with respect to portlandite is not
tion [54]. The hydration rate can therefore be calculated from the con- shown. From the beginning until the end of experiment, the silicate con-
ductivity (Fig. 9). It is worthwhile pointing out that during the entire centration decreases, or in other words, the solution becomes more and
experiment, contrary to some other experiments presented in this more undersaturated with respect to C3S, and respectively less and less
paper, portlandite does not precipitate. The evolution of saturation in- supersaturated with respect to C–S–H. At roughly the same time as the
dexes with respect to C3S and C–S–H is plotted in Fig. 10 but is less hydration rate peak (at about 2.5 h), a more pronounced concentration
marked than in the three previous reported experiments. A decrease drop occurs.
of C3S supersaturation is observed at approximately 8 h, which corre- The path followed over time by the activities of species involved in
sponds to the time of the maximum hydration rate (Fig. 9). The super- the coupling of reactions is commonly called the kinetic path. For the
saturation with respect to C–S–H also decreases at that time. C3S hydration, this path is usually represented in the silica-lime dia-
gram. The kinetic paths of the reported experiments are plotted in
3.6. Data from Garrault — hydration of C3S, water to solid ratio of 50 started Fig. 13. This plot reveals the direction followed by the system,
in a solution maintained at constant Ca(OH)2 concentration of 11 mmol/L i.e., which reaction in the coupling tends to drive the overall kinetics.
In addition, some considerations can be made on the precipitation of
The last experiment from the literature is C3S hydration carried out portlandite. Indeed, the massive precipitation of portlandite is always
under controlled conditions reported by Garrault in her thesis [52]. accompanied by a decrease of the concentration in calcium hydroxide.
The C3S sample is hydrated in 11 mmol/L of calcium hydroxide at a liq- At least, we can identify at which experimental point the portlandite
uid to solid ratio of 50. The concentration is maintained constant during is certainly present when this data is not provided in the articles exam-
the entire experiment by means of a dedicated set-up. In particular, the ined here.
solution is aspirated through a filter and distilled water is injected when All the experiments at low w/c reveal the same behavior: the kinetic
the conductivity rises above the reference conductivity, i.e. the conduc- path moves from the left to the right-hand side as a result of the increase
tivity of the 11 mM calcium hydroxide solution. The amount of calcium of the activity product {Ca2+}{OH−}2, but only as long as portlandite
ions removed over time is monitored, which allows for calculation of does not precipitate. Then it goes from the right to the left-hand side
the degree of hydration. The silicate concentration has been measured after the precipitation of portlandite. In the same time, before the
by spectrophotometry. The results are presented in Fig. 11. As the portlandite precipitation, the supersaturation with respect to C–S–H in-
creases and decreases as soon as portlandite precipitates. In Deschner's
experiment, the left to right displacement is not observable because the
portlandite already precipitated before the first experimental point. This
is also what the evolution of the saturation degrees reveals in Figs. 5, 6,
and 8. In Garrault's experiment, the activity product {Ca2+}{OH−}2 is
experimentally fixed at a constant and only the move towards C–S–H
solubility is observed as conversely the increase of the deviation from
the C3S solubility (see Fig. 12). In Nicoleau's experiment, only the left
to right displacement of the kinetic path is observed because at such a
dilution (w/c = 250), the activity product {Ca2 +}{OH−}2 does not
reach the critical supersaturation with respect to portlandite. However,
in that case, the deviation from the C–S–H solubility first increases and
then decreases.

4. Discussion

The data reported above allow ascribing what reaction rate limits
Fig. 10. Evolution of the undersaturation with respect to C3S and the supersaturation with the C3S hydration during the different identified steps and the conse-
respect to C–S–H and portlandite during the experiment shown in Fig. 9. quences on the evolution of its reactive surface area.
8 L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11

being much higher than the latter in the very beginning of hydration. In
this short time window, the hydration is limited by the precipitation
rate. In the more peculiar Garrault's experiment, because the sample
was hydrated in low calcium hydroxide concentration during the entire
hydration, a lot of C–S–H nuclei [18] have been nucleated which pro-
vides enough early CSH surface area to enable the system to immedi-
ately enter a regime limited by the dissolution.

4.3. The acceleratory period (step III)

In all the experiments during which the acceleration of hydration oc-


curs without increase of the calcium hydroxide concentration, the
undersaturation with respect to C3S always increases (hand-in-hand
with a decrease of supersaturation with respect to C–S–H) indicating a
system getting limited by the dissolution reaction. It applies to all the
experiments in paste where the precipitation of portlandite occurs typ-
Fig. 12. Evolution of saturation indexes calculated with respect to C3S and C–S–H ically in the beginning of the acceleration and also in dilute suspensions
solubilities during the experiment shown in Fig. 11 and reported in [52].
when the calcium concentration is maintained constant (Garrault's ex-
periment). In Nicoleau's experiment, the undersaturation briefly de-
4.1. The deceleratory period (step IV) creases during the first hour and then increases during the
acceleratory step until the peak maximum. Started in saturated lime so-
In all the experiments, whatever the w/c ratio, for pure C3S or ce- lution, this experiment does not lead to a large enough C–S–H/solution
ment, or whatever the initial lime concentration, the supersaturation interface at the acceleration onset to consume all the ions released by
with respect to C–S–H always decreases during the deceleratory period the dissolution of C3S. This is probably because of the very low amount
(step IV). Conversely, the undersaturation with respect to C3S increases. of initial C–S–H nuclei in saturated lime solution [52].
Clearly, during the step IV the kinetic path is always moving towards the The conclusion of these experimental observations is that, in cement
C–S–H solubility and departs from the C3S solubility. That means the pastes, the dissolution is the limiting reaction during the major part of
precipitation instantaneously consumes more ions than the dissolution the hydration peak. In these circumstances, the hydration peak has to
releases. In other words, the hydration rate is limited by the dissolution be shaped by the dissolution process.
rate in the deceleratory step.
4.4. Evolution of the dissolution surface area during steps III and IV

4.2. The induction period (step II) According to Eq. (11), the hydration rate is equal to the product of Ri
(β), the interfacial rate and I, the reactive interface area for both reac-
This period does not necessarily occur, it is not the case in Garrault's tions, i.e. the C3S dissolution and the C–S–H precipitation. Note that it
experiment for instance. But if such a period appears, the supersatura- is also true for portlandite if it precipitates. In the previous paragraph,
tion with respect to C–S–H increases and the undersaturation with re- the role of C3S dissolution on shaping the hydration peak has been
spect to C3S decreases, because the dissolution releases more ions highlighted. When the dissolution becomes limiting, the
than the precipitation can consume. It is certainly due to the difference undersaturation and consequently the interfacial dissolution rate con-
in the dissolution and precipitation interfacial surface areas, the former tinuously increase, that implies a non-monotonic variation of the C3S/

Fig. 13. Kinetic paths of the various experiments shown in this article are plotted in the Ca(OH)2-SiO2 solubility diagram. The solubilities of the relevant phases, i.e. C3S, C–S–H and
portlandite are represented. Within each path, the arrows indicate the direction of the time evolution. The concentration in calcium hydroxide cannot exceed about 33 mM without
instantaneous precipitation of portlandite. This limit defines a critical Ca(OH)2 activity product range, symbolized by a dashed box, where portlandite must precipitate. 33 mM
corresponds to the right-hand side of the box and to an immediate precipitation, the left-hand side corresponds to a lower metastability of the solution with respect to portlandite and
this latter should precipitate only after a couple of minutes. Within a single experimental data set, the identification of portlandite precipitation is pinned with a bigger dot. This point
corresponds to an inversion of the path direction.
L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11 9

Fig. 14. On the top-left corner, evolution of the heat flow during the hydration of alite in the presence of 2% magnesium chloride at a water to solid ratio of 0.5. Cryo-SEM investigation has
been performed and samples studied after 30 min ①, 2 h ② and finally 4 h ③. Some representative images ①, ② and ③ highlight the precipitation of C–S–H on alite grains. In ① some C–S–
H is already visible on the surface of the grain; amorphous brucite Mg(OH)2 has also precipitated. In ②, the height of the C–S–H layer is at least 180 nm, and in ③ about 600 nm.

solution reactive interface which has first to increase and then to de- 4.5. Late hydration
crease as a result of the increase and decrease of the hydration rate.
The increase of the dissolution surface area can appear surprising at The various elements raised previously converge towards the fact
first sight, since a monotonic decrease of the surface area is expected that alite dissolution is of paramount importance for shaping the hydra-
when an isotropic dissolution of grains is assumed. Yet, the dissolution tion curve, and that C–S–H precipitation is never the limiting step right
of C3S is characterized by the formation of etch-pits and it has been pro- after the end of the acceleratory step. The relative importance of both
posed that the extension of pits and their subsequent coalescence leads reactions is still to be determined for the late hydration. It could be con-
to non-monotonic variation of the surface area of dissolution [9]. A re- ceivable that, at relatively low w/c, C–S–H could be constrained in space,
cent NMR study supports the non-monotonic variation of the C3S sur- and then its precipitation would be even more limiting than the C3S dis-
face area and the role of etching [55]. The decrease of the dissolution solution rate. In such a case, due to the crystallization pressure, the su-
reactive interface can also result from the expected decrease of the over- persaturation would be expected to increase. Yet, Rothstein's,
all grain surface area. It could be argued too that the decrease of the C3S/ Lothenbach's and Longuet's data do not show any move back towards
solution interface results from the precipitation of C–S–H onto the C3S C3S solubility or conversely a deviation from C–S–H solubility even
surface as postulated in [23]. However, some studies show that alite after days or years of hydration. It indicates that C–S–H precipitation
grains are nearly covered with hydration products before the decelera- would never be limiting the hydration, and that the C3S dissolution
tion [20]. Another convincing example is shown in Fig. 14 which reports would be always the limiting kinetic step in the coupling. This conclu-
C3S hydration accelerated by the addition of soluble magnesium salt sion is not contradicted by the similar apparent activation energies re-
(leading to the precipitation of brucite). Cryo-SEM pictures taken during cently measured for the sole alite dissolution (~ 48 kJ/mol) [57], or
the acceleratory step III indicate that alite grains are covered by signifi- calculated for the sole alite dissolution (~53 kJ/mol) [58] and measured
cant amounts of C–S–H already at the beginning of hydration, and that a for the overall hydration (~51 kJ/mol) [59].
thick C–S–H layer is even formed before the deceleratory step. It has to be also mentioned that Thomas clearly demonstrated in this
On the contrary, when C3S grain surfaces are quasi-devoid of C–S–H last article that the apparent activation energy of C3S hydration is con-
as it is the case in the presence of large amounts of C–S–H nuclei [9,56] stant from the earliest age of hydration until at least 70% of hydration,
the typical shape of the hydration rate curve is maintained (Fig. 15) which tends to indicate that the same process always kinetically con-
whatever their amount. The nuclei lead to an increase of the C–S–H/so- trols the hydration. The fact that Thomas' value is very close to Juilland
lution interface area allowing the decrease of the supersaturation with et al.'s value, which was obtained for pure dissolution experiments fur-
respect to C–S–H and the increase of the undersaturation with respect ther supports our conclusions that the C3S dissolution becomes very
to C3S according to scheme (a) in Fig. 3. early the slowest step of hydration. Finally, it seems appropriate to
10 L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11

Acknowledgment

Luc Nicoleau is grateful to Stephen Farrington (BASF Corp. Cleve-


land) and Bruce Christensen (BASF SE) for the careful reading of this
paper. Luc Nicoleau thanks Sebastian Bergold for the stimulating discus-
sion and both authors acknowledge BASF Construction Solutions and
CNRS for the support necessary to carry out this analysis.

References
[1] A.J. Allen, J. Thomas, H.M. Jennings, Composition and density of nanoscale calcium-
silicate-hydrate in cement, Nat. Mater. 6 (2007) 311–316.
[2] A.C.A. Muller, K.L. Scrivener, A.M. Gajewicz, P.J. McDonald, Densification of C–S–H
measured by 1H NMR relaxometry, J. Phys. Chem. C 117 (2013) 403–412.
[3] H. Le Chatelier, Experimental Researches on the Constitution of Hydraulic Mortars,
McGraw Editions, New York, 1905.
Fig. 15. Hydration of alite at water to solid ratio of 0.5 with additions of calcium silicate
[4] J.J. Thomas, J.J. Biernacki, J.W. Bullard, S. Bishnoi, J.S. Dolado, G.W. Scherer, A. Lüttge,
hydrate nuclei. ξat peak, the degree of hydration reached at the maximum rate is Modeling and simulation of cement hydration kinetics and microstructure develop-
indicated beside the legend. ment, Cem. Concr. Res. 41 (12) (2011) 1257–1278.
[5] J.W.:. Bullard, H.M. Jennings, R.A. Livingston, A. Nonat, G.W. Scherer, J.S. Schweitzer,
K.L. Scrivener, J.J. Thomas, Mechanisms of cement hydration, Cem. Concr. Res. 41
(12) (2011) 1208–1223.
highlight another conclusion that arose in Thomas' article. Indeed, as [6] K.L. Scrivener, A. Nonat, Hydration of cementitious materials, present and future,
Cem. Concr. Res. 41(12), 651–665.
diffusion processes are rather insensitive to temperature, the activation [7] L. Nicoleau, A. Nonat, D. Perrey, The di- and tricalcium silicate dissolutions, Cem.
energies for diffusion are normally low. Since no significant decrease of Concr. Res. 47 (2013) 14–30.
the apparent activation energy is seen, at least until 70% of degree of hy- [8] L. Nicoleau, E. Schreiner, A. Nonat, Ion-specific effects influencing the dissolution of
tricalcium silicate, Cem. Concr. Res. 59 (2014) 118–138L.
dration, it shows that diffusion is not controlling the hydration rate. [9] M.A.B. Nicoleau, Analytical model for the alite (C3S) dissolution topography, J. Am.
Ceram. Soc. (2015), http://dx.doi.org/10.1111/jace.13647 (published online).
[10] H.N. Stein, J.M. levels, , Influence of silica on the hydration of 3CaO, SiO2, J. Appl.
Chem. 14 (1964) 338–346.
5. Conclusion [11] E.M. Gartner, H.M. Jennings, Thermodynamics of calcium silicate hydrates and their
solutions, J. Am. Ceram. Soc. 70 (10) (1987) 743–749.
In this paper, the evaluation of various pieces of published data en- [12] F. Begarin, G. Garrault, A. Nonat, L. Nicoleau, Hydration of lite containing aluminum,
Adv. Appl. Ceram. 110 (2011) 127–130.
abled a kinetic analysis of the chemical coupling established between [13] M. Avrami, Kinetics of phase change. I. General theory, J. Chem. Phys. 7 (1939)
C3S dissolution and C–S–H precipitation during C3S hydration. In partic- 1103–1112.
ular, based upon well-established principles, the evolution of ion con- [14] M. Avrami, Kinetics of phase change. II. Transformation—time relations for random
distribution of nuclei, J. Chem. Phys. 8 (1940) 212–224.
centrations indicates what limits the hydration kinetically. In the case
[15] M. Avrami, Kinetics of phase change. III. Granulation, phase change, and microstruc-
of C3S hydration, the deceleration start delineating the transition be- ture, J. Chem. Phys. 9 (1941) 177–184.
tween hydration steps III and IV, is always accompanied by a move of [16] J.J. Thomas, A new approach to modeling the nucleation and growth kinetics of
ion concentrations towards C–S–H solubility and a deviation from C3S tricalcium silicate hydration, J. Am. Ceram. Soc. 90 (10) (2007) 3282–3288.
[17] S. Bishnoi, K.L. Scrivener, Studying nucleation and growth kinetics of alite hydration
solubility. Thus, the deceleration is caused by dissolution process. using μic, Cem. Concr. Res. 39 (2009) 849–860.
During the acceleratory step III, the kinetic analysis also indicates [18] S. Garrault, A. Nonat, Hydrated layer formation on tricalcium and dicalcium silicate
that the dissolution is limiting and it highlights the necessary develop- surfaces: experimental study and numerical simulations, Langmuir 17 (2001)
8131–8138.
ment of the dissolution reactive interface to explain the acceleration of [19] K. Ioannidou, R.J.M. Pellenq, E. Del Gado, Controlling local packing and growth in
hydration. Both the increase of dissolution interface, as well as the sub- calcium-silicate-hydrate gels, Soft Matter 20 (2014) 1121–1133.
sequent deceleration, can be rationally explained by the formation and [20] A. Bazzoni, Study of Early Hydration Mechanisms of Cement by Means of Electron
Microscopy, Thesis of the EPFL (Lausanne), Switzerland, 2014.
coalescence of etch-pits. This new approach is in stark contrast with [21] M. Etzold, P.J. McDonald, A.F. Routh, Growth of sheets in 3D confinements — a
the usual models based on C–S–H growth. Furthermore, experimental model for the C–S–H mesostructure, Cem. Concr. Res. 63 (2014) 137–142.
evidence invalidates the role of C–S–H growth in the deceleration of [22] G.W. Scherer, Models of confined growth, Cem. Concr. Res. 42 (2012) 1252–1260.
[23] S. Garrault, L. Nicoleau, A. Nonat, Tricalcium silicate hydration modeling and numer-
the hydration. The conclusions drawn in this examination of literature ical simulations, Proceedings of the RILEM International Symposium on Concrete
data seem to be valid for a large binder composition range varying Modeling CONMOD'10, Lausanne, Switzerland June 2010, pp. 22–25, , http://dx.
from pure C3S to gray cement, and for various dilution conditions. In ad- doi.org/10.13140/RG.2.1.3622.2480.
[24] P. Barret, D. Bertrandie, Fundamental hydration kinetic features of the major cement
dition, the precipitation of portlandite is neither required for the accel-
constituents: Ca3SiO5 and βCa2SiO4, J. Chim. Phys. 83 (1986) 765–775.
eration nor for the deceleration but always leads to the immediate [25] J.W. Bullard, G.W. Scherer, J.J. Thomas, Time dependent driving forces and the kinet-
limitation by the dissolution reaction. Finally, there is no indication ics of tricalcium silicate hydration, Cem. Concr. Res. 74 (2015) 26–34.
that the hydration process becomes controlled by any other processes [26] A.C. Lasaga, J.M. Soler, J. Ganor, T.E. Burch, K.L. Nagy, Chemical weathering rate laws
and global cycles, Geochim. Cosmochim. Acta 58 (10) (1994) 2361–2386.
than C3S dissolution, even at later ages, neither by C–S–H growth nor [27] D. Nahon, E. Merino, Pseudomorphic replacement in tropical weathering: evidence,
by diffusion. geochemical consequences, and kinetic-rheological origin, Am. J. Sci. 297 (1997)
Some new perspectives arise from those statements. First, as the sur- 393–417.
[28] C.M. Pina, L. Fernández-Díaz, M. Prieto, A. Putnis, In situ atomic force microscope ob-
face dissolution topography contributes to the acceleratory step, the servations of a dissolution-crystallisation reaction: the phosgenite-cerussite trans-
prominent role of the C–S–H growth process becomes somewhat un- formation, Geochim. Cosmochim. Acta 64 (2) (2000) 215–221.
clear. Indeed, the concept of the autocatalytic growth [60], which stipu- [29] E. Merino, T. Dewers, Implications of replacement for reaction-transport modeling, J.
Hydrol. 209 (1998) 137–146.
lates that the C–S–H surface catalyzes precipitation, has never been [30] I. Prigogine, R. Defay, Chemical Thermodynamics, Longmans, Green and Co., London,
experimentally demonstrated and should be a future exciting topic. Sec- 1965.
ond, many interpretations about the effects of a large variety of addi- [31] A.C. Lasaga, Kinetic Theory in Earth Sciences, Princeton Series in Geochem, Princeton
Univ. Press, Princeton, New Jersey, 1998.
tives, which influences the hydration, for instance, were mostly
[32] V.A. Alekseyev, L.S. Medvedeva, N.I. Prisyagina, S.S. Meshalkin, A.I. Balabin, Change
grounded on modifications of the nucleation and growth of C–S–H. in the dissolution rates of alkali feldspars as a result of secondary mineral precipita-
For sure, the effects of these additives on C3S surface dissolution deserve tion and approach to equilibrium, Geochim. Cosmochim. Acta 61 (6) (1997)
1125–1142.
a closer look.
L. Nicoleau, A. Nonat / Cement and Concrete Research 86 (2016) 1–11 11

[33] C. Zhu, P. Lu, Alkali feldspar dissolution and secondary mineral precipitation in batch [46] L. Nicoleau, T. Gädt, L. Chitu, G. Maier, O. Paris, Oriented aggregation of calcium silicate
systems: 3. Saturation states of product minerals and reaction paths, Geochim. hydrate platelets by the use of comb-like copolymers, Soft Matter 9 (2013) 4864–4874.
Cosmochim. Acta 73 (2009) 3171–3200. [47] B. Lothenbach, F. Winnefeld, Thermodynamic modelling of the hydration of
[34] C. Zhu, P. Lu, Z. Zheng, J. Ganor, Alkali feldspar dissolution and secondary mineral Portland cement, Cem. Concr. Res. 36 (2006) 209–226.
precipitation in batch systems: 4. Numerical modeling of kinetic reaction paths, [48] D. Rothstein, J.J. Thomas, B.J. Christensen, H.M. Jennings, Solubility behavior of Ca-,
Geochim. Cosmochim. Acta 74 (2010) 3963–3983. S-, Al- and Si-bearing solid phases in Portland cement pore solutions as a function
[35] U. Ludwig, N.B. Singh, Hydration of hemihydrate of gypsum and its supersaturation, of hydration time, Cem. Concr. Res. 32 (2002) 1663–1671.
Cem. Concr. Res. 8 (1978) 291–300. [49] F. Deschner, F. Winnefeld, B. Lothenbach, S. Seufert, P. Schwesig, S. Dittrich, F. Goetz-
[36] F. Brandt, D. Borsbach, Bassanite (CaSO4, 0.5 H2O) dissolution and gypsum (CaSO4, 2 Neuenhoeffer, J. Neubauer, Hydration of Portland cement with high replacement by
H2O) precipitation in the presence of cellulose ethers, J. Cryst. Growth 233 (2001) fly ash, Cem. Concr. Res. 42 (10) (2012) 1389–1400.
837–845. [50] P. Longuet, L. Burglen, A. Zelwer, La phase liquide du ciment hydraté, Ciments et
[37] G. Dumazer, V. Narayan, A. Smith, A. Lemarchand, Modeling gypsum crystallization bétons, Rev. Mater. Constr. (1973).
on a submicrometric scale, J. Phys. Chem. C 113 (2009) 1189–1195. [51] L. Nicoleau, Physico-chemical interactions between latex dispersions and cement
[38] D. Daval, I. Martinez, J. Corvisier, N. Findling, B. Goffé, F. Guyot, Carbonation of Ca- phases, Thesis of the University of Burgundy (Dijon), France, 2004.
bearing silicates, the case of wollastonite: experimental investigations and kinetic [52] S. Garrault-Gauffinet, Etude expérimentale et par simulation numérique de la
modeling. cinétique de croissance et de la structure des hydrosilicates de calcium, produits
[39] D. Daval, I. Martinez, J.-M. Guigner, R. Hellmann, J. Corvisier, N. Findling, C. Dominici, d'hydratation des silicates tricalcique et dicalcique, Thesis of the University of Bur-
B. Goffé, F. Guyot, Mechanism of wollastonite carbonation deduced from micro- to gundy (Dijon), France, 1998.
nanometer length scale observations, Am. Mineral. 94 (2009) 1707–1726. [53] J. Haas, A. Nonat, From C–S–H to C–A–S–H: experimental study and thermodynamic
[40] A.C. Lasaga, Metamorphic reaction rate laws and development of isograds, Mineral. modelling, Cem. Concr. Res. 68 (2015) 124–138.
Mag. 50 (1986) 359–373. [54] P.W. Brown, E. Franz, G. Frohnsdorff, H.F.W. Taylor, Analyses of the aqueous phase
[41] W.M. Murphy, E.H. Oelkers, P.C. Lichtner, Surface reaction versus diffusion control of during the early C3S hydration, Cem. Concr. Res. 14 (1984) 257–262.
mineral dissolution and growth rates in geochemical processes, Chem. Geol. 78 [55] E. Pustovgar, R.P. Sangodkar, A.S. Andreev, M. Palacios, B.F. Chmelka, R.J. Flatt, J.-B.
(1989) 357–380. d.'’E. de Lacaillerie, Nat. Commun. 7 (2016) 10952.
[42] S.T. Bergold, F. Goetz-Neuenhoeffer, J. Neubauer, Quantitative analysis of C–S–H in [56] L. Nicoleau, New calcium silicate hydrate network, Transp. Res. Rec. 2 (2010) 42–51.
hydrating alite pastes by in-situ XRD, Cem. Concr. Res. 53 (2013) 119–126. [57] P. Juilland, E. Gallucci, Morpho-topological investigation of the mechanisms and ki-
[43] D.M. Kirby, J.J. Biernacki, The effect of water-to-cement ratio on the hydration kinet- netic regimes of alite dissolution, Cem. Concr. Res. 76 (2015) 180–191.
ics of tricalcium silicate cements: testing the two-step hydration hypothesis, Cem. [58] S.A. Grant, G.E. Boitnott, C.J. Korhonen, R.S. Sletten, Effet of temperature on hydra-
Concr. Res. 42 (2012) 1147–1156. tion kinetics and polymerization of tricalcium silicate in stirred suspensions of
[44] D. Damidot, A. Nonat, P. Barret, Kinetics of tricalcium silicate hydration in diluted CaO-saturated solutions, Cem. Concr. Res. 36 (2006) 671–677.
suspensions by microcalorimetric measurements, J. Am. Ceram. Soc. 73 (11) [59] J.J. Thomas, The instantaneous apparent activation energy of cement hydration mea-
(1990) 3319–3322. sured using a novel calorimetry-based method, J. Am. Ceram. Soc. 95 (10) (2012)
[45] J.J. Thomas, H.M. Jennings, J.J. Chen, Influence of nucleation seeding on the hydration 3291–3296.
mechanisms of tricalcium silicate and cement, J. Phys. Chem. C 113 (11) (2009) [60] E.M. Gartner, J.M. Gaidis, Hydration Mechanisms I, Materials Science of Concrete,
4327–4334. The American Ceramic Society, Westerville, 1989.

Potrebbero piacerti anche