Vous êtes sur la page 1sur 10

Chemical Geology 381 (2014) 144153

Contents lists available at ScienceDirect

Chemical Geology
journal homepage: www.elsevier.com/locate/chemgeo

Calcium carbonate and calcium sulfate precipitation, crystallization


and dissolution: Evidence for the activated steps and the mechanisms
from the enthalpy and entropy of activation values
Athinoula L. Petrou , Athina Terzidaki
Laboratory of Inorganic Chemistry, Department of Chemistry, University of Athens, Panepistimioupolis, 15771 Athens, Greece

a r t i c l e

i n f o

Article history:
Received 30 January 2014
Received in revised form 15 May 2014
Accepted 15 May 2014
Available online 22 May 2014
Editor: J. Fein
Keywords:
Free energy of activation
Enthalpy of activation
Entropy of activation
EyringPolanyi equation
Calcium carbonate
Calcium sulfate

a b s t r a c t
Calculation of the thermodynamic parameters Eact, H, S and G for the precipitation, crystallization and
dissolution processes of the salts CaCO3 and CaSO4, leads to very important conclusions about the activated
steps and the mechanisms. The G values are almost the same for all the processes at the same temperature, sug2+
and SO2
gesting that the electrostatic forces between the ions Ca2+ and CO2
3 (CaCO3) and Ca
4 (CaSO4) are the
most important factors governing the above processes. The values of H and S differ but the values of G
which refer to the overall transformations are the same and are independent of the various steps that take place
as well as the mechanisms (associative, dissociative). Dehydration and aquation of the ions are revealed by the
values of H and S. The precipitation of the two salts may take place both by a dissociative mechanism
(S N 0) or by an associative mechanism (S b 0). For processes taking place without the need for diffusion
2
1
, whereas when diffusion of
of the ions, (Ca2+, CO2
3 , SO4 ), the free energy of activation is about 85 kJ mol
1
the ions is necessary, an additional amount of ~20 kJ mol is required. This amount is the activation energy for
the diffusion. In the case of CaCO3, a wide range of values is found for H from 66.00 to 162.00 kJ mol1,
and for S from 501.00 to +248.00 J K1 mol1 while the G values cover only a small range from 75 to
90 kJ mol1. Values of 120, 131 and 132 kJ mol1 are reported for cases where retardation is caused due to the
presence of foreign compounds. In the case of CaSO4, a wide range of values is found for H from 6.00 to
122.00 kJ mol1 and for S from 342.00 to +117.00 J K1 mol1 while the G values fall in the narrow
range from 80.00 to 89.00 kJ mol1. A value of 126 kJ mol1 is reported for cases where retardation is caused
due to the presence of foreign compounds. The Eact values vary between 63 and 164 kJ mol1 for CaCO3 and between 8 and 184 kJ mol1 for CaSO4, demonstrating once again that the G value is more realistic, being almost
the same for similar processes. The various small differences for the values of G arise from the different ionic
strengths due to the concentration and charge of the foreign ions affecting the rate constants and thus the activation parameters. The pH also has an effect, as does the nature of the solvent. The very large and very small absolute
values of H and S suggest composite reactions. Composite reactions in the precipitation, crystallization and
dissolution processes are the ones where dehydration takes place (positive values of H and S) followed by
association of the ions (negative values of H and S). The algebraic sum of the relevant H and S values
gives the total value of H and S. Activation energy values reported in the literature for certain cases compared to the values reported in the presence of adducts do not indicate retardation. On the contrary, they suggest
acceleration of the reactions. The use of G instead of the Eact values is more realistic in showing the large retardation effect. The presence of salt, for example NaCl, causes retardation of the crystallization process.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Calcium carbonate (CaCO3) is an important material that occurs in
the natural environment as a constituent of many biological, geological
and ecological systems, while it is also involved in many industrial processes, for example in scale formation. For these reasons, precipitation,
crystallization and dissolution of calcium carbonate in aqueous
Corresponding author.
E-mail address: athpetrou@chem.uoa.gr (A.L. Petrou).

http://dx.doi.org/10.1016/j.chemgeo.2014.05.018
0009-2541/ 2014 Elsevier B.V. All rights reserved.

solutions has attracted the interest of many investigators in a wide variety of elds. Many studies have focused on the problems that are related
to the mechanism of its formation, and the role that various parameters
play in its formation in aqueous solutions has been studied (Koutsoukos
and Kontoyannis, 1984). Thermal decomposition has also been studied
by several investigators and various mechanisms have been reported
(Rajeswara Rao, 1993, and refs. therein).
Calcium sulfate (CaSO4) has also been intensively investigated since
many aspects of its reaction mechanisms remain unanswered
(Klepetsanis et al., 1999). Solution chemistry helps greatly in the

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

elucidation of the reaction mechanisms of the various calcium sulfate


forms (hydrates) that are formed in aqueous media. The precipitation
process includes nucleation and crystal growth that are two mechanistically different steps. A knowledge of the calcium sulfate precipitation
mechanism has also been very important for the effective prevention
of calcium sulfate scale (Klepetsanis and Koutsoukos, 1991). A kinetic
study has been conducted to measure and model the rate of calcium sulfate precipitation within porous media (Merdhah and Yassin, 2008).
In this work we aim to contribute to the efforts to solve the serious
problems that are related to CaCO3 and CaSO4 precipitation, crystallization and dissolution by providing values for the thermodynamic parameters in order to assist in the clarication of the corresponding
mechanisms. These parameters are obtained by the treatment of literature kinetic data using the EyringPolanyi equation. Thus we will
provide an alternative analysis of well-documented examples from
the literature on calcium carbonate and calcium sulfate precipitation,
crystallization and dissolution. Based on the values for enthalpy and
entropy of activation (H and S), we will provide evidence for
the existence of dehydration and aquation (hydration) steps and we
will also show that some processes of the mechanisms are composite.
We will also show that in some cases (depending on the experimental
conditions) the diffusion of the ions is one of the controlling factors in
the activation step, adding approximately 20 kJ mol1 to the value of
the free energy of activation. From the values of the free energy of
activation (G), we draw conclusions concerning the true energy
requirements of the processes.
It has been shown (Petrou, 2012) that the combination of enthalpy
and entropy of activation in the free energy of activation gives a more
realistic value of the energy requirements of the activation step that the
processes need in order to take place. The analysis has also shown that
similar processes have similar G values (Petrou, 2012). A detailed analysis has also been presented on how the values of S may be used to
suggest a mechanism for a process (Petrou and Economou-Eliopoulos,
2009b).
Our suggestion (Petrou, 2012) of using the corrected energy value
that is critical for geochemical processes, the G value, instead of the
Eact value, provides a way to address unexplained or queried geochemical results, and to give solutions to existing problems: (a) Based on the
Eact values authors suggest mechanisms. As we have already shown
(Petrou, 2012), the Eact value is not a realistic value, i.e. it does not
give the true energy requirements in geochemical processes. This
implies that the mechanisms that have been proposed based on the
Eact values (or the rate constants) may not be correct [Table 2, case
(o)]. This problem has been solved by introducing the G values. (b)
An explanation has also been given of why calculated activation energy
values (that are related only to the enthalpy of activation values) for
certain transformations deviate from the expected and observed energy
requirements that characterize the processes when the entropic component is substantial. Values of Eact that are intended to explain retardation of processes indicate acceleration instead [Table 1, case (e), Table 2,
case (q)]. This has also been solved by introducing G. (c) Different
values of Eact have been proposed in the literature for the various processes (precipitation, crystallization, dissolution) leading to various
conclusions. However, the values of G are similar and point to the
same conclusions. From the entropies of activation an insight into the
degrees of freedom of the activated complexes can be obtained and
hence information about the mechanisms of the processes may be
extracted (Petrou and Economou-Eliopoulos, 2009b).
2. The Eyring or activated complex theory relation free energy of
activation
The transition state or activated complex theory states that upon approach of reactant molecules some bonds start to lengthen while
other bonds start to form. These changes are accompanied by energy
changes and, nally, the reacting molecules achieve a specic

145

arrangement before they can be transformed into the products of the reaction. The specic, transient, arrangement that possesses a denite energy is known as the transition state. The transition state or activated
complex theory postulates that the reaction proceeds through formation of an activated complex that is in equilibrium with the reactants
(Petrou and Economou-Eliopoulos, 2009b; Petrou, 2012), and the rate
constant is given by:



k RT=Nh exp G =RT :

The transmission coefcient is generally close to unity and may be


ignored (Espenson, 1981). G is the Gibbs free energy of activation, h
is Planck's constant, R is the Universal gas constant and N is the
Avogadro's number. The term RT/Nh is independent of the nature of
the reactants and, at a given temperature, the rate of any reaction is
determined by G. By applying the thermodynamic relationship

H TS :

Eq. (1) can be expressed in terms of the activation enthalpy, , and


the activation entropy, S, thus:





k RT=Nh exp S =R exp H =RT :

The above Eq. (3) is the general form of the Eyring equation or Activated Complex Theory equation, which is also known as EyringPolanyi
equation in chemical kinetics. It relates the reaction rate constant to
temperature and its linear form is:





ln k=T =R 1=T ln R=Nh S =R :

This linear form is usually used for the suggestion of a mechanism of


a certain reaction in the following way: the reaction is performed at various temperatures and the reaction rate constant is measured. The plot
of ln (k/T) versus 1/T gives a straight line with slope H/R
from which the Enthalpy of Activation is derived and with intercept
ln(R/Nh) + S/R from which the Entropy of Activation is calculated.
The term ln (R/Nh) has a known constant value 23.76. The calculated
value of the Entropy of Activation suggests an Associative mechanism
when S is negative, a Dissociative mechanism when S is positive,
an Interchange mechanism when S is equal to zero (when the rate
constants refer to replacement reactions) (Petrou and EconomouEliopoulos, 2009b). Calculation of G is possible for the appropriate
reaction temperatures according to the Eq. (2).
3. Diffusion-controlled reactions
Reactions in solution in which every collision of the reacting
molecules leads to products are called diffusion-controlled reactions.
Their rates are limited only by diffusion that has activation energy
values ranging from 10 to 20 kJ mol1. The rate-constants for such (second order) reactions range from 109 to 1010 M1 s1 and are dependent
on the nature of the solvent (Katakis and Gordon, 1987).
Alkattan et al. (1998), have reported an apparent activation energy
for diffusion of 19 4 kJ mol1 which is somewhat higher than the
13.2 kJ mol1 obtained by Oelkers and Helgeson (1988) for the apparent activation energy of HCl diffusion. The differences between these results were assigned to the hydrodynamics of the different experimental
setups used to obtain the rates (Alkattan et al., 1998).

146

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

4. Application of the EyringPolanyi equation to calcium carbonate


and calcium sulfate precipitation, crystallization and dissolution
processes
We have applied the EyringPolanyi equation to literature data and
have calculated H, which is related to the activation energy Eact
(Petrou and Economou-Eliopoulos, 2009a) by the equation

H
Eact RT;

S, which is related to A of the Arrhenius equation, and G, which


combines H and S (G = H TS, Eq. (2)).
It has been shown that similar processes have similar G values (at
the same temperatures) but may have different enthalpic and entropic
contributions (Petrou, 2012). We will present and compare these activation parameters (along with the Eact) that have been calculated
from data reported in the literature for precipitation, crystallization
and dissolution of CaCO3 (Table 1) and CaSO4 (Table 2).

5Table 1

Kinetic data (k, T) and calculated thermodynamic parameters (Eact, H, S, G) for precipitation, crystallization and dissolution of calcium carbonate (CaCO3).
T (K)
1. Precipitation in aqueous solutions
Case (d)

Case (e)

Case (e*)

Case (f)

2. Crystallization in aqueous solutions


Case (g)

Case (h)

Case (i)

Case (j), kRS

Case (j), kRM


Case (k), calcite crystallization

Case (k), calcite nucleation (from induction times)

3. Dissolution in aqueous solutions


Case (l)

Case (m)

Case (n)
Carrara marble

Case (n)
Iceland spar

Case (o), strained

Case (o), normal

298.15
308.15
318.15
283
298
313
278
298
310
335.7
318.7
302.3

lnk(a)

1
E(b)
act kJ mol

H(c) kJ mol1

S(c) J K1 mol1

G(c) kJ mol1

5.12
4.446
1.485
5.265
3.932
2.318
5.10
4.78
4.483
15.61
16.76
17.91

142.2(c)
155(d)

139.6

177.8

84.8 at 308 K

72.1(c)
71.2(e)

69.7

42.5

82.8

13.52(c)

11.09

246.9

87.13

58.1(c)
46(f)

55.43

210.4

120.2 at 308 K

57.04(c)
57.1(g)

54.5

66.4

75 at 308 K

43.5(c)
43.1 3.8(h)

41.0

135.3

82.7 at 308 K

46.7(c)
28.9(c)
46 4(i)

44.2
26.5

120.3
181.9

81.3
82.5

163.9(c)

161.4

247.6

85.1

63.03(c)

65.52

501.2

88.86

73.6(c)
66 2(k)

71

40.78

58.44

73.7(c)
73 10(k)

71.3

59.75

89.7

34.7(c)
35.1(l)

32.2

233.2

104.0 at 308 K

20.22(c)
(m)

17.5

207.5

81.4 at 308 K

59.1(c)
54 4(n)

56.5

88.9

83.9 at 308 K

45.14(c)
46 4(n)

42.5

145.5

87.3 at 308 K

388.7

130.84

284.4
294.7
298
306.8
315.5
284
293.5
302.1
312.7
298
283
313
298
298
300.5
306
293
298
306
301
312,5
323
283
293
298

1.715
0.759
0.58
0.223
0.621
4.211
3.70
3.158
2.532
3.139
3.758
1.966
2.871
5.952
5.298
4.2
4.001
4.343
5.09
6
7
8
7.97
6.84
6.67

278
293
303
313
323
298
323
353
298
314
328
335
298
314
328
335
276
298
353
276
298
353

12.786
11.564
11.418
10.82
10.68
2.659
1.833
1.386
4.075
2.779
1.871
1.431
5.167
4.382
3.59
3.137
22.384
21.59
21.011
22.76
22.00
21.19

13.72(c)
11.12

383.6
16.1(c)

13.5

131.6

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

4.1. Processes involving CaCO3


4.1.1. Precipitation of CaCO3 in aqueous solutions
Calcium carbonate is encountered in several polymorphic crystalline
phases including its hexa and monohydrate forms, vaterite, aragonite
and calcite (Dalas and Koutsoukos, 1990). These differ in solubility
and they demonstrate sensitivity to temperature, pH and the presence
of foreign ions or compounds.
The precipitation of calcium carbonate in aqueous solutions depends
on solution supersaturation, and it has been studied at supersaturations
sufcient for spontaneous precipitation (Koutsoukos and Kontoyannis,
1984). This implies that the activation energy for the diffusion of ions
is not needed. The diffusion activation energy is about 1020 kJ mol1
(Katakis and Gordon, 1987). According to the authors, the high values
of the apparent activation energies, 155 kJ mol1, estimated from kinetic data of precipitation, suggest a surface-controlled mechanism [see
Table 1, case (d)]. Calculating Eact, H, S, and G from the data
of the above work we nd that both H and S are positive. G
has the value 84.8 kJ mol1 at 308 K. The positive value of S suggests
a dissociative mechanism. This can be explained by release of water
molecules from the ionic [Ca2+(aq), CO32(aq)]* environment in order
to form the precipitate [* the symbol (aq) means aquated, i.e. hydrated].
The high value of H suggests that a large amount of energy is needed
at the activation step (reactants to transition state step) in order to
break the bonds that exist between the ions and the water molecules
of the coordination sphere of the ions. The breaking of bonds requires
energy while the formation of bonds releases energy. Thus dehydration
is an endothermic main step in the reactants-transition state process
and governs the H and S values for the precipitation. The combination of the ions to form the electrostatic bond is an exothermic process and results in negative H0 and S0. Under the conditions of
precipitation described in case (d), the governing activation step is the
dehydration of the ions as indicated by the positive values of H and
S. Hydration of the ions should result in H and S values.
In case (e) the heterogeneous precipitation of aragonite from dilute
solutions is reported (Romanek et al., 2011). In the above study, the rate
was calculated directly by the mass of solid precipitated per unit time,
taking into consideration the change in surface area over the course of
the run. In the same work in Table 4, there are data for rate constants
from published data in articial and natural sea water studies, which
are also included in Table 1 [case (e*)]. The association of the ions in
case (e) may be governing the activated step since S has an overall
negative value (42.5 J K1 mol1). The G is 82.8 kJ mol1 (as expected). In case (e*) some exothermic equilibria preceding the ratedetermining step must be taking place (studies in articial and natural
sea water) lowering the value of H [from 69.7 in case e to
11.09 kJ mol 1 in case (e*)]. The value of S is more negative
[42.5 in case (e) vs. 246.9 J K1 mol1 in case (e*)] suggesting a
more organized structure in the transition state. The presence of foreign

147

ions is most probably causing these effects causing also retardation of


the reactions [G 82.8 kJ mol 1 in case (e) and 87.13 kJ mol1 in
case (e*)]. The higher value of G in case (e*) implies slower process,
the ionic strength playing its retarding role on the rate constants and
thus the G.
For the precipitation reaction of vaterite in aqueous solutions, an
apparent activation energy of 46 kJ mol1 was obtained [case (f)]
(Xyla et al., 1991). According to the authors, the precipitation, took
place via a polynuclear mechanism. The presence of EHDP (ethane-1hydroxyethylideno-1,2-di-phosphonic acid) had a strong retardation effect both on the induction times and on the subsequent rates of precipitation (Koutsoukos and Kontoyannis, 1984). Retardation implies higher
free energy of activation and this explains the higher G value
(120 kJ mol1) compared to approximately 80 kJ mol1 expected for
the precipitation of CaCO3 or 100 kJ mol1 expected for the precipitation
that includes also diffusion of species (~20 kJ mol1, activation energy
for the diffusion). The activation energy of 46 kJ mol1 [case (f)] compared to 155 kJ mol1 [case (d)] does not indicate the retardation caused
by EHDP. On the contrary, it implies acceleration of the reaction and it is
thus once again shown that the use of G instead of Eact is more realistic (Petrou, 2012) and corroborates the large retardation effect
(120.2 kJ mol1 vs. 84.8 kJ mol1). Data for the calculation of the
thermodynamic parameters, Eact, H, S, G, were taken from
Fig. 7 of the above work [case (f)].
4.1.2. Crystallization of CaCO3 in aqueous solutions
Crystallization can be categorized into the processes nucleation and
growth. Nucleation refers to the process of critical nucleus formation
and growth is the augmentation of those critical nuclei. Growth may
be a dominant process if the saturation level is not enough to overcome
the energy barrier for nucleation. When there are small concentrations
of crystals present, nucleation and growth both take place simultaneously (spontaneous or homogenous crystallization). The spontaneous crystallization process does not allow a reliable kinetic analysis to
be made since the implied assumption that homogeneous nucleation
takes place is doubtful (Nancollas and Purdie, 1968): nucleation is likely
to occur on impurity particles which offer available sites for crystal
growth. During spontaneous crystallization both the total number and
size distribution of the particles vary. All the above explain the variation
of the G values that we have calculated for the crystallization
processes for both CaCO3 and CaSO4.
Studying the kinetics of crystal growth of vaterite (CaCO3) in
aqueous solution [case (g)], Kralj et al. (1990) reported a value of
57.1 kJ mol1 for the activation energy. This, according to the authors,
means that the rate-determining mechanism is the release of water
molecules from the calcium ions as they jump into lattice positions.
Our calculations [Table 1, case (g)] based on the data of this work
suggest that a composite reaction is taking place since the absolute
value of S is small. This composite reaction may consist of: a) the

Notes to Table 1:
(a) Experimental data in units of T in K and of k in: s1 (cases d, f, k), nm s1 (case g), M1 s1/mg seed l1 (case h), M1 s1/mg seed/100 ml (case i), s1 cm2 (case l), mol m2 s1
(case m), cm s1 (case n), mol m2 s1 (case e), l2 s1 mol1 cm2 (case j, kRS), l2 s1 mol1 mg1(case j, kRM), mol cm2 s1 (case o).
(b) Calculated values.
(c) This work.
(d) Koutsoukos and Kontoyannis, 1984.
(e) Romanek et al., 2011.
(f) Xyla et al., 1991.
(g) Kralj et al., 1990.
(h) Wiechers et al., 1975.
(i) Nancollas and Reddy, 1971.
(j) Hasson et al., 2010.
(k) Rodriguez-Blanco et al., 2011.
(l) Sjoberg, 1976.
(m) Alkattan et al., 1998.
(n) Sjoberg and Rickard, 1984.
(o) Schott et al., 1989.

148

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

dehydration of the ions, which results in the increase of the degrees of


freedom and follows a dissociative mechanism with positive values of
S and H and b) the accommodation of the ions into lattice positions which means loss of degrees of freedom and follows an associative
mechanism with negative values of S and H. The algebraic sum of

the S values (for explanation see below) gives a small overall negative result. The H value being also the algebraic sum of the H
values of the two processes also has a small value. This value is obviously smaller than the value due to dehydration only [see case (d)]. This
means that, in this case, the association of the ions in the lattice happens

Table 2
Kinetic data (k, T) and calculated thermodynamic parameters (Eact, H, S, G) for precipitation, crystallization and dissolution of calcium sulfate (CaSO4).
lnk(a)

1
E(b)
act kJ mol

H(c) kJ mol1

1. Precipitation in aqueous solutions


Case (d)
293
303
318
333
Case (e)
303
313
323
333
Case (f)
298
308
313
Case (g)
352.6
311.43
298
Case (h)
363
343
323
Case (i)
322.58
303.03
294.12

6.624
4.634
2.534
0.394
4.094
3.401
2.639
2.303
12.3
11.2
10.66
11.354
11.864
12.284
2.6
2.9
3.1
5.113
5.310
5.436

124.3(c)
184 2(d)

121.7

2. Crystallization in aqueous solutions


Case (j)
347.8
346.6
344.8
Case (k)
288
298
303
308
318
Case (l)
298
323
343
363
Case (m)
318.5
307.7
288.2
Case (n)
328
318
308
298
Case (o)
303
323
343
363
1
303
Case (p)
Without polyacrylates
313
323
2
303
Case (p)
313
With polyacrylates
323
Case (q)
303.03
322.58
344.82
370.37
Case (r)
333
343
353
363

7.497
7.55
7.619
4.656
3.663
3.211
2.885
2.083
2.425
1.099
1.204
2.303
3.75
4.1
4.8
2.554
3.321
4.181
5.053
13.332
11.190
10.131
9.302
2.557
1.799
1.012
2.601
1.752
0.933
19.625
19.113
18.644
18.015
2.55
2.23
1.92
1.52

40.33(c)
65.2(j)

T (K)

III Dissolution in aqueous solutions


Case (s)

283
288
293
298
303

4.358
4.160
3.808
3.381
3.182

S(c) J K1 mol1
116.7

G(c) kJ mol1
85.8 at 308 K

51.62(c)
5171(e)

49

84.6(c)
84.30(f)

81.9

72.2

14.2(c)
(g)

11.5

307.8

106 at 308 K

12.12(c)
30(h)

9.3

242.63

84.03 at 308 K

8.84(c)
3050(i)

6.274

268.45

88.96

37.4

200.7

99 at 308 K

64.75(c)
62.8 2(k)

62.2

66.7

82.8 at 308 K

68.1(c)
52.6(l)

65.41

47.5

80.04 at 308 K

26.4(c)
26.930.7(m)
Average = 28.9
39(c)
66.9 6.3(n)

23.9

201.36

85.92
at 308 K

36.5

157.9

85.12
at 308 K

60.68(c)
62.8(o)

57.93

162.25

107.9
at 308 K

62.79(c)

60.2

67.63

81.03
at 308 K

67.832(c)
64.9(p2)

65.23

51.28

81.03
at 308 K

22.015(c)
13.7623.14(q)

20.39

341.45

125.56
at 308 K

33.9(c)
34 2(r)

31.05

173.83

84.59

44.6(c)
41.8 6.3(s)

42.2

132

82.9 at 308 K

117

(p1)

85.02 at 308 K

104.1 at 308 K

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

during the activation step, whereas in the case of precipitation [case (d)]
only dehydration occurs during the activation step. The G is approximately 75 kJ mol1 at 308 K.
Thus, dehydration, being an endothermic process, and crystal formation i.e. formation of (electrostatic) bonds, being an exothermic process,
both take place during the reactants-transition state step process
resulting in values of H and S smaller than those found for the
case where only dehydration occurs during the reactants-transition
state step process [precipitation, case (d)].
Reddy and Nancollas (1971), reporting on the kinetics of crystallization of calcium carbonate, state that the activation energy is relatively
high, 46.0 kJ mol1, and that the crystallization rate is independent of
the stirring speed. Wiechers et al. (1975, case (h)) suggest that the crystallization process conforms to the model for Ca2+ removal proposed by
Reddy and Nancollas (1971). The removal rate constant, k, is temperature dependent according to the Arrhenius relationship with Eact =
43.1 3.8 kJ mol 1. The high activation energy and other kinetic
parameters support the hypothesis that the crystallization process is
surface controlled.
Nancollas and Reddy (1971), in case (i) point out that the activation
energy for crystal growth of calcite is 46.0 4.2 kJ mol1,
thus supporting the proposed surface-controlled mechanism for
crystal growth. This value is considerably larger than the 17.6 kJ mol1
to be expected for a diffusion controlled reaction (Howard et al.,
1960), again pointing to the proposed interfacially controlled mechanism for calcium carbonate crystal growth. The assumptions made
of homogeneous nucleation and of simultaneous nucleation and
growth are, however, open to considerable criticism (Nancollas et al.,
1964). Data from Table 1 of the above work [case (i)] led us to
calculate H1 = 44.2 kJ mol 1, H2 = 26.5 kJ mol 1, S1 =
120.3 J K 1 mol 1, S2 = 181.92 J K 1 mol 1, G1 =
81.3 kJ mol1 at 308 K, G2 = 82.5 kJ mol1 at 308 K. The subscripts
1 and 2 refer to data that differ only according to mg seed crystal/
100 ml, the stirring rate (rpm) being the same.
A novel electrochemical system developed in case (j) (Hasson et al.,
2010) enables drastic reduction in the electrode area requirement by
directing the precipitation of calcium carbonate to occur in a seeds crystallization vessel rather than on the cathode. Kinetic coefcients reported in the literature are of two types: fundamental coefcients kRS based
on the actual crystallization area and coefcients kRM based on seeds
concentration. Literature values of the above kinetic coefcients kRS
and kRM are taken from Table 1 of case (j). The values of the two different kinetic coefcients result, as expected, in similar values of G
(85.1 vs. 88.86 kJ mol1 from kRS and kRM, respectively) but the values
of H and S are completely different implying completely different

149

processes and thus mechanisms. The H and S values resulting


from kRS are positive, but those resulting from kRM are negative.
Certainly the values from both coefcients suggest composite processes.
The rst include endothermic processes whereas the second include
exothermic ones which govern the total procedures. The very negative
value of S from kRM suggests a great degree of organization in all
the processes including the transition state step.
In case (k) a study of amorphous calcium carbonate ACC crystallization to calcite via vaterite is presented (Rodriguez-Blanco et al., 2011).
According to the authors, the crystallization process occurs in two
stages: rstly the particles of ACC rapidly dehydrate and crystallize to
form individual particles of vaterite; secondly the vaterite transforms
to calcite via a dissolution and reprecipitation mechanism with the
reaction rate controlled by the surface area of calcite. The Ecryst value
for calcite growth is found by the authors to be 66 2 kJ mol1. Several
other authors have derived activation energies for seeded calcite
growth ranging from ~39 to ~70 kJ mol1. The fact that our calculated
value of G is 58.4 kJ mol 1 for calcite crystallization, i.e., smaller
than the value found (by us) in all other cases (~80 kJ mol1), is easily
explained by the fact that in this crystallization process only the ions
are present in the solution. All other ions that are presCa2+ and CO2
3
ent in all the other cases are absent in this case. In this case the vaterite
transforms to calcite via dissolution and reprecipitation. It has been
found that retardation is caused due to the presence of foreign compounds. In this case there are no foreign compounds. So, the crystallization process should be faster than in all other cases. This implies a
smaller free energy of activation and this is actually happening. The
value of 58.44 kJ mol 1 is the smallest value which can be found for
calcite crystallization. In the other cases of Table 1, Na2CO3 and CaCl2
are the sources for CO23 and Ca2 +, respectively, whereas in case k,
CaCO3 was the source of both ions. The rest of the ions that are present
contribute in the ionic strength of the solutions which affects the values
of the rate constants and hence the Eact, H, S and G values making the rate slower and hence the G bigger. In this case S is positive (40.78 J K1 mol1) in agreement with the dehydration process
that the authors suggest based on other kind of scientic ndings. The
rate determining step is governed by the dehydration step resulting in
positive values of S. The contribution of dehydration only, results in
larger values of S [Table 1, case (d) and Table 2 case (d)]. The kinetic
data that are derived from induction times (calcite nucleation) result in
negative S and a value for G of 89.7 kJ mol1. One could argue
that the dissolution process in general does not involve foreign compounds, so should the G values be smaller? The smaller value is
~60 kJ mol1 (case (k), calcite crystallization). The free energy for dissolution is found to be ~80 kJ mol1 (Tables 1,2). The diffusion activation

Notes to Table 2:
(a) Experimental data in units of T in K and of k in: s1 (cases d, e, f, h, i,k, l, m, s), M1 s1 (cases g, k, p), m/s wt%2 (case j), l mm mol2 s1 (case n), mol/cm2 s (case o), mol m2 s1
(case q), l mol1 s1 (case r).
Case e*: for the supersaturation range between 2.5 and 11.
Case l*: activation energy for the nucleation of calcium sulfate dihydrate in 3 m NaCl solutions. The activation energy for case l was obtained using the induction time, tind.
(b) Calculated values.
(c) This work.
(d) Klepetsanis and Koutsoukos, 1991.
(e) Alimi et al., 2003.
(f) Merdhah and Yassin, 2008.
(g) Klepetsanis et al., 1999.
(h) Lancia et al., 1999.
(i) Prisciandaro et al., 2013.
(j) Jamialahmadi and Muller-Steinhagen, 2000.
(k) Liu and Nancollas, 1970.
(l) He et al., 1994.
(m) Prisciandaro et al., 2003.
(n) Liu and Nancollas, 1973.
(o) Smith and Sweett, 1971.
(p)1 and (p)2 Amjad and Masler, 1985.
(q) Yehia et al., 2011.
(r) Cetin et al., 2001.
(s) Liu and Nancollas, 1971.

150

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

2
energy is needed for the separation of the ions (Ca2+, CO2
3 , SO4 ) and
this has a value of ~ 20 kJ mol 1 increasing the 60 kJ mol1 value to
~80 kJ mol1 (as found). So the value of ~60 kJ mol1 is the free energy
of activation in the absence of foreign ions.

4.1.3. Dissolution of CaCO3 in aqueous solutions


Sjoberg (1976) describes a fundamental equation for calcite dissolution kinetics. He reports an apparent activation energy for the reaction
between 5 and 50 C of 35.1 kJ mol1 [case (l)]. The relatively high activation energy indicates, according to the author, that the reaction is
mainly chemically- rather than diffusion-controlled. Nancollas and
Reddy (1971) determined an activation energy of 46 4.2 kJ mol1
for the precipitation reaction and concluded that the rate-determining
step is a surface-controlled rather than transport-controlled precipitation. For a diffusion-controlled reaction, according to Burkin (1966),
the expected activation energy is about 16.7 kJ mol1.
From the data of the above work [case (l)] we calculate Eact =
34.67 kJ mol1, H = 32.19 kJ mol1, S = 233.2 J K1 mol1,
G = 104.0 kJ mol1 at 308 K.
Alkattan et al. (1998) reported an experimental study of calcite and
limestone dissolution rates as a function of pH from 1 to 3 (aqueous
HCl solutions) and temperature from 25 to 80 C [case (m)]. Dissolution
rates were determined for single calcite crystals, limestone sand and
compressed calcite powders. The corresponding rates of the three
different sample types were identical. Apparent rate constants and H+
diffusion coefcients were reported. According to the authors, the
apparent activation energy of 19 kJ mol1, is somewhat higher than
the 13.2 kJ mol1 obtained by Oelkers and Helgeson (1988), for the
apparent activation energy of HCl diffusion. Our calculations give a
dissolution value for G 81.4 kJ mol1 [Table 1, case (m)].
Sjoberg and Rickard (1984) found an activation energy 54
4 kJ mol1 for Carrara marble and 46 4 kJ mol1 for Iceland spar, determined at pH 8.4 in 0.7 M KCl solutions [cases (n)]. Our calculations
give G values 83.9 kJ mol1 for Carrara marble and 87.3 kJ mol1
for Iceland spar.
From the data of the work of Sjoberg (1976) and Alkattan et al.
(1998) we calculate negative values for S that suggest that the associative mechanisms are predominant. The large absolute value generally
suggests a composite reaction or process (Petrou, 2012).
The enthalpy of activation for the total (composite) transformation
is equal to the algebraic sum of the various enthalpies that accompany
all the steps involved in the composite process. This algebraic sum
may obviously be positive, zero or negative depending on the absolute
values of the various enthalpies that are involved. The same applies
for the entropy of activation.
Hydration-aquation (formation of bonds) is an exothermic process
(H0 is negative) and this is the reason that the H for this decomposition (dissolution) is so small. Hydration-aquation follows an associative mechanism (loss of degrees of freedom, increase of organization)
and thus the value of S is negative. The total S is the sum of a
negative value for the aquation of the cations, a negative value for
the aquation of the anions and a positive value for the dissociation of
the salt (dissociative mechanism), i.e. the separation of the calcium
and the carbonate ions (increase of degrees of freedom, loss of
organization).
In case (o) dissolution kinetics data of strained calcite are presented
(Schott et al., 1989). It is reported that interface-limited dissolution of
minerals occurs non-uniformly with preferential attack at sites of
excess surface energy such as dislocation, edges, point defects, and
microfractures. Strained crystals are predicted to show higher dissolution
rates due to the increased internal energy associated with dislocations
and due to enhanced nucleation of dissolution pits at dislocation outcrops
on the surface. Using calcite strained to different degrees, the authors
have observed a measurable rate enhancement relative to unstrained
crystals at temperatures from 3 to 80 C. The authors used large, single
crystals of calcite for which surface area and rate should be directly

related. From Fig. 4 of case (o) we take data for comparison of the rates
of dissolution of normal (non-strained) calcite and calcite strained to a
dislocation density of 0.5 1091.0 109 cm2 as a function of pH at 3,
25 and 80 C. The data were taken for pH = 5.5. Our calculations prove
that the G values are similar (130.8 vs. 131.6 kJ mol1). The values
of G suggest retardation in comparison with the other cases of
Table 1 and this may be attributed to the increased ionic strength due to
KCl and HCl present in the solution. The very negative S suggests a
great degree of organization in the transition state compared to the reactants. The values of H are very low suggesting composite reactions.
Similar effect of the ionic strength, that is retardation, is observed in
cases: Table 1 case (f) and Table 2 case (q).
4.2. Processes involving CaSO4
4.2.1. Precipitation of CaSO4 in aqueous solutions
An understanding of the mechanism of CaSO4 formation in aqueous
solutions where precipitation takes place spontaneously due to high
supersaturation is of great importance for the control of undesirable
scale deposits. The time period that elapses between the attainment of
supersaturation and the appearance of the rst crystals is called the
induction time (He et al., 1994). An empirical relationship similar to
the Arrhenius equation (temperature dependence of the rate constant)
was suggested (Liu and Nancollas, 1975) for the effect of temperature
on induction period.
For the spontaneous precipitation of calcium sulfate dihydrate at
conditions of sustained supersaturation, an apparent activation energy
of 184 2 kJ mol1 [Table 2, case (d)] has been reported (Klepetsanis
and Koutsoukos, 1991). The kinetic analysis according to the Arrhenius
equation yielded a high activation energy value indicative, according to
the authors, of surface-controlled mechanisms. According also to the
authors, calcium sulfate dihydrate is formed in aqueous supersaturated
solutions by a polynuclear mechanism. Using the kinetic data of the
above work, we have calculated Eact, H, S and G [Table 2,
case (d)]. The H and the S values are positive suggesting a dissociative mechanism, i.e. going from the reactants to the transition state
the number of the various species increases. This may mean that
water molecules are released from the ionic environments of Ca2+(aq),
and SO42 (aq) in order to form the precipitate. G has the value
85.8 kJ mol1 at 308 K [Table 2 case (d)]. Due to the supersaturation
conditions the diffusion activation energy of ~20 kJ mol1 is obviously
not needed. This value of G is comparable with the value of
84.8 kJ mol 1 at 308 K for calcium carbonate precipitation [Table 1,
case (d)]. The high value of H suggests that a large amount of energy
is needed in order to break the bonds that exist between the ions and
the water molecules in the coordination sphere of the ions. Thus
dehydration of the ions is a main step in the reactants-transition
state process and is the main component of the H and the S
values. Hydration of the ions should result in H and S values.
Alimi et al. (2003) studying the kinetics of the precipitation of calcium sulfate dihydrate in a desalination unit, reported an activation energy value of 51.047 kJ mol1 [Table 2, case (e)]. Using data from the
above work we calculate Eact = 51.62 kJ mol1, H = 49 kJ mol1,
S = 117 J K1 mol1, G = 85.02 kJ mol1 at 308 K. The negative value of S suggests an associative mechanism, i.e. going from the
reactants to the transition state the number of the various species
decreases. This may suggest that the water molecules are released
from the ionic environments [Ca 2 +(aq) , and SO 4 2 (aq)] before the
slow rate-determining step. The combination of the positive and
negative ions leading to negative S i.e. to decrease of the number
of the various species [decrease of the degrees of freedom], is the
rate-determining step.
In a laboratory study and a prediction of calcium sulfate at highsalinity formation water (Merdhah et. al., 2008) the activation energy is
found to be 84.30 kJ mol1. Calculating H, S and G from the
data of this work [Table 2, case (f)], we nd that S is negative, as

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

previously, suggesting an associative mechanism. The negative value of


S suggests that going from the reactants to the transition state the
number of particles overall decreases and that the combination of the
positive and negative ions [Ca2+, SO2
4 ] may be the rate-determining
step. The release of water molecules from the aquated ions [Ca2+(aq)
and SO42(aq)] (resulting in an increase in S) may have taken place
in a faster preceding step. The high-salinity formation water applied in
this experiment may have played a role in attracting the water of solvation and thus accelerating the release of water molecules in a step prior
to the rate-determining step of the precipitation process. G has the
value 104.14 kJ mol1 at 308 K [Table 2, case (f)] which suggests that
the diffusion activation energy (~20 kJ mol1) is needed (~85 +
~20 kJ mol1). According to the authors, a temperature rise from 40 to
90 C causes a decrease of CaSO4 solubility, that is, the solubility of
CaSO4 is overall an exothermic process. This implies that the heat of
aquation of the ions Ca2+ and SO2
4 , which is a process that releases
heat, is greater than the heat needed for separating the ions Ca2+ and
SO2
4 , i.e. the breaking of the lattice of the salt (CaSO4) which is an endothermic process.
In a study of the role of temperature in the spontaneous precipitation of calcium sulfate dihydrate in supersaturated solutions
(Klepetsanis et al., 1999), the authors maintain that surface diffusion
is the mechanism according to which the growth units are integrated
into the active sites of the supercritical nuclei. According to our calculations [Table 2, case (g)], diffusion is actually taking place as revealed by
the value of G, 106 kJ mol1 (~80 + ~20 kJ mol1 due to diffusion).
The H, S and G values were estimated by the application of the
Eyring equation on the data of the above work (Table 2). H has a very
small absolute value, 11.5 kJ mol1, due to a composite, possibly exothermic, process with negative H0, that precedes the activated step.
S is negative with very large absolute value (307.8 J K1 mol1)
implying again a composite process. The negative S value suggests
an associative mechanism which would mean that considerable organization occurs at the transition state. G is 106.0 kJ mol1 at 308 K. The
hydration may be the exothermic step that results in a negative component in the enthalpic term, thus decreasing its absolute value, and also
in a negative component in the entropic term, thus increasing its absolute value. The organization may be due to the association of the ions
(calcium and sulfate) along with water molecules (dihydrate salt).
Thus the composite process may be: a) the association of the ions,
Ca2+ and SO2
4 , and b) the association of water molecules to form the
dihydrated salt (hydration).
The dehydration process (release of water molecules from the coordination sphere of the ions) clearly does not occur at the transition step
since it would lead to a positive contribution to S, increasing thus its
very negative value to a less negative one, and also to a positive
contribution to H, increasing its very low positive value.
The G value of case (f) (104.1 kJ mol1) is comparable with the

G value of case (g), (10 kJ mol1), though their S values differ


by a factor of about 4.3 ( 72 versus 308 J K1 mol1) and their
H values differ by a factor of about 7.12 (81.9 versus 11.5 kJ mol1).
The value for G is the same in both cases, as expected (Petrou,
2012), since the same reaction is taking place.
In case (h), the dependence of the induction period (tind) on temperature made it possible to evaluate Eact as 30 kJ mol1 (Lancia et al.,
1999). Our calculations, based on the data of this work [case (h)], gave
a very small value for H (9.3 kJ mol 1) and a very negative value
for S ( 242.6 J K 1 mol 1). The G value is 84.03 kJ mol1 at
308 K (Table 2).
Prisciandaro et al. (2013) studied gypsum precipitation in industrial
equipment and made a comparison of different additives. We found
G values 88.96 kJ mol1 for gypsum precipitation without additives
[Table 2, case (i)]. Again the H value was very small (6.3 kJ mol1)
and the S value very negative (268.45 J K1 mol1).
The three cases (g), (h), (i) bear very small H values, very
negative S values and the G values are 106 kJ mol 1 for case

151

(g), 84 kJ mol1 for case (h) and 89 kJ mol1 for case (i). This indicates
that, for the three cases (g), (h) and (i), similar mechanisms must be
taking place with the only difference being that for case (g) the diffusion
of the ions is needed which adds an amount of ~ 20 kJ mol1 to the
activation energy.
4.2.2. Crystallization of CaSO4 in aqueous solutions
Jamialahmadi and Muller-Steinhagen (2000, case (j)), reporting
on the crystallization of calcium sulfate dihydrate from phosphoric
acid, state that the effect of temperature on the growth rate constant
follows an Arrhenius relationship with an activation energy of
65.2 kJ mol1. From the Arrhenius plot of growth rate constant we get
data from which we calculate H = 37.4 kJ mol 1, S =
200.72 J K1 mol1 and G = 99 kJ mol1 at 308 K [Table 2, case
(j)]. The negative S value suggests an associative mechanism and
its large absolute value suggests a composite process (see previous
cases).
Liu and Nancollas (1970), studied the kinetics of crystal growth of
calcium sulfate dihydrate [case (k)] and reported an activation energy
value for the surface controlled crystal growth of 62.8 2 kJ mol1.
The authors declare that this value is appreciably larger than the
value 18.8 kJ mol 1 (Howard et al., 1960) to be expected on the
basis of pure mass transport control. Using data from the above work
we calculate Eact = 64.75 kJ mol 1, H = 62.24 kJ mol 1, S =
66.7 J K1 mol1, G = 82.78 kJ mol1 at 308 K.
The nucleation kinetics of calcium sulfate dihydrate in NaCl solutions
up to 6 m and 90 C was studied by He et al. (1994, case (l)). Using data
from the above work at 25, 50, 70, 90 C, NaCl 3.0 m, CaSO4 (aq)
180.0 mm and the corresponding induction times (in s), we calculated
G = 80.04 kJ mol1. Rate constants were calculated from induction
times.
Prisciandaro et al. (2003, case (m)) propose that citric acid has
a strong retarding effect towards gypsum nucleation. Several values
for the interfacial tension and activation energy have been estimated
as a function of citric acid concentration and of temperature. The
average of the activation energy values reported by the authors is
28.9 kJ mol 1. The values with the above average are comparable
with respect to that previously found in the absence of any additives
(Eact = 30 kJ mol 1 , Lancia et al., 1999). Thus we could say that
there is no actual retarding effect. On the other hand, our calculations reveal that G is 85.92 kJ mol 1 , H = 23.9 kJ mol 1 ,
S = 201.36 J K 1 mol 1 , implying an associative mechanism
and a composite reaction. The G value does not suggest strong
retardation.
Studying the linear crystallization and the induction-period of the
growth of calcium sulfate dihydrate crystals, Liu and Nancollas (1973)
reported that the linear crystallization of well-formed crystals of calcium sulfate dihydrate from supersaturated solutions needs the activation energy of 66.9 6.3 kJ mol 1 for the fast-growing faces. Using
data from this work we nd G = 85.12 kJ mol1 at 308 K.
Smith and Sweett (1971, case (o)), studying the crystallization of
calcium sulfate dihydrate (CaSO42H2O) reported that the value of the
rate constant indicated that the rate-controlling step was dehydration
of calcium ions. This suggests a positive value for S and a dissociative
mechanism. However, the value of S is found to be negative [Table 2,
case (o)] suggesting an associative process. The G value is 107.9 at
308 K implying that diffusion of the ions is taking place in the ratedetermining step.
Amjad and Masler (1985, case (p)) suggest that their results indicate
that the crystallization of CaSO42H2O in the presence of polyacrylates is
preceded by an induction period. Crystal growth of CaSO42H2O
proceeds with a rate close to that in pure supersaturated solution. The
authors report an activation energy of 64.9 kJ mol1 for the polymer
containing solution which is in excellent agreement, according to the
authors, with the Eact value reported for the CaSO42H2O seeded growth
in the presence of phosphonates (Liu et al., 1975). Our results show that

152

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

polyacrylates do not affect the activation parameters, the G being


81.03 kJ mol1 in both cases, i.e. without and with polyacrylates. The
values of S and H differ slightly (Table 2).
Yehia et al. (2011, case (q)) studied the effects of some parameters affecting the crystallization rate of calcium sulfate dihydrate in sodium
chloride solution at different supersaturation, pH = 3, ionic strength
I = 0.5 M and at 2080 C, and they reported that the rate of crystallization was found to depend on the stirring rate, suggesting a diffusion
mechanism. In addition, the presence of Mg2+ or Al3+ retarded the rate
of crystallization to an extent proportional to the amounts present. Furthermore, the retardation effect was enhanced as the supersaturation decreased. The presence of sodium chloride decreases the supersaturation
and enhances the retardation. This explains the value of G that we
found, namely 125.6 kJ mol1 (in the absence of additives). Thus, the
presence of NaCl caused an increase in G, i.e. a decrease of k (retardaattract each
tion). This agrees with the fact that the ions Ca2+ and SO2
4
other with Coulombic forces and the presence of Na+, Cl among them
decreases the attractive forces and consequently decreases the rate constant, thus increasing G. The effect of the presence of the ions Na+,
Cl, that is the effect of the ionic strength on the rate constant, is
expressed by various equations (Katakis and Gordon, 1987). However,
the reported values of Eact, 13.7623.14 kJ mol1, do not support retardation. On the contrary, they suggest acceleration of the reaction (small Eact
values suggest large values for rate constants) [for the values of Eact of the
CaSO4 crystallization see Table 2, column Eact].
In case (r) the kinetics of gypsum formation and growth during the
dissolution of colemanite in sulfuric acid is presented (Cetin et al.,
2001). After the fast dissolution reaction of colemanite in aqueous sulfuric acid, the nucleation of the gypsum crystals rst occurs from the supersaturated solution and then the crystals grow on these nuclei. The
authors report an activation energy of 34 2 kJ mol1 for the crystal
growth of gypsum. Our calculated value of G is 84.59 kJ mol1 as expected. The S is negative as expected and the H is 31.05 kJ mol1.
4.2.3. Dissolution of CaSO4 in aqueous solutions
The kinetics of dissolution of calcium sulfate dihydrate crystals in
subsaturated solutions of the salt was studied at various temperatures
and there was a suggestion of lm diffusion as the rate-controlling
step (Liu and Nancollas, 1971, case (s)). The activation energy was
found to be 41.84 6.3 kJ mol1. This value, according to the authors,
is larger than the value 18.8 kJ mol 1 to be expected on the basis of
pure mass transport control. Although much of the evidence points to
lm diffusion as being the rate-controlling step, the authors support
that it is not possible to exclude the possibility of a more complicated
mechanism due to the relatively large activation energy. Our calculations based on data of the above work give: Eact = 44.6 kJ mol 1,
H = 42.2 kJ mol 1, S = 132 J K 1 mol 1, G =
82.9 kJ mol 1 at 308 K. The negative value of S suggests that the
transition state is more organized than the reactants.
5. Conclusions
The calculation of the thermodynamic parameters Eact, H, S
and G of the precipitation, crystallization and dissolution processes
for the salts CaCO3 and CaSO4, has demonstrated once again that, for
geochemical processes, the critical factor is G and not Eact (Petrou,
2012). The values of the above thermodynamic parameters lead to the
following conclusions:
The G values are almost the same for CaCO3 and CaSO4 precipitation, crystallization and dissolution at the same temperature. This suggests that the electrostatic forces between the ions Ca2+ and CO2
3
(CaCO3) and Ca2 + and SO24 (CaSO4) are important and that they
govern the above processes.
The precipitation of the two salts may take place both by a dissociative
mechanism, S N 0 [cases (d) of Tables 1 and 2, case (k) for calcite

crystallization, case (j) for kRS coefcient] and by an associative mechanism, S b 0 [all the other cases, Tables 1 and 2]. The G values
are approximately the same whereas the values of H and S are
different.
The fact that the G values are almost the same for cases (d) of
Tables 1 and 2, suggests that the rate-determining step is the same
and this must be the dehydration of CaCO3 and CaSO4. The positive
values of S suggest dissociative mechanisms and actually this is the
mechanism when water molecules are released from the coordination
spheres of the cations [Ca2 +aq] and anions [CO32 (aq) and SO42 (aq)].
This dehydration process involves breaking of bonds and demands energy that is revealed as large values for H. In all the other cases,
S is negative implying an associative mechanism, i.e. increased
order in the transition state.
When a process takes place without the need for diffusion of the
ions, the free energy of activation is about 85 kJ mol 1 whereas
when there is a need for diffusion of the ions an additional amount
of ~ 20 kJ mol 1 is required. This amount is the activation energy
for the diffusion. Thus, in total, ~ 100 kJ mol 1 is needed for these
processes.
In the case of CaCO3, a very wide range is found for the values for H
(66.00162.00 kJ mol1) and S (501.00 to +248.00 J K1
mol1) while the G values cover a narrow range (7590 kJmol1).
An additional amount of +20 kJmol1 is needed when diffusion of the
ions takes place. Values of 120, 131 and 132 kJ mol1 are observed for
cases [case (f), cases (o), Table 1] where retardation is caused due to
the presence of foreign compounds.
Similarly, in the case of CaSO4, a very wide range is found for the
values for H (6.00122.00 kJ mol1) and S (342.00 to
+117.00 J K1 mol1), while a narrow range is found for the G
values (80.0089.00 kJ mol1). There is an additional amount of
+20 kJ mol1 for processes that include diffusion of the ions. A value
of 126 kJ mol1 is observed for one case [case (q), Table 2] where retardation is caused due to the presence of foreign compounds.
The Eact values vary between 63 and 164 kJ mol1 for CaCO3 and between 8 and 184 kJ mol1 for CaSO4, demonstrating once more that
the G value is a more realistic parameter, being almost the same for
similar processes.
The various small differences that appear in the values of G are
due to differences in ionic strength (the concentration and charge
of the foreign ions) that affect the rate constants and thus the activation parameters. The pH also has an effect, as does also the nature of the solvent.
The large absolute values of H and S reveal composite reactions. A composite reaction is, for example, one for which dehydration takes place (positive values of H and S ) followed by
association of the ions (negative values of H and S). The algebraic sum of the values is the total value of H and S.
Acceleration/retardation. The activation energy value of 46 kJ mol1 for
case (f) compared to the value of 155 kJ mol1 for case (d) does not
indicate retardation caused by EHDP, but, on the contrary, implies
acceleration of the reaction. However, the use of G instead of Eact
has been shown to be more realistic (Petrou, 2012), indicating here
the large retardation effect. G for case (f) is 120.2 kJ mol1 and for
case (d) is 84.8 kJ mol1. The larger value of G suggests a slower reaction. Also, in case (q), the values of Eact reported by the authors do not
justify retardation, but, on the contrary, suggest acceleration of the
reaction since small Eact values suggest large values of rate constants.
Again the use of G instead of Eact explains the retardation [20 vs.
126 kJ mol1, case (q)].
The addition of salt, for example NaCl, causes retardation of the crystallization process [case (q)]. The S value, ( 341.45 J K1 mol1)
suggests a very organized transition state. This is due to the attraction
and Ca2+, respectively.
of Na+ and Cl to SO2
4

A.L. Petrou, A. Terzidaki / Chemical Geology 381 (2014) 144153

Acknowledgments
Many thanks are due to all the authors that are included in the list of
references, who, with their valuable data, the high-level scientic work
both in experiments and in explanations, have given us signicant assistance in providing additional evidence concerning the activated steps
and the mechanisms for precipitation, crystallization and dissolution
of CaCO3 and CaSO4. We thank Pauline Zannia for her assistance. We
also thank the University of Athens for the nancial support of this
work.
References
Alimi, F., Ell, H., Gadri, A., 2003. Kinetics of the precipitation of calcium sulfate dihydrate
in a desalination unit. Desalination 157, 916.
Alkattan, M., Oelkers, E.H., Dandurand, J.-L., Schott, J., 1998. An experimental study of
calcite and limestone dissolution rates as a function of pH from 1 to 3 and temperature from 25 to 80 C. Chem. Geol. 151, 199214.
Amjad, Z., Masler III, W.F., 1985. The inhibition of calcium sulfate dihydrate crystal
growth by polyacrylates and the inuence of molecular weight. Corrosion 85
(paper number 357).
Burkin, A.R., 1966. The Chemistry of Hydrometallurgical Processes. E. & F. N. Spon, (158
pp.).
Cetin, E., Eroglou, I., Ozkar, S., 2001. Kinetics of gypsum formation and growth during the
dissolution of colemanite in sulfuric acid. J. Cryst. Growth 231, 559567.
Dalas, E., Koutsoukos, P.G., 1990. Calcium carbonate scale formation and prevention in a
ow-through system at various temperatures. Desalination 78, 403416.
Espenson, J.H., 1981. Chemical Kinetics and Reaction Mechanisms. McGraw-Hill,
New York.
Hasson, D., Sidorenko, G., Semiat, R., 2010. Calcium carbonate hardness removal by a
novelelectrochemical seeds system. Desalination. http://dx.doi.org/10.1016/j.desal.
2010.06.036.
He, S., Oddo, J.E., Thomson, M.B., 1994. The nucleation kinetics of calcium sulfate
dihydrate in NaCl solutions up to 6 m and 90 C. J. Colloid Interface Sci. 162, 297303.
Howard, J.R., Nancollas, G.H., Purdie, N., 1960. Trans. Faraday Soc. 56, 278.
Jamialahmadi, M., Muller-Steinhagen, H., 2000. Crystallization of calcium sulfate
dihydrate from phosphoric acid. Dev. Chem. Eng. Miner. Process. 8 (5/6), 587604.
Katakis, D., Gordon, G., 1987. Mechanisms of Inorganic Reactions. Wiley-Interscience,
New York.
Klepetsanis, P.G., Koutsoukos, P.G., 1991. Spontaneous precipitation of calcium sulfate at
conditions of sustained supersaturation. J. Colloid Interface Sci. 143 (2), 299308.
Klepetsanis, P.G., Dalas, E., Koutsoukos, P.G., 1999. Role of temperature in the spontaneous
precipitation of calcium sulfate dihydrate. Langmuir 15, 15341540.
Koutsoukos, P.G., Kontoyannis, C.G., 1984. Precipitation of calcium carbonate in aqueous
solutions. J. Chem. Soc. Faraday Trans. 1 (80), 11811192.
Kralj, D., Brecevic, L., Nielsen, A.E., 1990. Vaterite growth and dissolution in aqueous solution I. Kinetics of crystal growth. J. Cryst. Growth 104, 793800.
Lancia, A., Musmarra, D., Prisciandaro, M., 1999. Measuring induction period for calcium
sulfate dihydrate precipitation. AIChE J. 45 (2), 390397.
Liu, S.-T., Nancollas, G.H., 1970. The kinetics of crystal growth of calcium sulfate dihydrate.
J. Cryst. Growth 6, 281289.

153

Liu, S.-T., Nancollas, G.H., 1971. The kinetics of dissolution of calcium sulfate dihydrate. J.
Inorg. Nucl. Chem. 33, 23112316.
Liu, S.-T., Nancollas, G.H., 1973. Linear crystallization and induction-period studies of the
growth of calcium sulphate dihydrate crystals. Talanta 20, 211216.
Liu, S.-T., Nancollas, G.H., 1975. A kinetic and morphological study of the seeded growth of
calcium sulfate dihydrate in the presence of additives. J. Colloid Interface Sci. 52, 593.
Merdhah, A.B..B., Yassin, A.A.M., 2008. Laboratory study and prediction of calcium
sulphate at high-salinity formation water. Open Petr. Eng. J. 1, 6273.
Nancollas, G.H., Purdie, N., 1964. Q. Rev. (Lond.) 28, 1.
Nancollas, G.H., Purdie, N., 1968. The kinetics of crystal growth. Q. Rev. (Lond.) 28, 120.
Nancollas, G.H., Reddy, M.M., 1971. The crystallization of calcium carbonate II. Calcite
growth mechanism. J. Colloid Interface Sci. 37, 824830.
Oelkers, E.H., Helgeson, H.C., 1988. Calculation of the thermodynamic and transport properties of aqueous species at high pressures and temperatures: aqueous tracer diffusion coefcients of ions to 1000 C and 5 kb. Geochim. Cosmochim. Acta 52, 6385.
Petrou, A.L., 2012. The free energy of activation as the critical factor in geochemical
processes. Chem. Geol. 308309, 5059.
Petrou, A.L., Economou-Eliopoulos, M., 2009a. The activation energy values estimated by
the Arrhenius equation as a controlling factor of platinum-group mineral formation.
Geochim. Cosmochim. Acta 73, 16251636.
Petrou, A.L., Economou-Eliopoulos, M., 2009b. Platinum-group mineral formation:
evidence of an interchange process from the entropy of activation values. Geochim.
Cosmochim. Acta 73, 56355645.
Prisciandaro, M., Lancia, A., Musmarra, D., 2003. The retarding effect of citric acid on
calcium sulfate nucleation kinetics. Ind. Eng. Chem. Res. 42, 66476652.
Prisciandaro, M., Amedeo, L., Mazziotti di Celso, G., Musmarra, D., 2013. Antiscalants for
gypsum precipitation in industrial equipments: comparison among different
additives. Chem. Eng. Trans. 32, 21372142.
Rajeswara Rao, T., 1993. Kinetic parameters for decomposition of calcium carbonate. Can.
J. Chem. Eng. 71, 481484.
Reddy, M.M., Nancollas, G.H., 1971. The crystallization of calcium carbonate: I.isotopic exchange and kinetics. J. Colloid Interface Sci. 36, 166.
Rodriguez-Blanco, J.D., Shaw, S., Benning, L.G., 2011. The kinetics and mechanisms of
amorphous calcium carbonate (ACC) crystallisation to calcite, via Vaterite. Nanoscale
3, 265.
Romanek, C.S., Morse, J.W., Grossman, E.L., 2011. Aragonite kinetics in dilute solutions.
Aquat. Geochem.. http://dx.doi.org/10.1007/s10498-011-9127-2.
Schott, J., Brantley, S., Crerar, D., Guy, C., Borcsik, M., Willaime, C., 1989. Dissolution kinetics of strained calcite. Geochim. Cosmochim. Acta 53, 373382.
Sjoberg, E.L., 1976. A fundamental equation for calcite dissolution kinetics. Geochim.
Cosmochim. Acta 40, 441447.
Sjoberg, E.L., Rickard, D.T., 1984. Temperature dependence of calcite dissolution kinetics
between 1 and 62 C at pH 2.7 to 8.4 in aqueous solutions. Geochim. Cosmochim.
Acta 48, 485493.
Smith, B.R., Sweett, F., 1971. The crystallization of calcium sulfate dihydrate. J. Colloid
Interface Sci. 37 (3), 612618.
Wiechers, H.N.S., Sturrock, P., Marais, G.V.R., 1975. Calcium carbonate crystallization
kinetics. Water Res. 9, 835845.
Xyla, A.G., Giannimaras, E.K., Koutsoukos, P.G., 1991. The precipitation of calcium
carbonate in aqueous solutions. Colloids Surf. 53, 241255.
Yehia, N.S., Ali, M.M., Kandil, K.M., El-Maddawy, M.M., 2011. Effects of some parameters
affecting the crystallization rate of calcium sulfate dihydrate in sodium chloride
solution. J. Am. Sci. 7 (6), 635643.

Vous aimerez peut-être aussi