You are on page 1of 9

J. Chem.

Thermodynamics 150 (2020) 106202

Contents lists available at ScienceDirect

J. Chem. Thermodynamics
journal homepage: www.elsevier.com/locate/jct

Salting-out precipitation of NaCl, KCl and NH4Cl in mixtures of water


and methanol described by the modified Pitzer model
Alessandro C. Galvão a,⇑, Yecid P. Jiménez b, Francisca J. Justel b, Weber S. Robazza a, Fernanda S. Donatti a
a
Laboratory ApTher – Applied Thermophysics, Department of Food and Chemical Engineering, Santa Catarina State University – UDESC, Pinhalzinho, Brazil
b
Departamento de Ingeniería Química y Procesos de Minerales, Facultad de Ingeniería, Universidad de Antofagasta, Antofagasta, Chile

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

Article history: Experimental data and modelling of electrolyte solutions are important in different fields. Separation and
Received 15 April 2020 purification of such materials depend on knowledge of solid-liquid equilibrium as well as the mathemat-
Received in revised form 29 May 2020 ical models allow the development of solution theories and process simulators. Therefore, the current
Accepted 30 May 2020
work presents the study of NaCl, KCl and NH4Cl solubility in binary liquid solutions formed by water
Available online 2 June 2020
and methanol at temperatures ranging from 288.15 K to 328.15 K. Solubility decreases with increasing
methanol concentration and increases with increasing temperature, with the exception of the NaCl,
Keywords:
where the temperature has a minimum effect on its solubility. The modified Pitzer model was applied
Phase equilibrium
Solubility
to describe the solid-liquid equilibrium and its successful implementation also described the salting-
Electrolytes out precipitation by the addition of an anti-solvent.
Modified Pitzer model Ó 2020 Elsevier Ltd.
Crystallization

1. Introduction electrolyte solutions. Accordingly, Wu et al. [11] proposed a mod-


ified Pitzer model for the prediction of the liquid-liquid equilib-
Crystallization is a separation technology commonly applied as rium of polymer + salt + water systems. This modified model was
a purification strategy. Its use, widely recognized by the chemical later successfully applied by Lovera et al. [12], for the correlation
[1] and pharmaceutical sectors [2,3] may also be incorporated to of the solid-liquid equilibrium of the LiCl + PEG 4000 + H2O,
the food industry [4,5]. The operation consists of changing the NaCl + PEG 4000 + H2O, and KCl + PEG 4000 + H2O systems.
chemical potential of the target component in a saturated solution The modified Pitzer model takes into account the existence of
with consequent precipitation. short-range and long-range interactions described by binary and
The new equilibrium state induced to produce crystals can be ternary parameters. The model has been applied to determine differ-
achieved by different methods including the use of an anti- ent thermodynamic properties of electrolyte solutions [13–17]. The
solvent. The process, also known as drowning-out or salting-out generated parameters are fundamental as guidance for the beha-
precipitation [6], is based on the addition of a second solvent that viour of multicomponent systems associated to separation units.
changes the solubility of the dissolved substance. The major advan- In the interest to collaborate with studies concerning solubility
tage of the procedure is that it can be conducted at room temper- and crystallization of electrolytes, the present work provides
ature [7]. experimental data of NaCl, KCl and NH4Cl solubility in mixtures
Many kinds of crystallization techniques have been investigated of water and methanol covering the entire molar composition of
[8] and all of them involve consistent information on solubility. the binary liquid solution and at temperatures ranging from
Besides the experimental study, mathematical modelling is always 288.15 to 328.15 K. The solid-liquid equilibrium was simulated
required to predict phase equilibrium changes. Regarding elec- by the modified Pitzer model, and the modelling parameters were
trolytes, the use of models is challenging due to the complex inter- employed to predict salting-out precipitation.
actions that take place in a mixture involving ions [9].
The Pitzer model [10] has been widely used for the prediction 2. Experimental
and correlation of solubility in systems containing one or more
salts in a single solvent, generally water. However, this original 2.1. Materials
form of the model is not successfully applied for mixed-solvent
All reagents employed in this research were of analytical grade
⇑ Corresponding author. and supplied by Sigma-Aldrich. Solutions were prepared using
E-mail address: alessandro.galvao@udesc.br (A.C. Galvão). methanol and doubly distilled deionized water (0.054 lS cm1)

https://doi.org/10.1016/j.jct.2020.106202
0021-9614/Ó 2020 Elsevier Ltd.
2 A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202

Table 1 3.2. The modified Pitzer model


Source and purity of the chemicals used in this work.

Component Source Mass Purification According to the modified Pitzer model [11], the activity coeffi-
fraction* method cients, c, may be written as described by Eq. (2) in which the super-
CH3OH Sigma-Aldrich (France) 0.999 – scripts, LR and SR, stand for long-range and short-range
NaCl Sigma-Aldrich (United 0.995 Drying contributions, respectively.
States)
KCl Sigma-Aldrich (Czech 0.995 Drying lnci ¼ lncLR
i þ lnci
SR
ð2Þ
Republic)
NH4Cl Sigma-Aldrich (Spain) 0.995 Drying For electrolytes 1:1, and for non-ionic components the long-
range contribution is defined respectively by Eqs. (3) and (4)
*Purity of the chemical reagents are provided by the suppliers.
wherein V i is the molar volume of pure non-ionic species i and d
without any further purification. Prior to the experiment, the salts represents the mixed-solvent density.
were dried in an electrical furnace for at least 2 h at 353.15 K and 1
AI2
kept in a desiccator until the beginning of the experiments. The lncLR
 ¼  1
ð3Þ
source and purity of the chemicals used are presented in Table 1. 1 þ bI2
    
2AV i d 1 1
2.2. Apparatus and procedure lncLR
1 1
i ¼ 3
1 þ bI2  1 þ bI2  2 ln 1 þ bI2 ð4Þ
b
An Ohaus analytical balance (model PA214CP) with a precision The ionic strength (I), expressed by Equation (5), depends on
of ±0.1 mg was used to prepare all the solutions by mass. 0
the concentration (molality) of species j (m ) and its charge (Z).
The experiments were conducted in a jacketed glass cell cou-
The Debye-Hückel constants A and b are calculated as described
pled with a thermostatic bath Tecnal (model TE-2005, precision
by Eqs. (6) and (7).
of ±0.1 K) to ensure isothermal conditions. Solid and liquid com-
pounds were added inside the equilibrium cell. To achieve satura- X
N
0

tion, the mixture was strongly agitated for 3 h, and after this I ¼ 0:5 m j Z 2j ð5Þ
j–w
period, 5 h of rest was applied to allow the phase equilibrium.
Three samples were removed from the liquid phase and solubil- 0:5
ity was determined by the gravimetric method, according to the 1:327757  105 d
A¼ ð6Þ
methodology previously reported [18]. The assays were conducted ðDTÞ1:5
at atmospheric pressure and temperatures of 288.15 K, 298.15 K,
0:5
308.15 K, 318.15 K, and 328.15 K. The solubility values were 6:359696d
expressed as mole fraction. b¼ ð7Þ
ðDTÞ0:5
The methodology was tested by evaluating the solubility of
NaCl, KCl and NH4Cl in pure water and the results were compared The properties of the mixed solvent D and d are empirically cal-
0
with the recommended values published by Lorimer et al. [19]. culated using Eqs. (8) and (9) in which /i is the salt-free volume
fraction of non-ionic species i in the liquid phase.
3. Thermodynamic framework X 0
D¼ /i Di ð8Þ
3.1. Solubility product X 0
d¼ /i di ð9Þ
Bearing in mind the definition of ci , the expression of solubility
Table 2 shows the physical properties of methanol and water
product of anhydrous univalent salts (K sp ) is represented by Eq. (1)
used to calculate the Debye-Hückel constants A and b. These phys-
wherein x2 and c represent the saturation mole fraction (solubility)
ical properties were taken from Albright and Gosting [20],
and the mean ionic activity coefficient of salt, respectively. An
Kudryavtsev et al. [21] and González et al. [22] for methanol, and
expression for c is provided by the modified Pitzer model originally
from Malmberg et al. [23] and Kell [24] for water.
reported for mixtures involving water, polymer and electrolyte.
The short-range contributions, lncSR
i , for the three components
K sp ¼ x22 c2 ð1Þ may be written as presented by Eqs. (10)–(12).

Table 2
Density (d), dielectric constant (Di), and molar volume (Vi) of pure substances at different temperatures (T).

T/K d/kgm3 Di Vi  106/m3mol1


b,c e a d
CH3OH H2O CH3OH H2O CH3OH H2O
288.15 796.09 999.0996 34.70 81.946 40.25 18.04
298.15 786.56 997.0449 32.66 78.304 40.73 18.07
308.15 777.00 994.0319 30.74 74.828 41.24 18.13
318.15 765.61 990.2132 28.92 71.512 41.85 18.20
328.15 757.85 985.6952 27.21 68.345 42.28 18.28
a
[20].
b
[21].
c
[22].
d
[23].
e
[24].
A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202 3

Fig. 1. Solubility (x2) as a function of temperature (T), (a) NaCl in water:  this work, s [25], h [26], d [27], D [28], j [29], e [30]; (b) KCl in water:  this work, s [25], h [27],
j [29], ▲ [31], D [32], d [33], * [34]; (c) NH4Cl in water:  this work, h [30], d [31], s [35], + [36], D [37], e [38], * [39]

M1 nw  0  C 111 ¼ l111 ð15Þ


lncSR
1 ¼ 2B11 r 1 m1 þ B12 r 1 m2 
2
I B12 r 1 m1 m2
1000 ns
0
þ 2B22 m22 þ 3C 111 r31 m21 þ 2C 112 r21 m1 m2 þ C 122 r 1 m22 ð10Þ
vc va
B22 ¼ kca þ kcc þ kaa ð16Þ
2v a 2v c
1h  0 0

lncSR ¼ B12 r1 m1 þ 4B22 m2 þ I B12 r 1 m1 þ 2B22 m2

2 ð11Þ 3 
i C /222 ¼ lcca v c þ lcaa v a ð17Þ
þC 112 r 21 m21 þ 2C 122 r 1 m1 m2 þ 2Cc222 m22 ðv a v c Þ
1
2

  
M3 nw 0 3 /
lncSR ¼  B r 2 2
m þ B þ IB r 1 m1 m2 þ 2v c v a C c222 ¼ C ð18Þ
3
1000
11 1 1 12
ns 12 2 222
 
nw 0
B22 þ IB22 m22 þ 2C 111 r 31 m31 þ 2C 112 r 21 m21 m2 ð12Þ B12 ¼ 2ðk1c v c þ k1a v a Þ ð19Þ
ns
 
þ2C 122 r 1 m1 m22 þ 2ðv c v a Þ2 C /222 m32 C 112 ¼ 3 l1cc v 2c þ l1aa v 2a
3
ð20Þ

The terms mi , B11 , C 111 , B22 , C /222 , C c222 , B12 , C 112 , C 122 are described C 122 ¼ 3 l1cc v 2c þ 2l1ca v c v a þ l1aa v 2a ð21Þ
by Eqs. (13)–(21). 0 0
The terms B12 and B22 are the derivatives of B12 and B22 with
ni
mi ¼ ði ¼ 1; 2Þ ð13Þ respect to I, respectively. Using the same treatment as in the orig-
nwater inal Pitzer model, for uni-bivalent salts, the overall second virial
coefficients can be further expressed by Eq. (22) wherein bð0Þ , bð1Þ
B11 ¼ k11 ð14Þ
and a have similar meanings as in Pitzer’s virial model.
4 A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202

x2 x2

T/K T/K

x2

T/K
Fig. 2. Solubility (x2) as a function of temperature (T), (a) NaCl in methanol:  this work, h [26], D [28], j [29], ▲ [40]; (b) KCl in methanol:  this work, j [29], r [40]; (c)
NH4Cl in methanol:  this work, j [41].

Table 3
Comparison of experimental data from the present work and literature values of the CH3OH + NaCl or KCl systems: solubility expressed as mole fraction (x) for different
temperatures (T) at 0.1 MPa.

T/K NaCl KCl

xExp xLit xExp xLit


298.15 0.0077 0.0075a; 0.0076b; 0.0076c; 0.0075d 0.0022 0.0023b; 0.0024c; 0.0023d
308.15 0.0074 0.0076a; 0.0074c 0.0025 0.0025c
318.15 0.0070 0.0070c 0.0027 0.0027c
328.15 0.0069 0.0068c 0.0029 0.0028c
a
[26].
b
[28].
c
[40].
d
[42]; Standard uncertainties u are uðTÞ = 0.1 K;uðpÞ = 0.01 MPa; ur ðxExp Þ = 0.02

2bij h    i
ð1Þ
The similar pattern of the data presented in Fig. 1 indicates that
1  1 þ a1;ij I2 exp a1;ij I2
ð0Þ 1 1
Bij ¼ bij þ ð22Þ the applied methodology is satisfactory for solid-liquid equilibrium
a2
1;ij
essays. A comparison with the recommended solubility values [19]
The correlation of the ternary solid-liquid equilibrium data shows deviation between 0.10% and 0.31% for the NaCl-H2O mix-
shows that it is not sensitive to the a parameter, and any value tures; between 0.12% and 1.13% for the KCl-H2O mixtures and
of a selected near the one given by Pitzer [10] is acceptable. There- between 0.19% and 1.74% for NH4Cl-H2O mixtures.
fore, this value was fixed asa1;12 ¼ a1;22 ¼ 2:0: Furthermore, Fig. 2 shows the solubility of sodium chloride,
potassium chloride and ammonium chloride in methanol.
The results presented in Fig. 2 show agreement with literature
4. Results and discussion data. This finding is illustrated by the comparison with some cho-
sen data as presents Table 3 for the solubility of NaCl and KCl in
4.1. Experimental results methanol. Considering the literature results shown in Table 3 as
reference, it is estimated that the experimental data has an average
Fig. 1 shows the solubility expressed as mole fraction of sodium deviation of approximately 1.7% for the solubility of NaCl in metha-
chloride, potassium chloride and ammonium chloride in water nol, and 3.1% for the solubility of KCl in methanol. Regarding the
along with published results since 1993. solubility of NH4Cl in methanol, Stenger [41] reported a value of
A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202 5

x2 x2

x1 x1
Fig. 3. Solubility (x2 ) of (a) NaCl and (b) KCl in mixtures of CH3OH (1) + H2O (3) at 298.15 K as a function of methanol mole fraction (x1 ) in the salt free mixtures:  this work,
h [26], D [28], + [42] Y  m[46].

Table 4 observed that for all the studied mixtures the NH4Cl presents the
Parameters of the modified Pitzer model for the binary CH3OH (1) + H2O (3) system at highest solubility and KCl presents the lowest solubility. The addi-
different temperatures.
tion of alcohol in water changes the dielectric constant of the mix-
T/K B11  102 C 111  102 Refs. AAD  102,a ture to a value lower than that of pure water and consequently the
273.15 184.400 0.247 [45] 7.130 solubility of the electrolytes decreases.
298.15 207.960 0.361 [46] 1.552 The effect of methanol as an anti-solvent on the solubility of
323.15 239.739 0.502 [47] 3.062 NaCl and KCl in water has been investigated by several authors
328.15 251.224 0.522 [47] 3.411
[26,28,40,42]. Fig. 3 illustrates the effect of methanol at 298.15 K.
333.15 259.582 0.559 [47] 5.061
After the application of a polynomial equation to adjust the litera-
PN

exp cal

ture data, the results of the fitting were used to estimate the devi-
a i

cw;i cw;i

AAD ¼ N ; N is the number of data used in the regression. ation of the experimental values in comparison to the literature
results. Regarding the solubility of NaCl, the average deviation
from the values of Akerlof and Turck [42], Pinho and Macedo
0.0206 at 296.65 K in contrast with the value of 0.0204 at 298.15 K
[28], and Shi et al. [26] are 0.64%, 0.84% and 0.95%, respectively.
found in this work.
Regarding the solubility of KCl, the average deviation from the val-
Table S1 (supplementary data) show the NaCl, KCl and NH4Cl
ues of Akerlof and Turck [42] and Pinho and Macedo [28] are 4.47%
solubility results in the solutions formed by water + methanol at
and 4.40%, respectively.
different temperatures. As can be seen, the solubility of the three
electrolytes in water increases with increasing temperature; the
solubility of KCl and NH4Cl increases in methanol with increasing 4.2. Parameters estimation
temperature, and the solubility of NaCl in methanol decreases with
increasing temperature. The unexpected behaviour for the solubil- The relationship between the solubility of the salt and solvent
ity of NaCl in methanol is explained [29] as the consequence of concentration is given by Eq. (1). Solving this equation for the sol-
association effect. For the range under investigation, the increase ubility requires knowing K sp value and the parameters of the activ-
in temperature is more noticeable for the mixtures involving NH4- ity coefficient model called the characteristic coefficients of the salt
ð0Þ ð1Þ
Cl followed by KCl and NaCl, respectively. Furthermore, it is (b22 , b22 and C /222 ) and the alcohol (B11 and C111), and the cross

aw

m -1
x1

Fig. 4. (a) Comparison of Y  m versus for the aqueous electrolyte mixtures at 308.15K: r NaCl, ▲ KCl, d NH4Cl and — calculated from the modified Pitzer model; (b)
comparison of water activity versus mole fraction of methanol at 298.15 K:  [46], calculated from the modified Pitzer model, — ideal solution.
6 A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202

Table 5
Parameters of the modified Pitzer model for the binary aqueous electrolyte solutions at different temperatures.

T/K ð0Þ
b22  102
ð1Þ
b22  102 C /222  102 K sp Refs. AAD  102 a

NaCl (2) + H2O (3)


288.15 4.639 1.115 0.330 1.420 [43] 0.079
298.15 5.409 1.240 0.250 1.506 [43] 0.082
308.15 5.986 1.246 0.187 1.575 [43] 0.084
318.15 6.413 1.164 0.135 1.625 [43] 0.086
328.15 6.721 1.017 0.093 1.647 [43] 0.087
KCl (2) + H2O (3)
288.15 1.587 1.266 0.243 0.237 [43] 0.028
298.15 2.178 0.894 0.195 0.308 [43] 0.023
308.15 2.638 0.672 0.158 0.391 [43] 0.019
318.15 2.988 0.552 0.129 0.478 [43] 0.019
328.15 3.248 0.505 0.106 0.574 [43] 0.021
NH4Cl (2) + H2O (3)
288.15 2.696 2.903 0.003 0.560 [44] 0.089
298.15 2.873 2.915 0.008 0.715 [44] 0.088
308.15 2.959 2.934 0.014 0.870 [44] 0.086
318.15 2.970 2.965 0.016 1.031 [44] 0.084
328.15 2.918 3.005 0.013 1.238 [44] 0.081

PN

exp cal

a i

cw;i cw;i

AAD ¼ N ;N is the number of data used in the regression.

Table 6
Cross parameters of the modified Pitzer model for the studied ternary mixtures at different temperatures.

T/K ð0Þ
b12  102
ð1Þ
b12  102 C 112  102 C 122  102 AAD  102 a

CH3OH (1) + NaCl (2) + H2O (3)


288.15 11.121 6.159 0.0017 0.102 0.19
298.15 7.978 4.794 0.0064 0.414 0.11
308.15 7.142 4.471 0.0068 0.452 0.12
318.15 6.417 4.103 0.0064 0.461 0.11
328.15 2.643 1.913 0.0118 0.734 0.18
CH3OH (1) + KCl (2) + H2O (3)
288.15 9.142 2.600 0.0043 0.974 0.002
298.15 8.800 2.723 0.0036 0.788 0.003
308.15 8.416 2.750 0.0035 0.658 0.004
318.15 7.886 2.706 0.0033 0.577 0.005
328.15 7.450 2.756 0.0029 0.532 0.007
CH3OH (1) + NH4Cl (2) + H2O (3)
288.15 5.118 5.307 0.0086 0.223 0.002
298.15 5.598 5.828 0.0053 0.130 0.004
308.15 4.803 5.546 0.0065 0.143 0.007
318.15 4.908 5.644 0.0046 0.093 0.015
328.15 6.035 6.542 0.0002 0.004 0.033
PN exp cal
a jx x j
AAD ¼ i 2N 2 ;N is the number of data used in the regression.

ð0Þ ð1Þ three electrolytes is similar. Fig. 4(b) shows the good agreement
virial coefficients between the salt and the alcohol (b12 , b12 , C112
ð0Þ ð1Þ between the experimental and calculated values of the water activ-
and C122). The parameters and b22 , b22
have different values C /222
ity for the CH3OH + H2O system. Also, it was observed that the
from those in the original Pitzer model and should be correlated.
system is far behaving as the ideal solution, with the increased
The binary parameters for the salt have been estimated by fit-
concentration of methanol.
ting data from the literature at T = (288.15, 298.15, 308.15,
The solubility product, K sp , for NaCl, KCl and NH4 were deter-
318.15 and 328.15) K, following the same procedure as used by
mined from the experimental data reported by Pitzer [43] and Ji
Wu et al. [11] and Lovera et al. [12]. The data for the NaCl + H2O
et al. [44]. These values are shown in Table 5. The unknowns
and KCl + H2O systems were taken from Pitzer [43] and for the
parameters in Eq. (1) were reduced to the solubility of the salt,
NH4Cl + H2O system were taken from Ji et al. [44]. Values of the
ð0Þ ð1Þ
ð0Þ ð1Þ x2, and the salt-alcohol cross parameters b12 , b12 , C112 and C122.
parameters b22 , b22 and C /222 are shown in Table 4. The second
In solving the equation for the solid-liquid equilibrium, the infor-
and third coefficients for methanol (B11 and C111) were calculated
mation from the solubility in the ternary system CH3OH + salt +
from the data reported by Brown et al. [45], Butler et al. [46] and
H2O has been fitted using the least squares criteria to find the
Kurihara et al. [47]. These parameters are shown in Table 4.
unknown parameters. Before minimization, Eq. (1) was linearized
The average absolute deviation values show a good agreement in
and Eq. (23) was used as objective function. The sum includes all
the correlation of the data of binary aqueous systems.
experimental data.
Fig. 4(a) shows the good agreement between experimental and
calculated data for NaCl, KCl and NH4Cl at 308.15 K (similar results N 
X 2
exp cal
were obtained to the other temperatures). It can be observed that OF ¼ lnx2  lnx2 ð23Þ
i
i¼1
the variation of the activity coefficient with the molality of the
A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202 7

Fig. 5. Solubility of (a) NaCl and (b) KCl in CH3OH (1) + H2O (3) mixtures at 298.15 K: } experimental data; — calculated data.

ð0Þ ð1Þ percentage, defining yield (Y) as mass of salt precipitate/mass of


The results of the estimated parameters (b12 , b12 , C112 and C122)
and the AAD are shown in Table 6. Fig. 5 shows the results of the solution. In the present work, the precipitated amount and anti-
solubility calculations. These results indicate a good agreement solvent percentage data were fitted to empirical equations and
with the experimental data at different temperatures. The validity maximized. Table 7 shows the anti-solvent percentages at different
of this model has been demonstrated for univalent salts and one temperatures that produces the maximum amount of precipitated.
mixed solvent, CH3OH + H2O. Table 7 shows that for NaCl, the optimum mass fraction of
Fig. 5 shows the significant effect of the anti-solvent in the methanol is practically the same for the five temperatures studied.
reduction of the salt solubility, for example at 0.4 mole fraction On the other hand, for KCl and NH4OH, the higher yield is achieved
of anti-solvent, the KCl solubility decreases by approximately at 328.15 K. This is evidently correct, given that the solubility of the
75%. According to this behaviour that present the three systems system increases with temperature and therefore at higher tem-
studied, quantifying the amount of salt precipitated (or the super- peratures the amount of dissolved salt is greater.
saturation) for different percentages of anti-solvent can provide
important information. This calculation can be carried out starting
from an initial equilibrium condition, in which the salt is saturated
in water. In this condition no salt precipitates. If any amount of
anti-solvent is added to this system, precipitation is produced, fol-
lowed by the establishing of a new equilibrium condition in which
the salt will be less in quantity and saturated again, but now in a
new water + anti-solvent medium. Substituting this new concen-
tration in Eq. (1) and developing it, the following expression to
quantify the precipitated amount is obtained.

K sp
P2  2x0 P þ x20  ¼0 ð24Þ
c2
In Eq. (24) x0 is the salt mole fraction in the initial saturated
solution, and P represents the precipitated amount (in mole frac-
tion) obtained when the anti-solvent is aggregated to the system.
The precipitated amount can be also expressed in g kg1 of solvent
considering the system composition. This procedure was applied to
the different anti-solvent percentages and the results obtained at
328.15 K, for the three systems studied are shown in Fig. 6. This fig-
ure shows that for the same anti-solvent weight percent, a higher
amount of KCl precipitates in comparison with NaCl and NH4Cl.
In general, a similar behaviour to that described in Fig. 6 was
observed at the other temperatures studied. From the value of P
is possible to calculate the percentage of anti-solvent that produces
the maximum precipitate. The usual way to determine this Fig. 6. Amount of salt precipitated versus anti-solvent weight percent at 328.15 K;
optimum condition is through a graphic yield versus anti-solvent j NaCl, ▲ KCl and d NH4Cl.
8 A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202

Table 7
Anti-solvent weight percentages w% that produces the maximum amount of precipitated P at different temperatures T.

T w1% P/g∙kg1 w1% P/g∙kg1 w1% P/g∙kg1


NaCl KCl NH4Cl
288.15 79.9 165.2 71.3 181.7 92.4 141.5
298.15 79.5 164.0 73.3 191.8 94.5 160.7
308.15 79.6 163.7 75.7 202.1 94.2 172.1
318.15 80.0 164.2 78.1 210.5 93.9 180.5
328.15 80.8 166.0 80.4 220.1 98.4 193.1

5. Conclusions [2] J. Wang, F. Li, R. Lakerveld, Process intensification for pharmaceutical


crystallization, Chem. Eng. Process. – Process Intensif. 127 (2018) 111–126,
https://doi.org/10.1016/j.cep.2018.03.018.
From the experimental results is concluded that the solubility of [3] T. Wang, H. Lu, J. Wang, Y. Xiao, Y. Zhou, Y. Bao, H. Hao, Recent progress of
NaCl, KCl and NH4Cl depends on temperature and composition of continuous crystallization, J. Ind. Eng. Chem. 54 (2017) 14–29, https://doi.org/
10.1016/j.jiec.2017.06.009.
the mixture solvents. The dependence on composition is related
[4] P. MacFhionnghaile, V. Svoboda, J. McGinty, A. Nordon, J. Sefcik, Crystallization
to the change on dielectric constant of solvent due to the addition diagram for anti-solvent crystallization of lactose: using design of experiments
of an anti-solvent. For all the studied mixtures the NH4Cl presented to investigate continuous mixing-induced supersaturation, Cryst. Growth Des.
17 (2017) 2611–2621, https://doi.org/10.1021/acs.cgd.7b00136.
the highest solubility and KCl presented the lowest solubility.
[5] M.H. Zamanipoor, R.L. Mancera, The emerging application of ultrasound in
The modified Pitzer model was successfully applied generating lactose crystallisation, Trends Food Sci. Technol. 38 (2014) 47–59, https://doi.
important binary and ternary parameters. The model performance org/10.1016/j.tifs.2014.04.005.
permitted a quantitative evaluation of an anti-solvent crystalliza- [6] S. Nalesso, M.J. Bussemaker, R.P. Sear, M. Hodnett, J. Lee, Development of
sodium chloride crystal size during anti-solvent crystallization under different
tion phenomenon, as product of this analysis was predicted that sonication modes, Cryst. Growth Des. 19 (2019) 141–149, https://doi.org/
the KCl presents the highest yield, which is in agreement with 10.1021/acs.cgd.8b01172.
the solubility data. [7] I. Dammak, M. Neves, H. Isoda, S. Sayadi, M. Nakajima, Recovery of polyphenols
from olive mill wastewater using drowning-out crystallization based
From the methodology proposed in the present work, it has separation process, Innov. Food Sci. Emerg. Technol. 34 (2016) 326–335,
been possible to obtain valuable information that could be useful https://doi.org/10.1016/j.ifset.2016.02.014.
in the design of the NaCl, KCl and NH4Cl crystallization processes [8] H. Lu, J. Wang, T. Wang, N. Wang, Y. Bao, H. Hao, Crystallization techniques in
wastewater treatment: an overview of applications, Chemosphere 173 (2017)
by means of the addition of an anti-solvent. Thus, in a future work, 474–484, https://doi.org/10.1016/j.chemosphere.2017.01.070.
it would be interesting to prove the reproducibility of the model [9] J.A. Fournier, W. Carpenter, L. De Marco, A. Tokmakoff, Interplay of ion–water
through experimental tests, for the validation of the model. and water–water interactions within the hydration shells of nitrate and
carbonate directly probed with 2D IR spectroscopy, J. Am. Chem. Soc. 138
(2016) 9634–9645, https://doi.org/10.1021/jacs.6b05122.
CRediT authorship contribution statement [10] K.S. Pitzer, Thermodynamics of electrolytes. I. Theoretical basis and general
equations, J. Phys. Chem. 77 (1973) 268–277, https://doi.org/10.1021/
j100621a026.
Alessandro C. Galvão: Conceptualization, Methodology, Writ- [11] Y.-T. Wu, D.-Q. Lin, Z.-Q. Zhu, L.-H. Mei, Prediction of liquid-liquid equilibria of
ing - original draft, Writing - review & editing. Yecid P. Jiménez: polymer salt aqueous two-phase systems by a modified Pitzer’s virial
equation, Fluid Phase Equilib. 124 (1996) 67–79, https://doi.org/10.1016/
Formal analysis, Writing - original draft, Writing - review & editing. S0378-3812(96)03066-X.
Francisca J. Justel: Formal analysis, Writing - original draft, Writ- [12] J.A. Lovera, A.P. Padilla, H.R. Galleguillos, Correlation of the solubilities of alkali
ing - review & editing. Weber S. Robazza: Investigation. Fernanda chlorides in mixed solvents: polyethylene glycol+H2O and Ethanol+H2O,
Calphad 38 (2012) 35–42, https://doi.org/10.1016/j.calphad.2012.03.002.
S. Donatti: Investigation.
[13] L. Ma, S. Li, Q. Zhai, Y. Jiang, M. Hu, Thermodynamic study of RbF/CsF in amino
acid aqueous solution based on the Pitzer, modified Pitzer, and extended
Declaration of Competing Interest Debye-Hückel models at 298.15 K by a potentiometric method, Ind. Eng.
Chem. Res. 52 (2013) 11767–11772, https://doi.org/10.1021/ie401411r.
[14] F. Deyhimi, M. Abedi, NaCl + CH3OH + H2O mixture: Investigation using the
The authors declare that they have no known competing finan- Pitzer and the modified Pitzer approaches to describe the binary and ternary
cial interests or personal relationships that could have appeared ion–nonelectrolyte interactions, J. Chem. Eng. Data 57 (2012) 324–329,
https://doi.org/10.1021/je201084a.
to influence the work reported in this paper. [15] G. Villca, F.J. Justel, Y.P. Jimenez, Water activity, density, sound velocity,
refractive index and viscosity of the (NH4)6Mo7O24 + poly(ethylene glycol) +
Acknowledgments H2O system in the temperature range from 313.15 to 333.15 K: experiment
and modelling, J. Chem. Thermodyn. 142 (2020) 105986, https://doi.org/
10.1016/j.jct.2019.105986.
The authors would like to thank CAPES (Coordenação de Aper- [16] Y.P. Jiménez, F.J. Justel, Measurement and modelling of water activities of the
feiçoamento de Pessoal de Nível Superior) and FAPESC (Fundação de CuSO4 + poly(ethylene glycol) + H2O system in the temperature range from
303.15 to 333.15 K, J. Chem. Thermodyn. 144 (2020) 106064, https://doi.org/
Amparo à Pesquisa e Inovação do Estado de Santa Catarina, project 10.1016/j.jct.2020.106064.
number 2019RT583) for the financial support. [17] J. Tang, S. Li, Q. Zhai, Y. Jiang, M. Hu, Osmotic and activity coefficient
investigation on the CsNO3+methanol+water and CsNO3+ethanol+water
ternary systems at 298.15K, J. Mol. Liq. 195 (2014) 205–211, https://doi.org/
Appendix A. Supplementary data 10.1016/j.molliq.2014.02.034.
[18] A.C. Galvão, W.S. Robazza, P.F. Arce, C. Capello, D.H. Hagemann, Experimental
study and modelling of citric acid solubility in alcohol mixtures, J. Food Eng.
Supplementary data to this article can be found online at 237 (2018) 96–102, https://doi.org/10.1016/j.jfoodeng.2018.05.032.
https://doi.org/10.1016/j.jct.2020.106202. [19] J.W. Lorimer, M. Salomon, C.L. Young, Solubility data series - Alakli metal and
ammonium chlorides in water and heavy water (binary systems), 1991.
[20] P.S. Albright, L.J. Gosting, Dielectric constants of the methanol-water system
References from 5 to 55° 1, J. Am. Chem. Soc. 68 (1946) 1061–1063, https://doi.org/
10.1021/ja01210a043.
[1] L. Peng, H. Dai, Y. Wu, Y. Peng, X. Lu, A comprehensive review of phosphorus [21] S.G. Kudryavtsev, A.N. Strakhov, O.V. Ershova, G.A. Krestov, Volume properties
recovery from wastewater by crystallization processes, Chemosphere 197 of the water-methanol system of different deutero-substitutions at 278–318-
(2018) 768–781, https://doi.org/10.1016/j.chemosphere.2018.01.098. K, Zhurnal Fiz. Khimii. 60 (1986) 2202–2205.
A.C. Galvão et al. / J. Chem. Thermodynamics 150 (2020) 106202 9

[22] E.J. González, L. Alonso, Á. Domínguez, Physical properties of binary mixtures [36] S. Sawamura, N. Yoshimoto, Y. Taniguchi, Y. Yamaura, Effects of pressure and
of the ionic liquid 1-methyl-3-octylimidazolium chloride with methanol, temperature on the solubility of ammonium chloride in water, High Press. Res.
ethanol, and 1-propanol at T = (298.15, 313.15, and 328.15) K and at P = 0.1 16 (1999) 253–263, https://doi.org/10.1080/08957959908200298.
MPa, J. Chem. Eng. Data 51 (2006) 1446–1452, https://doi.org/ [37] Y. Zeng, Z. Li, G.P. Demopoulos, Phase equilibria for the glycine–methanol–
10.1021/je060123k. NH4 Cl–H2O system, Ind. Eng. Chem. Res. 53 (2014) 16864–16872, https://doi.
[23] C.G. Malmberg, A.A. Maryott, Dielectric constant of water from 0 to 100 C, J. org/10.1021/ie502846m.
Res. Natl. Bur. Stand. 56 (1956) (1934) 1, https://doi.org/10.6028/jres.056.001. [38] A. Apelblat, The vapour pressures of saturated aqueous lithium chloride,
[24] G.S. Kell, Density, thermal expansivity, and compressibility of liquid water sodium bromide, sodium nitrate, ammonium nitrate, and ammonium chloride
from 0.deg. to 150.deg. Correlations and tables for atmospheric pressure and at temperatures from 283 K to 313 K, J. Chem. Thermodyn. 25 (1993) 63–71,
saturation reviewed and expressed on temperature scale, J. Chem. Eng. Data. https://doi.org/10.1006/jcht.1993.1008.
20 (1975) (1968) 97–105, https://doi.org/10.1021/je60064a005. [39] T. He, J. Sun, W. Shen, Y. Ren, Solid-liquid phase equilibria of quaternary
[25] F. Farelo, G. Von Brachel, H. Offermann, Solid-liquid equilibria in the ternary system NH4 + //C1 , SO4 2, H2PO4  –H2O and its subsystems NH4 + //C1
system nacl-kcl-h2o, Can. J. Chem. Eng. 71 (1993) 141–146, https://doi.org/ , SO4 2 –H2O, NH4 + //C1 , H2PO4  –H2O at 313.15 K, J. Chem.
10.1002/cjce.5450710119. Thermodyn. 112 (2017) 31–42, https://doi.org/10.1016/j.jct.2017.04.010.
[26] J. Shi, J. Hu, L. Li, Y. Guo, X. Yu, T. Deng, Solid-liquid phase equilibria of the [40] M. Li, D. Constantinescu, L. Wang, A. Mohs, J. Gmehling, Solubilities of NaCl,
ternary system (NaCl + CH3OH + H2O) at 298.15, 308.15, 318.15 K, and 0.1 KCl, LiCl, and LiBr in methanol, ethanol, acetone, and mixed solvents and
MPa, J. Chem. 2018 (2018) 1–8, https://doi.org/10.1155/2018/4849639. correlation using the LIQUAC model, Ind. Eng. Chem. Res. 49 (2010) 4981–
[27] Y. Gao, D. Zhao, S. Li, Q. Zhai, Y. Jiang, M. Hu, Measurement and correlation of 4988, https://doi.org/10.1021/ie100027c.
solubilities and solution thermodynamics for N, N-diethylformamide + MCl (M [41] V.A. Stenger, Solubilities of various alkali metal and alkaline earth metal
= Na, K, Rb, and Cs) + water systems in the temperature range 288.15–338.15 compounds in methanol, J. Chem. Eng. Data. 41 (1996) 1111–1113, https://doi.
K, J. Chem. Eng. Data 61 (2016) 1649–1656, https://doi.org/10.1021/acs. org/10.1021/je960124k.
jced.5b01043. [42] G. Åkerlöf, H.E. Turck, The solubility of some strong, highly soluble electrolytes
[28] S.P. Pinho, E.A. Macedo, Representation of salt solubility in mixed solvents: a in methyl alcohol and hydrogen peroxide—water mixtures at 25°, J. Am. Chem.
comparison of thermodynamic models, Fluid Phase Equilib. 116 (1996) 209– Soc. 57 (1935) 1746–1750, https://doi.org/10.1021/ja01312a074.
216, https://doi.org/10.1016/0378-3812(95)02889-7. [43] K.S. Pitzer, in: Activity Coefficients in Electrolyte Solutions, CRC Press, 2018,
[29] S.P. Pinho, E.A. Macedo, Solubility of NaCl, NaBr, and KCl in water, methanol, https://doi.org/10.1201/9781351069472.
ethanol, and their mixed solvents, J. Chem. Eng. Data. 50 (2005) 29–32, https:// [44] X. Ji, X. Lu, L. Zhang, N. Bao, J. Yanru Wang, B.C.-Y.Lu. Shi, A further study of
doi.org/10.1021/je049922y. solid–liquid equilibrium for the NaCl–NH4Cl–H2O system, Chem. Eng. Sci. 55
[30] Y. Zeng, Z. Li, Solubility measurement and modelling for the NaCl–NH4Cl– (2000) 4993–5001, https://doi.org/10.1016/S0009-2509(00)00113-5.
monoethylene glycol–H2O system from (278 to 353) K, J. Chem. Eng. Data. 60 [45] W. Brown, D.S. Martin, Activity coefficients of aqueous methanol solutions,
(2015) 2248–2255, https://doi.org/10.1021/acs.jced.5b00053. 1951.
[31] L. Zhang, Q. Gui, X. Lu, Y. Wang, J. Shi, B.C.-Y. Lu, Measurement of solidliquid [46] J.A.V. Butler, D.W. Thomson, W.H. Maclennan, The free energy of the normal
equilibria by a flow-cloud-point method, J. Chem. Eng. Data. 43 (1998) 32–37, aliphatic alcohols in aqueous solution. Part I. The partial vapour pressures of
https://doi.org/10.1021/je970176p. aqueous solutions of methyl, n-propyl, and n-butyl alcohols. Part II. The
[32] P. Sousa, A.M.C. Lopes, Solubilities of potassium hydrogen tartrate and solubilities of some normal aliphatic alcohols in water. Part III, J. Chem. Soc.
potassium chloride in water + ethanol mixtures, J. Chem. Eng. Data 46 674 (1933), 10.1039/jr9330000674.
(2001) 1362–1364, https://doi.org/10.1021/je010105x. [47] K. Kurihara, T. Minoura, K. Takeda, K. Kojima, Isothermal vapor-liquid
[33] L.-X. Wang, Q. Xia, J. Kang, M.-X. Du, G.-L. Zhang, F.-B. Zhang, Measurement equilibria for methanol + ethanol + water, methanol + water, and ethanol +
and correlation of solubilities of potassium chloride and potassium sulfate in water, J. Chem. Eng. Data 40 (1995) 679–684, https://doi.org/
aqueous glycerol solutions, J. Chem. Eng. Data 56 (2011) 3813–3817, https:// 10.1021/je00019a033.
doi.org/10.1021/je200238s.
[34] O. Chiavone-Filho, P. Rasmussen, Solubilities of salts in mixed solvents, J.
Chem. Eng. Data 38 (1993) 367–369, https://doi.org/10.1021/je00011a009.
[35] N. Tang, W. Shi, W. Yan, Modified method for measuring the solubility of
pharmaceutical compounds in organic solvents by visual camera, J. Chem. Eng.
Data 61 (2016) 35–40, https://doi.org/10.1021/acs.jced.5b00122. JCT 2020-265

You might also like