Sie sind auf Seite 1von 5

Ind. Eng. Chem. Res.

2008, 47, 1283-1287 1283

Boiling Point Rise Calculations in Sodium Salt Solutions


Marta Bialik,* Peter Sedin, and Hans Theliander
Forest Products and Chemical Engineering, Chalmers UniVersity of Technology, SE-412 96 Göteborg, Sweden

The boiling point rise (elevation) of aqueous industrial solutions is often regarded as an important property
with respect to chemical process design. This work shows an application of the Pitzer method for calculating
the activity coefficients to the estimation of the boiling point rise of single-component and multicomponent
electrolyte solutions. Good agreement between experimental and predicted values of the boiling point elevations
of solutions of several salts (NaOH, Na2CO3, Na2SO4, NaCl, Na2S2O3, Na2S, mixed NaOH-Na2CO3, and
mixed Na2CO3-Na2SO4) was obtained. A method for using the boiling point rise data to obtain ionic interaction
parameters for the Pitzer method is also shown.

Introduction Theory

Boiling point rise, also called boiling point elevation, has long Many properties of solutions, including boiling point eleva-
been recognized as an important property of numerous industrial tion, freezing point depression, and osmotic pressure, can be
solutions. Boiling point rise is defined as the difference in related to the vapor pressure of the solvent and hence to its
boiling temperatures of a given solution and its pure solvent activity. These properties are often called “colligative” properties
when measured at the same pressure. of a solution. In the rational system, the chemical potential,
µl, of a liquid solvent in a solution is expressed in terms
The estimation of a true boiling temperature is an important
step in the design of boiling, evaporation, and heat-exchange of the chemical potential of the pure liquid, µ0l , and its activity,
processes such as black liquor evaporation for pulp and paper al, as7
production, industrial crystallization, liquid waste handling, etc.
For this reason, many attempts have been made to develop a µl(T,p) ) gl(T,p) + RT ln ai ) µ0i (T,p) + RT ln ai (1)
reliable method for estimating the boiling point from measurable
solution properties.1 In the case of multicomponent industrial where al might be a function of temperature, pressure, and
solutions, especially those containing various organic com- composition. If the liquid solvent is in equilibrium with its vapor,
pounds (e.g., black liquor from the pulp and paper industry), the equilibrium conditions can be written as an equality in
empirical or semi-empirical methods relating density or solid chemical potential7
content to boiling point elevation have frequently been used.2
Dühring’s rule, which involves plotting the boiling temperature µ0i (T,p) + RT ln ai ) µvap(T,p) (2)
of a solution versus the temperature of the pure solvent at
different pressures, has also been applied.3 Because the difference between the chemical potential of
Estimating the boiling point elevation for multicomponent vapor, µvap, and the chemical potential of pure liquid, µ0l , at
industrial solutions is usually a challenging task. Both rigorous temperature T corresponds to the free energy of vaporization,
thermodynamic methods and empirical methods based on easily ∆G0vap, of the pure liquid at this temperature, eq 2 can be
measurable solution properties have been used for this purpose. rearranged to7
For the rigorous thermodynamic methods, the calculation
procedure is often very complicated because of the large number -∆G0vap
of chemical compounds present in the solution. Moreover, the ln a ) (3)
RT
exact composition of many industrial solutions (e.g., black
liquor) is often difficult to determine with sufficient accuracy, Differentiating eq 3 with respect to activity and applying the
and the appropriate thermodynamic parameters for less typical Gibbs-Helmholtz equation gives7
compounds are frequently not available. The empirical methods,

( )
on the other hand, usually have no support in solution theory 0
∆Hvap ∂T
1
and thus have very limited applications, modeling only one type )- (4)
of solution with acceptable precision. al RT2 ∂al p

The purpose of this work is to present a comprehensive set


of boiling point rise data from single-component and multi- where ∆H0vap is the heat of vaporization. This expression can
component salt solutions and to suggest a computational method then be integrated with the lower integration limit corresponding
for predicting boiling point rise values for industrial salt systems to pure liquid, boiling at temperature T0, and the upper limit to
of great interest. For this application, the Pitzer method for a solution with solvent activity a, boiling at temperature T. For
calculating the activity coefficients4,5 with interaction coef- simplicity, the heat of vaporization is often assumed to be
ficients from Pitzer5 and Weber6 was chosen, and the boiling independent of temperature within the small temperature range
point elevation was calculated using Pitzer’s osmotic coefficient. between T0 and T. Thus, the final equation takes the form7

( ) [ ]
∆H0vap 1 1 ∆H0vap 1 1
* To whom correspondence should be addressed. E-mail: bialik@ ln a ) - ) - (5)
chalmers.se. Fax: +46-31-772-2995. R T T0 R (T0 + θ) T0

10.1021/ie070564c CCC: $40.75 © 2008 American Chemical Society


Published on Web 01/23/2008
1284 Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008

( )x
where “0” refers to the properties of the pure solvent (water) 2πd0NA
1 e 3
and θ is the boiling point elevation. If the activity of the solvent Aφ ) (8)
is known, it is theoretically possible to use the eq 5 to predict 3 xDkT 1000
the boiling point rise of the solution.
There are numerous methods for calculating water activity where D is the dielectric constant of the solvent, d0 is the solvent
in aqueous solutions; however, only a few of them can be density, T is the temperature, k is the Boltzmann constant, NA
applied to high-temperature, concentrated, multicomponent is Avogadro’s number, and e is the unit charge. In addition to
systems. In this work, the semi-empirical Pitzer activity coef- its direct temperature dependence, AΦ includes two other
ficient method4,5 was chosen. This method has been widely used temperature-dependent quantities: D and d0. It is thus fairly
for estimating activity coefficients in applications ranging from easy to calculate AΦ for a desired temperature. The temperature-
the simple characterization of salt solutions to multifunctional dependent polynomials for the interaction coefficients in eq 7
commercial simulation packages. The method, combining the are often tabulated for common salt systems. This work uses
virial equation of state with the equation for potential energy the temperature-dependence coefficients for the Pitzer param-
between ions, utilizes empirical interaction parameters of the eters tabulated by Weber;6 for the systems whose coefficients
species present in the system. Consequently, the method requires were not available, only the Debye-Hückel constant modifica-
a set of well-established interaction parameters valid for the tion was applied. For the sake of simplicity, Chen’s formula,4
desired temperature, pressure, and concentration conditions. The which approximates the temperature dependence of Debye-
basic interaction parameters from Pitzer5 and Weber,6 presented Hückel constants excellently, was used in this work
in the Supporting Information, were used in this work (see
below).
For electrolyte solutions, the water activity calculation is
AΦ ) -61.44534 exp (T -273.15
273.15
)+
based on the definition of the practical osmotic coefficient, φ8 T - 273.15 2
2.864468[exp(
273.15 )]
+
-1000 ln a T
φ≡ (6) 183.5379 ln - 0.6820223(T - 273.15) +
273.15
Mw ∑
i)1
νimi
0.0007875695(T2 - 273.152) + 58.95788
273.15
(9)
T
where Mw is the molecular mass of the solvent (water), mi is Because the entire calculation procedure is thoroughly explained
the molality of the ith solute, and νi is the number of ions into by Pitzer,5 it is relatively simple to program it for easy use
which the solute dissociates. Pitzer5 proposed the following with a desktop computer or as part of more complex simula-
equation for the simplified calculation of the osmotic coefficient tion software. This work used a simple program written for

[
MATLAB (Mathworks, Inc., Natick, MA) software.
-AφI3/2 Estimation of Pitzer Parameters. The boiling point rise
(φ - 1) ) (2/ ∑i mi) + ∑c ∑a Intca + calculation procedure presented above is very interesting.
1 + 1.2I1/2

]
Because eq 5 shows a mutual relationship between the boiling
point elevation and the corresponding solvent activity, it is
∑ ∑ Intcc′(a) + ∑
c< c′
∑ Intaa′(c)
a< a′
(7) possible to use this procedure to calculate water activity in a
given aqueous solution if the value of a corresponding boiling
point rise is known. Once the water activity is known, it is
where AΦ is the Debye-Hückel constant describing solvent possible to calculate the osmotic coefficient in the Pitzer
properties and I is the total ionic strength. The first summa- equation according to eq 6. Consequently, reversing the Pitzer
tion term inside the square brackets represents the ionic equation for the osmotic coefficient (eq 7) makes it possible to
interactions in terms of Debye-Hückel theory. The second term convert a set of well-established boiling point rise data into a
summarizes double-interaction terms, involving ions of different set of Pitzer empirical interaction parameters for a given salt
signs (c, cation; a, anion). In analogy, the third and fourth terms solution. This procedure could be applied for estimating the
inside the square brackets summarize interactions of ions of Pitzer interaction parameters of complex solutions, especially
the same sign and also involve triple interaction terms (c-c′-a those involving atypical salts for which the original Pitzer
or a-a′-c). The equations for the subsequent interaction terms parameters are unknown (e.g., black liquor) and multicomponent
are given in the Supporting Information, and a detailed mixtures for which the higher-order interaction parameters have
description of the calculation procedure can be found else- not been identified.
where.4,5
Ionic activities in a typical electrolyte system are highly
Literature Data
temperature-dependent; the activity coefficients in an ambient-
temperature solution can be as much as an order of magnitude In this work, two literature sources of experimental data for
higher than those in a concentrated salt system boiling at boiling point elevations were used: the work of Makarov and
atmospheric pressure or higher. It is thus crucial that the activity Krasnikov,9 containing a set of Na2CO3-Na2SO4 solubility data
coefficient calculation method account for this temperature for systems boiling at atmospheric pressure, and that of
dependence correctly. The Pitzer method addresses this problem Theliander,10 which includes the experimental boiling point rise
in two ways:5 by introducing the temperature dependence into results from research on the physicochemical data of model
the solvent properties used in the calculation of the Debye- solutions for white liquor. The data include boiling point
Hückel constant and by also adjusting the empirical parameters elevation values for NaOH, Na2CO3, Na2SO4, NaCl, Na2S2O3,
responsible for the interaction of two oppositely charged ions. Na2S, mixed NaOH-Na2CO3, and mixed Na2CO3-Na2SO4
(The parameters are usually defined as temperature-dependent solutions. The values of the experimental and modeled boiling
polynomials.) The Debye-Hückel constant is defined as point rise values are summarized in the Supporting Information.
Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008 1285

the experimental data obtained for single-salt systems. The


modeling was done in two steps: First, the temperature
extension was applied only for the Debye-Hückel constant
(with temperature-independent interaction parameters from
Pitzer5), and then, the temperature-dependent interaction coef-
ficients from Weber6 were introduced. (For the systems not
tabulated by Weber,6 only the first step was performed). The
results from both steps are shown for comparison. As can be
seen, for both approaches, the overall fit is remarkably good,
especially for moderate concentration levels (2-3 molal): the
difference between the experimental and predicted values for
Figure 1. Boiling point rise as a function of molality: NaOH solution.
moderately concentrated NaOH, Na2CO3 and Na2SO4 solutions
is approximately equal to 0.2 °C, which can be compared to
the experimental error estimated to be 0.1 °C.
At higher concentrations, the predictions deteriorate some-
what, especially in the case of the NaOH solution modeled
without including the temperature dependency in Pitzer param-
eters, for which a significant discrepancy between the experi-
mental and predicted values can be observed at m ) 9. This is,
however, not surprising given that it is very difficult to make a
proper thermodynamic description of concentrated salt solutions
at elevated temperatures. Moreover, the application of the
temperature-independent Pitzer interaction parameters for the
Figure 2. Boiling point rise as a function of molality: Na2CO3 solution.
NaOH-H2O system is actually limited to 6 molal solutions.5
The model version that uses temperature-dependent interaction
parameters performs significantly better at higher NaOH
concentrations, which can be expected based on the fact that
the temperature behavior of highly concentrated NaOH solutions
has long been studied and evaluated.
The accuracy of the prediction of boiling point elevation
increased significantly with the introduction of the temperature
dependency into the interaction parameters for only two of the
four examined single-salt systems for which the temperature
extensions were available: NaOH and Na2SO4. For the NaCl
solution, no major improvement was observed within the
Figure 3. Boiling point rise as a function of molality: Na2SO4 solution.
examined composition range, whereas for the Na2CO3 solution,
the prediction was somewhat worse. Several possible reasons
for such behavior can be mentioned: First, both NaOH and
Na2SO4 are well-known and thoroughly examined salt systems
with well-tested parameters, whereas studies of Na2CO3 solu-
tions are less common. Second, the parameters provided by
Weber6 seem to pertain mostly to saturated solutions, so the
advantage of using them might not appear clearly at lower
concentration levels, as in the case of the examined NaCl
solution.
As can be seen in Figures 1-5, the accuracy of the predictions
of boiling point rises differs between the univalent and divalent
Figure 4. Boiling point rise as a function of molality: NaCl solution.
salt systems examined: The predictions for NaOH and NaCl
tend to be slightly more exact for the same molality level than
are those for Na2CO3, Na2SO4, and Na2S2O3. This trend can be
explained by the fact that the two univalent salts are among the
most comprehensively examined systems within the scope of
solution chemistry: their interaction parameters are thoroughly
verified. Moreover, the solutions of salts with divalent anions
have ionic strengths that are 3 times higher than those of
univalent-anion systems. Thus, relating the magnitude of error
in their boiling point rise prediction to molality alone might
not be a truly fair measure of the efficiency of the prediction.
Multicomponent Solutions. The model was also used to
Figure 5. Boiling point rise as a function of molality: Na2S2O3 solution.
predict the boiling point rises in multicomponent electrolyte
solutions, resulting in a remarkably good fit, especially for
Results and Discussion
moderate concentrations. Figures 6 and 7 show the experimental
Single-Salt Solutions. Figures 1-5 show comparisons be- and predicted values for boiling point elevation plotted versus
tween the modeled and calculated boiling point rise values and ionic strength, adopted as a total concentration measure. The
1286 Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008

Figure 6. Boiling point rise as a function of ionic strength: NaOH-Na2- Figure 8. Boiling point rise as a function of molality: Na2S solution.
CO3 solution.
Thus, the single-salt solution is actually a multicomponent
system under practical conditions and needs to be modeled as
such, which involves a need for higher-order interaction
parameters. The total salt concentration also changes slightly
by the amount of water molecules involved in the hydrolysis
reaction, so the first step in modeling a Na2S-H2O solution is
to recalculate the ionic concentrations.
Three different approaches to modeling of the Na2S-H2O
system were used in this work. The first approach involved
modeling the solution as if it did not contain any hydrosulfide
ions (i.e., only Na+ and OH- from water dissociation were
Figure 7. Boiling point rise as a function of ionic strength: Na2CO3-
assumed to be present). The purpose of this approach was to
Na2SO4 solution. assess the magnitude of the influence of the hydrosulfide ions
on the system’s behavior. As can be seen in Figure 8, ignoring
NaOH-Na2CO3 system shows an excellent fit, especially when the presence of SH- ions in the solution leads to an underes-
the temperature-dependent Pitzer parameters were used. Up to timation of the boiling point rise by roughly one-half of its value;
an ionic strength of 3, the differences between the experimental this suggests that these ions play a significant role in the
values and predictions from both versions of the model are still behavior of the system.
smaller than the estimated experimental error. However, such The second approach involved modeling the multicomponent
a good fit might possibly be due to the canceling of errors in system in the usual manner, but ignoring the higher-order terms
the individual parameters of the two single salts. for interactions between the two like-charged ions and for triple-
For the Na2CO3-Na2SO4 solution data from Makarov and ion interactions (empirical parameters for these interactions were
Krasnikov,9 the total concentration of salt was always kept at not known). Because no temperature-dependence coefficients
the saturation level; only the proportion between the two salts for Na+-HS- Pitzer interaction parameters were available, the
was varied. Given the fact that the solubility of sodium carb- behavior of OH- anions was, for consistency, also modeled
onate in boiling water is slightly higher than that of sodium using only a simple temperature dependence in the form of a
sulfate, the increasing ionic strength in Figure 7 roughly modification of the Debye-Hückel constant. As can be seen
reflects the increasing proportions of carbonate ions. Also, the in Figure 8, this modeling approach offers a remarkably good
model predictions for this system compare favorably with the fit, with the absolute difference between the experimental and
experimental data. The greatest discrepancies can be observed calculated values never exceeding the experimental error. Such
at both ends of the plot, which correspond to pure-salt solutions good agreement suggests that higher-order interactions do not
at saturation. It is worth noting that, whereas use of the play any significant role in the system behavior at these
temperature-independent Pitzer parameters tends to underesti- temperature and concentration levels.
mate the boiling point rises for solutions with higher Na2SO4- The third, and final, approach consisted of modeling the
to-Na2CO3 ratios, calculations including the temperature- behavior of all of the system anions as though they were
modified parameters overestimates them. For solutions rich in hydroxide ions. In other words, the presence of SH- ions was
sodium carbonate, the model predictions behave in exactly the ignored, and the amount of OH- ions was doubled (cf. the
opposite way. stoichiometry of the hydrolysis reaction). Surprisingly good
Na2S Solution. Aqueous solutions of sodium sulfide merit agreement was obtained; again, the differences between the
special attention. When dissolved in water at 25 °C, Na2S experimental and calculated values were almost negligible and
dissociates according to the following reaction scheme11 never exceeded the level of the experimental error. Such an
excellent fit can probably be explained by similarities in the
Na2S S Na+ + NaS- (pK1 ) 7.05) behavior of OH- and SH- ions.

NaS- S Na+ + S2- (pK2 ) 19) Conclusions


A technique for the calculation of boiling point elevation in
Because of the high pK value of the second dissociation step,
single-salt and multicomponent aqueous solutions was developed
the free S2- ion is virtually never present in a solution within
in this work. This routine is based on the Pitzer ionic activity
the operationally practical pH range. A hydrolysis reaction takes
coefficient method for the calculation of water activity and on
place instead, which can be summarized as
an osmotic coefficient calculation.
Predictions obtained using the method were compared with
Na2S + H2O S 2Na+ + HS- + OH- the experimental boiling point elevation values for NaOH,
Ind. Eng. Chem. Res., Vol. 47, No. 4, 2008 1287

Na2CO3, Na2SO4, Na2S, mixed NaOH-Na2CO3, and mixed a ) anion


Na2CO3-Na2SO4 solutions. Excellent predictions were observed i ) ith solute (ion)
for moderate concentrations in both single-component and l ) liquid (solvent)
multicomponent salt solutions. At significantly higher concen- vap ) vaporization
trations (solutions close to the saturation point), the predictions
deteriorate somewhat; nevertheless, the results can still be Superscript
recommended as a good first approximation in complex boiling 0 ) properties of the pure solvent
point rise calculations. It was also noticed that the predictions
for mixed systems were relatively better than those for pure Supporting Information Available: Tables showing the
salts; this fact was especially true for the saturated Na2CO3- Pitzer parameters for ionic interactions used in this work;
Na2SO4 solution. comparison of experimental and calculated boiling point rise
The modeling results for the complex solution of hydrolyzed values for single-salt systems; comparison of the experimental
Na2S salt suggest a very weak influence of higher-order ionic and calculated boiling point rise values for multicomponent salt
interactions on the solvent activity and the boiling point rise at systems; and summary of the Pitzer method for calculation of
the concentration levels analyzed, as well as a strong similarity the osmotic coefficient (PDFs). This material is available free
between hydroxide and hydrosulfide ions. The technique of charge via the Internet at http://pubs.acs.org.
presented for estimating boiling point rises also offers a method
for calculating the empirical interaction parameters for the Pitzer Literature Cited
method for systems for which the original parameters are
(1) Meranda, D.; Furter, W. F. Elevation of the Boiling Point of Water
unknown. by Salts at Saturation: Data and Correlation. J. Chem. Eng. Data 1997,
22, 315.
Nomenclature (2) Frederick, W. J.; Sachs, D. G.; Grady, H. J; Grace, T. M. Boiling
Point Elevation and Solubility Limit for Black Liquors. Tappi J. 1980, 63
a ) water activity (4), 151.
AΦ ) Debye-Hückel constant (3) Bujanovic, B.; Cameron, J. H. Effect of Sodium Metaborate on the
Boiling Point Rise of Slash Pine Black Liquor. Ind. Eng. Chem. Res. 2001,
d0 ) solvent density, kg/m3
40, 2518.
D ) dielectric constant of water (4) Zemaitis, J. F., Jr.; Clark, D. M.; Rafal, M.; Scrivner, N. C. Handbook
e ) unit charge; e ) 1.6022 × 10-19 C of Aqueous Electrolyte Thermodynamics; American Institute of Chemical
g ) free energy term in the definition of chemical potential Engineers (AIChE): New York, 1986.
∆Gvap ) free energy change of vaporization, J/mol (5) Pitzer, K. S. ActiVity Coefficients in Electrolyte Solutions, 2nd ed.;
CRC Press: Boca Raton, FL, 1991.
∆Hvap ) enthalpy change of vaporization, J/mol
(6) Weber, C. F. Phase Equilibrium Studies of Savannah River Tanks
I ) ionic strength and Feed Streams for the Salt Waste Processing Facility; Report ORNL/
Int ) ionic interaction term for the Pitzer method TM-2001/109, Oak Ridge National Laboratory, Oak Ridge, TN, 2001.
k ) Boltzmann constant; k ) 1.3807 × 10-23 J/K (7) Castellan, G. W. Physical Chemistry; Addison-Wesley: Reading,
m ) salt molality, mol/(kg of solvent) MA, 1971.
(8) Tester, J. W; Modell, M. Thermodynamics and Its Applications, 3rd
Mw ) molecular mass of water; Mw ) 18.02 kg/kmol
ed.; Prentice Hall PTR: Upper Saddle River, NJ, 1997.
NA ) Avogadro’s number; NA ) 6.022 × 1023 (9) Makarov, S. Z.; Krasnikov, S. N. The three-component equilibrium
p ) pressure conditions at the boiling point in the four-component system Na2SO4-
R ) gas constant; R ) 8.314 J/(mol K) Na2CO3-NaCl-H2O. IzV. Sekt. Fiz.-Khim. Anal., Inst. Obshch. Neorg.
T ) temperature, K Khim., Akad. Nauk SSSR 1956, 27, 367.
(10) Theliander, H. Omrörning i fast-fas Vätska, speciellt släcknings-
Greek Letters och kausticeringsoperationerna; Chalmers University of Technology:
Göteborg, Sweden, 1988 (in Swed.).
θ ) boiling point elevation, K (°C) (11) Handbook of Chemistry and Physics 2006-2007, 87th ed.; CRC
µ ) chemical potential Press: Boca Raton, FL, 2006 (electronic version).
φ ) osmotic coefficient
ν ) stoichiometric number of ions from dissociation ReceiVed for reView April 23, 2007
ReVised manuscript receiVed October 23, 2007
Subscripts Accepted November 27, 2007
c ) cation IE070564C

Das könnte Ihnen auch gefallen