Download as pdf or txt
Download as pdf or txt
You are on page 1of 3

Prediction of the Chemical Equilibrium Constant

for Peracetic Acid Formation by Hydrogen Peroxide


Milovan Janković and Snežana Sinadinović-Fišer*
Faculty of Technology, University of Novi Sad, 21000 Novi Sad, Serbia and Montenegro

ABSTRACT: An expression for temperature dependence of the model of the three-phase system has not. The parameters of the
chemical equilibrium constant for peracetic acid synthesis from mathematical models that describe the kinetics of the in situ
acetic acid and hydrogen peroxide in an aqueous solution, de- epoxidation reaction system can be classified into two groups.
rived on the basis of the van’t Hoff and Kirchhoff equations, was The first group contains parameters whose values can only be
proposed. The reverse trend of the chemical equilibrium constant determined experimentally, such as rate constants. The second
vs. temperature was apparent when the predicted values of the
comprises parameters that can be determined experimentally,
constant were compared with experimental ones taken from the
literature. However, using the proposed model to calculate the
calculated, or estimated, such as chemical equilibrium con-
chemical equilibrium constant resulted in better prediction of the stants, partition coefficients between phases, mass-transfer co-
equilibrium composition for peracetic acid synthesis at 297.5 K efficients, and interfacial areas. For application of the mathe-
than using experimental data from the literature. matical model, it is more convenient to use calculated model
Paper no. J10899 in JAOCS 82, 301–303 (April 2005). parameters than to determine them experimentally. For this rea-
son, in the present work we derived the equation for the tem-
KEY WORDS: Chemical equilibrium constant for the liquid perature dependence of the chemical equilibrium constant for
phase reaction, in situ epoxidation, mathematical model, per- the formation of peracetic acid in solution. The validity of the
acetic acid formation. proposed model was tested by comparing the values of the cal-
culated and experimentally determined equilibrium composi-
tions for peracetic acid synthesis.
The production of low-cost plasticizers for poly(vinyl chloride)
from natural and renewable sources, such as vegetable oils, is
achieved by epoxidation of the main constituent of these oils, THEORY
unsaturated TG, by percarboxylic acid. Because it is inexpen- A few two-phase models have been proposed in the literature
sive, peracetic acid is widely used as the epoxidation reagent to describe the catalyzed reaction of peracetic acid formation
and is prepared mostly by the in situ method, which is the from acetic acid and hydrogen peroxide under different in situ
safest. The in situ process involves a heterogeneous system: epoxidation conditions.
The acid-catalyzed formation of peracetic acid from acetic acid Rangarajan et al. (1) developed a two-phase model with
and hydrogen peroxide occurs in the water phase (using diluted local concentrations (C) of components in particular phases for
hydrogen peroxide), whereas the epoxidation reaction takes the in situ epoxidation of soybean oil in the presence of sulfu-
place in the oil phase, as follows (1): ric acid as the catalyst. The chemical equilibrium constant (KC)
for the formation of peracetic acid (P) from acetic acid (A) was
H+
CH 3 − COOH + H2O2 ←
→ CH 3 − COOOH + H2O [1] given for the water phase (W) by the following equation:
k+W CPW CH
W
2O
KCW = = [3]
R1 − CH = CH – R 2 + CH 3 − COOOH → R1 − CHOCH − R 2 + CH 3 − COOH [2] k−W CW W
A CH 2 O 2

The constant has been calculated using experimentally deter-


Depending on the catalyst applied for peracid formation, the mined values for the forward (k+W) and reverse (k−W) reaction
system is either two-phase (oil–water) or three-phase rate constants. The values obtained for the equilibrium constant
(oil–water–ion exchange resin). Additionally, the acid-cat- ranged between 0.7 and 5 and depended on the H2O2 concentra-
alyzed cleavage of oxirane occurs as a side reaction. tion, reactant ratio, and mineral (sulfuric) acid concentration.
Rigorous kinetic models of the three-step reaction system In the first, more rigorous kinetic model for the in situ epox-
for the epoxidation of unsaturated TG with peracid have been idation of soybean oil with peracetic acid in the presence of an
published in the last few years (1–4), although a heterogeneous ion-exchange resin catalyst (4), the Langmuir–Hinshel-
wood–Hougen–Watson approach was applied to model the re-
action of peracetic acid formation. The use of the overall con-
*To whom the correspondence should be addressed at Faculty of Technol-
ogy, Bul. Cara Lazara 1, 21000 Novi Sad, Serbia and Montenegro. E-mail: centration instead of the local phase concentrations of the reac-
ssfiser@uns.ns.ac.yu tants and the products, as well as the simultaneous calculation

Copyright © 2005 by AOCS Press 301 JAOCS, Vol. 82, no. 4 (2005)
302 M. JANKOVIĆ AND S. SINADINOVIĆ-FIŠER

of the kinetic parameters (by fitting the experimental data) in- coefficient of the component, and ξ is the reaction coordinate.
stead of determining them successively from the data of sepa- The only question related to the previous model was
rate experiments, led to a very large value for the chemical whether the UNIFAC method could be applied to determine the
equilibrium constant for peracetic acid formation (the order of activity coefficients of components in the mixtures that con-
magnitude of 109). In contrast to that suggested by the value of tained hydrogen peroxide and peracetic acid, as the data for the
KCW and a shifting of the equilibrium of peracetic acid forma- group contributions of these compounds could not be found in
tion to the right by the rapid consumption of peracetic acid in the literature (5). In the absence of specific parameters, in this
the epoxidation reaction, the reaction leading to peracetic acid work H2O2 was treated as if it consisted of two OH groups, and
formation cannot be considered irreversible. CH3COOOH was treated as if it consisted of CH3COO and OH
Musante et al. (3) provided an extensive model for peracetic groups. The interaction parameters for the OH group were de-
acid formation in the presence of an ion-exchange resin. The au- termined from the data on alcohols (5).
thors considered selective sorption and resin swelling and con- As mentioned, it is more convenient to calculate the chemi-
cluded that the component concentrations, i.e., activities (a), of cal equilibrium constant as the mathematical parameter in the
all components were different in the water (W) phase from those kinetic model than to determine its values at particular temper-
in the polymer, i.e., resin (R), phases. Accordingly, they defined atures experimentally. Therefore, this paper proposes a model
a two-phase system that consisted of a water phase with N com- to predict the chemical equilibrium constant (K) for peracetic
ponents in equilibrium with a high-viscosity liquid polymer acid formation from acetic acid and hydrogen peroxide in the
phase, which contained N + 1 components. (The N + 1st com- liquid phase.
ponent was the swollen polymer.) They used the UNIFAC (Uni- The chemical equilibrium constant (K) for the reaction in
versal Quasichemical Functional Group Activity Coefficients) the liquid phase was calculated via an integrated form of the
liquid–liquid equilibrium (LLE) method of group contributions van’t Hoff equation, using the standard enthalpy change for the
(5) to determine the activity coefficients of components in the reaction in the liquid phase. The latter was determined by inte-
water phase. The extended Flory–Huggins model (6) was ap- grating Kirchhoff’s law using Rowlinson’s equation for the
plied for the activities of the ith species in the polymer phase: heat capacity of the component (reactant or product) in the liq-
N +1 N +1
uid phase, tabulated polynomial equations for heat capacity in
ln aiR = 1 + ln ωi − ∑m ω + ∑χ ω −
ij j ij j
the ideal gas phase (7), and K at a standard temperature,
T0(KT ), determined from the following relation:
j =1 j =1 0
[4]
N +1 j −1
5  ∆GT 0,L
= −RgT0lnKT
∑∑ m 7 [8]
ik ω j ω k χ kj + ηVi  ω1R/ 3 − ω R  0 0
3 6 
j =1 k =1 where Rg is the ideal gas constant. The standard Gibbs’ free en-
where N indicates the number of species; aiR indicates the ac- ergy change for the liquid phase reaction at the temperature T0
tivity of the ith species in the resin; ωi ωj, ωk, and ωR are the (∆GT 0,L) was determined from the standard Gibbs’ free energy
0
volume fractions of the ith, jth, and kth species and of the resin of formation in the liquid state of all of the reaction compo-
in the polymer phase, respectively; mij and mik are the ratios of nents. Because Gibbs’ free energy of formation is usually given
the molar volume of the ith and jth, and ith and kth species, re- for the ideal gas state, the equation for values in the liquid state
spectively; Vi is the molar volume of ith species; χij and χkj rep- was derived using changes during transformation from the
resent the molecular interaction between components i and j, ideal gas state to the liquid state. In this paper, the Soave equa-
and k and j, respectively; and η represents the number of moles tion of state (8) was applied. To calculate the molar volume,
of active chains per unit volume. Rackett’s equation (9) was used for the saturated liquid, and
Since the standard Gibbs’ free energy of peracetic acid for- Tait’s equation (10) was used for the compressed liquid. The
mation cannot be defined with sufficient accuracy (3), the properties of the components were taken from Reid et al. (7),
chemical equilibrium constant and the interaction parameters except for hydrogen peroxide and peracetic acid data, which
of the Flory–Huggins model (χij) were determined simultane- were from the National Institute of Standards and Technology
ously by processing the experimental data. For these calcula- (11).
tions, the thermodynamic equilibrium condition for multicom- The final expression that offers a prediction model for the
pound sorption given by Equation 5, as well as the component temperature dependence (in degrees Kelvin) of the chemical
mass balance expressed by Equation 6 and the chemical equi- equilibrium constant for peracetic acid formation is as follows:
librium condition defined by Equation 7, was used:
K = exp(12.2324ln T − 0.0229913T +
aiW = aiR i = 1, 2, ..., N [5] 9.70452 × 10−6T 2 + 3045.76/T – 72.8758) [9]

niR + niW = ni,0 + λiξ i = 1, 2, ..., N [6]


RESULTS AND DISCUSSION
K = (aPRaH R
/aARaH O R)eq [7] According to Equation 9, the chemical equilibrium constants
2O 2 2
for peracetic acid formation calculated at 323, 333, and 343 K
where n indicates the number of moles, λ is the stoichiometric were 2.258, 2.091, and 1.952, respectively. Musante et al. (3)

JAOCS, Vol. 82, no. 4 (2005)


MATHEMATICAL MODEL PARAMETERS FOR EPOXIDATION 303

TABLE 1 well for this procedure. To calculate the activity coefficient, the
Comparison of Experimentally Determined and Calculated Values UNIFAC LLE model (5) was applied. The presence of sulfuric
of Peracetic Acid Equilibrium Content at 295.7 K
acid was neglected. The results, in terms of the equilibrium
Experimental Calculated content of peracetic acid, are given in Table 1. The lower SD,
Ref. 1 Ref. 3 This work
compared with those achieved when the data of Musante et al.
Chemical (3) were used, showed that using the derived model to predict
equilibrium constant 1.04a 2.91b
the chemical equilibrium constant, given by Equation 9, to-
at 295.7 K
gether with the UNIFAC LLE-calculated activity coefficients
Peracetic acid in the water phase, predicted the values of equilibrium compo-
content (30% H2O2), 8.6 5.43 9.49 sition for peracetic acid synthesis reasonably well, even when
mass% all previously mentioned simplifications were applied.
Peracetic acid
content (90% H2O2), 46.0 36.56 44.3 ACKNOWLEDGMENTS
mass%
This work is part of Project No. 1362 supported by the Ministry of
RMSDc — 7.04 1.34 Science and Environmental Protection of the Republic of Serbia,
AADd — 6.30 1.28 Serbia and Montenegro.
a
The correlation used to fit the Musante at al. (3) data is as follows: lnK =
7.02289 − 2065.575/T.
b
Value calculated from Equation 9. REFERENCES
N
1. Rangarajan, B., A. Havey, E.A. Grulke, and P.D. Culnan, Ki-
c
Root mean square deviation, RMSD = ∑( y exp
i − yicalc )2 / N
netic Parameters of a Two-Phase Model for in situ Epoxidation
i =1

N
of Soybean Oil, J. Am. Oil Chem. Soc. 72:1161–1169 (1995).
d
Average absolute deviation, AAD = 1
N ∑| y exp
i − yicalc | 2. Chou, T., and J. Chang, Acetic Acid as an Oxygen Carrier Be-
i =1 tween Two Phases for Epoxidation of Oleic Acid, Chem. Eng.
Commun. 41:253–266 (1986).
3. Musante, R.L., R.J. Grau, and M.A. Baltanás, Kinetics of Liq-
reported the values for K, determined experimentally at the uid-Phase Reactions Catalyzed by Acidic Resins: The Forma-
same temperatures, of 1.911, 2.18, and 2.778. It is apparent that tion of Peracetic Acid for Vegetable Oil Epoxidation, Appl.
the K values vs. the temperature trends of the predicted and Catal. A: General 197:165–173 (2000).
measured ones are inconsistent, i.e., in the opposite direction. 4. Sinadinović-Fišer, V.S., M. Janković, and Z.S. Petrović, Kinet-
ics of in situ Epoxidation of Soybean Oil in Bulk Catalyzed by
The validity of the model in Equation 9 for predicting the
Ion Exchange Resin, J. Am. Oil Chem. Soc. 78:725–731 (2001).
chemical equilibrium constant was tested by calculating the 5. Magnussen, T., P. Rasmussen, and A. Fredenslund, UNIFAC
equilibrium composition at 295.7 K and by comparing it with Parameter Table for Prediction of Liquid–Liquid Equilibria, Ind.
the equilibrium data reported in the literature for peracetic acid Eng. Chem. Process Des. Dev. 20:331–339 (1981).
formation by 30 and 90% hydrogen peroxide in the presence 6. Mazzotti, M., A. Kruglov, B. Neri, D. Gelosa, and M. Mor-
bidelli, A Continuous Chromatographic Reactor: SMBR, Chem.
of sulfuric acid as the catalyst (12). At the same temperature,
Eng. Sci. 51:1827–1836 (1996).
the equilibrium composition using data for K values given by 7. Reid, R.C., J.M. Prausnitz, and B.E. Poling, The Properties of
Musante et al. (3) was also calculated. For that purpose, we ex- Gases and Liquids, 4th edn., McGraw Hill, New York, 1987,
trapolated from the data of Musante et al. pp. 140, 667–668, 676–677.
Calculation of the equilibrium composition, i.e., determina- 8. Soave, G., Equilibrium Constants from a Modified Redlich–Kwong
Equation of State, Chem. Eng. Sci. 27:1197–1203 (1972).
tion of the number of moles that were reacted in Equation 1,
9. Rackett, H.G., Equation of State for Saturated Liquids, J. Chem.
was performed by solving the nonlinear equations as follows: Eng. Data 15:514–517 (1970).
10. Hankinson, R.W., and G.H. Thomson, Calculate Liquid Densi-
ties Accurately, Hydrocarbon Process. 9:277–283 (1979).
N λ  N λ λ 
K =  Π ai i  =  Π γ i i xi i  i = 1, 2, ..., N [10] 11. National Institute of Standards and Technology, Hydrogen Per-
i =1 eq i =1 eq oxide, http://webbook.nist.gov/ (accessed January 2004).
12. Greenspan, F.P., The Convenient Preparation of Per-acids, J.
Am. Chem. Soc. 68:907 (1946).
where γi is the activity coefficient and xi is the molar fraction
of the ith species. The modified Newtonian method worked [Received July 7, 2004; accepted March 24, 2005]

JAOCS, Vol. 82, no. 4 (2005)

You might also like