You are on page 1of 7

This is an open access article published under a Creative Commons Non-Commercial No

Derivative Works (CC-BY-NC-ND) Attribution License, which permits copying and


redistribution of the article, and creation of adaptations, all for non-commercial purposes.

Article

Cite This: J. Chem. Eng. Data 2019, 64, 3224−3230 pubs.acs.org/jced

Effect of Ethanol and Temperature on Partition Coefficients of Ethyl


Acetate, Isoamyl Acetate, and Isoamyl Alcohol: Instrumental and
Predictive Investigation
Ali Ammari* and Karin Schroen
Department of Agrotechnology and Food Science, Laboratory of Food Process Engineering, Wageningen University & Research,
Bornse Weilanden 9, 6708 WG Wageningen, The Netherlands
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: For alcoholic beverages such as beer, down-


stream processing for either dealcoholization or off-flavor
removal requires both quantitative data and suitable predictive
methods. Along with experimental investigations, we use a
Downloaded via 31.2.193.239 on September 19, 2020 at 09:34:08 (UTC).

method initially developed for studying the solubility of gases in


two or more miscible liquid solvents to monitor the effect of
ethanol on air−water partition coefficients of three major
flavors found in beer, namely, isoamyl alcohol, ethyl acetate,
and isoamyl acetate. In the ethanol concentration range
between 0 and 0.1 mole fraction, a slight, rather linear increase
in the Henry’s solubility coefficient was observed. This overall
behavior can be captured well using Henry coefficients for
aqueous binary and ternary systems together with the Wohl
expansion for excess Gibbs free energy coupled with the one-
parameter Margules equation. Based on the developed model, the Wohl’s expansion parameter for ethanol−water is introduced
as the solvent−solvent interaction parameter. The van ’t Hoff parameters for temperature dependence of Henry coefficients for
binary water−flavor solutions are determined in the range of 30 to 60 °C.

1. INTRODUCTION can be separated. For example, esters have higher affinity for
Volatile flavor compounds are small molecules with relatively carbon dioxide, whereas components with an alcoholic group
similar physiochemical characteristics, such as hydrophobicity can be removed through steam stripping.
or boiling points. These molecules are present at low Through stripping, it is (almost) impossible to target one
concentrations in a complex matrix present in beverages; single component due to physiochemical similarities between
therefore, separation or addition of these compounds to volatile compounds. This implies that control of beer flavor
enhance flavor profiles or develop new products is challenging. needs to take place in multistage processes. For instance, a
The market for alcohol-free beer is growing rapidly in stripping column separates the component of interest that
Western Europe, with a worldwide market projected to reach gives an off-flavor, and a consecutive stage recovers and
25 billion U.S. dollars by 2024.1 Alcohol-free beer can be recycles the “on-flavor” components that are removed with the
produced using yeasts that do not produce alcohol, which target molecule. It is evident that the effectiveness of the latter
affects the flavor composition compared to regular beer due to stage depends on the previous stages, and the interaction of
a different bioconversion. Alternatively, alcohol may be compounds needs to be charted in order to monitor the
removed from regular beer, but in the process, the flavors retention of flavors.
are expected to be removed as well, depending on their Methods for the prediction of nonideality and the phase
physiochemical interaction with the aqueous matrix. For more behaviors of simple systems comprising of liquid, gas, or vapor
information on the process details, we would like to refer to a have been reasonably well documented. For aqueous mixtures,
recent review on production processes by Brányik et al.2 comprising electrolytes and flavors, even commercial chemical
In current processes for the preparation of low-alcohol beer, engineering software can be used to predict nonideality,
two approaches are mostly used: single-stage and cost-effective although these databases use binary interaction parameters of
dealcoholization and multistage ethanol separation and flavor which it is questionable whether they are valid for such dilute
recovery followed by reconstitution of beer.2 In brewing systems or not. Furthermore, the effect of the food matrix on
industries, stripping is a popular postproduction technique for
both flavor control and dealcoholization due to its well-defined Received: November 25, 2018
and mild operational temperature and other conditions. Accepted: June 21, 2019
Depending on the stripping gas polarity, different volatiles Published: July 17, 2019

© 2019 American Chemical Society 3224 DOI: 10.1021/acs.jced.8b01125


J. Chem. Eng. Data 2019, 64, 3224−3230
Journal of Chemical & Engineering Data Article

jij G zyz
jjj zzz = A12 x1x 2 + A13x1x3 + A 23x 2x3
flavor has been limited to sensory evaluation. Data on E

k RT {123
quantitative analytical methods are very scarce and limited in
scope.
In the current work, we quantify the effect of ethanol and + (β0 + β1x1 + β2x 2)x1x 2x3 (6)
temperature on the partitioning of three major flavors found in
typical pilsner beer produced in the Netherlands and Belgium where x is the mole fraction of components, and β0, β1, and β2
(light ale). When considering a gas−liquid system at are adjustable parameters.6 As indicated in the literature,7 this
equilibrium, the ratio of the concentration of flavors in the approach does not lead to an invariance problem for a system
gas phase and liquid phase is constant. This constant is called consisting of water, ethanol, and ethyl acetate (one of our
the Henry’s law constant (HLC) and can be defined in either selected flavors), and we expect that given the low flavor
solubility or volatility terms concentration used, this will hold for all flavors under
investigation.
1 C The Gibbs energy and hence chemical potential are defined
K cc = = G
H cc CL (1) in relation to internal energy and entropy for which absolute
values are unknown but can be approached using the fugacity
where K and H are volatility and solubility coefficients, concept.8 For a system at the equilibrium condition, the
respectively, superscript cc indicates a dimensionless value fugacity in the liquid phase fLî is defined as
ij yz
fi ̂ = xiγifi 0 expjjj dpT zzz
based on concentrations, and CG and CL are the concentrations pT

j z
of the flavor in the gas and vapor phases, respectively. Binary viL
HLC has been tabulated for various components,3 and the
database is still expanding. By definition, in a binary water−
L

k
∫ pR
RT { (7)
flavor system at equilibrium, the chemical potential (μ) of R
where p is the reference pressure, f i is the liquid reference 0
flavor i is equal in both phases. fugacity of species i, and v is the partial molar volume of the
species in the solution. Since water (1) is present in the largest
μiG = μi L (2) quantity followed by ethanol (3) as a secondary solvent, the
symmetric standard state convention is assumed to be valid to
and by definition find their activity coefficients. This means that fugacity for
ij yp yz
μi 0 (p , T ) + RT lnjjjj i sT zzzz = μi 0 (p , T ) + RT ln(γixi)
solvents in their pure state is taken as a reference, complying

j p z
with Raoult’s law. The flavor (2) is referenced to the state of

k i {
“infinite dilution” therewith complying with Henry’s law9
(3) through an unsymmetric fugacity referencing method. For our

ij y
three components, eq 7 now becomes

v1dP zzz
where p, pT, and ps are the given, total, and saturated pressures,
j
expjjjj− zz
L

RT zz
respectively, R is the gas constant, T is the temperature, x and y f1̂ pT

j ∫
k {
are mole fractions in the liquid and gas phases, respectively, γ1 ≡
and γ is the activity coefficient. The partition coefficient based x 2f10 p1s (8)

ij y
v2dP zzz
on mole fraction can be defined as follows
j
expjjjj− zz
L

RT zz
f2̂ pT

j
pis γi
yi

k {
yx
Ki = = γ2* ≡ cc
xi pT x 2H21 p1s
(4) (9)

ij y
j v3dP zzz
expjjjj− zz
L

RT zz
and the Henry’s solubility coefficient (Hcc) as follows f3̂ pT

j
jij xi × ρsol zyz jij ρsol zyz

k {
γ3 ≡
j ∑ xj × Mwj z j ∑ xj × Mwj z
x3f 30

= k p ×y { = k {
p1s (10)
C Li
Hicc = pis γi
The asterisk denotes that the activity coefficient is normalized
CGi T i
using the unsymmetric convention. The reference pressure for
RT RT (5)
all components is taken as equal to the dominant solution
where ρ is the solution density, and Mw is the molar mass of vapor pressure, which is that of water. For solute molecules,
compounds. Equation 5 will be used in this work to derive O’Connell−Prausnitz10 showed that
HLC from activity coefficients available in the literature and
γ2* = γ2 × exp( −A12 ) (11)
databases.
To determine activity coefficients, excess Gibbs free energy and
is one of the useful approaches. Wohl presented a general cc
expression for the Gibbs free energy based on a power series H23
A 23 = A12 + cc
expansion of the effective volume fractions of the solution H21 (12)
combinations.4 Several methods have been developed based on
simplified Wohl expansion, such as Wilson, van Laar, Since for flavor compounds, by definition, G /RT = ∑i[xi ln E

UNIQUAC, Margules, etc., in which interaction parameters (γi)], integration of eq 6 gives


(A) of components i and j are used. In this work, Wohl’s ln γ2* = A12 [x1(1 − x 2) − 1] + A 23x3(1 − x 2) − A13x1x3
expansion is coupled with the one-parameter (two-suffix)
Margules equation for a ternary system of water (1), flavor (2), (13)
and ethanol (3) and defined as follows5 By assuming x2 = 0, substitution of eq 12 in eq 13 yields
3225 DOI: 10.1021/acs.jced.8b01125
J. Chem. Eng. Data 2019, 64, 3224−3230
Journal of Chemical & Engineering Data Article

Table 1. Chemicals and Their Properties


chemical name CAS purity (GC) molecular wt (g/mol) boiling point (°C) log Pa supplier
ethanol (SeccoSolv) 64-17-5 ≥99.9 46.07 78.37 −0.31 Merck
ethyl acetate (Chromasolv) 141-78-6 ≥99.7 88.11 77.1 0.73 Sigma-Aldrich
acetone (SupraSolv) 67-64-1 ≥99.8 58.08 56 −0.24 Merck
isoamyl alcohol (Emsure) 123-51-3 ≥99 88.148 131 1.42 Merck
isoamyl acetate (Emplura) 123-92-2 ≥99 130.19 141 2.25 Merck
water 18.013 100 Milli Q-Plus system
a
Experimental data.11

Figure 1. Effect of ethanol on HLC of (a) ethyl acetate, (b) isoamyl acetate, and (c) isoamyl alcohol regressed by eq 15 using experimental data.
(d) Ethanol−water interaction parameter A13 as a function of ethanol mole fraction. Markers are experimental data, and the values beneath them
are concentrations as volume percentage.
cc
H23 is based on the fact that the partition coefficient of a volatile
lim ln γ2* = x3 ln cc − A13x1x3 compound in a solution is not a function of the volume of the
x2 → 0 H21 (14)
solution. However, a larger volume of a solution makes
By recalling the definition of the activity coefficient of the volatiles more concentrated in the headspace. This change in
flavor in the solution (eq 9) and considering Hcc2M = f 2/x2 in a concentration can be detected by various headspace analyses.
dilute system, HLC of the flavor Hcc 2M can be found using the By using a gas chromatography method, the change in the
following equation reciprocal value of peak areas against the ratio of the vial total
cc x1 cc x3
volume over the liquid phase volume becomes a linear plot
H2ccM = (H21) × (H23) × exp( −A13x1x3) (15) with a slope of a′ and intercept of b′.
As described by Kolb and Ettre,12 HLC can be determined
Equation 15 suggests that the HLC of a flavor is a function
through
of the mole fraction of both water and ethanol as well as
solvent−solvent interaction (A13), which in itself is a function a′
of ethanol concentration. H cc =
b′ (16)
2. MATERIAL AND METHODS Therefore, in this work, various sample volumes of 0.1, 0.2,
2.1. Chemicals. All chemical are listed in Table 1 and used 0.5, 1, 2, and 5 mL were transferred to standard 20 mL
as supplied without further purification. headspace screw-neck vials supplied by VWR and were
2.2. Static Headspace Analysis. All HLC were incubated for 60 min. After that, 1 mL of the headspace
determined using phase ratio variation (PRV). This method sample was taken by a CombiPAL autosampler equipped with
3226 DOI: 10.1021/acs.jced.8b01125
J. Chem. Eng. Data 2019, 64, 3224−3230
Journal of Chemical & Engineering Data Article

Table 2. Parameters for Eq 15 at T = 30 °Ca


experimental values predicted data
i ii iv v, iii v, iii
component xflavor Hcc
21 Hcc
23 γ21iii γ23iii Hcc
21 Hcc
21 Hcc
23 γ21v γ23v
ethyl acetate 9.182 × 10−05 (1.836 × 10−06) 100.49 (5.53) 965.3 63.7 6.6 107.3 128.4 3197.5 66.4 2.66
isoamyl acetate 6.015 × 10−05 (1.203 × 10−06) 36.10 (1.79) 1181.3 2578.8 76.4 34.4 62.4 35453.7 1945.8 3.42
isoamyl alcohol 8.269 × 10−05 (1.654 × 10−06) 1125.31 (12.01) 8237.6 147.6 20.1 1142.2 1722.7 192787.7 129.2 1.15
a
The parameters were determined using (i) experimental data, (ii) extrapolated eq 5, (iii) eq 5, (iv) the literature,20 and (v) UNIFAC (Dortmund)
standard uncertainty values for experimental points, u(Hcc
21).

Figure 2. Binary HLC for (a) ethyl acetate, (b) isoamyl acetate, and (c) isoamyl alcohol in water as a function of temperature.

a Hamilton-Gastight 1002 syringe and was injected into the Henry solubility coefficients of flavor compounds with
same gas chromatograph (GC) mentioned earlier. The syringe corresponding regression lines and ethanol−water interaction
was heated 10 °C above the injection temperature to avoid parameters A13 (using eq 15). In the range of 0−0.05 mole
condensation of vapor. To create the −100 °C trap required fractions, the effect of ethanol on retention of flavors is minor,
for analysis, the GC was coupled with a CryoFocus-4 cold trap. which was in accordance with Conner et al.13 who reported
The initial temperature of the GC oven was kept at 40 °C for that activity coefficients for esters were not affected
30 s and was increased to maximum 160 °C with a significantly by ethanol concentrations below 17% (v/v) ≈
temperature-increase rate of 10 °C/s. We used DB-WAXetr, (0.0622 mole fraction), whereas they decreased at higher
a high-polarity polyethylene glycol (PEG) column from ethanol concentrations, which means a higher Henry’s
Agilent with a flame ionization detector (FID). solubility constant (Hcc) in our case. They attribute this to
All flavor concentrations in this study were produced by the formation of ethanol clusters that reduce hydrophobic
mixing 0.5 mL of flavor, taken using a 1 mL (±1%) GSM gas interactions, leading to partitioning into these ethanol-rich
tight syringe, with water in a 1000 mL (±0.4) volumetric flask, clusters.14−18
creating 0.05 vol %. HLC was determined at 30, 40, 50, and 60 The value of the interaction parameter A13 showed a fairly
°C for the binary flavor−water systems. linear increase with ethanol volume fraction. Please note that
for all components, attention must be given due to the fact that
the Hcc23 values are extrapolated (ultimately to a volume fraction
3. RESULTS AND DISCUSSION of 1 for ethanol) as illustrated in the appendix (Figure S1).
3.1. Effect of Ethanol on Henry Solubility Coefficients This co-determines the slope of A13; however, because we did
of Flavors. Figure 1 is constructed by experimental data and not see any significant difference in the value of A13 when
the proposed model (eq 15) with parameters presented in comparing flavors, we expect that these values are reliable. It is
Table 2. It illustrates the effect of ethanol concentration on important to note that applying binary parameters for
3227 DOI: 10.1021/acs.jced.8b01125
J. Chem. Eng. Data 2019, 64, 3224−3230
Journal of Chemical & Engineering Data Article

Table 3. Experimental HLC Data Using Headspace Analysis Binary Water−Flavor at Different Temperaturesa
Hcc(−)
component xflavor 30 °C 40 °C 50 °C 60 °C
−05 −06
ethyl acetate 9.182 × 10 (1.836 × 10 ) 100.49 (12.12) 62.98 (8.84) 40.89 (5.10) 26.55 (6.03)
isoamyl acetate 6.015 × 10−05 (1.203 × 10−06) 36.097 (5.79) 20.52 (2.68) 10.77 (1.26) 5.65 (1.09)
isoamyl alcohol 8.269 × 10−05 (1.654 × 10−06) 1125.31 (412.01) 593.56 (117.95) 308.20 (38.63) 160.03 (39.52)
a
The standard uncertainties for the temperature of the incubator u(T) = 0.5 °C and u(Hcc) are given in the table.

multicomponent systems by extending usual quadratic mixing where Hcc⊖ is the HLC at the reference temperature T⊖, ΔsolH
rules in equations of state to higher-order polynomials is the enthalpy of dissolution, and R is the universal gas
potentially suffers from the fact that these models are not constant. The negative or positive value in the exponent
invariant when a component is divided into two or more depends on how the HLC is defined. When applying nonlinear
identical subcomponents.19 There are modified mixing rules regression to the experimental data (Figure 2) using the least-
for multicomponent systems to overcome this problem.7 In squares method, the values tabulated in Table 4 are obtained.
this work, composition A13 serves not only as a so-called
solvent−solvent interaction parameter in the model but also Table 4. Henry’s Law Constants at the Reference
carries structural characteristics of an ethanol−water mixture. Temperature (25 °C) for the Binary Water−Flavor System
Even though we can argue about its degree of linearity, we Regressed from Gas−Liquid Equilibrium Data Given in
believe that this parameter is immune to the abovementioned Table 3
complications because of two main reasons. First, flavor-matrix
this work literature
systems are so dilute that they hardly have any interactions
with the solvents and each other; therefore, their mixtures can Δsol H Δsol H
be considered as ternary. Second, dissimilarity between the two component Hcc⊖ R Hcc⊖ R reference
solvents (water and ethanol) and solute (flavor) is large ethyl acetate 123 4500 146 5900 21
enough that it does not fall in the category of what is called 154 5500 28
“identical subcomponents” addressed by Michelsen and 89 4800 20
Kistenmacher.19 isoamyl acetate 49 5500 45 5000 25
3.2. Temperature Dependence of Henry’s Law 64 5000 24
Constant. Temperature is known to significantly alter the 64 5000 20
HLC particularly for those components with a low enthalpy of isoamyl alcohol 1628 7000 1710 7600 26
dissolution. Figure 2 shows HLC of the flavor compounds at 1834 8200 27
four experimental temperatures, 30, 40, 50, and 60 °C, 1140 7600 20
together with their corresponding regressed and predicted
data. As also shown in Table 3, the HLC has a higher standard
deviation at low temperatures, and this may originate from the To investigate the effect of experimental uncertainties on the
relatively low solubility of the flavors, which influences uncertainty of driven parameters using eq 17 the concept of
the propagation of uncertainties29 is applied and presented in
equilibrium quality. Figure 2a compares our experimental
Table 5.
data with that of Kutsuna et al.21and Fenclová et al.,22 both
using a column-stripping method. The method of Hilal et al.,20 4. CONCLUSIONS
which uses SPARC (SPARC Performs Automated Reasoning
in Chemistry)23 vapor pressure coupled with activity The applied predictive method allows us to describe air−water
coefficient models, relatively underestimates the HLC of partition coefficients of flavors in ethanoic solutions. The
ethyl acetate in water; however, it is in good agreement with retention of isoamyl acetate and ethyl acetate increased slightly
the prediction method of Mackay et al.,24 which is based on with increasing amount of ethanol, whereas these effects were
the ratio of vapor pressure over the solubility of isoamyl acetate
in water. Our data, especially at higher temperatures, is in Table 5. The Uncertainty of Propagation for Driven Data
better agreement with Meylan and Howard’s25 predictive (see Appendix B)
methods, which are based on bond contribution values.
In the absence of experimental data, we compared three Δsol H

different predictive methods: those of Nirmalakhandan and


component temp [°C] W (Hcc⊖) W ( R )
30 6.09 1057.73
Speece26 using quantitative structure−activity relationship 40 18.46 883.73
(QSAR), Kühne et al.27 using their novel model based on ethyl acetate
50 16.64 487.05
two-dimensional structure for organic compounds, and Hilal et 60 29.56 647.02
al.,20 which has already been mentioned above. 30 2.11 1048.24
Classically, the temperature dependency of the Henry’s law 40 6.63 830.74
constant is described by the approaching van ’t Hoff isoamyl acetate
50 5.38 463.49
introduced for equilibrium constants. The definition of Hcc 60 7.63 552.75
was as follows

i Δ Hi1 1 yy
30 45.91 684.35

H cc(T ) = H cc ⊖ × expjjjj( −) sol jjj − ⊖ zzzzzzz


40 368.98 1252.84

k R kT T {{
isoamyl alcohol
50 245.90 497.34
(17) 60 469.40 705.89

3228 DOI: 10.1021/acs.jced.8b01125


J. Chem. Eng. Data 2019, 64, 3224−3230
Journal of Chemical & Engineering Data Article

much stronger for isoamyl alcohol, which eventually were (11) Souza, E. S.; Zaramello, L.; Kuhnen, C. A.; Junkes, B. D. S.;
completely retained in ethanol. We found that the ethanol Yunes, R. A.; Heinzen, V. E. F. Estimating the Octanol/Water
concentration dependency of parameter A13 plays a pivotal role Partition Coefficient for Aliphatic Organic Compounds Using Semi-
in describing Henry’s law constants and found a similar Empirical Electrotopological Index. Int. J. Mol. Sci. 2011, 12, 7250−
dependency for the flavors tested, which may help translate our 7264.
findings to those of other flavors. The observed dependency of (12) Kolb, B.; Ettre, L. S. Static Headspace-Gas Chromatography:
A13 may originate from the structural changes reported by Theory and Practice; John Wiley & Sons: 2006, 308−311.
others.13−17 As expected, temperature affected the partition (13) Conner, J. M.; Birkmyre, L.; Paterson, A.; Piggott, J. R.
coefficient of flavors, which we successfully covered through a Headspace Concentrations of Ethyl Esters at Different Alcoholic
Strengths. J. Sci. Food Agric. 1998, 77, 121−126.
van ’t Hoff approach.


(14) Cipiciani, A.; Onori, G.; Savelli, G. Structural Properties of
Water-Ethanol Mixtures: A Correlation with the Formation of
ASSOCIATED CONTENT Micellar Aggregates. Chem. Phys. Lett. 1988, 143, 505−509.
*
S Supporting Information (15) D’Angelo, M.; Onori, G.; Santucci, A. Self-Association of
The Supporting Information is available free of charge on the Monohydric Alcohols in Water: Compressibility and Infrared
ACS Publications website at DOI: 10.1021/acs.jced.8b01125. Absorption Measurements. J. Chem. Phys. 1994, 100, 3107−3113.
(16) Price, W. S.; Ide, H.; Arata, Y. Solution Dynamics in Aqueous
Henry’s solubility coefficient of flavors (ethyl acetate, Monohydric Alcohol Systems. J. Phys. Chem. A 2003, 107, 4784−
isoamyl acetate, and isoamyl alcohol) graphed and 4789.
tabulated for different water−ethanol concentrations; (17) Asenbaum, A.; Pruner, C.; Wilhelm, E.; Mijakovic, M.; Zoranic,
the propagation of uncertainties for the regressed L.; Sokolic, F.; Kezic, B.; Perera, A. Structural Changes in Ethanol-
parameters used in temperature dependency of HLC Water Mixtures: Ultrasonics, Brillouin Scattering and Molecular
(PDF) Dynamics Studies. Vib. Spectrosc. 2012, 60, 102−106.


(18) Gereben, O.; Pusztai, L. Hydrogen Bond Connectivities in
AUTHOR INFORMATION Water-Ethanol Mixtures: On the Influence of the H-Bond Definition.
J. Mol. Liq. 2016, 220, 836−841.
Corresponding Author (19) Michelsen, M. L.; Kistenmacher, H. On Composition-
*E-mail: ali.ammari@wur.nl. Phone: +31(0)317 483396. Dependent Interaction Coefficeints. Fluid Phase Equilib. 1990, 58,
ORCID 229−230.
Ali Ammari: 0000-0003-2557-0037 (20) Hilal, S. H.; Ayyampalayam, S. N.; Carreira, L. A. Air−Liquid
Partition Coefficient for a Diverse Set of Organic Compounds:
Notes
Henry’s Law Constant in Water and Hexadecane. Environ. Sci.
The authors declare no competing financial interest.


Technol. 2008, 42, 9231−9236.
(21) Kutsuna, S.; Chen, L.; Abe, T.; Mizukado, J.; Uchimaru, T.;
ACKNOWLEDGMENTS Tokuhashi, K.; Sekiya, A. Henry’s Law Constants of 2,2,2-
This project was funded by the Institute for Sustainable Trifluoroethyl Formate, Ethyl Trifluoroacetate, and Non-Fluorinated
Processes (ISPT) in the Netherlands. Analogous Esters. Atmos. Environ. 2005, 5884.


(22) Fenclová, D.; Blahut, A.; Vrbka, P.; Dohnal, V.; Böhme, A.
REFERENCES Temperature Dependence of Limiting Activity Coefficients, Henry’s
Law Constants, and Related Infinite Dilution Properties of C4-C6
(1) Sellbyville, D. Worldwide Non-alcoholic Beer Market worth over $
25 billion by 2024: Global Market Insights, Inc. https://globenewswire. Isomeric n-Alkyl Ethanoates/Ethyl n-Alkanoates in Water. Measure-
com/news-release/2018/03/20/1442488/0/en/Worldwide-Non- ment, Critical Compilation, Correlation, and Recommended data.
alcoholic-Beer-Market-worth-over-25-billion-by-2024-Global-Market- Fluid Phase Equilib. 2014, 375, 347−359.
Insights-Inc.html (accessed Sep 16, 2018). (23) Hilal, S.; Karickhoff, S. Verification and Validation of the SPARC
(2) Brányik, T.; Silva, D. P.; Baszczyňski, M.; Lehnert, R.; Almeida E Model; US Environmental Protection Agency, 2003, No. March, 44.
Silva, J. B. A Review of Methods of Low Alcohol and Alcohol-Free (24) Mackay, D.; Shiu, W.-Y.; Ma, K.; Lee, S. C. Handbook of
Beer Production. J. Food Eng. 2012, 108, 493−506. Physical-Chemical Properties and Environmental Fate for Organic
(3) Sander, R. Compilation of Henry’s Law Constants (Version 4.0) Chemicals; 2nd Edition, Introduction and Hydrocarbons, CRC
for Water as Solvent. Atmos. Chem. Phys. 2015, 15, 4399−4981. press: 2006; Vol. 3.
(4) Wohl, K. Thermodynamic Evaluation of Binary and Ternary (25) Meylan, W. M.; Howard, P. H. Bond Contribution Method for
Liquid Systems. Trans. Am. Inst. Chem. Eng. 1946, 42, 215−249. Estimating Henry’s Law constants. Environ. Toxicol. Chem. 1991, 10,
(5) Villamañań , R. M.; Martín, M. C.; Villamañań , M. A.; Chamorro, 1283−1293.
C. R.; Segovia, J. J. Thermodynamic Properties of Binary and Ternary (26) Nirmalakhandan, N. N.; Speece, R. E. QSAR Model for
Mixtures Containing Di-Isopropyl Ether, 2-Propanol, and Benzene Predicting Henry’s Constant. Environ. Sci. Technol. 1988, 22, 1349−
atT= 313.15 K. J. Chem. Eng. Data 2010, 55, 2741−2745. 1357.
(6) Jackson, S. L. Extension of Wohl’s Ternary Asymmetric Solution (27) Kühne, R.; Ebert, R.-U.; Schüürmann, G. Prediction of the
Model to Four and n Components. Am. Mineral. 1989, 74, 14−17.
Temperature Dependency of Henry’s Law Constant from Chemical
(7) Mathias, P. M.; Klotz, H. C.; Prausnitz, J. M. Equation-of-State
Structure. Environ. Sci. Technol. 2005, 39, 6705−6711.
Mixing Rules for Multicomponent mixtures: The Problem of
(28) Fenclová, D.; Dohnal, V.; Vrbka, P.; Laštovka, V. Temperature
invariance. Fluid Phase Equilib. 1991, 67, 31−44.
(8) Smith, J. M. Introduction to Chemical Engineering Thermody- Dependence of Limiting Activity Coefficients, Henry’s Law
namics. J. Chem. Educ. 1950, 584. Constants, and Related Infinite Dilution Properties of Branched
(9) Brelvi, S. W.; O’Connell, J. P. Prediction of Unsymmetric (C3 and C4) Alkanols in Water. Measurement, Critical Compilation,
Convention Liquid-phase Activity Coefficients of Hydrogen and Correlation, and Recommended Data. J. Chem. Eng. Data 2007, 52,
Methane. AIChE J. 1975, 21, 157−160. 989−1002.
(10) O’Connell, J. P.; Prausnitz, J. M. Thermodynamics of Gas (29) Kline, S. J.; Mcclintock, F. A. Describing Uncertainties in a
Solubility in Mixed Solvents. Ind. Eng. Chem. Fund. 1964, 3, 347−351. Single Sample Experiment. Mech.Eng. 1953, 75, 3−8.

3229 DOI: 10.1021/acs.jced.8b01125


J. Chem. Eng. Data 2019, 64, 3224−3230
Journal of Chemical & Engineering Data


Article

NOTE ADDED AFTER ASAP PUBLICATION


This paper was published ASAP on July 17, 2019, with
incorrect Supporting Information. The corrected version was
reposted on July 19, 2019.

3230 DOI: 10.1021/acs.jced.8b01125


J. Chem. Eng. Data 2019, 64, 3224−3230

You might also like