You are on page 1of 10

International Journal of Refrigeration 103 (2019) 145–154

Contents lists available at ScienceDirect

International Journal of Refrigeration


journal homepage: www.elsevier.com/locate/ijrefrig

Molecular modeling of the solubility of low global warming potential


refrigerants in polyol ester lubricants
Wael A. Fouad a,∗, Lourdes F. Vega b
a
Department of Chemical Engineering, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
b
Gas Research Center and Department of Chemical Engineering, Khalifa University, P.O. Box 127788, Abu Dhabi, United Arab Emirates

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

Article history: Pentaerythritol esters (PECs) act as the main precursors in polyol ester (POE) lubricants. The solubility of
Received 12 January 2019 low global warming potential (GWP) refrigerants in linear chained PECs was studied in this work using
Accepted 5 April 2019
the perturbed chain form of the statistical associating fluid theory (PC-SAFT). The effect of PEC alkyl tail
Available online 10 April 2019
length on the solubility of various refrigerants was investigated. Results showed that the solubility of HFO
Keywords: refrigerants is weakly dependent on PEC molecular structure. On contrary, the solubility of other natu-
Lubricant ral and paraffin refrigerants significantly increased with increasing the alkyl tail length. Within the SAFT
Refrigerant framework, pure component and binary interaction parameters were found to correlate linearly with PEC
Low GWP molecular weight. These correlations were used to predict the solubility of refrigerants in a predefined
PC-SAFT PEC blend through component lumping. Sensitivity analyses showed that the latter simplification could
POE adequately describe the mixture phase behavior. The proposed model was then applied to fully and accu-
rately predict the solubility of carbon dioxide in commercial POE lubricants. Results demonstrate that the
proposed model can be used as a fully predictive tool to determine the solubility of refrigerants in com-
mercial POE oils, provided that the lubricant molecular weight is known, a clear step forward developing
adequate low GWP refrigerants-lubricant pairs.
© 2019 Elsevier Ltd and IIR. All rights reserved.

Modélisation moléculaire de la solubilité des frigorigènes à faible potentiel de


réchauffement planétaire dans les lubrifiants à base d’esters de polyol

Mots-clés: Lubrifiant; Frigorigène; Faible GWP; PC-SAFT; POE

1. Introduction The latter described process does not just affect the level of lubri-
cation needed for compression, but also adds to the resistance for
Lubrication oils are commonly used in rotating equipment to heat transfer in the evaporator. Consequently, gas–oil miscibility is
reduce friction and excessive heating due to the moving body parts one of the important criteria in selecting the proper lubricant for
in contact. In a typical vapor compression cycle, the lubrication oil a specified process (Ananthanarayanan, 2013).
frequently escapes the compressor and is swept along with the re- Ideally, the lubricant should remain miscible with the refrig-
frigerant used in the system. In this stage, the oil mist is carried erant in the evaporator for complete oil return as well as for
by the hot discharge gas to the condenser where it mixes with liq- efficient heat transfer. Historically, mineral oils were commonly
uid refrigerant and passes through the liquid line to the evaporator. used with dichlorodifluoromethane (R-12) at low pressure systems
As heat is being removed from the media to be cooled, the liquid due to their full miscibility. However, chlorofluorocarbons (CFCs)
refrigerant vaporizes leaving behind an oil film in the evaporator. were phased out based on the Montreal Protocol of 1987 ow-
ing to their ozone depletion potential (ODP). Consequently, 1,1,1,2-
tetrafluoroethane (R-134a) was utilized as a replacement for R-12

Corresponding author. but, in return, is immiscible in mineral oils. This has necessitated
E-mail address: wael.ahmed@kfupm.edu.sa (W.A. Fouad). the rapid development of synthetic oils that are compatible with

https://doi.org/10.1016/j.ijrefrig.2019.04.004
0140-7007/© 2019 Elsevier Ltd and IIR. All rights reserved.
146 W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154

Nomenclature

A Helmholtz free energy (J)


a2 reduced Helmholtz free energy of second-order per-
turbation expansion, Eq. (9)
a3 reduced Helmholtz free energy of third-order per-
turbation expansion, Eq. (9)
NA Avogadro’s number (particles/mol)
kB Boltzmann constant (J/K)
T temperature (K)
P pressure (Pa)
m number of segments per chain
ghs average radial distribution function of hard chain
fluid Fig. 1. Base chemical structure of pentaerythritol esters considered in this work.
d temperature dependent segment diameter (Å)
ɛ depth of pair potential (J)
σ segment diameter (Å) is one of the models used to study these systems. Due to the
η packing fraction lack of information on structure and composition, lubricant oils
Z compressibility factor were treated as a pseudocomponents within the SAFT framework.
ρ total number density of molecules (1/Å3 ) Huber et al. (2002) used manufacturer density data at 15.6 °C to fit
xp fraction of polar segments per chain the model parameters and calculate an effective molecular weight
μ dipole moment (D) for six POE ISO 32/68/100 oils. The simplified SAFT equation was
kij binary interaction parameter then used to model the solubility of HFCs in these oils. Lucia and
MW molecular weight, (g/mol) Luo (2002) used gas chromatography to characterize the molecular
x mole fraction weight of two POE oils and fitted the SAFT parameters to satu-
w mass fraction rated liquid density and vapor pressure estimates calculated using
group contribution methods (Elbro et al., 1991; Nath et al., 1976).
Superscripts
The model was then used to describe the solubility of R-134a in
id ideal gas
these oils. Neto and Barbosa (Neto and Barbosa, 2014, 2009) ap-
hc hard chain
plied the PC-SAFT EoS (Gross and Sadowski, 2001; Gross and Sad-
disp dispersion
owski, 2002) to study the solubility of R-744 in POE ISO 68 and
polar dipolar
alkylbenzene (AB) ISO 32, as well as the solubility of R-600a in
POE ISO 7 and AB ISO 5 synthetic oils. The SAFT parameters for
hydrofluorocarbons (HFCs). POE oils were found to be miscible oils under consideration were fitted to experimental density data
with HFCs, tolerant to small amounts of CFCs and mix well with at atmospheric pressure.
mineral oils. Consequently, converting a system with R-12-mineral To better understand the thermodynamics of these systems and
oil pair to R-134a-POE oil pair would involve a simple straight- its connection to the structure of the molecules integrating the
forward procedure (Ananthanarayanan, 2013). Recently, the Ki- mixtures, several research groups have recently studied the solu-
gali Amendment to the Montreal Protocol, signed at year 2016, bility of refrigerants in the precursors of these commercial lubri-
has legally bound governments to phase down the production cants. Many of the commercial POE oils are based on mixtures of
and consumption of HFCs by year 2024. Hydrofluoroolefins (HFO) PEC polymers that exist in the form of linear, branch and cyclic
such as trans-1,3,3,3-tetrafluoropropene (R-1234ze(E)) and 2,3,3,3- chains. Most of the recently measured solubility data focused on
tetrafluoropropene (R-1234yf) as well as natural refrigerants such studying the effect of varying the number of CH2 repeating units
as isobutane (R-600a) and carbon dioxide (R-744) were recom- (n) found in linear and branch chained PECs (see Fig. 1). Peng–
mended as low global warming potential (GWP) and non-ozone Robinson (PR) EoS (Peng and Robinson, 1976) and Huron-Vidal
depleting alternatives for R-134a. As a result, it is critical to exam- mixing rule (Orbey and Sandler, 1995) combined with the Wil-
ine the miscibility of the currently used lubrication oils with the son excess free-energy model (Wilson, 1964) were utilized as the
newly introduced low GWP refrigerant alternatives. primary tool for data correlation in most of these experimental
Absorption properties of R-1234ze(E), R-1234yf, R-600a and R- efforts. Razzouk et al. (2007) followed a different molecular ap-
744 in commercial lubricant oils have been reported in the past proach by applying the PC-SAFT EoS to model the density of pure
few years (Neto et al., 2014; Spatz, 2009; Fujitaka et al., 2010; compressed PECs up to 45 MPa. The SAFT parameters were deter-
Bobbo et al., 2014; Lee et al., 2016; Karnaz, 2014; Leck, 2009; Bock, mined through fitting a limited number of measured vapor pres-
2015; Fukuta et al., 2005; Neto and Barbosa, 2011; Neto and Bar- sure data (below 1 Pa) as well as to saturated liquid density es-
bosa, 2010; Neto and Barbosa, 2008; Zhelezny et al., 2007; Seeton timates obtained from Tammann–Tait correlations. The developed
et al., 20 0 0; Youbi-Idrissi et al., 20 03; Tsuji et al., 20 04; Bobbo et PC-SAFT model was later used by Fandiño et al. (2010) to model
al., 2006b; Bobbo et al., 2008a). However, the exact structures of the solubility of R-744 in linear chained PECs.
these studied oils, which consists of different organic compounds, In this work, we extend our previous efforts (Fouad and Vega,
antioxidant and antiwear additives, are unknown. Therefore, it has 2017, 2018a, 2018b, 2018c) on studying the molecular interactions,
been extremely difficult to analyze the influence of the oil struc- properties and applications of low GWP refrigerants by model-
ture on the system phase behavior. In addition, the task of devel- ing their mixtures with POE oils of different viscosity grades (ISO
oping any reliable thermodynamic model to predict the phase be- 32/46/68). To do so, we followed a systematic approach by first
havior remains a challenging endeavor. modeling the solubility of R-1234ze(E), R-1234yf, R-600a, RE-170
Owing to its superiority in modeling chains of hard or soft and R-744 in linear chained PECs using the polar version of the PC-
(Lennard–Jones) segments with highly directional attractive forces, SAFT equation of state (Polar PC-SAFT) (Dominik et al., 2005). We
the SAFT equation of state (EoS) (Chapman et al., 1988, 1989, 1990) show that the fitted binary interaction parameters (kij ) exhibit a
W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154 147

linear dependency on PEC molecular weight. Ignoring the presence The universal model constants (a0j , a1j , a2j , b0j , b1j , b2j ) were fit-
of additives and assuming that a typical POE oil can be described ted to pure component data of n-alkanes and are tabulated in
as a binary or a ternary mixture of linear chained PECs that can Gross and Sadowski’s work (Gross and Sadowski, 2001).
be lumped into a single pseudo-component, molecular parameters Finally, the change in free energy due to dipolar interactions
including kij s could be determined using correlations developed in is calculated using the Jog and Chapman, (1999) and Jog et al.
this work and applied to fully predict the solubility of refrigerants (2001) theory based on the dipolar term of Fischer and coworkers
in POE oils. Unlike other previous efforts, the proposed model does (Saager et al., 1991; Saager and Fischer, 1992; Müller et al., 1996).
not require any oil density or solubility data fitting to achieve re- This term was obtained by dissolving all the bonds in a chain and
sults. then applying the u-expansion to the resulting mixture of polar
and nonpolar spherical segments.
2. Molecular theory a2
a polar = (9)
1 − a3 /a2
The Polar PC-SAFT EoS (Dominik et al., 2005) used in this work,
where a2 and a3 are the second- and third-order terms in the per-
written for pure components in terms of the reduced residual
turbation expansion. These terms have the following form
Helmholtz free energy (ares ), is given as the sum of a hard-chain
contribution (ahc ), a dispersion contribution (adisp ), and a dipolar 2π ρ μ4
contribution (apolar ): a2 = − 2
(mi x pi )2 2i I2 (10)
9 ( kB T ) di
A Aid μ6
ares = − = ahc + adisp + a polar (1) 5 ρ2
NA kB T NA kB T a3 = π2 3
(mi x pi )3 3i I3 (11)
162 ( kB T ) di
where NA denotes Avogadro’s number, kB is Boltzmann’s con-
stant and T is the temperature. For the hard chain contribution, In the aforementioned equations, I2 and I3 are the angular pair
Chapman et al. (1988) (1989) (1990) developed an equation of state and triplet correlation functions, x pi is the fraction of polar seg-
applicable for mixtures of hard-sphere chains comprising m seg- ments in the chain and μi is the dipole moment for component i.
ments. As shown by Jog and Chapman (1999); Jog et al. (2001), they are
related to the corresponding pure fluid integrals by
ahc = mi ahs − (mi − 1 ) ln ghs
i ( di ) (2)
I2 = I2 (η, mi ) (12)
where mi is the number of segments (or spheres forming the
chain) in a molecule of component i, ghs (di ) is the radial pair dis-
i I3 = I3 (η, mi ) (13)
tribution function at contact for segments of component i, di is the
hard sphere diameter of component i, and the superscripts hc and Detailed explanations of the different terms of the molecular
hs indicate quantities of the hard-chain and hard-sphere systems, theory have been provided in previous publications (Gross and
respectively. Sadowski, 2001; Dominik et al., 2005) and the reader is referred
Gross and Sadowski (2001) used the perturbation theory of to them for further details. Table 1 illustrates pure component pa-
Barker and Henderson (1967) to calculate the contribution to the rameters for the refrigerants considered in this work. These param-
free energy due to dispersion (van der Waals) interactions between eters were previously fitted to saturated liquid density and vapor
molecules using a Lennard–Jones perturbing potential. pressure data.
 ε  Lubricants considered in this work were modeled following
adisp = −2π ρ [I1 (η, mi )]m2i σi3
i
kB T the work of Razzouk et al. (2007) in which PECs were modeled
 ε 2 as chains of homonuclear nonpolar segments using the origi-
−π ρ miC1 [I2 (η, mi )]m2i σi3
i
(3) nal PC-SAFT equation of state. The pure component parameters
kB T for pentaerytritol tetrapentanoate (PEC5, n = 3), pentaerytritol
and tetraheptanoate (PEC7, n = 5) and pentaerytritol tetranonanoate
 −1 (PEC9, n = 7) were determined through fitting a limited number of
∂ Z hc
C1 = 1 + Z hc + ρ (4) measured vapor pressure data (below 1 Pa) as well as to saturated
∂ρ liquid densities obtained from Tammann–Tait correlations. The ob-
where σ i and ɛi donate the segment diameter and dispersion en- tained molecular parameters are illustrated in Table 2. In reference
ergy of component i respectively. The value of C1 represents the to Fig. 2, Razzouk et al. (2007) showed that parameters follow
compressibility of the hard chain fluid (Zhc ) and I1 (η, mi ) and I2 (η, a linear trend as a function of molecular weight, as it happens
mi ) are given by power series in density where the coefficients are for other regular chemical families. As a result, parameters for
functions of the chain length and the packing fraction, η pentaerythritol tetrabutanoate (PEC4, n = 2), pentaerythritol tetra-
hexanoate (PEC6, n = 4) and pentaerythritol tetraoctanoate (PEC8,

6
n = 6) were obtained through interpolation using correlations
I1 (η, mi ) = a j ( m i )η j (5) shown in Fig. 2.
j=0

3. Results and discussion



6
I2 (η, mi ) = b j (mi )η j
(6)
j=0 The aim of this work is to develop a molecular model that can
predict the solubility of low GWP refrigerants in POE oils with dif-
where the coefficients aj and bj depend on the chain length ac- ferent viscosity grades. In order to do so, the solubilities of HFOs,
cording to R-600a, RE-170 and R-744 are investigated and compared as a
mi − 1 mi − 1 mi − 2 function of PEC chain length. The dipolar nature of R-1234ze(E),
a j ( mi ) = a0 j + a1 j + a2 j (7) R-1234yf and RE-170 is explicitly modeled with the perturbation
mi mi mi
theory while the dipolar and quadrupole nature of PECs and R-744,
mi − 1 mi − 1 mi − 2 respectively, are ignored due steric hindrance. From a molecular
b j ( mi ) = b0 j + b1 j + b2 j (8)
mi mi mi point of view, the dipole moment is placed on one of the chain
148 W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154

Table 1
Pure component molecular parameters for refrigerants considered in this work.

Refrigerant MW (g mol−1 ) mi σ i (Å) ɛi /kB (K) x pi μi (D) T range (K) Ref.

R-1234ze(E) 114.042 2.811 3.369 176.935 0.33 1.44 170–380 (Fouad and Vega, 2018c)
R-1234yf 114.042 2.358 3.631 179.953 0.33 2.01 230–360 (Fouad and Vega, 2018c)
RE-170 46.068 2.009 3.434 215.980 0.50 1.30 20 0–40 0 (Dominik et al., 2005)
R-600a 58.122 2.262 3.757 216.530 113–407 (Gross and Sadowski, 2001)
R-744 44.01 2.0729 2.7852 169.21 (Gross and Sadowski, 2001)

Table 2 25%, as shown in Table 4. This would allow accurate model predic-
Pure component PC-SAFT molecular parameters for the PECs
tions at temperatures where experimental data does not exist.
considered in this work.
Fig. 4 depicts the isothermal solubility of R-1234yf in linear
PEC MW (g mol−1 ) mi σ i (Å) ɛi /kB (K) PECs as a function of temperature. Excellent agreement with ex-
PEC4 416.51 9.949 3.825 254.602 perimental data (Wang et al., 2016; Sun et al., 2017a) was achieved
PEC5 472.62 10.533 3.972 261.810 using a single temperature independent kij (see Table 3), while
PEC6 528.73 11.280 4.070 266.545 adding a temperature dependent binary parameter did not improve
PEC7 584.84 12.104 4.141 269.990
the accuracy (see Tables 3 and 4). Despite their structural isome-
PEC8 640.94 12.611 4.244 275.966
PEC9 697.05 12.893 4.364 284.830 try, R-1234yf containing systems exhibit a strong positive devia-
tion from Raoult’s law in comparison to R-1234ze(E) systems. It
can be deduced that placing all the fluorine atoms on one side of
the π -bond, as in the case of R-1234yf, leads to stronger repulsive
spheres as derived by Jog and Chapman (1999) Jog et al. (2001).
interactions between the HFO and PEC molecules. The latter also
As a result, the effective contribution to the Helmholtz free energy
explains the higher volatility (vapor pressure) exhibited by pure R-
due to dipolar interactions decreases with increasing the chain
1234yf in comparison to R-1234ze(E). Although R-1234yf showed
length. Therefore, we expect that the dipolar nature of the bulky
full miscibility across all temperatures covered in this work, the
PEC molecules would exhibit minimal influence on the phase be-
phase diagrams indicate that R-1234yf is more susceptible to phase
havior. Knowledge gained from modeling the binary mixtures will
split in comparison to R-1234ze(E). This is in line with experi-
then be used to develop a predictive model to determine the gas-
mental observations by Smith (2014) for the immiscibility of R-
POE oil miscibility behavior without the need for any fitting pa-
1234ze(E) and R-1234yf in POE ISO 32 lubricants.
rameters.
The isothermal solubilities of R-600a, RE-170 and R-744 in PEC4
as a function of temperature are shown in Fig. 5. The solubil-
3.1. Solubility of refrigerants in pentaerythritol esters ity of R-744 in PEC5, PEC7 and PEC9 were previously modeled
by Fandiño et al. (2010) using the PC-SAFT EoS. In this work,
Fig. 3 demonstrates the isothermal solubility of R-1234ze(E) in we extend previous efforts to binary systems containing PEC4,
linear chained PECs as a function of temperature. The system ex- PEC6 and PEC8, thanks to the correlations of the molecular pa-
hibits a challenging S-shape type phase diagram with slight neg- rameters provided in Fig. 2. A temperature independent kij al-
ative deviations from Raoult’s law at low to moderate HFO con- lowed to adequately correlate the experimental data (Sun et al.,
centrations and positive deviations at high HFO concentrations. It 2015b, 2016; Fedele et al., 2009; Pernechele et al., 2009; Sun et
can be seen that a satisfactory agreement with experimental data al., 2017b; Fandiño et al., 2008; Bobbo et al., 2008b; Sun et al.,
(Sun et al., 2015c; Wang et al., 2015; Sun et al., 2014b; Wang et al., 2015a, 2014a), and values are illustrated in Table 3. As expected,
2016) was achieved using a temperature independent kij for each the solubility of isobutane exhibits positive deviation from Raoult’s
PEC (see Table 3). In spite of the overall good agreement, devia- law indicating that the phase behavior is dominated by weak van
tions increase with increasing the PEC chain length; considering der Waal’s interactions. On the contrary, the solubility of RE-170
a temperature linear dependent kij reduces deviations by at least and R-744 exhibit negative deviation from Raoult’s law. Molecular

a b c

Fig. 2. Molecular parameters of the studied PECs as a linear function of their molecular weight. Filled circles (•) indicate fitted parameters while filled squares () indicate
interpolated parameters.
W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154 149

Fig. 3. Isothermal solubility of R-1234ze(E) in (a) PEC5, (b) PEC7 and (c) PEC9 at different temperatures. Symbols represent experimental data (Sun et al., 2015c, 2014b); (−)
molecular model with temperature independent kij . See text for details.

Table 3
Temperature independent fitted binary interaction parameters for refrigerant + PEC binary mixtures.

R-1234ze(E) R-1234yf R-600a RE-170 R-744

kij AAD (%) kij AAD (%) kij AAD (%) kij AAD (%) kij AAD (%)

PEC4 −0.00629 4.466 0.0261 1.143 0.0261 1.143 −0.016 4.678 0.064 6.378
PEC5 0.00425 4.968 0.0260 1.360 0.0260 1.360 −0.0143 4.195 0.0881 4.980
PEC6 0.0126 6.098 0.0202 2.125 0.0202 2.125 −0.0082 5.993 0.089 8.490
PEC7 0.0215 6.160 0.0207 1.651 0.0207 1.651 −0.0056 4.709 0.1225 10.751
PEC8 0.0306 6.399 0.0178 2.875 0.0178 2.875 0.0 0 04 7.684 0.1119 8.597
PEC9 0.04 7.248 −0.0074 5.495 0.1329 5.850

Table 4
Regressed coefficients for calculating the temperature dependent binary in- simulations performed by Sugii et al. (2015) on the solubility of R-
teraction parameters for HFO + PEC binary mixturesa .
744 in PEC6 indicated that R-744 molecules are preferably attached
R-1234ze(E) R-1234yf to the double-bonded oxygen atoms of the PEC ester groups. More-
A × 104 B × 102 AAD (%) A × 104 B × 102 AAD (%) over, simulation results also indicated the presence of unspecific
sorption of R-744 in the alkane tail region of the PECs. The latter
PEC4 2.179 7.669 1.544 1.714 7.583 0.994
PEC5 1.500 4.347 3.172 1.494 6.332 1.308 explains the strong negative deviations from Raoult’s law shown
PEC6 1.869 4.684 3.723 2.086 7.386 1.671 in Fig. 5(b)–c). Cross association (hydrogen bonding between re-
PEC7 1.452 2.471 4.145 1.863 6.277 2.382 frigerants and PECs) and PECs long range dipolar interactions were
PEC8 1.821 2.732 4.272 1.297 3.441 2.399 not explicitly modeled in this work. Instead, these interactions are
PEC9 2.167 2.893 4.233
implicitly incorporated into the dispersion term and the binary in-
a
ki j (T ) = A · T (K )−B. teraction parameters.

a b c

Fig. 4. Isothermal solubility of R-1234yf in (a) PEC5, (b) PEC6 and (c) PEC7 at different temperatures. Symbols represent experimental data (Wang et al., 2016; Sun et al.,
2017a); (−) molecular model.
150 W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154

a b c

Fig. 5. Isothermal solubility of (a) R-600a, (b) RE-170 and (c) R-744 in PEC4 at different temperatures. Symbols represent experimental data (Sun et al., 2015b; Pernechele et
al., 2009; Sun et al., 2017b); (−) molecular model with temperature independent kij .

a b c

Fig. 6. Effect of PEC chain length on the isothermal solubility of (a) R-1234ze(E), (b) R-600a and (c) RE-170 at 313.15 and 343.15 K. Symbols represent experimental data and
solid lines represent model calculations.

a b c

Fig. 7. Comparison of refrigerant isothermal solubility in (a) PEC5, (b) PEC6 and (b) PEC7 at 313.15 K. Symbols represent experimental data and solid lines represent model
calculations.

3.2. Relationship between solubility and chemical structure solubility of this refrigerant in linear PEC blends using any of the
binary mixture data. In contrast, the solubility of R-600a and RE-
The solubility of the refrigerants considered in this work as 170 show stronger chain length dependence and increases with in-
a function of PEC chain length is depicted in Fig. 6. It can be creasing the PEC chain length. As a result, the aim should be to
seen that the solubility of R-1234ze(E) is weakly dependent on the increase the fraction of heavier PEC components when designing
PEC chain length. Therefore, it would be safe to approximate the POE lubricants for both refrigerants. Fig. 7 compares the solubility
W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154 151

a b c d

Fig. 8. Binary interaction parameters as a linear function of PEC molecular weight. (a) R-1234ze(E), (b) R-1234yf, (c) R-600a, (d) R-744.

Table 5
Deviations by the lumped model in predicting the solubility of R-1234ze(E) in a known PEC mixture. The unlumped model
is used as a reference.

PEC blend MW (g mol−1 ) kij T = 313.15 K T = 343.15 K

AMAXD (%) AAD (%) AMAXD (%) AAD (%)

25% PEC4 + 75% PEC5 458.59 0.00117 0.427 0.214 0.359 0.186
50% PEC6 + 50% PEC5 500.68 0.00801 2.808 2.249 1.770 1.613
75% PEC9 + 25% PEC5 640.94 0.0308 2.021 0.996 1.747 0.935
25% PEC4 + 75% PEC7 542.76 0.0149 4.033 2.834 3.633 2.638
50% PEC6 + 50% PEC7 556.79 0.0171 1.309 0.970 1.188 0.906
75% PEC9 + 25% PEC7 669.00 0.0354 0.816 0.953 0.756 0.857
25% PEC4 + 75% PEC9 626.92 0.0286 4.409 2.692 3.894 2.519
50% PEC6 + 50% PEC9 612.89 0.0263 1.899 1.060 1.666 0.997
75% PEC8 + 25% PEC9 654.97 0.0331 0.744 0.665 0.673 0.605
50% PEC5 + 25% PEC7 + 25% PEC9 556.78 0.0171 4.332 2.821 3.850 2.616
25% PEC5 + 50% PEC7 + 25% PEC9 584.84 0.0217 3.277 2.193 2.927 2.046
25% PEC5 + 25% PEC7 + 50% PEC9 612.89 0.0263 3.039 1.880 2.689 1.758
50% PEC4 + 25% PEC6 + 25% PEC8 500.67 0.0080 4.799 3.150 4.289 2.919
25% PEC4 + 50% PEC6 + 25% PEC8 528.73 0.0126 3.364 2.241 3.001 2.073
25% PEC4 + 25% PEC6 + 50% PEC8 556.78 0.0171 4.287 2.903 3.825 2.687

of these refrigerants in PECs at 313.15 K. As a general trend, R- equations shown in Figs. 2 and 8 can now be used to fully predict
1234yf exhibits the least solubility in POE oils while R-600a ex- the solubility of refrigerants in POE oils.
hibits the highest solubility. Interestingly, shifting to R-1234ze(E) A sensitivity analysis was carried out to examine the applica-
(the structural isomer of R-1234yf) leads to an enhancement in the bility of using a molecular weight dependent “lumped” model to
solubility. This goes back to differences in molecular interactions predict the solubility of HFO refrigerants in a predefined multicom-
explained earlier in this work. In addition, R-600a and RE-170 ex- ponent PEC mixture (unlumped model). Due to the lack to experi-
hibit similar solubilities up to intermediate concentrates. mental data, a set of predefined binary and ternary PEC systems
were first predicted using the molecular theory (unlumped) de-
3.3. Sensitivity analysis on component lumping scribed in Section 3.1. In this case, components of the PEC mixtures
were explicitly modeled, HFO-oil binary interaction parameters
Information provided by manufacturers on the characteristics were taken from Table 3 and kij values among PEC molecules were
of a certain lubrication oil are usually limited. In most cases, the assumed to be zero. The latter was used as a reference in Tables 5
information available to the public includes the kinematic viscos- and 6 to calculate the absolute maximum deviations (AMAXD) and
ity at 40 °C, density at around 20 °C and 1 atm, pour point and the absolute average deviations (AAD) in using the lumped model
flash point. The composition analyses of these commercial oils are to predict the solubility of R-1234ze(E) and R-1234yf respectively.
usually considered proprietary information. Consequently, the de- It can be seen that deviations between both models increase with
velopment of any predictive thermodynamic model to determine increasing the number of components in the mixture and decrease
the phase behavior of oil containing systems would be very chal- with increasing the system temperature. Deviations were highest
lenging. Eychenne and Mouloungui (1998) and Niedzielski (1976) at low HFO concentrations and remained almost constant up to in-
showed that the kinematic viscosities of POE oils at 40 °C are lin- termediate concentrations before dramatically decreasing at high
early correlated to their molecular weights. Razzouk et al. (2007) concentrations; note, however, that deviations were lower than 5%
used the correlation by Eychenne and Mouloungui (1998) to es- in all cases. Hence, the analyses indicate that the proposed lumped
timate the molecular weight of several POE oils with viscosities model is expected to provide good predictions for gas solubility in
up to 32 cSt. Assuming that these commercial oils can be modeled lubricants of unknown chemical composition.
as single pseudo-components (lumped model), correlations shown
in Fig. 2 were used to calculate their respective SAFT parameters. 3.4. Solubility of refrigerants in commercial POE oils – lumped model
The lumped model was able to accurately predict the oil density predictions
as a function of temperature. In this work, we extend the lumped
model application to refrigerant + oil binary mixtures. Fig. 8 de- Bobbo et al. (2006b) performed a gas-chromatographic anal-
picts that the temperature independent kij s provided in Table 3 fol- ysis to qualitatively characterize three commercial oils, namely:
low a linear correlation with PEC molecular weight. Consequently, Castrol Icematic SW32, ICI Emkarate RL32S and Mobil EAL Arctic
152 W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154

Table 6
Deviations by the lumped model in predicting the solubility of R-1234yf in a known PEC mixture. The unlumped model is
used as a reference.

PEC blend MW (g mol−1 ) kij T = 313.15 K T = 343.15 K

AMAXD (%) AAD (%) AMAXD (%) AAD (%)

25% PEC4 + 75% PEC5 458.59 −0.0161 0.915 0.671 0.810 0.623
50% PEC6 + 50% PEC5 500..8 −0.0110 2.132 0.836 2.552 1.137
75% PEC8 + 25% PEC5 598.86 0.0 0 098 3.191 2.261 2.831 2.127
25% PEC4 + 75% PEC7 542.76 −0.00585 1.140 0.915 1.065 0.906
50% PEC6 + 50% PEC7 556.79 −0.00414 0.0191 0.105 0.0415 0.117
75% PEC8 + 25% PEC7 626.92 0.00440 1.474 1.123 1.315 1.060
25% PEC4 + 75% PEC8 584.83 −0.0 0 073 5.386 3.833 4.802 3.627
50% PEC6 + 50% PEC8 584.84 −0.0 0 073 2.804 1.997 2.487 1.875
50% PEC4 + 25% PEC6 + 25% PEC8 500.67 −0.0110 5.355 3.630 4.786 3.438
25% PEC4 + 50% PEC6 + 25% PEC8 528.73 −0.00755 4.080 2.811 3.638 2.655
25% PEC4 + 25% PEC6 + 50% PEC8 556.78 −0.00414 5.075 3.557 4.526 3.363

Fig. 9. Isothermal solubility (mass fraction) of R-744 in Castrol Icematic (a) SW 32, (b) SW 46 and (c) SW 68 commercial lubricant. Symbols represent experimental data
(Bobbo et al., 2008a) and solid lines represent the lumped model predictions.

Table 7 POE oils of different viscosity grades. The only discrepancy exists
Predicted molecular parameters for commercial POE oils considered in
in Fig. 9(c) where the model predicts immiscibility at a slightly
this work using correlations shown in Figs. 2 and 5.
lower R-744 composition. As discussed earlier, experiments show
POE MW (g mol−1 ) mi σ i (Å) ɛi /kB (K) kCO2−POE that other commercial oils may contain a wider spectrum of com-
SW 32 554 11.580 4.114 268.855 0.101 ponents, hence model accuracy maybe influenced. Still, results ex-
SW 46 593 12.042 4.176 272.195 0.110 hibit the applicability of using the proposed model as an efficient
SW 68 634 12.529 4.235 275.441 0.119 tool for miscibility calculations with the oil molecular weight being
the only required input.

32. Results showed that the three oils were composed of differ- 4. Conclusions
ent components and that Castol’s SW32 contained the least num-
ber of components in the mixture. A matrix assisted laser desorp- The solubility of low GWP refrigerants in linear chained PECs
tion ionization analysis was performed to determine the molecular was modeled using the PC-SAFT equation of state. Using temper-
weight of the four main components found in SW32 (Bobbo et al. ature independent kij s, the model was able to adequately corre-
(2006a)). Results indicated the presence of PEC6 or PEB6, PEC7 or late the VLE with the highest deviations (up to 10%) scored for
PEB7 and PEC8 or PEB8 (PEB means that the alkyl chain of the R-744 + PEC binary systems. PEC mixtures with R-1234yf showed
ester is branched). It was not possible to distinguish between the positive deviations from Raoult’s law due to the presence of all
structural isomers of similar molecular weight. Further elementary the fluorine atoms on one side of the carbon double bond. The
analysis (amount of C, H and O) was performed by Bobbo et al. latter creates a cloud of negative charges that leads to strong re-
(2008a) to determine the molecular weights of Castrol Icematic pulsive interactions between the HFO and PEC molecules. On con-
SW32, SW46 and SW68. Estimated molecular weights of the three trary, mixtures with R-1234ze(E), RE-170 and R-744 exhibited neg-
oils with different viscosity grades are reported in Table 7. Con- ative deviations from Raoult’s law due to specific association inter-
sequently, available information on the constituents and molec- actions between the refrigerant and PEC molecules. The isothermal
ular weights of these commercial oils make them an attractive solubility of HFO refrigerants was found to be weakly dependent
option for examining the predictive power of the lumped model on the PEC alkyl tail length. As a result, binary mixture data can be
developed in this work. Correlations shown in Figs. 2 and 8 were used as a good representative for the phase behavior with POE oils.
used to determine the required model parameters illustrated in The solubility of R-600a, RE-170 and R-744 was found to increase
Table 7. In reference to Fig. 9, the lumped model was able to with increasing the PEC alkyl tail length. The opposite trend was
fully and accurately predict the solubility of R-744 in commercial found for HFO refrigerants. R-600a exhibited the highest solubility
W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154 153

in PECs while R-1234yf exhibited the least solubility. Molecular pa- Fujitaka, A., Shimizu, T., Sato, S., Kawabe, Y., 2010. Application of low global warm-
rameters used within the SAFT framework were found to be a lin- ing potential refrigerants for room air conditioner. In: Proceedings of the 2010
International Symposium on Next-generation Air Conditioning and Refrigeration
ear function of PEC molecular weight. Sensitivity analyses carried Technology. Tokyo, Japan.
out in this work showed that lumping multicomponent PEC mix- Fukuta, M., Yanagisawa, T., Omura, M., Ogi, Y., 2005. Mixing and separation charac-
tures into a single pseudo-component can adequately model the teristics of isobutane with refrigeration oil. Int. J. Refrig. 28, 997–1005.
Gross, J., Sadowski, G., 2001. Perturbed-chain SAFT: an equation of state based
mixture phase behavior. As a result, correlations developed here on a perturbation theory for chain molecules. Ind. Eng. Chem. Res. 40,
were used to predict the molecular parameters needed to compute 1244–1260.
the solubility of R-744 in commercial lubricants of different viscos- Gross, J., Sadowski, G., 2002. Application of the perturbed-chain SAFT equation of
state to associating systems. Ind. Eng. Chem. Res. 41, 5510–5515.
ity grades. Good agreement with experimental data was achieved
Huber, M.L., Lemmon, E., Friend, D.G., 2002. Modeling bubble points of mixtures of
without the need any of fitting parameters. Results demonstrate hydrofluorocarbon refrigerants and polyol ester lubricants. Fluid Phase Equilib.
that the proposed model can be used as a fully predictive tool to 194, 511–519.
Jog, P.K., Chapman, W., 1999. Application of Wertheim’s thermodynamic perturba-
determine the solubility of refrigerants in commercial POE oils pro-
tion theory to dipolar hard sphere chains. Mol. Phys. 97, 307–319.
vided that the lubricant molecular weight is known. Jog, P.K., Sauer, S.G., Blaesing, J., Chapman, W.G., 2001. Application of dipolar chain
theory to the phase behavior of polar fluids and mixtures. Ind. Eng. Chem. Res.
40, 4641–4648.
Acknowledgments Karnaz, J.A., 2014. Lubricant development to meet lower GWP refrigerant challenges.
In: Proceedings of the International Refrigeration and Air Conditioning Confer-
ence. West Lafayette, Purdue University Paper 1444.
W.A. Fouad acknowledges financial support for this work from
Leck, T.J., 2009. Evaluation of HFO-1234yf as a Potential Replacement for R-134a in
King Fahd University of Petroleum and Minerals under project Refrigeration Applications. In: Proceedings of the Third IIR Conference on Ther-
SR181017. L.F. Vega acknowledges financial support for this work mophysical Properties and Transfer Processes of Refrigerants. Boulder, CO.
from Khalifa University under project CIRA-121. Lee, B.-M., Kwon, J.-W., Kim, M.-H., 2016. Miscibility of POE and PVE oils with
low-GWP refrigerant R-1234ze (E). Sci. Technol. Built Eng. 22, 1263–1269.
Lucia, A., Luo, Q., 2002. Binary refrigerant–oil phase equilibrium using the simplified
References SAFT equation. Adv. Environ. Res. 6, 123–134.
Müller, A., Winkelmann, J., Fischer, J., 1996. Simulation studies on mixtures of dipo-
Ananthanarayanan, P., 2013. Basic Refrigeration and Air Conditioning, 4th ed. Mc- lar with nonpolar linear molecules II. A mixing rule for the dipolar contribution
Graw-Hill Education, New Delhi, India. to the Helmholtz energy. Fluid Phase Equilib. 120, 107–119.
Barker, J.A., Henderson, D., 1967. Perturbation theory and equation of state for fluids. Nath, J., Das, S., Yadava, M., 1976. On the choice of acentric factor. Ind. Eng. Chem.
II. A successful theory of liquids. J. Chem. Phys. 47, 4714–4721. Fundam. 15, 223–225.
Bobbo, S., Fedele, L., Pernechele, F., Stryjek, R., 2008a. solubility of CO2 in commer- Neto, M.A.M., Barbosa Jr, J.R., 2008. Solubility, density and viscosity of a mixture of
cial POE oils with different standard viscosity. In: Proceedings of the Interna- R-600a and polyol ester oil. Int. J. Refrig. 31, 34–44.
tional Refrigeration and Air Conditioning Conference. Purdue University, West Neto, M.A.M., Barbosa Jr, J.R., 2009. Phase and volumetric behaviour of mixtures of
Lafayette Paper 988. carbon dioxide (R-744) and synthetic lubricant oils. J. Supercrit. Fluids 50, 6–12.
Bobbo, S., Fedele, L., Stryjek, R., 2006a. Oil structure influence on the solubility of Neto, M.A.M., Barbosa Jr, J.R., 2010. Solubility, density and viscosity of mixtures of
carbon dioxide in POE lubricants. In: Proceedings of the International Refriger- isobutane (R-600a) and a linear alkylbenzene lubricant oil. Fluid Phase Equilib.
ation and Air Conditioning Conference. Purdue University, West Lafayette Paper 292, 7–12.
776. Neto, M.A.M., Barbosa Jr, J.R., 2011. Absorption of isobutane (R-600a) in lubricant oil.
Bobbo, S., Pernechele, F., Fedele, L., Stryjek, R., 2008b. Solubility measurements and Chem. Eng. Sci. 66, 1906–1915.
data correlation of carbon dioxide in pentaerythritol tetrahexanoate (PEC6). J. Neto, M.A.M., Barbosa Jr, J.R., 2014. Prediction of refrigerant-lubricant viscosity using
Chem. Eng. Data 53, 2581–2585. the general PC-SAFT friction theory. Int. J. Refrig. 45, 92–99.
Bobbo, S., Scattolini, M., Camporese, R., Fedele, L., Stryjek, R., 2006b. Solubility of Neto, M.A.M., França, R.M., Barbosa Jr, J.R., 2014. Convection-driven absorption of
carbon dioxide in some commercial POE oils. In: Proceedings of the Seventh IIR R-1234yf in lubricating oil. Int. J. Refrig. 44, 151–160.
Gustav Lorentzen Conference on Natural Working Fluids. Trondheim, Norway. Niedzielski, E.L., 1976. Neopentyl polyol ester lubricants-bulk property optimization.
Bobbo, S., Zilio, C., Scattolini, M., Fedele, L., 2014. R1234yf as a substitute of R134a Ind. Eng. Chem. Prod. Res. Dev. 15, 54–58.
in automotive air conditioning. Solubility measurements in two commercial PAG Orbey, H., Sandler, S.I., 1995. On the combination of equation of state and excess
oils. Int. J. Refrig. 40, 302–308. free energy models. Fluid Phase Equilib. 111, 53–70.
Bock J.J., 2015. Vapor–liquid equilibria of a low GWP refrigerant, R-1234ze (E), Peng, D.-Y., Robinson, D.B., 1976. A new two-constant equation of state. Ind. Eng.
mixed with a POE lubricant. Doctoral Dissertation, University of Illinois at Chem. Fundam. 15, 59–64.
Urbana-Champaign. Pernechele, F., Bobbo, S., Fedele, L., Stryjek, R., 2009. Solubility of carbon dioxide in
Chapman, W.G., Gubbins, K.E., Jackson, G., Radosz, M., 1989. SAFT: equation-of-state pentaerythritol tetrabutyrate (PEC4) and comparison with other linear chained
solution model for associating fluids. Fluid Phase Equilib 52, 31–38. pentaerythritol tetraalkyl esters. Int. J. Thermophys. 30, 1144–1154.
Chapman, W.G., Gubbins, K.E., Jackson, G., Radosz, M., 1990. New reference equation Razzouk, A., Mokbel, I., García, J., Fernandez, J., Msakni, N., Jose, J., 2007. Vapor pres-
of state for associating liquids. Ind. Eng. Chem. Res. 29, 1709–1721. sure measurements in the range 10–5Pa to 1Pa of four pentaerythritol esters:
Chapman, W.G., Jackson, G., Gubbins, K.E., 1988. Phase equilibria of associating flu- density and vapor–liquid equilibria modeling of ester lubricants. Fluid Phase
ids: chain molecules with multiple bonding sites. Mol. Phys. 65, 1057–1079. Equilib. 260, 248–261.
Dominik, A., Chapman, W.G., Kleiner, M., Sadowski, G., 2005. Modeling of polar sys- Saager, B., Fischer, J., 1992. Construction and application of physically based equa-
tems with the perturbed-chain SAFT equation of state. Investigation of the per- tions of state: part II. The dipolar and quadrupolar contributions to the
formance of two polar terms. Ind. Eng. Chem. Res. 44, 6928–6938. Helmholtz energy. Fluid Phase Equilib. 72, 67–88.
Elbro, H.S., Fredenslund, A., Rasmussen, P., 1991. Group contribution method for Saager, B., Fischer, J., Neumann, M., 1991. Reaction field simulations of monatomic
the prediction of liquid densities as a function of temperature for solvents, and diatomic dipolar fluids. Mol. Simul. 6, 27–49.
oligomers, and polymers. Ind. Eng. Chem. Res. 30, 2576–2582. Seeton, C., Fahl, J., Henderson, D., 20 0 0. Solubility, viscosity, boundary lubrication
Eychenne, V., Mouloungui, Z., 1998. Relationships between structure and lubricating and miscibility of CO2 and synthetic lubricants. In: Proceedings of 4th IIRGus-
properties of neopentylpolyol esters. Ind. Eng. Chem. Res. 37, 4835–4843. tav Lorentzen Conference on Natural Working Fluids. Purdue Purdue University,
Fandiño, O., López, E.R., Lugo, L., García, J., Fernández, J., 2010. Solubility of carbon West Lafayette, pp. 446–454.
dioxide in pentaerythritol ester oils. New data and modeling using the PC-SAFT Smith, G.L., 2014. Refrigerant/lubricant properties of new low GWP options. In: Pro-
model. J. Supercrit. Fluids 55, 62–70. ceedings of the 2014 ASHRAE Annual Conference. Seattle, Washington.
Fandiño, O., López, E.R., Lugo, L., Teodorescu, M., Mainar, A.M., Fernández, J., 2008. Spatz, M., 2009. HFO-1234yf technology update–part II. In: Proceedings of the VDA
Solubility of carbon dioxide in two pentaerythritol ester oils between (283 and Winter Meeting. Saalfelden, Austria.
333) K. J. Chem. Eng. Data 53, 1854–1861. Sugii, T., Ishii, E., Müller-Plathe, F., 2015. Solubility of carbon dioxide in pentaery-
Fedele, L., Pernechele, F., Bobbo, S., Scattolini, M., Stryjek, R., 2009. Solubility of car- thritol hexanoate: molecular dynamics simulation of a refrigerant–lubricant oil
bon dioxide in pentaerythritol tetraoctanoate. Fluid Phase Equilib. 277, 55–60. system. J. Phys. Chem. B 119, 12274–12280.
Fouad, W.A., Vega, L.F., 2017. The phase and interfacial properties of azeotropic re- Sun, Y., Wang, X., Gong, N., Liu, Z., 2014a. Solubility of dimethyl ether in pentaery-
frigerants: the prediction of aneotropes from moelcular theory. Phys. Chem. thritol tetrahexanoate (PEC6) and in pentaerythritol tetraoctanoate (PEC8) be-
Chem. Phys. 19, 8977–8988. tween (283.15 and 353.15) K. J. Chem. Eng. Data 59, 3791–3797.
Fouad, W.A., Vega, L.F., 2018a. On the anomalous composition dependence of vis- Sun, Y., Wang, X., Gong, N., Liu, Z., 2014b. Solubility of trans-1, 3, 3, 3-te-
cosity and surface tension in refrigerant blends. J. Mol. Liq. 268, 190–200. trafluoroprop-1-ene (R1234ze (E)) in pentaerythritol tetrapentanoate (PEC5)
Fouad, W.A., Vega, L.F., 2018b. Transport properties of HFC and HFO based refriger- in the temperature range from 283.15 to 353.15K. Int. J. Refrig. 48, 114–
ants using an excess entropy scaling approach. J. Supercrit. Fluids 131, 106–116. 120.
Fouad, W.A., Vega, L.F., 2018c. Next generation of low global warming poten- Sun, Y., Wang, X., Gong, N., Liu, Z., 2015a. Solubility Measurement and correlation
tial refrigerants: thermodynamic properties molecular modeling. AlChE J. 64, of isobutane with two pentaerythritol tetraalkyl esters between (293.15 and
250–262. 348.15) K. J. Chem. Eng. Data 60, 1504–1509.
154 W.A. Fouad and L.F. Vega / International Journal of Refrigeration 103 (2019) 145–154

Sun, Y., Wang, X., Gong, N., Liu, Z., 2015b. Solubility of dimethyl ether in pentaery- Tsuji, T., Tanaka, S., Hiaki, T., Saito, R., 2004. Measurements of bubble point pressure
thritol tetrabutyrate and in pentaerythritol tetrapentanoate. Comparison with for CO2+ decane and CO2+ lubricating oil. Fluid Phase Equilib. 219, 87–92.
other pentaerythritol tetraalkyl esters. J. Chem. Thermodyn. 87, 23–28. Wang, X., Sun, Y., Gong, N., 2016. Experimental investigations for the phase equilib-
Sun, Y., Wang, X., Gong, N., Liu, Z., 2015c. Solubility of trans-1, 3, 3, 3-tetraflu- rium of R1234yf and R1234ze (E) with two linear chained pentaerythritol esters.
oropropene (R1234ze (E)) in pentaerythritol ester heptanoic acid (PEC7) and J. Chem. Thermodyn. 92, 66–71.
in pentaerythritol tetranonanoate (PEC9) between 283.15K and 353.15K. Fluid Wang, X., Sun, Y., Kang, K., 2015. Experimental investigation for the solubility
Phase Equilib. 387, 154–159. of R1234ze (E) in pentaerythritol tetrahexanoate and pentaerythritol tetraoc-
Sun, Y., Wang, X., Lang, H., Jin, L., Liu, Z., 2016. Phase equilibrium behavior tanoate. Fluid Phase Equilib. 400, 38–42.
for methoxymethane+ pentaerythritol tetraheptanoate and methoxymethane+ Wilson, G.M., 1964. Vapor-liquid equilibrium. XI. A new expression for the excess
pentaerythritol tetranonanoate systems. J. Chem. Eng. Data 61, 3504–3509. free energy of mixing. J. Am. Chem. Soc. 86, 127–130.
Sun, Y., Wang, X., Wang, D., Jin, L., 2017a. Measurement and correlation for phase Youbi-Idrissi, M., Bonjour, J., Terrier, M., Meunier, F., Marvillet, C., 2003. Solubility of
equilibrium of HFO1234yf with three pentaerythritol esters from 293.15K to CO2 in a synthetic oil. In: Proceedings of the Twenty-first International Congress
348.15K. J. Chem. Thermodyn. 112, 122–128. of Refrigeration. Washington, D.C. August 17-22.
Sun, Y., Wang, X., Wang, D., Jin, L., Liu, Z., 2017b. Absorption of isobutane in three Zhelezny, P.V., Zhelezny, V.P., Procenko, D.A., Ancherbak, S.N., 2007. An experimen-
linear chained pentaerythritol esters between 293.15 and 348.15K. Int. J. Refrig. tal investigation and modelling of the thermodynamic properties of isobutane—
76, 118–125. compressor oil solutions: some aspects of experimental methodology. Int. J. Re-
frig. 30, 433–445.

You might also like