You are on page 1of 7

J. Chem.

Thermodynamics 128 (2019) 127–133

Contents lists available at ScienceDirect

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

Heat capacities of potential organic phase change materials


John A. Noël a, Mary Anne White a,b,⇑
a
Department of Chemistry, Dalhousie University, Halifax, Nova Scotia B3H 4R2, Canada
b
Clean Technologies Research Institute, Dalhousie University, Halifax, Nova Scotia B3H 4R2, Canada

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

Article history: The heat capacity of a material is an important factor for many applications, including phase change
Received 11 April 2018 material (PCM) selection for thermal energy storage. Aside from storing energy via the latent heat of a
Received in revised form 7 August 2018 phase transition, significant sensible heat can be stored using the material’s heat capacity outside the
Accepted 10 August 2018
transition temperature region. Furthermore, accurate heat capacity values are required to model and
Available online 12 August 2018
thereby optimize PCM heat storage systems. Various group additivity models can be used to estimate
the heat capacity of organic liquids. However, these methods do not distinguish isomers built of the same
Keywords:
base units. In the present study the liquid phase heat capacities of five isomeric, 12-carbon, linear,
Heat capacity
Phase change material
saturated fatty esters (potential PCMs) have been measured experimentally, and the results are compared
Ester to calculations from several group additivity methods. It was found that the heat capacities of the liquid
Group additivity esters are well described by the group additivity models, and that each of the esters had approximately
the same heat capacity at a given temperature, despite their structural differences. The heat capacity of
another isomer, dodecanoic acid, was also compared to the ester isomers, and found to be significantly
different. The group additivity methods also were able to accurately represent the heat capacity of the
acid isomer.
Ó 2018 Elsevier Ltd.

1. Introduction the enthalpy change of the phase transition is not accessed, and
much of the thermal storage capacity is lost [1]. Among potential
Phase change materials (PCMs) offer considerable promise for PCMs with suitable transition temperatures, the best material
storage of renewable energy via thermal energy storage and tem- can then be selected to maximize the volumetric or gravimetric
perature moderation [1]. PCMs store thermal energy isothermally thermal storage. Usually this means choosing a material with a
by utilizing the latent heat of a phase transition, typically melting high latent heat of fusion. The thermal energy, Q, stored over the
[1–3], but solid-solid transitions are also possible [4]. Coupled with temperature range of the application is given by:
a solar thermal collector, PCMs can store heat for use when the sun
Z T fus Z T2
is not available, for applications such as heating hot water or build-
Q¼ mcp;s ðT ÞdT þ mDfus H þ mcp;l ðT ÞdT ð1Þ
ings [1,5]. Alternatively, PCMs can be encapsulated within walls,
floors, or ceilings to increase the ‘thermal inertia’ of a room. In such T1 T fus

a situation, the PCM absorbs heat during the day and keeps the
room cooler for longer, thus acting as a thermal buffer and reduc- where T1 is the minimum temperature of the application, T2 is the
ing cooling energy requirements. The PCM then releases its stored maximum temperature of the application, Tfus is the melting tem-
energy at night when the temperature has dropped, with the over- perature (T1 < Tfus < T2) , m is the mass of the PCM, DfusH is the melt-
all effect of reducing temperature swings [1,6]. ing enthalpy change, and cp,s and cp,l are the specific heat capacities
While cost, safety, and long-term stability of PCMs are impor- of the solid and liquid phases. From Eq. (1), it can be seen that over a
tant concerns in material selection, the paramount consideration sufficiently broad temperature range, the heat capacity of the PCM
is the transition temperature. If the PCM does not undergo its can contribute significantly to the total energy stored [1]. As such,
phase transition within the temperature range of its application, the selection of a PCM with a high heat capacity, as well as a high
DfusH, can maximize the storage capacity of the system.
For good PCM selection, there must be accurate thermodynamic
⇑ Corresponding author at: Department of Chemistry, Dalhousie University, data on which the choice can be based. Accurate simulations of
Halifax, Nova Scotia B3H 4R2, Canada. PCM systems to optimize performance also rely on accurate ther-
E-mail address: mawhite@dal.ca (M.A. White). modynamic property inputs. There are many PCMs for which the

https://doi.org/10.1016/j.jct.2018.08.014
0021-9614/Ó 2018 Elsevier Ltd.
128 J.A. Noël, M.A. White / J. Chem. Thermodynamics 128 (2019) 127–133

melting temperature and the melting enthalpy change are avail- connectivity index which takes into account the masses of adjacent
able. However, there can be considerable variation in values pub- groups in the molecules. Ceriani et al. have put forth a group addi-
lished for the same compound [7–9]. Care must be taken to tivity method for determining heat capacities of hydrocarbons
consult the original literature to check the veracity of the data, commonly found in biodiesel and edible oils [23]. Their method
and the expected accuracy of the methods. Considerably less infor- has been found to accurately predict cp in many fatty compounds.
mation is available regarding the heat capacity of PCMs, making it Pilar et al. measured the heat capacities of 24 solid explosives by
difficult to fully model, and thereby optimize, thermal storage sys- DSC and compared their results to several group additivity estima-
tems that utilize PCMs [10]. Various calorimetric methods can be tions [24]. They found that the group additivity estimations could
used to determine heat capacity experimentally, but these require match some experiments within 1%, but had an average deviation
access to suitable instrumentation and a sample. For example, Jia of 11% at T = 298 K.
et al. have measured the heat capacities of six sugar alcohol PCMs, Group additivity can be used to estimate a range of thermody-
including three stereoisomers, by relaxation calorimetry [11]. namic properties in addition to the heat capacity. Much of the early
Chieruzzi et al. determined the heat capacity of a molten nitrate work on the estimation of thermodynamic properties by group
salt-based nanofluid PCM [12]. Lu et al. measured the heat capacity additivity was initiated by Benson, particularly the calculation of
of a solid-to-solid PCM, bis(1-octylammonium) tetrachlorocuprate transition temperatures, enthalpy changes and entropy changes
by adiabatic calorimetry [13]. In situations where experimental [25,26]. Benson also showed it to be possible to estimate the heat
data are insufficient or difficult to obtain, or a quick assessment capacity from vaporization enthalpy and entropy changes deter-
of the PCM would be helpful to decide whether to proceed with mined by group additivity [27]. Similar to the tables of functional
detailed investigations, it would be especially useful to have simple group contributions to the heat capacity of organic compounds,
methods to provide reasonable estimates of the thermodynamic Chickos et al. have provided tables of constituent contributions to
properties of PCMs, including the heat capacity. melting/fusion entropy and enthalpy changes for organic com-
The heat capacities of elemental solids at room-temperature are pounds, derived from 1858 compounds and tested on 260 more
well-approximated by the Dulong-Petit law [14]. The Dulong-Petit [28].
law states that all atoms of simple bodies have equal heat capacity, An attractive feature of group additivity estimations is that they
and in elemental solids at room temperature, a molar heat capacity are quick, usually simple, and provide results of reasonable accu-
of 3R, where R is the gas constant, is a good estimate. The racy. There are, nonetheless, more advanced methods for predicting
Neumann-Kopp law extends the Dulong-Petit model to com- heat capacities. For example, vibrational frequencies calculated via
pounds: the Neumann-Kopp law states that the heat capacity of ab initio methods can be used to estimate ideal gas heat capacities
a solid compound is equal to the sum of the heat capacities of its of organic compounds using the harmonic oscillator model with
constituent elements [14]. The Neumann-Kopp law works well high accuracy [29]. These estimates out-performed group additivity
for estimating the heat capacities of simple materials such as bin- methods for rigid molecules containing heteroatoms.
ary compounds or alloys, but often deviates from experimental When heat capacity additivity models do not represent the
results for more complex compounds in which chemical interac- experimental data well, new insights also can arise. An example
tions influence the physical properties. In such cases, methods of is the heat capacity of clathrates (binary materials with structures
group additivity can be used with considerable success. in which one component forms a lattice in which the other compo-
While the Neumann-Kopp law is based on the sum of contribu- nent is trapped). Sometimes the heat capacity of the guest mole-
tions from the individual elements, group additivity treats the heat cules and host lattice can be additive; an example is the hexakis
capacity as the sum of contributions from the constituent chemical (phenylthiol)benzene (HPTB) clathrate containing CBr4 [30]. How-
groups of the compound. In inorganic compounds, the constituent ever, the non-additivity of the guest and host heat capacity contri-
groups are often the base minerals. For example, the heat capacity butions in the case of the THF guest species in the corresponding
of sodalite (Na8Al6Si6O24Cl2) is well modelled from 100 to 1000 K clathrate hydrate indicates additional degrees of freedom for the
by summing the contributions of Na2O, Al2O3, a-SiO2, and NaCl THF guest molecules, making them ‘‘fluid-like” [31], and account-
[14]. This method works well for many mineral species. ing for its exceptionally low thermal conductivity [32] as the
Inorganic PCMs have high volumetric enthalpies of fusion, but heat-carrying phonons are perturbed by the motions of the THF
also have a number of drawbacks including potential for incongru- guests. These insights into a novel mechanism to reduce thermal
ent melting, corrosiveness, and supercooling [1]. Organic PCMs lar- conductivity opened an important new approach to more efficient
gely avoid these issues, while maintaining high enthalpies of thermoelectric materials [33], again contributing to capture of
fusion. As such, much interest has been focused on developing renewable energy.
organic PCMs. In organic compounds, the constituent groups are One promising class of PCMs for low-temperature (near or just
the functional groups of the molecule, and several methods for above ambient temperature) applications is linear, saturated fatty
determining the contributions of individual functional groups to esters [1]. Fatty esters have relatively high enthalpies of fusion
the heat capacity have been developed. Růžička and Domalski [34], and melting temperatures suitable for a variety of cooling
[15,16], Zábranský and Růžička [17], and Kolská et al. [18], have and refrigeration applications [1]. Linear esters have better packing
published extensive tables of temperature-dependent contribu- efficiency in the crystalline phase than branched esters, which
tions for numerous organic functional groups, to calculate the heat gives them higher values of DfusH. PCM interest is focussed on sat-
capacities of organic liquids with high accuracy. Chickos et al. have urated fats, as unsaturated fats are prone to oxidation at the double
developed a similar method for estimating heat capacities of both bond, which leads to poor long-term stability of the material.
solid and liquid organic compounds at room temperature [19]. In a recent study, we examined the influence of the distribution
Their method works well for both simple and more complex of methylene units between the two alkyl chains of isomeric linear
organic molecules such as sucrose. Su et al. have extensively esters on the melting temperature and on the melting enthalpy
reviewed a large number of group additivity methods useful for and entropy changes of the esters [34]. As isomers, the compounds
modeling working fluids used in thermodynamic cycles, such as were equal in molecular mass, all were linear, and all contained the
refrigeration [20]. Valderrama and coworkers [21] and Soriano same number of the same organic units: from a group additivity
et al. [22] have developed accurate models to determine the heat perspective, they were identical. It was found that esters with
capacities of ionic liquids. Valderrama’s model uses group additiv- longer uninterrupted alkyl chains had higher melting temperatures
ity in conjunction with a structural parameter called the mass and associated enthalpy changes than those with two shorter
J.A. Noël, M.A. White / J. Chem. Thermodynamics 128 (2019) 127–133 129

chains. For example, consider two 12-carbon isomers: ethyl decan- of the tests. The samples were measured in crimped aluminum
oate, with its 2-carbon and 10-carbon chain, has Tfus = 252 K and pans, and had masses of 6–9 mg. Melting points were determined
DfusH = 143 J mol1, while hexyl hexanoate, with two 6-carbon using onset temperatures and a scanning rate of 2 K min1. The
chains, has Tfus = 217 K and DfusH = 108 J mol1 [34]. The melting ASTM standard heat capacity determination method [41] is com-
entropy change, however, was largely independent of the alkyl parative, in which the sample data is compared to a reference
chain distribution. material with known Cp; in this case 6.5 mg of high-purity sapphire
The above example highlights one of the shortcomings of group crystallites (Calorimetry Conference [NBS-49]) was used. Each Cp
additivity methods: they fail to distinguish isomers containing the measurement consisted of three temperature scans performed
same elementary constituents. In the present work, we determine under identical conditions. The heat flow to the empty sample
experimentally the heat capacity of five 12-carbon, isomeric, liquid pan was measured first. Secondly, the heat flow to the reference
esters. The values from experiment are compared to values esti- material in a sample pan was measured. Finally, the heat flow to
mated by various group additivity methods [15–19,23]. With the the sample in a sample pan was measured. While it is recom-
exception of methyl esters [35,36], which are used in biodiesel, mended that the same sample pan be used for all three tempera-
there are few fatty ester liquids for which experimental heat capac- ture scans, we used three separate pans with well-matched
ity data are available, and therefore there is a need for heat capac- masses (<2% difference), and applied a correction factor as recom-
ity information for other esters as PCMs. mended in the standard method [41], to account for Al mass differ-
Fatty acids have been well-studied as PCMs for heat storage and ences. For each temperature scan, the DSC was held isothermally
load-levelling applications [1]. They offer high enthalpies of fusion, for 5 min at 203 K, ramped at 10 K min1 to 303 K, thus defining
cycle well [37], are non-toxic, and can be produced sustainably the temperature range of the experiment, and then held isother-
[38]. Dodecanoic acid, a promising PCM [37,39] and isomer of mally at 303 K for 5 min. The isothermal steps were included to
the five 12-carbon esters for which experimental heat capacity ensure that the samples had reached thermal equilibrium before
information is available [40] is also compared. The fatty acid has and after heating.
the same elemental composition as the esters, but differs in its con- The heat flow to the empty sample pan was subtracted from the
stituent chemical groups. heat flow to the reference material, and also to the sample, as back-
The goals of this work are thus twofold. First, we wish to deter- ground corrections. The differences between the heat flow signals
mine whether the distribution of methylene units between the two to the sample pan and the sapphire, Dst, and between the heat flow
alkyl chains of the esters influences the heat capacity, as it does signals to the sample pan and the sample, Ds, were then calculated
change the melting temperature and melting enthalpy of fusion as a function of temperature and used to calculate Cp. At any tem-
[34]. Secondly, we wish to know if the group additivity methods perature, the heat capacity of the sample, Cp(s), was found from [41]
describe the heat capacity of some isomers better than others,
Ds W st
based on the distribution of the constituents. The importance of C pðsÞ ¼ C pðstÞ
esters and saturated fatty acids for phase change materials provide Dst W s
 
the focus for this investigation. The information revealed will allow Ds ðW st  W pan Þ ðW s  W pan Þ
þ C pðpanÞ  ð2Þ
better modelling and optimization of PCM thermal energy storage Dst Ws Ws
systems, based on better prediction of PCM properties.
where Cp(st) and Cp(pan) are the heat capacities of the sapphire refer-
ence [41] and the empty sample pan, respectively, Wpan is the mass
2. Materials and methods of the empty sample pan used for the background subtraction, and
Ws and Wst are the masses of the sample pans containing the sample
The heat capacities of methyl undecanoate (Alfa Aesar, 99%), and sapphire reference material, respectively. The Cp(pan) term is
ethyl decanoate (Sigma-Aldrich, >98%), butyl octanoate (Aldrich), included to account for the minor mass differences between the
hexyl hexanoate (Sigma-Aldrich, >97%), and octyl butyrate sample pans used. The method was validated with indium, magne-
(Sigma-Aldrich > 98%) (Table 1) were measured according to the sium, and benzoic standards, compared to literature values, the rel-
ASTM standard test method designation E1269-11 [41]. All of the ative standard uncertainty was <10% [42]. For the purposes of
esters were used without further purification. Measurements were assessment of PCMs, and to see if additivity models accurately rep-
made using a TA Instruments DSC Q200 under a 25 mL min1 flow resent experimental heat capacities, this information is sufficiently
of helium gas. The instrument was calibrated for temperature with accurate. More accurate results could be obtained by other meth-
mercury and indium melting points as standards prior to the start ods, including by of use relaxation calorimetry which, with the heat

Table 1
Sample Sources and Purities.

Chemical Name Source Mass Fraction Purity (as received) Purification Method Method of Analysis
Test Materials
Octyl Butyrate Sigma-Aldrich >0.99a None GCe
Hexyl Hexanoate Sigma-Aldrich 0.97b None GC
Butyl Octanoate Aldrich >0.98c None MPt
Ethyl Decanoate Sigma-Aldrich 0.998a None GC
Methyl Undecanoate Alfa Aesar 0.986a None GC
Calibration and Standard Materials
Indium TA Instruments 0.99999d None –
Mercury TA Instruments 0.9999d None –
Sapphire Calorimetry Conference [NBS-49] 0.999d None –
a
Based on the manufacturer’s certificate of analysis indicating purity of 1.0.
b
As stated on the manufacturer’s product specification sheet.
c
Based on melting point relative to literature value; note that manufacturer does not record analytical data on this product.
d
As stated by the supplier.
e
Gas-chromatography
130 J.A. Noël, M.A. White / J. Chem. Thermodynamics 128 (2019) 127–133

capacity option of the Quantum Design Physical Properties Mea- group additivity methods for organic liquids. Table 2 presents the
surement System (PPMS), can give relative standard uncertainty measured and calculated heat capacities at T = 298 K, along with
of 2% for volatile samples [43,44]. Initial trials were run in a PPMS the deviations between experiment and calculation. The
but, with the very high volatility of the liquid esters, proved futile. temperature-dependent values of the heat capacities of the five
The need for a sealed pan to prevent the esters from volatilizing esters in the liquid phase are presented in the Supplemental
when placed under high vacuum would have required such a mas- Information.
sive seal that the container heat capacity would have dwarfed the Despite the differences in the distribution of methylene units
sample’s contribution to the total heat capacity. For further details, between the five isomers, each of the esters has the same heat
see Supplemental Information. capacity, within the standard uncertainty of the measurement,
over the temperature range examined. Isomeric esters exhibit dif-
ferences in melting temperature and in melting enthalpy change
3. Results and discussion
[34], but similar results are not observed when it comes to heat
capacity. The melting point and enthalpy change differences stem
The heat capacities of liquid octyl butyrate, hexyl hexanoate,
from increased stabilization of the crystalline phase, relative to the
butyl octanoate, ethyl decanoate, and methyl undecanoate were
liquid, by chain-chain interactions by isomers with longer alkyl
measured from 203 K to 303 K. The heat capacities of the esters
chains. Here, rather unsurprisingly, the relative thermodynamic
from their individual melting points to 303 K are plotted in
stabilities of the solid and liquid phases have little bearing on the
Fig. 1. All heat capacity values are given in units of the gas constant,
heat capacity of the liquid.
R, where R = 8.314 J K1 mol1. The experimentally determined
Fig. 2 compares the values of Cp for methyl undecanoate and
values are compared with values determined using five different
ethyl decanoate to previously published values [35,45,46]. Dzida
et al. [46] made their measurements using micro-DSC, while van
Bommel et al. [35] and Zaitsau et al. [46] used adiabatic calorime-
try. The results acquired in the present work by DSC underestimate
the heat capacity of the esters in both cases, but are still well
within the anticipated 10% uncertainty of the values from the other
methods. The results presented in this work share the same gen-
eral trend as the previously reported data, and accuracy within
10% is suitable for most PCM modelling applications. The deviation
likely arises due to low relative mass of the sample compared to
the mass of the sample pan used here.
The data from the adiabatic calorimetry and micro-calorimetry
studies are well-estimated by group additivity. The group additiv-
ity estimate for the heat capacity of a 12-carbon ester using the
method of Zábranský [17] also is plotted in Fig. 2, and describes
methyl undecanoate and ethyl decanoate equally well, with all
experimental data points within 1% of the estimate.
The heat capacity of linear esters in the liquid phase does
increase with increasing number of carbon atoms, and thereby
molar mass, as has been shown for series of methyl esters
[35,36]. Comparing our results, group additivity estimates for
esters[23], and heat capacity results for other esters [47–50] to
those for the methyl esters (Fig. 3), it appears that all linear, satu-
rated, fatty esters of the same molecular mass (i.e., number of car-
bons) have essentially equivalent heat capacity in the liquid phase.
Each of the five group additivity methods tested predicts the
Fig. 1. Heat capacities, Cp, in units of the gas constant, R, as a function of measured heat capacities of the five liquid 12-carbon esters with
temperature, T, of five 12-carbon esters in the liquid phase determined here by reasonable accuracy. At T = 298 K, no group additivity estimate
experiment [octyl butyrate ( ), hexyl hexanoate (d), butyl octanoate ( ), ethyl was further than 10% away from experiment, which is sufficient
decanoate ( ), methyl undecanoate ( )], and by the group additivity methods of
accuracy for PCM modelling. Depending on the method, the esti-
Ceriani (–––)[23], Kolská ( ) [18], Chickos ( ) [19], Růžička ( ) [15,16], and
Zábranský (. . .)[17]. A representative relative standard uncertainty (error bar of mates are on average 2–8% higher than the experiment at
±10%) is shown. Note that the method used here gives >50 data points per kelvin T = 298 K. Note that at T = 298 K, all five group additivity methods
and only selected data are shown here. appear to over-estimate the heat capacity, compared to our

Table 2
Measured and calculated heat capacities, Cp, of several 12-carbon esters at temperature T = 298 K and pressure p = 0.10 MPa, where heat capacities are expressed in units of the
gas constant, R. Note that the calculated values from the group additivity methods are independent of structure.a

Ester Experimental Ceriani 2009 Kolská 2008 Chickos 1993 Růžička 1993 Zábranský 2004 Average Difference (%)
Calculated 48.4 47.8 50.5 48.6 49.0
Octyl Butyrate 46.1 Difference 2.3 1.7 4.4 2.5 2.9 6
Hexyl Hexanoate 47.5 0.9 0.3 3.0 1.1 1.5 3
Butyl Octanoate 47.8 0.6 0 2.7 0.8 1.2 2
Ethyl Decanoate 46.1 2.3 1.7 4.4 2.5 2.9 6
Methyl Undecanoate 46.8 1.6 1.0 3.7 1.8 2.2 5
Average Difference (%) 3 2 8 4 5
a
Standard uncertainties, u, are ur(Cp) = 10%, u(T) = 0.1 K, u(p) = 10 kPa.
J.A. Noël, M.A. White / J. Chem. Thermodynamics 128 (2019) 127–133 131

Fig. 2. (a) Heat capacity, Cp, of methyl undecanoate as a function of temperature, T, determined in this work (d) and by van Bommel et al. ( ) [35] and compared to the group
additivity estimate of Zábranský (–––) [17] (b) Heat capacity of ethyl decanoate as a function of temperature determined in this work (d) and Zaitsau et al. ( ) [45] and Dzida
et al. ( ) [46] and compared to the group additivity estimate of Zábranský (–––) [17]. Note that the method used here gives >50 data points per kelvin and only selected data
are shown in these figures.

experimental values. In terms of the group additive methods’ abil- phase heat capacities, the heat capacity of the methyl ester of equal
ity to distinguish isomers, all of the ester isomers are predicted molar mass to the desired ester for the PCM application can be
well, within 6% of experiment on average. used a reasonably accurate surrogate. The heat capacities for many
With regard to PCM selection and modelling: as all linear satu- fatty acid methyl esters are known [35,36]. In the absence of exper-
rated esters of equal molar mass have approximately equal liquid imental information, group additivity methods provide estimates
of reasonably accuracy (<10% deviation at 298 K) for PCM use for
fatty esters.
Dodecanoic acid is an isomer of the five 12-carbon esters exam-
ined. It, however, is different from a group additivity perspective.
Comparison of the heat capacity of liquid dodecanoic acid deter-
mined by adiabatic calorimetry [40] to the group additivity estima-
tions shows that the acid is well described by the estimations as an
acid, whereas comparable ester estimates are lower (Fig. 4). Thus,
although dodecanoic acid is isomeric to the 12-carbon esters,
dodecanoic acid behaves differently with respect to heat capacity.
In addition, using group additivity, it is possible to distinguish
isomers when the constituent groups are altered, even when the
alteration is relatively minor (ester to acid).
Heat capacities of liquid saturated, n-fatty acids of even num-
bers of carbons in the range 8 to 20 carbons are, in general, mod-
elled well using group additivity. Comparing the experimental
results of Schaake [40] to four group additivity estimates
[15–18,23], there is little difference (within about 2%; see Fig. 5).
Saturated fatty acids are useful PCMs [39] and the present results
suggest that group additivity estimations provide acceptable heat
capacity estimates for liquid fatty acids as PCMs. Such estimates
can be used in modelling PCM systems at temperatures for which
experimental data are not available (e.g., Schaake gives heat capac-
ities at only two temperatures for liquid hexadecanoic acid,
octadecanoic acid and eicosanoic acid [40]).
In conclusion, we have determined experimentally the heat
Fig. 3. Heat capacities, Cp, in units of the gas constant, R, of linear saturated fatty capacity of five saturated 12-carbon linear ester isomers in the liq-
esters in the liquid phase (at T = 300 K, unless otherwise stated) from this work (j), uid phase from just above their melting points to T = 303 K, by DSC.
methyl esters from van Bommel ( ) [35], Pauly ( ) [36], Liu ( ) [47] (305.4 K), van The ASTM method used provided data of reasonable accuracy
Miltenburg ( ) [48], Agafonova (s) [49] (350 K), and Dzida (x)(298.13 K) [50], and
without requiring high vacuum, which is detrimental for such
group additivity estimates by the method of Ceriani ( ) [23]. The 20-carbon esters
(open circles) are given at T = 350 K as they are solid at T = 300 K. Error bar highly volatile liquid samples. In this temperature range, values
represents the relative standard uncertainty of the measurements ur(Cp) = 10%. of Cp of the five isomers are equal within the standard uncertainty
132 J.A. Noël, M.A. White / J. Chem. Thermodynamics 128 (2019) 127–133

experimental data are not available. Using group additivity, the Cp


of liquid fatty acids also can be accurately predicted and distin-
guished from their isomeric esters. Group additivity has been
shown to provide Cp values for potential organic phase change
materials that are of acceptable accuracy to be used in modelling
of heat storage systems for their optimization, in the absence of
experimental data.

Acknowledgements

J. A. N. acknowledges support from Dalhousie Research in


Energy, Advanced Materials and Sustainability (DREAMS), an
NSERC CREATE program, and an NSERC and a Nova Scotia scholar-
ship. M. A. W. acknowledges support from NSERC and the Clean
Technologies Research Institute at Dalhousie University. We also
acknowledge S. Kahwaji for assistance. The authors thank and
commend Professor Trevor Letcher on his abiding interest in chem-
ical thermodynamics and sustainable energy, and dedicate this
paper to him on the occasion of his 80th birthday.

Appendix A. Supplementary data

Supplementary data associated with this article can be found, in


Fig. 4. Heat capacity, Cp, in units of the gas constant, R, of liquid dodecanoic acid ( )
the online version, at https://doi.org/10.1016/j.jct.2018.08.014.
[40] as a function of temperature, T, compared to values estimated by group
additivity for dodecanoic acid [Ceriani (–––), Kolská (  ), Růžička (-  -), References
Zábranský (-  -)] and for 12-carbon esters [Ceriani ( ), Kolská (   ), Růžička
(-  -), Zábranský (-  -)] [15–18,23]. [1] J.A. Noël, S. Kahwaji, L. Desgrosseilliers, D. Groulx, M.A. White, Phase Change
Materials, in: T. Letcher (Ed.), Storing Energy, Elsevier, Amsterdam, 2016, pp.
249–272.
[2] A. Sharma, V.V. Tyagi, C.R. Chen, D. Buddhi, Review on thermal energy storage
with phase change materials and applications, Renew. Sust. Energy Rev. 13
(2009) 318–345.
[3] B. Zalba, J.M. Marin, L.F. Cabeza, H. Mehling, Review on thermal energy storage
with phase change: materials, heat transfer analysis and applications, Appl.
Therm. Eng. 23 (2003) 251–283.
[4] C.A. Whitman, M.B. Johnson, M.A. White, Characterization of thermal
performance of a solid-solid phase change material, Di-n-Hexylammonium
bromide, for potential integration in building materials, Thermochim. Acta 531
(2012) 54–59.
[5] A. Joseph, M. Kabbara, D. Groulx, P. Allred, M.A. White, Characterization and
real-time testing of phase-change materials for solar thermal energy storage,
Int. J. Energy Res. 40 (2016) 61–70.
[6] C. Arkar, S. Medved, Free cooling of a building using PCM heat storage
integrated into the ventilation system, Sol. Energy 81 (2007) 1078–1087.
[7] K. Pielichkowska, K. Pielichkowski, Phase change materials for thermal energy
storage, Prog. Mat. Sci. 65 (2014) 67–123.
[8] W. Acree, J.S. Chickos, Phase transition enthalpy measurements of organic and
organometallic compounds. sublimation, vaporization and fusion enthalpies
from 1880 to 2015. Part 1. C1  C10, J. Phys. Chem. Ref. Data. 45 (2016) 03101.
[9] W. Acree, J.S. Chickos, Phase transition enthalpy measurements of organic and
organometallic compounds and ionic liquids. sublimation, vaporization, and
fusion enthalpies from 1880 to 2015. Part 2. C11–C192, J. Phys. Chem. Ref.
Data. 46 (2017) 013104.
[10] Y. Cascone, M. Perino, Estimation of the thermal properties of PCMs through
inverse modelling, Energy Procedia 78 (2015) 1714–1719.
[11] R. Jia, K. Sun, R. Li, Y. Zhang, W. Wang, H. Yin, D. Fang, Q. Shi, Z. Tan, Heat
capacities of some sugar alcohols as phase change materials for thermal
energy storage applications, J. Chem. Thermodyn. 233–248 (2017).
[12] M. Chieruzzi, G.F. Cerritelli, A. MIliozzi, J.M. Kenny, Effect of nanoparticles on
heat capacity of nanofluids based on molten salts as PCM for thermal energy
storage, Nanoscale Res. Lett. 8 (2013) 448-.
Fig. 5. Experimental heat capacities, Cp, ( ) [40] and group additivity estimates [13] D.F. Lu, Y.Y. Di, Z.C. Tan, J.M. Dou, Low-temperature heat capacities, and
using the methods of Růžička ( ), Ceriani ( ), Kolská (▲), and Zábranský ( ) [15– thermodynamic properties of solid-solid phase change material bis(1-
Octylammonium) tetrachlorocuprate, J. Therm. Anal. Calorim. 111 (2013)
18,23] for saturated n-fatty acids of general formula CnH2n+1OOH, n = even, at
213–218.
temperature, T = 300 K (C8), T = 345 K (C10 to C16), or T = 355 K (C18 to C20) in units of
[14] L. Qiu, M.A. White, The constituent additivity method to estimate heat
the gas constant, R. Vertical dashed lines separate different temperatures. capacities of complex inorganic solids, J. Chem. Ed. 78 (2001) 1076–1079.
[15] V. Růžička Jr, E.S. Domalski, Estimation of the heat capacities of organic liquids
as a function of temperature using group additivity. I – hydrocarbon
of the measurement, despite the variation in the distribution of compounds, J. Phys. Chem. Ref. Data. 22 (1993) 597–618.
[16] V. Růžička Jr, E.S. Domalski, Estimation of the heat capacities of organic liquids
carbon atoms between the two alkyl chains. It is proposed that as a function of temperature using group additivity. II – compounds of carbon,
the heat capacities of liquid linear esters depend only on molar hydrogen, halogens, nitrogen, oxygen and sulfur, J. Phys. Chem. Ref. Data. 22
mass, not alkyl chain distribution. The Cp values of the liquid esters (1993) 619–657.
[17] M. Zábranzký, V. Růžička, Estimation of the heat capacities of organic liquids
are well described by group additivity, and as such, group additiv- as a function of temperature using group additivity: an amendment, J. Phys.
ity estimates provide a reasonable Cp value in circumstances when Chem. Ref. Data 33 (2004) 1071–1081.
J.A. Noël, M.A. White / J. Chem. Thermodynamics 128 (2019) 127–133 133

[18] Z. Kolská, J. Kukal, M. Zábranzký, V. Růžička, Estimation of the Heat capacity of [36] J. Pauly, A.C. Kouakou, M. Habrioux, K. Le Mapihan, Heat capacity
organic liquids as a function of temperature by a three-level group measurements of pure fatty acid methyl esters and biodiesels from 250 to
contribution method, Ind. Eng. Chem. Res. 47 (2008) 2075–2085. 390 K, Fuel. 137 (2014) 21–27.
[19] J.S. Chickos, D.G. Hesse, J.F. Liebman, A group additivity approach for the [37] L. Desgrosseilliers, C.A. Whitman, D. Groulx, M.A. White, Dodecanoic acid as a
estimation of heat capacities of organic liquids and solids at 298 K, Struct. promising phase-change material for thermal energy storage, Appl. Therm.
Chem. 4 (1993) 261–269. Eng. 53 (2013) 37–41.
[20] W. Su, L. Zhao, S. Deng, Group contribution methods in thermodynamic cycles: [38] J.A. Noël, P.M. Allred, M.A. White, Life cycle assessment of two biologically
physical properties estimation of pure working fluids, Renew. Sust. Energy produced phase change materials and their related products, Int. J. Life Cycle
Rev. 79 (2017) 984–1001. Assess. 20 (2015) 367–376.
[21] J.O. Valderrama, A. Toro, R.E. Rojas, Prediction of the heat capacity of ionic [39] S. Kahwaji, M.B. Johnson, A.C. Kheirabadi, D. Groulx, M.A. White, Fatty acids
liquids using the mass connectivity index and a group contribution method, J. and related phase change materials for reliable thermal energy storage at
Chem. Thermodynamics. 43 (2011) 1068–1073. moderate temperatures, Sol. Energy Mater. Sol. Cells. 167 (2017) 109–120.
[22] A.N. Soriano, A.M. Agapito, L.J.L.I. Lagumbay, A.R. Caparanga, M.H. Li, A simple [40] R.C.F. Schaake, J.C. van Miltenbeurg, C.G. de Kruif, Thermodynamic properties
approach to predict molar heat capacity of ionic liquids using group-additivity of the normal alkanoic acids II. Heat Capacities of Seven Even-Numbered
method, J. Taiwan Inst. Chem. Eng. 41 (2010) 307–314. Alkanoic Acids, J. Chem. Thermodynamics 14 (1982) 771–778.
[23] R. Ceriani, F. Gani, A.J.A. Meirelles, Prediction of heat capacities and heats of [41] ASTM International, E1269–11 Standard Test Method for Determining Specific
vaporization of organic liquids by group contribution methods, Fluid Phase Heat Capacity by Differential Scanning Calorimetry. 2011.
Equilibria. 283 (2009) 49–55. [42] S. Kahwaji, M.B. Johnson, A.C. Kheirabadi, D. Groulx, M.A. White, Stable, low-
[24] R. Pilar, J. Pachman, R. Matyás, P. Honcová, Comparison of heat capacity of cost phase change material for building applications: the eutectic mixture of
solid explosives by DSC and group contribution methods, J. Therm. Anal. decanoic acid and tetradecanoic acid, Appl. Energy 168 (2016) 457–464.
Calorim. 121 (2015) 683–689. [43] C.A. Kennedy, M. Stancescu, R.A. Marriott, M.A. White, Recommendations for
[25] S.W. Benson, J.H. Buss, Additivity rules for the estimation of molecular accurate heat capacity measurements using a Quantum Design physical
properties. Thermodynamic Properties, J. Chem. Phys. 29 (1958) 546–572. property measurement system, Cryogenics 47 (2007) 107–112.
[26] S.W. Benson, F.R. Cruickshank, D.M. Golden, G.R. Haugen, H.E. O’Neal, A.S. [44] R.A. Marriott, M. Stancescu, C.A. Kennedy, M.A. White, Technique for
Rogers, R. Shaw, R. Walsh, Additivity rules for the estimation of determination of accurate heat capacities of volatile, powdered, or air-
thermochemical properties, Chem. Rev. 69 (1969) 279–324. sensitive samples using relaxation calorimetry, Rev. Sci. Instrum. 77
[27] S.W. Benson, New methods for estimating the heats of formation, heat (096108) (2006) 1–3.
capacities, and entropies of liquids and gases, J. Phys. Chem. A. 103 (1999) [45] D.H. Zaitsau, Y.U. Paulechka, A.V. Blokhin, A.V. Yermalayeu, A.G. Kabo, M.R.
11481–11485. Ivanets, Thermodynamics of ethyl decanoate, J. Chem. Eng. Data. 54 (2009)
[28] J.S. Chickos, W.E. Acree, J.F. Liebman, Estimating solid-liquid phase change 3026–3033.
enthalpies and entropies, J. Phys. Chem. Ref. Data. 28 (1999) 1535–1673. [46] M. Dzida, S. Je˛zak, _ _
J. Sumara, M. Zarska, P. Góralski, High-pressure
[29] R.A. Marriott, M.A. White, Comparison of Ab initio and group additive ideal gas physicochemical properties of ethyl caprylate and ethyl caprate, J. Chem.
heat capacities, AIChE J. 51 (2005) 292–297. Eng. Data. 58 (2013) 1955–1962.
[30] D. Michalski, R.T. Perry, M.A. White, J. Phys.: Condens. Matter 8 (1996) 1647. [47] X. Liu, C. Su, X. Qi, M. He, Isobaric heat capacities of ethyl heptanoate and ethyl
[31] M.A. White, Thermal conductivity of tetrahydrofuran clathrate hydrate, J. de cinnamate at pressures up to 16.3 MPa, J. Chem. Thermodyn. 93 (2016) 70–74.
Physique (Paris) 48 (1987) 565–571. [48] J.C. van Miltenburg, H.A.J. Oonk, Thermal properties of ethyl undecanoate and
[32] J.S. Tse, M.A. White, Origin of glassy crystalline behaviour in the thermal ethyl tridecanoate by adiabatic calorimetry, J. Chem. Eng. Data. 50 (2005)
properties of clathrate hydrates: a thermal conductivity study of 1348–1352.
tetrahydrofuran hydrate, J. Phys. Chem. 92 (1988) 5006–5011. [49] L.E. Agafonova, R.M. Varushchenko, A.I. Druzhinina, O.V. Polyakova, Y.S.
[33] G.A. Slack, in: CRC Handbook of Thermoelectrics, CRC Press, Boca Raton, 1995, Kolesov, Heat capacity, saturation vapor pressure, and thermodynamic
p. 407. functions of ethyl esters of C3–C5 and C18 carboxylic Acids, Russian J. Phys.
[34] J.A. Noël, S. Kahwaji, M.A. White, Molecular structure and melting: Chem. A. 85 (2011) 1516–1527.
implications for phase change materials, Can. J. Chem. 96 (2018) 722–729, [50] M. Dzida, S. Je˛zak, _ _
J. Sumara, M. Zarska, P. Góralski, High pressure
https://doi.org/10.1139/cjc-2017-0578. physicochemical properties of biodiesel components used for spray
[35] M.J. van Bommel, H.A.J. Oonk, J.C. van Miltenburg, Heat capacity characteristics in diesel injection systems, Fuel. 111 (2013) 165–171.
measurements of 13 methyl esters of n-Carboxylic acids from methyl
octanoate to methyl eicosanoate between 5 K and 350 K, J. Chem. Eng. Data.
49 (2004) 1036–1042. JCT 2018-189

You might also like