You are on page 1of 11

Adsorption (2018) 24:345–355

https://doi.org/10.1007/s10450-018-9946-1

Confinement effect on enthalpy of fusion and melting point of organic


phase change materials in cylindrical nanospace of mesoporous silica
and carbon
Jihye Choi1 · Hirotaka Fujita1 · Masaru Ogura1 · Akiyoshi Sakoda1

Received: 8 December 2017 / Revised: 8 February 2018 / Accepted: 21 March 2018 / Published online: 26 March 2018
© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Organic phase change materials (PCMs) were successfully confined into mesopores of host materials independently via vapor
transportation to precisely investigate the changes in the enthalpy of fusion and the melting point of such confined PCMs
under various conditions. Paraffins, fatty acids, and fatty alcohols with long hydrocarbon chains were employed as guest
PCMs. Mesoporous silica SBA-15s and soft-templated mesoporous carbons with cylindrical mesopores were employed as
host materials of the guest PCMs. It was elucidated that mesopore diameter, functional groups of both PCMs and functional
groups of host materials result in significant changes in the enthalpy of fusion and the melting point of confined PCMs. Fur-
thermore, it was concluded that the host materials with mesopores of diameter 10–20 nm and minimum interaction between
PCM molecules and the functional group on the wall of mesopores of host materials are required to obtain an enthalpy of
fusion of confined PCMs as much as 50% of that in its bulk phase.

Keywords  Confinement effect · Phase change material · Enthalpy of fusion · Melting point

1 Introduction solid–liquid phase change is most commonly used because


it is easier to confine solid–liquid PCMs due to small volume
Owing to the depletion of fossil fuels and increased energy change as compared to gas–liquid or gas–solid PCMs (Mon-
consumption, it is important to not only develop new green dal 2008). There are many types of PCMs such as organics,
energies but also conserve energy. Specifically, efficient use inorganics, metals, and hydrates (Farid et al. 2004). For host
of energy such as heat recovery and thermal energy storage materials to confine solid–liquid PCMs in their structures,
are urgently required. Heat can be divided into three dif- polymer (Sari 2004), metallic voids (Li et al. 2015), and
ferent types: sensible heat (Dincer et al. 1997), latent heat micro-capsule (Sari et al. 2014), etc., were reported.
(Kadoono and Ogura 2014), and heat generated during a Recently, porous materials have newly been reported as a
chemical reaction (Aihara et al. 2001). Among them, latent novel host material owing to their high specific surface area,
heat enables us to store thermal energy at a high density level high structural stability, and most importantly, less leakage
within a narrow temperature range. In order to utilize the of liquefied PCMs owing to high surface tension derived
merits of latent heat of phase change, a phase change mate- from nanopores (Kadoono and Ogura 2014; Mitran et al.
rial (PCM) is employed in various heat storage applications 2015; Wang et al. 2016). In the past, melting and freezing
such as solar energy harvesting (Feng et al. 2011), building properties in a limited space such as a capillary have mainly
materials (Zhou et al. 2012) and enhancement of perfor- been researched using small model molecules such as water
mance of the adsorption process (Horstmeier et al. 2016; (Morishige and Iwasaki 2003), methane (Miyahara and Gub-
Li and Li 2015). Among various types of phase changes, bins 1997), and benzene (Jackson and Mckenna 1990). How-
ever, these molecules are not suitable for PCMs owing to
their high degree of super cooling and much lower or higher
* Masaru Ogura melting point as compared to room temperature. Currently,
oguram@iis.u‑tokyo.ac.jp the phase change behaviors of confined, long-chain organic
1
Institute of Industrial Science, The University of Tokyo, PCM molecules with less degree of super cooling and
Komaba 4‑6‑1, Meguro, Tokyo 153‑8505, Japan melting point near room temperature such as polyethylene

13
Vol.:(0123456789)

346 Adsorption (2018) 24:345–355

glycol (Min et al. 2015), paraffin (Nomura et al. 2015), and diameter, e.g., SBA-15-7.6 nm. SBA-15-7.6 nm was syn-
fatty acid (Kadoono and Ogura 2014) are under discussion. thesized by following the procedure proposed by Zhao et al.
Specifically, among the melting properties, decrease in the (1998). 4.0 g of pluronic P123 was added in 30 g of water
melting point of these PCM molecules is being intensively and 120 g of 2 M HCl aqueous solution. After P123 was
investigated. On the other hand, in addition to the melting totally dissolved in solution, 8.50 g of TEOS was added,
point, the enthalpy of fusion, which determines the amount and the mixture was stirred for 20 h at 35 °C. Subsequently,
of stored heat, is an important property for the practical the mixture was aged at 100 °C without stirring. Then, the
application of PCMs in thermal energy storage. However, solid product was filtered, washed with DI water and dried
only a few previous works dealt with the change of enthalpy at 100 °C for 12 h. Then, SBA-15 was calcined under air
of fusion. To the best of our knowledge, almost all PCM- flow for 6 h at 773 K. A heating rate was 1 K/min. SBA-
confined materials were prepared using a liquid impregna- 15-5.6 nm was synthesized under the same conditions as
tion method in the previous works (Mitran et al. 2015; Wang those of SBA-15-7.6 nm except for the change of aging
et al. 2016; Jackson and Mckenna 1990). Liquid impregna- temperature from 100 to 40 °C (Valange et al. 2009). SBA-
tion is a method of confining PCM into mesopores of host 15-12.5 nm was synthesized by following the procedure
material via evaporation of volatile liquid solvent from a proposed by Kruk et al. (Kruk 2012), wherein triisopropylb-
mixture of dissolved PCM and host material by heating or enzene (TIPB) was added as a swelling agent while mixing
evacuation. During this liquid impregnation step, host mate- P123 and HCl solution to expand the cylindrical micelle,
rial comes into contact with the dissolved liquid phase PCM, followed by stirred at 17 °C and aged at 120 °C.
and this result in PCM exists not only inside mesopore, but
also outside mesopore, i.e., on the external surface and inter- 2.2 Preparation of mesoporous carbons
granular space. Consequently, the enthalpy of fusion of con-
fined PCM in mesopore was discussed in terms of the units Soft-templated mesoporous carbon was synthesized by
of J/g-sample or J/g-PCM approximately (Mitran et al. 2015; using pluronic triblock copolymers ­EO100PO65EO100 (F127,
Wang et al. 2016; Jackson and Mckenna 1990) because the Sigma-Aldrich) as the surfactant for the cylindrical micelle
accurate amount of PCM existing inside the mesopore and to form a mesoporous structure and resorcinol as the carbon
the exact pore filling ratio of PCM were unknown. precursor. The synthesis procedure proposed by Libbrecht
In this study, we confined PCMs using vapor transpor- et al. (2015) was followed. Accordingly, 2.2 g of resorcinol
tation. With this method, PCMs can exist only inside the (Wako) and 2.2 g of F127 were mixed in 9 mL of ethanol
mesopores and this enables us to calculate the exact value of and 9 mL of 3M HCl, and stirred until F127 was fully dis-
the pore filling ratio. Using the pore filling ratio, the change solved in a closed vial. Then, 2.6 g of formaldehyde (Wako,
of enthalpy of fusion could be accurately calculated with 37 wt%) was added and the solution was stirred at 100 rpm
the unit of J/g-PCM, and the adequate and effective condi- for 15 min in the same closed vial. After the viscosity of the
tions for obtaining a high value of the latent heat of PCMs transparent solution increased, the solution was poured into
were determined. Paraffin, fatty alcohol, and fatty acid were a 10 cm-diameter petri dish. Next, ethanol was evaporated at
selected as PCMs in this study. Mesoporous silica SBA-15 room temperature for 6 h and subsequently cured at 60 °C in
and soft-templated mesoporous carbon, which possesses a an oven for 18 h. Subsequently, the resorcinol/formaldehyde
structure analogous to SBA-15, were prepared as the host resin deposited at the bottom of the petri-dish was scratched
materials for the PCMs. Various methods of confining off, and thereafter calcined under ­N2 flow. First, the tempera-
PCMs, pore diameters of the hosts, functional groups of ture was maintained at 350 °C for 2 h at a heating rate of
PCMs, and quantity of functional groups of host materials 1 K/min. Then, the temperature was increased to 800 °C for
are investigated as the key parameters toward achieving high 3 h to carbonize the resin at the rate of 2 K/min. Half of the
value of enthalpy of fusion. product was oxidized using ozone to increase the oxygen-
containing functional groups on the surface of carbon. The
obtained mesoporous carbon was denoted as MC and its
2 Experimental oxidized sample as OMC.

2.1 Preparation of mesoporous silica SBA‑15s 2.3 Characterization of SBA‑15s and mesoporous


carbons
SBA-15 was synthesized using triblock copolymer
­E O 20PO 70EO 20 (pluronic P123, Sigma-Aldrich) as the Nitrogen adsorption (Quantachrome, QUADRSORB evo™)
surfactant for cylindrical micelle to form a mesoporous was performed to analyze the total pore volume Vt , BET sur-
structure, with tetraethyl orthosilicate used as the silica face area SBET  , pore diameter d , and pore size distribution of
source. Each SBA-15 sample name was denoted by the pore the substances used in this study. Pre-treatment of nitrogen

13
Adsorption (2018) 24:345–355 347

adsorption of mesoporous samples was 300 °C for 3 h, and acid, tetradecanol, and dodecanol—were confined in each
that for PCM-confined mesoporous samples was 25 °C for SBA-15 sample and two species—octadecane and tetrade-
3 h. Vt is calculated by amount of nitrogen adsorbed when canol—were confined in MC and OMC.
the relative pressure is 0.95. The pore size distribution was Moreover, octadecane confined in SBA-15 samples with
calculated using the NLDFT method based on a cylindrical different mesopore diameters was prepared using the typi-
pore model. cal liquid impregnation method according to the method
X-ray diffraction (XRD) patterns were recorded on a dif- described in literature (Mitran et al. 2015) to compare the
fractometer (Rigaku, Rint-2100) operating at 40 kV and difference between the melting properties of PCM con-
40 mA using Cu Kα radiation. fined in mesopores using vapor transportation and liquid
Boehm titration was performed (Boehm et al. 1964) to impregnation.
characterize the quantity of surface functional groups on
mesoporous carbon. Accordingly, 0.1 g of MC or OMC was
2.5 Analysis of the melting properties of PCM
added to 0.1 M NaOH solution and 0.1 M ­NaHCO3 solution.
molecules inside the confined nanospace
Subsequently, the mixture was stirred at 300 rpm for 3 days
at 30 °C and carbon powder was recovered via filtration. The
The melting properties of confined PCM were analyzed
filtered solution was titrated using 0.1 M HCl solution. The
using differential scanning calorimetry (DSC, Shimadzu
amount of oxygen-containing functional groups was calcu-
DSC-60). Accordingly, 2–3 mg of PCM-confined sample
lated using titration (Scheibe et al. 2009).
was used to perform DSC at the heating/cooling rate of
The thermal properties of the PCM confined in SBA-15
10 K/min.
were measured using thermogravimetric analysis (TGA;
Rigaku, Thermoplus TG8120) in the temperature range of
room temperature to 250 °C at the heating rate of 1 K/min.
3 Results and discussion
2.4 Confinement of PCM molecules to mesopores 3.1 PCMs confined in silica hosts
of SBA‑15s and mesoporous carbons
Figure  1 shows the nitrogen adsorption–desorption iso-
PCMs were confined using vapor transportation. First, a host therms of synthesized mesoporous silica SBA-15-5.6,
material was placed in a vial and dried at 130 °C for more SBA-15-7.6, and SBA-15-12.5 nm and also the pore size
than 1 h, the weight of which was named as mb as used distributions. The corresponding XRD patterns are shown
below. And glass wool was used to cover the vial to prevent in Fig. 2. Table 1 shows the porous characteristics of each
host materials from escaping during the evacuation step. SBA-15 sample.
The vial with containing the host material was subsequently The micropore volume Vmicro of SBA-15 was calculated
placed in the upper part of a corked vacuum desiccator and a using the t-plot method. The mesopore volume Vmeso was
PCM was placed at the bottom of the desiccator separately. calculated as Vt − Vmicro . Three SBA-15 samples, SBA-15-
The desiccator was evacuated and maintained at 100–130 °C 5.6, SBA-15-7.6 and SBA-15-12.5 nm, with cylinder-shaped
in an oven for a few hours such that the vaporized PCM can mesopores were successfully obtained. The pore diameters
be adsorbed and condensed inside the mesopores of host of SBA-15-5.6, SBA-15-7.6, and SBA-15-12.5 nm calcu-
materials. The resultant solid was recovered, then weighed, lated using the NLDFT method were 5.6, 7.6, and 12.5 nm,
which was defined as ma used below. The vapor pressure of respectively. SBA-15-12.5 nm was observed to have a rela-
each PCM was approximately 1 mmHg at the 100–130 °C. tively wide pore size distribution as compared to those of
By measuring the weight of the sample before and after the SBA-15-5.6 and SBA-15-7.6 nm. This is because TIPB,
PCM was confined, the pore filling ratio w of PCM, which which works as a swelling agent, was heterogeneously
is the ratio of the volume of confined PCM to the total pore solved in the micelle. The XRD patterns of all the SBA-15
volume of SBA-15, was calculated as follows: samples showed a two-dimensional hexagonal periodicity
100(ma − mb ) originating from SBA-15.
w= (1) Nitrogen adsorption–desorption isotherms of octadecane-
𝜌mb Vt
confined SBA-15 are shown in Fig. 3.
where ma represents the weight of PCM-confined SBA- Also, Fig. 4 shows the pore filling ratio of confined PCMs
15, mb represents the weight of SBA-15, 𝜌 represents the with different pore diameters of SBA-15. The pore filling
density of PCM, and Vt represents the total pore volume of ratio was calculated using the weight change of the sam-
SBA-15 calculated from nitrogen adsorption. Six species of ple before and after confining PCM molecules using vapor
PCMs—octadecane, tetradecane, dodecanoic acid, decanoic transportation (1).

13

348 Adsorption (2018) 24:345–355

Fig. 1  a Nitrogen adsorp-


tion–desorption isotherms of
SBA-15s. b Pore size distribu-
tion of SBA-15 calculated by
the NLDFT method

tendency was observed in Fig. 4. Almost 100% of the pores


of SBA-15-5.6 and SBA-15-7.6 nm were filled with PCMs.
However, the pore filling ratio of SBA-15-12.5 nm was only
65%. These indicate that the confinement of PCM with 100%
pore filling ratio becomes difficult with the increase in pore
diameter. This is because the condensation of PCMs inside
the mesopores occur higher relative pressure with increase
of pore diameter. Since we confined PCM in the low vapor
pressure which was approximately 1 mmHg during the vapor
transportation, the relative pressure of PCM in the step of
vapor transportation was not high enough and this causing
difficulty in condensation and fulfillment of PCM inside
mesopores with increase of pore diameter.

3.2 Effect of confining methods on melting


properties
Fig. 2  Low angle XRD patterns of SBA-15s obtained in this study

Octadecane was confined in the mesopore of SBA-15 sam-


The vertical axis in Fig. 3 indicates the adsorbed volume ples using the vapor transportation or liquid impregnation
of nitrogen per weight of SBA-15, which was calculated methods to compare their effects on the transportation of
by using results from Fig. 4. Figure 3 shows that the total PCM. The melting properties of the octadecane confined in
amount of nitrogen adsorbed on each octadecane-confined SBA-15 prepared using the different methods are measured
SBA-15 was considerably reduced, when compared to the using the DSC as shown in Fig. 5.
total amount of nitrogen adsorbed on pure SBA-15 (Fig. 1), The samples prepared using vapor transportation showed
which suggests that PCMs were filled inside the mesopore. only one peak of phase change during both cooling and
However, the total volume of nitrogen adsorbed on octade- heating steps. However, the samples prepared using liquid
cane-confined SBA-15 was different for different pore diam- impregnation demonstrated several peaks. One near the peak
eters. The volume of nitrogen adsorbed on SBA-15 samples obtained from the samples prepared using vapor transpor-
increased with the increase in pore diameter. Also, same tation, and at least one more peak near the bulk reference,

Table 1  Characteristics of SBA- Sample SBET ­(m2/g) Vt ­(cm3/g) Vmicro ­(cm3/g) Vmeso ­(cm3/g) d (nm)
15 used in this study
SBA-15-5.6 nm 492 0.444 0.148 0.296 5.6
SBA-15-7.6 nm 941 1.093 0.115 0.978 7.6
SBA-15-12.5 nm 533 1.534 0.000 1.534 12.5

13
Adsorption (2018) 24:345–355 349

heterogeneously distributed also on the external surface or


in the intergranular space of SBA-15 owing to the excessive
weight loss of the sample prepared using liquid impregna-
tion. Moreover, the weight loss of the sample prepared using
liquid impregnation began at lower temperature as compared
to the sample prepared using vapor impregnation. This is
because bulk-derived octadecane molecules existing in the
sample prepared using liquid impregnation have weak inter-
action compared to octadecane confined by surface tension,
and hence, the vaporization of octadecane began in a lower
temperature range.
The observed effect of bulk-derived PCM on its melting
properties was a shift of both melting and freezing points
(Fig. 5). When bulk-derived PCM exists, the nano-confined
peaks shifted closer to bulk-derived peak at a high tempera-
ture. Since there was no difference between the two samples
except for their confining methods, the shift of both melting
Fig. 3  Nitrogen adsorption–desorption isotherms of octadecane-con-
fined SBA-15 and freezing points was caused by the method of loading of
PCM on the SBA-15 porous structure. The shift of the freez-
ing point might have been affected by the nucleation of bulk
PCMs (Kittaka et al. 2013). Furthermore, since bulk PCMs
exist on the external surface and in the intergranular space
of its substrate, the meniscus of the confined PCM could
be changed (Schreiber et al. 2001) and this possibly has an
influence on the melting point. As discussed above, confin-
ing PCM using vapor transportation is an effective method
to avoid the effect of excessive bulk residue on its melting
properties. Furthermore, using vapor transportation, the
exact value of pore filling ratio can be calculated as shown in
Fig. 4, and this is an important advantage because the exact
value of the enthalpy of fusion with the unit of J/g-PCM can
be elucidated using the accurate pore filling ratio.
Fig. 4  Pore filling ratio of PCMs confined in SBA-15s
3.3 Changes in enthalpy of fusion and melting
point of PCMs confined in silica host
which is indicated as a colored peak in Fig. 5, during the
cooling or heating steps. Therefore, these colored DSC Figure 7 shows the DSC curves of various organic PCMs
peaks, which only exist for the sample prepared using the confined in SBA-15 with different pore diameters during
liquid impregnation method, originated from a bulk state of heating.
PCM such as an external surface of SBA-15 or intergranu- The enthalpy of fusion ΔHf with the unit of J/g-PCM, and
lar space of SBA-15 grains because the host material first the melting point ­Tmelting are summarized in Table 2.
mixed with dissolved liquid PCM. DSC peaks that appeared ΔHf was calculated by using the pore filling ratio. Notably,
at a lower temperature originated from nano-confined PCM all the values of the enthalpy of fusion and melting points of
inside the mesopore of SBA-15. the confined PCMs were lower when compared with the cor-
The weight loss of both samples shown in Fig. 6 indicates responding values of the bulk PCMs. The enthalpy of fusion
the vaporization of octadecane via TGA analysis. and melting point increased with the increase in pore diameter,
The calculated weight ratio of octadecane to SBA- and became closer to the corresponding values of the bulk
15 when PCM was confined using vapor transportation PCMs. Specifically, in Fig. 7, no phase change was observed
was 66%, which is consistent with the pore filling ratio in when fatty alcohols (tetradecanol or dodecanol) were confined
Fig. 4. Further, the calculated weight ratio of octadecane to in SBA-15-5.6 nm. This can be explained in two ways: one is
SBA-15 when PCM was confined using liquid impregna- that the phase change of the confined fatty alcohol does not
tion was 109%, which is more than 100%. It can be con- occur owing to a strong interaction of PCM with the surround-
cluded that PCM confined using liquid impregnation was ing silica wall, and the other is that the enthalpy change of

13

350 Adsorption (2018) 24:345–355

Fig. 5  DSC curves of octadecane confined in SBA-15 by different methods: liquid impregnation and vapor transportation. a octadecane/SBA-
15-5.6 nm, b octadecane/SBA-15-7.6 nm, c octadecane/SBA-15-12.5 nm

where 𝛾sl represents the surface tension, ΔHf represents the


fusion enthalpy, Vm is the molar volume, Tbulk represents the
bulk melting point, 𝜃 represents the interfacial contact angle
between the nucleating phase and support surface, and r is
the radius of a pore. The Gibbs–Thomson equation shows
that an inverse of pore diameter (d−1) and the differential of
melting point ( ΔT ) have a linear relationship and the extrap-
olation of this linear relationship is converged to zero. The
Gibbs–Thomson equation is a remarkable indication, and
was used in many preceding researches (Christenson 2001;
Mitran et al. 2015). In this study, d−1 and ΔT have a linear
relationship. However, the extrapolation of d−1 and ΔT is not
converged to zero (Fig. 8a). The result of Fig. 8a was plotted
Fig. 6  TGA curves of octadecane-confined SBA-15 prepared by dif- when the pore diameters were at the scale of approximately
ferent confining method 10 nm and below.
Hence, this suggests that another appropriate parameter
should be added to the evaluation given by the Gibbs–Thom-
the phase change is too low to be detected by DSC. Also, two
son equation. Other parameters such as non-melting layer
endothermic peaks were observed when tetradecanol was con-
(Morishige et al. 2000; Christenson 2001) or surface rough-
fined in SBA-15-12.5 nm. This might be because phase transi-
ness (Tanaka et al. 2013) should be considered when phase
tion of tetradecanol occurred in 2 steps: solid–solid transition
change occurs in a nanoscale space.
and solid–liquid transition (Zeng et al. 2009). In this work, we
Figure 8b shows the plot of the enthalpies of fusion of
calculated the enthalpy of fusion of tetradecanol confined in
the confined PCMs versus the pore diameter of the host. The
SBA-15-12.5 nm by sum of the enthalpy of both peaks.
enthalpy of fusion also has a linear relationship with the pore
Typically, the difference in the melting points ( ΔT ) of the
diameter of the host. When the pore diameter increased, the
confined PCM is expressed by the Gibbs–Thomson equation
enthalpy of fusion of PCM became closer to the correspond-
(Christenson 2001) (2);
ing value of the bulk PCM. The possible reason for the lower
2Tbulk Vm 𝛾sl cos 𝜃 value of the enthalpy of fusion is the effect of the non-melt-
ΔT = (2) ing layer. A non-melting layer is composed of a few molec-
rΔHf
ular layers that freezing and melting do not occur owing
to the interaction between PCM molecules and the wall of
the host (Churaev et al. 1993). If the pore diameter of the

13
Adsorption (2018) 24:345–355 351

Fig. 7  DSC curves of a octadecane, b tetradecane, c dodecanoic acid, d decanoic acid, e tetradecanol, and f dodecanol confined in SBA-15s

Table 2  Melting properties of organic PCMs confined in SBA-15s


SBA-15 Octadecane Dodecanoic acid Tetradecanol Tetradecane Decanoic acid Dodecanol

ΔHf (J/g-PCM) 5.6 nm 39.1 19.1 – 54.6 9.0 –


7.6 nm 53.3 30.8 6.4 82.2 33.9 10.3
12.5 nm 103.3 65.9 48.4 124.2 64.3 69.5
Bulk 230 169 198 216 163 196
Tmelting (°C) 5.6 nm 9.3 − 8.86 – − 20.4 − 26.0 –
7.6 nm 11.7 3.6 6.4 − 12.4 − 4.1 − 4.1
12.5 nm 17.4 22.32 11.4 − 7.4 11.1 0.2
Bulk 26.5 44.6 36.0 6.2 30.0 24.0

host decreases, the volume of the non-melting layer inside pore is not exactly the same as that of the bulk liquid phase,
the pore increases, and consequently, the effect of the non- and this could be another possible reason for the decrease
melting layer is significant and the value of the enthalpy of in the enthalpy of fusion. From the results of Fig. 8b, it can
fusion reduces. Moreover, the state of liquid phase inside the be observed that the value of the enthalpy of fusion reaches

13

352 Adsorption (2018) 24:345–355

Fig. 8  a Experimental ΔTm val-


ues plotted as a function of pore
diameter d−1, and b the relative
enthalpy of fusion of each PCM
to bulk as a function of d 

approximately 40% that of the bulk PCM for pore diameters All six species of PCMs are shown to have a propor-
more than 10 nm. The value of enthalpy of fusion of tetrade- tional relationship to the pore diameter of the host except
cane confined in SBA-15-12.5 nm is almost 60% that of the for fatty alcohols confined in SBA-15-5.6 nm. However,
bulk PCM. According to the work by Jackson and Mckenna there is a difference between the relationship when paraf-
(1990), the enthalpy of fusion rapidly increases when the fin, fatty acids, and fatty alcohol with different functional
pore diameter d is below 10 nm, and it gradually increases groups confined in nanospace in Fig. 8. A silica wall con-
when d is in the range of approximately 10–25 nm. When d tains silanol groups. Since hydroxyl group and carboxylic
is larger than 25 nm, the value of enthalpy of fusion is satu- acid group interact more strongly with silanol group via
rated according to the reported results. From these results, it hydrogen bonding as compared to an alkane group, the
can be predicted that the enthalpy of fusion of the confined change of melting point and enthalpy of fusion of fatty
tetradecane will be saturated at ca. 60% even if d increases acids and fatty alcohol were more significant as compared
beyond 12.5 nm. Moreover, as discussed above, fully con- to paraffin. This indicates that, if the interaction with PCM
fined PCM within the mesopore is difficult to attain with molecules and the wall of host material becomes stronger,
the increase in pore diameter. Consequently, 10–20 nm is a the pore diameter will have a more significant effect on
preferable pore diameter for the confinement of PCM, result- the melting properties. Therefore, to achieve high enthalpy
ing in the achievement of high enthalpy of fusion inside the of fusion with host material containing hydrophilic walls
mesopore. such as silica, PCMs should be selected from hydrophobic
It is generally accepted that SBA-15 has not only periodic molecules such as paraffin.
mesopores but also a micropore between the mesopores.
It can be concluded that a phase change in the micropore 3.4 Characterization of soft‑templated mesoporous
does not occur owing to the effect of small pore diameter. carbons
Hence, the enthalpy of fusion with the unit of J/g-PCM cal-
culated from the weight of PCM existing only mesopores Nitrogen adsorption–desorption isotherms and pore size
can increase and further investigation is required. distribution of synthesized soft-templated mesoporous

Fig. 9  a Nitrogen adsorption


and desorption isotherm of
MC and OMC, and b pore size
distribution of MC and OMC
calculated by NLDFT method

13
Adsorption (2018) 24:345–355 353

carbon MC and OMC are shown in Fig. 9. The shape of 3.5 Changes in enthalpy of fusion and melting
adsorption branch, total amount adsorbed, pore diameter, point of PCMs confined in carbon host
and pore size distribution of MC and OMC was exactly
same, which indicates that oxidation with ozone does not The DSC curves of octadecane and tetradecanol confined in
undergo a physical change. MC and OMC, respectively, are shown in Fig. 11.
The calculated amount of surface oxygen-containing The enthalpy of fusion corresponds to the area of the
groups of MC and OMC using the Boehm titration method DSC peak, and the melting point is indicated as the point
is shown in Table 3. crossing the base line and a tangential line of the DSC peak
The amount of phenolic hydroxyl group and carboxylic in Fig. 11. The enthalpy of fusion and melting point of octa-
acid group on the surface of OMC increased via surface decane are almost the same irrespective of whether it is con-
oxidation with ozone. OMC has a 5–6 times larger amount fined in the porous structures of MC or OMC. However, a
of surface oxygen-containing groups as compared with noticeable difference can be observed between the values
MC. From the results mentioned above, MC and OMC can of the enthalpy of fusion and the melting point of tetrade-
be considered to have exactly the same parameters such as canol when confined in the pores of MC or OMC. Specifi-
pore shape, total pore volume, pore diameter, and pore size cally, with the increase in the amount of surface oxygen-
distribution with the exception of the amount of functional containing functional groups, the enthalpy of fusion of the
groups on the surface. The pore filling ratios of octadecane tetradecanol became considerably reduced. The calculated
and tetradecanol confined in MC and OMC, respectively, values of the enthalpies with the unit of J/g-PCM of confined
are shown in Fig. 10. octadecane and tetradecanol are listed in Table 4.
Table 3  Amount of the following functional groups exist on carbon ΔHf ,MC − ΔHf ,OMC represents the difference between
surface the enthalpies of fusion of MC and OMC. When octa-
decane was confined in OMC, the value of the enthalpy
Functional group Phenolic hydroxyl Carbox-
(mmol/g) ylic acid of fusion was lower only by 1.4 J/g-PCM as compared to
(mmol/g) that confined in MC. On the other hand, when tetrade-
canol was confined in OMC, the value of the enthalpy of
MC 0.047 1.086
fusion was lower by 18.5 J/g-PCM as compared to that
OMC 0.846 4.568
confined in MC. This decrease corresponds to approxi-
mately 50% of the enthalpy of fusion of tetradecanol con-
fined in MC because the amount of oxygen-containing

Table 4  The enthalpies of fusion of PCMs confined in mesoporous


carbons
PCM Host material ΔHf (J/g- ΔHf ,MC − ΔHf ,OMC
PCM) (J/g-PCM)

Octadecane MC 46.1 1.4


OMC 44.7
Tetradecanol MC 30.0 18.5
OMC 11.5
Fig. 10  Pore filling ratio of PCMs confined in mesoporous carbons

Fig. 11  DSC curves of a
octadecane and b tetradecanol
confined in OMC and MC

13

354 Adsorption (2018) 24:345–355

functional groups existing on the surface of OMC was 5–6 Acknowledgements  The authors would like to express sincere grati-
times larger than that on MC. The hydroxyl group of tet- tude to Professor Ryo Shirakashi and research associate Kiyoshi
Takano, IIS, the University of Tokyo, for their support in the experi-
radecanol interacts with the oxygen-containing groups of mental section of the DSC measurement.
carbon wall such as phenolic hydroxyl or carboxylic acid
groups. This interaction causes the tetradecanol molecules
near the wall of carbon host to be arranged irregularly, i.e.,
References
not in order. Moreover, these molecules are restricted from
moving freely. There might be an increase in the thickness Aihara, M., Nagai, T., Matsushita, J., Negishi, Y., Ohya, H.: Devel-
of the non-melting layer, and subsequently, the enthalpy opment of porous solid reactant for thermal-energy storage and
of fusion can decrease. On the contrary, in case of octa- temperature upgrade using carbonation/decarbonation reaction.
Appl. Energy 69, 225–238 (2001)
decane, which is an alkane with no functional group, the
Boehm, H.P., Diehl, E., Heck, W., Sappok, R.: Surface oxides of car-
enthalpy of fusion did not dramatically changed owing to bon. Angew. Chem. Int. Ed. 3, 669–677 (1964)
small interaction. Christenson, H.K.: Confinement effects on freezing and melting. J.
These results suggest that the interaction between a Phys. Condens. Matter 13, R95–R133 (2001)
Churaev, N.V., Bardasov, S.A., Sobolev, V.D.: On the non-freezing and
functional group on the PCM molecules and a functional
a silica surface water interlayers between ice and a silica surface.
group on the host material strongly affects the enthalpy Colloids Surf. A 79, 11–24 (1993)
of fusion of the confined PCM. If PCM molecules have Dincer, I., Dost, S., Li, X.: Performance analyses of sensible heat stor-
a functional group with strong interaction with a surface age. Int. J. Energy Res. 21, 1157–1171 (1997)
Farid, M.M., Khudhair, A.M., Razack, S.A.K., Al-Hallaj, S.: A review
functional group, reducing the amount of surface func-
on phase change energy storage: materials and applications.
tional groups on the wall of the host or substitution of PCM Energy Convers. Manag. 45, 1597–1615 (2004)
into the molecules not yet containing a strongly interacting Feng, L., Zhao, W., Zheng, J., Frisco, S., Song, P., Li, X.: The shape-
functional group are required to achieve a high latent heat. stabilized phase change materials composed of polyethylene gly-
col and various mesoporous matrices (AC, SBA-15 and MCM-
In particular, PCM, which is composed only of hydrocar-
41). Sol. Energy Mater. Sol. Cells 95, 3550–3556 (2011)
bon such as paraffin, is preferable to achieve a high latent Horstmeier, J.F., Gomez Lopez, A., Agar, D.W.: Performance improve-
heat under any conditions because it has no functional ment of vacuum swing adsorption processes for ­CO2 removal with
group to interact with. integrated phase change material. Int. J. Greenh. Gas Control 47,
364–375 (2016)
Jackson, C.L., Mckenna, G.B.: The melting behavior of organic mate-
rials confined in porous solids. J. Chem. Phys. 93, 9002 (1990)
Kadoono, T., Ogura, M.: Heat storage properties of organic phase-
4 Conclusions change materials confined in the nanospace of mesoporous SBA-
15 and CMK-3. Phys. Chem. Chem. Phys. 16, 5495–5498 (2014)
Kittaka, S., Takahara, S., Matsumoto, H., Wada, Y.: Low temperature
The melting behavior of a long-chain organic PCM in a phase properties of water confined in mesoporous silica MCM-
cylindrical mesopore of mesoporous silica and carbon host 41: thermodynamic and neutron scattering study. J. Chem. Phys.
is studied. The methods to confine PCM in the mesopore, 138, 204714 (2013)
Kruk, M.: Access to ultralarge-pore ordered mesoporous materials
pore diameter, functional groups of PCM molecules, and
through selection of surfactant/swelling-agent micellar templates.
functional groups of the wall of the host were investigated as Acc. Chem. Res. 45, 1678–1687 (2012)
the parameters that affect the melting behavior. Confinement Li, X., Li, Y.: Applications of organic phase change materials embed-
of PCM using vapor transportation is a promising method ded in adsorbents for controlling heat produced by charging and
discharging natural gas. Adsorption 21, 383–389 (2015)
to confining organic PCM only inside the mesopore. When
Libbrecht, W., Deruyck, F., Poelman, H., Verberckmoes, A., Thybaut,
the pore diameter decreases, the influence of the non-melting J., De Clercq, J., Van, P., Voort, D.: Optimization of soft templated
layer on the melting behavior becomes significant, and leads mesoporous carbon synthesis using definitive screening design.
to the decrease in the enthalpy of fusion of PCM. Moreover, Chem. Eng. J. 259, 126–134(2015)
Min, X., Fang, M., Huang, Z., Liu, Y., Huang, Y., Wen, R., Qian, T.,
when the interaction between PCM molecules and the wall
Wu, X.: Enhanced thermal properties of novel shape-stabilized
of the host material is strong (e.g., hydrogen bonding), the PEG composite phase change materials with radial mesoporous
PCM molecules cannot be aligned in order and their motion silica sphere for thermal energy storage. Sci. Rep. 5, 12964 (2015)
is strictly limited, resulting in the decrease in the enthalpy Mitran, R.A., Berger, D., Munteanu, C., Matei, C.: Evaluation of dif-
ferent mesoporous silica supports for energy storage in shape-
of fusion. Consequently, by selecting the host materials with
stabilized phase change materials with dual thermal responses. J.
mesopores of 10–20 nm in diameter and PCM molecules Phys. Chem. C 119, 15177–15184 (2015)
with a functional group that weakly interacts with the wall Miyahara, M., Gubbins, K.E.: Freezing/melting phenomena for Len-
of the host such as paraffin, enthalpy of fusion more than nard-Jones methane in slit pores: a Monte Carlo study. J. Chem.
Phys. 106, 2865 (1997)
50% of the value of the heat generated by the bulk PCM can
Mondal, S.: Phase change materials for smart textiles—an overview.
be obtained. Appl. Therm. Eng. 28, 1536–1550 (2008)

13
Adsorption (2018) 24:345–355 355

Morishige, K., Iwasaki, H.: X-ray study of freezing and melting of Tanaka, H., Hiratsuka, T., Nishiyama, N., Mori, K., Miyahara, M.T.:
water confined within SBA-15. Langmuir 19, 2808–2811 (2003) Capillary condensation in mesoporous silica with surface rough-
Morishige, K., Kawano, K., Hayashigi, T.: Adsorption isotherm and ness. Adsorption 19, 631–641 (2013)
freezing of Kr in a single cylindrical pore. J. Phys. Chem. B. 104, Valange, S., Palacio, R., Charmot, A., Barrault, J., Louati, A., Gabelica,
10298–10303 (2000) Z.: Nanoparticles of ­Fe2O3 inserted in SBA-15 silica at micropore
Nomura, T., Zhu, C., Sheng, N., Tabuchi, K., Sagara, A.: Shape-stabi- mouth level: an experimental evidence of the confinement effect.
lized phase change composite by impregnation of octadecane into J. Mol. Catal. A 305, 24–33 (2009)
mesoporous ­SiO2. Sol. Energy Mater. Sol. Cells 143, 424–429 Wang, J., Yang, M., Lu, Y., Jin, Z., Tan, L., Gao, H., Fan, S., Dong, W.,
(2015) Wang, G.: Surface functionalization engineering driven crystal-
Sari, A.: Form-stable paraffin/high density polyethylene composites lization behavior of polyethylene glycol confined in mesoporous
as solid-liquid phase change material for thermal energy storage: silica for shape-stabilized phase change materials. Nano Energy
preparation and thermal properties. Energy Convers. Manag. 45, 19, 78–87 (2016)
2033–2042 (2004) Zeng, J.L., Cao, Z., Yang, D.W., Xu, F., Sun, L.X., Zhang, L., Zhang,
Sari, A., Alkan, C., Bilgin, C.: Micro/nano encapsulation of some X.F.: Phase diagam of palmitic acid-tetradecanol mixtures
paraffin eutectic mixtures with poly(methyl methacrylate) shell: obtained by DSC experiments. J. Therm. Anal. Calorim. 95,
preparation, characterization and latent heat thermal energy stor- 501–505 (2009)
age properties. Appl. Energy 136, 217–227 (2014) Zhao, D., Feng, J., Huo, Q., Melosh, N., Fredrickson, G.H., Chmelka,
Scheibe, B., Borowiak-palen, E., Kalenczuk, R.J.: Oxidation and reduc- B.F., Stucky, G.D.: Triblock copolymer syntheses of mesoporous
tion of multiwalled carbon nanotubes—preparation and charac- silica with periodic 50 to 300 angstrom pores. Science 279, 548–
terization. Mater. Charact. 61, 185–191 (2009) 552 (1998)
Schreiber, A., Ketelsen, I., Findenegg, G.H.: Melting and freezing of Zhou, D., Zhao, C.Y., Tian, Y.: Review on thermal energy storage with
water in ordered mesoporous silica materials. Phys. Chem. Chem. phase change materials (PCMs) in building applications. Appl.
Phys. 3, 1185–1195 (2001) Energy 92, 593–605 (2012)

13

You might also like