You are on page 1of 19

Sustainable Cities and Society 51 (2019) 101767

Contents lists available at ScienceDirect

Sustainable Cities and Society


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

Experimental charging and discharging performance of alumina enhanced T


pentaerythritol using a shell and tube TES system

K.P. Venkitaraja,b, S. Suresha, , B. Praveena
a
Department of Mechanical Engineering, NIT Tiruchirappalli, Tiruchirappalli 620015, India
b
Department of Mechanical Engineering, College of Engineering, Adoor 691551, India

A R T I C LE I N FO A B S T R A C T

Keywords: The work presented in this paper is the results of an experimental study conducted on the thermal energy storage
Pentaerythritol (TES) performance of pentaerythritol (PE) added with alumina (Al2O3) nanoparticles. The charging and dis-
Solid–solid PCM charging experiments are performed using PE added with 0.5 and 1.0 wt.% of Al2O3 in a shell and tube type
Al2O3 nanoparticles thermal energy storage system. The experimental data are analyzed to obtain the charging/discharging effi-
Charging and discharging
ciencies and the overall efficiency of the TES system at different flow rates 2, 4, 6 LPM of the heat transfer fluid
Energy efficiency
Exergy efficiency
(HTF, Therminol-55). Exergy analysis based on the 2nd law of thermodynamics is also carried out in this ex-
perimental work. The results indicated that charging and discharging time is reduced due to the addition of
Al2O3 with PE at all volume flow rates of HTF. The maximum charging and discharging powers of 289.3 W and
230.8 W respectively are observed in the case of PE + 1.0% Al2O3 corresponding to the flow rate of 6 LPM. The
efficiencies of charging and discharging showed maximum values of 86.8% and 75.0% respectively at 6 LPM.
The overall energy efficiency of the thermal energy storage system found increased from 38.3% obtained in the
case of PE to 50.5% and 58.5% obtained for PE added with 0.5 wt.% and 1.0 wt.% of Al2O3 nanoparticles
respectively.

1. Introduction transformation involves latent heats comparable to the most effective


solid/liquid PCM. Several materials show the solid–solid reversible
The latent heat storage method using phase change materials (PCM) phase transition behavior, but only a handful has sufficient latent heat
have been employed for various applications such as solar, medical, to be a possible latent heat storage material. These solid–solid PCM are
textile, electronic, energy efficient buildings, air conditioning, in- categorized into two types, namely, organic solid–solid PCM and in-
dustrial and aerospace, etc. (Nkwetta & Haghighat, 2014). Extensive organic solid–solid PCM (Busico, Carfagna, Salerno, Vacatello, &
work is going on the application of PCM for thermal energy storage in Fittipaldi, 1980; Gu, Xi, Cheng, & Niu, 2010; Jiang, Ding, & Li, 2002;
heat exchangers (Anish, Mariappan, & Suresh, 2019; Seddegh, Joybari, Landi & Vacatello, 1975; Li et al., 1999; Li & Ding, 2007; Pielichowska
Wang, & Haghighat, 2017) and energy efficient buildings (Arcuri, & Pielichowski, 2010; Qi & Liu, 2006; Ruan, Li, & Hu, 1995; Ruiyun,
Spataru, & Barrett, 2017; Mourid, El Alami, & Kuznik, 2018). So- Dejun, Xian, & Jing, 1990; Xi, Gu, Cheng, & Wang, 2009).
lid–liquid PCM which are mostly used in the field of thermal energy Organic solid–solid PCM mainly comprises a particular group of
storage and heat transfer applications suffer from serious limitations hydrocarbon compounds, polyhydric alcohols, and polymers.
like expansion and leakage in their liquid phase. An easy way to avoid Polyalcohols or polyols are hydrocarbon compounds that have body-
these limitations is to use a PCM that undergoes a solid–solid phase centered tetrahedral molecular structure at low temperatures. These
transition. Solid–solid PCM undergo a solid to solid phase transition polyalcohols and their amine derivatives show a solid-state crystalline
which is associated with absorption and release of a large amount of transition because of the presence of metastable, vibrant hydrogen
heat. They change their crystalline structure from one lattice type to bonds between molecules in their crystal lattice. When the temperature
another at a definite temperature (Cao, Ye, & Yang, 2013) and this rises, these bonds break and the crystal structure changes from an

Abbreviations: PCM, phase change material; HTF, heat transfer fluid; PE, pentaerythritol; DSC, Differential Scanning Calorimetry; LPM, liter per minute; TES,
thermal energy storage; SSPCM, solid–solid PCM; TEM, transmission electron microscope; SEM, scanning electron microscope; XRD, X-ray diffraction; EDX, Energy
Dispersive X-ray; FESEM, field emission scanning electron microscope

Corresponding author at: Department of Mechanical Engineering, National Institute of Technology Tiruchirappalli, India.
E-mail addresses: venkitaraj@cea.ac.in (K.P. Venkitaraj), ssuresh@nitt.edu (S. Suresh), praveennitt15@gmail.com (B. Praveen).

https://doi.org/10.1016/j.scs.2019.101767
Received 6 April 2019; Received in revised form 7 July 2019; Accepted 5 August 2019
Available online 12 August 2019
2210-6707/ © 2019 Elsevier Ltd. All rights reserved.
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Notations and symbols Pc average charging power (W)


Pd average discharging power (W)
T temperature (°C) tc time for charging (min/s)
cp,p specific heat of the PCM (kJ/kg K) td time for discharging (min/s)
mp mass of the PCM (kg) ηc charging efficiency (%)
Qsp heat supplied (kJ) ηd discharging efficiency (%)
Qst heat stored (kJ) η overall energy efficiency (%)
Qre heat rejected (kJ) Th,i inlet temperature of hot HTF (°C)
Qsen sensible heat (kJ) Th,o outlet temperature of hot HTF
Qtrs solid–solid transition heat (kJ) T̄p,i mean initial temperature of PCM (°C)
ṁh mass flow rate hot fluid (kg/s) T̄p,f mean final temperature of PCM (°C)
ṁc mass flow rate of cold fluid (kg/s) Tc,o outlet temperature of cold HTF (°C)
cp,h specific heat of hot fluid (kJ/kg K) Tc,i inlet temperature of cold HTF (°C)
cp,c specific heat of cold fluid (kJ/kg K)

ordered tetrahedral phase to a disordered cubic phase accompanied tested composite PCM comprising carbonate-salt for medium and high-
with the absorption of a significant amount of energy, more significant temperature energy storage applications. Li et al. (2017) investigated
the disorder, more substantial is the amount of energy absorbed. Now, calcium chloride hexahydrate with aluminum oxide nanoparticles for
when the temperature is lowered, at the transition temperature, the thermal energy storage. Singh et al. (2017) experimentally studied the
hydrogen bonds reform and the materials regain its original tetrahedral thermal energy storage performance of Myo-Inositol based nano PCM.
structure which causes the stored thermal energy to get released (Singh They used Myo-Inositol added with 1%, 2% and 3% mass fractions of
et al., 2015). Polyalcohols also called as “plastic crystals,” which dis- CuO and Al2O3 nanoparticles. Based on the results of Differential
play crystalline disorder at elevated temperatures accompanied by the Scanning Calorimetry (DSC), Thermogravimetric analysis (TGA), and
absorption of thermal energy (Timmermans, 1961). Among poly- Fourier Infrared Transforms (FTIR) analysis, they commented that the
alcohols, the most exciting materials are pentaerythritol (PE), penta- nano-enhanced myo-inositol is a potential PCM for thermal energy
glycerine (PG), Neopentyl glycol (NPG) and Neopentyl alcohol (NPA). storage.
Certain amine derivatives of polyalcohols such as Tris (hydroxymethyl) Though there are numerous investigations reported on solid–liquid
aminomethane (TAM) and Aminoglycol (AMPL) were also very tested PCM, the papers on thermal energy storage performance of SSPCM with
and proved by various researchers as potential solid–solid PCM for heat transfer enhancement additives are decidedly less in number. Hu,
thermal energy storage applications (Benson, Burrows, & Webb, 1986; Zhao, Jin, and Chen (2014) studied the solid-state phase transition of
Murrill & Breed, 1970). PE added with nano-aluminum nitride. In our recent work (Venkitaraj
Most organic and some inorganic PCM have an inherently low & Suresh, 2019) we studied the thermal and chemical stability of PE
thermal conductivity which leads to poor charging and discharging added with 0.1 wt.% of metal oxide nanoparticles. We found that the
performances. Improving the thermal conductivity of PCM by disper- nanoparticles decreased the subcooling found during the cooling cycle.
sing thermally conductive materials is more widely used for enhancing We have earlier investigated the energy storage characteristics and
the effectiveness of the PCM-based TES systems. The majority of PCM, crystallization kinetics of PE added with indium (Venkitaraj & Suresh,
in general, has a very poor thermal conductivity which results in lower 2018).
charging and discharging rates. The literature reported the use of The experimental heat transfer analysis of PCM involves conducting
thermally conductive particles like nanoparticles as additives to en- charging and discharging performances of the PCM for evaluating the
hance the thermal energy storage performance of organic and inorganic energy efficiency of the thermal energy storage system. Numerous ex-
PCM. The research findings also indicated that the enhancement of perimental and numerical studies on the thermal performance of latent
thermal conductivity accompanied by a slight decrease in the phase heat storage (LHS) system have been reported in the literature. Among
change properties of PCM (Adorno & Silva, 2006). Siegel (1977) in- the various works reported, a sizeable number of research works dis-
vestigated the use of high conductivity particles to enhance the rate of cusses the various improvements in the thermal performances of the
solidification molten salt. He conducted an analysis to determine how LHS system. This includes geometrical design (Khan, Khan, & Ghafoor,
the solidification rate is influenced by the introduction of high con- 2016; Khodadadi & Zhang, 2001; Medrano et al., 2009), using metal
ductivity particles into a solidifying low conductivity material. He fins (Liu, Sun, & Ma, 2005), dispersion of thermally conductive particles
concluded that for reasonable concentrations such as 20% particles by (Liu, Su, Tang, & Fang, 2016), use of form stabled and encapsulated
volume, the heat removal rate for a plane geometry can be increased by PCM (Giro-Paloma, Martinez, Cabeza, & Fernandez, 2016; Liu, Rao,
10–20% depending on the ratio of particle to matrix conductivity. Zhao, Huo, & Li, 2015). Researchers have been trying different geo-
Cabeza, Mehling, Hiebler, and Ziegler (2002) and Py, Olives, and metric configurations of the LHS system for performing the charging
Mauran (2001) studied PCM/graphite enclosed in waterlogged metal and discharging processes. Khodadadi and Zhang (2001) numerically
modules. They observed a heat barrier between the metal modules and investigated the effect of convection on the melting process using a
water. Yin, Gao, Ding, and Zhang (2008) experimented paraffin/ex- spherical container. They observed a higher rate of heat transfer at the
panded graphite composite and reported that the thermal conductivity upper side of the sphere than at the lower side. Medrano et al. (2009)
enhanced about 17 fold compared to the thermal conductivity of pure conducted experimental heat transfer study of the PCM embedded in a
paraffin. Kim and Drzal (2009) reported that the presence of conductive graphite matrix using five different types of heat exchangers and they
graphite affects the phase change properties of PCM. Elgafy and Lafdi observed greater heat transfer in the case of pipe in pipe type heat
(2005) studied the performance enhancement of paraffin/carbon na- exchanger. Liu et al. (2005) studied the thermal energy storage analysis
nofibers (CNF) with the mass fraction of CNF varied between 1% and of stearic acid using vertical annulus storage system. They investigated
4%. Teng and Yu (2012) prepared composite PCM comprising paraffin the dependence of Reynolds number on the heat transfer performance
and 1.0, 2.0, and 3.0 wt.% of Al2O3, TiO2, and SiO2 nanoparticles. Their of stearic acid.
study revealed that TiO2 enhances the thermal performance of paraffin As mentioned in the discussion above, the geometric configuration
more efficiently compared to other nanoparticles. Ge et al. (2014) of the heat exchanger in TES is a very important matter. Among the

2
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

different types of heat exchangers, shell and tube type is very much a very little effect on the thermal performance due to the conduction
used by various researchers due to its simple installation requirements dominant heat transfer during the charging and discharging process.
and better heat transfer performance (Agyenim, Hewitt, Eames, & The result of varying the number and orientation of HTF tubes in using
Smyth, 2010). The numerical work reported by Wang, Zhang, Wang, a horizontal shell and tube heat exchanger on thermal performance was
and He (2013) involved a shell and tube heat exchanger to study the numerically studied by Luo, Yao, Yi, and Tan (2015). They observed
thermal energy storage performance of n-octadecane. Their results re- better thermal performance in the case of multi-tube arranged sym-
vealed that the charging time reduces significantly with the increase in metrically with respect to the center.
the inlet temperature of the HTF. They also reported that the charging The results of the experimental investigations using shell and tube
time reduces with the increase in the mass flow rate of the fluid. The heat exchangers can be further improved by the use of extended sur-
experimental study conducted by Ezan, Ozdogan, and Erek (2011) faces (fins) or by enhancing the thermal conductivity of PCM by adding
discussed the thermal storage performance of a shell and tube heat thermally conductive particles. Darzi, Jourabian, and Farhadi (2016)
exchanger using water as the HTF. They reported the effects varying the performed a numerical study of thermal performance of n-eicosane
inlet temperature, flow rate, shell diameter and shell material on the using a longitudinally finned horizontal shell and tube heat exchanger.
thermal performance of the system. They observed that the charging They reported an increase in the solidification when fins were em-
and discharging processes are dominated by free convection. The effect ployed. The results of the experimental and numerical study conducted
varying the inlet temperature on the thermal performance is more by Meng and Zhang (2017) for testing the thermal performance of
compared to the effect of varying the flow rate of the HTF. The nu- paraffin-copper foam composite using a rectangular shaped shell and
merical study conducted by Seddegh, Wang, and Henderson (2016) tube heat exchanger was used. They indicated the dependence of soli-
reported the thermal storage performance of paraffin using a shell and dification time on the temperature gradient between the HTF and PCM.
tube system. They tested the storage performance of the LHS system The experimental paper published by Pandiyarajan, Pandian,
corresponding to the horizontal and vertical orientation of the shell and Malan, Velraj, and Seenira (2011) discussed the thermal energy storage
tube heat exchanger and concluded that the geometrical orientation has performance of a finned shell and tube heat exchanger for IC engine

Fig. 1. (a) Molecular structure of PE molecule, (b) XRD pattern of PE, (c) SEM image of PE, and (d) photograph of PE.

3
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

exhaust heat recovery. Their study indicated about 10–15% of waste using pentaerythritol as the PCM. The objective of the work was to
heat recovery using the finned shell and tube TES system. They sug- determine the influence of the PCM to fins volume ratio on the per-
gested a cascade LHS system using multiple PCM for increasing the % formance of the LHS unit for cooking. They concluded that the charging
heat recovery. Khan and Khan (2017) investigated the charging and and discharging performances decrease with PE to fin volume ratio. In
discharging of paraffin using a longitudinally finned shell and tube heat one of our recent work (Venkitaraj, Suresh, & Venugopal, 2018), we
exchanger for domestic and industrial applications. They investigated experimentally investigated the thermal performance of pentaerythritol
the effect of inlet temperature and volume flow rate of HTF on phase added with Al2O3 nanoparticles for IC engine exhaust heat recovery
transition rate and mean power. by regulating the inlet temperature or applications. The experimental results showed that the addition of 0.1%
volume flow rate of HTF, the influence of overall thermal resistance was and 0.5% weight fractions of Al2O3 results in 36.36% and 73.39% in-
minimized. Mean discharge power is enhanced by 36.05% as the inlet crease in the amount of heat recovered compared to pure PE. The
temperature is reduced from 15 °C to 5 °C. Likewise, the mean discharge present work reports the energy and exergy analysis of charging and
power is improved by 49.75% as the volume flow rate was increased discharging processes using PE added with 0.5 and 1.0 wt.% of Al2O3
from 1.5 LPM to 3 LPM. They also observed that with an increase in using Therminol-55 as the heat transfer fluid for charging and dis-
volume flow rate, the discharge time of the equal amount of thermal charging of PE. A shell and tube heat exchanger consisting of three
energy 12.09 MJ was reduced by 24%. It as, therefore, concluded that horizontal circular tubes was used as the TES for conducting the energy
by adjusting operating conditions, the required demand for output storage and discharge experiments. As far as the knowledge of the au-
temperature and mean discharge power can be attained. Agarwal and thors concerned, there is no research work that performed the energy
Sarviya (2016) experimentally evaluated the thermal and heat transfer and exergy analysis of a solid–solid PCM based TES using therminol as
characteristics of paraffin using a shell and tube type LHS system during the HTF is reported in the literature.
charging and discharging processes. They used air as heat transfer fluid
(HTF) and determined the effectiveness of the LHS for drying of food
products. 2. Material preparation
Majority of energy storage applications involve the use of solid–li-
quid phase change materials for thermal energy storage. However, the Pentaerythritol [PE] is a polyalcohol having the chemical formula
solid–liquid PCM have disadvantages of volume expansion and leakage C5H12O4. PE is an organic solid–solid PCM that exhibits a structural
issues in their liquid phase. These limitations do not exist in the case of transition from tetragonal to cubic at a fixed temperature in the range
solid–solid PCM since these type of PCM exhibit structural changes 187–189 °C. At ordinary temperature, PE has the body-centered tetra-
involving enthalpy change in their solid state. Polyalcohols are a class gonal (BCT) molecular structure consisting of four C–CH2–OH frag-
of organic solid–solid PCM having very high enthalpy change of tran- ments (Fig. 1(a)). In a PE molecule, the middle carbon atom is tetra-
sition. But, being organic materials, the thermal conductivity of poly- gonally surrounded by four other carbon atoms. The PE molecules are
alcohols is low causing poor charging and discharging performances. arranged flat in a sheet parallel to the crystallographic c-plane. The
Therefore, to improve the heat transfer in organic PCM, thermally transition to the face-centered cubic (FCC) crystalline structure is ac-
conductive particles are employed. Nanoparticles are one type of heat companied by the absorption of the energy required for the breaking of
transfer enhancement additives that are widely used by various re- hydrogen bonds. During the phase transition, the hydrogen bonds break
searchers. Most of the experimental investigations reported in the lit- resulting in a change in the crystal structure. The separated molecules
erature discussed the charging and discharging performances of so- are held in their lattice sites by dispersion forces. During the dischar-
lid–liquid PCM. Again, to improve the thermal performance of the ging period, the hydrogen bonds reform by rejecting the stored energy.
system, the researchers made use of finned type shell and tube based Fig. 1(b) shows the X-ray diffraction (XRD) pattern of PE obtained using
TES. The papers published on the charging and discharging perfor- Bruker AXS D8 Advance X-ray diffractometer having Cu-kα1 radiation
mances of solid–solid PCM are limited. Some of the papers available in in the range of 0–90°. The reflections in the XRD pattern obtained were
literature reports study on the characterization, charging and dischar- recognized as matching to the tetragonal phase of PE using JCPDS
ging performances of pure solid–solid PCM. But, there is very little (Joint Committee on Powder Diffraction Standards). The Scanning
research work that studied the effect of heat transfer enhancement Electron Microscope (SEM) analysis was performed using Jeol JSM
additives on charging and discharging performances of polyalcohols is 6390LV SEM to study the microstructure of PE. The obtained SEM
reported in the literature. The work presented in this thesis is aimed at image showed that PE has a loose microstructure with a large number
filling this gap found in the literature. of individual lamellae on the surface (Fig. 1(c)). Pentaerythritol (ana-
NKhonjera et al. (2016) experimentally studied the charging and lytical reagent grade, purity 98.0%) in white crystalline powder form
discharging performance of rectangular type TES provided with fins was supplied by Alfa Aesar, USA (Fig. 1(d)).
The alumina nanoparticles supplied by Alfa Aesar, USA were used as

Fig. 2. (a) XRD pattern and (b) SEM image of Al2O3.

4
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

the additive to pentaerythritol in this experimental study. The specifi- the heterogeneous mixing of pentaerythritol and alumina was the fact
cations of the nanoparticles are as follows: alumina (Al2O3), NanoDur that the dispersion of Al2O3 nanoparticles was purely physical using a
99.5%, APS: 40–50 nm, SA: 32–40 m2/g, MW: 101.96. The XRD pattern ball-mill. The region circled in red color indicates the particle ag-
of Al2O3 nanoparticles confirmed its crystalline nature. The diffraction gregation and the area shown by blue arrow represent the region that is
peaks in the XRD pattern of Al2O3 nanoparticles (Fig. 2(a)) were deprived of nanoparticles. Fig. 4(b) and (c) shows the EDX spectra of
identified as those present in the case of hexagonal α-Al2O3. Fig. 2(b) the PE added with Al2O3 nanoparticles before and after thermal cycling.
exhibits the microstructure of alumina obtained using Jeol JSM 6360 The elemental mapping is shown in Fig. 4(b) also indicated the het-
Scanning Electron Microscope. erogeneous mixing of Al2O3 nanoparticles with PE. The EDX spectrum
The dimension, profile, growth, and distribution of the Al2O3 na- of the thermally cycled sample shown in Fig. 4(c) confirms that the
noparticles were studied by transmission electron microscopic (TEM) dispersion of alumina nanoparticles in PE was not much disturbed even
images. The images were obtained using Jeol/JEM 2100 having point after several thermal cycles.
and lattice resolution of 0.23 nm, and 0.14 nm respectively, and mag- Fig. 5 shows the Fourier Transform Infrared (FTIR) spectrum of the
nification ranging from 2000× to 1,500,000×. The TEM photographs PE + 0.5 wt.% Al2O3 before and after thermal cycling. Since the spectra
of the Al2O3 nanoparticles are shown in Fig. 3. The obtained images of the uncycled and cycled samples are identical, it can be confirmed
indicated the spherically shaped alumina nanoparticles have particle that there involved no chemical reaction between the PCM and the
sizes ranging from 14 to 58 nm. nanoparticles.
The alumina nanoparticles were added to pentaerythritol in 0.5%
and 1% weight fractions. The dispersion of nanoparticles in PCM was
ensured with the help of the low energy ball mill run at 200 rpm for 3. Thermal conductivity measurement
1.5 h. The dispersion stability of nanoparticles and particle aggregation
in the prepared PCM samples were analyzed by a field emission scan- The thermal conductivity of the PCM samples was determined by T-
ning electron microscope (FESEM) system to obtain the topographical history method (Venkitaraj & Suresh, 2018). The T-history analysis is
and elemental information of the PE/Al2O3 composite PCM. Elemental based on lumped capacitance theory in which a uniform temperature
identification with mapping to study the dispersion of nanoparticles in profile is maintained. In order to ensure that the lumped heat capacity
PE was carried out by Energy Dispersive X-ray (EDX) mapping. theory is applicable, the test conditions used in T-history analysis
Fig. 4(a) shows the FESEM image of PE added with Al2O3 nano- should have of Biot number (Bi) value less than unity. In T-history
particles. The image shows nanoparticle agglomeration at some places method, the variation of temperature of the test samples and the re-
with some areas without any presence of nanoparticles. The reason for ference sample with time during the cooling process is recorded. The
thermal conductivity values of the test samples are then determined by

Fig. 3. TEM images of Al2O3 nanoparticles.

5
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 4. (a) FESEM image, (b) EDX spectrum of PE with Al2O3 before thermal cycling, and (c) EDX spectrum of PE with Al2O3 after thermal cycling.

instrument had a temperature range from −130 °C to max. 450 °C and a


measurement range ± 350 mW at RT with a measurement resolution of
0.04 mW at room temperature, temperature accuracy of ± 0.2 °C and
temperature reproducibility of ± 0.1 °C. The PCM samples were sub-
jected to heating and cooling between the temperature range 30–280 °C
at 10 °C/min.

5. Charging and discharging performance

5.1. Experimental setup

The performance of pentaerythritol added with Al2O3 nanoparticles


in energy storage application was tested by conducting charging and
discharging experiments using a thermal energy storage system. A
schematic diagram of the experimental setup is shown in Fig. 6. The
experimental setup comprises a shell and tube type heat exchanger
filled with the PCM. The shell side of the heat exchanger filled the PCM.
Fig. 5. FTIR spectrum of PE added with Al2O3 nanoparticles. The fluid flowing through the copper tubes exchanges heat with the
PCM during the charging and discharging process. The hot thermal
fluid from the hot fluid container is pumped through the heat exchanger
comparing their cooling curves with the cooling curves of the reference
using a gear pump. The flow rate of the hot thermal fluid is controlled
samples. The detailed experimental requirements and the procedure of
using a bye pass valve provided in the delivery side of the gear pump. A
T-history method are given in our previously published article
rotameter is provided to monitor the flow rate of hot therminol flowing
(Venkitaraj & Suresh, 2018).
through the heat exchanger. The hot fluid after flowing through the
copper tubes flows back to the hot reservoir. During the flow through
4. Determination of enthalpy of transition the heat exchanger, the hot fluid loses its heat to the PCM surrounding
the copper tubes. This is the charging cycle of the experiment. The cold
Thermal property analysis in terms of latent heat capacity and phase line of the experimental setup consists of a cold fluid container cen-
transition temperature for pure and nano-enhanced PE were de- trifugal pump; rotameter, and an air-cooled heat exchanger.
termined using DSC (Mettler Toledo DSC 822e, Hong Kong). The During the discharge cycle, the cold thermal fluid is pumped

6
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 6. Schematic diagram of the experimental setup for heat transfer study.

through the heat exchanger using the centrifugal pump. The flow rate exchanger. The heat exchanger comprises three tubes of 20 mm inner
of the cold fluid is monitored and controlled by using a rotameter and a diameter, 2 mm thickness, and 45 cm effective length enclosed within a
valve provided in the delivery side of the centrifugal pump. The cold shell of 65 mm inner diameter, and 5 mm thickness. The tubes are made
fluid flowing through the copper tube receives heat stored by the PCM. of copper and the shell is made of mild steel. The shell insulated using
The cold fluid leaving from the other end of the heat exchanger flows glass wool and asbestos in order to minimize the heat loss to the sur-
back to the cold reservoir via an air-cooled heat exchanger. During the roundings. The quantity of PCM required for filling in space inside the
flow through the radiator, it loses heat. The temperature of the PCM shell estimated to be 1.22 kg. The heat transfer fluid used in the hot and
inside the shell at six different locations is also monitored using the K- cold circuits is Therminol-55. Some important properties of Therminol-
type thermocouples. The temperature of the PCM inside the shell at the 55 are shown in Table 1. The temperature of therminol oil maintained
leading end, middle and, trailing ends are also monitored using the K- at 200 °C by using two hand immersion heaters of 1500 W each. The
type thermocouples. A multi-channel data logger records the PCM and gear pump (Make: HMP pumps, power: 0.5 HP, size: ½″ × ½″, rpm:
the fluid temperatures during the charging and discharging process. 1440, maximum discharge: 30 LPM) was used to pump the hot ther-
The photograph of the actual experimental setup fabricated in- minol oil through the heat exchanger. The cold fluid pumped by using a
cluded as Fig. 7. Fig. 8 shows the photographs of the shell and tube heat monoblock centrifugal pump (Make: Lakshmi pumps, power: 0.5 HP,
exchanger before and after insulating. size: 1″ × 1″, rpm: 1400, max. discharge: 45 LPM). The flow rate of hot
Fig. 9 shows the constructional features of the shell and tube heat therminol oil is measured using a metal tube magnetic rotameter

Fig. 7. The photograph of the actual experimental setup.

7
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 8. Shell and tube heat exchanger.

(Make: Eureka, model: SSVS-MTS-4, range: 1.1–11 LPM). The flow rate Table 1
in the cold side of the experimental setup measured by using an acrylic Properties of Therminol-55.
body rotameter (Make: flow point, range: 1–10 LPM). Two metal con- Property Value
tainers, each 20-l capacity, were used as hot and cold fluid reservoirs.
The entire hot circuit, part of the cold circuit between the heat ex- Appearance Clear, yellow liquid
changer exit and air-cooled heat exchanger inlet are well insulated to Composition Synthetic hydrocarbon
Average molecular weight 320
minimize the heat loss to the surroundings. Density 864 kg/m3 @33 °C
815 kg/m3 @220 °C
Kinematic viscosity 19 mm2/s @ 40 °C
5.2. Procedure
Specific gravity 0.876
Specific heat 1.95 kJ/kg K @33 °C
The experiments performed using pure PE and PE added with 0.5 2.21 kJ/kg K @220 °C
and 1.0 wt.% of Al2O3. The detailed experimental procedure is given Normal boiling point 351 °C
below. The heat exchanger filled with pure PE on the shell side Maximum use temperature 315 °C
mounted in the experimental setup. The heaters in the hot fluid re-
servoir are switched on to heat the therminol oil above 200 °C. The hot
the heat exchanger by starting the centrifugal pump. The flow rate of
therminol oil pumped through the heat exchanger by switching on the
the cold therminol adjusted to 2 LPM by operating the valve. The
gear pump. The flow rate of the oil set at 2 LPM by regulating the bye
temperature of the cold therminol entering and leaving the heat ex-
pass valve. The temperature of the hot therminol entering and leaving
changer and the PCM temperatures at different locations recorded.
the heat exchanger and the PCM temperatures at different locations
The centrifugal pump switched off when the PCM temperature re-
recorded. The gear pump switched off when the PCM temperature re-
stored to its ambient temperature. This completes the discharging
corded crosses the solid–solid transition temperature of the PCM. This
process. The experiment repeated for hot and cold therminol flow rates
completes the charging process. The cold therminol oil pumped through

Fig. 9. Constructional details of the heat exchanger.

8
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

of 4 LPM and 6 LPM. The experiment is repeated for PE/LMA composite part of the exergy is stored in the PCM. The rate of exergy supplied by
PCM of 0.5 and 1.0 wt.% Al2O3. HTF during the charging period is computed using the following
equation.
5.3. Data reduction
T
˙ input = m˙ h cp,h ⎡ (Th,i − Th,o) − To ln ⎛⎜ h,i ⎞⎟ ⎤
Ex ⎢ ⎥
5.3.1. Energy analysis ⎣ ⎝ Th,o ⎠ ⎦
The heat supplied to the PCM is given by,
where To is the ambient temperature.
Qsp = m˙ h c p,h (Th,i − Th,o) tc, kJ (1) The rate of exergy stored in the PCM is obtained as
where ṁh is the mass flow rate (kg/s) of hot HTF, cp,h is the specific heat
of hot HTF, Th,i and Th,oare the inlet and the exit temperature of the hot ˙ stored = m p htrs ⎡1 − To ⎤ + m p c p,p ⎡ (Ttrs − T¯p,i ) − To ln ⎛⎜ Ttrs ⎞⎟ ⎤
Ex
⎢ Ttrs ⎥ ⎢ ¯ ⎥
HTF, and tc is the charging time. ⎣ ⎦ ⎣ ⎝ Tp,i ⎠ ⎦
The heat stored in the TES is calculated as Heat stored = sensible ¯
+ m p c p,p ⎡ ¯ ⎛ Tp,f ⎞ ⎤
heat + solid–solid transition heat. ⎢ (Tp,f − Ttrs) − To ln ⎜ Ttrs ⎟ ⎥
⎣ ⎝ ⎠⎦
Qst = Qsen + Qtrs = m p c p,p (T¯p,f − T¯p,i ) + m p htrs , kJ (2)
where Tp,i and Tp,f are the initial and final PCM temperatures and Ttrs is
mp – the mass of the PCM (pentaerythritol), T̄p,f – mean final PCM the PCM transition temperature.
temperature, T̄p,i – mean initial temperature of the PCM, htrs – enthalpy The rate of exergy extracted from the PCM by the HTF during the
of transition, cp,p is the specific heat of PCM. discharging period is computed using the equation,
The average charging power (Pc) is obtained as the ratio of the heat
T
stored (Qst) to the time for charging (tc). ˙ output = m˙ c cp,c ⎡ (Tc,o − Tc,i ) − To ln ⎛⎜ c,o ⎞⎟ ⎤
Ex ⎢ ⎥
Qst ⎣ ⎝ Tc,i ⎠ ⎦
i.e. Pc =
tc (3) The charging exergy efficiency (ψchar) is computed as the ratio of rate of
Now, charging efficiency is found as the ratio of the heat stored (Qst) by exergy stored by the PCM to the exergy supplied by the HTF during the
the PCM to the heat supplied (Qsp) by the HTF during the charging charging period. Similarly, discharging exergy efficiency (ψdischar) is
cycle. i.e., estimated as the ratio of the rate of exergy extracted from the PCM to
the exergy stored by the PCM.
Qst The overall exergy efficiency (ψ) of the thermal energy storage
Charging efficiency (%), ηc = × 100
Qsp (4) process is computed from the exergy efficiencies of charging and dis-
Heat rejected to the cold HTF, charging processes as ψ = ψchar × ψdischar.

Qre = m˙ c c p,c (Tc,o − Tc,i) td, kJ (5)


6. Results and discussion
where ṁc – mass flow rate (kg/s) of cold HTF, cp,c – specific heat of cold
HTF, Tc,i and Tc,o – the inlet and the outlet temperature of the cold HTF, 6.1. Thermal properties of nano-enhanced PE
and td – time taken for the discharge.
The average discharging power (Pd) is calculated as the ratio of the The thermal conductivity of the PE with and without the addition of
heat rejected (Qre) to the cold HTF to the discharging time (td). Al2O3 nanoparticles calculated using the temperature data obtained
from the T-history plots. In T-history method, the thermal properties of
Qre
i.e., Pd = the PCM samples were estimated by comparing their temperature plot
td (6)
with that of the glycerin which was taken as the reference medium. The
The discharging efficiency determined as the ratio of the amount of obtained thermal conductivity value of pure PE was 0.106 W/m K. The
heat rejected to the cold HTF (Qre) to the heat stored (Qst) by the PCM values obtained for the PE samples added with Al2O3 nanoparticles
during the discharging cycle, i.e., showed very significant changes in the thermal conductivity values.
The T-history results showed that the thermal conductivity of PE en-
Qre
Discharging efficiency (%), ηd = × 100 hanced by 33.0% and 51.9% respectively corresponding to 0.5 and
Qst (7)
1.0 wt.% of Al2O3 nanoparticles. The improved thermal conductivities
The overall energy efficiency of the thermal energy storage (TES) of PE + 0.5 wt.% Al2O3 and PE + 1.0 wt.% Al2O3 obtained as 0.141 W/
system used in this experimental study calculated by combining the m K and 0.161 W/mK respectively.
efficiencies computed separately for the charging and discharging The variation of thermal conductivity with temperature is also in-
processes (Jagadheeswaran, Pohekar, & Kousksou, 2010; Rezaei et al., vestigated using the T-history data recorded in the application range
2013). 40–205 °C of pentaerythritol. The whole range is divided into three
The overall energy efficiency of the TES system, η = ηc × ηd smaller temperature ranges: (i) from 40 to 95 °C with 77.5 °C as the
(8)
mean temperature, (ii) from 95 to 150 °C with 122.5 °C as the mean
temperature, and (iii) from 150 to 205 °C with 177.5 °C as the mean
5.3.2. Exergy analysis temperature, the range in which the solid–solid transition occurs in PE
The energy analysis explained in the previous section is based on the (Fig. 10). Table 2 shows the thermal conductivity measurements ob-
first law of thermodynamics. Even though it provides an estimation of tained for pure PE and PE added with 0.5 and 1.0 wt.% of Al2O3 na-
the performance of the thermal energy storage system, it does not re- noparticles in the above temperature ranges.
flect the quality of energy stored/recovered. This shortfall can be It can be understood from the tabulated results that the thermal
overcome by performing the exergy analysis based on the second law of conductivity of all the tested samples decreases with temperature.
thermodynamics. Table 3 displays the thermal conductivity values of the PCM samples
Exergy is the quality or usefulness of energy. It is the maximum before and after thermal cycling. It can be seen from the table that the
amount of work that can be generated by a system as it comes to thermal conductivity of the PE added with 0.5 and 1.0 wt.% of Al2O3
equilibrium with surrounding. nanoparticles were significantly greater than that of pure PE even after
During charging mode, the HTF transfers the exergy to PCM and 1000 thermal cycles. This meant that Al2O3 nanoparticles did not

9
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

6.3. Charging performance analysis

6.3.1. Charging time


The charging performance experiments were conducted to study the
effect of heat transfer enhancement additive (Al2O3 nanoparticles) on
the thermal energy storage performance of PE. For this, experiments
were conducted using pure PE and PE added with 0.5 and 1.0 wt.% of
Al2O3 nanoparticles. During the charging process, the HTF (therminol
oil) heated to about 225 °C and circulated through the heat exchanger.
The PCM temperature variation from the inlet to the exit of the heat
exchanger in the axial direction is recorded for evaluating the charging
performance of the PCM. Fig. 12 shows the temperature variation in the
case of pure PE for HTF flow rates of 2, 4, and 6 LPM. The variation of
the PCM temperature at six locations in the heat exchanger is plotted in
this figure. During the charging process, the PCM in the shell side of the
heat exchanger absorbs the heat of the hot HTF. The charging process
continued until all the temperatures recorded by the thermocouples
exceeded the solid–solid transition temperature of pure PE. The char-
Fig. 10. Variation thermal conductivity of PE + Al2O3 with temperature. ging process ended in 152 min when the HTF flow rate of 2 LPM
maintained through the heat exchanger. The charging period found
decreased to 103 min and 68 min when the therminol flow rate changed
Table 2
Thermal conductivity variation with temperature. to 4 LPM and 6 LPM respectively. The increase in charging time ob-
served was because of the increased energy supplied at a higher flow
Temperature range Average Thermal conductivity (W/m K)
rate.
(°C) Temp (°C)
PE PE + 0.5% PE + 1.0% The charging experiments repeated for PE added with 0.5 and
Al2O3 Al2O3 1.0 wt.% of Al2O3 nanoparticles. Fig. 13 shows the temperature dis-
tribution in the PCM during the charging process of PE added with
40–95 67.5 0.112 0.148 0.165 0.5 wt.% of Al2O3 nanoparticles. The charging of PE + 0.5 wt.% Al2O3
95–150 122.5 0.108 0.139 0.157
150–205 177.5 0.104 0.132 0.149
ended in 127 min when the HTF flow rate of 2 LPM maintained through
the heat exchanger. The charging period decreased to 89 min and
56 min when the therminol flow rates changed to 4 LPM and 6 LPM
Table 3 respectively.
The thermal conductivity of PE with Al2O3 nanoparticles before and after Fig. 14 shows the charging process in the case of PE + 1.0 wt.%
thermal cycling. Al2O3 at different flow rates of the HTF. The charging of PE + 1.0 wt.%
PCM samples Thermal conductivity (k) (W/m K)
Al2O3 ended in 112 min when the HTF flow rate of 2 LPM maintained
through the heat exchanger. The charging period decreased to 78 min
Before thermal cycling After 1000 thermal cycles and 50 min when the flow rates changed to 4 LPM and 6 LPM respec-
tively.
PE 0.106 0.096
(−9.4%)
Fig. 15 shows the % reduction in the charging time of PE due to the
PE + 0.5 wt.% Al2O3 0.141 0.128 addition of Al2O3 nanoparticles at different flow rates of the HTF. It can
(+33.0%) (+20.8%) be understood from the graph that the charging time of PE decreases
PE + 1.0 wt.% Al2O3 0.161 0.140 due to the addition of 0.5 and 1.0 wt.% of Al2O3 nanoparticles at all
(+51.9%) (+32.1%)
volume flow rates of the heat transfer fluid.

undergo any deterioration in thermal property and caused very sig-


nificant enhancement thermal conductivity of PE even after several
thermal cycles.

6.2. Enthalpy of transition

The enthalpy of solid–solid transition is an important thermal


characteristic parameter of pentaerythritol. Differential Scanning
Calorimetry (DSC) analysis of the PCM samples was performed using
Mettler Toledo DSC 822e, Hong Kong. The heat capacity and enthalpy
of the solid–solid transition of the PCM samples obtained using the DSC
data. Fig. 11 illustrates the DSC plot obtained for pure PE sample.
The two peaks seen in the DSC curve represent the phase change
process in PE. The first peak corresponds to the solid to solid phase
change and the second peak corresponds to the solid–liquid phase
change. Table 4 summarizes the temperature and enthalpy of the so-
lid–solid transition of pure PE and PE added with 0.5 and 1.0 wt.% of
Al2O3 nanoparticles during the charging and discharging processes.

Fig. 11. DSC curve of pure PE.

10
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Table 4
Enthalpy of the transition of the PCM samples.
Samples Heating Cooling
Solid–solid transition Solid–solid transition

Onset Peak End set Enthalpy change Onset Peak End set Enthalpy change
(°C) (°C) (°C) (kJ/kg) (°C) (°C) (°C) (kJ/kg)

PE 181.3 187.8 193.6 263.9 165.6 163.5 163.0 238.9


PE + 0.5% Al2O3 180.8 188.4 190.0 255.0 169.5 168.9 162.9 229.5
PE + 1.0% Al2O3 180.9 188.4 200.6 251.7 169.8 169.6 162.6 226.5

Fig. 12. Charging process of pure PE at different flow rates of HTF.

6.3.2. Charging power and charging efficiency This accounts for a charging efficiency of 68.9%. When the flow rate
The average energy stored in unit time is estimated as the charging further increased to 6 LPM, the heat supplied and heat stored estimated
power. Fig. 16 shows the average charging power estimated in the case as 907.8 kJ and 1256.8 kJ showing a charging efficiency of 72.2%. The
of pure PE and PE added with 0.5 wt.% and 1.0 wt.% of Al2O3 nano- charging efficiency estimated for PE + 0.5 wt.% Al2O3 and PE + 1.0 wt.
particles. % Al2O3 is also shown in Fig. 17. The results of PE + 0.5 wt.% Al2O3
It can be understood that the charging power of PE increases with indicated charging efficiencies of 72.4%, 76.7% and 80.7% at HTF flow
the addition of Al2O3 nanoparticles for all the heating rates considered rates of 2, 4 and 6 LPM respectively. The heat supplied and heat stored
in this study. The average energy storage power of pure PE corre- at these flow rates obtained as 1235.2 kJ, 1154.2 kJ, 1089.3 kJ, and
sponding to 2 LPM, 4 LPM, and 6 LPM flow rates estimated as 100.5 W, 894.4 kJ, 884.7 kJ, 878.7 kJ respectively. The results of PE + 1.0 wt.%
149.0 W, and 222.5 W respectively. The average energy storage power Al2O3 indicated charging efficiencies of 80.3%, 83.0% and 86.8% at
corresponding to 2 LPM, 4 LPM, and 6 LPM flow rates increased to HTF flow rates of 2, 4 and 6 LPM respectively. The heat supplied and
117.45 W, 165.7 W, and 261.5 W respectively when 0.5 wt.% of Al2O3 heat stored at these flow rates obtained as 1093.4 kJ, 1050.2 kJ,
nanoparticles were added to PE. When the weight fraction of Al2O3 1089.3 kJ, and 877.5 kJ, 872.1 kJ, 868.0 kJ respectively.
nanoparticles increased to 1.0%, the average energy storage power
corresponding to 2 LPM, 4 LPM, and 6 LPM flow rates estimated as 6.4. Discharging performance analysis
130.5 W, 186.3 W and 289.3 W respectively.
Fig. 17 shows the charging efficiency obtained in the case of pure PE 6.4.1. Discharging time
and PE added with 0.5 and 1.0 wt.% of Al2O3 nanoparticles at different To study the discharge performance of the PCM, the PCM is first
HTF flow rates. In the case of pure PE, an amount of 916.4 kJ energy charged to a temperature above the solid–solid transition point and
stored out of the 1418.2 kJ heat supplied at an HTF flow rate of 2 LPM, then allowed to cool back to the ambient conditions. During the dis-
indicating a charging efficiency of 64.6%. When the flow rate changed charging process, the HTF at room temperature circulated through the
to 4 LPM, 911.6 kJ of heat was stored out of 1322.8 kJ of heat supplied. heat exchanger. The PCM rejects the heat stored to the circulating HTF.

11
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 13. Charging process of PE + 0.5 wt.% Al2O3 at different flow rates of HTF.

Fig. 14. Charging process of PE + 1.0 wt.% Al2O3 at different flow rates of HTF.

12
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 18 shows the temperature variation in the case of pure PE for


HTF flow rates of 2, 4, and 6 LPM. The variation of the PCM tempera-
ture at six locations in the heat exchanger is plotted in this figure.
During the discharging, the PCM in the shell side of the heat exchanger
rejects the heat to the HTF. The discharging process continued until all
the temperatures recorded by the thermocouples recorded temperature
equal to the ambient condition. The charging process ended in 106 min
when a flow rate of 2 LPM maintained through the heat exchanger. The
charging period found decreased to 86 min and 65 min when the HTF
flow rate changed to 4 LPM and 6 LPM respectively. The decrease in
charging time observed was because of the increased heat absorption at
the higher flow rates.
The discharging performances of PE added with 0.5 and 1.0 wt.% of
heat transfer enhancement additives Al2O3 nanoparticles, LMA and
LMM are also analyzed in this section. Fig. 19 shows the temperature
distribution in the PCM during the discharging process of PE added
with 0.5 wt.% of Al2O3 nanoparticles. The results showed that the dis-
charging time corresponding to the HTF flow rate of 2 LPM decreased
from 106 min to 90 min. This indicated that the discharging time re-
duced by 15% due to the increased heat transfer rated resulted from the
Fig. 15. Reduction in charging time of PE due to Al2O3. addition of 0.5 wt.% of Al2O3 nanoparticles. The discharging time
found decreased from 86 min to 72 min when the flow rate changed to
4 LPM indicating a 16.3% decrease. The discharge time further de-
creased to 54 min indicating a 16.9% decrease when the flow rate ad-
justed to 6 LPM.
Fig. 20 shows the temperature distribution in the PCM during the
discharging process of PE added with 1.0 wt.% of Al2O3 nanoparticles.
The results showed that the discharging time corresponding to the HTF
rate of 2 LPM decreased from 106 min to 78 min. This indicated that the
discharging time reduced by 26.4% due to the increased heat transfer
rated resulted from the addition of 1.0 wt.% of Al2O3 nanoparticles. The
discharging time found decreased from 86 min to 65 min when the flow
rate changed to 4 LPM indicating a 24.4% decrease. The discharge time
further decreased to 57 min indicating a 26.2% decrease when the flow
rate adjusted to 6 LPM. Fig. 21 summarizes the % reduction in the
discharging time of PE due to the addition of 0.5 and 1.0 wt.% of Al2O3
nanoparticles.

6.4.2. Discharging power and efficiency


The average energy released per unit time to the circulating HTF is
Fig. 16. Average charging power at different HTF flow rates.
estimated as the discharging power. Fig. 22 shows the average dis-
charging power estimated in the case of pure PE and PE added with
0.5 wt.% and 1.0 wt.% of Al2O3 nanoparticles at different flow rates of
the HTF.
The experimental results showed that the discharging power of PE
increases with the addition of additives for all the heating rates con-
sidered in this study. The average energy discharging power of pure PE
corresponding to 2 LPM, 4 LPM, and 6 LPM flow rates estimated as
75.8 W, 98.8 W, and 136.5 W respectively.
The average discharging power corresponding to 2 LPM, 4 LPM, and
6 LPM flow rates increased to 101.1 W, 134.8 W, and 187.0 W respec-
tively when 0.5 wt.% of Al2O3 nanoparticles were added to PE. When
the weight fraction of Al2O3 nanoparticles increased to 1.0%, the
average energy discharging power corresponding to 2 LPM, 4 LPM, and
6 LPM flow rates estimated as 124.1 W, 153.9 W, and 230.8 W respec-
tively.
Fig. 23 shows the discharging efficiency obtained in the case of pure
PE and PE added with 0.5 and 1.0 wt.% of Al2O3 nanoparticles at dif-
ferent HTF flow rates. In the case of pure PE, an amount of 482.2 kJ of
energy discharged, out of the 916.4 kJ heat stored, at an HTF flow rate
Fig. 17. Charging efficiency at different flow rates. of 2 LPM. This indicated a discharging efficiency of 52.6%. When the
flow rate changed to 4 LPM, 510.0 kJ of heat was released to the HTF,
The PCM temperature variation from the inlet to the exit of the heat out of 911.6 kJ of heat available. This accounts for a discharging effi-
exchanger in the axial direction is recorded for evaluating the dis- ciency of 55.9%. When the flow rate further increased to 6 LPM, the
charging performance of the PCM. heat rejected to the HTF estimated as 532.2 kJ, out of 907.8 kJ heat
stored by the PCM during the charging cycle. This gave a discharging

13
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 18. Discharging process of PE at different flow rates of HTF.

Fig. 19. Discharging process of PE + 0.5 wt.% Al2O3 at different flow rates of HTF.

14
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

Fig. 20. Discharging process of PE + 1.0 wt.% Al2O3 at different flow rates of HTF.

Fig. 22. Average discharging power at different HTF flow rates.


Fig. 21. % reduction in discharging time of PE due to Al2O3 nanoparticles at
different HTF flow rates.
6 LPM calculated as 66.2%, 68.8%, and 75.0% respectively. The heat
rejected to the HTF at these flow rates obtained as 580.9 kJ, 600.1 kJ,
efficiency of 58.6%. and 650.9 kJ respectively. The heat stored in the previous heating cycle
The discharging efficiencies obtained for PE + 0.5 wt.% Al2O3 and was 877.5 kJ, 872.1 kJ, and 868.0 kJ corresponding to the flow rates 2,
PE + 1.0 wt.% of Al2O3 are also displayed in Fig. 23. The results of 4, and 6 LPM respectively.
PE + 0.5 wt.% Al2O3 indicated discharging efficiencies of 61.0%,
65.8% and 70.2% at HTF flow rates of 2, 4 and 6 LPM respectively.
During the discharging cycle, 545.9 kJ of heat rejected, out of 894.4 kJ 6.5. Overall energy efficiency
of heat available, to the HTF circulated at 2 LPM. The heat released at
the flow rate of 4 LPM estimated to be 582.3 kJ, out of 884.7 kJ heat The overall energy efficiency of the thermal energy storage (TES)
stored by the PCM during the charging period. The heat discharged at system used in this experimental study is calculated by combining the
6 LPM flow rate estimated to be 617 kJ, out of 878.7 kJ heat that was energy efficiencies computed separately for the charging and dischar-
stored during the heating cycle. The discharging efficiency of ging processes. Fig. 24 displays the overall efficiencies obtained using
PE + 1.0 wt.% Al2O3 corresponding to the HTF flow rates of 2, 4 and the different PCM samples for energy storage and release. The experi-
mental results showed that the TES system using pure PE gave overall

15
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

to the improved thermal conductivity of PE due to the presence of


conductive Al2O3 nanoparticles as reported in the section on thermal
property measurement. The maximum value of the overall efficiency
obtained corresponding to 1.0 wt.% of Al2O3 nanoparticles at HTF flow
rate of 6 LPM. This indicated that the charging and discharging occurs
more efficiently at higher weight % of the heat transfer enhancement
additives and at higher flow rates of the HTF. Higher the weight % of
Al2O3 nanoparticles, greater is the enhancement in the thermal con-
ductivity which resulted in the heat transfer at an enhanced rate.
During the charging process, the hot HTF transfers heat to the PCM
filled in the shell side of the heat exchanger. Initially, heat is transferred
from the circulating HTF to the inner surface of the tubes through
convection. The heat is then conducted through the tube material by
conduction. The PCM layer in contact with the outer surfaces of the
tube receives this heat by conduction and then flows outwards into the
bulk of the PCM. The continuous heating of the PCM due to the transfer
of the heat from the HTF causes the temperature to increase linearly.
When the temperature of PE reaches its solid–solid transition tem-
Fig. 23. Discharging efficiency of PE with additives at different flow rates. perature, the added heat causes the breaking of metastable hydrogen
bonds in the PE molecules. The breaking of hydrogen bonds is an en-
dothermic process involving a large enthalpy change. Thus, a con-
siderable amount of heat is absorbed by the PCM during the breaking of
bonds. At ordinary temperature, P PE has the body-centered tetragonal
(BCT) crystal structure. At the transition temperature, the crystal
structure changes to face-centered cubic (FCC). If the PCM is further
heated above this transition temperature, the PCM receives heat again
by sensible heating.
During the discharging process, the heat stored by the PCM is re-
jected to the circulating cold HTF. The heat from the outer layer of the
PCM in contact with the shell is conducted toward the center resulting
in the sensible cooling of the PCM. When the temperature of the PCM
reaches the solid–solid transition temperature, the hydrogen bonds
which were broken earlier get reforms and the PE molecule regains its
original BCT structure. This reformation process is accompanied by the
release of heat. When the temperature of the PCM reaches below the
transition temperature, further release of heat is because of the sensible
cooling of the PCM. The outside surface of the metal tube wall receives
heat by conduction from the PCM layer in contact with the tubes. Heat
is conducted through the material of the tubes and then transferred to
Fig. 24. Overall energy efficiency of the thermal energy storage system.
the cold HTF by convection from the inner surface of the tubes.
The addition of Al2O3 nanoparticles has resulted in increasing the
energy efficiencies of 34%, 38.5% and 42.3% corresponding to HTF thermal conductivity of PE. This caused an enhanced rate of heat
flow rates of 2 LPM, 4 LPM, and 6 LPM respectively. The average effi- transfer by conduction through the PCM. Subcooling in solid–solid PCM
ciency of the system in the flow rate range 2–6 LPM was estimated to be refers to the cooling of material below its transition temperature but
38.3%. The addition of Al2O3 nanoparticles to PE resulted in an in- without the solid to solid transition actually taking place. The large
crease in the overall efficiency of the system. The overall efficiency of
the TES system employing PE + 0.5 wt.% Al2O3 showed overall energy
efficiencies of 44.2%, 50.5%, and 56.7% respectively corresponding to
HTF flow rates of 2 LPM, 4 LPM, and 6 LPM. The average efficiency of
the system in the flow rate range 2–6 LPM found to be 50.5%. When the
Al2O3 wt.% increased to 1.0, the TES system exhibited overall effi-
ciencies of 53.2%, 59.1% and 65.1% when the HTF flow rates were
2 LPM, 4 LPM, and 6 LPM respectively. The average efficiency of
PE + 1.0 wt.% Al2O3 in the flow rate range 2–6 LPM found to be 58.5%.
Fig. 25 shows the % enhancement in the overall energy efficiency of
the TES system at different flow rates. The addition of 0.5 wt.% of Al2O3
nanoparticles to PE at HTF flow rates of 2, 4 and 6 LPM gave en-
hancement of 10.2%, 12%, and 14.4% respectively compared to pure
PE. When the weight fraction of Al2O3 nanoparticles was increased to
1.0%, the overall energy efficiency showed enhancement of 19.2, 20.6
and 22.8 at HTF flow rates of 2, 4 and 6 LPM respectively compared to
PE.
The results of the experimental investigation discussed in the pre-
ceding sections revealed that the incorporation of Al2O3 nanoparticles
enhanced the thermal energy storage performance of pentaerythritol.
The enhanced energy storage and release performance can be attributed Fig. 25. % enhancement in overall efficiency.

16
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

degree of sub-cooling may affect the performance of phase change


materials and limit their application. During subcooling, the storage
material behaves similar to a sensible storage material, and the storage
capacity is reduced. This undesired subcooling in PE was currently re-
duced by using Al2O3 nanoparticles as nucleating agents, also called
nucleators. These additives remain solid at all times and act as centers
of crystal growth for the material that undergoes the phase change. The
reduction in subcooling was attributed to being heterogeneous nu-
cleation which occurred along with the Al2O3 particles in the crystal-
lization process.

6.6. Exergy analysis

The data obtained from the charging and discharging experiments


are used to carry out the exergy analysis of the TES system. Exergy-
based performance evaluation of TES system gives more perspective
measure than energy based one as it reveals the true potential of the
system and the economic assessment of the storage/discharge process.
The exergy input, exergy stored and the exergy output estimated in the
case of PE, PE + 0.5 wt.% Al2O3 PE + 1.0 wt.% Al2O3 for flow rates of
2, 4 and 6 LPM of the HTF are summarized in Table 5.
The energy efficiency of a TES under ideal condition will be 100% as
there is no energy loss in the system. But in the actual case, there are Fig. 26. Lost and useful exergy.
always irreversibilities in the system which results in energy loss.
Therefore, energy efficiencies calculated do not reflect the useful en- results showed 12.6%, 15.1% and 18% increase in the overall exergy
ergy since the lost energy is not. In the case of exergy evaluation, both efficiency corresponding to the flow rates 2, 4 and 6 LPM respectively.
the useful and lost energies are taken into consideration. Due to this Amount of exergy lost at each test condition is estimated from the
reason, the exergy efficiency is much less than the energy efficiency for input and output exergy values and are presented in Fig. 23. The
all the test conditions mentioned above. The results show that addition quantification of lost exergy is very important in the life cycle cost
of the alumina nanoparticles has resulted in enhancing the exergy ef- evaluation of PCM.
ficiency of the TES system. For each PCM tested, the exergy efficiency is It can be understood from Fig. 26 that the fraction of useful exergy
found increasing with the increase in the flow rate of the HTF. The increases due to the incorporation of Al2O3 nanoparticles for all the
reason for the enhanced heat transfer occurred at a higher flow rate was flow rates of HTF. In other words, the fraction of lost exergy decreases
because of the increased turbulence in the flow. The increased turbu- due to the addition of Al2O3 nanoparticles.
lence in the region near to the tube wall caused very efficient fluid
mixing and efficient redevelopment of the thermal/hydrodynamic
boundary layer leading to the improvement in the convective heat 6.7. Experimental uncertainties
transfer (Chandrasekar, Suresh, & Chandra Bose, 2010; Suresh,
Venkitaraj, Selvakumar, & Chandrasekar, 2012). The experimental re- The main parameter monitored during T-history, and thermal sto-
sults summarized in Table 3 indicate that the addition of Al2O3 nano- rage/release performance tests were the PCM and HTF temperatures. In
particles to PE has resulted in an increase in the charging and dis- this experimental study, K type thermocouples connected to a multi-
charging exergy efficiency of the TES for all the flow rates of the HTF channel data acquisition system (KEYSIGHT 34972A LXI) were used to
tested. Table 5 also presents the overall exergy efficiency of the TES record the various temperatures. The uncertainties in the energy mea-
using pure PE and PE added with 0.5 and 1.0 wt.% of Al2O3 nano- surements were calculated using the temperature and flow rate un-
particles. In the case of pure PE, overall exergy efficiencies of 21.5%, certainties. The uncertainties in the flow meter reading are specified by
25.2%, and 29.1% have obtained corresponding to the HTF flow rates the manufacturer as ± 1% of full flow. The calibrated temperature
of 2, 4 and 6 LPM respectively. The addition of 0.5 wt.% of Al2O3 to PE sensors showed an uncertainty of ± 0.1 °C. Now the uncertainty in the
resulted in an increase in the exergy efficiency by 5.6%, 5.8% and 7.5% heat supplied/recovered (δQ) was calculated using the know un-
corresponding to the HTF flow rates of 2, 4 and 6 LPM respectively. certainty values of temperature (δT) and mass flow rate of the HTF (δṁ ).
When the weight fraction of Al2O3 was increased to 1.0%, the test Therefore,

Table 5
Input, stored and output exergy.
PCM Flow rate ˙ input
Ex ˙ stored
Ex ˙ output
Ex Charging exergy efficiency Discharging exergy efficiency Overall exergy efficiency
(LPM) (kJ/day) (kJ/day) (kJ/day) (%) (%) (%)
ψchar ψdischar ψ

PE 2 6655 3182 1440 47.8 45.2 21.6


4 9277 4693 2337 50.6 49.8 25.2
6 12,759 6979 3719 54.7 53.3 29.1

PE + 0.5 wt.% Al2O3 2 7086 3740 1930 52.8 51.6 27.2


4 9312 5223 2889 56.1 55.3 31
6 13,339 8191 4879 61.4 59.6 36.6

PE + 1.0 wt.% Al2O3 2 6897 4132 2356 59.9 57 34.2


4 9201 5903 3710 64.2 62.9 40.3
6 13,065 9059 6153 69.3 67.9 47.1

17
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

˙ ˙ T )2
δQ = tcp,htf (ΔT δm˙ )2 + 2(mδ References

where ṁ – the mass flow rate of the HTF, cp,htf – specific heat of the Adorno, A., & Silva, R. (2006). Effect of Ag additions on the reverse martensitic trans-
HTF, ΔT – the difference between the inlet and exit fluid temperatures, t formation in the Cu–10 mass% Al alloy. Journal of Thermal Analysis and Calorimetry,
83, 241–246.
– time of charging/discharging. Agarwal, A., & Sarviya, R. M. (2016). An experimental investigation of shell and tube
The uncertainty in the heat stored was estimated as latent heat storage for solar dryer using paraffin wax as heat storage material.
δQstored = Mcp,pcm (ΔT × δT) + M(δh), where δh is the uncertainty in Engineering Science and Technology, an International Journal, 19, 619–631.
Agyenim, F., Hewitt, N., Eames, P., & Smyth, M. (2010). A review of materials, heat
the enthalpy change given by the DSC equipment and M is the mass of transfer and phase change problem formulation for latent heat thermal energy sto-
the PCM and cp,pcm is the specific heat of the PCM. rage systems (LHTESS). Renewable and Sustainable Energy Reviews, 14, 615–628.
Using the above relations, the average uncertainties in heat sup- Anish, R., Mariappan, V., & Suresh, S. (2019, January). Experimental investigation on
melting and solidification behavior of erythritol in a vertical double spiral coil
plied, the heat stored and heat recovered in the HTF flow rate range thermal energy storage system. Sustainable Cities and Society, 44, 253–264.
considered in this work were estimated as 8.1%, 7.1%, and 8.5% re- Arcuri, B., Spataru, C., & Barrett, M. (2017, February). Evaluation of ice thermal energy
spectively. storage (ITES) for commercial buildings in cities in Brazil. Sustainable Cities and
Society, 29, 178–192.
Benson, D. K., Burrows, R. W., & Webb, J. D. (1986). Solid state phase transitions in
pentaerythritol and related polyhydric alcohols. Solar Energy Materials, 13, 133–152.
7. Summary and conclusion
Busico, V., Carfagna, C., Salerno, V., Vacatello, M., & Fittipaldi, F. (1980). The layer
perovskites as thermal energy storage systems. Solar Energy, 24, 575–579.
The effect of adding Al2O3 nanoparticles on the thermal energy Cabeza, L. F., Mehling, H., Hiebler, S., & Ziegler, F. (2002). Heat transfer enhancement in
storage performance of PE was studied by conducting the experiments water when used as PCM in thermal energy storage. Applied Thermal Engineering, 22,
1141–1151.
using pure PE and PE added with 0.5 and 1.0 wt.% of Al2O3 nano- Cao, F., Ye, J., & Yang, B. (2013). Synthesis and characterization of solid-state phase
particles. During the charging process, the hot HTF (therminol oil) at change material microcapsules for thermal management applications. Journal of
about 225 °C, and circulated through the heat exchanger at different Nanotechnology in Engineering and Medicine, 4, 040901.
Chandrasekar, M., Suresh, S., & Chandra Bose, A. (2010). Experimental studies on heat
flow rates of 2 LPM, 4 LPM, and 6 LPM. During the discharging process, transfer and friction factor characteristics of Al2O3/water nanofluid in a circular pipe
cold HTF (therminol oil) at about 30 °C, and circulated through the heat under laminar flow with wire coil inserts. Experimental Thermal and Fluid Science, 34,
exchanger at different flow rates of 2 LPM, 4 LPM, and 6 LPM. The main 122–130.
Darzi, A. A. R., Jourabian, M., & Farhadi, M. (2016). Melting and solidification of PCM
findings of the experimental study are given below enhanced by radial conductive fins and nanoparticles in cylindrical annulus. Energy
Conversion and Management, 118, 253–263.
• The results of the charging and discharging experiments indicated a Elgafy, A., & Lafdi, K. (2005). Effect of carbon nanofibres additives on thermal behavior of
phase change materials. Carbon, 43, 3067–3074.
significant reduction in the charging and discharging time of PE due
Ezan, M. A., Ozdogan, M., & Erek, A. (2011). Experimental study on charging and dis-
to the addition of 0.5 and 1.0 wt.% of the additives Al2O3 nano- charging periods of water in a latent heat storage unit. International Journal of
particles. Thermal Sciences, 50, 2205–2219.

• The charging and discharging power and efficiency of PE with Al2O3 Ge, Z., Ye, F., Cao, H., Leng, G., Qin, Y., & Ding, Y. (2014). Carbonate-salt-based com-
posite materials for medium- and high-temperature thermal energy storage.
nanoparticles found increased at all flow rates of the HTF. The Particuology, 15, 77–81.
maximum enhancement in the efficiency of charging and dischar- Giro-Paloma, J., Martinez, M., Cabeza, L. F., & Fernandez, A. I. (2016). Types, methods,
ging observed at the highest flow rate of 6 LPM. techniques, and applications for microencapsulated phase change materials (MPCM):


A review. Renewable and Sustainable Energy Reviews, 53, 1059–1075.
The charging efficiency found increased to 72.4%, 76.7% and 80.7% Gu, X., Xi, P., Cheng, B., & Niu, S. (2010). Synthesis and characterization of a novel
at HTF flow rates of 2, 4 and 6 LPM respectively when 0.5 wt.% of solid–solid phase change luminescence material. Polymer International, 59, 772–777.
Al2O3 nanoparticles was added to PE. The results of PE + 1.0 wt.% Hu, P., Zhao, P. P., Jin, Y., & Chen, Z. S. (2014). Experimental study on solid–solid phase
change properties of pentaerythritol (PE)/nano-AlN composite for thermal storage.
Al2O3 indicated a further increase in the charging efficiencies to Solar Energy, 102, 91–97.
80.3%, 83.0% and 86.8% corresponding to the HTF flow rates of 2, Jagadheeswaran, S., Pohekar, S. D., & Kousksou, T. (2010). Exergy based performance
4 and 6 LPM respectively. evaluation of latent heat thermal storage system: A review. Renewable and Sustainable

• The discharging efficiency found increased to 61.0%, 65.8% and


Energy Reviews, 14, 2580–2595.
Jiang, Y., Ding, E., & Li, G. (2002). Study on the transition characteristics of PEG/CDA
70.2% at HTF flow rates of 2, 4 and 6 LPM respectively when 0.5 wt. solid–solid phase change materials. Polymer Journal, 43, 117–122.
% of Al2O3 nanoparticles was added to PE. The results of Khan, Z., Khan, Z., & Ghafoor, A. (2016). A review of performance enhancement of PCM
based latent heat storage system within the context of materials, thermal stability and
PE + 1.0 wt.% Al2O3 indicated a further increase in the discharging
compatibility. Energy Conversion and Management, 115, 132–158.
efficiency to 66.2%, 68.8% and 75.0% corresponding to the HTF Khan, Z., & Khan, Z. A. (2017). An experimental investigation of discharge/solidification
flow rates of 2, 4 and 6 LPM respectively. cycle of paraffin in novel shell and tube with longitudinal fins based latent heat

• The mean value of the overall energy efficiency of the thermal en- storage system. Energy Conversion and Management, 154, 157–167.
Khodadadi, J. M., & Zhang, Y. (2001). Effects of buoyancy-driven convection on melting
ergy storage system using pure PE estimated to be 38.3%. The ad- within spherical containers. International Journal of Heat and Mass Transfer, 44,
dition of 0.5 and 1.0 wt.% of Al2O3 nanoparticles caused the mean 1605–1618.
value of overall efficiency to increase to 50.5% and 58.5% respec- Kim, S., & Drzal, L. T. (2009). High latent heat storage and high thermal conductive phase
change materials using exfoliated graphite nanoplatelets. Solar Energy Materials and
tively. Solar Cells, 93, 136–142.
• The overall efficiency of a thermal energy storage system employing Landi, E., & Vacatello, M. (1975). Metal-dependent thermal behavior Ln (n-
CnH2n+1NH3)2MCl4. Thermochimica Acta, 13, 441–447.
0.5 and 1.0 wt.% of Al2O3 nanoparticles showed very significant
Li, W., Zhang, D., Zhang, T., Wang, T., Ruan, D., Xing, D., et al. (1999). Study of so-
improvement at all flow rates of HTF considered in this study. lid–solid phase change of (n-CnH2n+1NH3)2MCl4 for thermal energy storage.
Therefore, it can be concluded that the use of 0.5 and 1.0 wt.% Thermochimica Acta, 326, 183–186.
Al2O3 nanoparticles for heat transfer enhancement results in very Li, W. D., & Ding, E. Y. (2007). Preparation and characterization of cross linking PEG/
MDI/PE copolymer as solid–solid phase change heat storage material. Solar Energy
significant improvement in the thermal performance of PE for long Materials and Solar Cells, 91, 764–768.
term energy storage applications. Li, X., Zhoua, Y., Nian, H., Zhang, X., Dong, O., Ren, X., et al. (2017). Advanced nano-
composite phase change material based on calcium chloride hexahydrate with alu-
minum oxide nanoparticles for thermal energy storage. Energy & Fuels, 3, 6560–6567.
Acknowledgments Liu, C., Rao, Z., Zhao, J., Huo, Y., & Li, Y. (2015). Review on nanoencapsulated phase
change materials: Preparation, characterization and heat transfer enhancement. Nano
Energy, 13, 814–826.
Authors express their sincere gratitude to the Centre for Engineering Liu, L., Su, D., Tang, Y., & Fang, G. (2016). Thermal conductivity enhancement of phase
Research and Development (KTU/RESEARCH 3/1199/2017 dated change materials for thermal energy storage: A review. Renewable and Sustainable
19.04.2017) for the financial assistance given for the execution of the Energy Reviews, 62, 305–317.
Liu, Z., Sun, X., & Ma, C. (2005). Experimental study of the characteristics of solidification
work reported in this paper.

18
K.P. Venkitaraj, et al. Sustainable Cities and Society 51 (2019) 101767

of stearic acid in an annulus and its thermal conductivity enhancement. Energy Seddegh, S., Joybari, M. M., Wang, X., & Haghighat, F. (2017). Experimental and nu-
Conversion and Management, 46, 971–984. merical characterization of natural convection in a vertical shell-and-tube latent
Luo, K., Yao, F. J., Yi, H. L., & Tan, H. P. (2015). Lattice Boltzmann simulation of con- thermal energy storage system. Sustainable Cities and Society, 35, 13–24.
vection melting in complex heat storage systems filled with phase change materials. Seddegh, S., Wang, X., & Henderson, A. D. (2016). A comparative study of thermal be-
Applied Thermal Engineering, 86, 238–250. haviour of a horizontal and vertical shell-and-tube energy storage using phase change
Medrano, M., Yilmaz, M. O., Nogues, M., Martorell, I., Roca, J., & Cabeza, L. F. (2009). materials. Applied Thermal Engineering, 93, 348–358.
Experimental evaluation of commercial heat exchangers for use as PCM thermal Siegel, R. (1977). Solidification of low conductivity material containing dispersed high
storage systems. Applied Energy, 86, 2047–2055. conductivity particles. International Journal of Heat and Mass Transfer, 20, 1087–1089.
Meng, Z., & Zhang, P. (2017). Experimental and numerical investigation of a tube-in-tank Singh, D. K., Suresh, S., Singh, H., Rose, B. A. J., Tassou, S., & Anantharaman, N. (2017).
latent thermal energy storage unit using composite PCM. Applied Energy, 190, Myo-inositol based nano-PCM for solar thermal energy storage. Applied Thermal
524–539. Engineering, 110, 564–572.
Mourid, A., El Alami, M., & Kuznik, F. (2018, August). Experimental investigation on Singh, H., Talekar, A., Chiena, W.-M., Shi, R., Chandra, D., Mishra, A., et al. (2015).
thermal behavior and reduction of energy consumption in a real scale building by Continuous solid-state phase transitions in energy storage materials with orienta-
using phase change materials on its envelope. Sustainable Cities and Society, 41, tional disorder – Computational and experimental approach. Energy, 91, 334–349.
35–43. Suresh, S., Venkitaraj, K. P., Selvakumar, P., & Chandrasekar, M. (2012). Effect of
Murrill, E., & Breed, L. (1970). Solid–solid phase transitions determined by differential Al2O3–Cu/water hybrid nanofluid in heat transfer. Experimental Thermal and Fluid
scanning calorimetry: Part I. Tetrahedral substances. Thermochimica Acta, 1, Science, 38, 54–60.
239–246. Teng, T.-P., & Yu, C.-C. (2012). Characteristics of phase-change materials containing
NKhonjera, L., Kuboth, M., Haagen, A. K., John, G., Kingondu, C., Bruggemann, D., et al. oxide nano-additives for thermal storage. Nanoscale Research Letters, 7, 611.
(2016). Experimental investigation of a finned pentaerythritol-based heat storage Timmermans, J. (1961). Plastic crystals: A historical review. Journal of Physics and
unit for solar cooking at 150–200 °C. Energy Procedia, 93, 160–167. Chemistry of Solids, 18, 1–8.
Nkwetta, D. N., & Haghighat, F. (2014, February). Thermal energy storage with phase Venkitaraj, K. P., & Suresh, S. (2018). Experimental study on the thermal storage per-
change material—A state-of-the-art review. Sustainable Cities and Society, 10, 87–100. formance and non-isothermal crystallization kinetics of pentaerythritol blended with
Pandiyarajan, V., Pandian, M. C., Malan, E., Velraj, R., & Seenira, R. V. (2011). low melting metal. Thermochimica Acta, 10, 75–89.
Experimental investigation on heat recovery from diesel engine exhaust using finned Venkitaraj, K. P., & Suresh, S. (2019). Effects of Al2O3, CuO and TiO2 nanoparticles on
shell and tube heat exchanger and thermal storage system. Applied Energy, 88, 77–87. thermal, phase transition and crystallization properties of solid–solid phase change
Pielichowska, K., & Pielichowski, K. (2010). Biodegradable PEO/cellulose based so- material. Mechanics of Materials, 128, 64–88.
lid–solid phase change materials. Polymers for Advanced Technologies, 22, 1633–1641. Venkitaraj, K. P., Suresh, S., & Venugopal, A. (2018). Experimental study on the thermal
Py, X., Olives, R., & Mauran, S. (2001). Paraffin/porous-graphite-matrix composite as performance of nano enhanced pentaerythritol in IC engine exhaust heat recovery
high and constant power thermal storage material. International Journal of Heat and application. Applied Thermal Engineering, 137, 461–474.
Mass Transfer, 44, 2727–2737. Wang, W. W., Zhang, K., Wang, L. B., & He, Y. L. (2013). Numerical study of the heat
Qi, C., & Liu, P. S. (2006). Structure and mechanical properties of shape memory poly- charging and discharging characteristics of a shell-and-tube phase change heat sto-
urethane based on hyperbranched polyesters. Polymer Bulletin, 57, 889–899. rage unit. Applied Thermal Engineering, 58, 542–553.
Rezaei, M., Anisur, M. R., Mahfuz, M. H., Kibria, M. A., Saidur, R., & Metselaar, I. H. S. C. Xi, P., Gu, X., Cheng, B., & Wang, Y. (2009). Preparation and characterization of a novel
(2013). Performance and cost analysis of phase change materials with different polymeric based solid–solid phase change heat storage material. Energy Conversion
melting temperatures in heating systems. Energy, 53, 173–178. and Management, 50, 1522–1528.
Ruan, D., Li, W., & Hu, Q. (1995). Phase diagrams of binary systems of alkyl ammonium Yin, H., Gao, X., Ding, J., & Zhang, Z. (2008). Experimental research on heat transfer
tetrachloro metallates (II). Journal of Thermal Analysis and Calorimetry, 45, 235–242. mechanism of the heat sink with composite phase change materials. Energy
Ruiyun, X., Dejun, K., Xian, E. C., & Jing, Z. (1990). Studies of solid–solid phase transi- Conversion and Management, 49, 1740–1746.
tions for (n-C18H37NH3)2MCl4. Thermochimica Acta, 164, 307–314.

19

You might also like