You are on page 1of 13

Solar Energy 174 (2018) 489–501

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

The error of neglecting natural convection in high temperature vertical T


shell-and-tube latent heat thermal energy storage systems
S. Saeed Mostafavi Tehrania,*, Gonzalo Diarceb, Robert A. Taylora
a
School of Mechanical and Manufacturing Engineering, University of New South Wales, Sydney, NSW 2052, Australia
b
ENEDI Research Group, Dpto. de Máquinas y Motores Térmicos, Escuela de Ingeniería de Bilbao, University of the Basque Country UPV/EHU, Rafael Moreno Pitxitxi 2,
Bilbao 48013, Spain

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

Keywords: There is little understanding of the relative importance of natural convection when designing latent heat thermal
High temperature energy storage (LHTES) systems based on geometric parameters and/or phase change material (PCM) properties.
Latent heat For high temperature shell-and-tube LHTES systems, this study aims: (i) to determine the error of ignoring
Phase change material (PCM) natural convection, and (ii) to quantify this error for different geometric parameters and PCM properties. In
Natural convection
particular, the study defines the circumstances under which natural convection is important and the error of
Error correlation
Shell-and-tube
choosing a ‘conduction-only modelling approach’. To do so, the performance of LHTES systems with nine
geometric aspect ratios and three commercial PCMs (of different melting points) were analyzed by means of a
validated CFD model.
The results showed that the error is a function of the process under analysis (melting or solidification) and the
ratio of stored/delivered energy divided by the maximum capacity of PCM (i.e. its effectiveness). Geometry also
plays a critical role in the relative importance of natural convection. The study demonstrates that a specific
system geometry (i.e. a dimensionless number defined based on the inner and outer radius as well as the length
R2 − ro2
of shell-and-tube geometry: S = 2ro L
) can be used to determine the relevance of natural convection. It was
found that regardless of PCM type, the error is of neglecting natural congestion is small if S < 0.005. For
r
S > 0.005, the error depends on the following non-dimensional groups: Lo , Ra, Ste, and Bi . As might be ex-
pected, the Rayleigh number was found to be the most influential group. Notably, a critical Rayleigh number
value (8 × 105) was found, below which the error of neglecting natural convection is < 1%. Finally, two cor-
relations were developed in order to quantify the error achieved – one for melting and another for solidification.

1. Introduction et al., 2012; Zalba et al., 2003). Advanced high temperature systems are
currently under development to increase the efficiency of concentrated
Thermal energy storage (TES) is a key component in intermittent solar power-tower (CSP-tower) plants, where the heat transfer fluid is
energy conversion cycles like solar energy plants, where there is a heated up to approximately 565 °C in order to produce electricity
mismatch between supply and demand (Tehrani et al., 2013a, 2013b). (Mostafavi Tehrani et al., 2018; Gil et al., 2010; Jacob et al., 2016;
While sensible heat storage currently dominates the market for this type Kuravi et al., 2013; Liu et al., 2016; Tehrani and Taylor, 2016; Tehrani
of TES technology (Dincer and Rosen, 2002; Kuravi et al., 2013; et al., 2017).
Seddegh et al., 2015b; Tehrani et al., 2017), latent heat thermal energy Among the different configurations of LHTES systems, shell-and-
storage (LHTES) systems have gained prominence in recent years as tube heat exchangers represent a promising and straightforward high
they represent a promising alternative to traditional TES systems temperature PCM design (Seddegh et al., 2018). As a result, this con-
(Cárdenas and León, 2013; Dutil et al., 2011). These systems use phase figuration is gaining interest (Agyenim et al., 2010; Nithyanandam and
change materials (PCMs), in a single or cascaded configuration (Tehrani Pitchumani, 2011; Tehrani et al., 2017). One of the most important
et al., 2018b), which store the latent heat of melting and release it upon challenges facing this kind of system is the geometric design optimi-
solidification. Compared to sensible heat storage, PCMs enable more zation (Li et al., 2017; Mahdavi et al., 2016; Nithyanandam and
compact designs, which can result in lower storage media costs (Liu Pitchumani, 2011, 2014; Tiari and Qiu, 2015) – task that is usually

*
Corresponding author.
E-mail addresses: s.mostafavi.tehrani@gmail.com (S.S. Mostafavi Tehrani), gonzalo.diarce@ehu.es (G. Diarce), robert.taylor@unsw.edu.au (R.A. Taylor).

https://doi.org/10.1016/j.solener.2018.09.048
Received 4 July 2017; Received in revised form 20 August 2018; Accepted 16 September 2018
Available online 21 September 2018
0038-092X/ © 2018 Elsevier Ltd. All rights reserved.
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Nomenclatures Greek symbols

A area [m2] β volumetric expansion coefficient [1/K]


CP specific heat at constant pressure [J/kg K] ρ density [kg/m3]
d inner diameter of latent heat storage unit [m] μ dynamic viscosity [kg/m s]
f liquid fraction [–]
h enthalpy [J/kg] Subscripts
H , ΔH latent heat of fusion [J/kg]
k thermal conductivity [W/m K] l liquid
L length of tank [m] r radial direction
ṁ mass flow rate [kg/s] z axial direction
p pressure [Pa]
Q amount of stored/discharged energy [J] Abbreviations
Q̇ amount of stored/discharged energy [W]
R outer radius of a cylinder [m] ALF average liquid fraction
ro inner radius of pipe [m] HTF heat transfer fluid
t total time of simulation [s] LHTES latent heat thermal energy storage
T temperature [K] Nu Nusselt number
Tm1 lower melting point [K] PCM phase change material
Tm2 upper melting point [K] Ra Rayleigh number
→v velocity [m/s] Ste Stefan number
TES thermal energy storage

performed via modelling. However, the accurate simulation of two- recirculation zone inside the liquid region. This phenomenon increases
phase heat transfer problems is complex because of the moving the heat transfer within the liquid PCM and causes non-uniformities in
melting/solidification boundary (Fornarelli et al., 2016). the solid-liquid interface and temperature distribution during melting.
To date, several computational fluid dynamics (CFD) models have Solidification, on the other hand, is mainly dominated by conduction,
been reported in the literature as for the modelling of phase change although initially (at high liquid fractions) convection and conduction
processes (Riahi et al., 2017; Tehrani et al., 2016b). Conduction-only transfer a similar amount of heat. Consequently, various researchers
models are much faster than full convection models, so they are often have investigated the fundamentals of natural convection.
used for design optimization. In fact, several recent studies only con- Fornarelli et al. (2016) compared the results obtained from con-
sider the conduction mechanism (See Table 1). However, this approach vective and pure conductive models of a high temperature shell-and-
can lead to non-negligible errors depending on the conditions. tube LHTES system for a CSP application during the melting process.
As demonstrated in a recent review on the topic (Dhaidan and The study confirmed that the convective motion increases the heat flux
Khodadadi, 2015), upon melting, natural convection drives a to the PCM, effectively increasing the heat transfer rate (e.g. reducing

Table 1
Selected shell-and-tube LHTES system studies.
Ref. Tm (°C) Melting/Solidification Convection Considered? Method L (m) R/ ro L/d

(Wang et al., 2015) 26 Melt. Yes 2D Num. – 1.2–5 25–150


(Wang et al., 2013) 26 Melt./Sol. Yes 2D Num. 1 2 78
(Tao and He, 2011) 26 Melt. Yes 2D Num. 1 1.8 83
(Trp, 2005; Trp et al., 2006) 26 Sol. No Exp./2D Num. 1 3.7 30
(Lacroix, 1993) 26 Melt. Yes 2D Num. 1 1.6 78
(Adine and El Qarnia, 2009; El Qarnia, 2009) 26 Melt./Sol. Yes 2D Num. 1 1.6 78
(Longeon et al., 2013) 35 Melt./Sol. Yes Exp./CFD 0.4 2.9 26
(Kalhori and Ramadhyani, 1985; Kemink and Sparrow, 1981) 36 Melt. Yes 2D Num. 0.2 8.4 10
(Mendes and Brasil, 1987) 36 Melt. Yes Exp. 0.17 6 5
(Esen et al., 1998) 29–46 Melt. No 2D Num. 3.2 1.09–1.2 20–48
(Fang and Chen, 2007) 20–80 Melt. No 2D Num. 2 2 100
(Seddegh et al., 2016) 51 Melt./Sol. Yes CFD 1 3.9 12
(Seddegh et al., 2015a) 58 Melt./Sol. Yes CFD 1 4 30
(Xiao and Zhang, 2015a, 2015b) 60 Melt./Sol. Yes Exp./2D Num./CFD 0.75 – 10.7
(Akgün et al., 2007) 75 Melt./Sol. Yes Exp. 0.5 3.3 17
(Wang et al., 2016) 121 Melt./Sol. Yes Exp. 0.9 2.4 20
(Fan et al., 2014) 169 Melt./Sol. No 2D Num. 1.4 1.25 87
(Pointner et al., 2016) 215 Melt. Yes 2D Num. 1 – –
(Fornarelli et al., 2016) 230 Melt./Sol. Yes CFD 0.5 5 35
(Tehrani et al., 2016b) 300–500 Melt./Sol. No 2D Num. 1–5 1.3–3 10–100
Present study 300–500 Melt./Sol. Yes CFD 0.5–2 1.5–3 15–60
(Riahi et al., 2017) 306 Melt./Sol. Yes CFD 0.6 3.15 29
(Muhammad et al., 2015a) 306 Melt./Sol. Yes CFD 0.9 5.41 155
(Tao and Carey, 2016) 400–600 Melt. No 2D Num. 1 2 40
(Li et al., 2013) 400–700 Melt. No 2D Num. 1.2 2 120
(Pirasaci and Goswami, 2016) 550 Sol. No 2D Num. 10–150 1–3 10–150
(Bellecci and Conti, 1993a, b) 734 Melt. No 2D Num. – 2.6–5 12–150
(Tao et al., 2014; Tao et al., 2012) 766 Melt. No 2D Num. 1.5 2 60

490
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

these factors comprehensively. In addition, the actual error of ignoring


convection has not yet been quantified. The present study addresses the
aforementioned aspects by comparing conduction-only and convection
models in a high temperature shell-and-tube LHTES system during both
melting and solidification. A full range of geometric parameters
R L
(0.5 ≤ L ≤ 2, 1.5 ≤ r ≤ 3, 15 ≤ d ≤ 60 ) and three high temperature
o
commercial PCMs (http://www.pcmproducts.net/, 2018): H325, H425,
H525 were considered. This enables a geometric and property depen-
dent quantification of the error in the form of some simple correlation
equations. These equations will help researchers decide when to neglect
natural convection for any high temperature shell-and-tube LHTES
system and/or to estimate the error of making this assumption. Fol-
lowing this work, Tehrani et al. (2018a) developed a general natural
convection Nusselt correlation to be used via effective thermal con-
ductivity method in conduction numerical models for the rapid mod-
elling of shell-and-tube LHTES system, which can minimize the error of
neglecting natural convection.

2. The high temperature shell-and-tube LHTES system

Fig. 1. A (vertical) single unit cylindrical shell-and-tube LHTES system. 2.1. System description, operational parameters and materials

the charging time). Longeon et al. (2013) investigated the effect of top The generic system analyzed in this study is based on a LHTES
and bottom HTF injection with respect to natural convection in a shell- system previously proposed by the authors in (Tehrani et al., 2016b). It
and-tube LHTES system. For both injection options, the PCM always consists of concentric tube(s) whereby the HTF and PCM are segre-
melted first at the top of the exchanger throughout its whole thickness gated, as shown in Fig. 1.
and natural convection transported the heat along the inner tube of the The HTF flows through the inner tube and exchanges heat with the
exchanger. The natural convection effect, however, was less significant PCM in the surrounding region (i.e. between the inner and outer tube
during solidification as compared to melting. Wang et al. (2016) per- shown in Fig. 1). During the charging process (melting), the hot HTF
formed an experimental study followed by CFD simulations for a flows from top to bottom. During the discharging (solidification) pro-
medium temperature shell-and-tube LHTES system. They demonstrated cess, the flow direction is bottom up. The top of the tank is denoted as
that as the melting process progresses, the molten PCM first occupies Z/L = 1 and the bottom of the tank is referred to as Z/L = 0. The op-
the top region and then spreads from top to bottom. Moreover, natural erational conditions of the shell-and-tube LHTES system were adopted
convection was more pronounced as the volume and temperature of the from the design point operational conditions of CSP-tower plants as
molten PCM region increased. The authors also explained that during tabulated in Table 2 (Tehrani and Taylor, 2016; Tehrani et al., 2016a).
solidification, the PCM starts to solidify from the bottom region at the The flow rate was also adopted from a previous work of the authors,
beginning, and then the solid-liquid interface uniformly progresses and represents the typical discharging flow rate inside each tube of an
from the inner tube to the outer shell. These results confirm the initial optimized shell-and-tube LHTES system in a CSP-tower plant (Tehrani
dominance of natural convection at high liquid fractions (Wang et al., et al., 2016b).
2016). In order to perform the study under realistic conditions, and to
Despite the number of papers that have considered natural con- cover the temperature gap required for CSP-tower plants (i.e.,
vection, there are still several that have neglected it, as demonstrated in 286–565 °C), three commercially available PCMs based on eutectic
Table 1. Trp et al. (2006) compared a 2D-numerical model with ex- mixtures of carbonate salts were selected: H325, H425 and H525
perimental results and concluded that heat conduction controls the heat (http://www.pcmproducts.net/, 2018). The commercial Solar Salt, a
transfer inside the PCM region in both melting and solidification. Apart eutectic nitrate salt mixture, was selected as the HTF (Tehrani and
from this study, which appropriately justified neglecting convection, Taylor, 2016; Tehrani et al., 2016b). The thermophysical properties of
most studies provide no justification when invoking this assumption all materials used in this study are given in Table 3.
(Esen and Ayhan, 1996; Esen et al., 1998; Fan et al., 2014; Fang and
Chen, 2007; Li et al., 2013; Pirasaci and Goswami, 2016; Tao et al., 2.2. Case studies: geometric parameters employed
2012, 2014; Tao and Carey, 2016; Tehrani et al., 2016b). Alternatively,
some numerical studies do consider natural convection through the As demonstrated in a previous publication of the authors (Tehrani
effective thermal conductivity approach (Adine and El Qarnia, 2009; El et al., 2016b), the maximum potential of a shell-and-tube LHTES system
Qarnia, 2009; Pointner et al., 2016; Tao and He, 2011; Wang et al., is only achieved if the geometric parameters are selected appropriately.
2013, 2015). Again, in most cases, no justification was provided for this
approach. Neglecting convection simply for the sake of simplification – Table 2
while tempting – can be a major source of error. Operating conditions, adopted from (Tehrani et al., 2016a, 2016b, 2017).
From the literature, it can be concluded that there is a clear lack of Name Unit Value
direction on when convection can (or cannot) be neglected and as to
how much of a role it plays in a generalized shell-and-tube LHTES HTF inlet temperature in charging (top °C 565
injection)
system. The studies dealing with natural convection have been mostly
HTF inlet temperature in discharging °C 286
limited to a specific geometry or PCM (See Table 1). However, the re- (bottom injection)
lative importance of natural convection is clearly affected by geometry Mass flow rate (charging/discharging) kg/s 0.0025
and the PCM properties (Agyenim et al., 2009; Ismail and Melo, 1998; Flow regime on HTF side – Laminar (Reynolds < 2200)
Longeon et al., 2013; Ng et al., 1998; Patankar and Ftamadhyani, 1977; Nusselt number at HTF side (HTF/PCM – 3.66
interface)
Wang et al., 2012, 2016) and no study published to date has evaluated

491
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Table 3 for the interface between the HTF and PCM continuum zones.
Thermophysical properties of the HTF and PCMs. • Adiabatic walls were used for the remaining external surfaces of the
HTF and PCMs properties HTF (Bauer PCMs (http:// PCM sub-domain.
Units et al., 2012) www.pcmproducts.net/, 2018)
The continuity equation for incompressible flows was employed
Name Solar Salt H325 H425 H525 within the domain [21]:
Dynamic viscosity kg/m s 0.00326 0.0035 0.0035 0.0035
∇→
v =0 (1)
Melting point °C 325 425 525
Density kg/m3 1820 2110 2100 2350 The enthalpy-porosity method was used to model the melting and
Specific heat J/kg K 1553 1505 1535 1560 solidification processes of the PCM (Muhammad et al., 2015a, 2015b;
capacity
Thermal W/m K 0.52 0.549 0.565 0.565
Voller et al., 1987). The term H is calculated by the sum of sensible and
conductivity latent enthalpies, as shown in Eq. (3). Eqs. (1)–(4) are combined to
Latent heat of J/kg – 80,000 220,000 155,000 obtain Eq. (5).
fusion
∂ (ρH )
Lower melting °C – 323 423 523 + ∇ ·(ρ→
v H ) = ∇ . (k∇T )
point (Solidus) ∂t (2)
Upper melting °C – 328 428 528
point H = h + f ΔH (3)
(Liquidus)
Thermal expansion 1/K – 0.0005 0.0005 0.0005 T
coefficient
h = href + ∫T ref
CP dT
(4)

f=0 if T < Tm1


Based on the findings of that study and on the values covered by the T − Tm1
if Tm1 ⩽ T ⩽ Tm2
f= Tm2 − Tm1
reviewed literature (Table 2), the following ranges of geometric para-
R L f=1 if T > Tm2
meters were herein employed: 0.5 ≤ L ≤ 2, 1.5 ≤ r ≤ 3, 15 ≤ d ≤ 60 . (5)
o
VHTF
Besides, the minimum adopted value of porosity (ε = is typically ) ∂ (ρh) ∂ρf ΔH
+ ∇ . (ρ→ −∇ . (ρ→
Vtotal
0.2 (Flueckiger et al., 2014; Showalter and Kolb, 2002), so that was v h) = ∇. (k∇T)− v f ΔH )
∂t ∂t (6)
adopted herein. Each combination of the geometric parameters corre-
sponds to a certain Case ID in Table 4. The outcomes of this study will The general form of the momentum equation employed is shown in
be therefore extracted from 54 simulations (nine case studies for each of Eq. (7) [21]. Where τ̄¯ represents the viscous stress tensor, g is the ac-
the 3 PCMs for both melting and solidification). celeration due to gravity (with a value of −9.81 m/s2, according to the
coordinates of Fig. 1), and S′ is a source term defined below in Eq. (8):
3. Modelling approach ∂ (ρ→v)
+ ∇ (ρ→→
v v ) = −∇p + ∇ (¯τ¯) + ρg + S′
∂t (7)
The influence of natural convection was elucidated by comparing
the results of a pure conduction model with an approach that includes (1−f )2
S '= vAmush
both conduction and convection within the PCM (referred, respectively, f3 + ∊ (8)
to as conduction and convection models hereafter). As such, for each 5
case study (Table 4), two simulations were run, one for each approach. Default values of 0.001 and 10 were adopted for the terms ∊and
The details of each of the models employed are outlined below. Amush , respectively (Fluent, 2011). The viscous stress tensor (τ̄¯ ) present
in Eq. (7) was re-arranged to be adapted to incompressible fluids, be-
3.1. Mathematical domain, governing equations and boundary conditions coming μ∇2 v . Consequently, the final version of the momentum equa-
tion employed is shown Eq. (9), where β represents the volumetric
A commercial CFD software package, ANSYS Fluent 15.0, was em- expansion coefficient of the PCM.
ployed to simulate the LHTES system under both modelling approaches, ∂→v
using the Leonardi cluster at the University of New South Wales ρref ⎛⎜ +→
v ∇→
v ⎞⎟ = −∇p + ∇ (μ∇2 v ) + ρref g−ρref β (T −Tref ) g
⎝ ∂t ⎠
(UNSW) (Fluent, 2011). For the two cases, transient simulations were
carried out by solving the continuity and the Navier–Stokes equations (1−f )2
+ 3 vAmush
for the PCM and the HTF. An overview of the modelled domain with the f +∊ (9)
boundary conditions applied is shown in Fig. 2. According to the The main difference between the convection and the conduction
symmetric nature of the shell and tube geometry, a 2D axisymmetric approaches comes from including the buoyancy effects within the PCM.
CFD approach was adopted (Tao and He, 2015). Two continuum fluid
zones were defined: the PCM and the HTF. The HTF and the PCM (in its Table 4
liquid state) always stayed within the laminar regime, based on the Case ID and geometric specifications.
maximum Rayleigh number achieved for the geometries and conditions
Case ID L a (m) R/ ro L/d L/ R−ro R2 − ro2 ε
involved – i.e. lower than 109 for all cases, which is within the laminar S=
2ro L
regime (Ghajar, 2011). As illustrated in Fig. 2, the boundary conditions
for the melting process were defined as follows (note that for the soli- 1 2 3 60 60 0.033 0.11
dification process, the inlet/outlet boundary conditions and the flow 2 2 2 60 120 0.013 0.24
3 2 1.5 60 240 0.005 0.44
direction were reversed):
4 1 3 30 30 0.067 0.11
5 1 2 30 60 0.025 0.24
• The inlet mass flow rate and temperature were specified at the top 6
7
1
0.5
1.5
3
30
15
120
15
0.010
0.133
0.44
0.11
section of the inner tube, and the pressure outlet was specified at the
8 0.5 2 15 30 0.050 0.24
bottom section of the inner tube.

9 0.5 1.5 15 60 0.021 0.44
The geometry is axisymmetric around the centreline.
• The wall was thermally coupled (termed as a shadow wall in Fluent) a
For all cases: Re < 2200, ro = 0.0166 m , NP = 1.

492
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Fig. 2. CFD domain and boundary conditions during melting.

For the convection approach, the well-known Boussinesq approxima- associated with the medium mesh (90,000 cells) had less than ∼1%
tion was employed to compute natural convection within the PCM (Al- difference with the finest mesh, confirming the reliability of mesh se-
Abidi et al., 2013; Brent et al., 1988; Seddegh et al., 2015a). Con- lected (i.e., with all PCMs). Note that the melting time is sensitive to
versely, for the pure conduction approach, these buoyancy effects were geometry (Case ID) and the thermophysical properties of the PCM
not included in the model. employed.

3.2. Numerical schemes, convergence criteria, and other considerations


3.4. Model validation
A pressure-based double-precision solver was selected to solve the
To the best of the authors’ knowledge, no experimental data are
set of equations with a PISO pressure–velocity coupling scheme.
available in the literature for the operational temperatures involved
Second-order upwind discretization schemes were applied on the mo-
(See Table 1). Therefore, a system operating at lower temperatures was
mentum and energy equations, while a PRESTO scheme was employed
used for validation purposes. Simulated results were compared against
for the pressure equation because second order schemes are in-
the experimental work performed by Longeon et al. (2013).
compatible with the enthalpy-porosity approach in ANSYS Fluent
In that work, the LHTES system consisted of two concentric cylin-
(Seddegh et al., 2015a, 2016).
ders, where the outer cylinder had a diameter of 44 mm and the inner
The numerical convergence of the model was checked based on the
diameter was 15 mm. The length of the whole system was 400 mm. The
scaled numerical residuals of all computed variables. A value of 10−6
HTF selected was water, which flowed through the inner tube. The PCM
was chosen for the continuity residual. This value assured that the mass
consisted of technical grade paraffin (RT35 by Rubitherm (Longeon
imbalance was less than 0.2% of the net flux through the domain. The
et al., 2013)) and it was placed within the outer tube. Its phase change
velocity residuals were set to 10−6 and the energy residual value was
temperature ranges from 30 °C to 36 °C and the mass employed was
10−9. Temperature and velocity magnitudes were monitored each
0.480 kg. According to the Reynolds number obtained from the flow
iteration at several points of the mesh to check convergence as well.
rate of the HTF inside the tube, 0.01 m/s, the flow regime was laminar
An adaptive time stepping method was used with a minimum time
(Ghajar, 2011). The charging of the system was simulated by injecting
step size of 10−4 s and a maximum of 1 s. This configuration minimized
the HTF at 53 °C (by the top) to a system that had initial PCM and HTF
the error attained, especially at the beginning of the process, when
temperatures of 22 °C.
larger time steps would not be able to capture the high heat flux values
The comparison was based on the experimental temperatures re-
due to the large initial temperature gradients inside the system. A first-
gistered in (Longeon et al., 2013) for thermocouples located in different
order implicit transient formulation scheme was employed because
radial positions – points a, b and c, placed 3, 6 and 9 mm away from the
second-order schemes are incompatible with adaptive time stepping.
inner tube, respectively. Their axial position was Z/L = 0.5 (the middle
of the tube). The comparison between the experiment and the simula-
3.3. Meshes tion is shown in Fig. 4, which has suitable agreement. The model was
able to successfully predict the time when the melting front reached the
Different mapped grid meshes with orthogonal quadrilateral cells axial position of the thermocouples. That moment can be identified in
were employed depending on the geometry under evaluation. In the Fig. 4 as the time where the PCM temperatures increase abruptly (at
radial direction, every mesh had 40 cells in the PCM zone and 20 cells ∼1 h of charging time). Thus, it was concluded that our CFD model can
in the HTF zone. The divisions in the axial direction varied according to be employed confidently for the purposes of this study.
the length of each geometry: 1500, 750 and 375 divisions for those
lengths with 2, 1 and 0.5 m, respectively. According to Table 4, Case ID
1 is the largest unit that has been considered in this study, while the
Case ID 9 is the smallest module. Thus, the number of elements in the
meshes employed were 22,500 cells (Case IDs: 7–9), 45,000 (Case IDs:
4–6) and 90,000 (Case IDs: 1–3). Each mesh was refined close to the
boundary walls where larger temperature and velocity gradients are
expected.
The numerical uncertainty of the grid was evaluated by the Grid
Convergence Method (GCI) (Diarce et al., 2014) for all 3 mesh sizes. As
an illustration, a detailed view of the mesh employed for the largest
geometry is shown in Fig. 3 (90,000 cells). For the GCI analysis, three
more meshes were built for Case ID 1, having 126,000, 90,000 and
65,000 elements. The melting time was chosen as the key variable to
evaluate the error. The results showed that the numerical error (GCI) Fig. 3. Mesh employed for the outlet of the CFD domain (Case ID 1).

493
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

melting are depicted in Figs. 7 and 8, respectively. One significant


feature of Fig. 7 is that the solid/liquid interface has a ‘top-hat’ shape in
the presence of natural convection, whereas it has a parabola shape for
the conduction modelling approach. In the presence of natural con-
vection, the PCM melts mostly from top to the bottom. However, with
the conduction model, the PCM melts mostly in the axial direction,
from the hot inner surface (i.e., PCM/HTF interface) to the outer cold
surface. The conduction model does not show perfect axial melting
because of the non-uniform heat transfer rates caused by the initial
temperature gradient of the HTF as it is replaced by the hot fluid in-
troduced at the pipe inlet. It should be noted that these results are
consistent with the literature for HTF top injection in melting (Longeon
et al., 2013; Seddegh et al., 2015a).
Fig. 4. Validation of the CFD model against experimental data provided in
(Longeon et al., 2013). When comparing Figs. 7 and 8 it can be seen that the influence of
natural convection in the H325 system is more significant than the
H525 system. This is more easily noticed in the less radially shaped
4. Results and discussion: conduction vs. convection solid/liquid interface of H525 compared to the H325 system in the
convection model. This can be explained by the fact that the H525
All cases from Table 4 were simulated under the boundary condi- system has a very high melting point, allowing conduction to dominate
tions outlined in Section 2.1 for the three PCMs of Table 3. since melting takes much longer. Indeed, the melting of the system with
H325 is characterized by high heat fluxes since the PCM melts earlier.
4.1. Temperature profiles in melting/solidification According to Fig. 9, the heat transfer rate to the PCM has an initial peak
that decays over time (with much of the initial increase corresponding
This section investigates the behavior of the temperature profiles to the replacement of the initial cold HTF in the pipe). When the PCMs
under the two (conduction and convection) modelling approaches. Case start melting (after ∼1 and ∼5 h for the H325 and H525 systems, re-
IDs 1 and 3 as well as PCMs H525 and H325 were selected as re- spectively), the heat transfer rates to the PCM (i.e., the driving force of
presentative examples. Fig. 5 shows their timewise temperature dis- natural convection) is much higher for the system with H325 compared
tributions upon melting, while Fig. 6 depicts the evolution of the to that of H525. Thus, natural convection circulation is much more
temperatures during solidification. The following conclusions can be significant for low melting point PCMs. As such, for relatively high
drawn: melting point PCMs (i.e., when the melting point is close to the
The slope of the temperature profiles in the liquid phase is larger
compared to the solid phase (i.e., convection compared to conduction),
indicating that natural convection enhances the heat transfer rate in the
liquid phase. Thus, the melting process shows more discrepancy be-
tween the temperature profiles of conduction and convection modelling
approaches than the solidification process.
The influence of natural convection is a function of the geometric
parameters. The figures suggest that natural convection effect in Case
ID 1 is more significant than Case ID 3 (i.e., the deviation of the con-
vection profiles from the conduction profiles is lower in Case ID 3). This
is because of the larger non-dimensional radius of Case ID 1 (R/ ro=3, 3),
compared to Case ID 3 (R/ ro=1.5).
The influence of natural convection is not always identical at the top
and the bottom of the tank. In Fig. 5a, for instance, the difference be-
tween the conduction and the convection models is more significant at
the top of the tank, where the PCM melts much faster. Therefore, the
influence of natural convection cannot be evaluated at just one specific
location inside the PCM tank. Similarly, the influence of natural con-
vection must be monitored over the whole process, as the deviation
between the conduction and convection model changes from the be-
ginning to the end of the process.
The results also suggest that for a given geometry (e.g., Case ID 1),
the thermo-physical properties of the PCM significantly affect the pre-
sence of the natural convection. For example, the discrepancies be-
tween conduction and convection models are much higher for H325 as
compared to that of H525 (due to their 200 °C difference in melting
point), in both melting and solidification. The reasons for this are dis-
cussed in the following section.

4.2. Solid/liquid interface and heat transfer rates on melting

To visually demonstrate the differences between the conduction and


convection modelling approaches and to realize how the presence of
natural convection can be influenced by PCM selection, the behavior of
H325 and H525 were compared in detail for Case ID 1. The timewise Fig. 5. Melting temperature histories of the PCM at different axial locations for:
contours of the average liquid fraction (ALF) for H325 and H525 upon (a) Case ID 1-H525, (b) Case ID 3-H525, (c) Case ID 1-H325.

494
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

4.3. Effectiveness and error

In order to quantify the error between the conduction and convec-


tion models, a performance criterion is required. The authors suggest
that the storage effectiveness is the best performance metric for LHTES
systems as it reflects the real utilization of the PCM and covers the
whole range of sensible and latent phases of PCM. Effectiveness is de-
fined by the ratio of stored/delivered energy compared to the maximum
potential of the PCM, as shown in Eq. (10). Note that the transient value
of stored/delivered heat by PCM (QPCM ̇ ) was calculated from the CFD
simulations (i.e., the heat flux values of Fig. 9), while the maximum
potential of PCM (QPCM , max ) was case study dependent as described by
Eq. (11) and Table 5.
t
∫0 Q̇PCM (t ) dt
ε (t ) =
QPCM , max (10)

where

QPCM , max = mPCM CPCM , S (Tm1−Tinitial ) + mPCM ΔH


+ mPCM CPCM , L (Tinlet −Tm2) (11)

As can be seen in Table 5, Case ID 1 has the highest QPCM , max (due to
having the largest PCM volume), while Case ID 9 has the lowest
QPCM , max (due to having the minimum PCM volume). The amount of
PCM will clearly affect the potential for natural convection to play an
important role, but the volume/capacity of PCM is not the correct cri-
terion to assess the influence of natural convection. As will be shown in
this section, Case ID 7 most dramatically shows the presence of natural
convection.
The heat transfer rate and instantaneous effectiveness of H325 (Case
ID 1) are depicted in Fig. 10 for melting and in Fig. 11 for solidification.
In both conduction and convection models, the heat transfer rate is
Fig. 6. Solidification temperature histories of the PCM at different axial loca- maximum at the very beginning of the process because of the maximum
tions for: (a) Case ID 1-H525, (b) Case ID 3-H525, (c) Case ID 1-H325. temperature difference between the inner and outer wall of the annulus.
After this point, the heat transfer rate is decreasing over time because
maximum operating temperature of the system), there is little deviation the temperature difference between the inner and outer wall of annulus
between the conduction and convection models. decreases until they reach equilibrium, which is the end of charging/
While the temperature evolutions and solid/liquid interfaces dis- discharging process.
cussed so far give a qualitative representation of the ‘error’ in ne- Another feature of Figs. 10 and 11 is that the heat transfer rate in
glecting natural convection, they do not provide a quantitative mea- the convection model is not always higher than that of conduction
sure. In the next subsection, a method to quantify the error between model for both melting and solidification. This occurs because con-
choosing conduction versus convection models is proposed. vection allows the system to reach equilibrium faster. Thus, the effec-
tiveness for the convection model is always higher than that of the
conduction model at any given time. The time difference between the
two modelling approaches is a function of the effectiveness ratio, so the
error between two approaches can defined as follows:

Fig. 7. Solid/liquid interface during melting of H325 (Case ID 1) for conduction (left) and convection (right).

495
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Fig. 8. Solid/liquid interface during melting of H525 (Case ID 1) for conduction (left) and convection (right).

to reach full capacity (∼100% effectiveness) as often as possible.


However, in real operation the system may frequently undergo partial
charging and discharging (ε ≪ 100%) (Tehrani et al., 2017). Since the
objective in this study is to quantify the error of neglecting convection,
a value of 98% effectiveness was considered, as it provides a realistic
maximum error. According to Table 6, at effectiveness values lower
than 98%, the error decreases for the melting process, while it increases
for the solidification process. Note that the maximum error in melting is
always higher than that of solidification so that the priority was given
to melting. In the next subsection, general error correlations will be
presented to predict the error at 98% effectiveness.

Fig. 9. Heat flux to the PCM on melting for the convection modelling approach; 4.4. Error correlations at 98% effectiveness
H325 vs. H525.
So far, it has been demonstrated that the relative error between
Table 5 conduction and convection models is sensitive to: (1) geometry, (2)
Maximum stored/delivered energy. PCM, and (3) effectiveness ratio. Given the fact that the 98% effec-
Parameter QPCM , max [MJ ]
tiveness represents the maximum error that can occur in the simula-
tions (during melting), this section aims to incorporate these para-
ID case study PCM 1: H 325 PCM 2: H 425 PCM 3: H 525 meters to develop error correlations for researchers who wish to: (a)
decide whether to consider convection or not and/or (b) determine the
1 14.8 19 19.3
maximum error by neglecting convection.
2 5.5 7.1 7.2
3 2.3 2.9 3
4 7.3 9.5 9.7 4.4.1. Error as a function of geometry
5 2.7 3.5 3.6
6 1.1 1.4 1.5
Table 7 summarizes the error values presented in the different Case
7 3.6 4.7 4.8 IDs during melting and solidification. The maximum error observed was
8 1.3 1.7 1.8 100%, which corresponded to the melting of H325 in Case ID 7.
9 0.5 0.7 0.75 A regression analysis of the error values as a function of various
R−r L r
shell-and-tube LHTES geometric parameters (i.e., L o , R − r , Lo ) was
o

tconvection−tconduction performed. The results, which will be herein omitted for the sake of
Error (%) = × 100, where t = time to reach #% R2 − r 2
tconvection brevity, indicated that the specific geometry, S = 2r L o , might be a way
o
to lump together the effect of all geometric parameters (see Fig. 12a and
of effectiveness (12)
b), since as this parameter increases, the error increases. If the specific
In Table 6, the error between the conduction and the convection geometry is small enough (S < 0.005), natural convection can be safely
modelling approaches is summarized for various values of effectiveness. neglected (e.g. with an error of < 5%), for the PCM types and boundary
It is evident that for the melting process, the maximum error occurs at conditions tested in this study.
higher values of the effectiveness, while the maximum error in dis- In Fig. 12, a simple linear correlation is presented that is suitable for
charging occurs at lower values of effectiveness. This result can be each specified PCM only and is a function of geometry (i.e., the un-
explained by considering that during melting natural convection effect certainty of correlations is within 30% as shown by the error bars on the
increases as time progresses (i.e., the error accumulates from when graphs). It can also be observed that the error between the conduction
PCM first melts until the end of process), while during solidification it and the convection modelling approaches is lower when a PCM with a
has more influence at the beginning, when all the PCM is in the liquid higher relative melting point (i.e., close to the maximum operating
state. temperature) is implemented.
Without selecting an appropriate effectiveness value, it is not pos-
sible to determine the actual error between conduction and convection 4.4.2. Error as a function of geometry and PCM properties
modelling approaches. Commonly, LHTES systems should be designed Despite the monotonic effect of geometric parameters on error

496
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Fig. 10. Comparison of conduction and convection model simulations during melting of H325 in terms of (a) heat transfer rate and (b) effectiveness (ID case 1).

Fig. 11. Comparison of conduction and convection model simulations during solidification of H325 in terms of (a) heat transfer rate and (b) effectiveness (Case ID 1).

Table 6
Time needed to attain specific effectiveness values for melting and solidification (Case ID 1, PCM H325).
Melting Solidification

Effectiveness (%) Required time of conduction Required time of convection Error (%) Required time of conduction Required time of convection Error (%)
model (h) model (h) model (h) model (h)

98 13.11a 8.71 51 16.18 15.18 7


95 10.45 7.19 45 14.27 12.62 13
85 7.08 5.25 35 9.95 8.11 23
65 4.29 3.58 20 5.16 4.08 26
40 2.3 2.10 11 2.5 2.08 18

a
This is a very large melting time for most practical applications, but the aim in this study is to cover a wide range.

(Fig. 12), the value of error is different for the different PCMs. To in- hHTF Lc π (R2−ro2 ) L R2−ro2
Bi = , where Lc = =
vestigate the influence of PCM properties as well as geometric para- kPCM 2πro L 2ro (14)
meters on the error, with the results of Table 7, several sets of non-
dimensional groups that normally plays a role in the melting/solidifi- kHTF Nu R2−ro2
Bi = , where Nu = 3.66 in laminar flow
cation problems were analyzed through a multi-variant regression kPCM 4ro2 (15)
analysis in Excel as described by Table 8. The non-dimensional numbers
hHTF Lc, Nu
were defined as follows: Nu = , where Lc, Nu = 2ro
kHTF (16)
gβ (Tr = ro−Tm) (R−ro)3
Ra = Gr × Pr = , where Tr = ro = Thot = 565 ∘C CP, PCM (Thot −Tm)
υα Ste = , where Thot = 565 ∘C
ΔH (17)
(13)
It is worth noting that the Ra number was calculated when the PCMs
are in their liquid phase – i.e., at the end of the melting process and the

497
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Table 7
Error at 98% effectiveness for various Case IDs and PCMs.
Case ID Melting Solidification

Error (%)

H325 H425 H525 H325 H425 H525

1 51 22 7 7 7 1
2 23 9 2 10 7 1
3 8 4 0 4 1 0
4 73 36 13 22 8 2
5 35 16 4 12 5 1
6 13 5 0 5 2 0
7 100 53 21 28 10 3
8 48 24 7 16 6 1
9 19 9 1 7 3 1

Fig. 13. Accuracy of global error correlations versus the error calculated by
simulations: (a) melting, (b) solidification.

Table 9
Range of Stefan and ro/ L numbers used in this study.
ID case study ro/ L H325 H 425 H525

1, 2, 3 0.008 Ste = 4.47 Ste = 0.97 Ste = 0.4


4, 5, 6 0.01
7, 8, 9 0.03

Table 10
Range of Biot and Rayleigh numbers used in this study.
H325 H 425 H525

Fig. 12. Error at 98% effectiveness as a function of geometry for each particular ID case R/ ro Bi Ra Bi Ra Bi Ra
study
PCM: (a) melting, (b) solidification.
1, 4, 7 3 6.93 1.51 6.91 8.94 6.73 3.16 E + 07
beginning of the solidification process (when the error is maximum). E + 08 E + 07
2, 5, 8 2 2.59 1.88 2.58 1.12 2.52 3.95 E + 06
Based on the natural convection literature for the vertical annular space
E + 07 E + 07
(Yunus A. Cengel, 2015; Prasad and Kulacki, 1984; El-Shaarawi and Al- 3, 6, 9 1.5 1.08 2.35 1.07 1.40 1.05 4.94 E + 05
Nimr, 1990; Al-Arabi et al., 1987; Abouali and Falahatpisheh, 2009; E + 06 E + 06
Afrand et al., 2015; Sankar et al., 2011; Keyhani et al., 1983), the
characteristic length should be the distance between the cold and hot

Table 8
Results of multi-variant regression analysis on different non-dimensional groups for the melting process.
Group Non-dimensional numbers Multiple Ra R Squaredb Adjusted R Squaredc Averaged

1 Ra, ro/ L, Bi, Ste 0.97 0.89 0.87 0.91


2 Ra, ro/ L 0.84 0.71 0.68 0.74
3 Ra, ro/ L, Ste 0.92 0.85 0.83 0.86
4 Ra, S 0.80 0.64 0.60 0.68
5 Ra, S, Ste 0.91 0.83 0.80 0.85
6 Ra, R−ro/ L 0.80 0.65 0.62 0.69
7 Ra, R−ro/ L, Ste 0.91 0.83 0.81 0.85
8 Ra, ro/ L, kHTF / kHTF 0.92 0.86 0.84 0.87
9 Ra, ro/ L, Ste , R2−ro2/4ro2 0.94 0.88 0.86 0.89
10 Ra, ro/ L, R/ ro 0.94 0.89 0.87 0.90
11 Ra, ro/ L, R/ ro, Ste 0.94 0.89 0.87 0.90

a
Square root of R-squared.
b
R-squared is a statistical measure of how close the data are to the fitted regression line.
c
The adjusted R-squared is a modified version of R-squared for the number of predictors in a model.
d
Average of Multiple R, R squared, and Adjusted R Squared.

498
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

surfaces for the enclosure, so L = R−ro was selected here. It should be determines the error between conduction-only and full convection
noted that if tube length of L was used in Rayleigh number definition modelling approaches. From the results obtained, we can report the
instead of the annular space gap of R−ro , the range of Rayleigh number following conclusions:
could vary from 105 < Ra < 108 to 1011 < Ra < 1013 for the geome-
tries under evaluation, which means the flow inside PCM should have (1) To quantify the discrepancy between conduction and convection
been treated as turbulent instead of laminar. However, using tube modelling approaches, the error was defined as the difference be-
length in Rayleigh number calculations is a common practice for fluid tween conduction and convection models in terms of the time re-
inside a tube/cylinder (Lakeh et al., 2015; Evans et al., 1968) - that quired to reach 98% charging/discharging ratio. The errors were
should not apply to annular spaces. As for the temperature in Rayleigh clearly higher during the melting process compared to solidification
number definition, the temperature of inner surface depends on the – i.e. 100% versus 30% maximum error values, for the cases stu-
axial location of PCM cell. However, at the end of the melting process, died.
the inner wall temperature of the annulus has already reached the HTF (2) The error was higher when tube length (L ) decreases and when the
inlet temperature. Thus, Tr = ro = 838 K was assumed in Rayleigh number non-dimensional radius R increases. The proposed specific geo-
rO
calculations. R2 − r 2
metry parameter, S = 2r L o , was shown to provide a good indicator
According to Table 8, the best set of non-dimensional groups for o
for the overall effect of geometry on the error. Using this parameter,
error correlations is Group 1. The same groups were adopted for the
it was found that for S < 0.005 the natural convection could be
analysis of solidification results. The general error correlations attained
neglected for the boundary conditions and PCM properties range
for that Group 1, including the coefficients and exponents achieved by
used in this study.
regression, are shown in Eqs. (18) and (19):
(3) The results of a multi-variant regression analysis revealed that the
ro 0.92
Errormelting (%) = 2.5 × 10−11Ste−0.17Bi−3.4Ra2.1 ()L
global error correlations depends on the following non-dimensional
r
groups: Lo , Ra, Ste , Bi . The error correlations were established for
[range of this study melting and solidification and can be consulted in Eqs. (18) and
: 1 < Bi < 7, 105 < Ra < 108, 0.4 < Ste < 4.5, 0.008 < ro/ L < 0.03] (19).
(4) While the error correlations are appropriate to quantify the error, a
(18)
critical Rayleigh number, Racritical = 8 × 105 , was obtained below
r 0.47 which the error is < 1% and the natural convection is negligible. It
Errorsolidification (%) = 1.2 × 10−4Ste 0.33Bi−1.36Ra0.84 ⎛ o ⎞
⎝L⎠ (19) is also worth noting that low Ra numbers are associated with the
Fig. 13 illustrates the accuracy of the proposed correlations for high melting point PCMs (i.e. low Stefan numbers).
melting and solidification, respectively. The error lines in both figures
were set at 30%. Note that the error bars (lines) on Figs. 12 and 13 refer The outcomes of the present study also will help researchers to
to the difference percentage of the error calculated by the correlation identify cases where natural convection is negligible so that they can
and the one obtained from CFD simulations. This indicates that the use (much) faster conductivity models instead, to conduct coarse opti-
maximum uncertainty of the proposed correlations is within ∼30%. mizations for early design and feasibility analyses.
Note that the error correlations tend to overestimate the error at large
error values (e.g. > 100%), but at this point it is clear that convection Acknowledgments
cannot be neglected.
Tables 9 and 10 summarize the range of non-dimensional groups This work was supported by a grant from the Transfield Foundation
that were covered in this study. Research Grant – Round 2. Gonzalo Diarce would like to acknowledge
According to Tables 7–10, the following conclusions can be drawn: the Spanish's Ministry of Economy and Competitiveness (MINECO) for
financial support through the ekimProVe research project

• As tube length increase (e.g. r /L becomes small), the error de-


o
(ENE2015–71083–R) as well as the Erasmus Mundus Panther exchange
program. The authors also would like to acknowledge the Faculty of
creases.
• Both Rayleigh and Biot number are sensitive to the values of R andr . o
Engineering at the University of New South Wales (UNSW) for pro-
viding the Leonardi cluster without which this work would not have
Table 10 implies that the non-dimensional radius can be represented
by them. Thus, as R/ ro decreases, the Rayleigh and Biot also decrease been possible.
– along with the error.
• As Stefan number decreases, the error reduces. For the same case References
IDs, the Rayleigh and Biot numbers are less for H525 as compared to
Abouali, O., Falahatpisheh, A., 2009. Numerical investigation of natural convection of
H325, which explains why the natural convection is less significant
Al2O3 nanofluid in vertical annuli. Heat Mass Transf. 46 (1), 15.
when deploying H525. Adine, H.A., El Qarnia, H., 2009. Numerical analysis of the thermal behaviour of a shell-
• The error values can be best estimated by employing the general and-tube heat storage unit using phase change materials. Appl. Math. Model. 33 (4),
2132–2144.
error correlations (Eqs. (18) and (19)). However, as all of these
Afrand, M., Farahat, S., Nezhad, A.H., Sheikhzadeh, G.A., Sarhaddi, F., Wongwises, S.,
parameters are linked with each other, the Rayleigh number (e.g. as 2015. Multi-objective optimization of natural convection in a cylindrical annulus
defined by Eq. (13)) can be selected as a representative of all in- mold under magnetic field using particle swarm algorithm. Int. Commun. Heat Mass
fluential parameters in the error value. For Ra < 8 × 105 (considered Transfer 60, 13–20.
Agyenim, F., Eames, P., Smyth, M., 2009. A comparison of heat transfer enhancement in a
here as a critical Rayleigh number), natural convection can be ne- medium temperature thermal energy storage heat exchanger using fins. Sol. Energy
glected (since the error in melting/solidification is < 1%). Note that 83 (9), 1509–1520.
this value was obtained through the linear interpolation of the error Agyenim, F., Hewitt, N., Eames, P., Smyth, M., 2010. A review of materials, heat transfer
and phase change problem formulation for latent heat thermal energy storage sys-
values for H525 in regards to the Rayleigh numbers indicated on tems (LHTESS). Renew. Sustain. Energy Rev. 14 (2), 615–628.
Table 10. Akgün, M., Aydın, O., Kaygusuz, K., 2007. Experimental study on melting/solidification
characteristics of a paraffin as PCM. Energy Convers. Manage. 48 (2), 669–678.
Al-Abidi, A.A., Mat, S., Sopian, K., Sulaiman, M.Y., Mohammad, A.T., 2013. Numerical
5. Conclusions study of PCM solidification in a triplex tube heat exchanger with internal and ex-
ternal fins. Int. J. Heat Mass Transf. 61, 684–695.
This study defines the conditions under which natural convection is Al-Arabi, M., El-Shaarawi, M., Khamis, M., 1987. Natural convection in uniformly heated
vertical annuli. Int. J. Heat Mass Transf. 30 (7), 1381–1389.
significant for high temperature shell-and-tube LHTES systems, and

499
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

Bauer, T., Breidenbach, N., Pfleger, N., Laing, D., Eckand, M., 2012. Overview of molten Liu, M., Tay, N.S., Bell, S., Belusko, M., Jacob, R., Will, G., Saman, W., Bruno, F., 2016.
salt storage systems and material development for solar thermal power plants. In: Review on concentrating solar power plants and new developments in high tem-
Proceedings of the 2012 National Solar Conference for (SOLAR 2012), Denver. perature thermal energy storage technologies. Renew. Sustain. Energy Rev. 53,
Bellecci, C., Conti, M., 1993a. Phase change thermal storage: transient behaviour analysis 1411–1432.
of a solar receiver/storage module using the enthalpy method. Int. J. Heat Mass Longeon, M., Soupart, A., Fourmigué, J.-F., Bruch, A., Marty, P., 2013. Experimental and
Transf. 36 (8), 2157–2163. numerical study of annular PCM storage in the presence of natural convection. Appl.
Bellecci, C., Conti, M., 1993b. Transient behaviour analysis of a latent heat thermal Energy 112, 175–184.
storage module. Int. J. Heat Mass Transf. 36 (15), 3851–3857. Mahdavi, M., Qiu, S., Tiari, S., 2016. Improvement of a novel heat pipe network designed
Brent, A., Voller, V., Reid, K.t.J., 1988. Enthalpy-porosity technique for modeling con- for latent heat thermal energy storage systems. Appl. Therm. Eng. 108, 878–892.
vection-diffusion phase change: application to the melting of a pure metal. Numeric. Mendes, P.S., Brasil, A.P., 1987. Heat transfer during melting around an isothermal
Heat Transf., Part A Appl. 13 (3), 297–318. vertical cylinder. J. Heat Transfer 109 (4), 961–964.
Cárdenas, B., León, N., 2013. High temperature latent heat thermal energy storage: phase Muhammad, M., Badr, O., Yeung, H., 2015a. CFD modeling of the charging and dis-
change materials, design considerations and performance enhancement techniques. charging of a shell-and-tube latent heat storage system for high-temperature appli-
Renew. Sustain. Energy Rev. 27, 724–737. cations. Numeric. Heat Transf., Part A: Appl. 68 (8), 813–826.
Dhaidan, N.S., Khodadadi, J., 2015. Melting and convection of phase change materials in Muhammad, M., Badr, O., Yeung, H., 2015b. Validation of a CFD melting and solidifi-
different shape containers: a review. Renew. Sustain. Energy Rev. 43, 449–477. cation model for phase change in vertical cylinders. Numeric. Heat Transf., Part A:
Diarce, G., Campos-Celador, Á., Martin, K., Urresti, A., García-Romero, A., Sala, J., 2014. Appl. 68 (5), 501–511.
A comparative study of the CFD modeling of a ventilated active façade including Mostafavi Tehrani, S. Saeed, Shoraka, Yashar, Nithyanandam, Karthik, Taylor, Robert,
phase change materials. Appl. Energy 126, 307–317. 2018. Transient modelling of shell- and -tube and packed bed PCM storage systems
Dincer, I., Rosen, M., 2002. Thermal Energy Storage: Systems and Applications. John integrated with a CSP plant: a techno-economic comparison with sensible alter-
Wiley & Sons. natives. Appl. Energy (in press).
Dutil, Y., Rousse, D.R., Salah, N.B., Lassue, S., Zalewski, L., 2011. A review on phase- Ng, K., Gong, Z., Mujumdar, A., 1998. Heat transfer in free convection-dominated melting
change materials: mathematical modeling and simulations. Renew. Sustain. Energy of a phase change material in a horizontal annulus. Int. Commun. Heat Mass Transfer
Rev. 15 (1), 112–130. 25 (5), 631–640.
El-Shaarawi, M., Al-Nimr, M., 1990. Fully developed laminar natural convection in open- Nithyanandam, K., Pitchumani, R., 2011. Analysis and optimization of a latent thermal
ended vertical concentric annuli. Int. J. Heat Mass Transf. 33 (9), 1873–1884. energy storage system with embedded heat pipes. Int. J. Heat Mass Transf. 54 (21),
El Qarnia, H., 2009. Numerical analysis of a coupled solar collector latent heat storage 4596–4610.
unit using various phase change materials for heating the water. Energy Convers. Nithyanandam, K., Pitchumani, R., 2014. Cost and performance analysis of concentrating
Manage. 50 (2), 247–254. solar power systems with integrated latent thermal energy storage. Energy 64,
Esen, M., Ayhan, T., 1996. Development of a model compatible with solar assisted cy- 793–810.
lindrical energy storage tank and variation of stored energy with time for different Patankar, S., Ftamadhyani, S., 1977. Analwsis of melting in the presence of natural
phase change materials. Energy Convers. Manage. 37 (12), 1775–1785. confection in the melt region.
Esen, M., Durmuş, A., Durmuş, A., 1998. Geometric design of solar-aided latent heat store Pirasaci, T., Goswami, D.Y., 2016. Influence of design on performance of a latent heat
depending on various parameters and phase change materials. Sol. Energy 62 (1), storage system for a direct steam generation power plant. Appl. Energy 162,
19–28. 644–652.
Evans, L., Reid, R., Drake, E., 1968. Transient natural convection in a vertical cylinder. Pointner, H., De Gracia, A., Vogel, J., Tay, N., Liu, M., Johnson, M., Cabeza, L.F., 2016.
AIChE J. 14 (2), 251–259. Computational efficiency in numerical modeling of high temperature latent heat
Fan, Z., Ferreira, C.I., Mosaffa, A., 2014. Numerical modelling of high temperature latent storage: comparison of selected software tools based on experimental data. Appl.
heat thermal storage for solar application combining with double-effect H2O/LiBr Energy 161, 337–348.
absorption refrigeration system. Sol. Energy 110, 398–409. Prasad, V., Kulacki, F., 1984. Natural convection in a vertical porous annulus. Int. J. Heat
Fang, M., Chen, G., 2007. Effects of different multiple PCMs on the performance of a Mass Transf. 27 (2), 207–219.
latent thermal energy storage system. Appl. Therm. Eng. 27 (5), 994–1000. Riahi, S., Saman, W.Y., Bruno, F., Belusko, M., Tay, N., 2017. Impact of periodic flow
Flueckiger, S.M., Iverson, B.D., Garimella, S.V., Pacheco, J.E., 2014. System-level simu- reversal of heat transfer fluid on the melting and solidification processes in a latent
lation of a solar power tower plant with thermocline thermal energy storage. Appl. heat shell and tube storage system. Appl. Energy 191, 276–286.
Energy 113, 86–96. Sankar, M., Park, Y., Lopez, J., Do, Y., 2011. Numerical study of natural convection in a
Fluent, A., 2011. ANSYS Fluent User's Guide. Ansys Inc, p. 2427. vertical porous annulus with discrete heating. Int. J. Heat Mass Transf. 54 (7–8),
Fornarelli, F., Camporeale, S., Fortunato, B., Torresi, M., Oresta, P., Magliocchetti, L., 1493–1505.
Miliozzi, A., Santo, G., 2016. CFD analysis of melting process in a shell-and-tube Seddegh, S., Tehrani, S.S.M., Wang, X., Cao, F., Taylor, R.A., 2018. Comparison of heat
latent heat storage for concentrated solar power plants. Appl. Energy 164, 711–722. transfer between cylindrical and conical vertical shell-and-tube latent heat thermal
Ghajar, A.C.Y.a.A.J., 2011. Heat and Mass Transfer: Fundamentals & Applications, fourth energy storage systems. Appl. Therm. Eng. 130, 1349–1362.
ed. McGraw-Hill, New York. Seddegh, S., Wang, X., Henderson, A.D., 2015a. Numerical investigation of heat transfer
Gil, A., Medrano, M., Martorell, I., Lazaro, A., Dolado, P., Zalba, B., Cabeza, L.F., 2010. mechanism in a vertical shell and tube latent heat energy storage system. Appl.
State of the art on high temperature thermal energy storage for power generation. Therm. Eng. 87, 698–706.
Part 1—Concepts, materials and modellization. Renew. Sustain. Energy Rev. 14 (1), Seddegh, S., Wang, X., Henderson, A.D., 2016. A comparative study of thermal behaviour
31–55. of a horizontal and vertical shell-and-tube energy storage using phase change ma-
http://www.pcmproducts.net/, 2018. <http://www.pcmproducts.net/>, LAST terials. Appl. Therm. Eng. 93, 348–358.
ACCESSED ON 15/04/2018. (Accessed 15/04/2018 2018). Seddegh, S., Wang, X., Henderson, A.D., Xing, Z., 2015b. Solar domestic hot water sys-
Ismail, K., Melo, C., 1998. Convection-based model for a PCM vertical storage unit. Int. J. tems using latent heat energy storage medium: a review. Renew. Sustain. Energy Rev.
Energy Res. 22 (14), 1249–1265. 49, 517–533.
Jacob, R., Belusko, M., Fernández, A.I., Cabeza, L.F., Saman, W., Bruno, F., 2016. Showalter, S.K., Kolb, W.J., 2002. Development of a molten-salt thermocline thermal
Embodied energy and cost of high temperature thermal energy storage systems for storage system for parabolic trough plants.
use with concentrated solar power plants. Appl. Energy 180, 586–597. Tao, Y.-B., Li, M.-J., He, Y.-L., Tao, W.-Q., 2014. Effects of parameters on performance of
Kalhori, B., Ramadhyani, S., 1985. Studies on heat transfer from a vertical cylinder, with high temperature molten salt latent heat storage unit. Appl. Therm. Eng. 72 (1),
or without fins, embedded in a solid phase change medium. J. Heat Transfer 107 (1), 48–55.
44–51. Tao, Y., Carey, V., 2016. Effects of PCM thermophysical properties on thermal storage
Kemink, R., Sparrow, E., 1981. Heat transfer coefficients for melting about a vertical performance of a shell-and-tube latent heat storage unit. Appl. Energy 179, 203–210.
cylinder with or without subcooling and for open or closed containment. Int. J. Heat Tao, Y., He, Y., 2011. Numerical study on thermal energy storage performance of phase
Mass Transf. 24 (10), 1699–1710. change material under non-steady-state inlet boundary. Appl. Energy 88 (11),
Keyhani, M., Kulacki, F., Christensen, R., 1983. Free convection in a vertical annulus with 4172–4179.
constant heat flux on the inner wall. J. Heat Transfer 105 (3), 454–459. Tao, Y., He, Y., 2015. Effects of natural convection on latent heat storage performance of
Kuravi, S., Trahan, J., Goswami, D.Y., Rahman, M.M., Stefanakos, E.K., 2013. Thermal salt in a horizontal concentric tube. Appl. Energy 143, 38–46.
energy storage technologies and systems for concentrating solar power plants. Prog. Tao, Y., He, Y., Qu, Z., 2012. Numerical study on performance of molten salt phase
Energy Combust. Sci. 39 (4), 285–319. change thermal energy storage system with enhanced tubes. Sol. Energy 86 (5),
Lacroix, M., 1993. Numerical simulation of a shell-and-tube latent heat thermal energy 1155–1163.
storage unit. Sol. Energy 50 (4), 357–367. Tehrani, S.S.M., Saffar-Avval, M., Kalhori, S.B., Mansoori, Z., Sharif, M., 2013a. Hourly
Lakeh, R.B., Lavine, A.S., Kavehpour, H.P., Wirz, R.E., 2015. Study of turbulent natural energy analysis and feasibility study of employing a thermocline TES system for an
convection in vertical storage tubes for supercritical thermal energy storage. integrated CHP and DH network. Energy Convers. Manage. 68, 281–292.
Numeric. Heat Transf., Part A: Appl. 67 (2), 119–139. Tehrani, S.S.M., Saffar-Avval, M., Mansoori, Z., Kalhori, S.B., Abbassi, A., Dabir, B.,
Li, Q., Tehrani, S.S.M., Taylor, R.A., 2017. Techno-economic analysis of a concentrating Sharif, M., 2013b. Development of a CHP/DH system for the new town of Parand: an
solar collector with built-in shell and tube latent heat thermal energy storage. Energy. opportunity to mitigate global warming in Middle East. Appl. Therm. Eng. 59 (1),
Li, Y., He, Y., Song, H., Xu, C., Wang, W., 2013. Numerical analysis and parameters op- 298–308.
timization of shell-and-tube heat storage unit using three phase change materials. Tehrani, S.S.M., Shoraka, Y., Diarce, G., Taylor, R.A., 2018a. An improved, generalized
Renew. Energy 59, 92–99. effective thermal conductivity method for rapid design of high temperature shell-and-
Liu, M., Saman, W., Bruno, F., 2012. Review on storage materials and thermal perfor- tube latent heat thermal energy storage systems. Renew. Energy.
mance enhancement techniques for high temperature phase change thermal storage Tehrani, S.S.M., Shoraka, Y., Nithyanandam, K., Taylor, R.A., 2018b. Cyclic performance
systems. Renew. Sustain. Energy Rev. 16 (4), 2118–2132. of cascaded and multi-layered solid-PCM shell-and-tube thermal energy storage

500
S.S. Mostafavi Tehrani et al. Solar Energy 174 (2018) 489–501

systems: a case study of the 19.9 MW e Gemasolar CSP plant. Appl. Energy 228, thermal energy storage unit. Appl. Therm. Eng. 26 (16), 1830–1839.
240–253. Voller, V., Cross, M., Markatos, N., 1987. An enthalpy method for convection/diffusion
Tehrani, S.S.M., Taylor, R.A., 2016. Off-design simulation and performance of molten salt phase change. Int. J. Numer. Meth. Eng. 24 (1), 271–284.
cavity receivers in solar tower plants under realistic operational modes and control Wang, S., Faghri, A., Bergman, T.L., 2012. Melting in cylindrical enclosures: numerical
strategies. Appl. Energy 179, 698–715. modeling and heat transfer correlations. Numeric. Heat Transf., Part A: Appl. 61 (11),
Tehrani, S.S.M., Taylor, R.A., Ghazani, A.S., Saberi, P., 2016a. Part load behavior of 837–859.
molten salt cavity receiver solar tower plants under storage mode operational mode. Wang, W.-W., Wang, L.-B., He, Y.-L., 2015. The energy efficiency ratio of heat storage in
In: ASME 2016 10th International Conference on Energy Sustainability collocated one shell-and-one tube phase change thermal energy storage unit. Appl. Energy 138,
with the ASME 2016 Power Conference and the ASME 2016 14th International 169–182.
Conference on Fuel Cell Science, Engineering and Technology. American Society of Wang, W.-W., Zhang, K., Wang, L.-B., He, Y.-L., 2013. Numerical study of the heat
Mechanical Engineers, pp. V001T004A027–V001T004A027. charging and discharging characteristics of a shell-and-tube phase change heat sto-
Tehrani, S.S.M., Taylor, R.A., Nithyanandam, K., Ghazani, A.S., 2017. Annual compara- rage unit. Appl. Therm. Eng. 58 (1), 542–553.
tive performance and cost analysis of high temperature, sensible thermal energy Wang, Y., Wang, L., Xie, N., Lin, X., Chen, H., 2016. Experimental study on the melting
storage systems integrated with a concentrated solar power plant. Sol. Energy 153, and solidification behavior of erythritol in a vertical shell-and-tube latent heat
153–172. thermal storage unit. Int. J. Heat Mass Transf. 99, 770–781.
Tehrani, S.S.M., Taylor, R.A., Saberi, P., Diarce, G., 2016b. Design and feasibility of high Xiao, X., Zhang, P., 2015a. Numerical and experimental study of heat transfer char-
temperature shell and tube latent heat thermal energy storage system for solar acteristics of a shell-tube latent heat storage system: Part I-charging process. Energy
thermal power plants. Renew. Energy 96, 120–136. 79, 337–350.
Tiari, S., Qiu, S., 2015. Three-dimensional simulation of high temperature latent heat Xiao, X., Zhang, P., 2015b. Numerical and experimental study of heat transfer char-
thermal energy storage system assisted by finned heat pipes. Energy Convers. acteristics of a shell-tube latent heat storage system: Part II–discharging process.
Manage. 105, 260–271. Energy 80, 177–189.
Trp, A., 2005. An experimental and numerical investigation of heat transfer during Cengel, Yunus A., 2015. Heat and Mass Transfer: Fundamentals & Applications, fifth ed.
technical grade paraffin melting and solidification in a shell-and-tube latent thermal McGraw Hill Education, New York.
energy storage unit. Sol. Energy 79 (6), 648–660. Zalba, B., Marı́n, J.M., Cabeza, L.F., Mehling, H., 2003. Review on thermal energy storage
Trp, A., Lenic, K., Frankovic, B., 2006. Analysis of the influence of operating conditions with phase change: materials, heat transfer analysis and applications. Appl. Therm.
and geometric parameters on heat transfer in water-paraffin shell-and-tube latent Eng. 23 (3), 251–283.

501

You might also like