You are on page 1of 19

Available online at www.sciencedirect.

com

ScienceDirect
Solar Energy 107 (2014) 770–788
www.elsevier.com/locate/solener

Optimization of an encapsulated phase change material thermal energy


storage system
K. Nithyanandam, R. Pitchumani ⇑
Advanced Materials and Technologies Laboratory, Department of Mechanical Engineering, Virginia Tech, Blacksburg, VA 24061-0238, United States

Received 10 December 2013; received in revised form 8 May 2014; accepted 8 June 2014

Communicated by: Associate Editor Halime Paksoy

Abstract

Thermal energy storage enables uninterrupted operation of a concentrating solar power (CSP) plant during periods of cloudy or inter-
mittent solar availability. This paper presents a transient computational analysis of an encapsulated phase change material (EPCM) ther-
mal energy storage (TES) system for repeated charging and discharging cycles to investigate its dynamic response. The influence of the
design and operating parameters on the dynamic charge and discharge performance of the system is analyzed to identify operating win-
dows that satisfy technoeconomic targets of storage cost less than $15/kW ht, round-trip exergetic efficiency greater than 95%, charge
time less than 6 h and a minimum discharge period of 6 h. Overall, this study illustrates a methodology for design and optimization
of encapsulated PCM thermal energy storage system (EPCM-TES) for a CSP plant operation.
Ó 2014 Elsevier Ltd. All rights reserved.

Keywords: Thermal energy storage; Encapsulated phase change materials; Concentrating solar power; Design window; Optimization

1. Introduction heat significantly increases the energy storage capacity


and reduces size of the TES system. However, a major tech-
Thermal energy storage system is essential to make con- nology barrier that is limiting the use of latent thermal
centrating solar power (CSP) generation competitive for energy of PCM is the higher thermal resistance provided
meeting current and future energy needs. Storing energy by its intrinsically low thermal conductivity, thus requiring
for future use allows the power plant to operate continu- large heat transfer surface area of interaction. Several tech-
ously during periods of intermittent sun, reduces the mis- niques to improve the thermal performance of latent ther-
match between the energy supply and demand by mal energy storage systems are reported in the literature;
providing load leveling and helps to conserve energy by notable among them are embedding heat pipes or ther-
improving the reliability and performance of CSP plants. mosyphons between the HTF and PCM (Nithyanandam
Thermal energy can be stored as either sensible or latent and Pitchumani, 2011, 2013a, 2014a, 2013b), dispersing
heat (Stekli et al., 2013). Most of the thermal energy high conductivity particles in the PCM (Mettawee and
storage systems in operation are based on sensible heat Assassa, 2007) and storing PCM within the framework of
storage. Storing heat in the form of latent heat of fusion porous metal foams (Nithyanandam and Pitchumani,
of phase change material (PCM) in addition to sensible 2014b; Zhao and Wu, 2011). A review of various tech-
niques employed to enhance the performance of high tem-
⇑ Corresponding author. Tel.: +1 540 231 1776; fax: +1 540 231 9100. perature latent thermal energy storage system is discussed
E-mail address: pitchu@vt.edu (R. Pitchumani). in Cárdenas and León (2013). One promising approach is

http://dx.doi.org/10.1016/j.solener.2014.06.011
0038-092X/Ó 2014 Elsevier Ltd. All rights reserved.
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 771

Nomenclature

b capsule wall thickness (m) 2 2 PCM cascaded EPCM-TES system


b* dimensionless capsule wall thickness 3 3 PCM cascaded EPCM-TES system
c specific heat (J/kg K) B bottom
C* cost per unit mass ($/kg) c capsule
C0 cost per unit length ($/m) C charging
C 00 cost per unit area ($/m2) D discharging
fL liquid fraction encap encapsulation
h convective heat transfer coefficient (W/m2 K) f heat transfer fluid
hsl latent heat of fusion of PCM (J/kg) F foundation
H height (m) I insulation
k thermal conductivity (W/m K) L latent
k* dimensionless thermal conductivity M middle
m mass flow rate (kg/s) p phase change material
Nu Nusselt number SS stainless steel
Pe electrical power output (MWe) t tank
Pr Prandtl number T top
Pth thermal power output (MWt) eff effective
Q energy (MJ)
r radial direction Greek Symbols
R radius (m) a storage capital cost ($/kW ht)
R* dimensionless radius c melt fraction
Re Reynolds number e porosity of the packed bed
t time (s) f exergetic efficiency (%)
t* dimensionless time h dimensionless temperature
T temperature (K) hm dimensionless melt temperature
Tm melting temperature (K) h0C dimensionless charging cut-off temperature
U superficial velocity of heat transfer fluid (m/s) h0D dimensionless discharging cut-off temperature
w thickness (m) k capacitance ratio
z axial direction l dynamic viscosity (kg/m s)
q density (kg/m3)
Subscripts and Superscripts w inverse Stefan number
1 1 PCM non-cascaded EPCM-TES system

to increase the heat transfer area between the HTF and in the thermocline storage system till date. Felix Regin et al.
PCM by incorporating the PCM mixture in small capsules (2009) also reported the modeling of thermocline energy
(Nithyanandam and Pitchumani, 2014c; Mathur et al., storage system with embedded PCM capsules. Wu and
2013; Pendyala, 2012). Fang (2011) analyzed the discharging characteristics of a
Several reports on the numerical modeling of sensible heat solar heat storage system with a packed bed of spherical cap-
storage in packed beds are found in the literature sules filled with myristic acid as PCM. The influence of HTF
(Schumann, 1929; Shitzer and Levy, 1983; Yang mass flow rate, inlet temperature and the porosity of packed
and Garimella, 2010; Beasley and Clark, 1984; Yang and bed were studied.
Garimella, 2010; Van Lew et al., 2011; Yang and In developing thermal storage technologies, the exerget-
Garimella, 2013; Wakao and Kagei, 1982). Ismail ic efficiency is sought to be high to ensure that heat quality
and Henriquez (1999) presented a mathematical model for is maintained after storage (Stekli et al., 2013). Previous
predicting the thermal performance of cylindrical storage investigations focused on latent thermal energy storage sys-
tank containing spherical capsules filled with water as tems have shown that cascading several PCMs in the order
PCM. The model was used to investigate the influence of of their decreasing melt temperatures from the hot HTF
the working fluid inlet temperature, flow rate and the mate- inlet side can result in higher heat transfer rates, as well
rial of spherical capsules of fixed (77 mm) diameter on the as improved exergetic efficiency due to a more uniform
solidification process. Felix Regin et al. (2008) and Singh temperature difference between the hot and cold media
et al. (2011) presented a brief review of the works performed (Wang et al., 1999; Gong and Mujumdar, 1997). For
772 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

CSP systems to be cost-competitive with other sources of


energy generation on the grid, without subsidy, the U.S.
Department of Energy’s SunShot Initiative calls for ther-
mal energy storage systems to have round-trip exergetic
efficiency greater than 95% and a storage capital cost less
than $15/kW ht for a minimum (maximum) discharge
(charge) period of 6 h (Stekli et al., 2013; SunShot, 2012).
The optimal design of an EPCM-TES system that meets
the above technoeconomic requirements is, therefore, of
considerable significance, and forms the focus of the pres-
ent study.
Although several articles on the thermocline tank
packed with sensible filler materials are reported, relatively
few works on the performance of an EPCM-TES system is
found in the literature. To the authors’ knowledge, a com-
prehensive study of the dynamic performance of the system
subjected to constraints dictated by the power plant opera-
tion is lacking in the literature. To this end, the objective of Fig. 1. Schematic illustration of: (a) encapsulated PCM thermal energy
the present study is to investigate the influence of the design storage system (EPCM-TES) and (b) capsule cross-section.
and operating parameters on the dynamic energy and exer- zones in a 3-PCM cascaded EPCM-TES system are repre-
gy performance of the EPCM-TES system using the model sented by H3,T, H3,M and H3,B, respectively (Fig. 1a). The
developed by the authors in Nithyanandam and inner radius of the capsules, represented by Rc, is filled with
Pitchumani (2014c). Based on the parametric studies and PCM, while the thickness of the capsule wall is denoted by
specified constraints on the minimum exergetic efficiency, b, as depicted in Fig. 1b.
maximum storage capital cost, minimum discharge time The non-dimensional governing energy equations for
and maximum charge time (SunShot, 2012), operating win- the HTF temperature, hf, and the encapsulated PCM tem-
dows are identified as a function of the various design and perature, hp, provided in Nithyanandam and Pitchumani
operating parameters for cascaded and non-cascaded (2014c) are as follows:
EPCM-TES system. A further contribution of the study
involves finding the optimum combination of design and @hf @hf @ 2 hf 3ð1  eÞ
e þ Re H Pr f ¼ 2 þ 
operating parameters of the EPCM-TES system subject @t @z @z ðRc þ b Þ3
to the aforementioned constraints. hp;r ¼Rc  hf
The paper is organized as follows: The mathematical n 
o n o ð1Þ
b 2
model, as developed in Nithyanandam and Pitchumani    
k w Rc ðRc þb Þ
þ  
NuðRc þb Þ
(2014c), is introduced in the next section followed by a dis-     
cussion of the results of the parametric studies in Section 3. @fL ðhp Þ @hp k p Nuconv @ 2 @hp
1þw ¼ r ð2Þ
Section 4 presents the optimization problem formulation @hp @t k @r @r
followed by discussion of the optimization results. where the nondimensional parameters are defined as
follows:
2. Numerical analysis
z  r  kf t b Rc
z ¼ ;r ¼ ;t ¼ 2
; b ¼ ; Rc ¼ ;
2.1. Model development Ht Ht qf c f H t H t H t

Tf  TC Tp  TC
hf ¼ ; hp ¼ ;
Fig. 1a illustrates the schematic of an EPCM-TES tank TC  TD TC  TD
of height Ht and radius Rt packed with spherical capsules hsl qp cp  k w  k p
filled with PCM. For the sake of clarity in illustration, an w¼ ;k ¼ ;k ¼ ;k ¼ ;
cp ðT C  T D Þ qf cf w k f p k f
ordered arrangement of capsules is depicted in Fig. 1a;
qf UH t lf cf
however, in reality, the packing scheme may vary and the ReH ¼ ; Prf ¼ ;
porosity of the packed bed for a fixed diameter of spherical lf kf
capsules can range from 0.25 to 0.48 (Hales, 2006). The hH t 1=2 1=3
Nu ¼ ¼ 2 þ 2:87½ðRc þ b ÞReH  Prf
solid arrows indicate the direction of hot HTF during kf
charging while the dotted arrows indicate the direction of þ 0:098ðRc þ b ÞReH Prf
1=2
ð3Þ
cold HTF flow through the storage system during discharg-
ing. For a 2-PCM cascaded configuration, the total tank The terms q, c, k and T correspond to density, specific
height is divided into two and the heights of the top and heat, thermal conductivity, and temperature, respectively,
bottom zone are denoted by H2,T and H2,B, respectively of either the HTF denoted by subscript f or the encapsu-
in Fig. 1a. Similarly, the height of top, middle and bottom lated PCM denoted by subscript p, hsl denotes the latent
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 773

heat of fusion of the PCM, t denotes time, U represents the fL = 0, hp < hm and fL = 1, hp > hm, where hm is the melt
superficial velocity of the heat transfer fluid entering the temperature of PCM. The region between fL = 0–1, termed
packed bed storage system, h represents the convective heat the mushy zone, is modeled as a pseudo-porous medium, in
transfer coefficient between the HTF and PCM capsules which the porosity increases from 0 to 1 as the PCM
and e denotes the porosity of the packed bed. The convec- absorbs latent heat at a constant temperature and melts.
tion Nusselt number, Nuconv determines the effective ther- In the present case, the enthalpy iterative updating scheme
mal conductivity of the PCM due to the buoyancy driven of the liquid fraction takes the following form for a pure
convection currents during melting of the PCM, which is PCM, which melts at constant temperature:
explained in detail in Nithyanandam and Pitchumani ai h i
(2014c). The interstitial Nusselt number, Nu, thermally
nþ1
wfLðiÞ n
¼ wfLðiÞ þ 0 d hnpðiÞ  F 1 ðfLðiÞ
n
Þ ð6Þ
ai
couples the HTF with the spherical PCM capsules. The
fluid-particle heat transfer interaction in packed beds has In the above equation, ai is the coefficient of hp(i) for the
been extensively studied and numerous correlations for radial nodal point i in the discretized energy equation of
the interstitial heat transfer coefficient exist in the literature the PCM, n is the iteration number, d is a relaxation factor,
(Wakao and Kagei, 1982). In the present study, the intersti- which is set to 0.01 for the present case, and F1 is the
tial Nusselt number is given by the above expression inverse of latent heat function which takes the value of
obtained from (Galloway and Sage, 1970). The validity hm for a pure substance.
of the expression is confirmed experimentally for the range The coupled system of governing equations in their non-
of ReH  ðRc þ b Þ < 2500 (Beasley and Clark, 1984). The dimensional form is solved using an implicit method with
other characteristic nondimensional parameters that the diffusive and convective fluxes discretized using hybrid
appear in Eq. (3) are the Reynolds number, ReH, Prandtl differencing scheme, the convective–diffusive fluxes in the
number of the HTF, Prf and the inverse Stefan number HTF governing equation (Eq. (1)) discretized using the
of the PCM, w. Subscripts C and D correspond to the power-law scheme, and the diffusive fluxes in the PCM gov-
hot inlet HTF temperature during charging and the cold erning equation (Eq. (2)) discretized using central differenc-
inlet HTF temperature during discharging, respectively. ing scheme (Patankar, 1980). After a systematic grid
Note that hot HTF enters the tank at z* = 0 during charg- refinement study, the EPCM-TES tank was discretized into
ing and the exit temperature of the HTF corresponds to an 200 equally spaced intervals in the axial direction and the
axial position of z* = 1 and vice-versa during discharge. encapsulated PCM within the spherical capsules was dis-
Further, the boundary conditions during the charging cretized into 10 uniform zones in the radial direction.
and discharging process are specified as: The nondimensional time step chosen for the study was
Dt* = 1.25  109. The iterative residuals at each time step
@hf 
hf ðz ¼ 0Þ ¼ 1; ðz ¼ 1Þ ¼ 0; Charging are converged to the order of 109 to eliminate iterative
@z errors from the solution.
@hf
hf ðz ¼ 1Þ ¼ 0;  ðz ¼ 0Þ ¼ 0; Discharging
@z
2.2. Performance metrics
@h p ½hp ðRc ; z Þ  hf ðz Þ=Rc
k p  ðr ¼ Rc ; 0 < z < 1Þ ¼ hn o n oi
@r b 2Rc
  
k w ðRc þb Þ
þ  
NuðRc þb Þ Since a thermal storage system installed in a CSP plant
is subjected to repeated charging and discharging pro-
@hp 
ðr ¼ 0; 0 < z < 1Þ ¼ 0 cesses, the dynamic behavior of the EPCM-TES system is
@r
analyzed in the present study. Initially, the tank is kept in
ð4Þ a fully discharged state as established by Eq. (5) and charg-
The numerical simulations start with a charge process ing takes place as hot HTF flows into the tank from the top
assuming that the tank is completely discharged initially, (Fig. 1a) until the temperature of the HTF flowing out of
resulting in the following initial conditions for the HTF the storage system reaches a certain maximum charging
and the PCM phase: cut-off temperature, h0C . Now, during the subsequent dis-
charge process, cold HTF enters from the bottom of the
hf ¼ 0; hp ¼ 0 ð5Þ
tank (Fig. 1a) and discharge process continues until the exit
The complete set of governing equations is discretized temperature of the HTF flowing out from the top of the
using the finite volume approach and the melting/solidifica- tank reaches a minimum discharge cut-off temperature, h0D .
tion of the PCM is modeled using the fixed-grid enthalpy- Commonly in CSP plants, the HTF exiting the storage
porosity method as presented by Voller et al. (1987). In the tank during discharge (charge) is fed into the power block
fixed grid enthalpy-based formulation, the liquid fraction (receiver) for superheat steam generation (solar energy cap-
in each computational cell, in conjunction with the temper- ture). The material temperature limitations of the heat
ature, hp, predicted by Eq. (2) for the encapsulated PCM, is transfer fluid and the solar receiver dictate a maximum
updated based on the enthalpy balance at each iteration cut-off temperature, h0C , for the HTF exit temperature of
within a time step. The liquid fraction fL(hp) is given by the thermocline system during charging. If the exit temper-
the Heaviside step function at hp = hm, represented as ature exceeds this value, a high flow rate is needed through
774 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

the receiver in order to keep the temperature of HTF in the calculated in terms of nondimensional variables because
receiver within allowable limits. The high flow rate, in turn, the logarithmic term in the denominator always tends to
increases parasitic and thermal losses (SAM 2012.5.11, infinity, since the nondimensional variable of TD, viz. hD
2012). Similarly, the strong dependence of the power block as defined in Eq. (3) equals zero. Hence the charge (TC)
Rankine cycle efficiency on the incoming HTF temperature and discharge (TD) temperatures for the calculation of
(SAM 2012.5.11, 2012) requires the termination of dis- exergetic efficiency are considered to be 574 °C and
charge process when the HTF exit temperature reaches a 290 °C, respectively (SAM 2012.5.11, 2012), typical of mol-
certain minimum dimensionless discharge cut-off tempera- ten salt power tower operation.
ture, h0D . The combined charge and discharge process is The outputs from the model comprise the transient axial
referred to as one cycle. In the present study, h0C and h0D variation of temperature in the HTF, the transient radial
are selected to be 0.39 and 0.74, respectively, representative variation of temperature and the melt fraction of PCM
of restrictions imposed on molten salt power tower CSP within the capsule at any axial location. The nondimen-
plant operation (SAM 2012.5.11, 2012). The performance sional operating and design parameters investigated in
of the EPCM-TES system is analyzed following repeated the present study and the corresponding default values
charge and discharge processes until a periodic steady state are presented in Table 1. The other default values pertain
is established, when the variations in output and perfor- to b* = 5  108; Prf = 5; k w ¼ 1; k p ¼ 1 and e = 0.25.
mance metrics from one cycle to another are negligible (less The nondimensional height of each individual zone in cas-
than 0.002%). It is to be noted that repeated charging and caded configuration is defined as H 0p;q ¼ H p;q =H t , where p
discharging processes with large temperature oscillations denotes the number of cascades and subscript q denotes
throughout the filler bed and tank wall may result in a either T (top zone), M (middle zone) or B (bottom zone).
build-up of undesirable mechanical stress in the tank shell. The range for each of the operating and design parameters
Structural failure of the tank occurs if the stresses in the included in the present study is determined from a litera-
tank shell exceed the yield strength of the shell material, ture review of various PCM properties (Cárdenas and
a condition known as thermal ratcheting. However, León, 2013; Kenisarin, 2010) and molten salt HTF proper-
Flueckiger et al. (2013), through comprehensive numerical ties (SAM 2012.5.11, 2012). Numerical simulations were
analysis, has shown that thermal ratcheting is not an issue conducted for various combinations of the aforementioned
and is not explored in the present study. design and operating parameters to analyze its effect on the
The cyclic exergetic efficiency of the EPCM-TES system charge time, discharge time and exergetic efficiency of the
defined as the ratio of exergy recovered by the HTF during EPCM-TES system at its periodic steady state. It was
discharging to the exergy gained by the HTF during charg- shown in Nithyanandam and Pitchumani (2014c) that the
ing of each cycle (Stekli et al., 2013): periodic state is achieved beyond the fourth cycle and
R tD h  i accordingly, the values from the simulation during the fifth
T ðz¼0;tÞ
tC
T f ðz ¼ 0; tÞ  T D  T ref  ln f T D  dt cycle are presented as being the periodic state results. For
f ¼ Rt h  i ð7Þ instance, Table 2 shows the results obtained for cycles 1–
C TC
0
T C  T f ðz ¼ H t ; tÞ  T ref  ln T f ðz¼H t ;tÞ
 dt 15 for default 3-PCM cascaded EPCM-TES system config-
uration. The variations in the performance metrics between
where Tref is the reference or ambient temperature, which is the current and preceding cycle are presented in brackets. It
assumed to be 300 K (Stekli et al., 2013) in the present is observed that the charge and discharge times equili-
analysis. The term, Tf (z = 0, t) in the numerator refers brated to a periodic steady state value by the end of fifth
to the exit temperature of the HTF at the top of the cycle. Negligible variations in the exergetic efficiency are
EPCM-TES tank during discharging while the term Tf observed beyond the fifth cycle. However, the selection of
(z = Ht, t) in the denominator refers to the temperature 0.002% as the periodic state (cycle 5) is justifiable as the rel-
of the HTF at the bottom of the EPCM-TES system, which ative difference in the exergetic efficiency values between
is the outlet for charging. The exergetic efficiency cannot be cycle 5 (97.01502%) and cycle 15 (97.01381%) is only

Table 1
Dimensionless design and operating parameters investigated (N-C: non cascaded; 2-C: 2-PCM cascaded; 3-C: 3-PCM cascaded).
System configuration Parameters Design intervals Default value
N-C, 2-C, 3-C Capsule radius, Rc 0:00025  Rc  0:0025 0.001
N-C, 2-C, 3-C Capacitance ratio, k 0:5  k  1:5 1
N-C, 2-C, 3-C Inverse Stefan number, w 0:0  w  2:0 0.5
N-C, 2-C, 3-C Reynolds number, ReH 10; 000  ReH  25; 000 25,000
N-C PCM melt temperature, hm 0:00  hm  1:00 0.75
2-C Bottom zone PCM melt temperature, hm,B 0:00  hm;B  0:75 0.25
2-C Top zone PCM melt temperature, hm,T 0:25  hm;T  1:00 0.75
3-C Bottom zone PCM melt temperature, hm,B 0:00  hm;B  0:50 0.25
3-C Middle zone PCM melt temperature, hm,M 0:25  hm;M  0:75 0.50
3-C Top zone PCM melt temperature, hm,T 0:50  hm;T  1:00 0.75
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 775

Table 2
Cyclic results obtained for default 3-PCM cascaded configuration (the value in brackets shows the variation in results between the current and the
preceding cycle).
Cycle Charge time, tC (105) Discharge time, tD (105) Exergetic efficiency, f (%)
1 1.205 1.119 90.21449
2 1.128 (6.83%) 1.112 (0.63%) 96.80934 (6.81%)
3 1.126 (0.18%) 1.111 (0.09%) 96.99082 (0.19%)
4 1.126 (0.00%) 1.111 (0.00%) 97.01631 (0.03%)
5 1.126 (0.00%) 1.111 (0.00%) 97.01502 (1.33  103%)
6 1.126 (0.00%) 1.111 (0.00%) 97.01433 (7.11  104%)
7 1.126 (0.00%) 1.111 (0.00%) 97.01403 (3.09  104%)
8 1.126 (0.00%) 1.111 (0.00%) 97.01390 (1.34  104%)
9 1.126 (0.00%) 1.111 (0.00%) 97.01385 (5.15  105%)
10 1.126 (0.00%) 1.111 (0.00%) 97.01383 (2.06  105%)
11 1.126 (0.00%) 1.111 (0.00%) 97.01382 (1.03  105%)
12 1.126 (0.00%) 1.111 (0.00%) 97.01381 (1.03  105%)
13 1.126 (0.00%) 1.111 (0.00%) 97.01381 (0.00%)
14 1.126 (0.00%) 1.111 (0.00%) 97.01381 (0.00%)
15 1.126 (0.00%) 1.111 (0.00%) 97.01381 (0.00%)

1.2  105, which is well within the numerical tolerance of the system. It is observed that for a capsule radius of
limits. Rc ¼ 0:00025, characterized by small thermal conduction
The validation of the model is presented in resistance for heat transfer from the HTF to PCM, the
Nithyanandam and Pitchumani (2014c) and is not repeated influence of Reynolds number on the exergetic efficiency
here in the interest of brevity. The numerical model pre- is not pronounced. As the capsule radius increases, the con-
sented in this section forms the basis of the parametric duction resistance within the PCM capsules increases
studies discussed in Section 3 and is also used in a numer- resulting in a slower extraction of both latent and sensible
ical optimization of the EPCM-TES system as formulated energy from the interior of PCM capsules. In addition, the
in Section 4. In the following section, references to the shorter residence time of the incoming HTF with increase
design and operating parameters correspond to the nondi- in ReH leads to a faster decrease of the HTF exit tempera-
mensional parameters, unless stated otherwise. ture (hf at z* = 0) during the discharge process of larger
capsule radius, thus resulting in a decrease of exergetic effi-
3. Results and discussion ciency with increase in ReH and Rc .
The effect of capacitance ratio on the charge time, dis-
3.1. Parametric studies charge time and exergetic efficiency of the EPCM-TES sys-
tem are illustrated in Fig. 2c and d. For a given Reynolds
The effect of the various design and operating parame- number, the charge time and discharge time is found to
ters on the performance of the EPCM-TES system is stud- increase with increase in capacitance ratio (Fig. 2c) due
ied by varying one parameter at a time with the rest of the to the increase in volumetric heat capacity of the system.
parameters at the default value presented in Table 1. The same observations are found with exergetic efficiency,
Fig. 2a–d illustrates the effect of the nondimensional although the roles are reversed, in that it is found that exer-
parameters on the various performance metrics of the stor- getic efficiency is greater for system with smaller capaci-
age system namely, the charge time, discharge time and tance ratio (Fig. 2d). This is attributed to the fact that
exergetic efficiency for a non-cascaded EPCM-TES system. even though the charge and discharge times are longer
The variations in charge time are represented by lines with for a higher capacitance ratio, concomitant with increase
markers, while the lines without markers denote the varia- in storage capacity, the HTF exit temperature during dis-
tions in discharge time. From Fig. 2a, it is observed that charge (hf at z* = 0) for system with higher capacitance
charge time and discharge time decreases with increase in ratio will be at the PCM melt temperature (hm = 0.75) for
Reynolds number and capsule radius due to the shorter res- longer duration of the discharge process, which is less than
idence time of the incoming HTF with increase in ReH for the highest temperature (hf = 1), due to the higher thermal
effective heat exchange between the HTF and PCM, and inertia compared to the system with smaller capacitance
increase in thermal conduction resistance in the PCM with ratio. This implies a decrease in the quality of energy being
increase in Rc , respectively. The decrease in the discharge delivered to the power block, reflecting a decrease in the
time is more pronounced than the decrease in charge time exergetic efficiency. Nevertheless, the decrease in exergetic
with increase in capsule radius, due to the conduction dom- efficiency with increase in capacitance ratio is inconsequen-
inated solidification of PCM during discharge, in contrast tial as depicted in Fig. 2d. For the chosen default PCM
to the buoyancy-driven convection assisted melting of melt temperature of 0.75, the driving force for PCM solid-
PCM during charging. Fig. 2b illustrates the effect of cap- ification, DhD = hm  hD is higher compared to that for
sule radius and Reynolds number on the exergetic efficiency PCM melting, DhC = hC  hm. This leads to an efficient
776 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

Fig. 2. Effect of capsule radius on (a) charge and discharge time, (b) exergetic efficiency, and capacitance ratio on (c) charge and discharge time and (d)
exergetic efficiency of non-cascaded EPCM-TES system for different Reynolds number. The lines with markers in Fig. 3a and c represent charge time and
lines without markers indicate discharge time.

heat exchange between the HTF and PCM during the dis- temperature from hm = 0.74, which is the discharging cut-
charge process and hence, a gradual decrease in the exit off temperature (h0D ), to hm = 1. It is seen that for
temperature of the HTF. This results in a longer discharge h0C < hm < h0D , the charge time, as depicted in Fig. 3a, is
time compared to the charge time for the results presented faster due to insignificant system utilization. As explained
in Fig. 2a and c. in Nithyanandam and Pitchumani (2014c), the latent
Fig. 3a–c illustrates the variations in charge time, dis- energy utilization of the EPCM-TES system for PCM melt
charge time and exergetic efficiency with Reynolds number temperature in the range h0C < hm < h0D , is found to be the
for various PCM melt temperatures. For any given least with only a portion of the PCM adjoining the capsule
Reynolds number, the charge and discharge time is found wall involved in the phase change process and the EPCM-
to be significantly higher for PCM with hm either between TES system is in operation only for a short duration due to
0  hm  h0C or between h0D  hm  1 (Fig. 3a and b). The faster approach of cut-off temperatures.
charge time in Fig. 3a is found to increase from hm = 0.0 Comparing Fig. 3a and b, it is seen that the discharge
to hm = 0.39, which is the charging cut-off temperature time trends follow a mirror image of the trends observed
(h0C ). This can be explained by the fact that the driving force for the charge time about the ordinate axis. Hence, from
for PCM melting which manifests as the temperature differ- Fig. 3a and b, it can be observed that charge time is max-
ence between the incoming hot HTF and the PCM melt imum at hm = 0.38 while the discharge time is maximum at
temperature (DhC = hC  hm) decreases, leading to slower hm = 0.75 for any given Reynolds number. As PCM melt
charge rate. For hm = 0.0, discharge process does not lead temperature increases from 0.74 h0D to 1, the temperature
to PCM solidification and hence the EPCM-TES system difference between the incoming cold HTF and the PCM
works only in the sensible regime during the cyclic opera- melt temperature driving the PCM solidification increases,
tion. This is also confirmed by calculating the latent utiliza- thus leading to faster discharge time. The slight decrease in
tion of the system, defined as the amount of latent energy discharge time from hm = 0.39 to hm = 0 is attributed to the
extracted from the system to the total latent energy capac- decreasing drive force for the solidification of PCM during
ity of the system (Nithyanandam and Pitchumani, 2014c), discharging, and the discharging cut-off temperature is
which amounts to zero. A detailed explanation of the para- attained faster. The exergetic efficiency is higher for PCMs
metric effects on the utilization of the EPCM-TES system with melt temperatures of either 0 or 1 (Fig. 3c) because of
can be found in Nithyanandam and Pitchumani (2014c). operation in the sensible regime accompanied by almost
The aforementioned reason also applies for hm = 1.0, with equal charge and discharge time (Fig. 3a and b). For
the only difference being no driving force exists for PCM PCM with melt temperature between h0C and h0D , the exer-
melting during charge process, thereby leading to a slight getic efficiency is relatively high, as clearly seen from
decrease in the charge time with increase in PCM Fig. 3c, because of similar reason stated for PCM melt
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 777

Fig. 3. Effect of PCM melt temperature on (a) charge time, (b) discharge time and (c) exergetic efficiency, and inverse Stefan number on (d) charge time,
(e) discharge time and (f) exergetic efficiency of non-cascaded EPCM-TES system for different Reynolds number.

temperatures of 0 and 1, in that, sensible energy transfer due to increase in volumetric storage capacity of the
between the HTF and PCM takes place predominantly EPCM-TES system. With further increase in inverse Stefan
compared to latent energy transfer. Within the PCM melt number, w > 0.25, the latent thermal inertia of the system
temperature range, h0D  hm < 1, the HTF exiting the tank also increases, which results in a slower solidification of
during the discharge process remains at the PCM melt tem- the PCM in the capsules and a meager increase in the
perature for longer duration and as the PCM melt temper- charge and discharge times are observed in Fig. 3d and e,
ature increases from h0D to 1, the quality of energy respectively. Comparing Fig. 3d and e, it can be inferred
(exergetic efficiency) fed to the power block increases. that the effect of inverse Stefan number on the increase in
The exergetic efficiency increases as the PCM melt temper- charge time for w > 0.25 is less pronounced than on dis-
ature decreases from h0C to zero, because the EPCM-TES charge time. It is primarily due to the fact that the default
system approaches sensible energy operation mode. Even PCM melt temperature value chosen (hm = 0.75) leads to
though the exergetic efficiency of PCM’s with melt temper- slower PCM melting due to weaker driving force for melt-
atures in the feasible operating range, namely 0 < hm  h0C ing (DhC = hC  hm). This limits the heat exchange rate
and h0D  hm < 1, is lesser than the rest, it is to be noted between the HTF and PCM resulting in faster approach
that the exergetic efficiency is still greater than the 95% tar- of the charge cut-off temperature, notwithstanding the
get considered here. Also, from Fig. 3a–c, it can be increase in volumetric storage capacity with increase in
observed that charge time, discharge time and exergetic inverse Stefan number. For the same reason, the charge
efficiency, respectively, for any given PCM melt tempera- time in Fig. 3d is found to be faster than the discharge time
ture decreases with increase in Reynolds number due to in Fig. 3e, except for the inverse Stefan number of zero for
the shorter residence time of the incoming HTF. which the charge and discharge times are approximately
Fig. 3d–f portrays the various performance metrics of the same. It is to be noted that an inverse Stefan number
the EPCM-TES system as a function of ReH for different of zero conforms to a sensible thermocline storage system,
inverse Stefan number, w. With initial increase in inverse for which the PCM melt temperature has no influence,
Stefan number, the charge and discharge time increases, resulting in equal charge and discharge times. For any
778 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

given Reynolds number, it is found from Fig. 3f that the


exergetic efficiency decreases with increase in inverse Stefan
number. At higher inverse Stefan number, longer discharge
time at the PCM melt temperature, which is lesser than the
highest temperature (hC = 1.0) leads to larger exergy
destruction and hence reduced exergetic efficiency.
From previous plots, it is observed that the exergetic
efficiency for almost all of the parametric values are
greater than 95% which was also verified for other para-
metric combinations and hence is not illustrated further
on. To this extent, Fig. 4a and b displays the effects of
top zone PCM melt temperature and bottom zone PCM
melt temperature, respectively, on the charge (with mark-
ers) and discharge (without markers) times of a 2-PCM
cascaded EPCM-TES system configuration for various
Reynolds number. From the variations of PCM melt tem-
perature at the top zone (Fig. 4a), it is observed that the
charge and discharge times are higher for PCMs with melt
temperature greater than 0.74, which is the discharge cut-
off temperature. However, it is seen that with subsequent
increase beyond hm,T = 0.75 in PCM melt temperature at
the top zone — for which the discharge time at
ReH = 10,000 is tD ¼ 2:69  105 compared to a discharge
time of tD ¼ 2:45  105 for hm,T = 0.74 — the discharge
time decreases due to impediment in the melting of
PCM as DhC decreases. The same phenomena of the
decrease in discharge time can be observed in Fig. 4b,
although for temperatures below 0.39, due to the decrease
in driving force for solidification as DhD (DhD = hm  hD)
decreases. It is also observed from Fig. 4b, that the dis-
charge and charge times decrease and remains constant
for PCM melt temperatures greater than hm,B = 0.39,
which is the charging cut-off temperature, as observed
for non-cascaded configuration in Fig. 3b. However, the
magnitude of decrease in discharge and charge times as
the PCM melt temperature in the bottom zone decreases
from charging cut-off temperature to higher melt tempera-
Fig. 4. Effects of (a) top zone PCM melt temperature, (b) bottom zone
tures in the range h0C < hm < h0D , is not as high as observed PCM melt temperature, and (c) top zone height of a 2-PCM cascaded
for a non-cascaded EPCM-TES system configuration EPCM-TES system on charge and discharge time for different Reynolds
(Fig. 3b) because of the fact that the PCM melt tempera- number. The lines with markers represent charge time and lines without
ture at the top zone (hm,T = 0.75) extends the system oper- markers indicate discharge time.
ation during discharge. As depicted in Fig. 4b, the charge
time is greater than the discharge time only for bottom
zone PCM melt temperature in the range 0.22–0.40. For temperature of the HTF (hf at z* = 1) and longer charge
hm,B < 0.22, the thermal drive force for the melting of time to attain hC .
PCM (DhC = hC  hm) increases, leading to a steeper Fig. 4c shows the effect of varying the height of the top
increase in the HTF exit temperature during charging (hf zone in a 2-PCM cascaded EPCM-TES system, H 02;T on
at z* = 1) and faster attainment of the charging cut-off the overall performance of the EPCM-TES system. At
temperature (h0C ). The charge time in Fig. 4b is observed H 02;T ¼ 1, the EPCM-TES system is packed with PCM
to be lesser than the discharge time for hm,B > 0.39, as undergoing solid–liquid phase change at hm = 0.75 leading
the PCM melt temperature close to the exit port for charg- to a longer (shorter) discharge (charge) time compared to
ing (z* = 1) falls outside the charging cut-off temperature the shorter (longer) discharge (charge) time obtained for
constraint. The charge time in Fig. 4a is found to be H 02;T ¼ 0 , when the EPCM-TES system is packed with
higher than the discharge time for all the top zone PCM PCM having a melt temperature of hm = 0.25, for any given
melt temperature, because the default PCM melt tempera- Reynolds number. This is attributed to the higher solidifica-
ture in the bottom zone at hm,B = 0.25, which is within the tion (melting) driving force for H 02;T ¼ 1 ðH 02;T ¼ 0Þ, which
range 0  hm  h0C , results in a gradual increase in the exit results in an efficient heat exchange between the HTF and
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 779

Fig. 5. Effect of (a) top, (b) middle and (c) bottom PCM melt temperature, and (d) top, (e) middle and (f) bottom zone height of 3-PCM cascaded EPCM-
TES system on charge and discharge time for different Reynolds number. The lines with markers represent charge time and lines without markers indicate
discharge time.

PCM, thus leading to a longer discharge (charge) time. temperatures above the default value, h0D < hm;T  1, the
Fig. 4c shows that as the zone height increases, the charge charge and discharge times gradually decrease due to the
and discharge times increase until H 02;T ¼ 0:5 (equally decrease in driving force for melting as Dhc decreases.
divided zones) after which it decreases as the system config- Fig. 5b shows that changing the PCM melt temperature
uration tends to a non-cascaded EPCM-TES system config- at the middle zone does not significantly affect the charge
uration at H 02;T ¼ 1. For instance at ReH = 10,000, the and discharge times, although the maximum charge and
discharge time of ECPM-TES system is tD ¼ 2:69  105 discharge times are achieved for hm,M = 0.5. This is attrib-
with H 02;T ¼ 0:5 and the total utilization of 2-PCM cascaded uted to the fact that for equally divided 3-zone PCM cas-
configuration is 82.95% with 98.39% of the energy coming caded configuration; PCM’s with equally spaced intervals
from the latent enthalpy of fusion of the PCM, while the in melt temperatures lead to a more uniform temperature
discharge time of non-cascaded configuration at difference between the hot and cold media. It is observed
H 02;T ¼ 1:0 is tD ¼ 2:29  105 with a total utilization of from Fig. 5b that the discharge time is higher than the
68.0% and a latent utilization of 38.84% only. charge time for PCM melt temperatures, hm,M > 0.5, as
Fig. 5a–c portrays the combined effects of Reynolds the thermal drive force for the melting of PCM (Dhc)
number and PCM melt temperature at the top, middle decreases, leading to faster attainment of charge cut-off
and bottom zone respectively, while Fig. 5d–f portrays temperature. Fig. 5c exhibiting the variations in bottom
the combined effects of Reynolds number and height of zone PCM melt temperature shows that a step decrease
the top, middle and bottom zone, respectively on the in charge and discharge time is observed when the melt
charge and discharge times for a 3-PCM cascaded temperature falls outside the range, 0 < hm;B  h0C . Within
EPCM-TES system configuration. The variations in charge the temperature range of 0 < hm;B  h0C , the charge and dis-
time are represented by lines with markers, while the lines charge time decreases gradually with decrease in melt tem-
without markers denote the variations in discharge time. perature from h0C to 0 due to the decrease in driving force
In Fig. 5a, for the top zone default PCM melt temperature for solidification as DhD decreases. As observed for a 2-
of hm,T = 0.75, the charge and discharge times are found to PCM cascaded configuration in Fig. 4b, the charge time
be the highest. For melt temperature just below the default, is higher than the discharge time only for bottom-zone
a step decrease in charge and discharge times are observed PCM melt temperatures between 0.22 and 0.39 ðh0C Þ.
because the melt temperature falls outside the range, Since the height of each of the zones cannot be varied
h0D  hm;T  1, leading to a faster attainment of the independently, in the present study, the heights of two
discharge and charge cut-off temperatures. For melt zones are varied simultaneously to analyze its influence
780 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

on the performance metrics. For instance in Fig. 5d, the for ReH = 10,000) compared to a 2-PCM cascaded config-
nondimensional heights, as defined in Section 2.2, of the uration with equally divided zone heights (tD ¼ 2:69  105
top zone and middle zone are varied simultaneously while for ReH = 10,000) while the discharge time of a non-cas-
the height of the bottom zone is fixed at the default value of caded configuration at the default parametric value is
0.335. Fig. 5d shows that as the top zone height increases tD ¼ 2:69  105 for ReH = 10,000. It is also observed from
from 0 to 0.2, the charge and discharge times gradually Fig. 5a–f that the charge (discharge) time decreases with
increase while the increase is higher when the top zone increase in Reynolds number due to insufficient residence
height increases from 0.2 to 0.335. With a top zone height time of the incoming hot (cold) HTF within the tank,
of 0, the EPCM-TES system becomes a 2-PCM cascaded resulting in poor heat exchange between the HTF and
configuration with the nondimensional middle zone height EPCM, and faster decay of the HTF exit temperatures.
of 0.667. Since the default PCM melt temperature of the
middle zone is 0.5, which falls outside the cut-off tempera- 3.2. Design window
ture ranges, 0  hm < h0C andh0C  hm < 1, the latent energy
utilization of the PCM is restricted, and hence the charge The effects of the nondimensional parameters on the sys-
and discharge times are the lowest for any given Reynolds tem performance, presented in Figs. 2–5, may be used to
number. As the height of the top zone filled with PCM of identify system design configurations that satisfy the tech-
melt temperature 0.75 increases — within the limits of noeconomic targets of (1) discharge time greater than 6 h,
h0D  hm < 1 — the discharge time increases and when the (2) charge time less than 6 h, (3) exergetic efficiency greater
top zone height increases from 0.2 to 0.335 the discharge than 95% and (4) storage cost less than $15/kW ht. To this
time increases drastically, for instance from 2.36  105 extent, properties of molten salt (60% NaNO3 + 40%
to 2.89  105 at ReH = 10,000, reflecting a uniformity in KNO3) (Nithyanandam and Pitchumani, 2014c; SAM
the temperature difference between the HTF and PCM 2012.5.11, 2012) and two different tank heights, 15 m and
leading to a better system utilization. However, the dis- 20 m, are considered for which the nondimensional time
charge time peaks at H 03;T ¼ 0:467ðH 03;B ¼ 0:200Þ, for pertaining to 6 h are 1.732  105 and 0.9932  105,
instance tD ¼ 2:92  105 at ReH = 10,000, although the respectively. In this study, design windows identifying
increase from H 03;T ¼ 0:335 is very subtle. The charge time regions that satisfy the technoeconomic targets (Stekli
followed the same trend as discharge time except that it et al., 2013; SunShot, 2012) are presented in the reminder
peaked for a top zone height of 0.335 of the section for a power tower CSP plant operating at
ðtC ¼ 2:91  105 at ReH ¼ 10; 000Þ while it was the lowest a rated capacity of 115 MWe within the temperature ranges
for H 03;T ¼ 0 (Fig. 5d). of 290 °C (TD) and 574 °C (TC). For a gross cycle conver-
Similarly from Fig. 5e, it is found that the charge and sion efficiency of 42.5%, 270.588 MWt of thermal power
discharge times are highest for middle zone height of input is necessary to deliver power at the rated capacity,
0.335. It is to be noted that the observation is different from for which the mass flow rate of the HTF required is
that portrayed in Fig. 5d, in that the highest discharge time 635.44 kg/s. Hence, for a fixed HTF flow rate and tank
was observed for a top zone height of 0.467 with a corre- height, changes in Reynolds number, ReH ¼ mH _ t =lf pR2t ,
sponding mid zone height of 0.2. While varying middle correspond to variations in the tank radius. Also, with
zone height to 0.2 in Fig. 5e, the top zone height remains the properties of the HTF fixed, the change in capacitance
at the default 0.335 and the bottom zone height is increased ratio and inverse Stefan number corresponds to variation
to 0.467, which is different from the analysis in Fig. 5d. in density and latent heat of fusion of the PCM,
From Fig. 5f, it is observed that until H 03;B ¼ 0:335, the respectively.
charge and discharge times increase with simultaneous It is understood from the SunShot vision study
increase and decrease in bottom and top zone heights, (SunShot, 2012) that to achieve a levelized cost of electric-
respectively, due to the availability of PCM at melt temper- ity of 6 cents/kW h, equivalent to the electricity cost asso-
atures between 0 < hm  hC and h0D  hm < 1. The dis- ciated with fossil fueled power plants, improvements to
charge time (Fig. 5f) is highest when the top and bottom all subsystems within a CSP plant is required which also
zone heights are equally distributed (H 03;B ¼ 0:335) imply- includes reducing the thermal storage cost from $27/kW ht
ing the significance of having PCMs with melt temperature (2-tank sensible storage) to $15/kW ht. The total cost of
at both the low and high temperature range below the EPCM-TES storage system in units of dollar (US$) can
charging cut-off temperature and above the discharging be expressed as the summation of storage material cost
cut-off temperature, respectively. The discharge time (HTF and PCM), container cost, encapsulation cost and
decreases with further increase in H 03;B , due to the decrease overhead cost. The ratio of the total cost to the total stor-
in the top zone height filled with PCM melt temperature of age capacity of the EPCM-TES system provides the stor-
0.75, thus resulting in quicker attainment of discharge cut- age cost in units of $/kW ht. The overhead cost,
off temperature. Hence the charge time also decreases with accounting for the miscellaneous costs such as electrical,
increase in bottom zone height for H 03;B ¼ 0:335. The dis- instrumental, piping, valves and fitting costs is assumed
charge time of a 3-PCM cascaded configuration with to be 10% of the storage material, container and encapsu-
equally divided zone heights is higher (tD ¼ 2:89  105 lation cost. The storage material cost can be calculated as:
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 781
h i
qp ð1  eÞC p þ qf eC f pR2t H t where the PCM cost per unit
mass (C p ) is assumed to be $0.75/kg, which is the average
cost of the available PCMs as tabulated in Kenisarin
(2010) and the HTF cost per unit mass (C f ) used in the
molten salt power tower CSP plant is taken to be $1/kg
(Glatzmaier, 2011). The EPCM-TES tank cost is calculated
as explained in Glatzmaier (2011) and Kelly and Kearney
(2006), which encompasses the material cost of the stainless
steel (321 SS), foundation cost and the insulation cost:
2
qSS H t ½pðRt þ wt Þ  pR2t   C SS þ pR2t C 00F þ 2pRt  H t C 00I , where
the subscript SS refers to stainless steel, the density of
which is taken as qSS = 7900 kg/m3 and wt refers to the
thickness of the tank, which is assumed to be 0.038 m.
The cost per kg of stainless steel (C SS ) is obtained to be
$5.43/kg from International Ltd. (2012), the calcium sili-
cate insulation cost (C 00I ) is assumed to be $235/m2 (Kelly
and Kearney, 2006), and the foundation cost (C 00F ) is taken
as $1,210/m2 (Kelly and Kearney, 2006; Herrmann et al.,
2004). The cost values reported here are consistent with
the values of two-tank molten salt thermal storage system
reported in Glatzmaier (2011) and Kelly and Kearney
(2006). The cost of encapsulation is usually measured in
terms of cost/kg of capsule and due to lack of information
about the effect of capsule radius on the encapsulation cost,
 Rc 0:3 
it is expressed as C encap ¼ 0:005  C encap;0 where the base
cost for encapsulating one kg of PCM in a capsule of radius
0.005 m (C encap;0 ) is assumed to be $0.75/kg (Mathur, 2012).
The encapsulation cost increases with increase in capsule
radius due to higher heat treatment and processing cost.
Also, the stronger buoyancy driven natural convection
current that will be prevalent inside the molten PCM of Fig. 6. Variation in storage cost with tank radius for different (a) capsule
radius, (b) PCM latent heat of fusion and (c) PCM density for 15 m tall
larger capsules during the melting process requires a
non-cascaded EPCM-TES system.
thicker shell material. The total storage capacity of the
EPCM-TES system can be calculated from: pR2t H t


qp cp ð1  eÞðT C  T D Þ þ qp ð1  eÞhsl þ qf cf eðT C  T D Þ . As the system increases with increase in latent energy capacity
seen from the derived expressions, the storage cost depends of the PCM. It is observed from Fig. 6c that the storage
on the density of PCM, latent heat of fusion of PCM, cost decreases with increase in density of the PCM as the
capsule radius and tank radius, while the PCM melt tem- storage capacity of the tank increases for a given tank
perature does not influence the storage cost. dimension. It is to be noted that the storage cost obtained
The plots in Fig. 6 present the storage cost of EPCM- is less for a tank height of 20 m (not illustrated here),
TES system as a function of tank radius for various para- because of the larger storage capacity for a given HTF
metric values considered. It is to be noted that the density mass flow rate compared to a tank height of 15 m, notwith-
and latent heat of fusion of the PCM is varied to reflect the standing the increase in material cost.
changes in capacitance ratio and inverse Stefan number, From the plots in Fig. 6, design criteria on the EPCM-
respectively. Fig. 6a–c shows the variation in storage cost TES system parameters may be established so as to achieve
as a function of capsule radius, latent heat of fusion and a storage cost of $15/kW ht as called for in the technoeco-
density of PCM respectively for a tank height of 15 m. In nomic target (SunShot, 2012). An example of this design
Fig. 6a, it is observed that storage cost increases with criterion in illustrated in Fig. 6a, which shows that for a
increase in capsule radius due to increase in encapsulation tank height of 15 m, there exists a lower bound on the tank
cost. As tank radius increases, the storage cost decreases radius, RLt , for a given capsule radius, so as to limit the stor-
because of the increase in storage capacity, notwithstand- age cost to within $15/kW ht. The lower bound on the tank
ing the increase in material, insulation and foundation cost. radius increases with increase in capsule radius, as indi-
In the analysis of the effect of PCM’s latent heat of fusion cated by the filled markers in Fig. 6a.
on the storage cost (Fig. 6b), it is found that the storage Based on the variations in nondimensional charge and
cost decreases with increase in latent heat of fusion, which discharge time with different nondimensional design and
is consistent with the reasoning that the storage capacity of operating parameters in Figs. 2–5, the system design
782 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

Fig. 7. Effect of tank radius on (a) charge and (b) discharge time of 15 m tall EPCM-TES system, and (c) charge and (d) discharge time of 20 m tall
EPCM-TES system.

configurations that satisfy the design targets of discharge the constraint on the minimum discharge time and the max-
time greater than 6 h and charge time less than 6 h can imum storage cost forms the lower limit on the tank radius
be identified. To illustrate this, the results obtained in for different values of the design parameters as indicated by
Fig. 2a is presented in a dimensional form for tank heights the solid lines and chain-dashed lines, respectively, in Fig. 8.
of 15 m and 20 m in Fig. 7a–d. As mentioned earlier, the Similarly, the dotted line is based on the maximum charge
properties of molten salt (60% NaNO3 + 40% KNO3) are time constraint, which forms the upper bound on the tank
considered to convert the nondimensional charge and dis- radius in Fig. 8. The constraint on the minimum exergetic
charge times defined in Eq. (3) into realistic values. A efficiency does not influence the design window since the cri-
power plant of rated capacity 115 MWe is considered so teria is met for all the design parameters as presented in
that the effect of Reynolds number can be captured as var- Figs. 2–4. The shaded area identifies regions of operation,
iation in tank radius for a given mass flow rate in a unified which satisfy the technical design requirements of a thermal
manner, as discussed in the beginning of Section 3.2. energy storage system. The design windows illustrated in
Fig. 7a reveals an upper bound on the tank radius, RUt , Fig. 8 are bounded by the larger of the two lower limits
for a given capsule radius so as to limit the charge time on the tank radius based on the discharge time and storage
to within 6 h while Fig. 7b shows a lower bound on the cost constraint and the upper limit on the tank radius based
tank radius, RLt , in order to achieve a discharge time greater on the charge time constraint.
than 6 h. The upper bound ðRUt Þ and lower bound ðRLt Þ on With increase in capsule radius, the design window
the tank radius is found to increase with increase in capsule shrinks to a narrow range as depicted in Fig. 8a, as the
radius as presented in Fig. 7a and b. With increase in the lower limit on the tank radius based on the discharge time
height of the tank, as illustrated for tank height of 20 m increases. This is attributed to the increase in thermal con-
in Fig. 7c and d, the upper and lower bounds on the tank duction resistance for solidification with increase in capsule
radius corresponding to the constraints on charge and dis- radius, thus requiring a larger tank radius with longer res-
charge time, respectively, are observed to decrease. Since idence time of the incoming HTF. The effect of capsule
the exergetic efficiency is greater than the target of 95%, radius on the increase in upper bound (charge time con-
for the different design and operation parameters consid- straint) is not pronounced due to the fact that melting is
ered in this study (Fig. 2b), it does not pose any constraint assisted by buoyancy driven convection currents effecting
to the selection of EPCM-TES tank radius. efficient storage of energy. Since the storage cost increases
Using the parametric studies shown in Figs. 2–6, with increase in capsule radius and tank radius (Fig. 6a),
Fig. 8a–f presents the design windows on the tank radius, the lower bound on the tank radius based on storage cost
Rt, which identify the regions where tD  6 h;tC  constraint (chain-dashed line) increases with increase in
6 h; f  95% and a  $15=kW ht can be achieved as a func- capsule radius and restricts the operation of EPCM-TES
tion of the different design parameters of EPCM-TES system filled with large capsule radius (Fig. 8a). Fig. 8b
system. As illustrated in Figs. 6 and 7, it is observed that provides a comparison of the effect of capsule radius on
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 783

Fig. 8. Operating windows as functions of (a) capsule radius for Ht = 15 m, (b) capsule radius for Ht = 20 m, (c) PCM latent heat of fusion for Ht = 15 m
and (d) PCM density for non-cascaded EPCM-TES system, and (e) bottom zone PCM melt temperature for 2-PCM cascaded, (f) bottom zone height for
3-PCM cascaded 15 m tall EPCM-TES system.

the operation window for a tank height of 20 m with charge time is obtained for lower tank and capsule radii.
respect to a tank height of 15 m portrayed in Fig. 8a. With The maximum discharge time, as marked by filled circles,
increase in tank height, favorable operation of the EPCM- is higher for tank height of 15 m (tD = 7.01 h) compared
TES system is obtained at smaller tank radius. Analyzing to a tank radius of 20 m (tD = 6.99 h). Since the charge
Fig. 8a and b, it is seen that for an EPCM-TES system and discharge time decreases with increase (decrease) in
of height 15 m (Fig. 8a), the storage cost does not restrict Reynolds number (tank radius) as depicted in Fig. 2a
the operation of EPCM-TES system until a capsule radius (Fig. 7a and b), the maximum discharge time obtained
of Rc = 31 mm, compared to EPCM-TES height of 20 m, for Ht = 20 m is lesser than that obtained for Ht = 15 m.
where the limit on the storage cost does not favor Likewise the minimum charge time (square markers) and
operation of EPCM-TES system with capsule radius, storage cost (circle markers) for Ht = 15 m, are obtained
Rc > 28 mm (Fig. 8b). to be 5.14 h and $11.64/kW ht, respectively, which are less
Within the feasible design window in Fig. 8a and b, the than the corresponding values obtained for Ht = 20 m
design point that gives the maximum discharge time (A) is namely, tD = 5.15 h and a = 11.98 kW ht.
marked by filled circle, which coincides with the minimum Fig. 8c illustrating the effect of latent heat of fusion of
storage cost design point (B). The minimum charge time PCM on the design window shows that upper bound on
design point (C) is marked by filled square. In Fig. 8a Rt dictated by the charge time constraint decreases slightly
and b, the maximum discharge time is obtained for higher while the lower bound on Rt decided by the discharge time
tank radius and lower capsule radius while the minimum constraint decreases initially and then settles to a constant
784 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

value. With increase in latent heat of fusion and thermal thus leading to infeasible operation of EPCM-TES system
inertia of the system, a smaller tank radius with faster between PCM melt temperatures of 0.22 and 0.40 in the
HTF flow velocity and correspondingly higher heat trans- bottom zone. At higher PCM melt temperatures, Tm(-
fer coefficient is required for efficient heat exchange hm) > 403.6 C (0.40), for a given Reynolds number, the dis-
between the HTF and PCM. While the increase in thermal charge time is higher compared to the charge time because
inertia affects the conduction dominated solidification rate, none of the zones contain PCM within the temperature
its influence on the natural convection assisted melting rate range of TD and T 0C to prolong the charging operation of
is not compelling, which results in an inconsequential effect EPCM-TES system. Hence, the tank radius satisfying the
on the upper bound set by the charge time constraint. The discharge time constraint is smaller compared to the tank
range of tank radius that provides feasible operation of the radius that satisfies the charge time constraint and a feasi-
storage system is broader for higher hsl. It should be noted ble design window is observed in Fig. 8e. The effect of top
that for an hsl < 162.8 kJ/kg, the constraint on the storage zone height of a 2-PCM cascaded configuration on the
cost restricts the storage system operation (Fig. 8c) due to operation window (not illustrated here) showed that the
the decrease in storage capacity of the EPCM-TES system. feasible design window is observed for larger top zone
From Fig. 8d, it is observed that as density of PCM heights, since a higher percentage of the storage system is
increases the upper bound and lower bound on the tank packed with PCM of higher melt temperature. A feasible
radius decreases. As PCM density increases, the storage operation window as a function of top zone PCM melt
capacity of the EPCM-TES system also increases, which temperature is not obtained because the charge time is
leads to an increase in charge and discharge time for a found to be greater than the discharge time for the entire
given tank radius. Due to the inverse (direct) dependence range of PCM melt temperature, as portrayed in Fig. 4a.
of charge and discharge time with Reynolds number (tank Fig. 8f portrays the effect of simultaneously varying the
radius), as depicted in Fig. 2c, the EPCM-TES system filled bottom and top zone heights on the design window. Feasi-
with capsules containing low PCM density has feasible ble design window is obtained for small bottom zone
design windows for larger tank radius and vice-versa. height, (H 03;B ) and large top zone height, H 03;T , when the
Due to the increase in storage cost with increase in PCM bulk of the storage system is filled with PCM of higher melt
density as revealed in Fig. 6c, it is observed that the design temperature. The design window shown in Fig. 8f is
window of EPCM-TES system narrows for density of obtained for bottom zone height corresponding to which,
PCM less than 1135 kg/m3. The maximum discharge time the discharge time is greater than the charge time as illus-
and minimum storage cost are obtained for higher PCM trated in Fig. 5f. Similar operation window as a function
density (circle marker) while the minimum charge time is of other zone heights and PCM melt temperatures can be
obtained for lower PCM density (square marker). obtained, but is not illustrated here in the interest of
The effects of other parameters of non-cascaded EPCM- brevity.
TES system on the design window, though not illustrated From the design windows on tank radius obtained for
here, may be summarized as follows: A very narrow design the various design parameters, the design configurations
window is observed for PCM with melt temperatures in the that maximizes the discharge time, and minimizes the
range T 0D  T m < T C and PCM melt temperature below T 0D charge time, storage capital cost are obtained. Table 3 lists
does not have a feasible design window. This is attributed the combinations of parameters culled from the parametric
to the fact that PCM with melt temperature below T 0D studies that: (1) maximizes discharge time, tD, (2) mini-
requires a higher charge time to attain the corresponding mizes charge time, tC, and (3) minimizes storage cost, a
discharge time (Fig. 3a and b). within the design window of EPCM-TES operation for
Fig. 8e illustrates the effect of varying PCM melt tem- non-cascaded, 2-PCM cascaded and 3-PCM cascaded con-
perature at the bottom zone of a 2-PCM cascaded figurations. The maximum or minimum values of the
EPCM-TES system on the operation window. For PCM objective function are listed in bold face and the corre-
with melt temperature less than 352.5 C (hm = 0.22) and sponding design configuration is highlighted in Table 3.
greater than 403.6 °C (hm = 0.4), feasible design window It is seen from Table 3a that of the cases studied, the
is observed as portrayed in Fig. 8e. As PCM melt temper- best combination of parameters that maximizes tD for
ature in the bottom zone increases from TD, the tank radius non-cascaded EPCM-TES system yields discharge time of
which satisfies the charge (dotted line) and discharge (solid 7.42 h for Ht = 15 m (Table 3a.1). It was observed that
line) time requirements also decreases reflecting the fact the tank radius, Rt corresponding to the best configuration
that higher charge and discharge time is achieved with decreases with increase in EPCM-TES height, which leads
increase in PCM melt temperature (Fig. 4b). Even though to a lower discharge time for Ht = 20 m and is not listed
the discharge and charge time follows the same trend with here. Also, it is observed that the maximum discharge time
variations in PCM melt temperature as portrayed in corresponds to case with higher latent heat of fusion with
Fig. 4b, beyond Tm = 352.5C (hm = 0.22), smaller Rey- the rest of the parameters at the default values, which is
nolds number or larger tank radius is required to satisfy represented by the filled circles (design point A) in
the constraint on the discharge time compared to the tank Fig. 8c. For a 2-PCM cascaded configuration, the maxi-
radius required to satisfy the constraint on the charge time, mum discharge time of 6.91 h is obtained for EPCM-TES
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 785

Table 3
Optimum design obtained from operating windows. The maximum and minimum values of the objective function are listed in bold face. The best system
design configuration for each of the objective function is italicized.
Objective hsl (kJ/kg) Tm (C) qp (kg/m3) Rt (m) Rc (mm) tC (h) tD (h) a ($/kW ht) f (%)
a. Maximum discharge time, tD (h)
1. 1-PCM (non-cascaded) Ht = 15.0 m – 852.0 503 1975.5 11.252 15 6.000 7.422 7.545 96.491
2. 2-PCM (cascaded) H2,T = 15.0 m Top 213.0 503 1975.5 11.276 15 6.001 6.909 13.719 97.407
H2,B = 0.0 m Bottom 213.0 361 1975.5
3. 3-PCM (cascaded) H3,T = 5.0 m Top 852.0 503 1975.5 8.902 15 6.001 6.858 11.365 95.537
H3,M = 5.0 m Mid 213.0 432 1975.5
H3,B = 5.0 m Bottom 213.0 361 1975.5
b. Minimum charge time, tC (h)
1. 1-PCM (non-cascaded) Ht = 15.0 m – 852.0 503 1975.5 10.040 15 4.859 6.001 7.725 96.464
2. 2-PCM (cascaded) H2,T = 15.0 m Top 213.0 503 1975.5 10.722 15 5.219 6.001 13.860 97.392
H2,B = 0.0 m Bottom 213.0 361 1975.5
3. 3-PCM (cascaded) H3,T = 5.0 m Top 852.0 503 1975.5 8.425 15 5.257 6.001 11.545 95.522
H3,M = 5.0 m Mid 213.0 432 1975.5
H3,B = 5.0 m Bottom 213.0 361 1975.5
c. Minimum storage cost, a ($/kW ht)
1. 1-PCM (non-cascaded) Ht = 15.0 m – 852.0 503 1975.5 11.252 15 6.001 7.422 7.545 96.491
2. 2-PCM (cascaded) H2,T = 7.5 m Top 213.0 503 1975.5 9.380 15 6.001 6.820 10.127 95.448
H2,B = 7.5 m Bottom 213.0 361 1975.5
3. 3-PCM (cascaded) H3,T = 5.0 m Top 852.0 503 1975.5 8.902 15 6.001 6.858 11.365 95.537
H3,M = 5.0 m Mid 213.0 432 1975.5
H3,B = 5.0 m Bottom 213.0 361 1975.5

system with top zone height of 15 m. Another important 4. Numerical optimization


observation is that the charge time of the EPCM-TES sys-
tem corresponding to maximum tD corresponds to 6 h for The best configurations obtained from the parametric
all the EPCM-TES system configurations namely non-cas- studies (Figs. 2–7) and the design windows (Fig. 8)
caded, 2-PCM and 3-PCM cascaded EPCM-TES system, reported in Table 3 are obtained from varying one param-
which sets the upper bound on the tank radius for various eter at a time with the rest of the parameters at the design
design parameters as presented in Fig. 8. Table 3b presents values. It is imperative that an optimization procedure
the configurations of the storage system, which provided varying all the design parameters simultaneously is con-
the minimum charge time and the optimum system config- ducted to find the optimum combination of design and
uration are the same as those obtained for maximum operating parameters while also achieving the technoeco-
discharge time presented in Table 3a. However, the mini- nomic targets.
mum charge time, for a given tank height is obtained for The goal of the optimization is to determine the design
smaller tank radius values corresponding to which the parameters of the EPCM-TES system so as to minimize
discharge time is 6 h. Among all the configurations, the the storage cost. The optimization is subject to operation
minimum charge time of tC = 4.86 h is obtained for constraints on the charge time, discharge time and exerget-
non-cascaded configuration (Table 3b.1). The design point ic efficiency defined previously. The decision variables
corresponding to the minimum charge time of the non- whose optimum values are sought to be determined ranges
cascaded configuration is represented by the filled squares from 5 variables for a non-cascaded system configuration
(design point C) in Fig. 8c. The minimum storage cost to 14 variables for a 3-PCM cascaded system configuration.
for the different configurations of storage system reported For a 3-PCM cascaded storage system the decision vari-
in Table 3c corresponds to system with highest latent heat ables are the density, latent heat of fusion, melt tempera-
of fusion and hence higher energy storage capacity. The ture of the PCM in each of the three zones, the height
minimum storage cost also corresponds to the upper bound ratio of each of the zones, the tank radius and the capsule
of the design window based on the charge time constraint radius. In this optimization study as explained in Sec-
(Fig. 8) with larger tank radius and higher volumetric stor- tion 3.2, a 115 MWe CSP plant is considered for which
age capacity. Among the parametric studies considered in the desired flow rate of the HTF (635.44 kg/s) is fixed
this study, the non-cascaded configuration has the mini- and the optimum Reynolds number pertains to the desired
mum storage cost, a = 7.55 kW ht in Table 3c.1, since the tank radius which minimizes the storage cost and hence is
energy storage capacity is at least 50% higher compared also considered to be a decision variable. The upper and
to a 2-PCM cascaded configuration (Table 3c.2) and at lower bounds of the decision variables considered for the
least 66.7% higher compared to a 3-PCM cascaded config- study are listed in Table 4 and the optimization problem
uration (Table 3c.3). can be written mathematically as:
786 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

Table 4 an improvement to the objective function. The Nelder–


Design parameters considered in the study. Mead simplex method combined with simulated annealing
Parameters Symbol Units Design intervals is designed to solve non-linear unconstrained problems.
Tank radius Rt m 7:5  Rt  12:5 Since the optimization problem in hand has several con-
Capsule radius Rc mm 3:75  Rc  37:5 straints, gk, those constraints are incorporated using a pen-
PCM density qp kg/m3 987:75  qp  3951:00 alty method (Bertsekas, 1999) into an augmented objective
PCM melt temperature Tm C 290  T m  574
PCM latent heat of fusion hsl kJ/kg 106:5  hsl  852:0
function to be minimized, such that the optimization prob-
Zone height fraction H0 – 0  H0  1 lem now becomes unconstrained:
X
Minimize
0
a þ pk maxðgk ; 0Þ ð9Þ
qp ;T m ;hsl ;H ;Rc ;Rt
k

Minimize a ð8aÞ where pk are the penalty factors. The main drawback of
qp ;T m ;hsl ;H 0 ;Rc ;Rt
penalty functions is that the penalty factors require a care-
subject to ful fine tuning, which determines the severity of penaliza-
tion to be applied in order to obtain the optimum
g1 ¼ tC  tcrit 6 0 solution within the feasible region. Although increasing
g2 ¼ tD  tcrit P 0 ð8bÞ the penalty parameters improves the solution accuracy,
g3 ¼ f  fcrit P 0 the unconstrained formulation can also become very ill-
conditioned with large gradients and abrupt function
The physical significance of the constraints are as fol- changes. For the present problem, a value of pk = 1000
lows: The storage system for a CSP plant should be was found to strictly enforce the constraints, without mak-
charged in 6 h period (tcrit) which is the equivalent of excess ing the optimization problem formulation ill-conditioned.
energy available from solar irradiation during the day for a For instance, the 14 decision variables in a 3-PCM cas-
115 MWe power plant. Similarly, the storage system should caded EPCM-TES system configuration leads to a simplex
assist in the generation of electricity for more than 6 h dur- formed by fourteen-dimensional space and consists of fif-
ing the nighttime when solar energy is unavailable. The teen vertices. Initial guess for the decision variables corre-
exergy efficiency takes care of the amount of useful thermal spond to one of the vertices, and adding scaled basis unit
energy charged that can be converted into electricity and vectors creates the other fifteen vertices, such that the
should be greater than 95% (Stekli et al., 2013). Since for hyper-surfaces are independent of each other. The numer-
the various parameters considered, the exergetic efficiency ical model (Section 2.2) and the cost model (Section 3.2)
is greater than 95% (Figs. 2 and 3) the constraint is are then invoked to determine the objective function and
imposed on the overall exergetic efficiency. Moreover, the constraints respectively, corresponding to each vertex at
upper limits of the density and latent heat of fusion every step of the optimization procedure. Upon completion
considered in Table 4 are within the realistic properties of of the optimization algorithm, the decision variables corre-
available PCM. For instance, inorganic salt compositions sponding to the primary vertex are the optimal decision
such as 20% LiF + 80% LiOH and 46% LiF + 44% variable vector and the corresponding objective function
NaF2 + 10% MgF2 have latent heat of fusion values of value taken as the optimal storage cost, a.
869 kJ/kg and 858 kJ/kg, respectively. Some of the PCMs Table 5 lists the designs corresponding to the optimiza-
with high densities are 45% NaBr + 55% MgBr2 tion problems formulated by Eqs. (8a) and (8b). The stor-
(qp = 3490 kg/m3) and CaI2 (qp = 3956 kg/m3). age cost obtained for a non-cascaded configuration from
The optimization problem represented by Eqs. (8a) and the numerical optimization are much lower than those
(8b) is solved using the Nelder–Mead simplex method obtained from the parametric studies, cf. a = 5.79/kW ht
(Nelder and Mead, 1965) combined with a simulated (Table 5a), and a = 7.55/kW ht (Table 3c.1) for tank height
annealing technique to improve the effectiveness of the of 15 m. Similarly, the storage costs obtained for a 2-PCM
search (Mawardi and Pitchumani, 2003; Press et al., cascaded EPCM-TES system configuration a = $5.74/
1992). The simplex search method is an algorithm that per- kW ht (Table 5b) is much lower compared to that obtained
forms continuous search for selecting a new point during from parametric studies, a = 10.13/kW ht (Table 3c.2) for
an optimization iteration, which guarantees objective func- tank height of 15 m. On further observation, the optimum
tion improvement. A simplex is defined as a convex hull of values for a 2PCM cascaded configuration are obtained for
N + 1 vertices in an N-dimensional space, representing the a configuration with the maximum storage capacity, i.e.
N decision variables that govern the objective function with the highest PCM density and latent heat of fusion
evaluation. The vertices are ranked, from best to worst, (Table 5b). The optimum design parameters obtained for
based on the corresponding objective function evaluations all the non-cascaded and cascaded configurations in Table 5
and the best vertex is defined as the primary vertex. Since show that smallest capsule radius of Rc = 3.75 mm is pre-
the primary vertex corresponds to a set of decision vari- ferred due to the low encapsulation cost. The storage cost
ables, which corresponds to the lowest objective function for a 3-PCM cascaded configuration obtained from the
evaluation, the finding of a new primary vertex constitutes optimization studies is relatively higher compared to a
K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788 787

Table 5
Optimum design parameters obtained from optimization study. The minimum storage cost obtained from the optimization study for different system
configurations are listed in bold face.
hsl (kJ/kg) Tm (°C) qp (kg/m3) Rt (m) Rc (mm) tC (h) tD (h) a ($/kW ht) f (%)
a. 1-PCM Ht = 15 – 852.0 560.37 3951.0 10.70 3.75 5.968 6.104 5.791 98.976
b. 2-PCM H2,T = 7.73 Top 852.0 571.72 3951.0 9.31 3.75 5.935 6.402 5.739 98.359
H2,B = 7.27 Bottom 852.0 501.01 3951.0
c. 3-PCM H3,T = 3.30 Top 852.0 568.88 3951.0 8.76 3.75 5.643 6.618 5.800 96.896
H3,M=6.42 Mid 852.0 503.00 3951.0
H3,B=5.28 Bottom 852.0 462.67 3951.0

2-PCM cascaded configuration (Table 5b and c). However, above can be summarized as follows: Smaller radii capsules
an optimum 3-PCM cascaded configuration gives the high- yield higher discharge and charge time of the EPCM-TES
est discharge time among all the cases, cf. tD = 6.62 h for system, higher Reynolds number and consequently higher
tank height of 15 m. Compared to the result of the para- mass flow rate leads to decrease in discharge and charge
metric studies, the storage cost obtained for a 3-PCM time, higher inverse Stefan number and capacitance ratio
cascaded configuration from the optimization study is correspond to an EPCM-TES system with higher thermal
approximately two times lower. The optimization results inertia resulting in an increased charge and discharge time.
for 2-PCM and 3-PCM cascaded configuration in For all the various parameters and EPCM-TES system
Table 5b and c, respectively, also show that a larger per- configurations (non-cascaded, 2-PCM cascaded and 3-
centage of the tank should be filled with PCM’s with melt PCM cascaded) the exergetic efficiency at the periodic
temperature in the range T 0D  T m 6 T C , to satisfy the con- quasi-steady state was found to be greater than 95%.
straint on exergetic efficiency. From the optimization study Based on a systematic parametric analysis on the vari-
it can be concluded that a 2-PCM cascaded configuration ous performance metrics, feasible operating regimes and
offers the lowest storage cost and 3-PCM cascaded config- design conditions are identified which meet the technoeco-
uration offers the highest discharge and lowest charge nomic requirements: (a) charge time less than 6 h, (b) dis-
times. Further improvement in cascading will lead to only charge time greater than 6 h, (c) exergetic efficiency
minor improvement as seen from the negligible difference greater than 95% and (d) storage cost less than $15/kW ht.
in results between 2-PCM and 3-PCM cascaded The best combination of parameters that maximized the
configurations. various performance metrics are tabulated in Table 3.
The present study involves a detailed investigation of the The maximum discharge time of 7.42 h and a minimum
EPCM-TES system performance to obtain the optimum storage cost of $7.55/kW ht are obtained for a non-cas-
design and operating parameters of non-cascaded and cas- caded EPCM-TES tank of height, 15 m and radius,
caded configurations that meet the technoeconomic 11.25 m, filled with 15 mm radius capsules containing
requirements for cost parity of solar-generated electricity. PCM of density, 1975.5 kg/m3 and latent heat of fusion,
The authors have also investigated the EPCM-TES system 852 kJ/kg. Since the parametric studies allowed for analyz-
performance integrated with CSP plant performance model ing the performance of EPCM-TES by varying only one
to obtain design and operation windows which result in a parameter at a time, the effect of varying all the parameters
levelized cost of electricity less than 6 cents/kW h, to be simultaneously is studied by means of an optimization pro-
on par with the electricity cost associated with fossil fueled cedure. The optimum combination of parameters that min-
power plants (Nithyanandam and Pitchumani, 2014d). imized the storage cost is presented in Table 5. The
Based on the findings of Muren et al. (2011), future studies minimum storage cost of $5.539/kW ht is obtained for a
will involve coupling the numerical model for EPCM-TES 2-PCM cascaded EPCM-TES tank of height 15 m and
system with an optimization scheme to determine the opti- radius 9.31 m and filled with capsules of radius 3.75 mm.
mum cascaded EPCM-TES system configuration for realis- Overall, this paper presents a methodology for model
tic weather conditions. based design and optimization of the EPCM-TES system
based on the systematic parametric studies and consider-
5. Conclusions ation of target design requirements on the dynamic opera-
tion of the system.
An EPCM-TES model accounting for axial variation of
temperature in the HTF and radial temperature variation
in the PCM at any axial position is solved and the effects Acknowledgement
of various nondimensional design and operating parame-
ters on the dynamic performance of the storage system This work was supported by a Grant from the U.S.
are analyzed. Important results pertaining to the analysis Department of Energy SunShot Initiative under Award
788 K. Nithyanandam, R. Pitchumani / Solar Energy 107 (2014) 770–788

Number DE-FG36-08GO18146. Their support is gratefully Nithyanandam, K., Pitchumani, R., 2013a. Computational studies on a
acknowledged. latent thermal energy storage system with integral heat pipes for
concentrating solar power. Appl. Energy 103, 400–415.
Nithyanandam, K., Pitchumani, R., 2013b. Thermal energy storage with
References heat transfer augmentation using thermosyphons. Int. J. Heat Mass
Transf. 67, 281–294.
Beasley, D.E., Clark, J.A., 1984. Transient response of a packed bed for Nithyanandam, K., Pitchumani, R., 2014a. Computational modeling of
thermal energy storage. Int. J. Heat Mass Transf. 27, 1659–1669. dynamic performance of a latent thermal energy storage system with
Bertsekas, D.P., 1999. Nonlinear Programming. Athena Scientific, embedded heat pipes. ASME J. Sol. Eng. 136, 011010-1–9.
Belmont, MA. Nithyanandam, K., Pitchumani, R., 2014b. Computational studies on
Cárdenas, B., León, N., 2013. High temperature latent heat thermal metal foam and heat pipe enhanced latent thermal energy storage.
energy storage: phase change materials, design considerations and ASME J. Heat Transf. 136, 051503-1–10.
performance enhancement techniques. Renew. Sustain. Energy Rev. Nithyanandam, K., Pitchumani, R., 2014c. Analysis of a latent thermo-
27, 724–737. cline storage system with encapsulated phase change materials for
Felix Regin, A., Solanki, S.C., Saini, J.S., 2008. Heat transfer character- concentrating solar power. Appl. Energy 113, 1446–1460.
istics of thermal energy storage system using PCM capsule: a review. Nithyanandam, K., Pitchumani, R., 2014d. Cost and performance
Renew. Sustain. Energy Rev. 12, 2438–2458. analysis of concentrating solar power systems with integrated latent
Felix Regin, A., Solanki, S.C., Saini, J.S., 2009. An analysis of a packed thermal energy storage. Energy 64, 793–810.
bed latent heat thermal energy storage system using PCM capsules: Patankar, S.V., 1980. Numerical Heat Transfer and Fluid Flow. Hemi-
numerical investigation. Renew. Energy 34, 1765–1773. sphere, Washington, DC.
Flueckiger, S., Yang, Z., Garimella, S.V., 2013. Review of molten salt Pendyala, S., 2012. Macroencapsulation of phase change materials for
thermocline tank modeling for solar thermal energy storage. Heat thermal energy storage. M.S Thesis, University of South Florida.
Transf. Eng. 34, 787–800. Press, W.H., Flannery, B.P., Teukolsky, S.A., Vettering, W.T., 1992.
Galloway, T.R., Sage, B.H., 1970. A model of the mechanism of transport Numerical Recipes in FORTRAN. Cambridge University Press, New
in packed, distended, and fluidized beds. Chem. Eng. Sci. 25, 495–516. York, NY.
Glatzmaier, G., 2011. Developing a cost model and methodology to System Advisor Model Version 2012.5.11 (SAM 2012.5.11), User Docu-
estimate capital costs for thermal energy storage. National Renewable mentation. National Renewable Energy Laboratory, Golden CO.
Energy Laboratory, NREL/TP-550053066, pp. 1–17. Schumann, T.E.W., 1929. Heat transfer: a liquid flowing through a porous
Gong, Z.X., Mujumdar, A.S., 1997. Thermodynamic optimization of the prism. J. Franklin Inst. 208, 405–416.
thermal process in energy storage using multiple phase change Shitzer, A., Levy, M., 1983. Transient behavior of a rock-bed thermal
materials. Appl. Therm. Eng. 17, 1067–1083. storage system subjected to variable inlet air temperature: analysis and
Hales, T.C., 2006. Historical overview of the Kepler conjecture, discrete & experimentation. ASME J. Sol. Energy Eng. 105, 200–206.
computational geometry. Int. J. Math. Comput. Sci. 36, 5–20. Singh, H., Saini, R.P., Saini, J.S., 2011. A review on packed bed solar
Herrmann, U., Kelly, B., Price, H., 2004. Two-tank molten salt storage for energy storage system. Renew. Sustain. Energy Rev. 14, 1059–1069.
parabolic trough solar power plants. Energy 29, 883–893. Stekli, J., Irwin, L., Pitchumani, R., 2013. Technical challenges and
MEPS International Ltd. <www.meps.co.uk> (viewed November–Decem- opportunities for concentrating solar power with energy storage.
ber 2012). ASME J. Therm. Sci. Eng. Appl. 5, 021011-1–12.
Ismail, K.A.R., Henriquez, J.R., 1999. Numerical and experimental study SunShot, Energy Efficiency and Renewable Energy, U.S. Department of
of spherical capsules packed bed latent heat storage system. Appl. Energy, 2012. SunShot Vision Study: February 2012. NREL Report
Therm. Eng. 19, 757–788. No. BK5200-47927; DOE/GO-102012-3037. <http://www.solar.ener-
Kelly, B., Kearney, D., 2006. Thermal storage commercial plant design for gy.gov/pdfs/47927.pdf> (accessed July 2014).
a 2-tank indirect molten salt system. National Renewable Energy Van Lew, J.T., Li, P., Chan, C.L., Karaki, W., Stephens, J., 2011. Analysis
Laboratory, NREL/SR-550040166, pp. 1–32. of heat storage and delivery of a thermocline tank having solid filler
Kenisarin, M.M., 2010. High-temperature phase change materials for material. J. Sol. Energy Eng. 133, 021003-1–10.
thermal energy storage. Renew Sustain. Energy Rev. 14, 955–970. Voller, V.R., Cross, M., Markatos, N.C., 1987. An enthalpy method for
Mathur, A., 2012. Personal Communication. Terrafore, Inc.. convection/diffusion phase change. Int. J. Numer. Methods 24, 271–
Mathur, A., Kasetty, R., Oxley, J., Mendez, J., Nithyanandam, K., 2013. 284.
Using encapsulated phase change salts for concentrated solar power Wakao, N., Kagei, S., 1982. Heat and Mass Transfer in Packed Beds.
plant. In: Pitchumani, R. (Ed.), Energy Procedia, Proceedings of Gordon and Breach Science, New York.
SolarPACES 2013, Las Vegas, Nevada, 7p. Wang, J., Chen, G., Jiang, H., 1999. Theoretical study on a novel phase
Mawardi, A., Pitchumani, R., 2003. Optimal temperature and current change process. Int. J. Energy Res. 23, 287–294.
cycles for curing of composites using internal resistive heating. ASME Wu, S., Fang, G., 2011. Dynamic performance of solar heat storage
J. Heat Transf. 125, 126–136. system with packed bed using myristic acid as phase change material.
Mettawee, E.S., Assassa, G.M.R., 2007. Thermal conductivity enhance- Energy Build. 43, 1091–1096.
ment in a latent heat storage system. Sol. Energy 81, 839–845. Yang, Z., Garimella, S.V., 2010. Thermal analysis of solar thermal energy
Muren, R., Arias, D., Chapman, D., Erickson, L., Gavilan, A., 2011. storage in a molten-salt thermocline. Sol. Energy. 84, 974–985.
Coupled transient system analysis: a new method of passive thermal Yang, Z., Garimella, S.V., 2010. Molten-salt thermal energy storage in
energy storage modeling for high temperature concentrated solar thermoclines under different environmental boundary conditions.
power systems. In: Proceedings of the ASME 5th International Appl. Energy 87, 3322–3329.
Conference on Energy Sustainability, Paper No. ES2011-54111, Yang, Z., Garimella, S.V., 2013. Cyclic operation of molten-salt thermal
Washington, DC, USA. energy storage in thermoclines for solar power plants. App. Energy
Nelder, J.A., Mead, R., 1965. A simplex method for function minimiza- 103, 256–265.
tion. Comput. J. (UK) 7, 308–313. Zhao, Z.Y., Wu, Z.G., 2011. Heat transfer enhancement of high
Nithyanandam, K., Pitchumani, R., 2011. Analysis and optimization of a temperature thermal energy storage using metal foams and expanded
latent thermal energy storage system with embedded heat pipes. Int. J. graphite. Sol. Energy Mater. Sol. Cells 95, 636–643.
Heat Mass Transf. 54, 4596–4610.

You might also like